14.01.2013 Views

Analytical Chemistry Chemical Cytometry Quantitates Superoxide

Analytical Chemistry Chemical Cytometry Quantitates Superoxide

Analytical Chemistry Chemical Cytometry Quantitates Superoxide

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Anal. Chem. 2010, 82, 6745–6750<br />

Letters to <strong>Analytical</strong> <strong>Chemistry</strong><br />

<strong>Chemical</strong> <strong>Cytometry</strong> <strong>Quantitates</strong> <strong>Superoxide</strong><br />

Levels in the Mitochondrial Matrix of Single<br />

Myoblasts<br />

Xin Xu and Edgar A. Arriaga*<br />

Department of <strong>Chemistry</strong>, University of Minnesota, Minneapolis, Minnesota 55455<br />

Triphenylphosphonium hydroethidine (TPP-HE) is a membrane-permeable<br />

probe that reacts with superoxide and<br />

forms hydroxytriphenylphosphonium ethidium (OH-TPP-<br />

E + ), a fluorescent product that has been previously<br />

used in qualitative measurements of superoxide production.<br />

In order to develop quantitative methods to<br />

measure superoxide, it is necessary to take into<br />

consideration the principles that drive TPP-HE accumulation<br />

into various subcellular compartments. In<br />

the mitochondria matrix, TPP-HE accumulation depends<br />

on the mitochondrial membrane potential, which<br />

varies from cell to cell. Here we address this issue by<br />

including rhodamine 123 (R123) as an internal mitochondrial<br />

membrane potential calibrant in chemical<br />

cytometry experiments. After loading with TPP-HE and<br />

R123, a single cell is lysed within a separation capillary<br />

and its contents are separated and detected by micellar<br />

electrokinetic capillary chromatography with laserinduced<br />

fluorescence detection (MEKC-LIF). Using<br />

theoretical arguments, we show that the ratio [OH-TPP-<br />

E + ]/[R123] is adequate to obtain a relative quantitation<br />

of mitochondrial matrix superoxide levels for each<br />

analyzed cell. We applied this method to single skeletal<br />

muscle myoblasts and determined that the steady state<br />

superoxide levels in the mitochondrial matrix is ∼(0.29<br />

( 0.10) × 10 -12 M. The development of this quantitative<br />

method is a critical step toward establishing the<br />

importance of reactive oxygen species in biological<br />

systems, including those relevant to aging and disease.<br />

<strong>Superoxide</strong> is a reactive oxygen species (ROS) that plays a<br />

decisive role in the generation of secondary reactive radicals, 1,2<br />

which could result in oxidative damage associated with multiple<br />

diseases and aging. 2-4 As a probe to detect superoxide, 5 triph-<br />

* Corresponding author. Phone: 1-612-624-8024. Fax: 1-612-626-7541. E-mail:<br />

arriaga@umn.edu.<br />

(1) St.-Pierre, J.; Buckingham, J. A.; Roebuck, S. J.; Brand, M. D. J. Biol. Chem.<br />

2002, 277, 44784.<br />

(2) Orrenius, S.; Gogvadze, V.; Zhivotovsky, B. Annu. Rev. Pharmacol. Toxicol.<br />

2007, 47, 143.<br />

(3) Ashok, B. T.; Ali, R. Exp. Gerontol. 1999, 34, 293.<br />

(4) Beckman, K. B.; Ames, B. N. Physiol. Rev. 1998, 78, 547.<br />

enylphosphonium hydroethidine (TPP-HE, also known as (aka)<br />

MitoSOX Red) has several advantages over commonly used<br />

probes such as dichlorodihydrofluorescein (H2DCF) and hydroethidine.<br />

TPP-HE is more selective than H2DCF 6 and forms<br />

the product hydroxytriphenylphosphonium ethidium (OH-TPP-<br />

E + ) upon reaction with superoxide. 5 Other reactive species also<br />

react with TPP-HE, but relative to the reaction with an<br />

equivalent level of superoxide the OH-TPP-E + formed is a small<br />

fraction (i.e., ∼0.4%, ∼6%, and ∼4% for the reactions with<br />

hydrogen peroxide, hydroxyl radical, and hypochlorite, respectively).<br />

7 In comparison with hydroethidine, 8 TPP-HE accumulates<br />

in the mitochondrial matrix of cells according to the<br />

mitochondrial membrane potential, 9 making it possible to<br />

detect superoxide in the matrix even in the presence of<br />

competing superoxide dismutase (SOD). 5,7 Previous studies<br />

based on the use of TPP-HE are, however, only qualitative because<br />

the dependence of TPP-HE accumulation on the mitochondrial<br />

membrane potential was not taken into account. 5,7,10-12 Variations<br />

in mitochondrial membrane potential between cells could cause<br />

different TPP-HE distributions and bias the measured ROS<br />

levels. 7,13<br />

Here we demonstrate that chemical cytometry can be used to<br />

quantitate superoxide levels in the mitochondrial matrix of single<br />

myoblasts. Single myoblasts were simultaneously treated with<br />

both probes, TPP-HE and rhodamine 123 (R123), before the<br />

(5) Robinson, K. M.; Janes, M. S.; Pehar, M.; Monette, J. S.; Ross, M. F.; Hagen,<br />

T. M.; Murphy, M. P.; Beckman, J. S. Proc. Natl. Acad. Sci. U.S.A. 2006,<br />

103, 15038.<br />

(6) McArdle, F.; Pattwell, D. M.; Vasilaki, A.; McArdle, A.; Jackson, M. J. Free<br />

Radical Biol. Med. 2005, 39, 651.<br />

(7) Xu, X.; Arriaga, E. A. Free Radical Biol. Med. 2009, 46, 905.<br />

(8) Zhao, H.; Joseph, J.; Fales, H. M.; Sokoloski, E. A.; Levine, R. L.; Vasquez-<br />

Vivar, J.; Kalyanaraman, B. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 5727.<br />

(9) Ross, M. F.; Kelso, G. F.; Blaikie, F. H.; James, A. M.; Cocheme, H. M.;<br />

Filipovska, A.; Da Ros, T.; Hurd, T. R.; Smith, R. A. J.; Murphy, M. P.<br />

Biochemistry (Moscow) 2005, 70, 222.<br />

(10) Mukhopadhyay, P.; Rajesh, M.; Yoshihiro, K.; Hasko, G.; Pacher, P.<br />

Biochem. Biophys. Res. Commun. 2007, 358, 203.<br />

(11) Han, Z.; Varadharaj, S.; Giedt, R. J.; Zweier, J. L.; Szeto, H. H.; Alevriadou,<br />

B. R. J. Pharmacol. Exp. Ther. 2009, 329, 94.<br />

(12) Watanabe, N.; Zmijewski, J. W.; Takabe, W.; Umezu-Goto, M.; Le Goffe,<br />

C.; Sekine, A.; Landar, A.; Watanabe, A.; Aoki, J.; Arai, H.; Kodama, T.;<br />

Murphy, M. P.; Kalyanaraman, R.; Darley-Usmar, V. M.; Noguchi, N. Am. J.<br />

Pathol. 2006, 168, 1737.<br />

(13) Zielonka, J.; Kalyanaraman, B. Free Radical Biol. Med. 2010, 48, 983.<br />

10.1021/ac101509d © 2010 American <strong>Chemical</strong> Society 6745<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/20/2010


micellar electrokinetic capillary chromatography with laserinduced<br />

fluorescence detection (MEKC-LIF) analysis. An intact<br />

cell was then introduced into a narrow-bore capillary and lysed<br />

therein. Subsequently, the R123 and OH-TPP-E + contents released<br />

from each cell were analyzed by MEKC-LIF. 7 Since both<br />

R123 and TPP-HE accumulate into mitochondria according to<br />

the mitochondrial membrane potential, the mole ratio of OH-<br />

TPP-E + /R123 is approximately proportional to the steady state<br />

superoxide concentration in the mitochondrial matrix of<br />

individual cells with polarized mitochondria. Thus, determination<br />

of this ratio make it possible to quantitate superoxide in<br />

single cells. In the future, this method could be used for<br />

investigating the role of superoxide in xenobiotic toxicity, aging,<br />

and disease.<br />

THEORY<br />

When cells are incubated with TPP-HE and R123, both probes<br />

accumulate in their mitochondria according to the mitochondrial<br />

membrane potential (∆ψ). 5,14 The relevant Nernst equations are<br />

Thus<br />

∆ψ )- RT<br />

nF ln [TPP-HE] inside<br />

[TPP-HE] outside<br />

∆ψ )- RT<br />

nF ln [R123] inside<br />

[R123] outside<br />

[TPP-HE] inside ) [R123] inside<br />

[TPP-HE] outside<br />

[R123] outside<br />

where [TPP-HE]inside and [R123]inside, and [TPP-HE]outside and<br />

[R123]outside are the concentrations of TPP-HE and R123 in the<br />

mitochondrial matrix and outside the mitochondria, respectively.<br />

In this study, [TPP-HE]outside is 10 µM and [R123]outside<br />

is 50 nM.<br />

The reaction of TPP-HE with superoxide and its relevant<br />

kinetics equations are<br />

k<br />

-<br />

TPP-HE + O2 f<br />

d[OH-TPP-E + ]<br />

dt<br />

OH-TPP-E +<br />

[OH-TPP-E + ] T ) k[TPP-HE]∫ 0<br />

(1)<br />

(2)<br />

(3)<br />

-<br />

) k[TPP-HE][O2 ] (4)<br />

T<br />

-<br />

[O2 ]dt (5)<br />

where k is the kinetic rate of the reaction of TPP-HE with<br />

superoxide (∼4 × 10 6 M -1 s -1 ), 5 T is the incubation time of the<br />

cells with TPP-HE (60 min), and ∫0 T [O2 - ]dt is the integral of<br />

steady state superoxide concentration over the entire incubation<br />

time. If [O2 - ] is constant during the incubation time,<br />

T<br />

- - ∫ [O<br />

0 2 ]dt ) [O2 ]averageT (6)<br />

(14) Scaduto, R. C.; Grotyohann, L. W. Biophys. J. 1999, 76, 469.<br />

6746 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

where [O2 - ]average is the average steady state superoxide<br />

concentration over the length of the incubation time.<br />

When a cell is lysed and analyzed by MEKC-LIF, the detected<br />

OH-TPP-E + is the total from both the matrix and outside the<br />

mitochondria of the cell. 7 Thus, the amount of OH-TPP-E + in<br />

each individual cell is<br />

m OH-TPP-E+ ) m OH-TPP-E+, inside + m OH-TPP-E+, outside<br />

where mOH-TPP-E + ,inside and mOH-TPP-E + ,outside are the amounts of OH-<br />

TPP-E + in the mitochondrial matrix and outside the mitochondria<br />

of a given cell, respectively. On the basis of the results<br />

from bulk analysis (Supporting Information, Part I; Figure S-1),<br />

the mOH-TPP-E + ,outside is a small fraction of the mOH-TPP-E + ,inside under<br />

basal conditions and upon treatments with rotenone and<br />

antimycin A. To simplify eq 7, the amount of OH-TPP-E + in one<br />

cell can be expressed as<br />

m OH-TPP-E+ ) (1 + x)m OH-TPP-E+, inside<br />

where x is the fraction of mOH-TPP-E + ,outside relative to mOH-TPP-E + ,inside.<br />

This fraction “x” represent the bias of the measurement due<br />

to the superoxide found outside the mitochondria. On the basis<br />

of bulk measurements, x is ∼0.048 under basal conditions,<br />

∼0.063 upon treatment with rotenone, and ∼0.071 upon<br />

treatment with antimycin A (Supporting Information, Part I).<br />

Note that eq 8 is not applicable when mitochondria are depolarized<br />

by carbonyl cyanide m-chlorophenylhydrazone (CCCP) treatment<br />

(Supporting Information, Part II).<br />

With the combination of eqs 5, 6, and 8,<br />

m ) OH-TPP-E+ (1 + x)k[TPP-HE] outside<br />

-<br />

mR123 [O2 ]average,insideT [R123] outside<br />

(9)<br />

where [O2 - ]average,inside is the average of steady state superoxide<br />

concentration in the mitochondrial matrix of single cells over<br />

the incubation time.<br />

On the basis of the MEKC-LIF calibration curves of OH-TPP-<br />

E + and R123, the following equations are yielded<br />

and<br />

(7)<br />

(8)<br />

m OH-TPP-E+ ) aA OH-TPP-E+ + m (10)<br />

m R123 ) bA R123 + n (11)<br />

where the respective slopes are a and b, and the respective<br />

intercepts are m and n, and the respective peak areas in the<br />

electropherograms are AOH-TPP-E + and AR123.<br />

Thus,<br />

-<br />

[O2 ]average,inside ) K<br />

1 (aA + m)<br />

OH-TPP-E+<br />

(1 + x) (bAR123 + n)<br />

(12)<br />

where K ) ([R123]outside)/(k[TPP-HE]outsideT) ≈ 3.47 × 10 -13 M.<br />

EXPERIMENTAL SECTION<br />

<strong>Chemical</strong>s. Rhodamine 123 (R123), TPP-HE (aka MitoSOX<br />

Red), and tetramethylrhodamine methyl ester (TMRM) were


purchased from Invitrogen-Molecular Probes (Eugene, OR). OH-<br />

TPP-E + was synthesized and purified following a previously<br />

described procedure. 15 Its purity was verified by MEKC-LIF.<br />

Its concentration was determined spectrophotometrically using<br />

the reported molar extinction coefficient at 478 nm. 5,15 OH-<br />

TPP-E + is stable for at least 1 month when stored at -20 °C. 15<br />

All the other reagents were obtained from Sigma-Aldrich (St.<br />

Louis, MO). The MEKC running buffer contained 10 mM<br />

sodium borate and 1 mM cetyltrimethylammonium bromide<br />

(CTAB) (pH 9.3). All buffers were made with Milli-Q deionized<br />

water and filtered through 0.22 µm filters before use.<br />

Cell Culture. Rat muscle L6 myoblast cell line was obtained<br />

from the American Tissue Culture Collection (Manassas, VA). The<br />

cells were cultured in Dulbecco’s modified eagle medium (DMEM)<br />

containing 10% (v/v) calf serum and 10 µg/mL gentamicin at 37<br />

°C and 5% CO2. The cells were maintained by splitting them<br />

every 3-4 days.<br />

Cell Treatments. The cultured myoblasts were incubated with<br />

10 µM TPP-HE and 50 nM R123 in DMEM at 37 °C for 1 h. When<br />

needed, before incubation in the presence of TPP-HE and R123,<br />

cells were treated with either 1 mM tiron, 50 µM carbonyl cyanide<br />

m-chlorophenylhydrazone (CCCP), 5 µM rotenone, or 5 µM<br />

antimycin A. After treatments, the cells were washed two times<br />

with DMEM, trypsinized, and resuspended in PBS for either whole<br />

cell lysate or single cell analysis.<br />

MEKC-LIF. To do the whole cell lysate analysis, ∼2 × 10 6<br />

cells in PBS were centrifuged down at 600g for 5 min. The cell<br />

pellet was then dissolved in running buffer, treated with 2 mg/<br />

mL proteinase K and 400U/mL DNase 1, and analyzed by<br />

MEKC-LIF following previously described procedures. 7 Briefly,<br />

separations were carried out at -400 V/cm in MEKC running<br />

buffer. 7 The 488 nm line (12 mW) of an argon-ion laser (Melles<br />

Griot, Irvine, CA) was used for excitation, and fluorescence<br />

was selected with a bandpass filter transmitting in the 607-662<br />

nm range (Omega Optical, Brattleboro, VT).<br />

To do single-cell analysis, a cell was injected into a 150 µm<br />

o.d., 30 µm i.d. fused silica capillary (Polymicro Technologies,<br />

Phoenix, AZ) following previously reported procedures. 16,17 After<br />

the cell lysis in the capillary, the MEKC-LIF procedure was done<br />

as described for bulk analysis. Upon completion of the separation,<br />

the capillary was washed for 2 min with 0.1 M NaOH and 5 min<br />

with MEKC running buffer between runs.<br />

Separate calibration curves for OH-TPP-E + and R123, (area<br />

in electropherogram versus amount injected) were built by<br />

injecting standards of these compounds. The resulting calibration<br />

curves for OH-TPP-E + and R123 were<br />

m ) (122.4 ( 0.4)A + (1.0 ( 0.6);<br />

OH-TPP-E+ OH-TPP-E+<br />

(R 2 ) 0.99) (13)<br />

mR123 ) (12.9 ( 0.0)AR123 + (0.3 ( 0.1); (R 2 ) 0.99)<br />

(14)<br />

where m is given in attomoles, A is the peak area in the<br />

electropherogram, and the errors are standard deviations. These<br />

(15) Zielonka, J.; Vasquez-Vivar, J.; Kalyanaraman, B. Nat. Protoc. 2008, 3, 8.<br />

(16) Krylov, S. N.; Starke, D. A.; Arriaga, E. A.; Zhang, Z.; Chan, N. W.; Palcic,<br />

M. M.; Dovichi, N. J. Anal. Chem. 2000, 72, 872.<br />

(17) Anderson, A. B.; Gergen, J.; Arriaga, E. A. J. Chromatogr., B: Anal. Technol<br />

Biomed. Life Sci. 2002, 769, 97.<br />

Figure 1. Electropherograms of TPP-HE oxidation products and<br />

R123 in the whole cell lysate (∼2 × 10 6 myoblasts) and the blank<br />

buffer. Separations were performed in a 42 cm-long capillary at -400<br />

V/cm in MEKC running buffer. The 488 nm line of an argon-ion laser<br />

was used for excitation, and a 607-663 nm bandpass filter was used<br />

for detection. Traces have been offset vertically for clarity. The<br />

respective electrophoretic mobilities of OH-TPP-E + and R123 were<br />

-(6.89 ( 0.11) × 10 -4 and -(6.41 ( 0.09) × 10 -4 cm 2 V -1 s -1 (n )<br />

3). Separation efficiencies were calculated to be ∼260 000 for OH-<br />

TPP-E + and ∼31 000 for R123.<br />

equations were used to carry out calculations described by the<br />

eqs 10 and 11. The limits of detection (LOD at signal/noise ) 3)<br />

were ∼2 amol for OH-TPP-E + and ∼0.2 amol for R123. It would<br />

be possible to obtain a better LOD (e.g., zeptomole levels) for<br />

R123 by using a different bandpass filter (e.g., centered at 530<br />

nm). This alternative was not pursued here.<br />

Data Analysis. The MEKC electropherograms were analyzed<br />

using Igor Pro 5.0 software (Wavemetrics, Lake Oswego, OR).<br />

<strong>Superoxide</strong> concentrations in single myoblasts under each condition<br />

were reported as mean ± standard deviation (SD). A student’s<br />

t test was used to determine statistical significance of the data,<br />

with p values of


of OH-TPP-E + in the whole cell lysate (cf. eq 13), on average<br />

each cell contains ∼7.7 amol. However, this bulk analysis is biased<br />

because it ignores that cells do not necessarily have the same<br />

membrane potential. 18-20<br />

It is important to bear in mind that superoxide dismutases<br />

(SODs) are present in the matrix (Mn-SOD) and outside the<br />

mitochondria (Cu, Zn-SOD). 5 In general SODs transform superoxide<br />

into hydrogen peroxide (k ∼ 2 × 10 9 M -1 s -1 ), 21-23<br />

outcompeting the relatively slow reaction of superoxide with the<br />

TPP-HE probe (i.e., k ∼ 4 × 10 6 M -1 s -1 ). 5 However, in the<br />

mitochondrial matrix, the mitochondrial membrane potential<br />

increases the TPP-HE concentration ∼100-fold relative to the<br />

cytosol 7 that partially compensates for its relative slow reaction<br />

rate constant with superoxide. The concentration of Mn-SOD<br />

in the matrix has been reported to be ∼20 µM. 24 On the basis<br />

of this value and the kinetic rate constants of TPP-HE and Mn-<br />

SOD with superoxide (and assuming that there are no other<br />

reactions consuming superoxide), the estimated rate of OH-<br />

TPP-E + accumulation is ∼10% of the formation rate of H2O2<br />

by Mn-SOD. This simplified kinetic analysis argues that, at this<br />

relative rate of formation, the detected amounts of OH-TPP-<br />

E + are adequate to calculate [O2 - ]average,inside (eq 12), even in<br />

the presence of competing Mn-SOD.<br />

On the other hand, outside the mitochondria there is no<br />

enhanced accumulation of TPP-HE. Given a concentration of Cu,<br />

Zn-SOD ∼6 µM, 25 the rate of OH-TPP-E + accumulation outside<br />

the mitochondria is only ∼0.3% of the rate of H2O2 formed by<br />

Cu, Zn-SOD. This calculation further confirms that OH-TPP-<br />

E + produced outside the mitochondria does not contribute<br />

significantly to the total OH-TPP-E + detected (i.e., “x” ineq8<br />

is small; cf. Supporting Information, Part I).<br />

Single Cell <strong>Superoxide</strong> Quantitative Analysis. We used<br />

chemical cytometry to analyze mitochondrial matrix superoxide<br />

levels in single cells (Figure 2). In order to correct for variations<br />

in mitochondrial membrane potentials observed among individual<br />

cells, cells were treated with TPP-HE and R123. The MEKC-LIF<br />

separation and detection of OH-TPP-E + and R123 in single<br />

myoblasts provides the measurements needed to determine<br />

superoxide at the single-cell level (cf. eq 12).<br />

The detection of mitochondrial matrix superoxide at the single<br />

cell level was further confirmed by treatment with tiron and CCCP.<br />

Tiron is an SOD mimetic that scavenges intracellular superoxide<br />

in cultured cells. 6,26 As expected, OH-TPP-E + was not detected<br />

in single myoblasts treated with tiron (Figure 2). The R123 peak<br />

was still detected showing that tiron treatment does not compro-<br />

(18) Johnson, L. V.; Walsh, M. L.; Bockus, B. J.; Chen, L. B. J. Cell Biol. 1981,<br />

88, 526.<br />

(19) Ludovico, P.; Sansonetty, F.; Corte-Real, M. Microbiology 2001, 147, 3335.<br />

(20) Heerdt, B. G.; Houston, M. A.; Augenlicht, L. H. Cancer Res. 2005, 65,<br />

9861.<br />

(21) Klug, D.; Rabani, J.; Fridovich, I. J. Biol. Chem. 1972, 247, 4839.<br />

(22) Pick, M.; Rabani, J.; Yost, F.; Fridovich, I. J. Am. Chem. Soc. 1974, 96,<br />

7329.<br />

(23) Behar, D.; Czapski, G.; Rabani, J.; Dorfman, L. M.; Schwarz, H. A. J. Phys.<br />

Chem. 1970, 74, 3209.<br />

(24) Quijano, C.; Hernandez-Saavedra, D.; Castro, L.; McCord, J. M.; Freeman,<br />

B. A.; Radi, R. J. Biol. Chem. 2001, 276, 11631.<br />

(25) Chang, L. Y.; Slot, J. W.; Geuze, H. J.; Crapo, J. D. J. Cell Biol. 1988, 107,<br />

2169.<br />

(26) Yamada, J.; Yoshimura, S.; Yamakawa, H.; Sawada, M.; Nakagawa, M.; Hara,<br />

S.; Kaku, Y.; Iwama, T.; Naganawa, T.; Banno, Y.; Nakashima, S.; Sakai, N.<br />

Neurosci. Res. 2003, 45, 1.<br />

6748 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 2. Electropherograms of TPP-HE oxidation products and<br />

R123 in individual myoblasts under basal conditions and upon<br />

treatments with tiron and CCCP. Separations and detection conditions<br />

were same to those described for Figure 1.<br />

mise the mitochondrial membrane potential. CCCP is a proton<br />

ionophore that dissipates the mitochondrial membrane potential. 27,28<br />

Upon CCCP treatment both R123 and TPP-HE concentrations in<br />

the matrix decrease ∼100-fold relative to untreated cells (Supporting<br />

Information, Parts II and III). The expected R123 amount<br />

in a CCCP-treated cell is ∼0.2 amol of R123 (Supporting Information,<br />

Part III), which is undetectable in the MEKC-LIF system<br />

used in this study (i.e., limit of detection ∼0.2 amol of R123).<br />

Furthermore, the low TPP-HE concentration in CCCP-treated cells<br />

significantly decreases the rate of formation of OH-TPP-E + inside<br />

the mitochondria (cf. eq 5; Supporting Information, Part II). The<br />

expected amount of OH-TPP-E + formed in a CCCP-treated<br />

myoblast is ∼0.4 amol, which is undetectable in the MEKC-<br />

LIF that was used in this study (i.e., limit of detection ∼2 amol<br />

of OH-TPP-E + ). Thus, the concomitant disappearance of the<br />

R123 and the OH-TPP-E + peaks in CCCP-treated myoblasts<br />

(Figure 2) demonstrates that the mitochondrial membrane<br />

potential is vital to the accumulation of TPP-HE and formation of<br />

OH-TPP-E + in the mitochondrial matrix. 5<br />

On the basis of fluorescence microscopy, individual myoblasts<br />

display ∼39% variation in mitochondrial membrane potential<br />

(Supporting Information, Part V; Figure S-3). In MEKC-LIF, the<br />

ratio of the areas of OH-TPP-E + and R123 corrects for variations<br />

in mitochondrial membrane potential between individual cells<br />

(cf. eq 12). This calculation is based on the assumptions that both<br />

R123 and TPP-HE accumulate into the mitochondria according<br />

to a Nernstian behavior (cf. eqs 1 and 2) and that the amount of<br />

superoxide product outside the mitochondria is only a small<br />

fraction of that in the mitochondrial matrix (cf. eqs 7 and 8;<br />

Supporting Information, Part I). On the basis of MEKC-LIF, the<br />

steady state superoxide concentration in the mitochondrial matrix<br />

of single myoblasts under basal conditions was ∼(0.29 ± 0.10) ×<br />

10 -12 M with a RSD of 35% (n ) 12) (Figure 4). This variance<br />

is statistically different from that obtained if the signal of OH-<br />

(27) Heytler, P. G.; Prichard, W. W. Biochem. Biophys. Res. Commun. 1962, 7,<br />

272.<br />

(28) Lim, M. L.; Minamikawa, T.; Nagley, P. FEBS Lett. 2001, 503, 69.


Figure 3. Electropherograms of TPP-HE oxidation products and<br />

R123 in individual myoblasts under basal conditions and upon<br />

treatments with rotenone and antimycin. Separations and detection<br />

conditions were same to those described for Figure 1.<br />

TPP-E + is not corrected by membrane potential (RSD ∼69%)<br />

(F-test: p < 0.05). In summary, the simultaneous monitoring<br />

of OH-TPP-E + and R123 provides a correction for variations in<br />

membrane potential and provides the means to quantitate<br />

mitochondrial matrix superoxide levels in single cells.<br />

Effects of Mitochondrial Respiratory Inhibitors. The mitochondria<br />

is considered a major source of superoxide in cells,<br />

where superoxide is mainly released by complex I and III in the<br />

mitochondrial electron transport chain (ETC). 1,2 Rotenone and<br />

antimycin A are two respiratory inhibitors that block electron<br />

transfer through complex I and III, respectively, thereby stimulating<br />

superoxide production. 29,30 Although recent studies have<br />

utilized the TPP-HE probe to investigate the effects of rotenone<br />

and antimycin A on intracellular superoxide release in single cells<br />

by fluorescent microscopy and/or flow cytometry, 5,10-12 these<br />

studies are qualitative because they have not considered the effect<br />

of mitochondrial membrane potential.<br />

Here we demonstrated that chemical cytometry is suitable<br />

to quantify mitochondrial matrix superoxide levels in individual<br />

myoblasts upon their treatment with either rotenone or antimycin<br />

A (Figure 3). Some studies have reported such inhibitors<br />

at elevated concentrations may alter the membrane potential<br />

of isolated mitochondria. 31-34 In this report, bulk analyses<br />

showed that the treatments with rotenone and antimycin A did<br />

appear to alter the mitochondria membrane potentials because<br />

there were no significant changes in the ratio of<br />

mOH-TPP-E + ,outside to mOH-TPP-E + ,inside compared to basal conditions<br />

(29) Grivennikova, V. G.; Vinogradov, A. D. Biochim. Biophys. Acta 2006, 1757,<br />

553.<br />

(30) Muller, F. L.; Liu, Y. H.; Van Remmen, H. J. Biol. Chem. 2004, 279, 49064.<br />

(31) Petit, P. X.; O’Connor, J. E.; Grunwald, D.; Brown, S. C. Eur. J. Biochem.<br />

1990, 194, 389.<br />

(32) Kataoka, M.; Fukura, Y.; Shinohara, Y.; Baba, Y. Electrophoresis 2005, 26,<br />

3025.<br />

(33) Tzung, S. P.; Kim, K. M.; Basanez, G.; Giedt, C. D.; Simon, J.; Zimmerberg,<br />

J.; Zhang, K. Y.; Hockenbery, D. M. Nat. Cell Biol. 2001, 3, 183.<br />

(34) De Bari, L.; Atlante, A.; Guaragnella, N.; Principato, G.; Passarella, S.<br />

Biochem. J. 2002, 365, 391.<br />

Figure 4. Effects of rotenone and antimycin A on the superoxide<br />

levels detected in the mitochondrial matrix of single myoblasts. Data<br />

of individual cells as well as means ( SD are presented for each<br />

condition (n ) 12 cells for basal, 10 for rotenone, and 12 for antimycin<br />

A). * stands for p < 0.05 vs basal.<br />

(Supporting Information, Part I). Thus, eq 12 is adequate to<br />

determine mitochondrial matrix superoxide levels following<br />

inhibitor treatments.<br />

In single myoblasts treated with rotenone or antimycin A, the<br />

steady state superoxide concentrations in the mitochondrial matrix<br />

were (0.97 ± 0.61) × 10 -12 M(n ) 10) and (2.15 ± 1.20) × 10 -12<br />

M(n ) 12), respectively (Figure 4). These values are ∼3.5- and<br />

8.2-fold higher compared to basal levels. The magnitude of such<br />

changes are comparable with the ∼2 to 3-fold change observed<br />

in aged cells or cells associated with diabetic hyperglycemia<br />

relative to their respective controls. 35-37 Therefore, the use of<br />

inhibitors demonstrates that chemical cytometry is a suitable<br />

approach to quantitate mitochondrial superoxide levels expected<br />

in aging and disease related studies.<br />

CONCLUDING REMARKS<br />

Here we report the quantitation of mitochondrial matrix<br />

superoxide levels in single cells using a MEKC-based chemical<br />

cytometry method that is not biased by the mitochondrial<br />

membrane potential. On the basis of eq 12, the ratio of OH-<br />

TPP-E + and R123 signals corrects for variation in membrane<br />

potential of mitochondria among individual cells. According<br />

to this method, the mitochondrial matrix superoxide levels<br />

of single myoblasts are in the picomolar range. The method<br />

could also be extended to investigate other cellular systems<br />

such as cultured single skeletal muscle fibers, 38,39 whose<br />

superoxide levels were analyzed qualitatively in our earlier<br />

(35) Chen, H.; Cangello, D.; Benson, S.; Folmer, J.; Zhu, H.; Trush, M. A.; Zirkin,<br />

B. R. Exp. Gerontol. 2001, 36, 1361.<br />

(36) Klamt, F.; Gottfried, C.; Tramontina, F.; Dal-Pizzol, F.; Da Frota, M. L., Jr.;<br />

Moreira, J. C.; Dias, R. D.; Moriguchi, E.; Wofchuk, S.; Souza, D. O.<br />

Neuroreport 2002, 13, 1515.<br />

(37) Nishikawa, T.; Edelstein, D.; Du, X. L.; Yamagishi, S.-i.; Matsumura, T.;<br />

Kaneda, Y.; Yorek, M. A.; Beebe, D.; Oates, P. J.; Hammes, H.-P.; Giardino,<br />

I.; Brownlee, M. Nature 2000, 404, 787.<br />

(38) Pye, D.; Palomero, J.; Kabayo, T.; Jackson, M. J. J. Physiol. 2007, 581,<br />

309.<br />

(39) Palomero, J.; Pye, D.; Kabayo, T.; Spiller, D. G.; Jackson, M. J. Antioxid.<br />

Redox Signaling 2008, 10, 1463.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6749


study. 40 The method may also be modified to monitor superoxide<br />

released outside the mitochondria, which would be<br />

particularly useful to investigate cell lines derived from Cu, Zn-<br />

SOD-deficient animal models. 41,42 Long-term application of<br />

chemical cytometry to quantify superoxide levels may find wide<br />

applications to the fields of toxicology, aging, and oxidativestress<br />

related disease. 4,43,44<br />

ACKNOWLEDGMENT<br />

The National Institutes of Health supported this work through<br />

Grant R01-AG-20866. We thank Dr. Margaret Donoghue for<br />

providing comments on the manuscript.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

(40) Xu, X.; Thompson, L. V.; Navratil, M.; Arriaga, E. A. Anal. Chem. 2010,<br />

82, 4570–4576.<br />

(41) Reaume, A. G.; Elliott, J. L.; Hoffman, E. K.; Kowall, N. W.; Ferrante, R. J.;<br />

Siwek, D. F.; Wilcox, H. M.; Flood, D. G.; Beal, M. F.; Brown, R. H., Jr.; Received for review June 7, 2010. Accepted July 4, 2010.<br />

Scott, R. W.; Snider, W. D. Nat. Genet. 1996, 13, 43.<br />

AC101509D<br />

(42) Veerareddy, S.; Cooke, C. L.; Baker, P. N.; Davidge, S. T. Am. J. Physiol.<br />

Heart Circ. Physiol. 2004, 287, H40.<br />

(43) Thompson, L. V. Exp. Gerontol. 2009, 44, 106. (44) Amacher, D. E. Curr. Med. Chem. 2005, 12, 1829.<br />

6750 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010


Anal. Chem. 2010, 82, 6751–6755<br />

Resolving Disulfide Structural Isoforms of IgG2<br />

Monoclonal Antibodies by Ion Mobility Mass<br />

Spectrometry<br />

Dhanashri Bagal, † John F. Valliere-Douglass, ‡ Alain Balland,* ,‡ and Paul D. Schnier* ,†<br />

Molecular Structure, Amgen, Thousand Oaks, California 91320, and Process and Product Development, Amgen,<br />

Seattle, Washington 98119<br />

Recombinant monoclonal antibodies are an important<br />

class of therapeutic agents that have found widespread<br />

use for the treatment of many human diseases. Here, we<br />

have examined the utility of ion mobility mass spectrometry<br />

(IMMS) for the rapid characterization of disulfide<br />

variants in intact IgG2 monoclonal antibodies. It is shown<br />

that IMMS reveals 2 to 3 gas-phase conformer populations<br />

for IgG2s. In contrast, a single gas-phase conformer is<br />

revealed using IMMS for both an IgG1 antibody and a Cys-<br />

232 f Ser mutant IgG2, both of which are homogeneous<br />

with respect to disulfide bonding. This provides strong<br />

evidence that the observed IgG2 gas-phase conformers<br />

are related to disulfide bond heterogeneity. Additionally,<br />

IMMS analysis of redox enriched disulfide isoforms allows<br />

assignment of the mobility peaks to established disulfide<br />

bonding patterns. These data clearly illustrate how IMMS<br />

can be used to quickly provide information on the higher<br />

order structure of antibody therapeutics.<br />

The overall structure of the immunoglobulin G (IgG) family<br />

is organized in 12 subdomains, each closed by an intrachain<br />

disulfide bond. 1 Heavy chain (HC) and light chain (LC) are<br />

connected by interchain disulfide bonds to form a covalent<br />

complex of the form (HC-LC)2. IgG1, IgG2, and IgG4 isotypes<br />

share greater than 90% sequence homology in their constant<br />

domains but differ significantly in the hinge region. IgG1 and IgG4<br />

hinge core sequences are very similar with two cysteines on each<br />

heavy chain involved in interheavy chain connection, whereas<br />

IgG2 is unique in presenting four cysteine residues in the hinge<br />

region, notably two consecutive residues, Cys-232 and Cys-233<br />

(amino acid numbering of Kabat et al.), 2 that have no equivalent<br />

in any other immunoglobulin subclass. Researchers at Amgen<br />

recently reported that these residues confer distinctive structural<br />

features to the human IgG2 isotype resulting in the formation of<br />

disulfide-related structural isoforms. 3,4<br />

Three distinct structural isoforms (IgG2-A, IgG2-B, and<br />

IgG2-A/B) 3,4 specific to human IgG2s were revealed by chro-<br />

* To whom correspondence should be addressed. E-mail: ballanda@amgen.com<br />

(A.B.); pschnier@amgen.com (P.D.S.).<br />

† Molecular Structure.<br />

‡ Process and Product Development.<br />

(1) Padlan, E. A. Adv. Protein Chem. 1996, 49, 57–133.<br />

(2) Kabat, E. A.; Wu, T. T.; Perry, H. M.; Gottesman, K. S.; Foeller, C. Sequences<br />

of Proteins of Immunological Interest, 5th ed.; U.S. Public Health Service,<br />

NIH: Washington, DC., 1991.<br />

matographic and electrophoretic methods including capillary<br />

electrophoresis with sodium dodecyl sulfate (CE-SDS), 3,5 reversedphase<br />

high performance liquid chromatography (RP-HPLC), 4 and<br />

cation exchange chromatography (CEX). 3,6 IgG2-A corresponds<br />

to the classical model with independent Fab and Fc domains<br />

connected by the hinge (Figure 1a). 7 IgG2-B is a symmetrical form<br />

with HC and LC covalently linked to the hinge by disulfide bridges<br />

(Figure 1b). IgG2-A/B is an asymmetrical form intermediate<br />

between A and B. Detailed analysis of each IgG2 kappa structural<br />

isoforms showed that the different interchain disulfide bond<br />

arrangements involved only four residues: the cysteine in constant<br />

region one of the heavy chain (CH1), Cys-127 (Kabat numbering),<br />

the cysteine at the C-terminus of the light chain, Cys-214, and<br />

two cysteines in the upper hinge region, specific to the IgG2<br />

subclass, Cys-232 and Cys-233. The precise cysteine connectivity<br />

of each structural form was established by partial reductionalkylation<br />

and mass spectrometry (MS) 8 and Edman sequencing<br />

MS. 9<br />

Modeling of the IgG2 sequence based on the three-dimensional<br />

antibody structure places the four cysteines in close spatial<br />

proximity, supporting the concept that a variable arrangement of<br />

these residues could generate IgG2 structural isoforms. Two<br />

specific cysteine-to-serine mutants were designed at positions 232<br />

and 233 to disrupt potential disulfide rearrangements. 10 These<br />

mutants both exhibited no significant difference in expression and<br />

potency characteristics when compared to wild type IgG2 but<br />

proved structurally homogeneous with respect to the disulfide<br />

bonding of the IgG2-A type. 10<br />

(3) Wypych, J.; Li, M.; Guo, A.; Zhang, Z.; Martinez, T.; Allen, M. J.; Fodor, S.;<br />

Kelner, D. N.; Flynn, G. C.; Liu, Y. D.; Bondarenko, P. V.; Ricci, M. S.;<br />

Dillon, T. M.; Balland, A. J. Biol. Chem. 2008, 283, 16194–16205.<br />

(4) Dillon, T. M.; Ricci, M. S.; Vezina, C.; Flynn, G. C.; Liu, Y. D.; Rehder,<br />

D. S.; Plant, M.; Henkle, B.; Li, Y.; Deechongkit, S.; Varnum, B.; Wypych,<br />

J.; Balland, A.; Bondarenko, P. V. J. Biol. Chem. 2008, 283, 16206–16215.<br />

(5) Guo, A.; Han, M.; Martinez, T.; Ketchem, R. R.; Novick, S.; Jochheim, C.;<br />

Balland, A. Electrophoresis 2008, 29, 2550–2556.<br />

(6) Zhang, Y.; G. A.; Novick, S.; Jochheim, C.; Boyce, J. M.; Gerhart, M.; Qin,<br />

X.; Gombotz, W. I. Bioprocessing J. 2003, (Nov-Dec), 37–43.<br />

(7) Milstein, C.; Frangione, B. Biochem. J. 1971, 121, 217–225.<br />

(8) Martinez, T.; Guo, A.; Allen, M. J.; Han, M.; Pace, D.; Jones, J.; Gillespie,<br />

R.; Ketchem, R. R.; Zhang, Y.; Balland, A. Biochemistry 2008, 47, 7496–<br />

7508.<br />

(9) Zhang, B.; Harder, A. G.; Connelly, H. M.; Maheu, L. L.; Cockrill, S. L.<br />

Anal. Chem. 2009, 82, 1090–1099.<br />

(10) Allen, M. J.; Guo, A.; Martinez, T.; Han, M.; Flynn, G. C.; Wypych, J.; Liu,<br />

Y. D.; Shen, W. D.; Dillon, T. M.; Vezina, C.; Balland, A. Biochemistry 2009,<br />

48, 3755–3766.<br />

10.1021/ac1013139 © 2010 American <strong>Chemical</strong> Society 6751<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/21/2010


Figure 1. Illustration of the hinge region disulfide bonding pattern<br />

of human (a) IgG2-A and (b) IgG2-B antibodies.<br />

The ability to rapidly detect and characterize IgG2 isoforms is<br />

of great interest, as it may help to facilitate the transition of new<br />

IgG2 molecules from discovery into development and ultimately<br />

commercialization. In recent years, mass spectrometry has played<br />

an increasingly important role in the analytical characterization<br />

of IgG therapeutics. Mass spectrometry is now widely used to<br />

confirm the intact molecular weight of IgGs, 11,12 establish their<br />

glycosylation profile, 13,14 and confirm 15 or establish 16 the primary<br />

(11) Gadgil, H. S.; Pipes, G. D.; Dillon, T. M.; Treuheit, M. J.; Bondarenko, P. V.<br />

J. Am. Soc. Mass Spectrom. 2006, 17, 867–872.<br />

(12) Brady, L. J.; Valliere-Douglass, J.; Martinez, T.; Balland, A. J. Am. Soc. Mass<br />

Spectrom. 2008, 19, 502–509.<br />

(13) Damen, C. W. N.; Chen, W.; Chakraborty, A. B.; van Oosterhout, M.;<br />

Mazzeo, J. R.; Gebler, J. C.; Schellens, J. H. M.; Rosing, H.; Beijnen, J. H.<br />

J. Am. Soc. Mass Spectrom. 2009, 20, 2021–2033.<br />

(14) Olivova, P.; Chen, W.; Chakraborty, A. B.; Gebler, J. C. Rapid Commun.<br />

Mass Spectrom. 2008, 22, 29–40.<br />

(15) Ren, D.; Pipes, G. D.; Hambly, D.; Bondarenko, P. V.; Treuheit, M. J.; Gadgil,<br />

H. S. Anal. Biochem. 2009, 384, 42–48.<br />

(16) Bandeira, N.; Pham, V.; Pevzner, P.; Arnott, D.; Lill, J. R. Nat. Biotechnol.<br />

2008, 26, 1336–1338.<br />

6752 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

structure with a high degree of detail. Although MS is not<br />

routinely used to characterize higher order structural elements<br />

in IgGs, MS coupled with hydrogen/deuterium exchange was<br />

recently demonstrated as a method to characterize the conformational<br />

dynamics of IgG1 antibodies in solution. 17<br />

Ion mobility mass spectrometry (IMMS) has shown great<br />

promise as an intact protein separation and analysis methodology<br />

to probe higher order structural elements including the overall<br />

size/shapeofbiopolymersandlargemacromolecularassemblies. 18-28<br />

Recently, Waters Corporation commercialized an ion mobility<br />

mass spectrometer (Synapt) based on traveling waves (T-Wave). 29<br />

In the T-Wave implementation of ion mobility, ion separation<br />

occurs when a sequence of dc pulses push ions through the<br />

mobility cell in the presence of an inert gas at relatively high<br />

pressure. 29,30 The ability of an ion to “surf” the T-wave depends<br />

on its collision cross section (CCS). Ions with compact structures<br />

are pushed through the mobility cell faster than ions with more<br />

elongated structures. In this work, we present evidence that<br />

T-Wave IMMS can be used to separate disulfide variants of intact<br />

IgG2 antibodies. Attractive features of the method include high<br />

sensitivity (µg sample consumption), minimal sample preparation,<br />

and fast analysis time (minutes).<br />

EXPERIMENTAL SECTION<br />

Human monoclonal antibodies mAb#1 (IgG2), mAb#2 (IgG1),<br />

and mAb#3 (IgG2) were produced recombinantly in Chinese<br />

hamster ovary (CHO) cells and purified at Amgen. All additional<br />

reagents were purchased from Sigma-Aldrich (St. Louis, MO)<br />

unless otherwise specified. For control experiments, constant<br />

region two (CH2) domain N-glycans were removed by adding<br />

1500 U of PNGase F (New England Biolabs, Ipswich, MA) per<br />

100 µg of protein and incubating at 37 °C for 16 h.<br />

Disulfide isoforms IgG2-A and IgG2-B were selectively enriched<br />

using methods established by Dillon et al. 4 Briefly, to<br />

enrich isoform B, IgG2s were incubated in 200 mM Tris buffer<br />

(17) Houde, D.; Arndt, J.; Domeier, W.; Berkowitz, S.; Engen, J. R. Anal. Chem.<br />

2009, 81, 2644–2651.<br />

(18) Clemmer, D. E.; Hudgins, R. R.; Jarrold, M. F. J. Am. Chem. Soc. 1995,<br />

117, 10141–10142.<br />

(19) Bohrer, B. C.; Merenbloom, S. I.; Koeniger, S. L.; Hilderbrand, A. E.;<br />

Clemmer, D. E. Annu. Rev. Anal. Chem. 2008, 1, 293–327.<br />

(20) Ruotolo, B. T.; Giles, K.; Campuzano, I.; Sandercock, A. M.; Bateman, R. H.;<br />

Robinson, C. V. Science 2005, 310, 1658–1661.<br />

(21) Kaddis, C. S.; Loo, J. A. Anal. Chem. 2007, 79, 1778–1784.<br />

(22) Leary, J. A.; Schenauer, M. R.; Stefanescu, R.; Andaya, A.; Ruotolo, B. T.;<br />

Robinson, C. V.; Thalassinos, K.; Scrivens, J. H.; Sokabe, M.; Hershey,<br />

J. W. B. J. Am. Soc. Mass Spectrom. 2009, 20, 1699–1706.<br />

(23) Kim, H. I.; Kim, H.; Pang, E. S.; Ryu, E. K.; Beegle, L. W.; Loo, J. A.;<br />

Goddard, W. A.; Kanik, I. Anal. Chem. 2009, 81, 8289–8297.<br />

(24) Atmanene, C. D.; Wagner-Rousset, E.; Malissard, M.; Chol, B.; Robert, A.;<br />

Corvaïa, N.; Dorsselaer, A. V.; Beck, A.; Sanglier-Cianferani, S. Anal. Chem.<br />

2009, 81, 6364–6373.<br />

(25) Ruotolo, B. T.; Hyung, S.-J.; Robinson, P. M.; Giles, K.; Bateman, R. H.;<br />

Robinson, C. V. Angew. Chem., Int. Ed. 2007, 46, 8001–8004.<br />

(26) Hilton, G. R.; Thalassinos, K.; Grabenauer, M.; Sanghera, N.; Slade, S. E.;<br />

Wyttenbach, T.; Robinson, P. J.; Pinheiro, T. J. T.; Bowers, M. T.; Scrivens,<br />

J. H. J. Am. Soc. Mass Spectrom. 2010, 21, 845–854.<br />

(27) Schenauer, M. R.; Meissen, J. K.; Seo, Y.; Ames, J. B.; Leary, J. A. Anal.<br />

Chem. 2009, 81, 10179–10185.<br />

(28) Thalassinos, K.; Grabenauer, M.; Slade, S. E.; Hilton, G. R.; Bowers, M. T.;<br />

Scrivens, J. H. Anal. Chem. 2008, 81, 248–254.<br />

(29) Pringle, S. D.; Giles, K.; Wildgoose, J. L.; Williams, J. P.; Slade, S. E.;<br />

Thalassinos, K.; Bateman, R. H.; Bowers, M. T.; Scrivens, J. H. Int. J. Mass<br />

Spectrom. 2007, 261, 1–12.<br />

(30) Shvartsburg, A. A.; Smith, R. D. Anal. Chem. 2008, 80, 9689–9699.


(pH 8.0) with cysteine and cystamine at concentrations of 6 and<br />

1 mM, respectively. Isoform-A was enriched by incubating the<br />

IgG2 under the same conditions but with the addition of 1.0 M<br />

guanidinium chloride (GuHCl) to the buffer. The samples were<br />

protected from light at 2-8 °C for approximately 48 h.<br />

The IgG2 mAb#3 was subjected to Cysf Ser mutagenesis at<br />

position 232 as described by Allen et al. 10 Site-directed mutagenesis<br />

was performed using the QuickChange XL site-directed mutagenesis<br />

kit (Stratagene, La Jolla, CA). The mutant was stably transfected into<br />

a serum-free suspension-adapted CHO cell line. 31 Following production,<br />

the mAb was purified by protein A affinity chromatography.<br />

Incorporation of the expected mutation was verified on the purified<br />

molecule by size exclusion chromatography (SEC) MS. 12<br />

For IMMS analysis, IgG samples were buffer exchanged and<br />

concentrated using Vivaspin 30 kDa molecular weight cutoff filters<br />

(GE Healthcare, Buckinghamshire, UK). For experiments performed<br />

under native conditions, the IgGs were diluted using 160 mM<br />

ammonium acetate (pH not adjusted) to a final working concentration<br />

of ∼3 µM.<br />

Mass spectrometry experiments were performed using a<br />

hybrid ion mobility quadrupole time-of-flight MS (Synapt, Waters<br />

Inc., Milford, MA) equipped with a nanoelectrospray ionization<br />

(ESI) source using metal coated borosilicate glass capillaries<br />

(nanoflow probe tips, long thin walled, Waters Corporation).<br />

Solution flow rates of ∼75 nL/min and an ESI capillary voltage of<br />

∼1.3 kV were used for all experiments. The source temperature<br />

was 50 °C, and the pressure of the vacuum/backing region was<br />

3.5 mbar. Each ion mobility mass spectrum was acquired from<br />

m/z 4000-8000 every 2 s; approximately 12 counts per scan were<br />

observed. The signal was typically averaged for approximately 10<br />

min. Gentle source conditions were used to minimize gas phase<br />

unfolding of the protein (sample cone: 40 V, trap voltage: 10 V,<br />

transfer lens: 12 V, bias: 25; cone gas: 30 L/Hr; P TRAP(Ar): 0.0175<br />

mbar). Nitrogen was used as the mobility carrier gas, and the<br />

following parameters were found to give optimal ion mobility<br />

separation (PIMS: 0.5 mbar, wave velocity: 300 m/s, wave height:<br />

9.8 V). The instrument was mass calibrated using a 50 µg/µL<br />

CsI solution. Waters’ raw data files were translated to Matlab<br />

binary files using software developed in-house. Data processing<br />

and plotting were performed in Matlab (The MathWorks Inc.<br />

Natick, MA) and Igor Pro (WaveMetrics Inc., Lake Oswego, OR).<br />

RESULTS AND DISCUSSION<br />

Antibodies belonging to the IgG2 subclass exist as a group of<br />

distinct isoforms with different disulfide connectivities between<br />

the Fab domain and the hinge region of the molecule. The goal<br />

of this study was to investigate the ability of IMMS to successfully<br />

separate these discrete IgG2 isoforms.<br />

A nano-ESI ion mobility mass spectrum for a solution of 3 µM<br />

mAb#1 (theoretical MW for the most abundant glycoform (G1F/<br />

G1F): 149821.50 Da) in 160 mM ammonium acetate is shown<br />

Figure 2a. A narrow charge state distribution centered on the 24+<br />

ion is observed. Ion mobility spectra for all MAbs analyzed were<br />

acquired with minimal acceleration voltages (sample cone: 40 V,<br />

trap: 10 V, transfer: 12 V). These gentle tuning conditions were<br />

found to provide optimal resolution in the ion mobility dimension<br />

(31) Rasmussen, B.; Davis, R.; Thomas, J.; Reddy, P. Cytotechnology 1998, 28,<br />

31–42.<br />

Figure 2. Ion mobility TOF mass spectra of (a) mAb#1 and (c) mAb#2.<br />

The normalized ion mobility intensities are graphed as contour plots with<br />

9 levels from 4% to 100% intensity. Extracted arrival time distributions<br />

for the 26+ charge states of (b) mAb#1 and (d) mAb#2.<br />

(vide infra). However, with these conditions, the mass spectral<br />

peaks are relatively broad making it impossible to observe the<br />

individual glycoforms of the IgG. The glycoforms for these IgG<br />

molecules can, however, be readily resolved using higher source<br />

voltages (data not shown). Robinson and co-workers have previously<br />

observed that optimal conditions for the mass and mobility<br />

dimensions are often not compatible on the Synapt instrument<br />

when analyzing large protein assemblies. 32<br />

Ion mobility separation of mAb#1 (Figure 2a) reveals two distinct<br />

conformer populations. For example, the mobility spectrum for the<br />

26+ charge state (Figure 2b) shows two distinct peaks with a 1.1<br />

ms difference in drift times (9.3 and 10.4 ms, respectively). Similar<br />

distributions are observed for each charge state. However, with the<br />

tuning conditions employed here, the best mobility resolution was<br />

achieved for the 27+, 26+, and 25+ charge states. To elucidate if<br />

the heterogeneity of the sugar chains on the antibody influences the<br />

observed gas-phase conformers, deglycosylated mAb#1 was analyzed.<br />

The same distribution of gas-phase conformers is revealed in the<br />

ion mobility mass spectrum of mAb#1 with the sugars released<br />

(Figure S1a, Supporting Information). This clearly demonstrates that<br />

these multiple conformers are not caused by sugar heterogeneity<br />

but instead are related to conformational differences in the protein<br />

chains. The arrival time distributions for each charge state virtually<br />

overlaps with that of the glycosylated protein. For example, the major<br />

peaks in the arrival time distribution plot of the 26+ charge state of<br />

deglycosylated mAb#1 are only shifted by -0.13 ms compared to<br />

(32) Ruotolo, B. T.; Benesch, J. L. P.; Sandercock, A. M.; Hyung, S.-J.; Robinson,<br />

C. V. Nat. Protoc. 2008, 3, 1139–1152.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6753


that of the glycosylated IgG (Figure S1b, Supporting Information).<br />

This indicates that the glycosylated and deglycosylated antibodies<br />

have nearly identical gas-phase collision cross sections.<br />

The relative abundance of each gas-phase conformer can be<br />

roughly estimated by fitting the arrival time distribution to a sum<br />

of log-normal functions and integrating the relative area of each<br />

peak. The log-normal function is an asymmetric Gaussian function<br />

whose logarithm is normally distributed. 33 The log-normal function<br />

was used because it was empirically found to best fit the<br />

experimental peak shape of this ion mobility data. For the 26+<br />

charge state, the normalized area of peaks 1 and 2 from the best<br />

fits are 42% and 58%, respectively. Similar areas are observed for<br />

other charge states. These abundances are similar to the relative<br />

peak areas observed in electropherograms of IgG2s separated with<br />

capillary electrophoresis. We and others have shown CE-SDS to<br />

be a resolving technique for the separation of IgG2 structural<br />

isoforms. 3,5,35 However, this correlation alone does not signify that<br />

the resolved gas-phase conformers are definitively due to disulfide<br />

variants.<br />

To elucidate whether the gas-phase conformers observed for<br />

IgG2 molecules are related to their disulfide connectivity, we<br />

analyzed an IgG1 antibody, mAb#2 (theoretical MW for the most<br />

abundant glycoform (G1F/G0F):148408.0 Da), as a control (Figure<br />

2c). The most significant difference between human IgG1 and<br />

IgG2 subclasses is the primary structure of the hinge region,<br />

resulting in the absence of disulfide related isoforms in the IgG1. 3,4<br />

In contrast to the IgG2 mobility data, for each charge state of<br />

mAb#2, the arrival time profile is relatively narrow and consists<br />

of a single uniform distribution (Figure 2c). For example, the<br />

arrival time profile for the 26+ charge state of mAb#2 (Figure<br />

2d) shows a single peak at 8.7 ms. This arrival time profile is<br />

representative of the distribution observed for each of the charge<br />

states. This suggests that the multiple conformers observed of<br />

mAb#1 are due to disulfide variants in the antibody.<br />

To further demonstrate that the observed gas-phase conformers<br />

are indeed IgG2 disulfide variants, individual disulfide isoforms<br />

were selectively enriched in a refolding experiment using redox<br />

chemistry employing cysteine/cystamine. Dillon et al. previously<br />

demonstrated that isoform IgG2-A and IgG2-B can be redoxenriched<br />

by refolding with and without 1 M GuHCl, respectively. 4<br />

Under redox conditions in buffer alone, IgG2-A is refolded to IgG2-<br />

B. Liu et al. demonstrated that a slow conversion of IgG2-A to<br />

IgG2-B also occurs in vivo. 34 Isoform conversion toward IgG2-A<br />

requires in vitro refolding in presence of low levels of chaotropic<br />

reagents. 4 Figure 3a,b shows the arrival time distributions for the<br />

+26 charge state for redox enriched mAb#1 in the presence of<br />

guanidine (isoform A) and redox enriched mAb#1 in the absence<br />

of guanidine (isoform B), respectively. In contrast to IMMS for<br />

the untreated Mab#1 antibody (Figure 3, dashed line), a single<br />

abundant conformer is observed for each of these enriched<br />

isoforms. As a control, the IgG1 antibody, mAb#2, was also<br />

subjected to the same redox refolding protocol with or without<br />

guanidine. All treated IgG1 samples have identical arrival time<br />

(33) Brown, R. Personal Eng. Instrum. News 1991, 8, 51–54.<br />

(34) Liu, Y. D.; Chen, X.; Enk, J. Z.; Plant, M.; Dillon, T. M.; Flynn, G. C. J. Biol.<br />

Chem. 2008, 283, 29266–29272.<br />

(35) Lacher, N. A.; Wang, Q.; Roberts, R. K.; Holovics, H. J.; Aykent, S.; Schlittler,<br />

M. R.; Thompson, M. R.; Demarest, C. W. Electrophoresis 2010, 31, 448–<br />

458.<br />

6754 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 3. Arrival time distributions for the 26+ charge states of IgGs<br />

treated with redox reagents (cystamine, cysteine). (a,b) mAb#1 and<br />

(c,d) mAb#2 (control). Samples represented in (a) and (c) had 1 M<br />

GuHCl added to the buffer. The dashed line shows the arrival time<br />

distribution for the 26 + charge state of untreated mAb#1.<br />

distribution profiles as the untreated IgG1 molecule (Figure 3c,d),<br />

demonstrating that this refolding protocol does not affect the<br />

overall tertiary structure of the antibody. Comparing the untreated<br />

IgG2 IMMS trace with the enriched isoform distributions (Figure<br />

3a,b) identifies peaks 1 (9.3 ms) and 2 (10.4 ms) in the mobility<br />

spectra as isoform A and isoform B, respectively. The relatively<br />

late arrival time of isoform B indicates that this form of the<br />

antibody has a larger gas-phase collision cross section compared<br />

to isoform A. The IMMS resolution of the isoforms correlates with<br />

the previously observed capillary electrophoresis separation. 5,8,35<br />

With CE, IgG2 isoforms were resolved into two peaks, with IgG2-A<br />

migrating more rapidly than the IgG2-B isoform. The intermediate<br />

IgG2-A/B forms were split between the two peaks, IgG2-A/B1<br />

migrating with -A and IgG2-A/B2 with -B. 8<br />

While the disulfide connectivities of IgG2 isoforms have been<br />

well characterized, 3,4,8,9 only limited information is available<br />

describing the overall tertiary structure of human IgGs. To<br />

investigate if the relative ordering of gas-phase collision cross<br />

sections (IgG1 ≈ IgG2-A < IgG2-B) correlates with IgG solution<br />

structures, collision cross sections were calculated for two IgG<br />

structures using Waters’ CCS software. 36 The calculated cross<br />

section of an IgG antibody, with a disulfide bonding pattern<br />

consistent with the B isoform, is 8385 Å 2 (unpublished results).<br />

This value is 3% smaller than the calculated cross section of<br />

an IgG1 antibody (protein data bank code 1HZH, 8653 Å 2 ). 37<br />

(36) Williams, J. P.; Lough, J. A.; Campuzano, I.; Richardson, K.; Sadle, P. J.<br />

Rapid Commun. Mass Spectrom. 2009, 23, 3563.<br />

(37) Saphire, E. O.; Parren, P. W.; Pantophlet, R.; Zwick, M. B.; Morris, G. M.;<br />

Rudd, P. M.; Dwek, R. A.; Stanfield, R. L.; Burton, D. R.; Wilson, I. A. Science<br />

2001, 293, 1155–1159.


Figure 4. Ion mobility mass spectra of (a) mAb#3 and (b) Cys f Ser Mab#3 mutant. The extracted arrival time distributions for the 25+ charge<br />

states of these molecules are shown in (c) and (d), respectively.<br />

These limited calculations using individual static structures do<br />

not explain the relatively late arrival time of the IgG2-B isoform<br />

and suggest that the overall three-dimensional structure of<br />

these gas-phase ions may be quite different than the solution<br />

structure. While the IMMS results demonstrate that ion<br />

mobility separates covalent disulfide mediated IgG2 structural<br />

isoforms, the overall three-dimensional structure of these gasphase<br />

ions is presently not clear.<br />

To investigate the generality of these ion mobility results,<br />

measurements were also performed on a different IgG2 antibody<br />

(mAb#3) which also exists as an ensemble of disulfide mediated<br />

isoforms (Figure 4a). Ion mobility separation of mAb#3 reveals<br />

two abundant gas-phase conformer populations and one minor<br />

conformer, for each charge state. These three populations are<br />

readily apparent in the arrival time distribution for the 25+ charge<br />

state (Figure 4c) which shows three distinct peaks with drift times<br />

of 9.5, 10.3, and 11.9 ms. The relative abundance is determined<br />

by fitting the arrival time distribution to a sum of log-normal<br />

functions of each conformer and is roughly estimated to be 46%,<br />

49%, and 5%. Figure 4b, shows the ion mobility mass spectrum<br />

for a mutant form of this antibody with a single point mutation<br />

introduced at amino acid 232 (Cys f Ser) by site-directed<br />

mutagenesis. A single narrow peak is observed for the arrival time<br />

distribution of each charge state. The arrival time profile of the<br />

25+ charge state is shown in Figure 4d as an illustrative example<br />

demonstrating a homogeneous population, with a single peak at<br />

9.3 ms. This peak corresponds to the gas-phase conformer with<br />

the more compact structure (isoform A, vide supra). This is<br />

consistent with previous studies demonstrating that this mutant<br />

is homogeneous with respect to disulfide bonding and of the<br />

IgG2-A type. 10,38 This result also unambiguously confirms our<br />

assertion that the multiple IMMS peaks observed in this study of<br />

selected IgG2s can be correlated with the disulfide bonding<br />

patterns in these molecules.<br />

(38) Lightle, S.; Aykent, S.; Lacher, N.; Mitaksov, V.; Wells, K.; Zobel, J.; Oliphant,<br />

T. Protein Sci. 2010, 19, 753–762.<br />

CONCLUSIONS<br />

The results of the present study demonstrate that ion mobility<br />

as a shape-selective separation methodology can be used to detect<br />

disulfide heterogeneity in large (150 kDa) intact IgG2 antibodies.<br />

Two to three gas-phase conformers are observed by ion mobility<br />

for IgG2 antibodies. These gas-phase conformers were maintained<br />

with deglycosylated IgG2s. Analysis of redox refolded IgG2s as<br />

well as an IgG2 with a Cys f Ser single point mutation clearly<br />

demonstrates that the observed gas-phase conformers are related<br />

to disulfide variants. Ion mobility is fast (millisecond measurements),<br />

sensitive (nanomole), and amenable to high throughput<br />

automation. IMMS is a powerful new methodology for the<br />

characterization of intact antibodies and may be useful to routinely<br />

fingerprint higher order structure of these protein biopharmaceuticals<br />

in the near future. We are currently extending these<br />

measurements to investigate the utility of IMMS for the analysis<br />

of IgG2s containing lambda light chains as well as to directly<br />

characterize the binding of antigen targets to individual disulfide<br />

isoforms of IgG2 antibodies.<br />

ACKNOWLEDGMENT<br />

We are grateful to Allen Sickmier and Leszek Poppe for<br />

insightful discussions, Keith Richardson (Waters Inc.) for providing<br />

the Waters’ CCS software, Mike Berke, Rick Stanton, and<br />

Mikhail Toupikov for help with the data analysis software, and<br />

Mike Treuheit, Dean Pettit, Peter Grandsard, and Philip Tagari<br />

for championing and supporting this work. We also thank our<br />

Amgen colleagues, whose names are listed in the references, for<br />

their expert contributions to the collective knowledge built<br />

recently on IgG2 isoforms.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 19, 2010. Accepted July 6, 2010.<br />

AC1013139<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6755


Anal. Chem. 2010, 82, 6756–6763<br />

An Enzymatic Microreactor Based on Chaotic<br />

Micromixing for Enhanced Amperometric<br />

Detection in a Continuous Glucose Monitoring<br />

Application<br />

Byeong-Ui Moon, †,‡ Sander Koster, § Klaas J. C. Wientjes, † Radosław M. Kwapiszewski, |<br />

Adelbert J. M. Schoonen, † Ben H. C. Westerink, † and Elisabeth Verpoorte* ,‡<br />

Biomonitoring and Sensoring, Pharmaceutical Analysis, Department of Pharmacy, University of Groningen, Antonius<br />

Deusinglaan 1, P.O. Box 196, 9700 AD Groningen, The Netherlands, TNO Quality of Life, Utrechtseweg 48,<br />

3700 AJ Zeist, The Netherlands, and Department of Microbioanalytics, Faculty of <strong>Chemistry</strong>, Warsaw University of<br />

Technology, Noakowskiego 3, Warsaw, 00-664, Poland<br />

The development of continuous glucose monitoring systems<br />

is a major trend in diabetes-related research. Small,<br />

easy-to-wear systems which are robust enough to function<br />

over many days without maintenance are the goal. We<br />

present a new sensing system for continuous glucose<br />

monitoring based on a microreactor incorporating chaotic<br />

mixing channels. Two different types of chaotic mixing<br />

channels with arrays of either slanted or herringbone<br />

grooves were fabricated in poly(dimethylsiloxane) (PDMS)<br />

and compared to channels containing no grooves. Mixing<br />

in channels with slanted grooves was characterized using<br />

a fluorescence method as a function of distance and at<br />

different flow rates, and compared to the mixing behavior<br />

observed in channels with no grooves. For electrochemical<br />

detection, a thin-film Pt electrode was positioned at the<br />

end of the fluidic channel as an on-chip detector of the<br />

reaction product, H 2O2. Glucose determination was<br />

performed by rapidly mixing glucose and glucose<br />

oxidase (GOx) in solution at a flow rate of 0.5 µL/min<br />

and 1.5 µL/min, respectively. A 150 U/mL GOx<br />

solution was selected as the optimum concentration<br />

of enzyme. In order to investigate the dependence of<br />

device response on flow rate, experiments with a<br />

premixed solution of glucose and GOx were compared<br />

to experiments in which glucose and GOx were reacted<br />

on-chip. Calibration curves for glucose (0-20 mM, in<br />

the clinical range of interest) were obtained in channels<br />

with and without grooves, using amperometric detection<br />

and a 150 U/mL GOx solution for in-chip reaction.<br />

1. INTRODUCTION<br />

Diabetes mellitus is a widespread disease causing heart<br />

disease, weight loss, blurry vision, neurological disorders, and<br />

* Corresponding author. E-mail: e.m.j.verpoorte@rug.nl, Telephone: +31-50-<br />

363-3337; fax: +31-50-363-7582.<br />

† Biomonitoring and Sensoring, University of Groningen.<br />

‡ Department of Pharmacy, University of Groningen.<br />

§ TNO Quality of Life.<br />

| Warsaw University of Technology.<br />

6756 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

even death. 1 Proper management of blood glucose is thus of<br />

crucial importance for diabetic patients. The conventional way<br />

blood glucose determinations are carried out involves the<br />

finger-prick method. Usually, diabetic patients measure their<br />

own blood glucose several times per day by applying a drop of<br />

blood to a portable device. However, this intermittent monitoring<br />

does not yield the full range of information necessary to<br />

effectively control glucose levels 24 h a day. The current<br />

research trend is thus toward real-time in vivo monitoring over<br />

longer periods (days). 2-5 Type 1 diabetic patients in particular<br />

could benefit from a portable glucose sensor to keep their<br />

glucose values within a reasonable range. Theoretically, a<br />

portable glucose sensor could include an insulin pump in a<br />

feedback loop to serve as an artificial pancreas.<br />

Electrochemical detection is commonly used for glucose<br />

analysis. 6 The advantages of electrochemical detection include<br />

sensitivity and ease of interfacing with detection electronics.<br />

Several reports incorporating electrochemical detection on a<br />

microchip have been published for the detection of glucose. 7-9<br />

The principle of the enzymatic reaction involved is shown in<br />

eqs 1 and 2. Glucose, O2 and glucose oxidase (GOx) react to<br />

generate hydrogen peroxide (H2O2) at concentrations proportional<br />

to the original glucose concentrations. The H2O2<br />

is then oxidized under applied potential into O2 and H + , with<br />

the resulting electrons being recorded as an electrical<br />

current.<br />

(1) Patient Education Association. http://www.nlm.nih.gov/medlineplus/<br />

tutorials/diabetesintroduction/htm/index.htm. (Accessed July 12, 2010).<br />

(2) Klonoff, D. C. Diabetes Care 2005, 28, 1231–1239.<br />

(3) Feldman, B.; Brazg, R.; Schwartz, S.; Weinstein, R. Diabetes Technol. Ther.<br />

2003, 5, 769–779.<br />

(4) Lutgers, H. L.; Hullegie, L. M.; Hoogenberg, K.; Sluiter, W. J.; Dullaart,<br />

R. P. F.; Wientjes, K. J.; Schoonen, A. J. M. Neth. J. Med. 2000, 57, 7–12.<br />

(5) Tamada, J. A.; Garg, S.; Jovanovic, L.; Pitzer, K. R.; Fermi, S.; Potts, R. O.<br />

JAMA, J. Am. Med. Assoc. 1999, 282, 1839–1844.<br />

(6) Updike, S. J.; Hicks, G. P. Nature 1967, 214, 986–988.<br />

(7) Wang, J.; Chatrathi, M. P.; Tian, B.; Polsky, R. Anal. Chem. 2000, 72, 2514–<br />

2518.<br />

(8) Wilke, R.; Büttgenbach, S. Biosens. Bioelectron. 2003, 19, 149–153.<br />

(9) Yamaguchi, A.; Jin, P.; Tsuchiyama, H.; Masuda, T.; Sun, K.; Matsuo, S.;<br />

Misawa, H. Anal. Chim. Acta 2002, 468, 143–152.<br />

10.1021/ac1000509 © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/21/2010


GOx<br />

glucose + O298 gluconic acid +<br />

0.7 V<br />

H2O298 H 2 O 2 (GOx: glucose oxidase) (1)<br />

2H + + O 2 + 2e -<br />

The reaction may be carried out by mixing the reagents in solution<br />

to produce H2O2 which is then subsequently detected by<br />

integrated electrodes, 7,8 or by immobilizing GOx directly onto<br />

the detection electrode surface to react with glucose in solution. 9<br />

The electrochemical detectors in these microchips displayed fast<br />

response times and low detection limits. Immobilizing GOx<br />

certainly reduces its consumption significantly compared to<br />

devices in which GOx is added in solution. However, the major<br />

disadvantage of this approach is that the enzyme layer is prone<br />

to biofouling in in vivo applications, which might cause a decrease<br />

in activity over time and poor performance as a result. 10,11<br />

Over the past decade, a number of glucose sensor designs<br />

have been developed and become available for continuous<br />

monitoring purposes. 12-16 Mastrototaro et al. 14 reported an in<br />

vivo needle-type continuous glucose monitoring system (CGMS),<br />

which is currently being used by consumers as a kit (MiniMed).<br />

The detector in this system is based on the reaction of immobilized<br />

glucose oxidase on the electrode with glucose. Schoemaker et<br />

al. 15 and Wientjes et al. 17 have both reported a subcutaneous<br />

continuous glucose monitoring system (SCGM) based on a<br />

microdialysis probe and using the solution-based enzymatic<br />

reaction with glucose. They reported reliable measurements in<br />

diabetes patients over periods of four days 18 up to two weeks. 19,20<br />

However, due to their bulky size and the need for connecting<br />

tubes between system components, these systems exhibited<br />

relatively long physical lag times of 30 min or more. Pickup et<br />

al. 13 reviewed the clinical use of glucose monitoring systems and<br />

introduced new technology for noninvasive glucose sensing, based<br />

on near-infrared spectroscopy and fluorescence. Wentholt et al. 21<br />

(10) Gerritsen, M.; Jansen, J. A.; Lutterman, J. A. Neth. J. Med. 1999, 54, 167–<br />

179.<br />

(11) Wisniewski, N.; Moussy, F.; Reichert, W. M. Fresenius’ J. Anal. Chem. 2000,<br />

366, 611–621.<br />

(12) Maran, A.; Crepaldi, C.; Tiengo, A.; Grassi, G.; Vitali, E.; Pagano, G.; Bistoni,<br />

S.; Calabrese, G.; Santeusanio, F.; Leonetti, F.; Ribaudo, M.; Di Mario, U.;<br />

Annuzzi, G.; Genovese, S.; Riccardi, G.; Previti, M.; Cucinotta, D.; Giorgino,<br />

F.; Bellomo, A.; Giorgino, R.; Poscia, A.; Varalli, M. Diabetes Care 2002,<br />

25, 347–352.<br />

(13) Pickup, J. C.; Hussain, F.; Evans, N. D.; Sachedina, N. Biosens. Bioelectron.<br />

2005, 20, 1897–1902.<br />

(14) Mastrototaro, J. J. Diabetes Technol. Ther. 2000, 2, 13–18.<br />

(15) Schoemaker, M.; Andreis, E.; Roper, J.; Kotulla, R.; Lodwig, V.; Obermaier,<br />

K.; Stephan, P.; Reuschling, W.; Rutschmann, M.; Schwaninger, R.;<br />

Wittmann, U.; Rinne, H.; Kontschieder, H.; Strohmeier, W. Diabetes Technol.<br />

Ther. 2003, 5, 599–608.<br />

(16) Suzuki, H.; Tokuda, T.; Miyagishi, T.; Yoshida, H.; Honda, N. Sens. Actuators,<br />

B 2004, 97, 90–97.<br />

(17) Wientjes, K. J. C.; Grob, U.; Hattemer, A.; Hoogenberg, K.; Jungheim, K.;<br />

Kapitza, C.; Schoonen, A. J. M. Diabetes Technol. Ther. 2003, 5, 615–620.<br />

(18) Kapitza, C.; Lodwig, V.; Obermaier, K.; Wientjes, K. J. C.; Hoogenberg, K.;<br />

Jungheim, K.; Heinemann, L. Diabetes Technol. Ther. 2003, 5, 609–614.<br />

(19) Wientjes, K. J.; Vonk, P.; Vonk-van Klei, Y.; Schoonen, A. J.; Kossen, N. W.<br />

Diabetes Care 1998, 21, 1481–1488.<br />

(20) Schoonen, A. J. M.; Schmidt, F. J.; Hasper, H.; Verbrugge, D. A.; Tiessen,<br />

R. G.; Lerk, C. F. Biosens. Bioelectron. 1990, 5, 37–46.<br />

(21) Wentholt, I. M. E.; Hoekstra, J. B. L.; DeVries, J. H. Diabetes Technol. Ther.<br />

2007, 9, 399–409.<br />

(2)<br />

reported an overview of the current applications and clinically<br />

relevant aspects of continuous glucose monitors (CGMs), with<br />

emphasis on the calibration procedure, interpretation of continuous<br />

glucose data, and some important limitations. Overall, these<br />

authors concluded that improved accuracy, reliability for longer<br />

periods of time, miniaturization, and cost-effectiveness are the<br />

main issues which need to be considered in the further development<br />

of continuous glucose monitoring systems. 21<br />

Chip-based microfluidic technologies are a good alternative to<br />

improve on conventional monitoring approaches such as the<br />

SCGM described above. Microfluidic systems, also termed<br />

“miniaturized total analysis systems (µTAS)” 22 or “lab-on-a-chip”,<br />

are now widely used in analytical chemistry and biological<br />

applications. Advantages of these systems include dramatically<br />

reduced consumption of chemical reagents, faster reaction times<br />

and cost-effectiveness. Micro SCGM systems can thus be envisaged<br />

which exploit the solution-based reaction of GOx with<br />

glucose in nL volumes in micrometer-sized channels for glucose<br />

sensing. The speed of this analysis is determined by the efficiency<br />

of mixing reagents in the microfluidic channels. Mixing at the<br />

micrometer scale is a challenge, as flows are generally extremely<br />

well-defined and laminar. Mixing of two solution streams in a<br />

straight microchannel is possible only through means of diffusion,<br />

a passive molecular transport process which is very slow. There<br />

has therefore been an enormous amount of research done in the<br />

past decade or so on how to implement efficient mixing at the nL<br />

scale. So-called passive micromixers are generally preferred for<br />

many micro analytical flow systems, since these elements do not<br />

require the application of an external force to achieve mixing. A<br />

large number of passive micromixers have been reported, including<br />

a planar laminar flow mixer, 23 a cross-shaped micromixer 24<br />

and a droplet mixer. 25 The approach chosen for our work is one<br />

based on chaotic mixing, first described by Stroock et al. 26,27<br />

Mixing is achieved through the incorporation of an array of<br />

microgrooves into a microchannel. Flow over the groove array<br />

assumes a helical or corkscrew pattern, in which the contact area<br />

between two adjacent solutions is increased dramatically to<br />

facilitate mixing by diffusion.<br />

The long-term goal of the present project is to realize an<br />

autonomous, portable sensing system for continuous in vivo<br />

glucose monitoring, based on the reaction in solution of GOx with<br />

glucose to produce H 2O2. To accomplish this, we have designed<br />

a miniaturized glucose sensing system based on microdialysis<br />

sampling and lab-on-a-chip technology. In this system, nL<br />

amounts of sample and enzyme rapidly mix and react. As highrecovery<br />

microdialysis requires flow rates in tissue less than 1<br />

µL/min, 17 we have adapted the microreactor dimensions<br />

according to these conditions. In this paper, we describe a new<br />

application for chaotic mixing, that is, the efficient and fast<br />

mixing of GOx and glucose for reaction. To that end, either<br />

(22) Reyes, D. R.; Iossifidis, D.; Auroux, P.-A.; Manz, A. Anal. Chem. 2002, 74,<br />

2623–2636.<br />

(23) Melin, J.; Gimenez, G.; Roxhed, N.; van der Wijngaart, W.; Stemme, G.<br />

Lab Chip 2004, 4, 214–219.<br />

(24) Wong, S. H.; Bryant, P.; Ward, M.; Wharton, C. Sens. Actuators, B 2003,<br />

95, 414–424.<br />

(25) Paik, P.; Pamula, V. K.; Fair, R. B. Lab Chip 2003, 3, 253–259.<br />

(26) Stroock, A. D.; Dertinger, S. K. W.; Ajdari, A.; Mezic, I.; Stone, H. A.;<br />

Whitesides, G. M. Science 2002, 295, 647–651.<br />

(27) Stroock, A. D.; Dertinger, S. K.; Whitesides, G. M.; Ajdari, A. Anal. Chem.<br />

2002, 74, 5306–5312.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6757


slanted or herringbone grooves were fabricated in poly(dimethylsiloxane)<br />

(PDMS) chips, and the mixing characteristics<br />

of these channels compared to channels without grooves. A<br />

thin-film Pt electrode at the end of the fluidic channel served<br />

as electrochemical detector of the reaction product, H2O2.<br />

2. MATERIALS AND METHODS<br />

2.1. <strong>Chemical</strong>s and Reagents. All chemicals were analytical<br />

reagent-grade. Fluorescein and potassium iodide (KI) were<br />

purchased from Sigma-Aldrich (Germany) and used to prepare<br />

0.002 M fluorescein solution and 0.2 M KI solution, respectively.<br />

Potassium ferricyanide (K3Fe(CN)6) and potassium ferrocyanide<br />

(K4Fe(CN) 6) were supplied by Sigma-Aldrich (Germany) and<br />

used to prepare a1mMsolution in a 20 mM phosphate buffer<br />

(pH 7.2). pH was measured using a pH meter (inoLab pH 720,<br />

WTW, Germany). The phosphate buffer was prepared by<br />

mixing solutions of NaH2PO4 · H2O (63 mL) and Na2HPO4 · 2H2O<br />

(100 mL) containing 0.1 M potassium nitrate (KNO3) asa<br />

supporting electrolyte. H2O2 (30%) was supplied by VWR (The<br />

Netherlands) and prepared freshly every day at various<br />

concentrations in phosphate buffer. D-glucose and GOx were<br />

supplied from Merck (Germany) and used to prepare 100 mM<br />

and 5000 U/mL stock solutions, respectively. D-glucose was<br />

stored in a refrigerator at 2 °C and used for 1 month. GOx<br />

was stored at -20 °C and used for 3 months. All solutions were<br />

prepared with 18 MΩ ultrapure water purified in an Arium 611<br />

(Sartorius Stedim Biotech, Germany).<br />

2.2. Microfluidic Chip Fabrication. The first microchannels<br />

were constructed by standard microfabrication and replicated in<br />

the silicone rubber, PDMS (Sylgard 184, Dow Corning, U.S.). The<br />

chip layout and design were drawn using the program, Clewin<br />

(Wieweb software, Hengelo, The Netherlands). The structure on<br />

the silicon master, which served as a mold, was processed with<br />

two steps of standard photolithography (Figure 1, right column). 28<br />

A 4 in. p-type (100) silicon wafer (Si-Mat, Germany) (525 µm<br />

thickness) was employed as a substrate. The silicon wafer was<br />

first cleaned sequentially with acetone, isopropyl alcohol, and<br />

deionized water, and dried with N2 gas. The wafer was then<br />

treated with hexamethyldisilazane (HMDS) (Sigma-Aldrich,<br />

Germany) in a vacuum desiccator for 30 min to improve<br />

adhesion of the photoresist (PR). A thick positive PR, AZ4562<br />

(Microchemicals GmbH, Germany), was coated on the silicon<br />

wafer using a spin-coater at 1000 rpm for 3 s and baked at 100<br />

°C for 30 min. The resulting PR layer was 35 µm thick. After<br />

rehydration at ambient temperature for 3 h, the coated wafer<br />

was exposed to ultraviolet (UV) light (365 nm, 10 mW/cm 2 )<br />

using a photomask printed on a high-resolution transparency<br />

to pattern the microchannels (resolution 3,810 dpi; Pro-Art BV,<br />

Groningen, The Netherlands). The exposed PR was then<br />

removed by dipping in a developer solution (AZ351B: deionized<br />

water (DI) ) 1:3, Microchemicals GmbH, Germany) and<br />

agitating the beaker for 20 min, leaving behind PR ridges for<br />

microchannel replication. The wafer was then rinsed in DI<br />

water and dried with N 2 gas. Contrary to usual protocol, no<br />

postbake step was undertaken for this layer. Rather, the<br />

substrate was again exposed to UV light for 10 s using a second<br />

transparency photomask to pattern the groove arrays. The<br />

(28) Kim, J.; Heo, J.; Crooks, R. M. Langmuir 2006, 22, 10130–10134.<br />

6758 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 1. Fabrication process of microfluidic chip. See text for<br />

details.<br />

patterned PR was again immersed in a developer (AZ351B: DI<br />

) 1:4) solution for 6 min to form groove structures on top of<br />

the microchannel structures.<br />

PDMS resin and a curing agent were used for microchannel<br />

replication in a PDMS slab. The two liquids were mixed at a<br />

weight ratio of 10:1 PDMS: curing agent, and the solution left to<br />

stand at ambient conditions for 20 min to remove air bubbles.<br />

Afterward, the mixture was poured over the prepared silicon<br />

master. A polycarbonate alignment piece was used together with<br />

fused-silica capillaries to form access holes to the microchannels<br />

during the replication process. Holes (300 µm in diameter) were<br />

drilled in the polycarbonate using a CNC machine (Sherline, U.S.)<br />

at points corresponding to the locations of the required inlets/<br />

outlets when this piece was aligned with the silicon master. Fusedsilica<br />

capillaries were then inserted through the holes and aligned<br />

on the microchannel structures on the silicon master. The<br />

prepolymer mixture was allowed to cure against the master at<br />

50 °C for4h.<br />

2.3. Thin-Film Electrode Formation. Platinum (Pt) sensing<br />

electrodes were formed on a glass wafer by a standard photolithography<br />

and lift-off process (Figure 1, left column). A 4 in., 525-<br />

µm-thick borofloat glass wafer (Telic, U.S.) was used as a substrate<br />

for the electrode patterning. The glass wafer was first cleaned<br />

using a piranha solution (96% H 2SO4: 30% H2O2 ratio of 3:1), then<br />

thoroughly rinsed in DI water and dried with N2 gas. (WARN-<br />

ING: Piranha solution must be handled with extreme caution,<br />

as it is highly oxidizing and reacts explosively when it comes<br />

into contact with easily oxidized organic solvents. This solution<br />

must be used in a well-ventilated fumehood with absolutely no<br />

organic solvents in the vicinity. Explosive reactions also occur<br />

in the presence of trace amounts of metals such as Pt, Ag, and<br />

Mn. Metal tweezers must not be used when cleaning wafers<br />

in piranha.) The wafer was then treated with HMDS in a<br />

vacuum desiccator for 30 min. An image reversal PR 5214E<br />

(Microchemicals GmbH, Germany) was coated using a spin-


coater at 2000 rpm for 20 s, and the substrate was placed on a<br />

hot plate at 105 °C for 50 s to soft-bake the resist and remove<br />

solvents. Subsequently, the substrate was exposed to UV light<br />

(365 nm, 10 mW/cm 2 ) for 2 s using a photomask with electrode<br />

pattern. After reversal baking of the substrate at 115 °C for 2<br />

min, a flood exposure was performed for 20 s without mask to<br />

solubilize unexposed areas, resulting in a negative image of<br />

the mask. Finally, the substrate was immersed in a developer<br />

solution (AZ351B: DI water, ratio 1:4) to remove soluble PR<br />

and expose the glass surface where the electrodes were to be<br />

formed.<br />

A lift-off process was performed to obtain Pt sensing electrodes<br />

on the patterned glass wafer. A Ti thin-film layer (20 nm) was<br />

deposited over the entire wafer by E-beam evaporation (Temescal,<br />

U.S.) under a high vacuum of 9 × 10 -6 Torr. Ti was deposited<br />

first, as it adheres better to glass than Pt, and in turn, Pt<br />

adheres well to Ti. Subsequently, a Pt thin-film layer (150 nm)<br />

was grown by E-beam evaporation at 2 × 10 -5 Torr. The glass<br />

substrate was then immersed in acetone to dissolve the<br />

remaining PR layer, thereby simultaneously lifting off the thin<br />

Pt/Ti film on top of it. The metal remained behind in the<br />

predefined electrode regions. Groups of three Pt electrodes<br />

were patterned; the patterned electrode dimensions were 100<br />

× 1500 µm, and the distance between two electrodes was 30<br />

µm.<br />

2.4. Chip Integration. The dimensions of the channel and<br />

grooves on the structured silicon master were determined using<br />

a stylus profilometer (Veeco Instrument BV, The Netherlands)<br />

before casting PDMS. There are two inlets and one outlet (waste)<br />

to the microreactor channel. The two inlet channels are half as<br />

wide as the microreactor channel after the Y-junction, to ensure<br />

a constant linear flow velocity throughout the device upon<br />

introduction of two solutions at the same flow rate. A 1:1 dilution<br />

of solutions will also result under these conditions of operation.<br />

The inlet channels are both 100 µm wide, 35 µm deep, and 5 mm<br />

in length. A ruler is located along the channel to show the distance<br />

from the Y-junction. The total length of channel from the Y-junction<br />

is 22 mm, while the width and depth of the channel are 200 and<br />

35 µm, respectively, and groove widths are 50 µm. The 122<br />

grooves in the mixing channel were 6-8 µm deep and were<br />

patterned at an angle of 45° with respect to the central axis of the<br />

channel. The total volume of the mixing/reaction channel is about<br />

150 nL.<br />

In order to bond the PDMS slab to the glass chip with the<br />

electrodes, the PDMS was smoothly peeled off from the master<br />

and cut. The two devices were first treated with UV-generated<br />

ozone for 15 min to oxidize the PDMS surface, which creates<br />

hydrophilic properties and improves bonding strength. 29 Subsequently,<br />

the PDMS slab and glass were immediately aligned under<br />

a microscope and brought into contact with each other. The<br />

assembled chip was then placed on a hot plate at 140 °C and<br />

allowed to cool down to room temperature, after which the chip<br />

was irreversibly bonded. The whole bonding procedure was<br />

carried out in the cleanroom. Figure 2 shows a fabricated<br />

microfluidic chip and examples of the three different types of<br />

PDMS channels with integrated electrodes investigated in this<br />

(29) Berdichevsky, Y.; Khandurina, J.; Guttman, A.; Lo, Y. H. Sens. Actuators, B<br />

2004, 97, 402–408.<br />

Figure 2. (a) Photograph of the detection region of a fabricated<br />

microfluidic chip. (b) Three different types of PDMS channels with<br />

integrated electrodes: from top to bottom, channels with no grooves,<br />

slanted grooves, and herringbone grooves. (c) Schematic diagram<br />

of grooves in a channel structure. Note that the grooves were 6-8<br />

µm deep.<br />

study, containing either no grooves, slanted grooves, or herringbone<br />

grooves.<br />

2.5. Experimental Setup for Fluorescence. In order to<br />

characterize mixing, fluorescence detection was used to gain<br />

quantitative information about the degree of mixing along the<br />

length of the channel. A fluorescence microscope, model “DMIL”<br />

(Leica Microsystems, The Netherlands), was equipped with a 10×<br />

objective, a mercury arc lamp, and a CCD camera. The mixing<br />

process was visualized using a fluorescein filter, which is made<br />

for 488 nm excitation and 518 nm emission. Fluorescence<br />

quenching was carried out for quantitative analysis using 2 mM<br />

fluorescein and 200 mM potassium iodide (KI). KI quenches the<br />

fluorescence produced by fluorescein when it is excited at 488<br />

nm. Fluorescein and KI were introduced separately through the<br />

inlets into the Y-junction at various flow rates, and the image was<br />

captured with a CCD camera using a 20.1 ms exposure time, a<br />

gamma setting of 1.0, and a gain of 1.0. The images were acquired<br />

at different distances from the Y-junction and analyzed by<br />

determining the standard deviation (SD) of the intensity distribution<br />

across the entire channel using Lx95P image analysis<br />

software. Flow rates were controlled by syringe pumps (ProSense,<br />

The Netherlands) connected with silica capillaries to the chip<br />

inlets using 350-µm-o.d. and 250-µm-i.d. silica capillaries.<br />

2.6. Experimental Setup for Electrochemical Detection.<br />

Cyclic voltammetry and chronoamperometry were carried out<br />

using a potentiostat, “Electrochemical Analyzer” (CH Instrument,<br />

U.S.), interfaced to a computer. Two thin-film Pt electrodes were<br />

used, one as a working electrode (WE), the other as a counter<br />

electrode (CE). The Pt electrodes were located 21 mm downstream<br />

from the Y-junction. The active area of the Pt WE was<br />

0.02 mm 2 (200-µm-wide channel ×100-µm-wide electrode). A<br />

Ag/AgCl wire was used as a reference electrode (RE) and<br />

positioned in the reservoir at the end of the device. The RE<br />

was prepared by dipping a short piece of Ag wire (250-µmdiameter)<br />

for 5 s into a crucible containing melted silver<br />

chloride (Sigma-Aldrich, Germany). For the glucose analysis<br />

experiments, a 1:3 flow rate ratio of glucose sample to GOx<br />

solution was introduced through the two inlets, and measurements<br />

were carried out under continuous flow conditions<br />

controlled by the syringe pumps. (For microdialysis experi-<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6759


Figure 3. Electrochemical detection experimental setup.<br />

ments, the microdialysis probe outlet will be connected to the<br />

glucose sample inlet of the device using a simple, low-deadvolume<br />

tubing connection.) For the comparison of system<br />

response to premixed glucose-GOx solutions and direct reaction<br />

of glucose and GOx in the channel, premixed solution was<br />

prepared using a 1:3 volume ratio of 5.6 mM glucose and 150<br />

U/mL GOx in a vial. After adding the two solutions, the vial<br />

was gently shaken 10 times and introduced in the reactor<br />

channel through both inlets approximately 5 min later after<br />

shaking. A two-position actuator switching valve (Valco Instrument<br />

Co., Inc.) was adapted to alternatively introduce buffer<br />

solution and glucose sample solution to measure background<br />

and sensing signals, respectively. Once the flow rate had been<br />

set each time, the valve was switched from buffer solution to<br />

sample solution. The same solution was used for measurements<br />

at each flow rate, and current measurements were made<br />

continuously as flow rate was varied. All experiments were<br />

carried out at room temperature, and data were saved directly<br />

on a computer. A schematic diagram of the electrochemical<br />

detection experimental setup is shown in Figure 3.<br />

3. RESULTS AND DISCUSSION<br />

3.1. Characterization of the Micromixers. A fabricated<br />

micromixer was characterized by capturing images at different<br />

distances from the Y-junction using a fluorescence microscope<br />

and determining the variation in fluorescence intensity across the<br />

width of the channel. The standard deviation (SD) was calculated<br />

using the following eq 3: 30<br />

SD ) � 1<br />

N<br />

N ∑ i)1<br />

(x i - x¯) 2<br />

where xi is the gray-scale intensity value of pixel i, and x¯ is the<br />

mean intensity value of pixels across the entire channel.<br />

Quantitative values of the SD of mixing efficiency varied<br />

between 0.5 for completely unmixed solutions (variation of<br />

intensity from 0 for KI solution to 100% fluorescence for<br />

fluorescein) and 0 for completely mixed solutions (intensity<br />

equal across the channel).<br />

Figure 4 shows the SD of intensity versus distance as a function<br />

of flow rate, comparing channels with slanted grooves and no<br />

(30) Xia, H. M.; Wan, S. Y. M.; Shu, C.; Chew, Y. T. Lab Chip 2005, 5, 748–<br />

755.<br />

6760 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(3)<br />

Figure 4. Comparison of microfluidic mixers having no grooves and<br />

slanted grooves as a function of distance from the Y-junction and at<br />

different flow rates (one device, n ) 3). The flow rate in each case is<br />

the total flow rate in the mixing channel. The flow rates of fluorescein<br />

and KI at the inlets are each half of the total flow rate.<br />

grooves. Flow rates that fall in the range required to achieve high<br />

glucose recoveries in microdialysis sampling of subcutaneous<br />

tissue were considered. The results showed that grooved channels<br />

have a higher mixing efficiency compared to channels without<br />

grooves. In fact, mixing in channels with slanted grooves is almost<br />

complete at a distance of 1 cm along the channel at total flow<br />

rates ranging from 0.4 to 2 µL/min (Reynolds number ) 0.08 at<br />

1 µL/min). This result confirms that grooves formed in the<br />

channel ceiling at an oblique angle with respect to flow direction<br />

enhance the mixing rate of two adjacent solution streams.<br />

Differences in measured standard deviations of fluorescence<br />

intensity at shorter distances along the channel are most likely<br />

due to the reproducibility of the experiment itself. The depth and<br />

width of the grooves were designed based on Stroock’s geometric<br />

parameters (relative groove height to channel height, R < 0.3;<br />

channel height, h E width of the channel, w) and have values of<br />

6-8 and 50 µm, respectively, as shown in Figure 2 (c). 27 These<br />

initial results of mixing characterization were used as important<br />

information for further mixing and reaction experiments.<br />

3.2. Cyclic Voltammograms of Three Different Types of<br />

Microfluidic Channel. Cyclic voltammograms were carried out<br />

to check the electrochemical properties of the microfluidic chip.<br />

A ferri/ferrocyanide couple provides an ideal electrochemical<br />

model for a preliminary chip test with integrated electrodes. The<br />

redox couple ferri/ferrocyanide reaction is as follows (eq 4):<br />

3- - 4-<br />

Fe(CN) 6 + e a Fe(CN)6 E° ) 0.356 V (4)<br />

First, the redox curve was examined with a stationary flow when<br />

1mM(K3Fe(CN)6) was introduced in the channel at a scan<br />

range of -0.3 to 0.7 V. The oxidation and reduction peaks were<br />

observed at 0.21 V (35 nA current) and 0.10 V (56 nA current)<br />

respectively, at a scan rate of 100 mV/s (data not shown). For<br />

cyclic voltammetry under continuous flow conditions, 1 mM<br />

Fe(CN)6 3- /1 mM Fe(CN)6 4- was prepared in 20 mM phosphatebuffered<br />

solution (pH 7.2) containing 0.1 M potassium nitrate<br />

as a supporting electrolyte. Experiments were performed at a


Figure 5. Comparison of cyclic voltammograms for the three different<br />

types of microfluidic channel. A 20 mM phosphate-buffered solution<br />

(pH 7.2) containing 1 mM Fe(CN)6 3- /Fe(CN)6 4- in 0.1 M potassium<br />

nitrate was used at a flow rate of 1 µL/min and a scan rate of 20<br />

mV/s.<br />

flow rate of 1 µL/min and a scan rate of 20 mV/s. The cyclic<br />

voltammograms recorded over an applied potential range of<br />

0-0.6 V versus Ag/AgCl are shown in Figure 5. In these early<br />

devices, the array of mixing grooves extended over the entire<br />

length of the channel, including the integrated electrodes. Higher<br />

oxidation and reduction currents were observed for channels with<br />

slanted or herringbone grooves than for unstructured channels.<br />

This was probably caused by higher local velocities over the<br />

electrode surfaces in the chaotic mixers, leading to thinner<br />

diffusion layers at these surfaces and thus improved analyte<br />

delivery and higher currents. Because of the observed, rather<br />

uncontrolled local flow effect on detector response, grooves were<br />

not structured over the electrodes in later devices used in the<br />

rest of this study. Instead, the groove patterns ended more than<br />

1 mm upstream from the electrodes to reduce the flow effect on<br />

the electrode surface, as shown Figure 2(b). Comparing structured<br />

channels, the herringbone grooves did not yield a significant<br />

improvement in electrochemical signal with respect to the slanted<br />

grooves in the channel. It was therefore decided to focus further<br />

studies on comparing channels with no grooves to channels<br />

containing slanted grooves. This was because channels with<br />

slanted grooves were easier to fabricate than channels with<br />

herringbone grooves; aligning the photomask for slanted grooves<br />

in the second photolithographic step was simpler than aligning<br />

the mask for herringbone grooves.<br />

A linear sweep voltammogram was recorded by introducing 6<br />

mM H 2O2 in 0.1 M KNO3 solution, in order to decide on an<br />

applied potential for glucose measurement. The onset of<br />

oxidation current was observed at 0.3 V and showed a plateau<br />

around 0.7 V, at a total flow rate of 1 µL/min and a scan rate<br />

of 100 mV/s (data not shown). Therefore, this applied potential<br />

was selected for subsequent glucose experiments.<br />

3.3. Optimization of GOx Concentration. Prior to the<br />

optimization of GOx concentration, electrode variation from one<br />

chip to the next was characterized using a premixed solution of<br />

20 mM glucose and GOx solution. One example of each mixer<br />

type (grooved and ungrooved) was selected, and signal variation<br />

was tested three times with each device. The results of these<br />

Figure 6. Amperometric response arising from reaction of 20 mM<br />

glucose with GOx at different concentrations in channels without<br />

grooves and with slanted grooves (one device, n ) 3). The flow rates<br />

of glucose and GOx were 0.5 and 1.5 µL/min, respectively; an<br />

example of raw data showing background signal and sensing signal<br />

is shown in the inset.<br />

experiments showed that although current values varied from one<br />

device to the next, current values fell within the same range. Two<br />

devices, one with a grooved micromixer, the other an ungrooved<br />

micromixer, were selected for subsequent experiments to compare<br />

the effect of no grooves versus slanted grooves in the channel.<br />

Since the concentration of GOx plays an important role in the<br />

reaction, various concentrations of GOx solutions were tested in<br />

the two channel types. To date, only a few papers have reported<br />

on-chip glucose sensing by mixing glucose sample and free<br />

enzyme in solution rather than using immobilized enzyme on<br />

electrode surfaces. 8,31-33 Pijanowska et al. 32 developed a flowthrough<br />

amperometric sensor implemented in a silicon-glass<br />

sandwich construction having a volume of 5 µL. The glucose<br />

sample was reacted in a reaction cell using only two different<br />

enzyme activities, 22 and 132 U/mL GOx, with a higher sensitivity<br />

being obtained with the latter concentration. Wang et al. 7 proposed<br />

a microchip capillary electrophoresis (MCE) approach for glucose<br />

sensing, in which GOx solution was injected together with glucose<br />

sample from separate reservoirs into the separation channel. A<br />

75 U/mL GOx solution proved sufficient for these researchers to<br />

obtain a calibration curve.<br />

Figure 6 shows the result of amperometric response arising<br />

from reaction of 20 mM glucose with GOx at different concentrations<br />

in channels with no grooves and slanted grooves. The flow<br />

rates were set at 0.5 µL/min glucose and 1.5 µL/min GOx (1:3<br />

mixing ratio) in order to avoid the oxygen depletion effect. 34 This<br />

is an effect which has been observed at higher concentrations of<br />

glucose for both electrochemical sensors utilizing immobilized<br />

GOx and devices based on the reaction of GOx and glucose in<br />

solution. The sensitivity of sensor response decreases because<br />

(31) Wang, J.; Chatrathi, M. P.; Collins, G. E. Anal. Chim. Acta 2007, 585,<br />

11–16.<br />

(32) Pijanowska, D. G.; Sprenkels, A. J.; Olthuis, W.; Bergveld, P. Sens. Actuators,<br />

B 2003, 91, 98–102.<br />

(33) Böhm, S.; Pijanowska, D.; Olthuis, W.; Bergveld, P. Biosens. Bioelectron.<br />

2001, 16, 391–397.<br />

(34) Wientjes, K. J. C. Ph.D. Dissertation, University of Groningen, Groningen,<br />

The Netherlands, 2000.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6761


the amount of oxygen available in the solution is not sufficient<br />

for full conversion of glucose conversion to H2O2 with enzyme<br />

(see also eq 1). In effect, oxygen is depleted from the solution<br />

by the enzyme reaction. To prevent the effect of oxygen depletion<br />

in measurements employing immobilized enzymes on electrodes,<br />

electrodes have been prepared with a semipermeable coating to<br />

reduce the amount of glucose diffusing to the electrode. Other<br />

approaches have involved the use of mediators, molecules which<br />

take on the role of oxygen in the reaction. 3,35 In early experiments,<br />

we also observed that the signal was dramatically decreased and<br />

calibration curves were not linear at higher concentrations of<br />

glucose when the mixing ratio between glucose and GOx was<br />

1:1; instead, the signal tended to level off. It was decided to<br />

increase the flow rate of GOx with respect to glucose sample to<br />

a ratio of 3:1, diluting the effective glucose concentration in the<br />

channel by a factor of 4 to prevent oxygen depletion.<br />

All experiments were performed three times and a repeatable<br />

output signal was observed (as shown in the inset of Figure 6).<br />

As the concentrations of GOx solution increased, oxidation current<br />

gradually increased from approximately 20 nA (10 U/mL) to reach<br />

a maximum of 100 and 120 nA (150 U/mL) in channels with no<br />

grooves and slanted grooves, respectively. At higher GOx concentrations,<br />

the current leveled off and then decreased in both<br />

channels with no grooves and slanted grooves. These results<br />

indicated that increasing concentrations of enzyme provided better<br />

conversion to H2O2 and product within a reaction time of<br />

approximately 4 s. However, at concentrations of GOx greater<br />

than 150 U/mL, the system exhibited less stable signal, as<br />

reported elsewhere; 8 this could be due to the interaction of<br />

GOx and H2O2 produced. Therefore, a 150 U/mL concentration<br />

of GOx solution was chosen as an optimum condition for<br />

further experiments. Compared with slanted grooves in the<br />

channel, the absence of grooves in the channel resulted in<br />

lower current values. It can be concluded that the chaotic flow<br />

pattern results in more efficient mixing caused by transverse<br />

flow in the channel.<br />

3.4. Comparison of System Response for Premixed<br />

Glucose-GOx Solutions and Reaction of Glucose and GOx<br />

in the Channel. In order to investigate dependence on flow rate,<br />

the response measured for premixed glucose-GOx solutions was<br />

compared with that measured for the reaction of glucose and GOx<br />

in the mixing/reaction channel as a function of flow velocity. For<br />

the latter experiments, a 5.6 mM glucose sample and a 150 U/mL<br />

GOx solution were introduced at a 1:3 mixing ratio in a slanted<br />

grooves channel and the flow velocity was varied from 0.95 mm/s<br />

(0.4 µL/min) to 14.3 mm/s (6 µL/min). The amperometric<br />

responses for premixed solutions and reaction in the mixing<br />

channel are presented in Figure 7. There are two dominant factors<br />

in this continuous flow-through reaction system which affect<br />

detector response, namely flow velocity and reaction time. The<br />

effect of flow velocity was measured by introducing premixed<br />

solution (5.6 mM glucose sample and 150 U/mL GOx solution<br />

had already reacted to produce H 2O2) in both inlets and varying<br />

the flow rate. The results showed that as the flow velocity<br />

increased, the recorded current values also increased gradually.<br />

It was also observed that current values increased as a function<br />

(35) Dixon, B. M.; Lowry, J. P.; O’Neill, R. D. J. Neurosci. Methods 2002, 119,<br />

135–142.<br />

6762 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 7. Amperometric response in a channel with slanted grooves,<br />

arising from a 1:3 reaction ratio of 5.6 mM glucose and 150 U/mL<br />

GOx as a function of flow velocity (from 0.4 to 6 µL/min). The response<br />

due to premixed solution (5.6 mM glucose and 150 U/mL GOx) was<br />

compared with the glucose-GOx reaction in the mixing channel (a<br />

single device was used for both curves, n ) 3 for each point).<br />

of flow velocity when experimenting with 0.5 mM H2O2<br />

solutions (data not shown). This is very different from the<br />

decreasing signal observed for GOx immobilized on electrodes<br />

at higher flow rates, which Lowry et al. ascribed to the removal<br />

of H2O2 from the electrode surface. 36 On the other hand, Wu<br />

et al. 37 observed currents which increased as a function of flow<br />

rate at two closely spaced oxygen sensors in a microchannel.<br />

This latter report supports our hypothesis that the observed<br />

increase in current values for premixed solutions is due to the<br />

thinner diffusion layers at the electrode surface which result<br />

at higher flow rates. These layers are stagnant, and molecules<br />

must diffuse through them to reach the electrode surface. As<br />

diffusion layers decrease in thickness, the time required for<br />

diffusion to the electrode surface also decreases, resulting in<br />

the delivery of analyte to that surface to be less limited by<br />

diffusion. Higher rates of electron exchange and thus higher<br />

recorded currents are the result.<br />

In the case of glucose reaction with GOx in the mixing channel,<br />

a decreasing signal as a function of flow velocity was observed.<br />

This decrease in current is primarily due to lower H2O2 concentrations,<br />

the result of shorter reaction times. The maximum<br />

reaction of glucose and GOx was observed at very low flow<br />

rates (lower than 0.4 µL/min). Thus, based on reaction times<br />

of 22 and 4 s (0.4 and 2 µL/min, respectively) from the<br />

Y-junction to the electrodes, the most efficient conversion of<br />

glucose with GOx can be achieved either at very low flow rates<br />

and/or in longer reaction channels.<br />

3.5. Calibration Curve. A calibration curve was obtained<br />

when a 150 U/mL GOx solution was reacted with glucose sample<br />

in a channel at a total flow rate of 2 µL/min, with a 1:3 flow rate<br />

(and thus 1:4 dilution) ratio of glucose and GOx. Figure 8 shows<br />

the amperometric response as a function of glucose concentration<br />

(36) Lowry, J. P.; McAteer, K.; El Atrash, S. S.; Duff, A.; O’Neill, R. D. Anal.<br />

Chem. 1994, 66, 1754–1761.<br />

(37) Wu, J.; Ye, J. Lab Chip 2005, 5, 1344–1347.


in channels with either no grooves or slanted grooves. The data<br />

exhibit a linear relationship between current and glucose concentrations<br />

of 0-20 mM, with a sensitivity of 5.3 and 6.9 nA/mM<br />

in channels with no grooves and slanted grooves, respectively.<br />

The current density was calculated by dividing sensitivity by the<br />

active area of the working electrode, yielding 26.5 and 34.5 mA/<br />

M · cm2 for channels with no grooves and slanted grooves,<br />

respectively. The linearity of the curves was tested by plotting<br />

observed current values versus predicted current values and<br />

examining the distribution of points around the resulting<br />

diagonal line (data not shown). In both cases, points were very<br />

symmetrically distributed around the line. The slope of the plot<br />

for the no-groove case was 0.9992, with an R2 of 0.9983. In the<br />

groove case, the slope was 1.0032, with an R2 Figure 8. Calibration curves obtained using 150 U/mL GOx solution<br />

at a total flow rate of 2 µL/min (0.5 µL/min glucose: 1.5 µL/min GOx)<br />

in two channel types, without grooves and with slanted grooves (one<br />

device, n ) 3). Glucose concentrations were in the clinical range of<br />

interest.<br />

of 0.9989. In both<br />

cases, excellent linearity was thus observed.<br />

In our approach, the flow velocity and time for reaction of<br />

glucose sample and GOx enzyme strongly affect the sensitivity<br />

(38) Ghosal, S. Anal. Chem. 2002, 74, 771–775.<br />

(39) Mei, Q.; Xia, Z.; Xu, F.; Soper, S. A.; Fan, Z. H. Anal. Chem. 2008, 80,<br />

6045–6050.<br />

of the sensor signal, allowing flow conditions to be tuned to obtain<br />

optimal results. One of the big advantages of this microfluidic<br />

reactor approach is that by varying flow rates of glucose and GOx,<br />

it is possible not only to avoid the oxygen depletion effect without<br />

any electrode treatment but to tune the sensitivity for the<br />

application. However, continuous perfusion of glucose and GOx<br />

to the device does lead to a slightly decreased signal over time.<br />

This might be caused by GOx adsorbing either onto the microchannel<br />

or electrode surface as described elsewhere. 38 This effect<br />

will be further investigated as part of the development of systems<br />

for long-term measurement of glucose in in vivo applications.<br />

4. CONCLUSION<br />

We have successfully demonstrated an enzymatic glucose<br />

reactor based on chaotic mixing in a microfluidic channel network<br />

for continuous glucose monitoring. Together with another recent<br />

report describing improved lucerifase detection in a chaotic<br />

mixer, 39 our example of glucose detection is one of the first<br />

examples of this type of mixer being applied to the enhancement<br />

of a biochemical reaction at the nL scale. A linear calibration curve<br />

was obtained in both microchannels with slanted grooves and no<br />

grooves, using a 150 U/mL GOx enzyme solution. Higher<br />

sensitivity was obtained with slanted groove arrays compared to<br />

micromixers with no grooves, due to enhanced mixing. The<br />

factors which determine sensitivity are flow velocity and extent<br />

of reaction. Though there is a loss of GOx due to the continuous<br />

flow to the outlet, this disadvantage is alleviated by the use of<br />

micofluidics for nL liquid handling and the application of low flow<br />

rates (2 µL/min for the optimized system). The low flow rates<br />

used are also compatible with microdialysis sampling and are<br />

required to achieve a high recovery of glucose from the subcutaneous<br />

tissue and reduce solution consumption. The possible<br />

influence of interfering substances such as ascorbic acid, uric acid,<br />

and acetaminophen on glucose determination are now under<br />

investigation, prior to testing the system in vivo in rats.<br />

Received for review January 8, 2010. Accepted June 24,<br />

2010.<br />

AC1000509<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6763


Anal. Chem. 2010, 82, 6764–6769<br />

Development of a Compound-Specific Carbon<br />

Isotope Analysis Method for 2-Methyltetrols,<br />

Biomarkers for Secondary Organic Aerosols from<br />

Atmospheric Isoprene<br />

Qiang Li, § Wu Wang,* ,†,‡ Hong-Wei Zhang, ‡ Yang-Jun Wang,* ,† Bing Wang, | Li Li, † Huai-Jian Li, †<br />

Bang-Jin Wang, † Jie Zhan, † Mei Wu, ⊥ and Xin-Hui Bi @<br />

Institute of Environmental Pollution and Health, Shanghai University, 200444 Shanghai, China, Southern Medical<br />

University, 510515 Guangzhou, China, School of Stomatology, Tongji University, 200072 Shanghai, China, Hefei<br />

Artillery Academy, 230002 Hefei, China, Laboratory of Human Micromorphology, Qingdao University Medical College,<br />

266071 Qingdao, China, and State Key Laboratory of Organic Geochemistry, Guangzhou Institute of Geochemistry,<br />

Chinese Academy of Sciences, 510640 Guangzhou, China<br />

The stable carbon isotope compositions of 2-methyltetrols,<br />

biomarker compounds for secondary organic<br />

aerosols formed from isoprene in the atmosphere, have<br />

been determined by gas chromatography/combustion/<br />

isotope ratio mass spectrometry (GC/C/IRMS). In this<br />

work, isoprene with various δ 13 C values was used to<br />

produce 2-methyltetrols via an oxidation reaction<br />

with hydrogen peroxide in sulfuric acid under direct<br />

sunlight. The target compounds with different stable<br />

carbon isotope compositions were then derivatized<br />

by methylboronic acid with a known δ 13 C value and<br />

measured by GC/C/IRMS. With δ 13 C values of 2-methyltetrols<br />

and methylboronic acid predetermined,<br />

isotopic fractionation is evaluated for the derivatization<br />

process. Through reduplicate δ 13 C measurements,<br />

the carbon isotope analysis achieved excellent<br />

reproducibility and high accuracy with an average<br />

error of


indicators of atmospheric processing of volatile organic compounds;<br />

9 for example, carbon isotope ratio measurements allow<br />

the calculation of the extent of photochemical processing of<br />

isoprene in the atmosphere. 10 Furthermore, Rudolph et al. 11<br />

developed a gas chromatography/combustion/isotope ratio mass<br />

spectrometry (GC/C/IRMS) technique to measure the stable<br />

carbon isotope ratios of isoprene and its gas-phase oxidation<br />

products, MACR and MVK, and to gain insight into the atmospheric<br />

oxidation of isoprene. Very recently, this group has<br />

reported studies on the stable carbon kinetic isotope effects (KIE)<br />

of the reactions of isoprene, MACR, and MVK with OH radicals,<br />

as well as with ozone in the gas phase. 12-14 These data are<br />

valuable for obtaining insight into the role of loss processes in<br />

determining the atmospheric mixing ratios. However, there are<br />

no isotopic studies of isoprene SOA products in the aerosol phase.<br />

Methylboronic acid (MBA) has been reported to determine<br />

natural 13 C abundances of monosaccharides by derivatizing<br />

adjacent hydroxyl groups of monosaccharides followed by N,Obis(trimethylsilyl)trifluoroacetamide<br />

(BSTFA) derivatization of<br />

the remaining single OH groups. 15,16 Recently, Boschker et al. 17<br />

reported a versatile method for stable carbon isotope analysis of<br />

carbohydrates by high-performance liquid chromatography/<br />

isotope ratio mass spectrometry. To develop a compound-specific<br />

isotope analysis method for 2-methyltetrols, marker compounds<br />

of photooxidation products of isoprene, a technique from our<br />

previous work was adapted. 18 MBA was used as the derivatization<br />

reagent prior to gas chromatography/combustion/isotope ratio<br />

mass spectrometry (GC/C/IRMS). The derivatizing C from the<br />

reagent accounts for 29% of the analyte in the boronates, but 70%<br />

in the trimethylsilyl (TMS) derivatives. Therefore, the sensitivity<br />

of the MBA method should be high. The oxidation reaction of<br />

isoprene, atmospheric sampling, accuracy, and reproducibility of<br />

the method will be discussed in detail, and the stable carbon<br />

isotope effects during the procedure will be evaluated. δ 13 C data<br />

for atmospheric 2-methyltetrols will also be presented to<br />

demonstrate the practical utility of this method.<br />

EXPERIMENTAL SECTION<br />

Materials. Isoprene was obtained from three suppliers: Fluka,<br />

Sigma-Aldrich (>98% pure, M1); Alfa-Aesar (Lancaster, England)<br />

(99% pure, M2); and Toyo Kasei Kogyo Co. (Osaka, Japan) (99%<br />

pure, M3). Hydrogen peroxide (30% in water) was purchased from<br />

Guoyao (Shanghai, China). Methylboronic acid (MBA) was<br />

(9) Rudolph, J.; Czuba, E. Geophys. Res. Lett. 2000, 27, 3865–3868.<br />

(10) Rudolph, J.; Anderson, R. S.; Czapiewski, K. V.; Czuba, E.; Ernst, D.;<br />

Gillespie, T.; Huang, L.; Rigby, C.; Thompson, A. E. J. Atmos. Chem. 2003,<br />

44, 39–55.<br />

(11) Iannone, R.; Koppmann, R.; Rudolph, J. J. Atmos. Chem. 2007, 58, 181–<br />

202.<br />

(12) Rudolph, J.; Czuba, E.; Huang, L. J. Geophys. Res. 2000, 105, 29323–29346.<br />

(13) Iannone, R.; Koppmann, R.; Rudolph, J. Atmos. Environ. 2008, 38, 4093–<br />

4098.<br />

(14) Iannone, R.; Koppmann, R.; Rudolph, J. Atmos. Environ. 2009, 43, 3103–<br />

3110.<br />

(15) van Dongen, B.; Schouten, S.; Damste, J. Rapid Commun. Mass Spectrom.<br />

2001, 15, 496–500.<br />

(16) Gross, S.; Glaser, B. Rapid Commun. Mass Spectrom. 2004, 18, 2753–2764.<br />

(17) Boschker, H. T. S.; Moerdijk-Poortvliet, T. C. W.; van Breugel, P.;<br />

Houtekamer, M.; Middelburg, J. J. Rapid Commun. Mass Spectrom. 2008,<br />

22, 3902–3908.<br />

(18) Wang, W.; Li, L.; Li, H.; Zhang, D.; Wen, S.; Jia, W.; Wang, B.; Sheng, G.;<br />

Fu, J. Rapid Commun. Mass Spectrom. 2009, 23, 2675–2678.<br />

purchased from ABCR GmbH and Co. KG (Karlsruhe, Germany)<br />

(97% pure) and recrystallized three times from a benzene/acetone<br />

mixture (3:1). MBA of the same lot number was used for all<br />

derivatizations. Anhydrous pyridine (99% pure) was supplied by<br />

Acros Organics (Geel, Belgium). BSTFA [N,O-bis(trimethylsilyl)<br />

trifluoroacetamide] was purchased from Pierce (Rockford, IL). All<br />

solvents employed were HPLC grade.<br />

Preparation of Standard 2-Methyltetrols. 2-Methyltetrols<br />

were made by photooxidation of isoprene, which we conducted<br />

by exposing a 30 mL flask with a mixture of 5 mL of 30% H2O2<br />

and 5 mL of isoprene (0.05 mol, 3.4 g) to sunlight. A few drops<br />

of sulfuric acid (0.1 M) was added until the pH of reaction<br />

mixture was between 1 and 2. The resulting mixture was then<br />

maintained with continuous and vigorous stirring in sunlight<br />

for4h; 19 15 mg of barium carbonate was added to 1 mL of the<br />

reacted solution for neutralization. After centrifugation, the<br />

supernatant was dried, and a slightly yellow oil (2.5 g, 36% yield)<br />

was obtained. It worth noting that exposure to sunlight is<br />

crucial for the production of 2-methyltetrols.<br />

The purification of crude 2-methyltetrols was performed<br />

according to a procedure reported by Wang et al. 20 The purity of<br />

2-methyltetrols was verified by gas chromatography/mass spectrometry<br />

(GC/MS) after they had been derivatized with BSTFA,<br />

and the δ 13 C value was determined by elemental analyzer/<br />

isotope ratio mass spectrometry (EA/IRMS).<br />

Derivatization of 2-Methyltetrols. The derivatization technique<br />

of 2-methyltetrols with methylboronic acid was adapted from<br />

Wang et al.; 18 100 µL of a solution of 2-methyltetrols (approximately<br />

1 mg/mL in methanol) was dried under a gentle<br />

nitrogen flow, and then 5 mL of a solution of 1 mg of methylboronic<br />

acid in 10 mL of anhydrous pyridine was added. The molar<br />

ratio of MBA to 2-methyltetrols was ∼10:1. The mixture was<br />

allowed to react at 60 °C for 60 min. It should be noted that the<br />

pretreatment of pyridine with 4 Å molecular sieves in excess is<br />

crucial for a successful MBA derivatization. The δ 13 C value of<br />

methylboronic derivatives was determined by gas chromatography/combustion/isotopic<br />

ratio mass spectrometry (GC/C/<br />

IRMS).<br />

Measurements of the δ 13 C Value of Standard Isoprene.<br />

The method for determining the δ 13 C value of isoprene was as<br />

follows. 21 Isoprene (1 mL) was sealed in a2mLglass vial with<br />

an open screw cap containing a Teflon-lined silica septum. After<br />

∼1 h for equilibrium, 15 µL gas samples from the glass bottle<br />

were injected into the split/splitless injection port of the gas<br />

chromatograph using a Hamilton gastight locking syringe.<br />

Aerosol Sampling. Samples were collected in boreal-temperate<br />

Changbai Mountain Forest Ecosystem Research Station in Jilin<br />

Province and subtropical Dinghu Mountain Nature Reserve in<br />

Guangdong Province. The details for sampling sites have been<br />

described previously. 6 The sampling occurred during the summer<br />

when the meteorological conditions and the maximum solar<br />

radiation, as well as high temperatures, were favorable for the<br />

photooxidation of isoprene. A high-volume PM2.5 air sampler<br />

(19) Santos, L.; Dalmazio, L.; Eberlin, M.; Claeys, M.; Augusti, R. Rapid Commun.<br />

Mass Spectrom. 2006, 20, 2104–2108.<br />

(20) Wang, W.; Vas, G.; Dommisse, R.; Loones, K.; Claeys, M. Rapid Commun.<br />

Mass Spectrom. 2004, 18, 1787–1797.<br />

(21) Yu, Y.; Wen, S.; Feng, Y.; Bi, X.; Wang, X.; Peng, P.; Sheng, G.; Fu, J. Anal.<br />

Chem. 2006, 78, 1206–1211.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6765


(Graseby-Andersen) was operated at a flow rate of 1.13 m3 /<br />

min, and Whatman quartz fiber filters (20.3 cm × 25.4 cm) were<br />

used; 24 h samplers were collected. All filters were baked at<br />

550 °Cfor4htoremove organic contaminants. After collection,<br />

the filters were stored at -20 °C until they were analyzed. Part<br />

of the filters was extracted with methanol. The solvent was<br />

then concentrated, filtered, and finally dried completely. All<br />

samples were derivatized using the same procedure as described<br />

above and measured by GC/C/IRMS.<br />

<strong>Analytical</strong> Systems. The gas chromatography/mass spectrometry<br />

(GC/MS) instrument consisted of a Hewlett-Packard<br />

(Fullerton, CA) model 6890 gas chromatograph equipped with a<br />

DP-5MS device (30 m × 0.25 mm inside diameter, 0.25 µm film<br />

thickness), coupled to a Hewlett-Packard model 5975MSD quadrupole<br />

analyzer. Data were acquired and processed with Chem-<br />

Station (Hewlett-Packard). The temperature program was as<br />

follows: initial temperature at 100 °C held for 5 min, a gradient of<br />

3 °C/min up to 200 °C, a gradient of 30 °C/min up to 290 °C,<br />

held for 2 min. The mass spectrometer was operated in the<br />

electron ionization mode at 70 eV and an ion source temperature<br />

of 150 °C. Full scan mode was used in the mass range of m/z<br />

50-420.<br />

Two kinds of gas chromatography/combustion/isotopic ratio<br />

mass spectrometry (GC/C/IRMS) systems were used in this<br />

study. The analysis of MBA derivatives was performed on an HP<br />

6890 GC system (Agilent, Santa Clara, CA) equipped with a DP-<br />

5MS device (30 m × 0.25 mm × 0.25 µm), connected to an isotope<br />

ratio mass spectrometer (Isoprime, GV instrument, Manchester,<br />

U.K.). The oven temperature program was as follows: initially at<br />

60 °C, a gradient of 4 °C/min up to 110 °C, held for 2 min, a<br />

gradient of 50 °C/min up to 290 °C, held for 2 min. The injector<br />

was set at 250 °C in splitless mode. Helium was used as the carrier<br />

gas at 1.5 mL/min. CO2 with a known δ13C value (-26.65‰)<br />

was used as the external reference gas. The combustion<br />

furnace containing the CuO catalyst and the reduction oven<br />

containing the Cu catalyst were kept at 940 and 650 °C,<br />

respectively. The temperature of the interface between the GC<br />

and combustion furnace was set at 290 °C. The reproducibility<br />

and accuracy of carbon isotopic analyses were evaluated<br />

routinely every day using 10 laboratory isotopic standards (C12,<br />

C14, C16, C18, C20, C22, C25, C28, C30, and C32 n-alkanes supplied<br />

by Indiana University, Bloomington, IN) with known isotopic<br />

values (-31.89, -30.67, -30.53, -31.02, -32.24, -32.77,<br />

-28.49, -32.11, -33.05, and -29.41‰, respectively). 2-Methyltetrol<br />

methylboronates prepared in the laboratory were<br />

analyzed 10 times and used as laboratory standards. For δ13C analysis of isoprene, the same GC/C/IRMS system described<br />

above and a CP-PoraPLOT Q column (25 m × 0.32 mm ×<br />

10 µm, Varian) were used. The GC was run in a split ratio of<br />

∼40:1, and the injector was set at 150 °C. The initial oven<br />

temperature was held at 120 °C for 1 min, followed by a<br />

gradient of 20 °C/min up to 195 °C, at which point the<br />

temperature was held for 10 min. Standard CO2 (δ13C )<br />

-26.65‰) was used as the external reference gas. CH4 with a<br />

known δ13C value (-36.30‰) was used as the laboratory<br />

isotopic standard to check the reproducibility and accuracy of<br />

6766 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

the carbon isotopic analysis routinely. 22 For both analysis<br />

systems, the sextuple analysis of laboratory isotopic standards<br />

indicated excellent accuracy and reproducibility of carbon<br />

isotopic analysis [the corresponding standard deviation ranged<br />

from 0.13 to 0.30‰, and the deviation between the measured<br />

data and the predetermined data ranged from -0.02 to 0.13‰<br />

(Table 1S of the Supporting Information)].<br />

Elemental analyzer/isotope ratio mass spectrometry (EA/<br />

IRMS) was performed as follows. Standard 2-methyltetrols and<br />

recrystallized methylboronic acid were put into cleaned tin<br />

capsules and weighed, repectively. Capsules containing weighed<br />

samples were placed in the CE (Wigan, U.K.) EA1112 C/N/S<br />

analyzer and burned at 960 °C in an O2 atmosphere in a<br />

combustion tube. Combustion gases were swept through a<br />

reduction oven and entered a GC column where CO2 was<br />

separated from other gases. Then the CO2 passed through a<br />

Conflo III interface (Finnigan, Waltham, MA) and entered a<br />

DELTA plus XL mass spectrometer (Thermo Finnigan MAT,<br />

Bremen, Germany) where it was compared to the reference<br />

CO2 with a known δ 13 C value (-29.10‰, calibrated against the<br />

NBS-22 reference material with a δ 13 C value of -29.70‰).<br />

During every batch of analyses, an empty tin capsule was<br />

analyzed as a blank to check the background, and the carbon<br />

black sample with a known δ 13 C value (-36.91‰) was used to<br />

evaluate the reproducibility and accuracy. The standard deviation<br />

of analysis and the deviation between the measured data<br />

and the predetermined data were less than 0.3‰ (Table 2S of<br />

the Supporting Information).<br />

All 13 C: 12 C ratios are expressed in conventional delta (δ)<br />

notation, which is the per mil (‰) deviation from the standard<br />

Vienna Pee Dee Belemnite (VPDB) reference point.<br />

RESULTS AND DISCUSSION<br />

δ 13 C Analysis of Standard 2-Methyltetrols. Figure 1<br />

shows a typical total ion current (TIC) GC/MS chromatogram<br />

from isoprene oxidation products derivatized by BSTFA. Identification<br />

of these two major compounds was made by comparison<br />

of the retention time and mass spectra with those published in<br />

the literature. 1,6,20 Figure 2 presents the TIC GC/MS chromatogram<br />

of 2-methyltetrol-MBA derivatives. The mass spectra of<br />

these three compounds are very similar, as could be expected<br />

for diastereoisomers. The mass spectrum of compound 2 reveals<br />

a tiny molecular ion (m/z 184). Characteristic ions at m/z 99 and<br />

85 correspond to the cleavage of the C-C bond at the C2 position.<br />

Other ions at m/z 69, 57, and 43 can be explained by the further<br />

fragment of the ion at m/z 85 or 99.<br />

The δ 13 C values of methylboronic acid and standard 2-methyltetrols<br />

were determined by EA/IRMS to be -24.96 ± 0.06‰<br />

(eight measurements) and -32.40 ± 0.14‰ (n ) 6, from M1),<br />

-29.84 ± 0.12‰ (n ) 6, from M2), and -32.36 ± 0.18‰ (n )<br />

6, from M3) (Table 1).<br />

δ 13 C Analysis of MBA Derivatives. In derivatization, the<br />

preparation of 2-methyltetrol-MBA derivatives alters the original<br />

stable isotope compositions of the 2-methyltetrols. To correct for<br />

the introduction of carbon during derivatization, it is necessary<br />

to assess the isotopic reproducibility of the derivatization method.<br />

(22) Wen, S.; Feng, Y.; Yu, Y.; Bi, X.; Wang, X.; Sheng, G.; Fu, J. Environ. Sci.<br />

Technol. 2005, 39, 6202–6027.


Figure 1. GC/MS total ion chromatograms obtained for the reaction mixtures of isoprene with H2O2 and sulfuric acid derivatized by BSTFA and<br />

spectra for products (insets) 1 (2-methylthreitol) and 2 (2-methylerythritol).<br />

Figure 2. GC/MS total ion chromatogram obtained for 2-methyltetrols derivatized by MBA and mass spectra for the products (insets). Compounds<br />

1 and 2 correspond to 2-methylerythritol-MBA derivatives, and compound 3 corresponds to a 2-methylthreitol-MBA derivative.<br />

Table 1. Stable Carbon Isotopic Compositions of Isoprene, 2-Methyltetrols (2-MT), and MBA Derivatives Measured<br />

and Predicted in the Derivatization Reaction<br />

supplier measured isoprene b,c<br />

measured 2-MT b,d<br />

The reproducibility of the carbon isotope composition for three<br />

2-methyltetrols (with different δ 13 C values) was evaluated. Their<br />

δ 13 C values and those of the corresponding MBA derivatives<br />

are listed in Table 1. The analytical errors (standard deviation)<br />

obtained for six EA/IRMS analyses of 2-methyltetrols produced<br />

from isoprene from the same supplier ranged from 0.12 to 0.18‰,<br />

with an average of 0.15 ± 0.03‰, while for GC/C/IRMS analyses<br />

of MBA derivatives, the analytical errors were from 0.15 to 0.22‰,<br />

with an average of 0.17 ± 0.04‰. The reproducibility compares<br />

well with those obtained in the derivatization of fatty acids. 23<br />

Isotopic Effects of the Method. According to eq 1, the<br />

theoretical δ 13 C values of methylboronates can be calculated<br />

(23) Abrajano, T. A.; Murphy, D. E., Jr.; Fang, J.; Comet, P. A.; Brooks, J. M.<br />

Org. Geochem. 1994, 21, 611–617.<br />

δ 13 C a<br />

measured methylboronates b,c,e<br />

calculated methylboronates b,f<br />

M1 -24.54 ± 0.18 (n ) 7) -32.40 ± 0.14 (n ) 6) -29.98 ± 0.22 (n ) 6) -30.27 0.29<br />

M2 -23.72 ± 0.23 (n ) 7) -29.84 ± 0.12 (n ) 6) -28.08 ± 0.15 (n ) 6) -28.24 0.16<br />

M3 -25.19 ± 0.19 (n ) 7) -32.36 ± 0.18 (n ) 6) -30.35 ± 0.16 (n ) 6) -30.25 -0.10<br />

a Stable carbon isotopic compositions reported in per mil relative to PDB. b Arithmetic means and standard deviations. c δ 13 C values determined<br />

by GC/C/IRMS. d δ 13 C values determined by EA/IRMS. e Derivatized by methylboronic acid with a δ 13 C value of -24.96 ± 0.06‰ (eight<br />

measurements) determined by EA/IRMS. f On the basis of mass balance relationship eq 1. g Calculated δ 13 C - measured δ 13 C.<br />

according to stoichiometric mass balance and reflect the<br />

relative contributions of carbon from 2-methyltetrols, methylboronic<br />

acid, and their respective δ 13 C values:<br />

δ 13 C methylboronate ) f MBA δ 13 C MBA + f 2-methyltetrol δ 13 C 2-methyltetrols<br />

(1)<br />

where fMBA and f2-methyltetrol are the molar fractions of carbon<br />

in methylboronate arising from underivatized 2-methyltetrol<br />

and MBA reagent, respectively. Here, fMBA ) 2 /7, and<br />

f2-methyltetrols ) 5 /7. The stable carbon isotopic compositions<br />

obtained for 2-methyltetrols and derivatives are listed in Table<br />

1. The measured data for methylboronates were compared with<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

∆ g<br />

6767


Figure 3. GC/C/IRMS chromatogram of 2-methyltetrol-MBA derivatives: (a) standard derivative and (b) extract of the fine size fraction of a<br />

24 h Hi-Vol aerosol sample collected at Dinghu, China, on August 2, 2006. Compounds 1 and 2 are 2-methylerythritol-MBA derivatives, and<br />

compound 3 is the 2-methylthreitol-MBA derivative.<br />

those predicted by eq 1. The predicted and measured δ 13 C values<br />

of methylboronates agreed well with each other.<br />

According to Rieley’s discussion of the kinetic isotope effect, 24<br />

when a bond containing the carbon atom under consideration is<br />

changed in the rate-determining step, the primary isotopic effect<br />

is the most significant. If no carbon bond changed in the ratedetermining<br />

reaction or if no carbon-containing bond is involved<br />

in this step, there should not be a primary isotope effect on the<br />

δ 13 C value. In this work, four hydroxyl groups from two<br />

molecules of MBA react with four hydroxyl groups of 2-methyltetrol<br />

and eliminate four molecules of water. No other carboncontaining<br />

bonds, except C-O bonds of 2-methyltetrols, are<br />

involved in the reaction. The difference between measured and<br />

calculated δ 13 C values of methylboronic derivatives ranged from<br />

-0.10 to 0.29‰ (Table 1). Accuracy was well within the isotope<br />

technical specification (±0.5‰). It should be pointed out that the<br />

analytical error of the calculated data for underivatized 2-methyltetrol<br />

(usually expressed as the standard deviation, S) could be<br />

calculated by the following equation:<br />

2<br />

S2-methyltetrol ) (1/f2-methyltetrol ) 2 2<br />

Smethylboronate +<br />

(fMBA /f2-methyltetrol ) 2 2<br />

SMBA where fMBA and f2-methyltetrol are the same as in eq 1 and S2-methyltetrol,<br />

SMBA, and Smethylboronate are the analytical standard deviations of<br />

(24) Rieley, G. Analyst 1994, 119, 915–919.<br />

6768 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(2)<br />

2-methyltetrols, MBA, and methylboronates, respectively. In<br />

this work, SMBA was 0.06‰ and Smethylboronate was 0.17 ± 0.04‰<br />

(0.15-0.22‰). According to eq 2, the standard deviation of the<br />

measured δ 13 C value of 2-methyltetrol ranged from 0.21 to<br />

0.31‰. The results imply that the method is promising and<br />

introduces no isotopic fractionation in the derivatization process,<br />

as confirmed by Table 1. The stable carbon isotopic<br />

determination exhibited high precision and good reproducibility.<br />

On the other hand, precursor isoprene was approximately<br />

6.12-7.86‰ more enriched in 13 C than 2-methyltetrols (Table<br />

1). This might reflect isotope fractionation during the photochemical<br />

oxidation reaction. Rudolph et al. 12 reported that the stable<br />

carbon isotope fractionation for the reaction of isoprene with OH<br />

radicals in the gas phase was 6.94 ± 0.80‰, very similar to the<br />

value in this work.<br />

To fully evaluate the usefulness of relating precursor isoprene<br />

δ 13 C values to those measured in aerosol, one must investigate<br />

the potential for the primary kinetic isotope effect (KIE)<br />

involving the carbon-carbon bond cleavage products (such as<br />

methylglyceric acid, glyoxal, and methylglyoxal) to influence<br />

the δ 13 C value of 2-methyltetrols. This is especially true since<br />

the reaction pathways for formation of 2-methyltetrols remain<br />

unclear. Gas-phase and mixed-phase mechanisms involving the<br />

aqueous aerosol phase have been considered. 1-3<br />

Measurements of Atmospheric Samples. The target compounds<br />

in samples were well-separated (Figure 3), while the<br />

2-methylerythritol derivative gave rise to double peaks, due to the


Table 2. Concentrations and Stable Carbon Isotopic Compositions of 2-Methyltetrols at Two Forested Sites<br />

sampling site sampling date concentration of 2-methyltetrols (ng/m 3 ) measured methylboronates b<br />

formation of both E and Z isomers during the reaction. From our<br />

previous work, 18 the erythro form of tetrols, erythritol, eluted<br />

earlier than the threo form, threitol. Under the same conditions,<br />

peaks 1 and 2 were identified as (Z)- and (E)-2-methylerythritol-MBA<br />

derivatives, respectively, while peak 3 was the<br />

2-methylthreitol-MBA derivative. The latter derivative gives rise<br />

to only one tiny peak, due to the stereo hindrance of the 2-methyl<br />

group. The δ 13 C value of the 2-methyltetrol-MBA derivatives<br />

was obtained by defining these three peaks as one using the<br />

integral tools of IRMS.<br />

It could be seen from Table 2 that δ 13 C values of 2-methyltetrols<br />

were distinctly different for the two sites. It has been<br />

documented that there were significant differences in carbon<br />

isotope ratios among C3, C4, and CAM plants, e.g., -35 to<br />

-23‰ for C3 with an average of -26‰, -14 to -10‰ for C4<br />

with an average of -13‰, and intermediate carbon isotope<br />

compositions for CAM. 25 The δ 13 C of 2-methyltetrols might<br />

reveal the relation of atmospheric aerosol with the vegetation<br />

types and be used as an indicator of proportions of C3 to C4<br />

plants in vegetation.<br />

CONCLUSIONS<br />

The stable carbon isotope compositions of 2-methyltetrols,<br />

biomarker compounds of isoprene in atmospheric aerosols, were<br />

determined by GC/C/IRMS. MBA was used as the derivatization<br />

reagent which allows only 29% contribution of the analyzed C.<br />

The δ 13 C values of 2-methyltetrols are calculated on the basis<br />

of the stoichiometric mass balance equation among 2-methyltetrols,<br />

methylboronic acid, and methylboronate derivatives.<br />

Excellent reproducibility and accuracy are achieved without<br />

carbon isotopic fractionation. Moreover, photosynthesis of<br />

2-methyltetrols through photochemical oxidation of isoprene<br />

(25) Cerling, T. E.; Wang, Y.; Quade, J. Nature 1993, 361, 344–345.<br />

is introduced. The δ 13 C values of atmospheric 2-methyltetrols<br />

were determined for two forest aerosols. Considering the<br />

complex formation pathways of 2-methyltetrols, this technique<br />

will be helpful for full field application and will have potential<br />

in differentiation between homogeneous and heterogeneous<br />

oxidation processes. The application of this method may<br />

provide additional information about the sources and sinks of<br />

atmospheric isoprene. Further experiments on photochemical<br />

oxidation in a smog chamber are warranted.<br />

ACKNOWLEDGMENT<br />

This work was funded by the Knowledge Innovation Program<br />

of the Chinese Academy of Sciences (Grant kzcx2-yw-139), the<br />

Natural Science Foundation of China (Grants 20677036, 20877051,<br />

and 40873073), the Innovation Program of Shanghai Municipal<br />

Education Commission (10YZ08), the Shanghai Municipal Health<br />

Bureau (054088), Shanghai Leading Academic Disciplines (S30109),<br />

and the Scientific Research Foundation for Returned Overseas<br />

Chinese Scholars, State Education Ministry, China. We thank the<br />

anonymous referee for insightful comments. We also thank Dr.<br />

Jia Wanglu (State Key Laboratory of Organic Geochemistry,<br />

Guangzhou Institute of Geochemistry, Chinese Academy of<br />

Sciences) for his technical assistance.<br />

SUPPORTING INFORMATION AVAILABLE<br />

δ 13 C values (‰) of the Indiana reference n-alkanes measured<br />

by GC/C/IRMS (Isoprime) (Table 1S) and δ 13 C values of the<br />

reference materials of CH4 and carbon black (Table 2S). This<br />

material is available free of charge via the Internet at<br />

http://pubs.acs.org.<br />

Received for review January 25, 2010. Accepted April 19,<br />

2010.<br />

AC100214P<br />

δ 13 C a<br />

calculated 2-MT c<br />

Changbai July 27, 2007 76.29 -24.27 ± 0.13 (n ) 3) -24.00 ± 0.18<br />

July 28, 2007 108.8 -24.79 ± 0.61 (n ) 3) -24.73 ± 0.86<br />

Dinghu August 1, 2006 65.30 -27.29 ± 0.10 (n ) 3) -28.22 ± 0.14<br />

August 2, 2006 83.54 -26.93 ± 0.46 (n ) 3) -26.99 ± 0.64<br />

a Stable carbon isotopic compositions reported in per mil relative to PDB. b Arithmetic means and standard deviations. Derivatized by methylboronic<br />

acid with a δ 13 C value of -24.96 ± 0.06‰ (eight measurements) determined by EA/IRMS. δ 13 C values determined by GC/C/IRMS. c Arithmetic<br />

means of calculated δ 13 C values based on mass balance relationship eq 1 and standard deviations (S) calculated according to eq 2.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6769


Anal. Chem. 2010, 82, 6770–6774<br />

Surface-Enhanced Raman Spectroscopy as a Tool<br />

for Detecting Ca 2+ Mobilizing Second Messengers<br />

in Cell Extracts<br />

Elina A. Vitol, † Eugen Brailoiu, ‡ Zulfiya Orynbayeva, § Nae J. Dun, ‡ Gary Friedman, † and<br />

Yury Gogotsi* ,|<br />

Department of Electrical and Computer Engineering, and Department of Materials Science and Engineering, Drexel<br />

University, Philadelphia, Pennsylvania 19104, and Department of Pharmacology, Temple University, Philadelphia,<br />

Pennsylvania 19140<br />

Understanding of calcium signaling pathways in cells is<br />

essential for elucidating the mechanisms of both normal<br />

cell function and cancer development. Calcium messengers<br />

play the crucial role for intracellular Ca 2+ release.<br />

We propose a new approach to detecting the calcium<br />

second messenger nicotinic acid adenine dinucleotide<br />

phosphate (NAADP) in cell extracts using surfaceenhanced<br />

Raman spectroscopy (SERS). Currently available<br />

radioreceptor binding and enzymatic assays require<br />

extensive sample preparation and take more than<br />

12 h. With a SERS sensor, NAADP can be detected in<br />

less than 1 min without any special sample preparation.<br />

To the best of our knowledge, this is the first<br />

demonstration of using SERS for calcium signaling<br />

applications.<br />

Calcium signaling is one of the fundamental cellular processes<br />

involved in any cell metabolic and physiologic activity. 1 Calcium<br />

signals convey information from the cell plasma membrane to<br />

intracellular targets. The mechanism of calcium concentration<br />

modulations is a complex problem associated with calcium influx<br />

from the extracellular matrix and release from intracellular stores<br />

mobilized by calcium messengers. Calcium signaling pathways<br />

of two calcium messengers, inositol trisphosphate (IP3) 2,3 and<br />

cyclic adenine dinucleotide ribose (cADPR), 4 have been studied<br />

extensively in different types of cells. Nicotinic acid adenine<br />

dinucleotide phosphate (NAADP) has a unique physiological role<br />

in cells 5 in the release of Ca 2+ from acid-filled calcium stores<br />

* To whom correspondence should be addressed. E-mail: gogotsi@drexel.edu.<br />

† Department of Electrical and Computer Engineering, Drexel University.<br />

‡ Temple University.<br />

§ School of Biomedical Engineering, Science and Health System, Drexel<br />

University.<br />

| Department of Materials Science and Engineering, Drexel University.<br />

(1) Patel, S.; Churchill, G. C.; Galione, A. Biochem. J. 2000, 352, 725–729.<br />

(2) Taylor, C. W.; Thorn, P. Curr. Biol. 2001, 11, R352–R355.<br />

(3) Cancela, J. M.; Gerasimenko, O. V.; Gerasimenko, J. V.; Tepikin, A. V.;<br />

Petersen, O. H. EMBO J. 2000, 19, 2549–2557.<br />

(4) Guse, A. H.; da Silva, C. P.; Berg, I.; Skapenko, A. L.; Weber, K.; Heyer, P.;<br />

Hohenegger, M.; Ashamu, G. A.; Schulze-Koops, H.; Potter, B. V. L.; Mayr,<br />

G. W. Nature 1999, 398, 70–73.<br />

(5) Lee, H. C.; Aarhus, R. J. Biol. Chem. 1995, 270, 2152–2157.<br />

6770 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

through two-pore channels 1, 2, and 3. 6,7 NAADP is the least<br />

investigated Ca 2+ mobilizing second messenger, because of the<br />

lack of widely accessible and efficient techniques for detecting<br />

and quantifying its concentration in cells. Enzymatic bioassays<br />

and radioreceptor binding assays are the primary methods<br />

which have been used for detecting NAADP in cell extracts. 8,9<br />

The enzymatic assay 9 requires NAADP to be first converted to<br />

nicotinamide adenine dinucleotide phosphate (NADP) using ADPribosyl<br />

cyclase, which is followed by two enzymatic cycling<br />

reactions of oxidation/reoxidation of NADP. 10 Diaphorase, the<br />

enzyme for reoxidation of NADP to nicotinamide adenine dinucleotide<br />

phosphate (NADPH), also serves as a catalyst for conversion<br />

of the reaction indicator resazurin to a highly fluorescent resorufin.<br />

The latter is then used for fluorimetric assessment of the NAADP<br />

concentration. Importantly, the described assay requires a very<br />

high purity of all of the components and takes more than 12 h. 10<br />

The radioreceptor binding assay is less time-consuming and can<br />

be conducted without extensive sample purification, 8 but due to<br />

the need for unique specialized equipment, the availability of this<br />

method is extremely limited.<br />

Here we present an alternative, label-free technique for the<br />

detection of NAADP enabled by surface-enhanced Raman spectroscopy<br />

(SERS). 11-13 SERS enhances Raman scattering due to<br />

the amplification of the electric field around metal nanostructures.<br />

14,15 Solutions of metal colloids have been used for SERS, 16,17<br />

but in some cases they show relatively poor data reproducibility<br />

resulting from uncontrollable aggregation of colloidal particles. 17–19<br />

For this reason, SERS sensors with metal nanostructures fixed<br />

(6) Brailoiu, E.; Churamani, D.; Cai, X. J.; Schrlau, M. G.; Brailoiu, G. C.; Gao,<br />

X.; Hooper, R.; Boulware, M. J.; Dun, N. J.; Marchant, J. S.; Patel, S. J. Cell<br />

Biol. 2009, 186, 201–209.<br />

(7) Calcraft, P. J.; Ruas, M.; Pan, Z.; Cheng, X. T.; Arredouani, A.; Hao, X. M.;<br />

Tang, J. S.; Rietdorf, K.; Teboul, L.; Chuang, K. T.; Lin, P. H.; Xiao, R.;<br />

Wang, C. B.; Zhu, Y. M.; Lin, Y. K.; Wyatt, C. N.; Parrington, J.; Ma, J. J.;<br />

Evans, A. M.; Galione, A.; Zhu, M. X. Nature 2009, 459, 596–U130.<br />

(8) Lewis, A. M.; Masgrau, R.; Vasudevan, S. R.; Yarnasaki, M.; O’Neill, J. S.;<br />

Garnham, C.; James, K.; Macdonald, A.; Ziegler, M.; Galione, A.; Churchill,<br />

G. C. Anal. Biochem. 2007, 371, 26–36.<br />

(9) Graeff, R.; Lee, H. C. Biochem. J. 2002, 367, 163–168.<br />

(10) Gasser, A.; Bruhn, S.; Guse, A. H. J. Biol. Chem. 2006, 281, 16906–16913.<br />

(11) Kneipp, K.; Wang, Y.; Kneipp, H.; Perelman, L. T.; Itzkan, I.; Dasari, R.;<br />

Feld, M. S. Phys. Rev. Lett. 1997, 78, 1667–1670.<br />

(12) Moskovits, M. Rev. Mod. Phys. 1985, 57, 783–826.<br />

(13) Nabiev, I. R.; Morjani, H.; Manfait, M. Eur. Biophys. J. 1991, 19, 311–316.<br />

(14) Vitol, E. A.; Orynbayeva, Z.; Bouchard, M. J.; Azizkhan-Clifford, J.; Friedman,<br />

G.; Gogotsi, Y. ACS Nano 2009, 3, 3529–3536.<br />

10.1021/ac100563t © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/26/2010


on a substrate are preferred. 20-24 The SERS sensor employed in<br />

this work is comprised of a glass substrate coated with gold<br />

nanoparticles. Similar sensors fabricated in the form of glass<br />

nanopipets have been recently demonstrated for minimally invasive<br />

in situ intracellular SERS measurements. 14 In the future, the<br />

results of this study could be potentially extended to NAADP<br />

detection inside cells using the SERS-based approach enabled by<br />

SERS-active nanopipets.<br />

EXPERIMENTAL SECTION<br />

Cell Culture. Breast cancer SkBr3 cells were grown in<br />

McCoy’s 5A modified medium, supplemented with 10% fetal<br />

serum, streptomycin, and penicillin. Cells were purchased from<br />

ATCC.<br />

Acid Extraction of NAADP. For the extraction of NAADP<br />

we used the protocol reported by Lewis et al. 8 All chemicals were<br />

purchased from Sigma-Aldrich. Briefly, SkBr3 cells were treated<br />

with trypsin and suspended in the cell medium. Before treatment<br />

with the agonists, cells were preincubated for 30 min with BAPTA-<br />

AM, 1,2-bis(2-aminophenoxy)ethane-N,N,N′,N′-tetraacetic acid tetrakis(acetoxymethyl<br />

ester). Then in the presence of BAPTA-AM<br />

cells were stimulated for 20 min with the agonists, according to<br />

the technique described in ref 8. This technique has been shown<br />

to trigger the prolonged NAADP synthesis for at least 20 min<br />

during the agonist stimulation. This results in highly amplified<br />

NAADP production. Here we used histamine, adenosine triphosphate<br />

(ATP), and acetylcholine, all at a 5 µM concentration. The<br />

reaction was stopped by adding 0.75 M ice-cold HClO4. Next the<br />

cells were disrupted by sonication and then kept on ice for 10<br />

min. The disrupted cells were centrifuged at 9000g for 10 min.<br />

Supernatant was neutralized with 1 M KHCO3 and vortexed.<br />

The resulting KClO4 precipitate was removed by centrifugation<br />

at 9000g for 10 min. Samples were stored at -80 °C for later<br />

analysis.<br />

Fabrication of the SERS Sensor. Microscope glass slides<br />

were cut into 1 cm × 1 cm pieces and sonicated in a mixture of<br />

NaOH and ethanol. After being washed with plenty of 15 MΩ<br />

deionized water, the slides were dried at room temperature. Next<br />

the slides were dip coated with 0.001% poly-L-lysine, dried at room<br />

temperature for 24 h, and then coated with gold nanoparticles by<br />

dipping them in the gold colloid for 3 h. Poly-L-lysine promotes<br />

the adhesion of the gold nanoparticles to the glass surface. The<br />

mechanism of nanoparticle attachment is based on the electrostatic<br />

interaction between the negatively charged particles and<br />

(15) Vo-Dinh, T.; Yan, F.; Wabuyele, M. B. J. Raman Spectrosc. 2005, 36, 640–<br />

647.<br />

(16) Ivleva, N. P.; Wagner, M.; Horn, H.; Niessner, R.; Haisch, C. Anal. Chem.<br />

2008, 80, 8538–8544.<br />

(17) Kneipp, J.; Kneipp, H.; McLaughlin, M.; Brown, D.; Kneipp, K. Nano Lett.<br />

2006, 6, 2225–2231.<br />

(18) Chourpa, I.; Lei, F. H.; Dubois, P.; Manfait, M.; Sockalingum, G. D. Chem.<br />

Soc. Rev. 2008, 37, 993–1000.<br />

(19) Willets, K. A. Anal. Bioanal. Chem. 2009, 394, 85–94.<br />

(20) McFarland, A. D.; Van Duyne, R. P. Nano Lett. 2003, 3, 1057–1062.<br />

(21) Hartschuh, A.; Qian, H.; Meixner, A. J.; Anderson, N.; Novotny, L. Surf.<br />

Interface Anal. 2006, 38, 1472–1480.<br />

(22) Haynes, C. L.; Van Duyne, R. P. J. Phys. Chem. B 2003, 107, 7426–7433.<br />

(23) Shoute, L. C. T.; Bergren, A. J.; Mahmoud, A. M.; Harris, K. D.; McCreery,<br />

R. L. Appl. Spectrosc. 2009, 63, 133–140.<br />

(24) Deckert, V.; Zeisel, D.; Zenobi, R.; Vo-Dinh, T. Anal. Chem. 1998, 70, 2646–<br />

2650.<br />

Figure 1. (a) Scanning electron micrograph of the SERS sensor,<br />

(b) close-up view of the gold nanoparticles on the SERS sensor, (c)<br />

extinction spectrum of the SERS sensor, and (d) SERS spectrum of<br />

NAADP and a background spectrum from the substrate.<br />

positively charged NH2 functional groups of poly-L-lysine. 14,25,26<br />

After fabrication, the substrates were imaged with a scanning<br />

electron microscope to confirm the nanoparticle distribution on<br />

the surface. SEM images were collected with a field emission Zeiss<br />

Supra 50VP scanning electron microscope at a low accelerating<br />

voltage (0.7-2 kV) without any conductive coating. In addition,<br />

the UV-vis extinction spectra of the substrates were measured<br />

using a home-built setup employing a fiber-optic spectrometer,<br />

HR-4000, Ocean Optics.<br />

SERS Measurements. Raman spectroscopy was performed<br />

using a micro-Raman spectrometer (Renishaw, RM 1000) equipped<br />

with a 632.8 nm HeNe laser (1800 lines/mm grating) and a diode<br />

InGaAs laser operating at 785 nm wavelength (1200 lines/mm<br />

grating). The lasers are manufactured by Renishaw Inc., U.K. The<br />

laser source was focused on the sample through a long working<br />

distance 50× objective to a spot size of approximately 2 µm. The<br />

typical sample volume was 1 µL. The acquisition time for all<br />

spectra was 10 s. Data analysis was performed using the Renishaw<br />

Wire 2.0 software. Experimental data were analyzed using principal<br />

component analysis 27 in the Matlab environment.<br />

RESULTS AND DISCUSSION<br />

Testing the SERS Sensor for Distinguishing between<br />

Different Secondary Ca 2+ Mobilizing Messengers: NAADP,<br />

cADPR, and IP3. Figure 1a shows the SEM image of the SERS<br />

sensor, with the close-up view presented in panel b. The average<br />

diameter of the nanoparticles is on the order of 50 nm. Assembly<br />

of the SERS sensor is based on the wet chemistry two-step<br />

protocol. First, the glass substrates are coated with a positively<br />

charged polymer (poly-L-lysine) layer. The functionalized substrates<br />

are then coated with a monolayer of negatively charged<br />

gold nanoparticles through electrostatic binding from the gold<br />

(25) Freeman, R. G.; Grabar, K. C.; Allison, K. J.; Bright, R. M.; Davis, J. A.;<br />

Guthrie, A. P.; Hommer, M. B.; Jackson, M. A.; Smith, P. C.; Walter, D. G.;<br />

Natan, M. J. Science 1995, 267, 1629–1632.<br />

(26) Nie, S.; Emory, S. R. Science 1997, 275, 1102–1106.<br />

(27) Jackson, J. E. A User’s Guide to Principal Components; John Wiley: New<br />

York, 1991.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6771


Figure 2. SERS spectra of aqueous solutions of NAADP, cADPR,<br />

and IP3. The bottom spectrum was collected from the mixture of all<br />

three Ca 2+ mobilizing second messengers (3 µM concentration of<br />

each). The excitation laser wavelength was 633 nm. The data<br />

acquisition time was 10 s. The data were collected immediately after<br />

1 µL of the sample was placed on the SERS sensor. The spectra<br />

are normalized and offset for clarity.<br />

colloid. The average interparticle distance is controlled by the time<br />

that the substrates are exposed to the gold colloid. Here, the<br />

average distance is approximately 75 nm. The UV-vis spectrum<br />

of the SERS-enabled substrate, with the maximum extinction at<br />

around 540 nm, is shown in Figure 1c.<br />

The selectivity of the SERS sensor for calcium messengers<br />

was studied with three different samples with 10 µM concentration:<br />

NAADP, IP3, and cADPR. The mixture of all three<br />

messengers was also analyzed. Figure 1 shows that each<br />

messenger has its characteristic SERS spectrum. Moreover, the<br />

analysis of the mixture of all three messengers (bottom spectrum<br />

in Figure 2) shows that it is possible to distinguish the features<br />

of each component.<br />

For example, the adenine moiety of NAADP represents itself<br />

in the spectrum of the mixture with the 733 cm -1 peak, similar<br />

to that observed in the spectrum of the control NAADP<br />

solution. The contribution from cADPR and IP3 to the mixture<br />

spectrum is confirmed by the presence of 898 cm -1 , 1257 cm -1<br />

(amide II), and 1416 cm -1 (C-H stretch) peaks, which are<br />

present in the spectra collected from pure solutions. The 898<br />

cm -1 peak in the spectrum of cADPR can be attributed to<br />

ribose. 17,28 Interestingly, although cADPR contains adenine, it<br />

shows only as a weak signal at 733 cm -1 . This is likely due to<br />

the circular structure of the cADPR molecule, where adenine<br />

is located between two ribose groups. Further analysis of the<br />

Raman spectra is beyond the scope of this work. The key result<br />

demonstrated above is that detection of a specific analyte<br />

(NAADP in our case) is clearly possible using spectral<br />

signatures as a whole obtained from SERS-enabled substrates.<br />

It is important to note that, for the enzymatic cycling assay,<br />

samples must be purified from NADP, which interferes with<br />

(28) Kneipp, J.; Kneipp, H.; Rice, W. L.; Kneipp, K. Anal. Chem. 2005, 77, 2381–<br />

2385.<br />

6772 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 3. SERS spectra of aqueous solutions of NAD, NADP, and<br />

NAADP at a 100 µM concentration. The data were collected using<br />

the 785 nm excitation laser; the signal acquisition time was 10 s.<br />

NAADP detection. SERS, however, makes it possible to distinguish<br />

between NAADP and its metabolites, such as nicotinamide<br />

adenine dinucleotide (NAD), NADP, and cADPR. Such specificity<br />

to the molecular structure is unattainable, as far as we know, by<br />

any other technique previously applied for NAADP detection.<br />

Figure 3 illustrates the difference between the SERS spectra of<br />

NAD, NADP, and NAADP at a 100 µM concentration. Furthermore,<br />

SERS substrates employed in this work can be utilized in<br />

a wide range of excitation wavelengths. Figures 2 and 3 clearly<br />

illustrate that SERS signals obtained with 633 and 785 nm<br />

excitations have good signal-to-noise ratios. Therefore, the substrates<br />

can be used in conjunction with different lasers, depending<br />

on availability and on the demands of a particular application.<br />

Multiwavelength spectroscopy can also be employed to further<br />

improve differentiation of NAADP spectral signatures using<br />

appropriate pattern recognition.<br />

SERS Detection of an Agonist-Induced Change of the<br />

NAADP Concentration in Cells. Next we studied an agonistinduced<br />

change of the NAADP concentration in breast cancer<br />

SkBr3 cells using the SERS sensor. An increase of the NAADP<br />

concentration was triggered by treating cells with three different<br />

agonists with a 5 µM concentration: ATP, acetylcholine, and<br />

histamine. The protocol for inducing NAADP concentration<br />

modulation results in a final NAADP concentration that is<br />

significantly increased as compared to its basal level. 8,10 The latter<br />

has been estimated to be on the order of tens of nanomolar. The<br />

acid extraction protocol established in ref 8 was used to obtain<br />

NAADP samples. Figure 4a shows the SERS spectrum collected<br />

from the untreated cell extracts, denoted as the control, and those<br />

of the treated cells (Figure 4b-d). Multiple spectra were acquired<br />

from each sample for further analysis. Importantly, the data<br />

collected with the SERS sensor show a good repeatability, as can<br />

be seen in Figure 4. This is expected given the fixed configuration<br />

of the gold nanoparticles on the sensor’s surface. 29,30<br />

(29) Hering, K.; Cialla, D.; Ackermann, K.; Dorfer, T.; Moller, R.; Schneidewind,<br />

H.; Mattheis, R.; Fritzsche, W.; Rosch, P.; Popp, J. Anal. Bioanal. Chem.<br />

2008, 390, 113–124.


Figure 4. SERS analysis of NAADP concentration modulation in cell extracts. Each graph contains 12 spectra collected at different locations<br />

on the SERS sensor for each sample to demonstrate the data reproducibility. (a) SERS spectrum of the untreated cells, marked as the control.<br />

(b-d) SERS spectra of cell extracts with induced NAADP release by treating cells with (b) histamine, (c) ATP, and (d) acetylcholine. All three<br />

agonists had a concentration of 5 µM. The sample volume used in this experiment was on the order of 2 µL. The data acquisition time was 10 s.<br />

A 785 nm excitation laser was used. The spectra are normalized and offset for clarity.<br />

Figure 5. (a) Concentration-dependent SERS spectra of an aqueous solution of NAADP. The bottom spectrum represents the background<br />

signal from the SERS sensor. (b) Principal component analysis of SERS data collected on cells treated with acetylcholine, an aqueous solution<br />

of 100 µM NAADP, and untreated control cells. Each point in the principal component space represents an SERS spectrum with the distance<br />

between data points proportional to the degree of similarity between the spectra. (c) Pareto chart showing the amount of information about data<br />

variability explained by the first two principal components. The results demonstrate that there is a strong correlation between the SERS spectra<br />

of cells with the modulated NAADP concentration and that of the pure 100 µM NAADP solution.<br />

In order to quantify the NAADP concentration in treated cells,<br />

we compare their SERS spectra with the reference NAADP spectra<br />

collected from the pure NAADP aqueous solution. Concentrationdependent<br />

SERS spectra of NAADP are presented in Figure 5a.<br />

(30) Zou, S. L.; Schatz, G. C. Chem. Phys. Lett. 2005, 403, 62–67.<br />

As is often the case in SERS, the spectral signature changes with<br />

concentration. 31 The 733 cm -1 peak of adenine, for example,<br />

dominates the spectrum at higher concentrations. The reduc-<br />

(31) Kim, S. K.; Joo, T. H.; Suh, S. W.; Kim, M. S. J. Raman Spectrosc. 1986,<br />

17, 381–386.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6773


tion in the relative peak intensity for lower concentrations can<br />

be caused by reorientation of NAADP molecules with respect<br />

to the gold nanoparticles’ surfaces in the SERS sensor. 31,32 The<br />

correlation between the SERS spectra of cells treated with<br />

acetylcholine and that of pure NAADP was conducted using<br />

principal component analysis (PCA). 33 Principal component analysis<br />

is a technique which minimizes the dimensionality of the<br />

analyzed data array and permits assessment of the degree of<br />

correlation between large data sets. It is one of the most widely<br />

used methods in chemometrics, and it has been demonstrated to<br />

be efficient for analyzing SERS data. 34-36<br />

According to the PCA results (Figure 5b), the control data<br />

acquired on untreated cells form a cluster which is denoted as<br />

“A”, away from the data of the treated cells, denoted as “B”. This<br />

confirms that the SERS sensor employed here distinguishes<br />

between the cells producing different amounts of NAADP.<br />

Furthermore, there is a clear correlation between the SERS<br />

spectra of the treated cells and that of the aqueous solution of<br />

100 µM NAADP. This concentration is within the range of the<br />

expected induced NAADP concentration increase, according to<br />

the protocol that was used in this work. While this concentration<br />

(32) Barhoumi, A.; Zhang, D. M.; Halas, N. J. J. Am. Chem. Soc. 2008, 130,<br />

14040–14041.<br />

(33) Jolliffe, I. T. Principal Component Analysis, 2nd ed.; Springer: New York,<br />

2002.<br />

(34) Eliasson, C.; Loren, A.; Engelbrektsson, J.; Josefson, M.; Abrahamsson, J.;<br />

Abrahamsson, K. Spectrochim. Acta, Part A 2005, 61, 755–760.<br />

(35) Pearman, W. F.; Fountain, A. W. Appl. Spectrosc. 2006, 60, 356–365.<br />

(36) Hedegaard, M.; Krafft, C.; Ditzel, H. J.; Johansen, L. E.; Hassing, S.; Popp,<br />

J. R. Anal. Chem. 2010, 82, 2797–2802.<br />

6774 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

is higher than the one which can be detected with enzymatic and<br />

radioreceptor binding assays, there is a significant advantage of<br />

time efficiency and accessibility of the SERS-based method.<br />

CONCLUSION<br />

Label-free NAADP detection and quantification in cell extracts<br />

is enabled by SERS, which permits the rapid detection of NAADP<br />

with a 100 µM concentration without any special sample purification<br />

or labeling. Importantly, this concentration does not represent<br />

a limit for SERS sensing of second calcium messengers. We were<br />

able to successfully detect 10 nM concentrations of NAADP in<br />

aqueous solution, which is on the order of basal levels of NAADP<br />

in cells, suggesting that intracellular SERS detection of the calcium<br />

messengers is possible.<br />

ACKNOWLEDGMENT<br />

This work was supported by a W. M. Keck Foundation grant<br />

to establish the W. M. Keck Institute for Attofluidic Nanotube-<br />

Based Probes at Drexel University, by the Pennsylvania Nanotechnology<br />

Institute (NTI) through Ben Franklin Technology<br />

Partners of Southeastern Pennsylvania, and by NIH Grants HL<br />

90804 and HL 90804-01A2S1 to E.B. Raman spectroscopy analysis<br />

and scanning electron microscopy were conducted at the Centralized<br />

Research Facilities (CRF) at Drexel University.<br />

Received for review March 2, 2010. Accepted July 2, 2010.<br />

AC100563T


Anal. Chem. 2010, 82, 6775–6781<br />

Highly Sensitive Fluorescent Method for the<br />

Detection of Cholesterol Aldehydes Formed by<br />

Ozone and Singlet Molecular Oxygen<br />

Fernando V. Mansano, Rafaella M. A. Kazaoka, Graziella E. Ronsein, Fernanda M. Prado,<br />

Thiago C. Genaro-Mattos, Miriam Uemi, Paolo Di Mascio, and Sayuri Miyamoto*<br />

Departamento de Bioquímica, Instituto de Química, Universidade de São Paulo,<br />

CP26077, CEP 05513-970, São Paulo, SP, Brazil<br />

Cholesterol oxidation gives rise to a mixture of oxidized<br />

products. Different types of products are generated according<br />

to the reactive species being involved. Recently,<br />

attention has been focused on two cholesterol aldehydes,<br />

3�-hydroxy-5�-hydroxy-B-norcholestane-6�-carboxyaldehyde<br />

(1a) and 3�-hydroxy-5-oxo-5,6-secocholestan-6-al<br />

(1b). These aldehydes can be generated by ozone-, as well<br />

as by singlet molecular oxygen-mediated cholesterol oxidation.<br />

It has been suggested that 1b is preferentially<br />

formed by ozone and 1a is preferentially formed by singlet<br />

molecular oxygen. In this study we describe the use of<br />

1-pyrenebutyric hydrazine (PBH) as a fluorescent probe<br />

for the detection of cholesterol aldehydes. The formation<br />

of the fluorescent adduct between 1a with PBH was<br />

confirmed by HPLC-MS/MS. The fluorescence spectra of<br />

PBH did not change upon binding to the aldehyde.<br />

Moreover, the derivatization was also effective in the<br />

absence of an acidified medium, which is critical to avoid<br />

the formation of cholesterol aldehydes through Hock<br />

cleavage of 5r-hydroperoxycholesterol. In conclusion,<br />

PBH can be used as an efficient fluorescent probe for the<br />

detection/quantification of cholesterol aldehydes in biological<br />

samples. Its analysis by HPLC coupled to a<br />

fluorescent detector provides a sensitive and specific way<br />

to quantify cholesterol aldehydes in the low femtomol<br />

range.<br />

Cholesterol (cholest-5-en-3�-ol) is a neutral lipid found in the<br />

cellular membranes of mammals. 1 Cholesterol is susceptible to<br />

oxidation mediated by enzymatic and nonenzymatic mechanisms. 2-5<br />

The nonenzymatic oxidation can be mediated by reactive oxygen<br />

species. 2 Several oxidized products of cholesterol have been characterized<br />

including hydroperoxides, epoxides, and aldehydes. 2,6,7<br />

These oxysterols have been detected in biological tissues and their<br />

formation has been associated to neurodegenerative and cardio-<br />

* To whom correspondence should be addressed. Phone: (55) (11) 3091-<br />

3810 (x261). Fax: (55) (11) 3815-5579. E-mail: miyamoto@iq.usp.br.<br />

(1) Maxfield, F. R.; Tabas, I. Nature 2005, 438, 612–621.<br />

(2) Brown, A. J.; Jessup, W. Atherosclerosis 1998, 142, 1–28.<br />

(3) Björkhem, I.; Diczfalusy, U. Arterioscler. Thromb. Vasc. Biol. 2002, 22,<br />

734–742.<br />

(4) Garenc, C.; Julien, P.; Levy, E. Free Radical Res. 2010, 44, 47–73.<br />

(5) Schroepfer, G. J., Jr. Physiol. Rev. 2000, 80, 361–554.<br />

(6) Smith, L. L. Lipids 1996, 31, 453–487.<br />

(7) Girotti, A. W. J. Photochem. Photobiol. B 1992, 13, 105–118.<br />

vascular diseases. 2-5 Recently, attention has been focused on<br />

cholesterol aldehydes that can be formed by the oxidation of<br />

cholesterol by ozone 8 and singlet molecular oxygen. 9,10<br />

The ozonation of cholesterol produces several oxidized products,<br />

in special two aldehydes (Figure 1), the cholesterol 5,6secosterols,3�-hydroxy-5�-hydroxy-B-norcholestane-6�-carboxyaldehyde<br />

(1a) and 3�-hydroxy-5-oxo-5,6-secocholestan-6-al (1b). 11,12<br />

Wentworth and co-workers showed the presence of 1a and 1b<br />

in atherosclerotic plaques 8 and LDL oxidized with several oxidants.<br />

13 These aldehydes have been also detected in neurodegenerative<br />

diseases, like Lewy body dementia 14 and Alzheimer<br />

disease. 15 The role of 1a and 1b in the pathogenesis of<br />

cardiovascular and neurodegenerative diseases has been investigated.<br />

In vitro studies have shown that cholesterol aldehydes<br />

can covalently modify proteins, as well as, accelerate their<br />

aggregation, as in the case of amyloid �-peptide formation 15-17<br />

and R-synuclein 14 . Further studies have shown that covalent<br />

modification of apo-B by 1b causes this protein to misfold,<br />

rendering the LDL particle more susceptible to macrophage<br />

uptake. 18 Moreover, some reports have also shown that<br />

cholesterol aldehydes can induce apoptosis in macrophages<br />

and cardiomyoblasts. 19,20 The induction of apoptosis in cardi-<br />

(8) Wentworth, P., Jr.; Nieva, J.; Takeuchi, C.; Galve, R.; Wentworth, A. D.;<br />

Dilley, R. B.; DeLaria, G. A.; Saven, A.; Babior, B. M.; Janda, K. D.;<br />

Eschenmoser, A.; Lerner, R. A. Science 2003, 302, 1053–1056.<br />

(9) Brinkhorst, J.; Nara, S. J.; Pratt, D. A. J. Am. Chem. Soc. 2008, 130, 12224–<br />

12225.<br />

(10) Uemi, M.; Ronsein, G. E.; Miyamoto, S.; Medeiros, M. H. G.; Di Mascio,<br />

P. Chem. Res. Toxicol. 2009, 22, 875–884.<br />

(11) Gumulka, J.; Smith, L. L. J. Am. Chem. Soc. 1983, 105, 1972–1979.<br />

(12) Jaworski, K.; Smith, L. L. J. Org. Chem. 1988, 53, 545–554.<br />

(13) Wentworth, A. D.; Song, B. D.; Nieva, J.; Shafton, A.; Tripurenani, S.;<br />

Wentworth, P. Chem. Commun. 2009, 3098–3100.<br />

(14) Bosco, D. A.; Fowler, D. M.; Zhang, Q.; Nieva, J.; Powers, E. T.; Wentworth,<br />

P.; Lerner, R. A.; Kelly, J. W. Nat. Chem. Biol. 2006, 2, 249–253.<br />

(15) Zhang, Q.; Powers, E. T.; Nieva, J.; Huff, M. E.; Dendle, M. A.; Bieschke,<br />

J.; Glabe, C. G.; Eschenmoser, A.; Wentworth, P., Jr.; Lerner, R. A.; Kelly,<br />

J. W. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 4752–4757.<br />

(16) Scheinost, J. C.; Wang, H.; Boldt, G. E.; Offer, J.; Wentworth, P. Alzheimer<br />

Dis. 2008, 47, 3919–3922.<br />

(17) Usui, K.; Hulleman, J. D.; Paulsson, J. F.; Siegel, S. J.; Powers, E. T.; Kelly,<br />

J. W. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 18563–18568.<br />

(18) Takeuchi, C.; Galve, R.; Nieva, J.; Witter, D. P.; Wentworth, A. D.; Troseth,<br />

R. P.; Lerner, R. A.; Wentworth, P. Biochemistry 2006, 45, 7162–7170.<br />

(19) Sathishkumar, K.; Gao, X.; Raghavamenon, A. C.; Parinandi, N.; Pryor, W. A.;<br />

Uppu, R. M. Free Radical Biol. Med. 2009, 47, 548–558.<br />

(20) Gao, X.; Raghavamenon, A. C.; D’Auvergne, O.; Uppu, R. M. Biochem.<br />

Biophys. Res. Commun. 2009, 389, 382–387.<br />

10.1021/ac1006427 © 2010 American <strong>Chemical</strong> Society 6775<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/21/2010


Figure 1. Structures of cholesterol carboxyaldehyde (1a), cholesterol secocholestanal (1b) and their corresponding fluorescent adducts formed<br />

upon derivatization with 1-pyrenebutyric hydrazide (2a and 2b).<br />

omyoblasts was associated with the induction of ROS generation<br />

and the activation of extrinsic and intrinsic pathway. 19<br />

Besides ozone, recent studies have evidenced the generation<br />

of cholesterol aldehydes in the reaction of cholesterol with singlet<br />

molecular oxygen. The proposed mechanisms involve Hockcleavage<br />

of cholesterol 5R-hydroperoxide (5R-OOH) in the presence<br />

of acids 9 or cholesterol dioxetane decomposition. 10 In both<br />

cases, 1a was detected as the major product. Supporting these<br />

studies, Tomono et al. 21 have also detected 1a as the major<br />

product in the incubation of cholesterol with human myeloperoxidase<br />

in the presence H2O2 and chloride ions, a system that<br />

is suggested to generate singlet molecular oxygen in activated<br />

neutrophils. 22<br />

Several methods have been used for the determination of<br />

cholesterol aldehydes, including mass spectrometry, 10 UVvisible,<br />

8,13,23 and fluorescence detection, 14,21 most of them coupled<br />

to liquid chromatography. Although mass spectrometry based<br />

methods have shown a good sensitivity, this technique requires<br />

specialized people and expensive equipment. UV-visible detection<br />

methods using 2,4-dinitrophenylhydrazine (DNPH) have been<br />

widely used to detect and quantify aldehydes derived from lipid<br />

peroxidation. However, this method is usually carried out in a<br />

strong acidic media, which is able to induce the cleavage of<br />

cholesterol 5R-OOH to form 1a and 1b, 9 therefore leading to an<br />

overestimation of the real aldehyde concentration in biological<br />

samples. An alternative method used to detect aldehydes involves<br />

their reaction with fluorescent probes and the detection of the<br />

fluorescent adducts formed. The fluorescent methods have been<br />

used to increase detection limits and exclude interferences. The<br />

fluorescent probe described in the literature for cholesterol 5,6secosterols<br />

detection is dansyl hydrazine. Bosco and co-workers<br />

used this probe in acidic media to detect cholesterol aldehydes<br />

in brain. 14<br />

(21) Tomono, S.; Miyoshi, N.; Sato, K.; Ohba, Y.; Ohshima, H. Biochem. Biophys.<br />

Res. Commun. 2009, 383, 222–227.<br />

(22) Kiryu, C.; Makiuchi, M.; Miyazaki, J.; Fujinaga, T.; Kakinuma, K. FEBS<br />

Lett. 1999, 443, 154–158.<br />

(23) Wang, K. Y.; Bermudez, E.; Pryor, W. A. Steroids 1993, 58, 225–229.<br />

6776 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

This manuscript describes a new fluorescence-based method<br />

to detect and quantify cholesterol aldehydes using the probe<br />

1-pyrenebytiric hydrazine (PBH) (Figure 1). This method proved<br />

to be highly sensitive and has the advantage of minimizing the<br />

effect of interfering compounds and not requiring acidic media.<br />

The use of acid conditions showed to cause an overestimation of<br />

1a in a sample containing cholesterol 5R-OOH. Using PBH as<br />

the fluorescent probe we could detect 1a in the low femtomolar<br />

range by HPLC coupled to fluorescence detector. Moreover, this<br />

methodology allowed detecting and calculating the ratio of 1a and<br />

1b formed by the oxidation of cholesterol exposed to singlet<br />

molecular oxygen and ozone, respectively.<br />

EXPERIMENTAL SECTION<br />

Reagents. Cholesterol, silica gel (200-400 mesh, 60 Å),<br />

1-pyrenebutyric hydrazide (PBH), deuterated chloroform (CDCl3),<br />

sodium phosphate monobasic and dibasic were purchased from<br />

Sigma (St. Louis, MO). Methylene blue was purchased from<br />

Merck (Rio de Janeiro, Brazil). All the other solvents were of<br />

HPLC grade and were acquired from Mallinckrodt Baker<br />

(Phillipsburg, NJ). The water used in the experiments was<br />

treated with the Nanopure Water System (Barnstead, Dubuque,<br />

IA).<br />

Synthesis of 3�-Hydroxy-5�-hydroxy-B-norcholestane-6�carboxyaldehyde<br />

(1a). 1a was synthesized by photooxidation<br />

of cholesterol in the presence of methylene blue as described<br />

previously. 10 Briefly, 200 mg of cholesterol dissolved in 20 mL of<br />

chloroform was mixed with 250 µL of methylene blue (10 mM in<br />

methanol). This solution was cooled at 4 °C and irradiated under<br />

continuous agitation using two tungsten lamps (500 W) during<br />

2.5 h. The formation of cholesterol oxidation products were<br />

checked by thin-layer chromatography using ethyl acetate and<br />

isooctane (1:1, v/v) as the eluent. 1a was purified from the<br />

reaction mixture by flash column chromatography, using silica<br />

gel and a gradient of hexane and ethyl eter. The purified 1a was<br />

analyzed and quantified in CDCl 3 by NMR spectroscopy using<br />

DRX500 instrument, AVANCE series (Bruker-Biospin, Rheinstetten,<br />

Germany) as described previously. 10 For the quantifica-


tion, 2 µL of proprionaldehyde were added as the internal<br />

standard into 750 µLofthe1a solution in CDCl3. The 1 H signal<br />

corresponding to the aldehyde group of the internal standard<br />

and 1a appeared at 9.73 and 9.62 ppm, respectively. The relative<br />

signals of 1 H of aldehydes were integrated by Mestre C<br />

software (Figure S1 in the Supporting Information (SI)).<br />

Synthesis of 3�-Hydroxy-5-oxo-5,6-secocholestan-6-al (1b).<br />

1b was prepared by ozonation of cholesterol as described by<br />

Wentworth et al. 8 and Wang et al. 24 Ozone was produced at a<br />

rate of 100 mg/h by passing pure oxygen through AquaZone<br />

PLUS 200 instrument (Red Sea Fish Pharm. Ltd., Houston, TX).<br />

Oxygen flow rate was set to 10 mL/min. A solution of 10 mg/mL<br />

of cholesterol in chloroform was cooled at dry ice temperature<br />

and oxidized by bubbling ozone for 5 min. After oxidation,<br />

chloroform was evaporated with nitrogen gas and the residue was<br />

stirred for 2hatroom temperature with Zn powder (19.4 mg) in<br />

water-acetic acid (1:19 v/v; 3.1 mL). Dichloromethane (15 mL)<br />

was added and the mixture was washed five times with deionized<br />

water (5 × 15 mL). The organic phase was evaporated to dryness<br />

in vacuo. The residue was dissolved in isopropyl alcohol and kept<br />

at -80 °C for further analysis. The characterization of 1b was<br />

performed by NMR spectroscopy using a DRX500 instrument,<br />

AVANCE series (Bruker-Biospin, Rheinstetten, Germany) operating<br />

at 11.7 T. Cholesterol ozonation followed by Zn reduction in<br />

acetic acid gave 1b as a major aldehyde, as confirmed by 1 H NMR<br />

analysis (δ, ppm, CDCl3): δ 9.607 (s, J ) 0.5 Hz, 1H, CHO),<br />

4.466 (s, 1H, H-3), 3.097 (dd, J ) 13.5, 4.0 Hz, 1H, H-4e), 1.010<br />

(s, 3H, CH3-19), 0.671 (s, 3H, CH3-18) (Figure S2 in the SI).<br />

Derivatization of Cholesterol Aldehydes with PBH. The<br />

optimal conditions for the derivatization of cholesterol aldehydes<br />

were established using the purified 1a. Time-course of PBH<br />

adduct formation with 1a was monitored by incubating 100 µM<br />

1a with 2 mM PBH in isopropanol at 37 °C under continuous<br />

agitation for up to 8 h. Aliquots of 1 µL of the reaction mixture<br />

were taken at specified times and analyzed by HPLC coupled to<br />

fluorescence detector. The ideal concentration of the probe was<br />

determined by incubating 1 µM of1a in the presence of 1-1000<br />

µM of PBH in isopropanol at 37 °C under continuous agitation<br />

for 6 h. The effect of the pH in the formation of 2a was analyzed<br />

by incubating 1 µM of1a with 50 µM PBH in a reaction system<br />

consisted of isopropanol:10 mM phosphate buffer at pH 5.7, 7.4,<br />

and 8.0 (90:10, vol/vol). Quantitative analysis of 1a was done by<br />

incubating an aliquot of sample with 600 µM of the probe in<br />

isopropanol containing 1 mM phosphate buffer at pH 7.4 (neutral<br />

condition) or 0.1 mM HCl (acid condition) at 37 °C for6h.<br />

Analysis of the Fluorescent Adduct by HPLC Coupled<br />

with Fluorescence Detector. HPLC analysis was carried out on<br />

a Shimadzu Prominence system (Tokyo, Japan), consisted of LC-<br />

20AT pumps, SIL-20AV autosampler, RF-10Axl fluorescence detector,<br />

and a CBM-20A controller. One microliter of the sample was<br />

injected into a reversed-phase column Synergi C18 (50 × 4.6 mm,<br />

2.5 µm particle size, Phenomenex, Torrance, CA). The HPLC<br />

mobile phase consisted of water (A) and methanol (B), and the<br />

flow rate was 1 mL/min. The separation of the fluorescent adducts<br />

was done using the following condition: 86% B for 5 min, 86-92%<br />

in 1 min, 92% B for 15 min, and 92-86% B in 1 min. The excitation<br />

and emission wavelengths were fixed at 339 and 380 nm,<br />

respectively. 25 For the analysis 1 µL was injected through the<br />

autosampler. Data were acquired at high sensitivity and gain of<br />

1× in the fluorescence detector. Fluorescence spectra were<br />

acquired by the fluorescence detector RF-551 (Shimadzu, Japan).<br />

HPLC data were processed using the LC solution software<br />

(Shimadzu, Japan).<br />

Analysis of the Fluorescent Adducts by HPLC-MS/MS.<br />

The PBH fluorescent adducts formed with 1a and 1b were<br />

analyzed by a HPLC system connected to a UV-visible detector<br />

(SPD 10 AVVP, Shimadzu, Kyoto, Japan) and a Quattro II triple<br />

quadrupole tandem mass spectrometer (Micromass, Manchester,<br />

UK) (HPLC-MS/MS). HPLC separation was carried out as<br />

described above. The eluent from the column was monitored at<br />

342 nm and 10% of the HPLC flow rate was directed into the mass<br />

spectrometer. The fluorescent adducts were detected using<br />

electrospray ionization (ESI) in the positive ion mode. The source<br />

and desolvation temperature of the mass spectrometer were set<br />

at 100 and 200 °C, respectively. The cone voltage was set to 50 V,<br />

the extractor cone voltage was set to 5 V and the capillary and<br />

the electrode potentials were set at 4.5 and 0.5 kV, respectively.<br />

Collision energy was set at 20 eV for 2a and 30 eV for 2b. Fullscan<br />

data was acquired over a mass range of 100-900 m/z. Data<br />

was processed by means of the MassLynx NT software.<br />

Cholesterol Oxidation Induced by Photooxidation, Ozone<br />

and HOCl/H 2O2. Cholesterol (10 mg/mL in chloroform) was<br />

photooxidized in the presence of methylene blue for 2.5 h. An<br />

aliquot of this sample was taken for cholesterol aldehyde determination.<br />

The ozonation of cholesterol (10 mg/mL in chloroform)<br />

was conducted essentially as described for the synthesis of 1b.<br />

After 5 min ozonation, chloroform was evaporated by nitrogen<br />

gas and the residue was dissolved in isopropanol. An aliquot of<br />

this sample was diluted and used for cholesterol aldehyde<br />

determination. The oxidation of cholesterol by HOCl/H 2O2<br />

system was conducted by incubating a 10 mM cholesterol<br />

solution containing 1% ethanol in 10 mM phosphate buffer at<br />

pH 7.4 with 1 mM HOCl and H2O2, under continuous agitation<br />

at 37 °C for 5 min.<br />

Method Validation. The limit of quantification (LOQ) was<br />

established as the amount of adduct formed that generated a signal<br />

6-fold higher than baseline. The limit of detection (LOD) was<br />

established as the amount of adduct formed that generated a signal<br />

3-fold higher than baseline. The reproducibility was checked by<br />

intra- and interday analysis. Interday analyses were conducted on<br />

three consecutive days.<br />

RESULTS<br />

Synthesis and Characterization of 1a. Cholesterol carboxyaldehyde<br />

(1a) was synthesized by photooxidation of cholesterol. 10<br />

After the reaction, the oxidized products were purified by flash<br />

column chromatography and 1a was isolated. The identity of 1a<br />

was confirmed by NMR spectroscopy as described by Uemi et<br />

al. 10 The 1 H NMR analysis showed a doublet peak at 9.62 ppm<br />

consistent with the presence of the aldehyde group in 1a<br />

structure (Figure S1 in the SI). 1a was quantified by 1 H NMR<br />

analysis using proprionaldehyde as the internal standard as<br />

described in the experimental section.<br />

Characterization of 2a Fluorescent Adduct. Using the<br />

purified 1a sample, several experiments were conducted to<br />

(24) Wang, K. Y.; Bermudez, E.; Pryor, W. A. Steroids 1993, 58, 225–229. (25) Morsel, J. T.; Schmiedl, D. Fresenius´ J. Anal. Chem. 1994, 349, 538–541.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6777


Figure 2. Analysis of 2a fluorescent adduct, formed in the reaction<br />

of 1a with PBH by HPLC coupled to fluorescence detector. HPLC/<br />

fluorescence chromatogram obtained for the analysis of 2a using<br />

excitation at 339 nm and emission at 380 nm. For the derivatization,<br />

5 pmol of 1a was incubated with PBH 100 µM in a final volume of<br />

100 µL isopropanol at 37 °C for 6 h under agitation. Analysis was<br />

done by injecting 1 µL of the reaction mixture into the HPLC. The<br />

inset shows the time dependent formation of 2a during incubation of<br />

100 µM of1a in the presence of 2 mM of PBH for up to 8 h.<br />

establish the optimal conditions for the reaction with PBH.<br />

Incubations of 1a with PBH were conducted in isopropanol at 37<br />

°C under continuous agitation. HPLC analysis using fluorescence<br />

detection showed the appearance of an intense peak at 12.5 min,<br />

corresponding to the fluorescent adduct 2a (Figure 2). The pH<br />

effect on 2a adduct formation was analyzed. Incubations conducted<br />

at pH 5.7, 7.4, and 8.5 showed that this adduct was<br />

preferentially formed at pH 5.7 (Figure S3 in the SI), which is in<br />

accordance with favorable formation of Schiff base adduct at<br />

slightly acidic conditions.<br />

A time-course analysis of 2a showed a time-dependent increase<br />

of the peak area up to 4 h, reaching a plateau after this time (inset,<br />

Figure 2). Based on this, the incubation time for the derivatization<br />

was fixed to 6 h. The ideal concentration of PBH for the reaction<br />

was established by incubating 1a (1 µM) in the presence of<br />

1-1000 µM of PBH. The fluorescent adduct formation reached<br />

its maximum with PBH concentration higher than 50 µM.<br />

The fluorescence of 2a was characterized to establish the<br />

optimal excitation and emission wavelengths. Fluorescence analysis<br />

of 2a showed a spectrum similar to the original probe with<br />

excitation and emission wavelengths maximum at 339 and 380<br />

nm, respectively (Figure 3). Thus, indicating that adduct formation<br />

does not alter the fluorescence spectra of the probe.<br />

The formation of 2a was also confirmed by HPLC-MS/MS.<br />

The mass spectrum of 2a acquired by ESI in the positive ion mode<br />

showed peaks corresponding to 2a molecular ion ([M+H] + ) and<br />

its sodium adducts ([M+Na] + )atm/z 703 and 725, respectively<br />

(Figure 4A and 4B). Two other peaks at m/z 685 and 667 were<br />

also detected, corresponding to the loss of one ([M+H-H2O] + )<br />

and two water molecules ([M+H-2H2O] + ), respectively. These<br />

ions were also observed in the MS/MS spectrum of m/z 703<br />

(Figure 4C). The MS/MS spectrum also showed other fragment<br />

ions at m/z 431, 398, 365, and 303. The first three, correspond to<br />

the ions formed by the loss of pyrene butyric (PB) group<br />

6778 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 3. Fluorescence excitation and emission spectra of PBH and<br />

2a. Both compounds showed the same excitation and emission<br />

maximum wavelengths at 339 and 380 nm, respectively.<br />

([M+H-PB] + , m/z 431), PBH and two protons ([M+H-PBH-<br />

2H] + , m/z 398) and PBH and two water molecules<br />

([M+H-PBH-2H2O] + , m/z 365), respectively. The fragment<br />

ion at m/z 303 corresponds to the positively charged PBH ion.<br />

Reproducibility and Stability. The intra- and interday reproducibility<br />

of the assay procedure was determined by evaluating<br />

three individual reactions containing 1a (1 pmol; 1 µM final<br />

concentration) and PBH (50 µM). The relative standard deviation<br />

for the intra- and interday reproducibility at this concentration was<br />

1.7% and 7.6%, respectively.<br />

The stability of 2a adduct was evaluated in triplicate reactions<br />

of 1a (1 pmol; 1 µM final concentration) with PBH (50 µM). After<br />

derivatization, these samples were kept in vials inside the<br />

autosampler at 4 °C and the 2a peak was monitored up to 6 h.<br />

There was no significant change in the peak area of the<br />

fluorescence signal during this period.<br />

Standard Curve, Limit of Detection and Limit of Quantification.<br />

A calibration curve was constructed using 5-100 fmol<br />

(50-1000 nM) of 1a in triplicate reaction in a final volume of 100<br />

µL. Peak area of 2a increased linearly over this concentration<br />

range with R 2 value of 0.9962 and Y ) 6.64 × 10 3 X + 1.65 ×<br />

10 4 . The limit of detection (LOD) and quantification (LOQ) of<br />

1a was 10 fmol (10 nM) and 20 fmol (20 nM), respectively.<br />

Application of PBH for Cholesterol Aldehyde Detection<br />

in Oxidized Cholesterol Samples. The newly developed fluorescence-based<br />

method was applied for the detection and quantification<br />

of 1a in cholesterol samples oxidized by photooxidation<br />

or ozone. Cholesterol photooxidation gave rise to a major peak at<br />

12.5 min (Figure 5A). On the other hand, cholesterol ozonation<br />

yielded an intense peak at 10.3 min as well as some other smaller<br />

peaks, including the same peak at 12.5 min observed for the<br />

photooxidation (Figure 5B). Peak assignment was done by<br />

comparison of the retention times of the peaks with those obtained<br />

for the standard samples of 1a (Figure 5C) and 1b (Figure 5D).<br />

In this way, the peaks observed at 10.3 and 12.5 min were assigned<br />

to the fluorescent adducts 2b and 2a, respectively. Additionally,<br />

the identity of the peaks was confirmed by HPLC-MS/MS analysis<br />

(Figure S4 in the SI). It should be noted that the overall retention<br />

times for the analysis by HPLC-MS/MS were increased by almost<br />

2 min compared to the analysis by HPLC/fluorescence detection.<br />

This retention time delay was due to the differences in room


Figure 4. Analysis of 2a by HPLC-MS/MS using ESI in the positive ion mode. The cone voltage and the collision energy were set to 50 V and<br />

20 eV, respectively. Selected ion chromatogram of the ion at m/z 703 (A). Spectrum of the peak corresponding to 2a at 14 min (B). Fragment<br />

ion spectrum of m/z 703 (C).<br />

Figure 5. Analysis of cholesterol aldehydes formed in the photooxidation<br />

(A) and ozonation (B) of cholesterol using PBH. An aliquot<br />

of the sample was reacted with PBH (600 µM) for 6 h and analyzed<br />

by HPLC coupled to fluorescence detector. For peak assignment<br />

standard samples containing 1a (C) or 1b (D) were also reacted with<br />

PBH and analyzed.<br />

temperatures, which was about 5 °C lower in the case of HPLC-<br />

MS/MS analysis.<br />

The major fluorescent adduct detected in cholesterol ozonation<br />

was further characterized by HPLC-MS/MS analysis (Figure 6).<br />

Selected ion mass chromatogram of the ion at m/z 703 showed a<br />

single peak at 12 min. (Figure 6A). Mass spectrum for this peak<br />

showed two major ions, one at m/z 703 and other at m/z 725<br />

(Figure 6B), which corresponds to the molecular ion ([M+H] + )<br />

and the sodium adduct ([M+Na] + )of2b (Figure 6B). Collision<br />

induced dissociation of the ion at m/z 703, showed the same<br />

fragment ions observed for 2a, suggesting that the peak corresponds<br />

to its isomer, 2b (Figure 6C). Fragments ions observed<br />

at m/z 685 and 667 are formed by the loss of one and two water<br />

molecules from 2b, respectively. The ions at m/z 431, 398, 383,<br />

and 365 are formed by the loss of PB, PBH and two protons, PBH<br />

and one water, and PBH and two water molecules from 2b,<br />

respectively (Figure 6C).<br />

Aiming to get a quantitative data on cholesterol aldehyde<br />

formation, a calibration curve constructed for 1a was used to<br />

determine its concentration. For the quantification, oxidized<br />

cholesterol samples were diluted to get a final cholesterol<br />

concentration of 219 µM. The amounts of 1a detected in the<br />

samples by PBH method at neutral pH were 19.1 ± 4.0 µM and<br />

0.7 ± 0.3 µM for the photooxidation and ozonation, respectively.<br />

This corresponds to a 1a yield of 8.7% for the photooxidation and<br />

0.3% for the ozonation. For comparison, 1a quantification was also<br />

done using dansyl hydrazine method in the presence of acid (0.1<br />

mM HCl) (method in the SI). The concentrations of 1a determined<br />

by this method were 35.2 ± 0.7 µM and 1.0 ± 0.1 µMinthe<br />

photooxidation and ozonation, respectively. This provides 1a<br />

yields of 16% for the photooxidation and 0.4% for the ozonation.<br />

As can be clearly noticed, the 1a yield determined by the dansyl<br />

hydrazine method was overestimated by almost 2-folds for the<br />

photooxidation. The same result was obtained when PBH derivatization<br />

was conducted in the presence of acid. Under this<br />

condition, the 1a concentration was 32.6 ± 0.7 µM, which<br />

corresponds to a 1a yield of 15% in the photooxidation. These<br />

results are consistent with the fact that photooxidized cholesterol<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6779


Figure 6. Analysis of 2b by HPLC-MS/MS using ESI in the positive ion mode. The cone voltage was set to 50 V and the collision energy was<br />

set to 30 eV. Selected ion chromatogram for the ion at m/z 703 (A). Mass spectrum of the peak corresponding to 2b at 12 min (B). Fragment<br />

ion spectrum of m/z 703 (C).<br />

samples contain cholesterol 5R-OOH, which in the presence of<br />

acids is easily converted to 1a. 9<br />

Moreover, the sensitivity of PBH and dansyl hydrazine<br />

methods were also compared (Figure S5 in the SI). PBH method<br />

conducted at neutral pH was almost 10 times more sensitive than<br />

dansyl hydrazine method. This difference was even larger when<br />

PBH reaction was carried out with acid, reaching a 90 times higher<br />

sensitivity.<br />

PBH was also used for the quantification of ChAld formed<br />

during cholesterol oxidation promoted by the HOCl/H 2O2 reaction<br />

system. This is a biologically relevant reaction system that<br />

is known to generate stoichiometric amounts of singlet molecular<br />

oxygen. The formation of 2a was analyzed by HPLC<br />

coupled with fluorescence detector (Figure 7). The complete<br />

reaction system containing cholesterol and HOCl/H2O2 showed<br />

the appearance of an intense fluorescent peak corresponding<br />

to 2a. The fluorescent adduct was quantified using the<br />

calibration curve and a value of 0.2 µM of1a was found. This<br />

corresponds to a yield of 0.2%.<br />

DISCUSSION<br />

Cholesterol is oxidized in the presence of reactive oxygen<br />

species generating aldehydes, hydroperoxides and epoxides. 6,7<br />

Recently, two cholesterol aldehydes have attracted attention, the<br />

cholesterol carboxyaldehyde (1a) and cholesterol secoaldehyde<br />

(1b). The formation of 1a and 1b was first described in the<br />

oxidation of cholesterol with ozone. 11,12 Wentworth and co-workers<br />

reported the presence of 1a and 1b in atherosclerotic plaques<br />

and related the detection of these aldehydes as an evidence for<br />

ozone generation in human tissues. 8 On the other hand, two recent<br />

studies identified the generation of 1a in the reaction of cholesterol<br />

with singlet molecular oxygen. 9,10 Indeed, Brinkhorst and<br />

6780 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 7. Analysis of cholesterol aldehydes generated by the<br />

reaction of cholesterol with H2O2 and HOCl using PBH. HPLC/<br />

fluorescence chromatograms obtained for cholesterol (A), and cholesterol<br />

incubated with H2O2 (B), HOCl (C), and H2O2 + HOCl (D).<br />

Reactions were carried out with 100 µM cholesterol in the presence<br />

of 1 mM HOCl, 1 mM H2O2 and 100 µM PBHfor6hat37°C under<br />

continuous agitation.<br />

co-workers reported the formation of 1a from cholesterol 5Rhydroperoxide<br />

by Hock cleavage in acid media 9 and Uemi and<br />

co-workers reported the formation of 1a in the reaction of<br />

cholesterol with singlet molecular oxygen generated either by<br />

photooxidation or by the thermodecomposition of endoperoxides.<br />

10<br />

In this study, we have developed a new detection method for<br />

cholesterol carboxyaldehyde, 1a, using the fluorescent probe<br />

PBH, which contains the highly fluorescent pyrene group. The


eaction of 1a with PBH can be conducted in the absence of acids,<br />

which is important to avoid the formation of 1a from acid catalyzed<br />

cholesterol 5R-OOH decomposition. 9 Bosco and co-workers 14 used<br />

dansyl hydrazine as a fluorescent label to detected 1a and 1b in<br />

a brain tissue sample from Lewy body disease. However, the<br />

derivatization of cholesterol aldehydes with dansyl hydrazine is<br />

conducted in the presence of strong acids, such as, 5 mM sulfuric<br />

acid. 14 In fact, we have done a comparative analysis of the two<br />

fluorescent probes (PBH and dansyl hydrazine) to detect 1a in<br />

vitro and observed that the method using dansyl hydrazine can<br />

overestimate the basal level of 1a, especially in samples containing<br />

cholesterol 5R-OOH. The 1a yields in photooxidized cholesterol<br />

samples estimated by PBH (neutral pH) and dansyl hydrazine<br />

were 8.7% and 16.0%, respectively. This data indicates that the<br />

derivatization in the absence of acid is a critical point that should<br />

be considered when estimating 1a basal level in samples,<br />

especially when they were oxidized by singlet molecular oxygen.<br />

When comparing dansyl hydrazine and PBH methods, the latter<br />

was approximately 10 times more sensitive for 1a detection<br />

(Figure S5 in the SI). This difference was even larger when PBH<br />

derivatization was carried out at acid condition, clearly evidencing<br />

the superiority of PBH method compared to dansyl hydrazine in<br />

terms of sensitivity.<br />

A number of studies has detected the formation of cholesterol<br />

aldehydes in vitro, 9,10,21 as well as in vivo. 8,14-16 In this study, we<br />

have used PBH to analyze the formation of 1a during cholesterol<br />

oxidation mediated by ozone and singlet molecular oxygen. The<br />

reaction of cholesterol with ozone yielded 1a and 1b and also<br />

other minor unidentified products. We could not quantify exactly<br />

the amount of 1b due to the lack of an appropriate pure standard<br />

for it. Nonetheless, the relative peak intensities suggest that<br />

cholesterol ozonation yields 1b as a major product and a small<br />

amount of its aldolization product (Figure 5). The estimated ratio<br />

of 1a:1b in the ozonation was 1:10. Similar results were described<br />

for the ozonation of human LDL, where a relative yield of 1:4 was<br />

found. 13 On the other hand, cholesterol oxidation promoted by<br />

singlet molecular oxygen yielded 1a as the major product<br />

consistent with the data described by Brinkhorst et al. 9 and Uemi<br />

et al. 10 The estimated yield of 1a in the photooxidation of<br />

cholesterol was approximately 8.7% and the ratio of 1a:1b was<br />

165:1. These values are in agreement with those reported for the<br />

photooxidation of human LDL using hematoporphyrin IX for 14 h<br />

where only 1a was detected. 13 The incubation of cholesterol in<br />

(26) Khan, A. U.; Kasha, M. J. Am. Chem. Soc. 1970, 92, 3293–3300.<br />

the presence of H2O2 and HOCl also gave 1a as the major<br />

product, consistent with the oxidation of cholesterol promoted<br />

by singlet molecular oxygen. It is known that the reaction of<br />

H2O2 with HOCl generate stoichiometric amounts of singlet<br />

molecular oxygen 26 and this reaction is suggested to play an<br />

important role during inflammatory conditions. 22 The predominant<br />

formation of 1a over 1b by this reaction was also reported<br />

by Tomono and co-workers. 21 They incubated cholesterol (100<br />

µM) in the presence of H2O2 (100 µM) and NaOCl (100 µM)<br />

and detected five times more 1a than 1b using dansyl<br />

hydrazine as a fluorescent probe. 21<br />

CONCLUSIONS<br />

We have developed a new sensitive method for the detection<br />

of cholesterol aldehydes using PBH as a fluorescent probe. This<br />

new methodology allows detecting and quantifying the relative<br />

amounts of 1a and 1b with high sensitivity, which is important<br />

to assess the relative contributions of ozone and singlet molecular<br />

oxygen to the oxidation of cholesterol in biological systems.<br />

Derivatization of cholesterol aldehydes using PBH can be conducted<br />

without the addition of strong acids, which is important<br />

for the determination of the basal level of cholesterol aldehydes<br />

present in biological tissues from different sources. In conclusion,<br />

the easy and reliable method developed in this study can help to<br />

investigate the formation cholesterol aldehydes in biological<br />

system, which is critical to clarify their relevance in disease<br />

progression.<br />

ACKNOWLEDGMENT<br />

This work was supported by the Brazilian research funding<br />

institutions FAPESP (Fundação de Amparo à Pesquisa do Estado<br />

de São Paulo), CNPq (Conselho Nacional para o Desenvolvimento<br />

Científico e Tecnológico), INCT de Processos Redox em Biomedicina<br />

- Redoxoma, and Pró-Reitoria de Pesquisa-USP.<br />

SUPPORTING INFORMATION AVAILABLE<br />

NMR analysis of 1a and 1b. HPLC-MS/MS detection of 2a<br />

and 2b. HPLC-fluorescence data showing the pH effect on 2a<br />

formation and the sensitivity of dansyl hydrazine and PBH<br />

methods. Dansyl hydrazine derivatization method. This material<br />

is available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review March 10, 2010. Accepted July 8,<br />

2010.<br />

AC1006427<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6781


Anal. Chem. 2010, 82, 6782–6789<br />

Magnetic Nanoparticle Enhanced Surface Plasmon<br />

Resonance Sensing and Its Application for the<br />

Ultrasensitive Detection of Magnetic<br />

Nanoparticle-Enriched Small Molecules<br />

Jianlong Wang, Ahsan Munir, Zanzan Zhu, and H. Susan Zhou*<br />

Department of <strong>Chemical</strong> Engineering, Worcester Polytechnic Institute, 100 Institute Road, Worcester,<br />

Massachusetts 01609<br />

Magnetic nanoparticles (MNPs) have been frequently<br />

used in bioseparation, but their applicability in bioassays<br />

is limited due to their extremely small size so that<br />

sensitive detection is difficult to achieve using a general<br />

technique. Here, we present an amplification technique<br />

using MNPs for an enhanced surface plasmon resonance<br />

(SPR) bioassay. The amplification effect of carboxyl group<br />

modified Fe3O4 MNPs of two sizes on SPR spectroscopy<br />

is first demonstrated by assembling MNPs on amino<br />

group modified SPR gold substrate. To further evaluate<br />

the feasibility of the use of Fe3O4 MNPs in enhancing<br />

a SPR bioassay, a novel SPR sensor based on an<br />

indirect competitive inhibition assay (ICIA) is developed<br />

for detecting adenosine by employing Fe3O4<br />

MNP-antiadenosine aptamer conjugates as the amplification<br />

reagent. The results confirm that Fe3O4<br />

MNPs can be used as a powerful amplification agent<br />

to provide a sensitive approach to detect adenosine by<br />

SPR within the range of 10-10 000 nM, which is<br />

much superior to the detection result obtained by a<br />

general SPR sensor. Importantly, the present detection<br />

methodology could be easily extended to detect other<br />

biomolecules of interest by changing the corresponding<br />

aptamer in Fe3O4 MNP-aptamer conjugates. This<br />

novel technique not only explores the possibility of the<br />

use of SPR spectroscopy in a highly sensitive detection<br />

of an MNP-based separation product but also offers a<br />

new direction in the use of Fe3O4 MNPs as an amplification<br />

agent to design high performance SPR biosensors.<br />

Over the past few decades, magnetic nanoparticles (MNPs)<br />

have been receiving increasing attention due to their unprecedented<br />

advantages such as higher surface-to-volume ratio for<br />

chemical binding, minimum disturbance to attached biomolecules,<br />

faster binding rates, higher miscibility, and higher specificity. 1,2<br />

These characteristics of MNPs render them easier labeling by<br />

biomolecules, as well as easier binding with its target analytes.<br />

* Corresponding author. Tel.: 508-831-5275. Fax: 508-831-5936. E-mail:<br />

szhou@wpi.edu.<br />

(1) Pankhurst, Q. A.; Connoliy, J.; Johns, S. K.; Dobson, J. J. Appl. Phys. 2003,<br />

36, 167–181.<br />

(2) Gijs, M. A. M. Microfluid. Nanofluid. 2004, 1, 22–40.<br />

6782 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Up to now, all kinds of biomolecules including DNA and RNA, 3<br />

protein and peptides, 4 cell and virus 5-7 have been separated and<br />

concentrated under an external magnet by the use of MNPs as<br />

carriers. Compared with the springing up of MNPs in bioseparation,<br />

the application of MNPs in bioassays is limited because their<br />

size is too small to be detected by general techniques developed<br />

for detecting magnetic microbeads. In order to extend the<br />

application of MNPs in bioassays, some novel techniques had been<br />

used to detect MNPs, such as the electrochemical method, 8 IR<br />

spectroscopy, 9 fluorescence spectroscopy, 10 magnetic atomic force<br />

microscopy (AFM), 11 magnetic resonance imaging (MRI), 12<br />

bio-bar-code, 13 etc. However, these methods either need labeling<br />

MNPs by electroactive probes or fluorescence molecules or need<br />

expensive experiment setups, thereby limiting them to be used<br />

on a benchtop scale and cannot be used for simple, in situ, and<br />

cost-effective detection of real samples.<br />

Here, we investigate the application of surface plasmon<br />

resonance (SPR) spectroscopy for fast, ultrasensitive, and in situ<br />

detection of the MNP-enriched biomolecules. SPR being a surfacesensitive<br />

characterization method not only can be used for<br />

analyzing the kinetic data including the equilibrium constant and<br />

the association and dissociation parameters between biomolecules<br />

by simulating SPR kinetic curves but also can be used in situ to<br />

detect the concentrations of biomolecules with high sensitivity<br />

and selectivity. 14,15 The surface plasmon used in SPR spectroscopy<br />

is highly sensitive to changes in the effective refractive index or<br />

the thickness of the test medium in the vicinity of the metal<br />

surface, especially for the molecules with high mass change.<br />

(3) Obata, K.; Tajima, H.; Yohda, M.; Matsunaga, T. Pharmacogenomics 2002,<br />

3, 697–708.<br />

(4) Safarik, I.; Safarikova, M. BioMag. Res. Technol. 2004, 2, 7.<br />

(5) Pamme, N. Lab Chip 2006, 6, 24–38.<br />

(6) Zakhireh, J.; Gomez, R.; Esserman, L. Eur. J. Cancer 2008, 44, 2742–2752.<br />

(7) Safarik, I.; Safarikova, M. J. Chromatogr., B 1999, 722, 33–53.<br />

(8) Hsing, I. M.; Xu, Y.; Zhao, W. T. Electroanalysis 2007, 19, 755–768.<br />

(9) Ravindranath, S. P.; Mauer, L.; DebRoy, C.; Irudayaraj, J. Anal. Chem. 2009,<br />

81, 2840–2846.<br />

(10) Song, Y. J.; Zhao, C.; Ren, J. S.; Qu, X. G. Chem. Commun. 2009, 1975–<br />

1977.<br />

(11) Arakaki, A.; Hideshima, S.; Nakagawa, T.; Niwa, D.; Tanaka, T.; Matsunaga,<br />

T. Biotechnol. Bioeng. 2004, 88, 543–546.<br />

(12) Perez, J. M.; Josephson, L.; O’Loughlin, T.; Hogemann, D.; Weissleder, R.<br />

Nat. Biotechnol. 2002, 20, 816–820.<br />

(13) Li, Y.; Hong Cu, Y. T.; Luo, D. Nat. Biotechnol. 2005, 23, 885–889.<br />

(14) Li, X.; Wei, X. L.; Husson, S. M. Biomacromolecules 2004, 5, 869–876.<br />

(15) Li, X.; Husson, S. M. Biosens. Bioelectron. 2006, 22, 336–348.<br />

10.1021/ac100812c © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/20/2010


Numerous references had demonstrated nanoparticles (NPs)<br />

could greatly enhance the sensitivity of SPR spectroscopy due to<br />

the large molecular weight of nanoparticles in spite of the fact<br />

that the size of nanoparticles is so small that it could not be<br />

observed by other techniques. Several kinds of NPs including Au<br />

NPs, 16-20 SiO2 NPs, 21 Pd NPs, 22 and Pt NPs 23 had been applied<br />

to increase the SPR sensitivity for detecting all kinds of<br />

biomolecules. However, the application of MNPs in the SPR<br />

field is still limited. Considering the high refractive index and<br />

the high molecular weight of MNPs, 24 it is possible to design<br />

an excellent SPR biosensor using MNPs as an amplification<br />

reagent. Once the amplifying effect of MNPs for a SPR signal<br />

is demonstrated, it can then be proved that SPR will be a<br />

powerful candidate for detecting MNP-based separation products.<br />

There are very few works that have been done to study the<br />

SPR response of MNPs, and most of them focus on utilizing<br />

commercial strepavidin-conjugated MNPs for signal amplification.<br />

25,26 Obviously, biotin needs to be attached on a SPR substrate<br />

surface for the further binding of strepavidin-conjugated MNPs,<br />

which limits the extensive application of MNPs in the SPR field.<br />

To further understand the SPR response of MNPs and extend<br />

the application of SPR in detecting MNP labeled biomolecules<br />

and their separation product, in this work, we study the SPR<br />

response of the carboxyl group modified Fe3O4 MNPs by<br />

nonspecifically adsorbing the Fe3O4 MNPs on amino group<br />

modified SPR gold substrate. The carboxyl groups on Fe3O4<br />

MNPs allow the MNPs to be easily functionalized by all kinds<br />

of biomolecules for extensive applications. Our results demonstrate<br />

that the monolayer adsorption of Fe3O4 MNPs could<br />

result in a big SPR angle shift with a low optical loss. On the<br />

basis of the amplification effect of Fe3O4 MNPs, we further<br />

demonstrate SPR spectroscopy can be used to sensitively detect<br />

Fe3O4 MNP-enriched small molecules by an indirect competitive<br />

inhibition assay (ICIA). In this case, Fe3O4 MNPs labeled<br />

by antiadenosine aptamer are used both as the enrichment<br />

reagent of adenosine and the amplification reagent of SPR<br />

spectroscopy.<br />

EXPERIMENTAL SECTION<br />

Materials. Adenosine, uridine, cytidine, guanosine, ethanolamine,<br />

6-mercaptohexan-1-ol (MCH), 1-ethyl-3-(3-dimethylaminopropyl)carbodiimide<br />

hydrochloride (EDC), N-hydroxysuccinimide<br />

(NHS), thrombin, FeO(OH), oleic acid, 1-octadecene, acetone,<br />

(16) Golub, E.; Pelossof, G.; Freeman, R.; Zhang, H.; Willner, I. Anal. Chem.<br />

2009, 81, 9291–9298.<br />

(17) Riskin, M.; Tel-Vered, R.; Lioubashevski, O.; Willner, I. J. Am. Chem. Soc.<br />

2009, 131, 7368–7378.<br />

(18) Lioubashevski, O.; Chegel, V. I.; Patolsky, F.; Katz, E.; Willner, I. J. Am.<br />

Chem. Soc. 2004, 126, 7133–7143.<br />

(19) Zayats, M.; Pogorelova, S. P.; Kharitonov, A. B.; Lioubashevski, O.; Katz,<br />

E.; Willner, I. Chem.sEur. J. 2003, 9, 6108–6114.<br />

(20) Wang, J. L.; Munir, A.; Zhou, H. S. Talanta 2009, 79, 72–76.<br />

(21) Luckarift, H. R.; Balasubramanian, S.; Paliwal, S.; Johnson, G. R.; Simonian,<br />

A. L. Colloids Surf., B 2007, 58, 28–33.<br />

(22) Lin, K. Q.; Lu, Y. H.; Chen, J. X.; Zheng, R. S.; Wang, P.; Ming, H. Opt.<br />

Express 2008, 16, 18599–18604.<br />

(23) Beccati, D.; Halkes, K. M.; Batema, G. D.; Guillena, G.; de Souza, A. C.;<br />

van Koten, G.; Kamerling, J. P. ChemBioChem 2005, 6, 1196–1203.<br />

(24) Grigoriev, D.; Gorin, D.; Sukhorukov, G. B.; Yashchenok, A.; Maltseva, E.;<br />

Moehwald, H. Langmuir 2007, 23, 12388–12396.<br />

(25) Teramura, Y.; Arima, Y.; Iwata, H. Anal. Biochem. 2006, 357, 208–215.<br />

(26) Soelberg, S. D.; Stevens, R. C.; Limaye, A. P.; Furlong, C. E. Anal. Chem.<br />

2009, 81, 2357–2363.<br />

chloroform, poly(maleic anhydride-alt-1-octadecene) (molecular<br />

weight: 30 000-50 000) and 2-(2-aminoethoxy)-ethanol were purchased<br />

from Sigma and used as received. Sodium hydrogen<br />

phosphate heptahydrate, potassium dihydrogen phosphate, and<br />

sodium chloride were ordered from Alfa Aesar. All DNA molecules<br />

were obtained from Integrated DNA Technologies (IDT). The<br />

sequence of the adenosine-binding aptamer was 5′-NH 2-C6-AGA<br />

GAA CCT GGG GGA GTA TTG CGG AGG AAG GT-3′<br />

(aptamer), the sequence of its partial complementary strand<br />

was 5′-SH-C6-ACC TTC CTC CGC-3′ (ss-DNA). DNA solutions<br />

were prepared by dissolving DNA in 50 mM, pH 8.0 Tris-HCl<br />

buffer including 138 mM NaCl. Different concentrations of<br />

adenosine and 1 mM uridine, cytidine, and guanosine were all<br />

prepared in the Tris-HCl buffer. All glassware used in the<br />

experiment was cleaned in a bath of freshly prepared 3:1 HCl/<br />

HNO3 (aqua regia) and rinsed thoroughly in H2O prior to use.<br />

(Caution: Aqua regia solution is dangerous and should be<br />

handled with care.)<br />

Synthesis of Monodisperse Fe3O4 MNPs. Monodisperse<br />

Fe3O4 MNPs were synthesized by the pyrolysis of iron carboxylate<br />

in the organic phase. 27 In brief, a mixture of FeO(OH),<br />

oleic acid, and 1-octadecene was refluxed at 320 °C for1h<br />

under a nitrogen atmosphere. During this process, the solution<br />

changed its color from turbid black to black. The resulting<br />

MNPs were precipitated with acetone and collected by centrifuge<br />

at 4000g. After that, Fe3O4 MNPs were further purified<br />

by repeated extraction of the precipitate with CHCl3/acetone<br />

(1:10) until a powder of Fe3O4 MNPs was obtained. The powder<br />

of Fe3O4 MNPs was stored at room temperature for further<br />

application.<br />

Forming Soluble Fe3O4 MNPs by Phase Transfer. Fe3O4<br />

MNPs were transferred to a PBS solution according to Yu’s<br />

work with minor modifications. 28 Carboxy group modified<br />

amphiphilic polymers was first prepared by mixing poly(maleic<br />

anhydride-alt-1-octadecene) with 2-(2-aminoethoxy)-ethanol (molar<br />

ratio 1:120) in chloroform overnight. Then, the monodisperse<br />

Fe3O4 MNPs (purified and dispersed in chloroform) were<br />

dispersed in the carboxy group modified amphiphilic polymer<br />

solution, and the mixture was stirred overnight at room<br />

temperature (molar ratio of Fe3O4/polymer was 1:10). After<br />

that, PBS buffer (pH 8.0, 10 mM) was added to the chloroform<br />

solution of the complexes with at least a 1/1 volume ratio;<br />

chloroform was then gradually removed by rotary evaporation<br />

at 35 °C and water-soluble carboxy group modified Fe3O4<br />

MNPs were obtained in a clear and dark purple solution. This<br />

transfer process had a 100% efficiency, and no residue was<br />

observed. The original concentrations of ∼14.51 and ∼32.82<br />

nm soluble Fe3O4 MNPs analyzed by atomic absorption<br />

spectroscopy are 205.4 and 16.2 nM, respectively, which will<br />

be used to prepare other concentrations of Fe3O4 MNP<br />

solutions by dilution.<br />

Synthesis of Fe3O4 MNP-Aptamer Conjugates. The<br />

monodisperse and soluble Fe3O4 MNPs with ∼32.82 nm were<br />

diluted into pH 8.0 PBS buffer with a final concentration of 1.6<br />

nM. Then, 1 mg of EDC and 1 mg of NHS were added to 5<br />

(27) Yu, W. W.; Falkner, J. C.; Yavuz, C. T.; Colvin, V. L. Chem. Commun. 2004,<br />

20, 2306–2307.<br />

(28) Yu, W. W.; Chang, E.; Sayes, C. M.; Drezek, R.; Colvin, V. L. Nanotechnology<br />

2006, 17, 4483–4487.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6783


mL of Fe3O4 MNP solution under stirring for 0.5 h to activate<br />

the carboxyl group on the surface of Fe3O4 MNPs. After that,<br />

100 µL of 37.7 µM amino-modified antiadenosine aptamer was<br />

added in this solution, and they were allowed to react for 2 h<br />

to immobilize aptamer on the surface of Fe3O4 MNPs. After<br />

that, 1 M ethanolamine was added for 1 h to block the<br />

unreacted carboxyl groups. Then, this solution was centrifuged<br />

at 14 000g at room temperature for 25 min twice to remove<br />

the free amino-aptamer. At last, the Fe3O4 MNPs were<br />

dispersed in 5 mL of pH 8.0 PBS buffer and stored at 4 °C.<br />

In Situ SPR Measurement. The SPR experiments were done<br />

using Eco Chemie Autolab SPR systems (Brinkmann Instruments,<br />

New York). 29,30 It works with a laser diode fixed at a wavelength<br />

of 670 nm, using a vibrating mirror to modulate the angle of<br />

incidence of the p-polarized light beam on the SPR substrate. The<br />

instrument was equipped with a cuvette. A gold sensor disk (25<br />

mm in diameter) was mounted on the hemicylindrical lens (with<br />

index-matching oil) to form the base of the cuvette. The cuvette<br />

could contain sample with adjustable volume from 10 to 1000 µL.<br />

An O-ring (3 mm inner diameter) between the cuvette and disk<br />

prevents leakage. An autosampler (Eco Chemie) with a controllable<br />

aspirating-dispensing-mixing pipet was used to add<br />

samples into the cuvette and provide constant mixture by aspiration<br />

and dispensing during measurements. This experimental<br />

arrangement maintains a homogeneous solution and reproducible<br />

hydrodynamic conditions. The injection rate and mixing rate for<br />

all samples were 10 and 40 µL/s, respectively, with the total<br />

volume for all samples dispensed in the SPR cell equal to 40 µL.<br />

This setup allows us to measure the SPR angle shift in millidegrees<br />

(m°) as a response unit to quantify the binding amount of<br />

macromolecules to the sensor surface.<br />

Details of the experiment are as follows. The SPR gold film<br />

was initially immersed into the thiolated ss-DNA solution for 12 h<br />

in order to assemble the monolayer of ss-DNA. Then, the modified<br />

gold film was thoroughly rinsed with 50 mM Tris-HCl buffer<br />

and water to remove the weakly adsorbed ss-DNA. Then, the ss-<br />

DNA modified SPR gold film was immersed in 100 µM 6-mercaptohexanol<br />

for 1htoblock the uncovered gold surface. This gold<br />

film was used as a sensing surface to detect the amount of aptamer<br />

possessing ss-DNA structure on Fe 3O4 MNP-aptamer conjugates,<br />

which can be adjusted by the concentration of adenosine<br />

added to the Fe3O4 MNP-aptamer conjugate solution. The<br />

detection procedure is made up of two steps. First, Fe3O4<br />

MNP-aptamer conjugate solution was mixed with different<br />

concentrations of adenosine for 30 min. After that, adenosine<br />

bound with Fe3O4 MNP-aptamer conjugates were separated<br />

and enriched by centrifuging at 14 000g for 1 h twice. The<br />

precipitation was dispersed in PBS again. The resulting solution<br />

was injected into the SPR cell, and the SPR angle-time curve<br />

was recorded. In order to reduce the disturbance of DNA<br />

denaturation that resulted from the regenerating process for<br />

the detecting results, we change a new substrate after each<br />

detection. The modification of each gold substrate is carried<br />

out under the same experimental condition. To confirm the<br />

reproducibility of the detection result, different concentrations<br />

(29) Wang, J. L.; Zhou, H. S. Anal. Chem. 2008, 80, 7174–7178.<br />

(30) Wang, J. L.; Munir, A.; Li, Z. H.; Zhou, H. S. Biosens. Bioelectron. 2009,<br />

25, 124–129.<br />

6784 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

of the MNPs and adenosine are repeatedly detected for three<br />

times.<br />

RESULTS AND DISCUSSION<br />

Characterization of Fe3O4 MNPs. Although soluble Fe3O4<br />

MNPs could be easily synthesized by coprecipitation of aqueous<br />

Fe 2+ /Fe 3+ salt solutions with the addition of a base under<br />

an inert atmosphere at room temperature or at elevated<br />

temperature, the size, shape, and composition of the MNPs<br />

very much depends on the type of salts used (e.g., chlorides,<br />

sulfates, nitrates), the Fe 2+ /Fe 3+ ratio, the reaction temperature,<br />

the pH value, and ionic strength of the media. 31-33 Furthermore,<br />

the Fe3O4 MNPs are also easy to aggregate. Inspired by the<br />

synthesis of high-quality semiconductor nanocrystals and<br />

oxides in nonaqueous media by thermal decomposition, 34-36<br />

monodisperse Fe3O4 MNPs with controlled size has essentially<br />

been synthesized through thermal decomposition of ion<br />

compounds in high-boiling organic solvents. 37,38 Here, we<br />

synthesize Fe3O4 MNPs by the pyrolysis of iron carboxylate in<br />

an organic phase. By changing the ratio of FeO(OH) and oleic<br />

acid, two kinds of Fe3O4 MNPs are prepared. Figure 1A,B<br />

provides the TEM images of the prepared Fe3O4 MNPs and their<br />

size distribution. The average size of Fe3O4 MNPs derived from<br />

Figure 1A,B are 14.51 nm (n ) 300 particles) and 32.82 nm<br />

(n ) 138 particles), respectively. Importantly, both kinds of Fe3O4<br />

MNP size distributions are narrow, which indicates the<br />

prepared Fe3O4 MNPs are monodisperse. After transferring<br />

Fe3O4 MNPs from organic reagent to water solution by the<br />

amphiphilic polymer, Fe3O4 MNPs show a good stability in both<br />

PBS and Tris buffer due to the large hydrodynamic size of<br />

polymer (data not shown), which are all beneficial for the<br />

acquisition of accurate and repeated SPR analytical results.<br />

SPR Response and Concentration Dependence of Fe3O4<br />

MNPs. The basis of a particle-enhanced bioassay is that biomolecular<br />

interaction events lead to particle immobilization, i.e., more<br />

immobilized proteins yield higher particle coverage. 39 Currently,<br />

most SPR instruments are able to quantitatively detect the<br />

concentration of biomolecules through calculating the SPR angle<br />

shift enhanced by the binding of nanoparticle. Considering our<br />

synthesized Fe3O4 MNPs are protected by negative polymer,<br />

2-mercaptoethyamine is used on SPR Au substrates for the<br />

adsorption of Fe3O4 MNPs. The thiol group binds to the Au<br />

surface, leaving the amine group free to bind with carboxyl<br />

group in soluble Fe3O4 MNPs by electrostatic interaction.<br />

(31) Lu, A. H.; Salabas, E. L.; Schuth, F. Angew. Chem., Int. Ed. 2007, 46, 1222–<br />

1244.<br />

(32) Kang, Y. S.; Risbud, S.; Rabolt, J. F.; Stroeve, P. Chem. Mater. 1996, 8,<br />

2209–2211.<br />

(33) Massart, R. IEEE Trans. Magn. 1981, 17, 1247–1248.<br />

(34) Murray, C. B.; Norris, D. J.; Bawendi, M. G. J. Am. Chem. Soc. 1993, 115,<br />

8706–8715.<br />

(35) Peng, X.; Wickham, J.; Alivisatos, A. P. J. Am. Chem. Soc. 1998, 120, 5343–<br />

5344.<br />

(36) O’Brien, S.; Brus, L.; Murray, C. B. J. Am. Chem. Soc. 2001, 123, 12085–<br />

12086.<br />

(37) Sun, S.; Zeng, H.; Robinson, D. B.; Raoux, S.; Rice, P. M.; Wang, S. X.; Li,<br />

G. J. Am. Chem. Soc. 2004, 126, 273–279.<br />

(38) Redl, F. X.; Black, C. T.; Papaefthymiou, G. C.; Sandstrom, R. L.; Yin, M.;<br />

Zeng, H.; Murray, C. B.; O’Brien, S. P. J. Am. Chem. Soc. 2004, 126, 14583–<br />

14599.<br />

(39) Lyon, L. A.; Pena, D. J.; Natan, M. J. J. Phys. Chem. B 1999, 103, 5826–<br />

5831.


Figure 1. TEM images of the prepared Fe3O4 MNPs: (A) 14.51 nm and (B) 32.82 nm.<br />

Figure 2. Variation of SPR angle-time curves with the concentration of Fe3O4 MNPs: (A) 14.51 nm; (B) 32.82 nm. (C) Comparison of the<br />

variation of SPR angle shift with the concentration of Fe3O4 MNPs of 14.51 and 32.82 nm, respectively.<br />

Figure 2 illustrates the changes that occur in the SPR angle shift<br />

as a function of the concentration for 14.51 nm (Figure 2A) and<br />

32.82 nm (Figure 2B) Fe3O4 MNPs. In the case of 14.51 nm<br />

Fe3O4 MNPs (Figure 2A), the SPR angle shifts gradually increase<br />

from 6.4 to 1111.06 m° with the increase of Fe3O4 MNP<br />

concentrations from 0.016 to 1.6 nM. It is evident that a higher<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6785


Figure 3. AFM images of 14.51 nm Fe3O4 MNP (A) and 32.82 nm Fe3O4 MNP (B) modified SPR gold substrate. The concentration of Fe3O4<br />

MNPs is 1.6 nM, and the assembly time is 10 min.<br />

coverage of Fe3O4 MNPs on SPR gold film could be reached<br />

at a higher concentration due to the rapid diffusion adsorption.<br />

It should be pointed out that the SPR angle shift resulting from<br />

the adsorption of Fe3O4 MNPs is much higher than that of the<br />

value resulting from the adsorption of most biomolecules under<br />

the same concentration. 40,41 It means Fe3O4 MNPs greatly<br />

enhance the signal of SPR spectroscopy. Importantly, the SPR<br />

angle shift could be further increased when 32.82 nm Fe3O4<br />

MNPs are used. Figure 2B shows the SPR angle shift curves<br />

resulting from the adsorption of 32.82 nm Fe3O4 MNPs with the<br />

same concentration sequence as 14.51 nm Fe3O4 MNPs. With<br />

the increase of Fe3O4 MNP concentrations from 0.016 to 1.6<br />

nM, the SPR angle shifts increase from 79.22 to 2479.79 m°.<br />

To further understand the size effect of Fe3O4 MNPs on the<br />

SPR response, the relation between SPR angle shifts and the<br />

concentrations of Fe3O4 MNPs with the two sizes are compared<br />

in Figure 2C. Obviously, much larger angle shifts are observed<br />

for 32.82 nm Fe3O4 MNPs because of its larger moleculer<br />

weight. Besides that, the affinity constants and surface coverage<br />

for both kinds of Fe3O4 MNPs could also be obtained from<br />

Figure 2C. From the data of Figure 2C, we can see that the<br />

plasmon resonance shifts with the increasing Fe3O4 MNP<br />

concentration and a simple Langmuir isotherm 42 could be used<br />

to fit the data. The Langmuir equation used to fit the data is<br />

given by<br />

KA [C] Fe3O4 ∆λ ) ∆λmax1 + KA [C] Fe3O4 Where ∆λ is the angle shift caused by the adsorption of Fe3O4<br />

MNPs, ∆λmax is the angle shift which will be observed at<br />

saturation, KA is the apparent equilibrium affinity constant, and<br />

[C]Fe3O4 is the concentration of Fe3O4 MNPs. Using the above<br />

equation, KA values for 14.51 and 32.82 nm Fe3O4 MNPs are<br />

(40) Mullett, W. M.; Lai, E. P. C.; Yeung, J. M. Methods 2000, 22, 77–91.<br />

(41) Hoa, X. D.; Kirk, A. G.; Tabrizian, M. Biosens. Bioelectron. 2007, 23, 151–<br />

160.<br />

(42) Lee, H. J.; Wark, A. W.; Corn, R. M. Langmuir 2006, 22, 5241–5250.<br />

6786 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(1)<br />

calculated to be around 7.14 × 10 8 and 2.63 × 10 9 M -1 ,<br />

respectively, which are similar to the value of NPs as found in<br />

literature. 43 The saturation responses ∆λmax of 14.51 and 32.82<br />

nm are calculated to be 2165 and 3134 0 m, respectively. It can<br />

be seen that 32.82 nm Fe3O4 MNPs have slightly higher affinity<br />

for adsorption on the sensing surface due to their larger<br />

molecular weight; therefore, they show larger angle shifts as<br />

well as higher saturation response (∆λmax) compared to 14.51<br />

nm Fe3O4 MNPs. In our following experiments, we, therefore,<br />

use 32.82 nm Fe3O4 MNPs. The ratio (∆λ)/(∆λmax) gives the<br />

fraction of surface coverage, and if the bulk concentration of<br />

[C]Fe3O4 is equal to (1)/(KA), half of the surface sites will be<br />

occupied ((∆λ)/(∆λmax) ) 0.5). On the basis of the experimental<br />

results and the fitting parameters, we calculate that 80%<br />

of the coverage ((∆λ)/(∆λmax) ) 0.8) will be obtained when<br />

32.82 nm Fe3O4 MNPs with a concentration of 1.6 nM are used.<br />

This concentration will be used in our following experiments<br />

because the changes in SPR angle shift (∆λ) will be very small<br />

even if we further increase the concentration of Fe3O4 MNPs.<br />

To demonstrate that a dense monolayer of MNPs could be<br />

formed on a SPR substrate within 10 min using a 1.6 nM Fe3O4<br />

MNP solution as assembly solution, AFM is used to evaluate<br />

the surface morphology of MNP modified SPR gold substrate.<br />

Figure 3A,B shows the AFM images of 14.51 nm Fe3O4 MNPs<br />

and 32.82 nm Fe3O4 MNPs on SPR gold substrate, respectively.<br />

As seen from the AFM figures, the nanoislands are formed<br />

after MNPs are assembled. The diameter of nanoislands in<br />

Figure 3A is smaller than that of the value in Figure 3B due to<br />

different diameter of MNPs. However, the dense layer of MNPs<br />

on SPR gold substrate has been observed from Figure 3 for both<br />

kinds of Fe3O4 MNPs at the present experimental condition,<br />

which indicates the adsorption of Fe3O4 MNPs on SPR gold<br />

substrate is fast and the present experimental condition is<br />

suitable for further experiments.<br />

(43) Liao, W. S.; Chen, X.; Yang, T. L.; Castellana, E. T.; Chen, J. X.; Cremer,<br />

P. S. Biointerphases 2009, 4, 80–85.


Figure 4. Schematic representation of the SPR biosensor for the detection of the small molecules.<br />

Fe3O4 MNP Enhanced SPR Sensing for the Detection<br />

of Small Molecules. To further demonstrate the practicability<br />

of the amplification effect of Fe3O4 MNPs in an enhancing SPRbased<br />

bioassay, a novel SPR sensor based on indirect competitive<br />

inhibition assay (ICIA) for the detection of adenosine is<br />

constructed. The principle of this SPR sensor is shown in<br />

Figure 4. The partial complementary thiolated ss-DNA of antiadenosine<br />

aptamer is first immobilized on SPR gold film as a<br />

sensing surface. When Fe3O4 MNP-antiadenosine aptamer<br />

conjugate solution is added to the SPR cell in the absence of<br />

adenosine, Fe3O4 MNP-antiadenosine aptamer conjugates will<br />

be adsorbed to the SPR sensor by the DNA hybridization<br />

reaction and result in a huge change of SPR signal due to the<br />

amplification effect of Fe3O4 MNPs. However, the change of<br />

SPR signal will decrease after Fe3O4 MNP-antiadenosine<br />

aptamer conjugates bind with adenosine. This is because<br />

adenosine reacts with antiadenosine aptamer in Fe3O4 MNPantiadenosine<br />

aptamer conjugates and changes its structure<br />

from ss-DNA to tertiary structure, which cannot hybridize with<br />

its partial complementary ss-DNA immobilized on the SPR gold<br />

surface. Thus, the change of SPR signal will decrease with the<br />

increase of the number of Fe3O4 MNP-antiadenosine aptamer<br />

conjugates possessing tertiary structure, which is proportional<br />

to the concentration of adenosine.<br />

The essential prerequisite of this detection is to prepare Fe3O4<br />

MNP-antiadenosine aptamer conjugates. For this purpose,<br />

EDC and NHS are used as a carboxyl activating agent for the<br />

coupling of primary amines in aptamer and carboxyl groups<br />

in polymer coated MNPs. To demonstrate that MNPs have<br />

been labeled by an aptamer successfully, FT-IR spectroscopy<br />

is used to evaluate the surface modification of MNPs before<br />

and after aptamer labeling. Figure 5 shows the IR absorbance<br />

spectra of Fe3O4 MNPs before and after the aptamer label. For<br />

the polymer coated Fe3O4 MNPs, the absorbance peak near<br />

1710 cm -1 is assigned to υ (CdO) (stretch vibration of<br />

carbonyl) and amide I, peaks at 1550 cm -1 are assigned to<br />

amide II, and the peaks at 2854 and 2925 cm -1 are assigned to<br />

υs (C-H2) and υas (C-H2) (symmetric and asymmetric stretch<br />

vibrations of C-H2), respectively. In the region of 3200-3570<br />

Figure 5. FT-IR of polymer coated Fe3O4 MNPs before and after<br />

aptamer labeling.<br />

cm -1 , a peak due to the O-H stretch is expected. Several new<br />

peaks are observed after MNPs are labeled by the aptamer.<br />

The peaks at 1048 and 1211 cm -1 are attributed to stretching<br />

vibrations of the PO 2- in the aptamer. It should be pointed out<br />

that we are not able to confirm the presence of DNA-related<br />

peaks at 1550 cm -1 because the peak overlaps with amide II.<br />

After Fe3O4 MNP-antiadenosine aptamer conjugates are<br />

prepared successfully, the SPR detection is carried out<br />

according to the principle described in Figure 4. Figure 6A<br />

shows the SPR angle-time curves of the separation products<br />

obtained after Fe3O4 MNP-antiadenosine aptamer conjugates<br />

are reacted with different concentrations of adenosine for<br />

30 min. In the absence of adenosine, Fe3O4MNP-antiadenosine<br />

aptamer conjugates in solution directly hybridize with ss-<br />

DNA immobilized on SPR gold film, resulting in the largest<br />

SPR angle shift (∼1082.94 m°). This angle shift is much<br />

larger than that of the angle shift resulting from the binding<br />

of ss-DNA or most of the protein. The SPR angle shift<br />

resulting from the binding of Fe3O4 MNP-antiadenosine<br />

aptamer conjugates decreases (∼820.14 m°) after Fe3O4<br />

MNP-antiadenosine aptamer conjugates are reacted with<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6787


Figure 6. (A) SPR angle-time curves of the separation products obtained after Fe3O4 MNP-antiadenosine aptamer conjugates are reacted<br />

with different concentrations of adenosine for 30 min. Inset: The linear relationship between the logarithms of adenosine concentrations and the<br />

SPR angle shift resulting from the binding of Fe3O4 MNP-antiadenosine aptamer conjugates. (B) SPR angle-time curves of 1 µM antiadenosine<br />

aptamer without adenosine (red line) and after react with 1 mM adenosine for 30 min (black line).<br />

10 nM adenosine because parts of aptamers on Fe3O4 MNPs<br />

react with adenosine and form its tertiary structure, which<br />

cannot hybridize with its complementary ss-DNA. With the<br />

further increase of the concentration of adenosine added to<br />

the Fe3O4 MNP-antiadenosine aptamer conjugate solution,<br />

the SPR angle shift continuously decreases until most of the<br />

antiadenosine aptamer binds with adenosine. By analyzing<br />

the change of SPR angle shift with the concentrations of<br />

adenosine, a good linear relationship between the logarithms<br />

of adenosine concentrations and the SPR angle shift is<br />

obtained with a range of 10-1 × 10 4 nM (shown in the insert<br />

of Figure 6A). This detection result is comparable with most<br />

other aptasensors 44 and is a little lower than that of detection<br />

results from SPR aptasensor which utilize Au NPs as an<br />

amplification reagent 20,29 but much superior to the detection<br />

results obtained by a general SPR sensor based on a molecularly<br />

imprinted technique. 45,46 It should be pointed out that the<br />

SPR angle shift only decreases to ∼107.33 m° even when the<br />

aptamer reacts with 1 mM adenosine. This nonspecific adsorption<br />

may come from the stereohindrance effect of aptamer<br />

possessing different structures. After adenosine is added to<br />

Fe3O4 MNP-aptamer conjugates, most free-coiled aptamers<br />

react with adenosine and form its tertiary structure. However,<br />

the formation of a tertiary structure of an aptamer may<br />

inhibit the further reaction between its adjacent free-coiled<br />

aptamer and adenosine. To clearly show the amplification<br />

effect of Fe3O4 MNPs, as a comparison, the SPR response<br />

resulting from the binding of aptamer without MNP is also<br />

studied. The black line in Figure 6B is the SPR angle shift<br />

resulting from the binding of 1 µM antiadenosine aptamer after<br />

antiadenosine aptamer reactes with 1 mM adenosine for 0.5 h.<br />

Only a 24.6 m° SPR angle shift is observed. Although the SPR<br />

angle shift resulting from the binding of antiadenosine aptamer<br />

could increase to 65.7 m° (Red line in Figure 6B) in the absence<br />

of adenosine, this value is still much lower than that of the<br />

(44) Wang, Y. L.; Wei, H.; Li, B. L.; Ren, W.; Guo, S. J.; Dong, S. J.; Wang, E. K.<br />

Chem. Commun. 2007, 5220–5222.<br />

(45) Taniwaki, K.; Hyakutake, A.; Aoki, T.; Yoshikawa, M.; Guiver, M. D.;<br />

Robertson, G. P. Anal. Chim. Acta 2003, 489, 191–198.<br />

(46) Yoshikawa, M.; Guiver, M. D.; Robertson, G. P. J. Mol. Struct. 2005, 739,<br />

41–46.<br />

6788 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 7. SPR angle-time curves of the separation products<br />

obtained after Fe3O4 MNP-antiadenosine aptamer conjugates are<br />

reacted with different analytes for 30 min.<br />

value resulting from the binding of Fe3O4 MNP-antiadenosine<br />

aptamer conjugates (1082.94 m°). Besides sensitivity, the<br />

specification of aptamer promises the selectivity of the<br />

present SPR sensor for adenosine. Figure 7 exhibits SPR<br />

angle-time curves of the separation products obtained after<br />

Fe3O4 MNP-antiadenosine aptamer conjugates are reacted<br />

with different analytes for 30 min. It could be easily seen<br />

that the addition of Fe3O4 MNP-antiadenosine aptamer<br />

conjugates which are reacted with 1 × 10 6 nM cytidine,<br />

guanosine, and uridine result in big SPR angle shifts.<br />

However, The SPR angle shift only increases a little after<br />

Fe3O4 MNP-antiadenosine aptamer conjugates are reacted<br />

with 1 × 10 6 nM adenosine. These results not only demonstrate<br />

that Fe3O4 MNPs could greatly enhance the SPR signal<br />

but also, more importantly, give us an important indication<br />

that SPR spectroscopy could be an excellent candidate for<br />

detecting an MNP-based separation product.<br />

CONCLUSION<br />

In summary, the SPR response of the carboxyl group<br />

modified Fe3O4 MNPs of two different sizes onto an amino<br />

group modified SPR gold substrate has been studied. The<br />

results show the monolayer adsorption of Fe3O4 MNPs could


esult in a big SPR signal change with a low optical loss. To<br />

evaluate the practicability of the use of Fe3O4 MNPs in<br />

enhancing the SPR signal for biosensing, a novel SPR sensor<br />

based on ICIA for the detection of adenosine is constructed<br />

using Fe3O4 MNP-antiadenosine aptamer conjugates as the<br />

amplification reagents. The experimental results demonstrate<br />

that the SPR sensor possesses a good sensitivity and a high<br />

selectivity for adenosine. Importantly, by changing the kind<br />

of biomolecules labeled by MNPs, the present detection<br />

method will be able to explore new applications of SPR<br />

spectroscopy for the detection of a large variety of MNPbased<br />

separation products. At the same time, the technique<br />

demonstrated in this work could also offer a new direction<br />

in designing high performance SPR biosensors for sensitive<br />

and selective detection of small molecules by the amplification<br />

effect of MNPs.<br />

ACKNOWLEDGMENT<br />

This work is supported by National Science Foundation (EEC-<br />

0823974). Authors thank Prof. Camesano Terri and her graduate<br />

student Sena Ada for assistance with AFM characterization.<br />

Received for review March 10, 2010. Accepted July 8,<br />

2010.<br />

AC100812C<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6789


Anal. Chem. 2010, 82, 6790–6796<br />

Improved Sensitivity Mass Spectrometric<br />

Detection of Eicosanoids by Charge Reversal<br />

Derivatization<br />

James G. Bollinger, † Wallace Thompson, † Ying Lai, ‡ Rob C. Oslund, † Teal S. Hallstrand, ‡<br />

Martin Sadilek, † Frantisek Turecek, † and Michael H. Gelb* ,†,§<br />

Departments of <strong>Chemistry</strong>, Medicine, and Biochemistry, University of Washington, Seattle, Washington 98195<br />

Combined liquid chromatography-electrospray ionization-tandem<br />

mass spectrometry (LC-ESI-MS/MS) is a<br />

powerful method for the analysis of oxygenated metabolites<br />

of polyunsaturated fatty acids including eicosanoids.<br />

Here we describe the synthesis of a new derivatization<br />

reagent N-(4-aminomethylphenyl)pyridinium (AMPP) that<br />

can be coupled to eicosanoids via an amide linkage in<br />

quantitative yield. Conversion of the carboxylic acid of<br />

eicosanoids to a cationic AMPP amide improves sensitivity<br />

of detection by 10- to 20-fold compared to negative<br />

mode electrospray ionization detection of underivatized<br />

analytes. This charge reversal derivatization allows detection<br />

of cations rather than anions in the electrospray<br />

ionization mass spectrometer, which enhances sensitivity.<br />

Another factor is that AMPP amides undergo considerable<br />

collision-induced dissociation in the analyte portion rather<br />

than exclusively in the cationic tag portion, which allows<br />

isobaric derivatives to be distinguished by tandem mass<br />

spectrometry, and this further enhances sensitivity and<br />

specificity. This simple derivatization method allows prostaglandins,<br />

thromboxane B2, leukotriene B4, hydroxyeicosatetraenoic<br />

acid isomers, and arachidonic acid to<br />

be quantified in complex biological samples with limits<br />

of quantification in the 200-900 fg range. One can<br />

anticipate that the AMPP derivatization method can be<br />

extended to other carboxylic acid analytes for enhanced<br />

sensitivity detection.<br />

Liquid chromatography coupled to electrospray ionization<br />

tandem mass spectrometry (LC-ESI-MS/MS) has emerged as a<br />

powerful method to detect oxygenated derivatives of fatty acids<br />

including eicosanoids (for example, refs 1-3). With these methods<br />

it is possible to analyze a large collection of eicosanoids in a single<br />

LC-ESI-MS/MS run. These lipid mediators are detected by single<br />

* To whom correspondence should be addressed. Michael H. Gelb, Depts.<br />

of <strong>Chemistry</strong> and Biochemistry, Campus Box 35100, University of Washington,<br />

Seattle, WA 98195. Phone: 206 525-8405. Fax: 206 685-8665. E-mail:<br />

gelb@chem.washington.edu.<br />

† Department of <strong>Chemistry</strong>.<br />

‡ Department of Medicine.<br />

§ Department of Biochemistry.<br />

(1) Dickinson, J. S.; Murphy, R. C. J. Am. Soc. Mass Spectrom. 2002, 13, 1227.<br />

(2) Kita, Y.; Takahashi, T.; Uozumi, N.; Nallan, L.; Gelb, M. H.; Shimizu, T.<br />

Biochem. Biophys. Res. Commun. 2005, 330, 898.<br />

(3) Buczynski, M. W.; Stephens, D. L.; Bowers-Gentry, R. C.; Grkovich, A.;<br />

Deems, R. A.; Dennis, E. A. J. Biol. Chem. 2007, 282, 22834.<br />

6790 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

reaction monitoring (SRM) in which precursor ions are isolated<br />

in the first stage of the mass spectrometer followed by collisioninduced<br />

dissociation to give fragment ions, which are detected<br />

after an additional stage of mass spectrometer isolation. The<br />

current limit of quantification for these analytes is in the ∼10-20<br />

pg range. This sensitivity level is appropriate for studies with<br />

cultured cells in vitro or with relatively large tissue samples, but<br />

it is not sufficient for studies with smaller volume samples such<br />

as joint synovial fluid or bronchoalveolar lavage fluid from<br />

experimental rodents. Given the importance of oxygenated fatty<br />

acid derivatives in numerous medically important processes such<br />

as inflammation and resolution of inflammation, we sought to<br />

improve the LC-ESI-MS/MS sensitivity of detection of these lipid<br />

mediators using a widely available analytical platform.<br />

For reasons that are not well understood, cations generally<br />

form gaseous ions better than anions in the electrospray ionization<br />

source of the mass spectrometer. Additionally, for underivatized<br />

carboxylic acids it is required to add a weak organic acid to the<br />

chromatographic mobile phase, i.e., formic acid, so that the<br />

carboxylic acid is kept in its protonated state, which allows it to<br />

be retained on the reverse-phase column to ensure chromatographic<br />

separation. However, the presence of the weak acid offsets<br />

the formation of carboxylate anions in the electrospray source<br />

because the weak acid carries most of the anionic charge in the<br />

electrospray droplets, and thus formation of analyte anions is<br />

suppressed. We reasoned that conversion of the carboxylic acid<br />

to a fixed-charge cationic derivative would lead to improved<br />

detection sensitivity by ESI-MS/MS. Charge-reversal derivatization<br />

of carboxylic acids with quaternary amines has been explored in<br />

previous work (for example, refs 4-6). However, these reagents<br />

utilize organic cations that tend to fragment by collision-induced<br />

dissociation near the cationic site. Fragmentation in the derivatization<br />

tag is not desirable because analytes that form isobaric<br />

precursor ions and that comigrate on the LC column will not be<br />

distinguished in the mass spectrometer if they give rise to the<br />

same detected fragment ion. This loss of analytical specificity is<br />

a serious problem when analyzing complex biological samples.<br />

Fragmentation in the analyte portion rather than in the tag portion<br />

(4) Lamos, S. M.; Shortreed, M. R.; Frey, B. L.; Belshaw, P. J.; Smith, L. M.<br />

Anal. Chem. 2007, 79, 5143.<br />

(5) Yang, W.; Adamec, J.; Regnier, F. E. Anal. Chem. 2007, 79, 5150.<br />

(6) Pettinella, C.; Lee, S. H.; Cipollone, F.; Blair, I. A. J. Chromatogr., B: Anal.<br />

Technol. Biomed. Life Sci. 2007, 850, 168.<br />

10.1021/ac100720p © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/22/2010


also reduces chemical noise, which also enhances sensitivity of<br />

detection.<br />

In this study we report the design and synthesis of a new<br />

cationic tag and show that it can be quantitatively attached via an<br />

amide linkage to the carboxyl group of eicosanoids by a simple<br />

derivatization procedure. We then show that the derivatized<br />

eicosanoids can be analyzed by LC-ESI-MS/MS with limits of<br />

quantification that are well below those reported for underivatized<br />

eicosanoids.<br />

EXPERIMENTAL METHODS<br />

Synthesis of AMPP. Pyridine (40 mmol, 3.2 mL) was<br />

dissolved in 46 mL of absolute ethanol followed by the addition<br />

of 1-chloro-2,4-dinitrobenzene (40 mmol, 8.2 g, Aldrich). The<br />

mixture was heated with a reflux condenser at 98 °C for 16 h<br />

under nitrogen. After cooling, ethanol was removed by rotary<br />

evaporation, and the crude product was recrystallized by dissolving<br />

in a minimal amount of hot ethanol and allowing the solution to<br />

slowly cool. The product N-2,4 dinitrophenyl pyridinium chloride<br />

was isolated as a yellow solid in 62% yield, and its identity was<br />

confirmed by melting point analysis (189-191 °C observed,<br />

189-190 °C reported 7 ). N-2,4-Dinitrophenyl pyridinium chloride<br />

(16.8 mmol, 4.76 g) was dissolved in 70 mL of ethanol-pyridine<br />

(3:1). 4-[(N-Boc) -amino-methyl] aniline (33.6 mmol, 7.56 g,<br />

Aldrich) was added, and the reaction mixture was heated under<br />

a reflux condenser at 98 °C under nitrogen for 3 h. After cooling,<br />

700 mL of water was added to precipitate 2,4-dinitroaniline. After<br />

filtration, the filtrate was concentrated to dryness by rotary<br />

evaporation, and the product was isolated as a brown oil. This oil<br />

was treated with 112 mL of 25% (v/v) trifluoroacetic acid in<br />

dichloromethane for 30 min at room temperature. The mixture<br />

was concentrated by rotary evaporation, and the solid was<br />

triturated twice with benzene to remove excess trifluoroacetic acid.<br />

The mixture was again concentrated by rotary evaporation. The<br />

residue was dissolved in a minimal amount of heated ethanol, the<br />

solution was allowed to cool for ∼5 min, and then diethyl ether<br />

was added with swirling until the solution started to cloud up.<br />

The mixture was transferred to the freezer (-20 °C) and left<br />

overnight. The mixture was allowed to warm to room temperature<br />

and then decanted. The obtained mother liquor was treated with<br />

additional diethyl ether as above to give additional AMPP solid.<br />

The solids were combined and triturated with diethyl ether. The<br />

solid was dried under vacuum to give 3.30 g of AMPP as a brown<br />

solid, 59% yield. 1 H NMR (300 MHz, D6-DMSO) 9.34 (d, 2H),<br />

8.81 (t, 1H), 8.53 (broad, 3H), 8.33 (t, 2H), 7.95 (d, 2H), 7.80<br />

(d,2H), 4.22 (s, 2H) ( 1 H NMR spectra are shown in the<br />

Supporting Information, estimated purity of AMPP is >95%).<br />

Preparation of Eicosanoid Stock Solutions. The following<br />

eicosanoid standards from Cayman <strong>Chemical</strong>s were used (PGE2,<br />

PGD2, PGF2R, 6-keto-PGF1R, TxB2, 5(S)-HETE, 8(S)-HETE,<br />

11(S)-HETE, 12(S)-HETE, 15(S)-HETE, LTB4, arachidonic acid,<br />

D4-PGE2, D4-PGD2, D4-PGF2R, D4-6-keto-PGF1R, D4-TXB2, D8-<br />

5(S)-HETE, D4-LTB4, D8-arachidonic acid). Stock solutions of<br />

eicosanoids were prepared at a concentration of 100 pg/µL in<br />

absolute ethanol and stored at -80 °C under Ar in Teflon<br />

septum, screw cap vials. Serial dilutions of the stock solutions<br />

were made in absolute ethanol for standard curve and extrac-<br />

(7) Park, K. K.; Lee, J.; Han, D. Bull. Korean Chem. Soc. 1985, 6, 141.<br />

tion analysis. Internal standards were diluted to a working stock<br />

of 5 pg/µL in absolute ethanol.<br />

Preparation of Samples Prior to Derivatization with<br />

AMPP. Standard Curves. Each sample contained 50 pg of each<br />

internal standard and various amounts of nonisotopic eicosanoids<br />

(added from the stock solutions described above) transferred to<br />

a glass autosampler vial insert (Agilent catalog no. 5183-2085).<br />

Solvent was removed with a stream of nitrogen, and the residue<br />

was derivatized with AMPP as described below.<br />

Mouse Serum. Analysis of endogenous eicosanoids in serum<br />

was carried out with commercial mouse serum (Atlantic Biologicals<br />

catalog no. S18110). A volume of 1, 5, or 10 µL of serum was<br />

placed in a glass autosampler vial insert. Two volumes of methanol<br />

(LC/MS, JT Baker catalog no. 9863-01) containing 50 pg of each<br />

internal standard were added. The vial insert was mixed on a<br />

vortex mixer for ∼10 s. The concentration of methanol was<br />

lowered to 10% (v/v) by addition of purified water (Milli-Q,<br />

Millipore Corp.), and the samples were loaded via a glass Pasteur<br />

pipet onto a solid phase extraction cartridge (10 mg Oasis-HLB,<br />

Waters catalog no. 186000383). The cartridges were previously<br />

washed with 1 mL of methanol and then 2 × 0.75 mL of 95:5<br />

water-methanol. After sample loading, the sample tube was rinsed<br />

with 200 µL of purified water-methanol (95:5, v/v), and this was<br />

added to the cartridge. The cartridge was washed with 2 × 1mL<br />

of water-methanol (95:5, v/v). Additional solvent was forced out<br />

of the cartridge solid phase by applying medium pressure N2<br />

(house N2 passed through a 0.2 µm cartridge filter) for a few<br />

seconds. Column eluant receiver vials (Waters Total Recovery<br />

autosampler vials, Waters catalog no. 186002805) were placed<br />

under the cartridges. The cartridges were then eluted with<br />

methanol (1 mL). All cartridge steps were carried out using a<br />

vacuum manifold (Waters catalog no. WAT200606) attached<br />

to a water aspirator. Solvent was removed by placing the<br />

receiver vials in a centrifugal evaporator (Speed-Vac). These<br />

processed samples were derivatized with AMPP (see below)<br />

without storage.<br />

Lung Epithelial Cells. The University of Washington Institutional<br />

Review Board approved the studies involving human<br />

subjects, and written informed consent was obtained from all<br />

participants. Primary bronchial epithelial cells were isolated from<br />

a volunteer with asthma during a bronchoscopy using a nylon<br />

cytology brush of cells from subsegmental airways. To establish<br />

primary culture, the epithelial cells were seeded into a culture<br />

vessel coated with type 1 collagen in bronchial epithelial basal<br />

media (BEBM, Lonza, Allendale, NJ) supplemented with bovine<br />

pituitary extract, insulin, hydrocortisone, gentamicin, amphotericin<br />

B, fluconazole, retinoic acid, transferrin, triiodothyronine, epinephrine,<br />

and human recombinant epidermal growth factor<br />

(serum-free BEGM) and maintained at 37 °C in a humidified<br />

incubator. After expansion in vitro, passage 2 epithelial cells were<br />

grown to >90% confluence on a 12-well plate. The medium was<br />

changed to 200 µL of Hanks balanced salt solution, and the cells<br />

were treated with either calcium ionophore (A23187, 10 µM in<br />

DMSO) or a DMSO-containing control solution for 20 min at 37<br />

°C. The synthesis of eicosanoids was stopped by the addition of<br />

4 volumes of ice-cold methanol with 0.2% formic acid, and samples<br />

were stored at -80 °C until processed. The number of epithelial<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6791


cells was 5.0 × 10 5 cells per well. The studies were performed<br />

in duplicate.<br />

Frozen samples were thawed on ice, 50 pg of each internal<br />

standard was added, and Milli-Q water was added to bring the<br />

methanol to 10% (v/v). The samples were processed by solidphase<br />

extraction (Oasis-HLB) as described for mouse serum<br />

samples.<br />

Rat 3Y1 Cells. 3Y1 cells were a gift from Dr. H. Kuwata<br />

(Department of Health <strong>Chemistry</strong>, School of Pharmaceutical<br />

Sciences, Showa University). 3Y1 cells were maintained in<br />

complete medium consisting of Dulbecco’s Modified Eagle<br />

Medium (DMEM) with low glucose (Invitrogen catalog no. 11885-<br />

084) containing 10% heat inactivated, fetal bovine serum (FBS)<br />

(Invitrogen catalog no. 16140), 100 units/mL penicillin, and 100<br />

µg/mL streptomycin (Invitrogen catalog no. 15140) in plastic<br />

tissue culture dishes (Nunc catalog no. 172958) in a humidified<br />

atmosphere of 5% CO 2 at 37 °C. Passage of cells was performed<br />

using 0.25% trypsin/EDTA (Invitrogen catalog no. 25200-056).<br />

For PGE2 analysis, 3Y1 cells were plated at 5 × 10 4 cells/<br />

well in a 24-well plate (Nunc catalog no. 142475) in 1 mL of<br />

complete medium and incubated overnight. The medium was<br />

then replaced with DMEM containing 2% FBS. After 24 h<br />

incubation, the medium was removed from each well and<br />

replaced with 1 mL DMEM containing 2% FBS and the<br />

cytosolic phospholipase A2-R inhibitor Wyeth-2 (compound 10<br />

from ref 9) or DMSO vehicle control (final concentration in<br />

each well did not exceed 0.1% (v/v)). Cells were incubated for<br />

20 min at room temperature. Mouse interleukin-1� (1 ng/well)<br />

and human tumor necrosis factor-R (1 ng/well) (R & D Systems<br />

catalog no. 401-ML and 210-TA) were both added to each well<br />

in 10 µL of DMEM containing 2% FBS (negative control wells<br />

received only 10 µL of DMEM containing 2% FBS). Cells were<br />

incubated for 48 h at 37 °C in a humidified atmosphere of 5%<br />

CO 2. The supernatant was then carefully removed from each<br />

well and transferred into a glass test tube and placed on ice<br />

for immediate analysis. PGE2 levels were measured by an<br />

enzyme immunoassay kit (Cayman <strong>Chemical</strong> catalog no.<br />

500141) according to the manufacturer’s instructions using 50<br />

µL of medium above the cells. A sample of medium (50 µL)<br />

was also analyzed by LC-ESI-MS/MS as follows. To the<br />

medium was added 2 volumes of methanol containing 50 pg<br />

of each internal standard, methanol was reduced to 10% v/v<br />

by addition of Milli-Q water, and samples were submitted to<br />

solid-phase extraction using Oasis-HLB cartridges as described<br />

for mouse serum samples.<br />

Derivatization with AMPP. To the residue in the Waters<br />

Total Recovery autosampler vial was added 10 µL of ice-cold<br />

acetonitrile (JT Baker catalog no. 9017-03)-N,N-dimethylformamide<br />

(Sigma catalog no. 227056) (4:1, v:v). Then 10 µL of icecold<br />

640 mM (3-(dimethylamino)propyl)ethyl carbodiimide hydrochloride<br />

(TCI America catalog no. D1601) in purified water<br />

was added. The vial was briefly mixed on a vortex mixer and<br />

placed on ice while other samples were processed as above. To<br />

each vial was added 20 µL of 5 mM N-hydroxybenzotriazole<br />

(8) Kita, Y.; Takahashi, T.; Uozumi, N.; Shimizu, T. Anal. Biochem. 2005, 342,<br />

134.<br />

(9) McKew, J. C.; Lee, K. L.; Chen, L.; Vargas, R.; Clark, J. D.; Williams, C.;<br />

Clerin, V.; Marusic, S; Pong, K. Inhibitors of Cytosolic Phospholipase A2.<br />

U.S. Patent 7,557,135 B2, July 7, 2009.<br />

6792 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(Pierce catalog no. 24460)-15 mM AMPP in acetonitrile. The vials<br />

were mixed briefly on a vortex mixer, capped, and placed in a 60<br />

°C incubator for 30 min. The cap was replaced with a split-septum<br />

screw cap (Agilent catalog no. 5185-5824) for autoinjection onto<br />

the LC-ESI-MS/MS. Samples were analyzed on the same day.<br />

Samples were kept in the autosampler rack at 4 °C while queued<br />

for injection.<br />

LC-ESI-MS/MS Analysis. Some studies were carried out<br />

with a Waters Quattro Micro triple quadrupole mass spectrometer,<br />

a 2795 Alliance HT LC/autosampler system, and the QuanLynx<br />

software package. Chromatography was carried out with a C18<br />

reverse-phase column (Ascentis Express C18, 2.1 mm × 150 mm,<br />

2.7 µm, Supelco catalog no. 53825-U). Solvent A is 95% H 2O/5%<br />

CH3CN/1% acetic acid, and solvent B is CH3CN/1% acetic acid.<br />

The solvent program is (linear gradients) 0-1.0 min, 95-78%<br />

A; 1.0-7.0 min, 78-74% A; 7.0-7.1 min, 74-55% A; 7.1-12.1<br />

min, 55-40% A; 12.1-13.0 min, 40-0% A; 13.0-15.0 min, 0%<br />

A; 15.0-15.1 min, 0-95% A; 15.1-20.1 min, 95% A. The flow<br />

rate is 0.25 mL/min.<br />

Similarly, some studies were carried out with a Waters Quattro<br />

Premier triple quadrupole mass spectrometer interfaced to an<br />

Acquity UPLC. Solvent A is 100% water (Fisher Optima grade<br />

catalog no. L-13780)-0.1% formic acid (Fluka catalog no. 94318),<br />

and solvent B is CH3CN (Fisher Optima grade catalog no.<br />

L-14338)-0.1% formic acid. The same LC column was used but<br />

with a modified solvent program (linear gradients): 0-1.0 min,<br />

95% A; 1.0-2.0 min, 95%-85% A; 2.0-2.1 min, 85%-74% A;<br />

2.1-6.0 min, 74%-71% A; 6.0-6.1 min, 71%-56% A; 6.1-10.0<br />

min, 56% A; 10.0-14.0 min, 56%-0% A; 14.0-14.1 min, 0%-95%<br />

A; 14.1-18.0 min, 95% A. Supplemental Material Tables 1 and<br />

2 in the Supporting Information give the autosampler and ESI-<br />

MS/MS data collection parameters for the Waters Quattro Micro<br />

triple quadrupole and the Waters Quattro Premier triple quadrupole<br />

mass spectrometers, respectively.<br />

For comparison purposes, we also carried out LC-ESI-MS/<br />

MS analysis of underivatized eicosanoids in the negative ion mode.<br />

The same LC column, solvents A and B, and flow rate were used<br />

as for the AMPP amides. The solvent program was slightly<br />

modified as follows: For the Waters Quattro Micro, we used 0-1.0<br />

min, 95-63% A; 1.0-7.0 min, 63%A; 7.0-7.1 min, 63-38% A;<br />

7.1-12.1 min, 38-23% A; 12.1-13.0 min, 23-0% A; 13.0-15.0 min,<br />

0% A; 15.0-15.1 min, 0-95% A; 15.1-20.1 min, 95% A. For the<br />

Waters Quattro Premier, we used 0-1.0 min, 95% A; 1.0-2.0 min,<br />

95%-85% A; 2.0-2.1 min, 85%-59% A; 2.1-6.0 min, 59% A; 6.0-6.1<br />

min, 59%-38% A; 6.1-11 min, 38%-23% A; 11-11.1 min, 23%-0%<br />

A; 11.1-14 min, 0% A; 14.0-14.1 min, 0%-95% A; 14.1-19.0 min,<br />

95% A. Values of m/z for precursor and fragment ions were as<br />

published, 8 and cone voltages and collision energies were optimized<br />

for each instrument and for each analyte in the usual way<br />

(values not shown but similar to those published 8 ).<br />

Eicosanoid Recovery Studies. We measured the recovery<br />

of eicosanoids following the sample workup procedure given<br />

above. Recovery studies were done using either 30 mg Strata-X<br />

(Phenomenex Cat. 8B-S100-TAK-S) or 10 mg Oasis-HLB (Waters<br />

Cat. 186000383) cartridges with 50 or 5 pg of each eicosanoid in<br />

phosphate buffered saline. For these recovery studies, 50 pg of<br />

each internal standard was added to samples just prior to<br />

derivatization with AMPP (i.e., post-solid phase extraction).


Table 1. Extraction Yields, Liquid Chromatography Retention Times, and Tandem Mass Spectrometry Parameters<br />

for Eicosanoid AMPP Amide Molecular Species<br />

eicosanoid<br />

extraction<br />

yield (%) a<br />

LC retention<br />

time (min) b<br />

limit of<br />

quantitation in<br />

positive mode<br />

(pg) (CV) c<br />

Recoveries were obtained by comparing the LC-ESI-MS/MS peak<br />

integrals to those obtained from a sample of eicosanoids that were<br />

derivatized with AMPP and injected directly onto the LC column<br />

without sample processing. Recovery yields are given in Table 1.<br />

The peak areas for the internal standards were similar in all<br />

samples studied (not shown) showing that the eicosanoid recoveries<br />

were similar regardless of sample matrix.<br />

RESULTS AND DISCUSSION<br />

Design and Preparation of AMPP Amides. Our goal was<br />

to develop a simple derivatization procedure to convert the<br />

carboxylic acid terminus of eicosanoids to a cationic group so that<br />

mass spectrometry could be carried out in positive mode. The<br />

tag should be designed so that fragmentation occurs in the analyte<br />

portion rather than in the tag portion so that the power of tandem<br />

mass spectrometry can be used for analytical specificity and<br />

sensitivity. Our computational analysis of substituted N-pyridinium<br />

methylamino derivatives 10 indicated that these had only moderate<br />

dissociation energies for loss of pyridine and other single bond<br />

cleavages in the linker and thus might not be suitable for our<br />

purposes, which require fragmentation in the fatty acyl chain.<br />

Therefore, the charge tag was redesigned to strengthen the<br />

pyridinium N-C bond by inserting a phenyl ring as a linker. The<br />

limit of<br />

quantitation in negative<br />

mode (pg) d precursorion e (m/z)<br />

fragment<br />

ion e (m/z)<br />

cone<br />

voltage f (V)<br />

collision<br />

energy f (eV)<br />

6-keto-PGF1R 102/87 5.4/4.1 5/0.3 (6.8) 110/10 537 239 65/80 55/55<br />

PGF2R 64/100 6.7/5.2 5/0.5 (9.1) 110/7 521 239 60/75 43/52<br />

PGE 2 67/74 6.8/5.3 5/0.3 (5.5) 100/7 519 239 60/75 40/45<br />

PGD 2 63/67 7.2/5.7 5/0.6 (7.5) 120/7 519 307 60/75 40/48<br />

TxB2 86/86 6.2/4.6 5/0.2 (2.3) 90/4 537 337 65/70 43/42<br />

LTB 4 42/63 9.8/7.7 5/0.5 (10.3) 80/7 503 323 50/60 35/35<br />

5(S)-HETE 42/65 10.4/9.7 5/0.4 (10.4) 120/10 487 283 55/65 37/34<br />

8(S)-HETE 40/75 10.3/9.1 5/0.2 (6.7) 120/10 487 295 55/65 37/40<br />

11(S)-HETE 31/66 10.2/8.8 5/0.3 (8.9) 80/5 487 335 55/65 32/30<br />

12(S)-HETE 69/62 10.2/8.9 10/0.9 (6.6) 140/10 487 347 55/65 35/35<br />

15(S)-HETE 35/53 10.1/8.5 5/0.4 (11.1) 200/10 487 387 55/65 33/34<br />

arachidonic acid 73/113 11.7/12.4 10/1 (5.6) no data 471 239 55/65 40/50<br />

D 4-6-keto-PGF1R 5.4/4.1 541 241 65/80 55/55<br />

D 4-PGF 2R 6.7/5.2 525 241 60/75 43/52<br />

D 4-PGE2 6.8/5.3 523 241 60/75 40/45<br />

D 4-PGD2 7.1/5.6 523 311 60/75 40/48<br />

D 4-TxB 2 6.2/4.6 541 341 65/70 43/42<br />

D 4-LTB4 9.8/7.7 507 325 50/60 35/35<br />

D 8-5(S)-HETE 10.4/9.6 495 284 55/65 37/34<br />

D 8-arachidonic acid 11.7/12.4 479 239 55/65 40/50<br />

a The first number is for the Waters Oasis HLB cartridge, and the second number is for the Phenomenex Strata-X cartridge. b The first number<br />

is for the Waters Quattro Micro, and the second number is for the Waters Quattro Premier. c LOQ values are for eicosanoid AMPP amides in<br />

positive mode. The first number is for the Waters Quattro Micro, and the second number is for the Waters Quattro Premier. The values in<br />

parentheses are the CVs based on 6 independent 15 µL injections of 0.075 pg/µL on the Waters Quattro Premier (see the main text). d LOQ values<br />

are for underivatized eicosanoids analyzed by LC-ESI-MS/MS in negative mode. The first number is for the Waters Quattro Micro, and the<br />

second number is for the Waters Quattro Premier. e m/z values listed are calculated monoisotopic values. The actual values used are derived<br />

from instrument tuning, which is instrument dependent. f Cone voltages and collision energies were optimized for each analyte. These numbers<br />

are instrument dependent. The first number is for the Waters Quattro Micro instrument, and the second number is for the Waters Quattro Premier<br />

instrument.<br />

Figure 1. Structure of AMPP and an AMPP amide along with the reagents used for derivatization.<br />

phenyl ring also serves to enhance the derivative’s interaction with<br />

the LC reverse phase column, which is necessarily used to<br />

minimize any matrix effects during ESI-MS/MS analysis. Density<br />

functional theory calculations (B3LYP/6-31G*) indicated that<br />

homolytic cleavage of the benzylic CH 2-N bond in the charge<br />

tag was a high-energy process that required >350 kJ mol -1 of<br />

dissociation energy and was deemed not to out compete<br />

fragmentations in the fatty acyl chain (Supplementary Material<br />

Figure 1 in the Supporting Information). A side reaction that<br />

could not be evaluated with the model system shown in Supplementary<br />

Material Figure 1 in the Supporting Information is<br />

elimination of 4-(N-pyridyl)benzylamine, which gives rise to the<br />

m/z 183 marker fragment ion. An amide linkage of the charge<br />

tag to the analytic carboxylic acid was preferred over an ester<br />

since the former are generally more resistant to fragmentation in<br />

the mass spectrometer. 11 The final feature of the design of AMPP<br />

is its ease of synthesis in pure form.<br />

We developed a simple synthesis of highly pure N-(4-aminomethylphenyl)-pyridinium<br />

(AMPP, Figure 1) and used it as a new<br />

derivatization reagent. We developed a simple method to convert<br />

(10) Chung, T. W.; Turecek, F. Int. J. Mass Spectrom. 2008, 276, 127.<br />

(11) McLafferty, F. W.; Turecek, F. Interpretation of Mass Spectra, 4th ed.;<br />

University Science Books: Mill Valley, CA, 1993.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6793


Figure 2. LC-ESI-MS/MS traces of eicosanoids after workup from a solution in phosphate buffered saline: (A) early eluting eicosanoids and<br />

(B) latter eluting eicosanoids.<br />

the carboxyl group of lipid mediators to the AMPP amide using<br />

a well-known carbodiimide coupling reagent. Coupling conditions<br />

were optimized to give near quantitative formation of the AMPP<br />

amide. The was shown by converting the fluorescent fatty acid<br />

10-pyrene-decanoic acid into its AMPP amide and examining the<br />

reaction mixture by fluorimetric HPLC and ESI-MS. No remaining<br />

10-pyrenedecanoic acid was detected, and the AMPP amide was<br />

formed in >95% yield with


scanning of AMPP amides of the analytes are shown in the<br />

Supporting Information Supplementary Material Figure 6A-D.<br />

In all the spectra, some cleavage of the AMPP tag occurs giving<br />

rise to the peaks at m/z ) 169 and 183. To best ensure analytical<br />

specificity, we quantified the eicosanoids using a fragment in the<br />

analyte portion rather than in the AMPP tag. For PGD2, we used<br />

the m/z ) 307 fragment, which is likely due to cross-ring<br />

cleavage of the cyclopentanone ring and represents the most<br />

abundant non-tag fragment ion. For PGE2 and PGF2R, we used<br />

m/z ) 239 due to cleavage between C3 and C4. For AA and<br />

6-keto-PGF1R, we used m/z ) 239, again due to C3-C4<br />

cleavage. The major nontag fragment for TxB2 is m/z ) 337<br />

due to cross-ring cleavage. For all HETE species, we could<br />

identify a unique fragment ion (see Table 1) to avoid cross<br />

contamination of MS/MS signals due to partially unresolved LC<br />

peaks. Finally, for LTB4, we chose m/z ) 323, which is the most<br />

abundant non-tag fragment ion but has a nonobvious origin.<br />

Next we studied a mixture of eicosanoids in phosphate buffered<br />

saline and optimized pre-ESI-MS/MS sample workup and LC-ESI-<br />

MS/MS conditions. Critical for a successful method with ultrasensitive<br />

analyte detection is high yield recovery of analytes from<br />

the sample prior to LC-ESI-MS/MS analysis. We found that solid<br />

phase extraction with a rapidly wettable matrix (Oasis HLB<br />

cartridges, Waters Inc. or Strata-X cartridges, Phenomenex)<br />

combined with analyte elution with methanol gave high yield<br />

recovery of all eicosanoids analyzed (Table 1). The data in Table<br />

1 is based on 50 pg of each analyte submitted to recovery studies.<br />

When the same recovery studies were carried out with 5 pg of<br />

eicosanoid, results were essentially identical (data not shown).<br />

Inferior recoveries were obtained using elution of the solid phase<br />

cartridges with acidified methanol or if the sample was subjected<br />

to liquid-liquid extraction or protein precipitation using various<br />

solvents (Supporting Information). Figure 2A,B shows a typical<br />

LC-ESI-MS/MS run of an eicosanoid mixture. All of the analytes<br />

examined elute from the LC column in less than 11 min. All<br />

HETEs except 11(S)-HETE and 12(S)-HETE are baseline resolved.<br />

The partial overlap of the 11(S)- and 12(S)-HETEs is not a concern<br />

because completely selective fragment ions are being monitored.<br />

The LC elution profile and peak shape for both the derivatized<br />

and underivatized analytes (not shown) are similar. This shows<br />

that the presence of the quaternary ammonium group of the<br />

AMPP tag has no negative effect on the LC. TxB 2 gives rise to<br />

the typical broad peak shape due to interconversion between<br />

the two hemiacetal isomers. Also similar to underivatized<br />

analysis, the fragments used to monitor PGD2 and PGE2<br />

(isobars of each other) are not completely unique to each<br />

species, but baseline LC resolution of these two eicosanoids<br />

resolves the issue. The same is true for the isobaric pair TxB2<br />

and 6-keto-PGF1R.<br />

To ensure high yield conversion of eicosanoids to AMPP<br />

amides with minimal formation of N-acyl-ureas, we used a large<br />

excess of derivatization reagents. The AMPP tag and the EDCI<br />

coupling reagent elute in the void volume (not shown), and the<br />

HOBt elutes in the large time window between PGD2 and LTB4<br />

(not shown). Thus removal of excess derivatization reagents<br />

is not necessary prior to LC-ESI-MS/MS.<br />

Standard curve analysis was performed starting below the limit<br />

of quantification up to 2 orders of magnitude above this limit. In<br />

Table 2. Coefficient of Variation (%) for LC-ESI-MS/MS<br />

Analysis of Eicosanoid AMPP Amides<br />

intrasample a<br />

intersample<br />

(10 µL PBS) b<br />

intersample<br />

(10 µL serum) c<br />

6-keto-PGF1〈 4.9 16.9 not detected<br />

TxB2 2.4 3.0 3.7<br />

PGF2 〈 8.1 6.5 6.4<br />

PGE2 8.0 9.2 not detected<br />

PGD2 4.5 8.3 not detected<br />

LTB4 7.3 17.2 7.3<br />

5(S)-HETE 3.3 2.8 6.4<br />

8(S)-HETE 6.7 20.3 13.7<br />

11(S)-HETE 5.5 12.2 8.8<br />

12(S)-HETE 5.2 10.7 6.3<br />

15(S)-HETE 4.0 5.8 12.4<br />

arachidonic acid 12.7 28.2 7.9<br />

a Coefficient of variation for the analysis of the same sample of 1.25<br />

pg/µLofeicosanoidstandardsinjectedthreetimesontotheLC-ESI-MS/MS.<br />

b Coefficient of variation for the analysis of three independent samples<br />

of 1.25 pg/µL of eicosanoid standards (50 pg of standards spiked into<br />

phosphate buffered saline and worked up as described in the<br />

Experimental Methods section for LC-ESI-MS/MS.). c Coefficient of<br />

variation for three separate serum samples spiked with internal<br />

standards only.<br />

all cases, the mass spectrometry response was linear from the<br />

limit of quantification up to the highest amount analyzed (19 pg<br />

on-column for the Waters Quattro Micro and 5-10 pg on-column<br />

fortheWatersQuattroPremier)(SupplementalMaterialFigure2A,B<br />

in the Supporting Information). We define limit of quantification<br />

as the amount of eicosanoid AMPP amide needed to give an ESI-<br />

MS/MS signal that is 10-fold the noise (as determined using the<br />

Waters MassLynx software). The limits of quantification for all<br />

eicosanoid AMPP amides are listed in Table 1. The values using<br />

the Waters Quattro Micro are 5 pg for all eicosanoids except 12(S)-<br />

HETE (10 pg) and AA (10 pg). With the Waters Quattro Premier,<br />

the values are 0.2-0.5 pg for all eicosanoids except 12(S)-HETE<br />

(0.9 pg) and AA (1 pg). We also determined limit of quantification<br />

of underivatized eicosanoids analyzed by LC-ESI-MS/MS in<br />

negative ion mode. The values are shown in Table 1 for the Waters<br />

Quattro Micro and the Waters Quattro Premier. The new method<br />

using AMPP amides is found to be 10- to 20-fold more sensitive<br />

on both instruments. This constitutes a significant improvement<br />

in detection sensitivity of eicosanoids.<br />

To validate the above stated estimates of the limit of quantification,<br />

we injected eicosanoids standard mixtures six times on the<br />

Waters Quattro Premier, where each analyte was injected at<br />

roughly twice the estimated limit of quantification. Values of CV<br />

for analyte to internal standard peak ratios for each analyte are<br />

listed in Table 1. These CVs validate our estimates for the limits<br />

of quantification by demonstrating that eicosanoid analysis can<br />

be performed with reasonable reproducibility proximal to these<br />

stated values.<br />

Next, we more thoroughly investigated the reproducibility of<br />

the AMPP derivatization method for eicosanoid analysis. The<br />

results are summarized in Table 2. Three repetitive LC-ESI-MS/<br />

MS analyses of an identical sample (19 pg on-column levels of<br />

each eicosanoid) gave coefficients of variation in the range of the<br />

2.4-12.7% range. We also prepared three independent mixtures<br />

of eicosanoids (50 pg each) in phosphate buffered saline, submitted<br />

each to sample workup, and analyzed each by LC-ESI-MS/<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6795


Table 3. Eicosanoid Levels (pg/µL) in Mouse Serum and Bronchial Epithelial Cells a<br />

serum (µL) TxB2 PGE2 PGD2 PGF2〈 LTB4 5(S)-HETE 8(S)-HETE 11(S)-HETE 12(S)-HETE 15(S)-HETE<br />

Arachi-donic<br />

Acid<br />

1 46 15 198 748 393 437 11 345 240 28 623<br />

1 49 11 340 1075 545 646 17 840 703 23 246<br />

5 43 13 174 792 490 350 14 116 617 14 900<br />

5 43 11 169 986 469 350 16 512 672 16 631<br />

10 51 9 255 971 586 536 15 108 749 12 200<br />

10 39 13 248 958 593 416 14 487 732 15 476<br />

bronchial<br />

epithelial<br />

cellsb 4 (7) 105 (195) 25 (22) 170 (300) 5 (19) 880 (1210)<br />

a Eicosanoids not listed in the table were not detected. b The first number is for unstimulated cells, and the number in parentheses is for<br />

stimulated cells. Values are for 0.47 million cells per sample stimulated in 200 µL of Hank’s balanced salt solution.<br />

MS (19 pg on-column) (Table 2). Coefficients of variation ranged<br />

from 2.8 to 28.2%. We also analyzed three independent serum<br />

aliquots (10 µL), and the coefficient of variation ranged from 3.7<br />

to 13.7%.<br />

We then analyzed the set of endogenous eicosanoids present<br />

in mouse serum in order to test the AMPP derivatization method<br />

on a complex biological sample. Representative LC-ESI-MS/MS<br />

traces are shown in the Supporting Information Supplementary<br />

Material Figure 7 for a relatively high abundant serum eicosanoid<br />

(TxB 2) and a relatively low abundant eicosanoid (PGF2R).<br />

Traces for the full set of eicosanoids and internal standards<br />

are shown in the Supporting Information Supplementary Material<br />

Figure 3. Eicosanoids levels are shown in Table 3 for multiple<br />

runs and with different volumes of serum analyzed.<br />

As another example of low-level eicosanoid analysis in a<br />

complex biological matrix, we analyzed eicosanoid formation in<br />

calcium ionophore stimulated primary bronchial epithelial cells.<br />

Results are summarized in Table 3, and selected ion traces for<br />

the full set of eicosanoids are shown as Supplementary Material<br />

Figure 4 in the Supporting Information.<br />

Finally, we cross-validated the AMPP method with a commercially<br />

available antibody-based assay kit. We analyzed PGE 2<br />

production by Rat 3Y1 cells and compared the results obtained<br />

via LC-ESI-MS/MS analysis of AMPP amides with those<br />

obtained from the EIA kit. Results are summarized in the<br />

Supporting Information Supplementary Material Figure 8.<br />

There is strong agreement between the two methods.<br />

CONCLUSIONS<br />

We developed a simple derivatization procedure to convert<br />

carboxylic acids to AMPP amides and showed that this significantly<br />

improves the sensitivity for detection of eicosanoids by<br />

LC-ESI-MS/MS. Antibody-based quantifications of eicosanoids<br />

6796 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

have a limit of quantification around 1 pg, and the method requires<br />

that each analyte be quantified in a single assay well. Thus, if 12<br />

eicosanoids are to be analyzed, one requires a sample containing<br />

∼12 pg of each species. The AMPP amide method disclosed in<br />

the current study can detect all 12 eicosanoid species in a single<br />

sample containing ∼0.3-1 pg of each eicosanoid and is thus more<br />

than an order of magnitude more sensitive than antibody based<br />

detection. Previously reported LC-ESI-MS/MS detection of<br />

underivatized eicosanoids in negative ion mode provide a limit of<br />

quantification in the 10-20 pg range 1–3 (see also our data in Table<br />

1), and thus our method is more than an order of magnitude more<br />

sensitive than these methods. We are currently studying the use<br />

of AMPP to improve the detection sensitivity for other oxygenated<br />

fatty acids including cysteinyl-leukotrienes, lipoxins, resolvins, and<br />

protectans. Analysis of the full spectrum of fatty acids should be<br />

feasible based on the results with AA reported in this study. This<br />

new method based on AMPP amides should find widespread use<br />

in the quantification of lipid mediator levels where sample supply<br />

is limited.<br />

ACKNOWLEDGMENT<br />

This work was supported by grants from the National Institutes<br />

of Health (Grant HL50040 to M.H.G. and Grant HL089215 to<br />

T.S.H.) and by grants from the National Science Foundation<br />

(Grants CHE-0750048 and CHE-0349595 to F.T.).<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review March 19, 2010. Accepted July 8,<br />

2010.<br />

AC100720P


Anal. Chem. 2010, 82, 6797–6806<br />

δ 13 C Stable Isotope Analysis of Atmospheric<br />

Oxygenated Volatile Organic Compounds by Gas<br />

Chromatography-Isotope Ratio Mass Spectrometry<br />

Brian M. Giebel,* Peter K. Swart, and Daniel D. Riemer<br />

University of Miami, Rosenstiel School of Marine and Atmospheric Science, 4600 Rickenbacker Causeway,<br />

Miami, Florida 33149<br />

We present a new method for analyzing the δ 13 C isotopic<br />

composition of several oxygenated volatile organic<br />

compounds (OVOCs) from direct sources and ambient<br />

atmospheric samples. Guided by the requirements for<br />

analysis of trace components in air, a gas chromatograph<br />

isotope ratio mass spectrometer (GC-IRMS)<br />

system was developed with the goal of increasing<br />

sensitivity, reducing dead-volume and peak band broadening,<br />

optimizing combustion and water removal, and<br />

decreasing the split ratio to the isotope ratio mass<br />

spectrometer (IRMS). The technique relies on a twostage<br />

preconcentration system, a low-volume capillary<br />

reactor and water trap, and a balanced reference gas<br />

delivery system. The instrument’s measurement precision<br />

is 0.6 to 2.9‰ (1σ), and results indicate that<br />

negligible sample fractionation occurs during gas sampling.<br />

Measured δ 13 C values have a minor dependence<br />

on sample size; linearity for acetone was 0.06‰ ng C -1<br />

and was best over 1-10 ng C. Sensitivity is ∼10 times<br />

greater than similar instrumentation designs, incorporates<br />

the use of a diluted working reference gas<br />

(0.1% CO2), and requires collection of >0.7 ng C to<br />

produce accurate and precise results. With this detection<br />

limit, a 1.0 L sample of ambient air provides<br />

sufficient carbon for isotopic analysis. Emissions from<br />

vegetation and vehicle exhaust are compared and show<br />

clear differences in isotopic signatures. Ambient<br />

samples collected in metropolitan Miami and the<br />

Everglades National Park can be differentiated and<br />

reflect multiple sources and sinks affecting a single<br />

sampling location. Vehicle exhaust emissions of ethanol,<br />

and those collected in metropolitan Miami, have<br />

anomalously enriched δ 13 C values ranging from -5.0<br />

to -17.2‰; we attribute this result to ethanol’s origin<br />

from corn and use as an additive in automotive fuels.<br />

Oxygenated volatile organic compounds (OVOCs) such as<br />

methanol, ethanol, acetaldehyde, and acetone are gases found<br />

throughout the troposphere that influence atmospheric chemistry<br />

in many ways. These compounds act as a source of radicals and<br />

a sink for the hydroxyl radical (OH), participate in tropospheric<br />

* Corresponding author. Fax: 305-421-4689. E-mail: bgiebel@rsmas.miami.edu.<br />

ozone formation, and are precursors to formaldehyde and CO. 1<br />

Mixing ratios for OVOCs are typically at the low parts per billion<br />

by volume (ppbv) level and depend on sampling location and<br />

season. 2-5 Most atmospheric OVOC measurements have reported<br />

information on ambient levels, source emission strengths, and flux<br />

rates, 2,3,5-9 with two individual OVOCs, methanol and acetone,<br />

receiving the majority of focus thus far.<br />

Methanol is the second most abundant organic gas in the<br />

atmosphere after methane and its global budget has been studied<br />

extensively. 3,10-12 Emissions from vegetation are the single largest<br />

source to the atmosphere and are estimated between 75-312 Tg<br />

year -1 . Other sources of methanol exist, including fossil fuel<br />

combustion, biomass burning, plant decay, and in situ atmospheric<br />

production via oxidation of methane. Combined, these<br />

sources are estimated at


estimated source strength of 40-95 Tg year -1 , primary biogenic<br />

emissions comprise 22%-40% of acetone’s presence in the<br />

atmosphere. 15 Whereas secondary sources, such as the photooxidation<br />

of nonmethane hydrocarbons (NMHCs), contribute<br />

the largest fraction and are estimated to account for 24%-96%<br />

of acetone’s total source strength. 10,14,15 Sinks for acetone consist<br />

of an oxidation mechanism by OH and direct photolysis, and<br />

acetone’s estimated atmospheric lifetime against each of these<br />

loss pathways is 60 days. 13,14<br />

The value of stable isotope measurements in the unique<br />

isotopic compositions which exist for individual sources and in<br />

the fractionation associated with atmospheric removal. The<br />

distinctions arise from isotope fractionation effects which occur<br />

during chemical and physical processes. There are two types of<br />

fractionation effects, kinetic effects and equilibrium effects. Both<br />

kinetic and equilibrium effects influence the isotopic composition<br />

of gas phase products produced during photosynthesis, fermentation,<br />

and secondary metabolic processes within vegetation. Kinetic<br />

effects, however, are principally responsible for isotopic compositions<br />

resulting from unidirectional removal processes where<br />

equilibrium is not achieved, as is the case of in situ chemical<br />

oxidation by OH or photolysis. The use of stable isotopes of<br />

carbon, oxygen, and hydrogen is a valuable tool to enhance our<br />

understanding of global patterns for OVOCs, a unique class of<br />

hydrocarbons that originate from not only primary biogenic and<br />

anthropogenic sources but also secondary sources. Collectively,<br />

the measured average isotopic composition of an OVOC in the<br />

atmosphere, coupled with isotopic characterization of its major<br />

sources and sinks, can be used as a means to elucidate the<br />

chemistry of these compounds within the atmosphere, their<br />

formation pathways, and reduce uncertainty in their budgets. An<br />

excellent review by Goldstein and Shaw contains more details on<br />

the use of stable isotopic data and its application to atmospheric<br />

budgets of VOCs. 16<br />

Herein, we restrict the discussion to variations in the stable<br />

isotopes of carbon, 13 C and 12 C. The stable isotopic composition<br />

of a sample is expressed as a ratio (R) of 13 Cto 12 C. During<br />

analysis, the ratio of a sample is compared to that of Vienna<br />

Pee Dee Belemnite (V-PDB) by way of a working reference<br />

gas using an isotope ratio mass spectrometer (IRMS). The ratio<br />

of the sample is reported in delta (δ) notation as a per mil (‰)<br />

difference of the sample compared to the reference: δ 13 C (‰)<br />

) ((RSAMPLE/RSTANDARD) - 1) × 10 3 . If a compound has a<br />

negative value, then it contains less 13 C than the standard and<br />

is said to be isotopically light (or depleted); if a compound has<br />

a positive value then it is isotopically heavy (or enriched).<br />

Application of stable isotopic techniques in atmospheric<br />

chemistry has focused on more abundant components of the<br />

atmosphere, notably, CO2, CH4, and CO. The use of isotopes<br />

led to an increased understanding of sources, sinks, and<br />

seasonal cycling of these compounds. 17-24 Rudolph and his<br />

(15) Mao, H.; Talbot, R.; Nielsen, C.; Sive, B. Geophys. Res. Lett. 2006, 33.<br />

(16) Goldstein, A. H.; Shaw, S. L. Chem. Rev. 2003, 103, 5025–5048.<br />

(17) Allison, C. E.; Francey, R. J. J. Geophys. Res. 2007, 112, DOI: 10.1029/<br />

2006JD007345.<br />

(18) Mak, J. E.; Manning, M. R.; Lowe, D. C. J. Geophys. Res. 2000, 105, 1329–<br />

1336.<br />

(19) Mak, J. E.; Kra, G. Chemosphere 1999, 1, 205–218.<br />

(20) Mak, J. E.; Kra, G.; Sandomenico, T.; Bergamaschi, P. J. Geophys. Res. 2003,<br />

101, 14415–14420.<br />

6798 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

colleagues are credited as the first to use gas chromatographyisotope<br />

ratio mass spectrometry (GC-IRMS) to analyze NMHCs<br />

in atmospheric gases. 25 These compounds are difficult to measure<br />

because their mixing ratios are low; they require separation from<br />

the atmospheric matrix and need to be combusted to CO2 before<br />

transfer to the IRMS. For this reason, analyses are performed<br />

using a gas chromatograph and combustion oven coupled to<br />

an IRMS operating under continuous flow conditions. Rudolph<br />

et al. sampled a variety of locales, ranging from pristine<br />

noncontaminated air to heavily contaminated air dominated by<br />

automobile emissions. 27 Results from their work showed clear<br />

distinctions between the δ 13 C signatures of NMHCs sampled<br />

at each location. The signatures reflected the local sources,<br />

meteorological conditions, and fractionation mechanisms acting<br />

at each site. Additional studies of NMHCs include studying<br />

specific sources, oxidation processes, and other locations. 26-28<br />

Applications of GC-IRMS techniques for analysis of OVOCs<br />

are restricted in scope and focus on one or two compounds. 29-39<br />

This is due to additional difficulties associated with sampling and<br />

analyzing OVOCs, including low mixing ratios, tendency for<br />

sample to become lost on sampling surfaces caused by OVOCs’<br />

polar nature, difficulty in selectively removing water and CO2 from<br />

the sample matrix without perturbing the OVOCs, and optimizing<br />

chromatographic separation in the presence of associated<br />

NMHCs. To circumvent some of these difficulties, derivatization<br />

has been used as a technique to make the polar compounds<br />

amenable to analysis. 29-31,36,37 Four studies of OVOCs<br />

using direct GC-IRMS are available, two for formaldehyde 34,35 and<br />

two for acetaldehyde. 32,39<br />

The technique presented here allows for direct conversion of<br />

several OVOCs to CO2 for measurement of δ 13 C from small<br />

sample volumes and a range of sample types, including<br />

(21) Quay, P.; Stutsman, J.; Wilbur, D.; Snover, A.; Dlugokencky, E.; Brown, T.<br />

Global Biogeochem. Cycles 1999, 13, 445–461.<br />

(22) Rice, A. L.; Gotoh, A. A.; Ajie, H. O.; Tyler, S. C. Anal. Chem. 2001, 73,<br />

4104–4110.<br />

(23) Tyler, S. C. In Stable Isotopes in Ecological Research; Rundel, P. W.,<br />

Ehleringer, J. R., Nagy, K. A., Eds.; Springer-Verlag: New York, 1989.<br />

(24) Tyler, S. C.; Rice, A. L.; Ajie, H. O. J. Geophys. Res. 2007, 112.<br />

(25) Rudolph, J.; Lowe, D. C.; Martin, R. J.; Clarkson, T. S. Geophys. Res. Lett.<br />

1997, 24, 659–662.<br />

(27) Tsunogai, U.; Yoshida, N.; Gamo, T. J. Geophys. Res. 1999, 104, 16033–<br />

16040.<br />

(26) McCauley, S. E.; Goldstein, A. H.; DePaolo, D. J. Proc. Natl. Acad. Sci.<br />

U.S.A. 1999, 96, 10006–10009.<br />

(28) Rudolph, J.; Czuba, E.; Huang, L. Geophys. Res. Lett. 2000, 27, 3865–3868.<br />

(29) Guo, S.; Wen, S.; Wang, X.; Sheng, G.; Fu, J.; Hu, P.; Yu, Y. Atmos. Environ.<br />

2009, 43, 3489–3495.<br />

(30) Guo, S.; Wen, S.; Wang, X.; Sheng, G.; Fu, J.; Jia, W.; Yu, Y.; Lu, H. Rapid<br />

Commun. Mass Spectrom. 2007, 21, 1809–1812.<br />

(31) Guo, S. J.; Wen, S.; Zu, G. W.; Wang, X. M.; Sheng, G. Y.; Fu, J. M. Chin.<br />

J. Anal. Chem. 2008, 36, 19–23.<br />

(32) Jardine, K.; Karl, T.; Lerdau, M.; Harley, P.; Guenther, A.; Mak, J. E. Plant<br />

Biol. 2008, 9999.<br />

(33) Keppler, F.; Kalin, R. M.; Harper, D. B.; McRoberts, W. C.; Hamilton, J. T. G.<br />

Biogeosciences 2004, 1, 123–131.<br />

(34) Rice, A. L.; Quay, P. Environ. Sci. Technol. 2009, 43, 8752–8758.<br />

(35) Rice, A. L.; Quay, P. D. Anal. Chem. 2006, 78, 6320–6326.<br />

(36) Wen, S.; Feng, Y. L.; Yu, Y. X.; Bi, X. H.; Wang, X. M.; Sheng, G. Y.; Fu,<br />

J. M.; Peng, P. A. Environ. Sci. Technol. 2005, 39, 6202–6207.<br />

(37) Wen, S.; Yu, Y. X.; Guo, S. J.; Feng, Y. L.; Sheng, G. Y.; Wang, X. M.; Bi,<br />

X. H.; Fu, J. M.; Jia, W. L. Rapid Commun. Mass Spectrom. 2006, 20, 1322–<br />

1326.<br />

(38) Yamada, K.; Hattori, R.; Ito, Y.; Shibata, H.; Yoshida, N. Geophys. Res. Lett.<br />

2009, 36.<br />

(39) Jardine, K.; Harley, P.; Karl, T.; Guenther, A.; Lerdau, M.; Mak, J. E.<br />

Biogeosciences 2008, 5, 1559–1572.


Figure 1. Schematic view of the preconcentration system and GC-IRMS. The GC-IRMS relied on a low-volume capillary reactor, water trap,<br />

and a balanced reference gas delivery system (WRG, working reference gas).<br />

emissions from vegetation and automobiles and rural and urban<br />

ambient air. The system consists of a two-stage trapping<br />

procedure using a carbon sorbent followed by cryofocusing. A<br />

low-volume capillary reactor, water trap, and open split combined<br />

with high-speed chromatography enhance sensitivity and<br />

improve peak resolution.<br />

EXPERIMENTAL SECTION<br />

Reference and Calibrant Gases. A working CO2 reference<br />

gas calibrated against the V-PDB carbonate standard was used<br />

for determination of the 13 C/ 12 C ratio of OVOCs using IRMS.<br />

To achieve proper instrument response, a 0.1% CO2 working<br />

reference was prepared and calibrated. A cylinder of research<br />

grade CO2 was used to create a gas phase subsample, from<br />

which, a dilution of 0.1% CO2 was made in helium. The isotopic<br />

values for pure CO2, its subsample, and the 0.1% CO2 dilution<br />

were measured using a Finnigan MAT 251 IRMS with conventional<br />

dual inlet and a Europa Scientific 20-20 IRMS linked to<br />

an automated nitrogen carbon analyzer (ANCA) under continuous<br />

flow. Henceforth, the 0.1% CO2 gas will be referred to as<br />

working reference gas.<br />

The primary calibrant gas mixture used for testing and<br />

development was prepared gravimetrically. 40 Liquid compounds<br />

with purity g98% were obtained from Sigma-Aldrich and included<br />

methanol, ethanol, propanal, acetone, methyl ethyl ketone, 2-pentanone,<br />

and 3-pentanone. Microliter volumes of the compounds<br />

were discharged directly into a clean, evacuated aluminum<br />

cylinder (Luxfer Gas Cylinders, Riverside, CA) with a stainless<br />

steel valve (Ceoedux, Mount Pleasant, PA). The cylinder was<br />

pressurized with high-purity BIP Technology N2 gas (all compressed<br />

gases utilized in this work were obtained from Airgas<br />

South, Keenesaw, GA) to achieve desired mixing ratios for all<br />

components. The mixing ratio and purity of the mixture were<br />

(40) Apel, E. C.; Calvert, J. G.; Greenberg, J. P.; Riemer, D.; Zika, R.; Kleindienst,<br />

T. E.; Lonneman, W. A.; Fung, K.; Fujita, E. J. Geophys. Res. 1998, 103,<br />

22281–22294.<br />

verified via GC-flame ionization detector (FID) and GC/mass<br />

spectrometry (MS). Final mixing ratios for each individual<br />

component fell between 365 ppbv (2-pentanone) and 931 ppbv<br />

(methanol) with an associated error of 5% for each component.<br />

The calibration gas mixture was diluted further by dynamic<br />

dilution into moist zero air when needed. 41<br />

Low-pressure single and multicomponent gases were made<br />

similarly to the calibrant gas mixture except an evacuated<br />

electropolished stainless steel canister (Entech Instruments, Simi<br />

Valley, CA) was used and pressurized with 55 pounds per square<br />

inch absolute (psia) UHP helium for dilution. Mixing ratios for<br />

the single mixed gases were selected to equal that of the working<br />

reference gas.<br />

Elemental Analyzer. A Europa Scientific 20-20 IRMS linked<br />

to an automated nitrogen carbon analyzer (ANCA) was used to<br />

obtain the δ 13 C signature of each raw liquid compound used to<br />

prepare the calibrant gas mixture. Microliter volumes of liquid<br />

reagents were loaded on a sorbent (Chromosorb W, Advanced<br />

Minerals, Goleta, CA) contained within a small tin capsule. The<br />

capsule was folded over itself to prevent sample loss and<br />

quickly dropped into the furnace to begin the analysis. Raw<br />

compounds were compared against a working lab standard of<br />

pure CO2 calibrated against V-PDB.<br />

Preconcentration System. A custom preconcentration system<br />

was installed on the inlet of the gas chromatograph and included<br />

two manual stainless steel six-port rotary valves (1/16 in. fitting,<br />

0.40 mm i.d.; all rotary valves in this work were obtained from<br />

Vici Valco Instruments, Houston, TX) connected in series. Fused<br />

silica-lined stainless steel tubing joined the entire precononcentration<br />

system (1.59 mm o.d., 0.51 mm i.d.; Silcosteel tubing, all<br />

Silcosteel tubing in this work were obtained from Restek, Bellefonte,<br />

PA). A schematic of the complete analytical design appears<br />

in Figure 1.<br />

(41) Apel, E. C.; Hills, A. J.; Lueb, R.; Zindel, S.; Eisele, S.; Riemer, D. D. J.<br />

Geophys. Res. 2003, 108.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6799


The first rotary valve (RV1) used an injection loop (14.3 cm ×<br />

1.6 mm o.d. × 1.02 mm i.d.; Silcosteel tubing) yielding a volume<br />

of 0.12 cm 3 . Rotary valve 1 allowed for analysis of working<br />

reference gas for leak check purposes and single-component<br />

gases in helium to test the system with individual OVOCs.<br />

Sample flow rates through RV1 were set at 3 cm 3 min -1 and<br />

regulated upstream by a 0-100 standard cubic centimeters per<br />

minute (sccm) mass flow controller (MFC, no. 1) (all MFCs<br />

used in this work were obtained from Unit/Celerity, San Jose,<br />

CA).<br />

The second valve (RV2) allowed for loading of the multicomponent<br />

calibrant gas during development tests and samples for<br />

analysis. The calibrant gas or samples connected to a simple valve<br />

system giving the operator ease of on/off control and an additional<br />

helium input used to purge the sample loop. Flow rates through<br />

RV2 were 50 cm 3 min -1 and controlled by an oil pump (Alcatel<br />

Vacuum Products, Smyrna, GA) with a liquid nitrogen vapor<br />

trap and a 0-200 sccm MFC (no. 2) located downstream. The<br />

sampling loop of RV2 included a carbon adsorbent trap which<br />

is nonretentive for the major components of air (N2, O2, CO2,<br />

and H2O) while trapping OVOCs. The carbon adsorbent trap<br />

consisted of a section of Silcosteel (30.5 cm × 3.2 mm o.d. ×<br />

2.1 mm i.d) packed with 24.6 cm of Carboxen 1016 (Supleco,<br />

St. Louis, MO). Carboxen 1016 retained all the analytes of<br />

interest, completely desorbed them when heated, and contributed<br />

negligible amounts of CO2 during the desorption process.<br />

Typically, 1.0 L volumes were sampled over 20 min. Helium<br />

purged the sampling loop and adsorbent trap (50 cm 3 min -1 ) for<br />

5 min to remove unwanted gases and water that may remain<br />

after loading. After the purge procedure, RV2 was manually<br />

switched allowing carrier gas (0.7 cm 3 min -1 ) from the GC to<br />

back flush the adsorbent trap for 10 min. During this step, the<br />

adsorbent trap was desorbed with resistive heating at 200 °C<br />

using high temperature heater tape, thermocouple, and a<br />

CN76000 temperature controller (Omega Engineering, Stamford,<br />

CT).<br />

Desorbed material was cryo-focused as it passed through an<br />

open loop trap (46 cm, 0.56 mm o.d., 0.28 mm i.d., Siltek/Sulfinert<br />

tubing, Restek, Bellefonte, PA) contained in liquid nitrogen. Before<br />

injection, RV2 was switched to remove the carbon adsorbent trap<br />

from the carrier gas flow path. The cryo-focuser was heated with<br />

boiling water to inject the compounds onto the GC column.<br />

GC-IRMS System. An Agilent 6890 gas chromatograph was<br />

fitted with a DB-624 capillary chromatography column (20 m ×<br />

0.18 mm i.d. × 1 µm film thickness, 6% cyanopropylphenyl/94%<br />

dimethyl polysiloxane, Agilent Technologies, Santa Clara, CA).<br />

Ultrahigh purity helium entering the GC and IRMS passed<br />

through liquid nitrogen as a final measure to scrub it of possible<br />

contaminants, including trace CO2. The GC operated in constant<br />

flow mode with a nominal carrier flow rate set to 0.7 cm 3 min -1 .<br />

Carrier flow was measured at the open split located immediately<br />

upstream of the IRMS ion source. Temperature programming<br />

for the GC oven started at 30 °C for 3 min followed by a ramp<br />

of 10 °C min -1 and held at a final temperature of 125 °C for 7.5<br />

min. The total GC run time was 23 min.<br />

A capillary reactor used for combusting OVOCs to CO2 was<br />

constructed of fused silica tubing and contained two catalytic<br />

wires. One was an alloy blend consisting of Cu/Mn/Ni (84%/<br />

6800 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

12%/4%, 0.1 mm diameter, Alfa-Aesar, Ward Hill, MA), the other<br />

is Pt (99.99%, 0.1 mm diameter, Sigma-Aldrich, St. Louis, MO).<br />

Details about the fabrication and conditioning of the reactor<br />

are provided in the Supporting Information.<br />

Trace amounts of H2O, created with CO2 during the combustion<br />

process, can interfere by contributing signal at m/z -45<br />

and -46 due to protonation of CO2 within the ion source. A<br />

continuous length of deactivated fused silica (∼1.5 m × 0.36<br />

mm o.d. × 0.25 mm i.d., Supelco, St. Louis, MO) was connected<br />

to the downstream end of the capillary reactor and served as<br />

a transfer line, water trap, and link to the open split. At the<br />

midpoint of the transfer line a single loop (∼10-15 cm in<br />

length) created a water trap when immersed in a dry ice-ethanol<br />

slurry (-55 to -65 °C) and prevented water from reaching<br />

the ion source. Sample derived CO 2 continued to the outlet<br />

located in the open split. The looped trap was removed from<br />

the slurry after every run, and condensed water was driven off<br />

with helium carrier gas and by warming the loop to 100 °C for<br />

several minutes.<br />

A custom designed open split (Figure 1) was positioned<br />

adjacent to the ion source of a Europa Scientific GEO 20-20<br />

IRMS. 42 The open split resided within a chamber that housed four<br />

distinct outlets: two primary outlets for (1) the GC effluent and<br />

(2) the helium blank/0.1% CO2 reference gas; two secondary<br />

outlets for (1) tuning the IRMS ion beams and (2) supplying<br />

the chamber assembly with an inert atmosphere of helium<br />

(MFC no. 4, 10 cm 3 min -1 ). Details of the open split and<br />

working reference gas delivery system are contained in the<br />

Supporting Information. A length of deactivated fused silica (∼3<br />

m × 0.36 mm o.d., 0.10 mm i.d. SGE, Austin, TX) produced a<br />

capillary leak to the ion source and was manually moved between<br />

the various outlets located in the open split as needed.<br />

Europa Scientific GEO 20-20 IRMS. The IRMS was tuned<br />

daily with working reference gas and operated with an accelerating<br />

voltage of 2.5 kV. The ion trap and filament emission currents<br />

were set to approximately 600 and 1300 µA, respectively. The<br />

vacuum system of the IRMS created a 0.2 cm 3 min -1 flow rate<br />

through the restriction capillary leading to the ion source while<br />

maintaining an operating source pressure of 4.5 × 10 -6 Torr.<br />

When combined, the carrier flow rate (0.7 cm 3 min -1 ) exiting<br />

the GC outlet and the ion source flow (0.2 cm 3 min -1 ) resulting<br />

from the capillary leak produced a split ratio of 3.5 and reduced<br />

the amount of carbon transferred to the IRMS by approximately<br />

70%.<br />

Data Collection and Analysis. Data for time (t) and mass to<br />

charge ratios (m/z) 44, 45, and 46 were recorded at 1.0 Hz to a<br />

PC using acquisition software provided by the IRMS manufacturer.<br />

Three separate third party software packages were used (i.e.,<br />

Thermo Grams AI/8.0, Origin Lab Origin Pro 8, and Microsoft<br />

Excel 2007) for data analysis. Details of the data manipulation are<br />

described in the Supporting Information.<br />

Source and Ambient Sampling. Two types of sources were<br />

investigated in this study, natural emissions released from five<br />

tropical plants and exhaust from an automobile without a catalytic<br />

converter. Plant emissions were studied using static enclosure<br />

methods with (1) clipped and (2) intact vegetation. For the (1)<br />

clipped vegetation experiments, a branch was physically removed<br />

(42) Sacks, G. L.; Zhang, Y.; Brenna, J. T. Anal. Chem. 2007, 79, 6348–6358.


from the specimen and enclosed in a Teflon sampling bag with a<br />

Nupro valve fitting (SS-4H, Swagelok Co., Solon, OH). Hydrocarbonfree<br />

zero-air flushed the enclosed branch before being zip-tied<br />

shut and returned to direct sunlight for a period of 20-60 min.<br />

After this period, the bag and branch were brought into the lab<br />

and connected to the preconcentration system via the gas<br />

manifold. Sample volumes between 100-200 cm 3 were loaded<br />

using the RV2 loop and carbon sorbent. For (2) intact branch<br />

experiments, a single branch on the specimen was enclosed<br />

with the Teflon bag assembly and flushed with zero-air before<br />

being zip-tied shut and left in ambient light for a period 20-60<br />

min. Instead of clipping the branch and returning the bag to<br />

the laboratory, an evacuated canister was connected to the<br />

bag’s valve and its contents were removed. As with the clipped<br />

branch, sample volumes between 100-200 cm 3 were loaded<br />

for analysis.<br />

Automobile emissions from the right exhaust bank of a 1972<br />

International Scout were collected. The Scout lacks a catalytic<br />

converter and was operated at constant cruising and load conditions<br />

(∼2000 rpm, ∼80 kph) during sample collection. A piece of<br />

stainless steel tubing (∼1.5 m length) was secured to the inside<br />

of the passenger side exhaust and brought to the inside of the<br />

passenger compartment for connection to an evacuated canister.<br />

Samples were at atmospheric pressure after collection.<br />

Ambient whole air samples were collected at three locations<br />

in South Florida. Everglades National Park was chosen to<br />

represent a rural atmosphere dominated more by biogenic sources<br />

than anthropogenic sources. A partially enclosed lower roadway<br />

of Miami International Airport was sampled to ascertain a<br />

collective measure of concentrated and fresh vehicular emissions;<br />

Miami’s Financial District represented a more semiurban environment.<br />

Samples from the Everglades were pressurized to approximately<br />

30 pounds per square inch gauge (psig) by a DC<br />

powered metal bellows pump (MB-302, Senior Flexonics, Sharon,<br />

MA) into an evacuated canister; airport and financial district<br />

samples were collected at atmospheric pressure. Sample volumes<br />

for analysis ranged between 100-200 cm 3 for the airport and<br />

financial district to 1.0 L for Everglades National Park.<br />

RESULTS AND DISCUSSION<br />

Guided by the requirements for analysis of trace components<br />

in air, a GC-IRMS system was developed with the goal of analyzing<br />

OVOCs in ambient air. This included developing a preconcentration<br />

stage, increasing sensitivity, reducing dead-volume and peak<br />

band broadening, optimizing combustion and water removal, and<br />

decreasing the split ratio to the IRMS. An extended discussion<br />

focusing on these developments, including images of the working<br />

reference gas delivery system and open split, are available in the<br />

Supporting Information. These developments allowed for analysis<br />

of several OVOCs in ambient air at mixing ratios representative<br />

of a large portion of the troposphere ranging from urban to<br />

remote. Figure 2 shows the range of mixing ratios present in the<br />

atmosphere for methanol, ethanol, acetone, and MEK and the<br />

corresponding nanograms of C transmitted to the ion source with<br />

this analytical technique. 3,6,9,12,14,43 All but the most remote<br />

portions of the atmosphere can be investigated using this method.<br />

Reference Gas Analyses. Working Reference Gas. The diluted<br />

working reference gas was prepared carefully in a series of steps<br />

in order to monitor possible fractionation during the dilution<br />

Figure 2. Range of atmospheric OVOC mixing ratios expected in<br />

large portions of the troposphere and their relation to carbon<br />

transmitted to the IRMS ion source. The shaded area corresponds<br />

to the method detection limit.<br />

process. Results and further discussion of the preparation procedure,<br />

including an interlab comparison performed with Stony<br />

Brook University, can be found in the Supporting Information<br />

(Working Reference Gas Analyses and Table S-1).<br />

System Diagnostic Leak Assessments. Rotary Valve 1 (RV1) was<br />

used for various diagnostics tests. One test evaluated the leak<br />

tightness of the complete sample path through the preconcentration<br />

system, gas chromatograph, reactor assembly, water trap,<br />

and open split delivery to the IRMS. We compared working<br />

reference gas injected at the preconcentration system to that of<br />

its normal injection site at the open split. In the absence of<br />

detectable leaks in the sample path, daily offsets of 0.3-0.5‰ were<br />

routinely observed. On occasions when a microcrack was suspected<br />

of breaching the capillary reactor, offsets >2‰ were<br />

observed. This test was performed daily to evaluate the integrity<br />

of the instrumentation. If the offsets were 0.3-0.5‰, the concurrent<br />

data for the day were corrected by the appropriate amount.<br />

If the offset exceeded 0.5‰, the capillary reactor was replaced.<br />

Dynamic Range and Linearity. A 6 L electropolished stainless<br />

steel bulb with a dip tube assembly served as an exponential<br />

dilution flask to test the dynamic range and linearity of the method<br />

and IRMS in the absence of combustion. The bulb contained a<br />

1% CO 2 mixture in helium made from the same subsampled<br />

CO2 used in the production of working reference gas. A diluant<br />

flow of helium entered the steel bulb through the dip tube at<br />

a rate of ∼125 cm 3 min -1 and the outflow was plumbed to RV1.<br />

The total analysis occurred over a 5.5 h period broken into 9<br />

segments with the introduction of working reference gas. The<br />

amount of carbon reaching the ion source was ∼0.1-80 ng.<br />

The δ 13 C values over this range are expressed as a difference<br />

of the measured (and corrected) exponentially diluted CO2 from<br />

the working reference gas’ accepted value and are displayed<br />

in Figure 3a. Of particular interest is the appearance of a positive<br />

offset, ∼0.46‰, from zero. Efforts were made to minimize<br />

fractionations during the gas transfers, and the gases were made<br />

from the same stock. Despite this effort, the offset still persisted.<br />

(43) Apel, E. C.; Emmons, L. K.; Karl, T.; Flocke, F.; Hills, A. J.; Madronich, S.;<br />

Lee-Taylor, J.; Fried, A.; Weibring, P.; Walega, J.; Richter, D.; Tie, X.;<br />

Mauldin, L.; Campos, T.; Weinheimer, A.; Knapp, D.; Sive, B.; Kleinman,<br />

L.; Springston, S.; Zaveri, R.; Ortega, J.; Voss, P.; Blake, D.; Baker, A.;<br />

Warneke, C.; Welsh-Bon, D.; de Gouw, J.; Zheng, J.; Zhang, R.; Rudolph,<br />

J.; Junkermann, W.; Riemer, D. D. Atmos. Chem. Phys. 2010, 10, 2353–<br />

2375.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6801


Figure 3. (A,B) Results for exponentially diluted (A) CO2 and (B) acetone samples. δ 13 C values are expressed as a difference of the diluted<br />

CO2 and acetone from the accepted value of the 0.1% working reference gas and acetone value obtained on the elemental analyzer. An offset,<br />

opposite in sign but almost equal magnitude, exists for CO2 (∼0.5‰) and acetone (∼-0.6‰). For CO2, this is thought to be the result of an<br />

ambient leak whereby atmospheric CO2 enters the system. For acetone, incomplete combustion within the capillary reactor may contribute to<br />

the observed negative offset.<br />

Table 1. Tabulated δ 13 C Values for OVOCs Used in This Work a<br />

elemental analyzer<br />

liquid compounds<br />

mean<br />

±1σ<br />

(‰)<br />

GC-IRMS<br />

single-component gas<br />

mean<br />

±1σ<br />

(‰)<br />

95%<br />

confidence<br />

(±‰)<br />

%<br />

error<br />

The positive offset for CO2 was likely a result of small<br />

amounts of ambient CO2 becoming entrained in the system<br />

and thereby enriching the measured δ 13 C value. For large<br />

sample sizes this appeared to have a minimal effect. However,<br />

the effect became magnified for sample sizes below 1 ng C,<br />

where deviations up to 2‰ are observed. The linearity over<br />

this range, determined by ordinary linear regression, was<br />

0.01‰ ng C -1 . This suggests that variation of the δ 13 C signature<br />

with sample size is negligible and requires 10 ng C to induce<br />

a change of 0.1‰. Accuracy and precision were best over the<br />

range of ∼0.2 to ∼20 ng C. Samples containing more than 20<br />

ng C and less that 0.2 ng C were enriched in 13 C.<br />

Liquid Compounds and Single-Component Gases. Lowpressure<br />

single-component gas samples entered RV1 at a rate of<br />

3cm 3 min -1 and were loop injected onto the GC-IRMS with no<br />

cryo-focusing before the chromatographic column. Ten injec-<br />

GC-IRMS<br />

low-pressure 7-component<br />

mean<br />

±1σ<br />

(‰)<br />

95%<br />

confidence<br />

(±‰)<br />

%<br />

error<br />

GC-IRMS<br />

high-pressure 7-component pooled GC-IRMS<br />

mean<br />

±1σ<br />

(‰)<br />

95%<br />

confidence<br />

(±‰)<br />

%<br />

error<br />

mean<br />

±1σ<br />

(‰)<br />

95%<br />

confidence<br />

(±‰)<br />

methanol –35.3 ± 0.1 –34.1 ± 0.1 0.1 3.5 –33.0 ± 0.1 0.1 6.5 –34.4 ± 2.8 2.1 2.5 –34.0 ± 1.6 0.7 3.7<br />

ethanol –29.2 ± 0.2 –27.8 ± 0.8 0.6 4.8 –26.2 ± 0.8 1.3 10.2 –26.5 ± 0.5 0.4 9.2 –26.6 ± 1.4 0.6 8.7<br />

propanal –32.8 ± 0.1 –31.9 ± 0.2 0.1 2.7 –35.0 ± 0.6 1.0 –6.9 –27.4 ± 0.8 0.6 16.5 –30.5 ± 2.9 1.2 6.9<br />

acetone –27.5 ± 0.2 –28.5 ± 0.7 0.5 –3.7 –27.9 ± 0.2 0.3 –1.4 –27.6 ± 0.2 0.1 –0.4 –28.0 ± 0.6 0.2 –1.7<br />

MEK –23.2 ± 0.2 –23.5 ± 0.9 0.6 –1.3 –25.5 ± 0.7 1.1 –10.1 –22.5 ± 0.3 0.2 –3.0 –23.5 ± 1.2 0.5 –1.2<br />

2-pentanone –25.0 ± 0.3 –26.1 ± 0.2 0.1 –4.3 –33.7 ± 1.1 1.8 –34.8 –28.4 ± 0.6 0.5 –13.6 –28.6 ± 2.8 1.1 –14.3<br />

3-pentanone –30.7 ± 0.2 –30.7 ± 0.6 0.4 –0.1 –34.3 ± 0.3 0.5 –11.6 –32.7 ± 0.7 0.5 –6.5 –32.3 ± 1.5 0.6 –5.1<br />

sample number n ) 3 n ) 10 n ) 4 n ) 9 n ) 23<br />

diluent He<br />

Instrument Variables<br />

He N2 sample loop RV1 RV1 RV2<br />

cryo-focus no yes yes<br />

carbon sorbent no no yes<br />

zero air dilution no no yes<br />

a Accuracy and precision is traced from the elemental analyzer through the final design of the GC-IRMS system. Different variables tested<br />

during each phase are listed. The system’s total precision was calculated between 0.6 and 2.9‰ when compared to the values obtained on the<br />

elemental analyzer.<br />

6802 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

%<br />

error<br />

tions of each gas were compared to six injections of working<br />

reference gas. Single-component gases tested the GC-IRMS<br />

instrumentation by comparison to the isotopic values obtained<br />

for the raw liquids on the elemental analyzer (ANCA). The<br />

values for the raw liquids served as the basis of all our<br />

comparisons. The calculated percent difference between the<br />

two measurements ranged between -0.1 and 4.8%, and the<br />

results are listed in Table 1.<br />

Acetone was suitable to test the dynamic range and linear<br />

response of the IRMS with the added step of combustion. Acetone<br />

was chosen as the analyte because it showed consistent and<br />

excellent reproducibility across all aspects of this study. The<br />

experimental design was identical to that of the CO2 experiment<br />

described previously but with one exception, the acetone<br />

mixture had a starting carbon equivalent of 0.1% before the<br />

diluent flow of helium was added. A 1% mixture of acetone was


Figure 4. (A) IRMS mass-44 chromatographic response for the high-pressure, seven-component gas. (B) An enlarged plot shows reasonable<br />

separation for all seven components. The first peak observed in part B is a combination of CO2 collected during the zero-air dilution of the<br />

calibrant gas and CO2 created by Carboxen 1016 during the desorption process. The peak immediately preceding methanol was identified as<br />

the OVOC acetaldehyde and is an artifact background from the adsorption trap.<br />

avoided for two reasons. First, ambient samples are not<br />

expected to be greater than 0.1%, and second, there is more<br />

concern for what happens to measured isotopic signatures as<br />

smaller sample concentrations are approached. The amount<br />

of carbon reaching the ion source was determined similarly to<br />

the CO2 test and ranged between ∼0.8 and 12 ng. The δ 13 C<br />

values over this range are expressed as a difference of the<br />

measured and corrected exponentially diluted acetone from<br />

the value obtained on the elemental analyzer (Figure 3b).<br />

The linearity over this range, determined by ordinary linear<br />

regression, was 0.06 ‰ ng C -1 . For acetone, this indicates<br />

that sample size can influence measured δ 13 C and that a<br />

change between 1 and 10 ng C can induce a noticeable shift<br />

of 0.6‰ in measured δ 13 C. Accuracy and precision were best<br />

over the range of 0.2-10 ng C.<br />

Also worth noting is the apparent negative offset for acetone,<br />

∼-0.56‰, compared to the positive offset for CO2, ∼0.46‰. Raw<br />

data for both experiments were corrected by 0.5‰ and 0.4‰<br />

for acetone and CO2, respectively. However, the offsets still<br />

exist. Entrainment of ambient CO2 did not appear to affect<br />

acetone because of its separation on the chromatographic<br />

column. The negative offset for acetone was likely related to<br />

incomplete combustion within the capillary reactor. Thermodynamic<br />

principles support 12 C being combusted before 13 C;<br />

thus, if combustion was incomplete, we would observe a lighter<br />

δ 13 C value.<br />

Calibrant Gas Analyses. Low-Pressure Seven-Component Gas<br />

Mixture. A low-pressure seven-component gas mixture in helium<br />

was used preliminarily to test chromatographic conditions in the<br />

absence of the carbon sorbent by using the RV1 loop (Table 1).<br />

This was a logical step between the use of single-component gases<br />

and a gravimetrically prepared, high-pressure, seven-component<br />

calibration gas in nitrogen. The low-pressure seven-component<br />

gas mixture flowed through the RV1 loop for 5 min prior to<br />

starting the analysis. The flow rate (3 cm 3 min -1 ) was maintained<br />

by MFC (no. 1) upstream of RV1. After the initial 5 min purge<br />

period, RV1 was manually switched and the gas within the<br />

injection loop was diverted through RV2 and cryogenically<br />

focused in liquid nitrogen for an additional 5 min before<br />

injection into the chromatographic column. Of particular note<br />

are the values obtained for 2- and 3-pentanone, which are<br />

depleted in 13 C compared to both the liquid compounds and<br />

the single component gas mixtures. This may indicate an<br />

unknown effect resulting from the analytical column. The<br />

percent error between this measurement technique and that<br />

performed on the elemental analyzer for the pure liquid<br />

compounds ranges between 1.4 and 35%.<br />

Gravimetric Seven-Component Gas Mixture. One of the main<br />

goals of this work was to develop a GC-IRMS system capable of<br />

measuring OVOCs over the dynamic range found in the atmosphere.<br />

To mimic ambient levels of these compounds in the<br />

atmosphere, the high-pressure calibrant gas was diluted into moist<br />

zero-air using a dynamic dilution system. Dilution produced<br />

mixing ratios between ∼18.6 ppbv (methanol) and 7.3 ppbv (2pentanone)<br />

for all components. The diluted calibrant was connected<br />

directly to the gas manifold (Figure 1). Using the range<br />

of mixing ratios produced after the high-pressure calibrant gas<br />

was diluted in zero-air (7.3-18.6 ppbv), the volume of air<br />

concentrated (1.0 L), and the open split dilution (∼30%), we<br />

calculated ∼2.5-5 ng C were delivered to the ion source for all<br />

components. Results for nine replicate analyses are presented in<br />

Table 1, and an example of the chromatographic response appears<br />

in Figure 4. Reasonable agreement exists for all seven components<br />

compared to the liquid reagents analyzed on the elemental<br />

analyzer; the margin of error between these two measurements<br />

ranged between 0.4 and 16.5%. The components with the two<br />

largest errors were propanal (16.5%) and 2-pentanone (13.6%). Both<br />

of these peaks are the leading peak in a pair (propanal/acetone<br />

and 2-penatnone/3-penatanone), and perhaps the later eluting<br />

compounds influence the measured δ 13 C values of the earlier<br />

compounds. This is supported by the observation that analysis<br />

of the single-component gases for the same compounds on the<br />

GC-IRMS had a lower error (


Table 2. δ 13 C Values for Compounds Emitted from Various Tropical Plants and a Fossil Fuel Combustion Source a<br />

sand live oak<br />

Quercus geminata<br />

orange<br />

Citrus sinensis<br />

lemon<br />

Citrus limon<br />

plant type<br />

Students’ t test, and the results show that differences between<br />

the raw material on the elemental analyzer and the gases on the<br />

GC-IRMS system are significant. However, the percent difference<br />

between the two procedures is small; less than 9% error exists<br />

between the pooled results and those of the raw liquid compounds<br />

for all compounds except 2-pentanone, which has an associated<br />

error of 14.3%.<br />

Source Measurement Results. The δ 13 C results from the<br />

automotive exhaust and tropical plants are presented in Table<br />

2. A representative chromatogram from the automotive exhaust<br />

sample is shown in the Supporting Information (Figure S-1) and<br />

serves as a good example of a complex sample matrix and the<br />

system’s peak resolution. We have also incorporated, in some<br />

instances, for both source and ambient samples, results for the<br />

OVOC acetaldehyde even though analysis of this compound is<br />

influenced by a variable artifact background. 9,41 Data for NMHCs,<br />

isoprene, benzene, and toluene are made known because they<br />

were easily identifiable in the chromatograms and offer a<br />

comparison to previously measured NMHCs δ 13 C values. 44-46<br />

In general, a clear distinction exists between plant and fossil fuel<br />

emissions, and as would be expected based on the range of δ 13 C<br />

values for oils and fuels worldwide, 16,46 fossil fuel derived OVOC<br />

emissions were more enriched in 13 C compared to the plant<br />

results.<br />

Of particular note are the substantially enriched δ 13 C values<br />

(-5.0 ± 0.4‰) for ethanol, compared to the other compounds<br />

in the fossil fuel combustion experiments. Ethanol, derived<br />

from corn, is currently being blended into gasoline in mixtures<br />

g10% (v/v). It has been reported that biofuels consume >25%<br />

of U.S. corn production and that ethanol constitutes 99% of all<br />

biofuels in the United States. 47,48 Utilizing C4 photosynthesis,<br />

which discriminates less against 13 C, corn and other C4 plants<br />

are generally enriched in the isotope compared to C3 plants.<br />

(44) Iannone, R.; Koppmann, R.; Rudolph, J. J. Atmos. Chem. 2007, 58, 181–<br />

202.<br />

(45) Rudolph, J.; Anderson, R. S.; Czapiewski, K. V.; Czuba, E.; Ernst, D.;<br />

Gillespie, T.; Huang, L.; Rigby, C.; Thompson, A. E. J. Atmos. Chem. 2003,<br />

44, 39–55.<br />

(46) Rudolph, J.; Czuba, E.; Norman, A. L.; Huang, L.; Ernst, D. Atmos. Environ.<br />

2002, 36, 1173–1181.<br />

(47) Barnett, M. O. Environ. Sci. Technol. 2010, 44, 5330-5331.<br />

(48) Farrell, A. E.; Plevin, R. J.; Turner, B. T.; Jones, A. D.; O’Hare, M.; Kammen,<br />

D. M. Science 2006, 311, 506–508.<br />

philodendron<br />

Philodendron selloum<br />

sea grape<br />

Coccoloba uvifera Keppler et al. 33<br />

fossil fuel<br />

combustion<br />

acetaldehyde -29.9 (2.3) -25.7 (0.1) -22.4 (1.4)* -17.5 (0.5) -30.7 (1.1)* -24.9 (2.2) -20.9 (0.4)<br />

methanol -41.9 (3.1) -59.7 (2.9) -37.8 (2.6)* -27.5 (0.5) -30.7 (1.0)* -68.2 (11.2) -16.9 (1.3)<br />

ethanol -41.5 (0.8) -37.5 (0.3) -30.6 (0.2) -36.5 (0.2)* -29.4 (2.6) -5.0 (0.4)<br />

propanal -25.6 (2.7)<br />

isoprene off scale b -26.9 (3.7) -35.2 (3.5) -33.8 (2.6) -23.0 (2.6)* -16.7 (1.2) -32.6 (0.9)*<br />

acetone -35.7 (4.1) -37.4 (2.4) -32.8 (1.2) -38.8 (1.1) -29.3 (1.5)* -33.8 (0.8) -31.3 (0.8)* -28.1 (2.5) -25.6 (0.5)<br />

2-pentanone -35.2 (1.4)<br />

benzene -26.9 (0.3)<br />

toluene -27.5 (0.6)<br />

a Also included are values for prepped and incubated biogenic samples from Keppler et al.. 33 All values are reported as the average (standard<br />

deviation). All samples n ) 5, except the fossil fuel source where n ) 3. All biogenic samples are wounded/clipped branches, except where noted<br />

(*), which represents an intact branch on the sample specimen. The fossil fuel source was collected from a 1972 Scout International with no<br />

catalytic converter at a constant cruise. b Isoprene was present; however, it saturated the detectors and the signal response was off scale and the<br />

δ 13 C value could not be calculated.<br />

6804 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Bulk carbohydrate analyses between the two plant types show<br />

an enrichment of ∼15‰ in carbohydrates extracted from C4<br />

plant material. 49 Investigations of industrially produced ethanol<br />

originating from corn have been shown to have δ 13 C values of<br />

-10.71 ± 0.31‰. 49 The values we observed in the Scout samples<br />

are ∼5‰ heavier and, considering the widespread use of<br />

ethanol (7.5 billion gallons are expected to be used in fuel by<br />

2012 48 ), may serve as a tracer for transportation related sources<br />

to the atmosphere.<br />

Some biogenic samples in this study, such as sand live oak<br />

and orange citrus, had substantially depleted values for methanol<br />

and agree with incubated emissions from various deciduous trees<br />

and grasses made by Keppler et al. 33 (Table 2). However, this<br />

observation is not consistent across all samples and suggest that<br />

variations in δ 13 C values may result from interspecies differences,<br />

microbe interaction on the leaf’s surface, prey/injury<br />

response, the potential presence of a methanol utilization<br />

pathway which oxidizes methanol to formaldehyde and formic<br />

acid/formate, 50,51 and other lesser known metabolic, formation,<br />

and loss pathways within plants. 52 Finally, a wound response may<br />

be observed between the clipped and intact philodendron and sea<br />

grape samples. In one distinct case, acetaldehyde emitted from<br />

clipped sea grape specimens were enriched by ∼4‰ compared<br />

to the fossil fuel emissions.<br />

Ambient Measurement Results. Considerable differences<br />

in δ 13 C are observed between ambient sampling locations<br />

(Table 3). Results from Miami International Airport are reflective<br />

of an averaged value for fresh vehicular sources. The measured<br />

δ 13 C range for airport samples is between -12.3 ± 3.7‰<br />

(ethanol) to -35.3 ± 1.7‰ (3-pentanone). With the exception<br />

of ethanol, which has a δ 13 C value consistent with its C4 plant<br />

source, and 2- and 3-pentanone, the measured range at the<br />

airport agrees with that established for NMHCs from transportation-related<br />

sources by Rudolph, namely, -21.9 to<br />

-31.3‰. 25 Our acetaldehyde value is consistent with the range<br />

(49) Ishida-Fujii, K.; Goto, S.; Uemura, R.; Yamada, K.; Sato, M.; Yoshida, N.<br />

Biosci., Biotechnol., Biochem. 2005, 69, 2193–2199.<br />

(50) Cossins, E. A. Can. J. Biochem. 1964, 42, 1793–1802.<br />

(51) Gout, E.; Aubert, S.; Bligny, R.; Rebeille, F.; Nonomura, A. R.; Benson, A. A.;<br />

Douce, R. Plant Physiol. 2000, 123, 287–296.<br />

(52) Fall, R. In Reactive Hydrocarbons in the Atmosphere; Hewitt, C. N., Ed.;<br />

Academic Press: San Diego, CA, 1999; pp 43-97.


Table 3. Ambient Measurement Results for Samples<br />

Collected from Metropolitan Miami and Everglades<br />

National Park a<br />

Miami<br />

International<br />

Airport<br />

δ 13 C ± 1σ (‰)<br />

Miami<br />

Financial<br />

District<br />

Everglades<br />

National<br />

Park<br />

acetaldehyde -26.7 ± 0.7 -26.8 ± 1.2 -19.0 ± 2.7<br />

methanol -36.3 ± 3.7<br />

ethanol -12.3 ± 3.7 -17.2 ± 4.1<br />

isoprene -30.3 ± 2.1<br />

propanal -28.4 ± 1.5 -26.2 ± 2.4<br />

acetone -31.0 ± 3.5 -26.6 ± 0.4 -23.7 ± 0.4<br />

MEK -28.3 ± 2.1 -25.9 ± 1.9<br />

2-pentanone -34.8 ± 6.5 -29.4 ± 0.1<br />

3-pentanone -35.3 ± 1.7 -37.8 ± 1.8<br />

toluene -33.7 ± 2.0<br />

a Miami International Airport, n ) 5; Miami financial district, n )<br />

4; Everglades National Park, n ) 3.<br />

presented by Wen et al. who measured values via a derivatization<br />

procedure of ∼-21.0‰ and ∼-29.2‰ for samples<br />

collected at a bus station and petrochemical refinery, respectively.<br />

36<br />

Some observations at Miami’s Financial District are between<br />

2.2 and 4.4‰ enriched in 13 C compared to the same compounds<br />

at Miami International Airport, and again we observe an<br />

anomalously enriched value for ethanol (-17.2 ± 4.1‰).<br />

Samples from the airport are general δ 13 C values we can expect<br />

for OVOCs from transportation related sources without addition<br />

from other sources and losses caused by solar radiation and<br />

reaction with OH. Miami’s Financial District is located within<br />

0.1 mile of Biscayne Bay and 1 mile of the Port of Miami and<br />

was dominated by an onshore breeze during the sample<br />

collection. Therefore, we can expect values from the financial<br />

district to be enriched since the δ 13 C signature for each<br />

compound will reflect a combination of vehicular, biogenic, and<br />

possibly marine sources and, additionally, losses attributable<br />

to reactivity with OH and photolysis. Isotopic values for samples<br />

from the financial district are bound within the reported range<br />

of -15.8 to -37.4‰ for NMHCs sampled at a moderately<br />

polluted waterfront in Wellington, New Zealand. 25<br />

In comparison with automobile exhaust (Table 2), the mean<br />

values observed at Miami International Airport and Miami’s<br />

financial district are generally depleted in 13 C. The two most<br />

obvious differences among these samples that may influence<br />

the observations are the fuel source and the presence of a<br />

catalytic converter. Emissions collected at the airport are a mix<br />

of refined petroleum and diesel, whereas the Scout International<br />

was fueled by unleaded gasoline. Furthermore, vehicle<br />

emissions at the airport are assumed to be produced by engines<br />

having a catalytic converter. However, the Scout lacked a<br />

converter, and the speeds of the engines producing the<br />

emissions were very different. Traffic through the airport’s<br />

lower roadway moved at an idle pace and rarely exceeded 15<br />

mph. The Scout samples were obtained with the engine under<br />

significant load and at a constant revolution per minute (2000<br />

rpm) and cruise speed (80 kph). To our knowledge, no studies<br />

exist showing how the presence of a catalytic converter or<br />

engine speed may influence the δ 13 C of emitted hydrocarbons.<br />

Samples from Everglades National Park spanned a large range<br />

from -19.0 to -36.3‰. Measured methanol from within the<br />

National Park was -36.3 ± 3.7‰, considerably depleted and<br />

consistent with other values obtained in the tropical plant<br />

enclosure studies (i.e., sand live oak δ13CMethanol )-41.9 ± 3.1‰)<br />

and with the results presented earlier from Keppler et al. 33<br />

Similarly, δ13C values for isoprene released from C3 plants<br />

range from -26 to -29‰. 45 Isoprene values at the National<br />

Park are lighter (-30.3 ± 2.1‰) than the range presented by<br />

Rudolph et al. However, when the precision of the measurement<br />

is considered, the isoprene values measured from the<br />

Everglades’ samples overlap the range observed with that<br />

previous work. Acetone and acetaldehyde values from within<br />

the National Park are more enriched than anticipated. The<br />

mean δ13C values for these compounds are -23.7‰ and<br />

-19.0‰, respectively. Each are enriched approximately 7.5‰<br />

compared to samples collected at Miami International Airport<br />

and are fairly consistent with samples from Miami’s financial<br />

district and fossil fuel combustion.<br />

When estimated atmospheric lifetimes (τ) are considered for<br />

these compounds in the troposphere for losses caused by<br />

OH OH reactivity with OH (τacetone ) 66 days; τacetaldehyde ) 11 h) and<br />

hν hν photolysis (τacetone ) 38 days; τacetaldehyde ) 5 days), these<br />

observations can be explained, especially for the enrichment<br />

of acetaldehyde over acetone (∼5‰). Few studies of ambient<br />

δ13C for acetaldehyde exist, 29,36 and only one exists for acetone. 37<br />

For samples collected within a biosphere reserve in China, Guo<br />

et al. measured acetaldehyde values between -31.6 and -34.9‰.<br />

These values are depleted in 13C compared to our measurements.<br />

However, they report weak photolytic loss of formaldehyde in<br />

the same study, and considering formaldehyde’s lifetime<br />

against photolysis is shorter (4 h) compared to acetaldehyde<br />

(5 days), we assume this to be true for acetaldehyde at the<br />

same location.<br />

Guo et al. used a derivatization method to calculate δ13C values<br />

for acetone collected at a forested site (∼-31‰) and at the<br />

top of a 10 m building influenced by vehicle emissions<br />

(∼-26‰). The acetone values from Everglades National Park<br />

are enriched by 2-7‰ compared to the values presented by<br />

Guo et al. Isotopic values obtained from the forest may reflect<br />

the signature of fresh acetone emissions from biomass, while<br />

values for Everglades National Park samples may be more<br />

strongly influenced by photochemistry. The measured values<br />

for acetone and acetaldehyde from within the Everglades may<br />

also indicate contributions from in situ atmospheric production<br />

via oxidation and photolysis of higher order hydrocarbons. An<br />

exact assessment to separate direct emissions from photochemical<br />

production and loss is not possible at this time since<br />

fractionations associated with these pathways are not known.<br />

CONCLUSIONS<br />

A new method for measuring δ 13 C values of low-molecular<br />

weight OVOCs from direct sources and ambient samples was<br />

developed. The method incorporated a carbon sorbent, a lowvolume<br />

capillary reactor, water trap, and balanced working<br />

reference gas delivery system. The method’s total precision<br />

ranged between 0.6 and 2.9‰, and negligible sample fractionation<br />

occurred while sampling and trapping gases. Further<br />

testing showed that measured δ 13 C values had little dependence<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6805


on sample size (0.06 ‰ ng C 1- ), and linearity was best over<br />

the range of 1-10 ng C. The method was sensitive, requiring<br />

>0.2 ng C into the ion source to produce accurate and precise<br />

results. The analysis of ambient samples required small sample<br />

volumes, with ∼1.0 L of gas providing sufficient carbon for<br />

analysis.<br />

Clear distinctions in δ 13 C were observed between emissions<br />

released from plants and automobiles. In particular, ethanol<br />

emissions from automotive exhaust and metropolitan Miami<br />

were significantly enriched in 13 C. This is related to ethanol’s<br />

C4 plant origin and use as a fuel additive. Ambient samples<br />

can be differentiated, but the variation in δ 13 C values was not<br />

as great as for the source samples. Ambient samples suffer<br />

from additional complexity with multiple sources and sinks<br />

affecting single sampling locations. Clearly, more studies of<br />

sources and ambient sampling are required to define and<br />

characterize OVOCs in the troposphere along with laboratory<br />

studies to determine the kinetic isotope effects associated with<br />

OVOCs’ in situ production and loss from reaction with OH and<br />

photolysis. As it stands now, this technique can be used to<br />

differentiate OVOC sources and to assess the carbon isotopic<br />

6806 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

values for OVOCs in ambient air. It should serve as a useful<br />

way to investigate transformations of organic gases in the<br />

atmosphere.<br />

ACKNOWLEDGMENT<br />

We thank Tom Brenna and Herbert Tobias for helpful discussions<br />

in developing this method and Rich Iannone for providing<br />

a template for raw data calculations. We acknowledge John Mak<br />

and Zhihui Wang for the working reference gas interlab comparison.<br />

We appreciate the efforts of Kevin Polk and his 1972<br />

International Scout. Finally, we gratefully acknowledge the helpful<br />

comments made by two anonymous reviewers and support<br />

provided by NSF Grant No. 0450939.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review March 23, 2010. Accepted July 6,<br />

2010.<br />

AC1007442


Anal. Chem. 2010, 82, 6807–6813<br />

Direct Voltammetric Analysis of DNA Modified with<br />

Enzymatically Incorporated 7-Deazapurines<br />

Hana Pivoňková, † Petra Horáková, †,‡ Miloslava Fojtová, †,§ and Miroslav Fojta* ,†<br />

Institute of Biophysics, v.v.i., Academy of Sciences of the Czech Republic, Královopolská 135,<br />

CZ-612 65 Brno, Czech Republic, Department of <strong>Analytical</strong> <strong>Chemistry</strong>, Faculty of <strong>Chemical</strong> Technology, University of<br />

Pardubice, Studentská 573, CZ-532 10 Pardubice, Czech Republic, and Department of Functional Genomics and<br />

Proteomics, Institute of Experimental Biology, Faculty of Science, Masaryk University, Kotlárˇská 2,<br />

CZ-611 37 Brno, Czech Republic<br />

Nucleic acids studies use 7-deazaguanine (G*) and 7-deazaadenine<br />

(A*) as analogues of natural purine bases<br />

incapable of forming Hoogsteen base pairs, which prevents<br />

them from being involved in DNA triplexes and<br />

tetraplexes. Reduced propensity of the G*- and/or A*modified<br />

DNA to form alternative DNA structures is<br />

utilized, for example, in PCR amplification of guanine-rich<br />

sequences. Both G* and A* exhibit significantly lower<br />

potentials of their oxidation, compared to the respective<br />

natural nucleobases. At carbon electrodes, A* yields an<br />

oxidation peak which is by about 200-250 mV less<br />

positive than the peak due to adenine, but coincides with<br />

oxidation peak produced by natural guanine residues. On<br />

the other hand, oxidation signal of G* occurs at a potential<br />

by about 300 mV less positive than the peak due to<br />

guanine, being well separated from electrochemical signals<br />

of any natural DNA component. We show that<br />

enzymatic incorporation of G* and A* can easily be<br />

monitored by simple ex situ voltammetric analysis of the<br />

modified DNA at carbon electrodes. Particularly G* is<br />

shown as an attractive electroactive marker for DNA,<br />

efficiently incorporable by PCR. While densely G*-modified<br />

DNA fragments exhibit strong quenching of fluorescence<br />

of SYBR dyes, commonly used as fluorescent<br />

indicators in both gel staining and real time PCR applications,<br />

the electrochemical detection provides G*-specific<br />

signal suitable for the quantitation of the amplified DNA<br />

as well as for the determination of the DNA modification<br />

extent. Determination of DNA amplicons based on the<br />

measurement of peak G* ox is not affected by signals<br />

produced by residual oligonucleotide primers or primary<br />

templates containing natural purines.<br />

Electrochemical techniques are increasingly applied in the area<br />

of nucleic acids sensing (reviewed in refs 1-4). Nucleic acids<br />

* To whom correspondence should be addressed. E-mail: fojta@ibp.cz .<br />

† Academy of Sciences of the Czech Republic.<br />

‡ University of Pardubice.<br />

§ Masaryk University.<br />

(1) Fojta, M. In Electrochemistry of Nucleic Acids and Proteins. Towards<br />

Electrochemical Sensors for Genomics and Proteomics; Palecek, E., Scheller,<br />

F., Wang, J., Eds.; Elsevier: Amsterdam, 2005, pp 386-431.<br />

(2) Fojta, M.; Jelen, F.; Havran, L.; Palecek, E. Curr Anal Chem 2008, 4, 250–<br />

262.<br />

possess intrinsic electroactivity due to the presence of electrochemically<br />

oxidizable or reducible nucleobases, 2,4 making it<br />

possible to analyze them electrochemically without any labeling.<br />

Indeed, various label-free electrochemical techniques were proposed<br />

for the detection of DNA damage 1 or DNA hybridization. 3<br />

In spite of these efforts, application of various redox indicators<br />

and labels proved useful particularly in sequence-specific DNA<br />

sensing requiring reliable discrimination between two complementary<br />

strands (e.g., target DNA and hybridization probe 5,6 ),<br />

specific determination of newly synthesized pieces or fragments<br />

of DNA (in primer extension or PCR-based assays 7-9 ) or even<br />

identification of a single nucleobase incorporated at a specific<br />

position (in SNP typing 9-11 ). Generally, introducing electroactive<br />

tags producing “new” specific electrochemical responses (not<br />

yielded by natural DNA components) increases specificity of the<br />

assays considerably. Electrochemically active moieties can be<br />

incorporated into nucleic acids during chemical oligonucleotide<br />

synthesis, postsynthetically by chemical modification of “natural”<br />

nucleic acids 5,6 or using modified nucleoside triphosphates<br />

(dNTPs) and DNA polymerases. 7-12 The latter approach represents<br />

a versatile way to facile construction of labeled or otherwise<br />

functionalized nucleic acids and to efficient sequence-specific DNA<br />

sensing. 12 A critical prerequisite for these applications is the ability<br />

of a DNA polymerase to use modified dNTPs as substrates for<br />

efficient incorporation without losing sequence-specificity. C7substituted<br />

7-deazapurines and C5-substituted pyrimidines are<br />

usually acceptable substrates for (at least some) DNA polymerases<br />

(3) Palecek, E.; Fojta, M. Talanta 2007, 74, 276–290.<br />

(4) Palecek, E.; Jelen, F. In Electrochemistry of Nucleic Acids and Proteins.<br />

Towards Electrochemical Sensors for Genomics and Proteomics.; Palecek, E.,<br />

Scheller, F., Wang, J., Eds.; Elsevier: Amsterdam, 2005, pp 74-174.<br />

(5) Flechsig, G. U.; Reske, T. Anal. Chem. 2007, 79, 2125–2130.<br />

(6) Fojta, M.; Kostecka, P.; Trefulka, M.; Havran, L.; Palecek, E. Anal. Chem.<br />

2007, 79, 1022–1029.<br />

(7) Brazdilova, P.; Vrabel, M.; Pohl, R.; Pivonkova, H.; Havran, L.; Hocek, M.;<br />

Fojta, M. Chem.-Eur. J. 2007, 13, 9527–9533.<br />

(8) Patolsky, F.; Weizmann, Y.; Willner, I. J. Am. Chem. Soc. 2002, 124, 770–<br />

772.<br />

(9) Vrabel, M.; Horakova, P.; Pivonkova, H.; Kalachova, L.; Cernocka, H.;<br />

Cahova, H.; Pohl, R.; Sebest, P.; Havran, L.; Hocek, M.; Fojta, M. Chem.-<br />

Eur. J. 2009, 15, 1144–1154.<br />

(10) Cahova, H.; Havran, L.; Brazdilova, P.; Pivonkova, H.; Pohl, R.; Fojta, M.;<br />

Hocek, M. Angew. Chem., Int. Ed. 2008, 47, 2059–2062.<br />

(11) Horakova, P.; Simkova, E.; Vychodilova, Z.; Brazdova, M.; Fojta, M.<br />

Electroanalysis 2009, 21, 1723–1729.<br />

(12) Hocek, M.; Fojta, M. Org. Biomol. Chem. 2008, 6, 2233–2241.<br />

10.1021/ac100757v © 2010 American <strong>Chemical</strong> Society 6807<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/23/2010


Scheme 1. Top: Formulas of Purine Nucleobases (A,<br />

G) and Their 7-Deaza Analogues (A*, G*) a<br />

in primer extension (PEX) experiments. 12,13 Nevertheless, many<br />

of the nucleobase conjugates have displayed less facile incorporation<br />

at adjacent positions 7,9,10 and only some of them have been<br />

efficiently used in PCR. 13,14<br />

The 7-deazapurines, 7-deazaguanine (G*), or 7-deazaadenine<br />

(A*) (Scheme 1) are substitutes of standard purine nucleobases<br />

incorporable into DNA by PCR. Although the rate of G* and A*<br />

incorporation by DNA polymerases has been reported 15 to be<br />

lower, compared to the “parent” purines, both of these nucleobase<br />

analogues allow efficient sequence-specific DNA amplification by<br />

PCR and it has been possible to prepare PCR products with a<br />

high density of the corresponding modification. The 7-deazapurines<br />

are able to form Watson-Crick base pairs, maintaining<br />

pairing specificity of the respective natural nucleobases, but cannot<br />

form Hoogsteen pairs due to absence of the N7 atom (which is<br />

substituted by CH group, Scheme 1). Thus, the 7-deazapurines<br />

cannot be involved in triplex and tetraplex DNA structures. 16<br />

Reduced tendency of the “deaza-modified” DNA to adopting<br />

multistranded conformations has been utilized to improve PCR<br />

amplification of G-rich sequences such as (CGG)n repeat, 17 the<br />

length of which is analyzed during molecular diagnostics of<br />

fragile X syndrome. Similarly, 7-deaza-8-azaguanine has been<br />

used to create G-rich DNA probes with reduced propensity to<br />

aggregate and improved specificity. 18 Absence of the nitrogen<br />

atom at the 7-position in G* was also utilized in a study of<br />

sequence-specificity of DNA modification with cisplatin 19 and<br />

in studies of echinomycin-DNA interaction modes. 20 Moreover,<br />

DNA substituted with G* or A* has been shown to resist<br />

cleavage with certain endonucleases. 15<br />

a<br />

Atoms at 7-position are highlighted in red. Bottom: Watson-Crick<br />

and Hoogsteen base pairing. N7 is involved in the Hoogsteen pairing<br />

of natural purines.<br />

(13) Cahova, H.; Pohl, R.; Bednarova, L.; Novakova, K.; Cvacka, J.; Hocek, M.<br />

Org. Biomol. Chem. 2008, 6, 3657–3660.<br />

(14) Raindlova, V.; Pohl, R.; Sanda, M.; Hocek, M. Angew. Chem., Int. Ed. 2010,<br />

49, 1064–1066.<br />

(15) Seela, F.; Roling, A. Nucleic Acids Res. 1992, 20, 55–61.<br />

(16) Palecek, E. Crit. Rev. Biochem. Mol. Biol. 1991, 26, 151–226.<br />

(17) Cao, J.; Tarleton, J.; Barberio, D.; Davidow, L. S. Mol. Cell. Probes 1994,<br />

8, 177–180.<br />

(18) Kutyavin, I. V.; Lokhov, S. G.; Afonina, I. A.; Dempcy, R.; Gall, A. A.; Gorn,<br />

V. V.; Lukhtanov, E.; Metcalf, M.; Mills, A.; Reed, M. W.; Sanders, S.;<br />

Shishkina, I.; Vermeulen, N. M. J. Nucleic Acids Res. 2002, 30, 4952–4959.<br />

(19) Cairns, M. J.; Murray, V. Biochim. Biophys. Acta 1994, 1218, 315–321.<br />

(20) Sayers, E. W.; Waring, M. J. Biochemistry 1993, 32, 9094–9107.<br />

6808 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Table 1. Oligonucleotides Used in This Work a<br />

prim rnd 5′-CATGGGCGGCATGGG-3′<br />

prim noG 5′-TACTCATCATATCAA-3′<br />

temp rnd16 5′-CTAGCATGAGCTCAGTCCCATGCCGCCCATG-3′<br />

temp noG 5′-AATATAAATATATTGATATGATGAGTA-3′<br />

p53-for 5′-GAGGTTGTGAGGCGCTGCCC-3′<br />

p53-rev 5′-TCCTCTGTGCGCCGGTCTCT-3′<br />

a The template strands (temp rnd16 , temp noG ) used in PEX experiments<br />

were 5′end-biotinylated.<br />

Compared to the standard purines, 7-deazapurines exhibit<br />

significantly lower potentials of their oxidation. 9,21,22 G* is thus<br />

easily photooxidized by intercalated ethidium and this photooxidation<br />

has been interrogated as a function of distance, nucleotide<br />

sequence and integrity of π-stacking in studies of DNA-mediated<br />

charge transfer. 21 The G* ability of being selectively oxidized by<br />

a redox mediator with relatively low redox potential, such as<br />

Ru(dmb)3 3+/2+ (dmb )4,4′-dimethyl-2,2′-bipyridine), compared<br />

to A* together with natural G requiring stronger oxidants, such<br />

as Ru(bpy)3 3+/2+ (bpy )2,2′-bipyridine), has been utilized in a<br />

PCR-coupled electrochemical technique proposed for parallel<br />

detection of two genes. 22 Mediated electrooxidation of G* (or<br />

G) was also applied in an indirect electrochemical real time<br />

monitoring of PCR via measuring consumption of the respective<br />

dNTPs. 23 To our best knowledge, direct electrochemical<br />

analysis of DNA with incorporated G* or A* as electrochemically<br />

oxidizable tags has not been reported to date.<br />

In this paper we report on adsorptive transfer stripping<br />

voltammetric analysis of DNA with enzymatically incorporated G*<br />

or A* residues at a carbon electrode. We show that, particularly<br />

G* producing a specific signal separated from those yielded by<br />

natural DNA components, can be utilized as an excellent electroactive<br />

label suitable for monitoring of DNA amplification by PCR.<br />

MATERIALS AND METHODS<br />

Material. Synthetic ODNs (Table 1) were purchased from<br />

VBC genomics (Austria). Templates used in experiments involving<br />

the magnetoseparation procedure were biotinylated at their 5′<br />

ends. Plasmid pT77 bearing wild type p53 cDNA insert 24 (used<br />

as primary template for PCR amplification of the 347-bp fragment)<br />

was isolated from E. coli cells using Qiagen Plasmid Purification<br />

Kit and linearized with EcoR I restrictase (Takara). Streptavidincoated<br />

magnetic beads (MBstv) were purchased from Novagen<br />

(Germany), DyNAzyme II DNA Polymerase from Finnzymes<br />

(Finland), Pfu DNA Polymerase from Promega (U.S.), unmodified<br />

nucleoside triphosphates (dATP, dTTP, dCTP and dGTP),<br />

SYBR Green I, SYBR Gold and Stains-All reagent from Sigma,<br />

7-deaza-dGTP and 7-deaza-dATP from Jena Bioscience. Other<br />

chemicals were of analytical grade.<br />

Primer Extension (PEX). The primer (0.7 µM) was mixed<br />

with corresponding template ODN (0.7 µM), dNTPs (100 µM<br />

each; composition of the dNTP is specified in the text and Figure<br />

legends for individual experiments) and the DyNAzyme II DNA<br />

(21) Kelley, S. O.; Barton, J. K. Chem. Biol. 1998, 5, 413–425.<br />

(22) Yang, I. V.; Ropp, P. A.; Thorp, H. H. Anal. Chem. 2002, 74, 347–354.<br />

(23) Defever, T.; Druet, M.; Rochelet-Dequaire, M.; Joannes, M.; Grossiord, C.;<br />

Limoges, B.; Marchal, D. J. Am. Chem. Soc. 2009, 131, 11433–11441.<br />

(24) Hupp, T. R.; Meek, D. W.; Midgley, C. A.; Lane, D. P. Cell 1992, 71, 875–<br />

886.


Polymerase (1 U per sample). Reactions were carried out at 60 °C<br />

for 30 min.<br />

MBstv Magnetoseparation Procedure. The PEX products<br />

were captured at MBstv via biotin tags. Then, 50 µL aliquots of<br />

the PEX reaction mixtures were added to the MBstv (25 µL of<br />

the stock suspension washed twice with 100 µL of 0.3 M NaCl,<br />

10 mM Tris-HCl, pH 7.4, buffer H). The mixture was incubated<br />

on a shaker for 30 min at 20 °C. Then the beads were washed<br />

three times with 100 µL of PBS (0.14 M NaCl, 3 mM KCl, 4<br />

mM sodium phosphate, pH 7.4) containing 0.01% Tween20,<br />

three times with 100 µL of the buffer H and resuspended in<br />

deionized water (50 µL). The extended primers were released<br />

by heating at 75 °C for 2 min. Prior to the electrochemical<br />

measurements, NaCl in total concentration of 0.2 M was added<br />

to the samples.<br />

Preparative PCR. Amplification of the DNA fragment: 500<br />

ng of the pT77 template was mixed with p53-for and p53-rev<br />

primers (0.5 µM each), Pfu DNA Polymerase (3 U) and mix of<br />

dNTPs (125 µM each) in total volume of 100 µL. The PCR involved<br />

30 cycles if not stated otherwise (denaturation 94 °C/90 s,<br />

annealing 60 °C/120 s, polymerization 72 °C/180 s) and was run<br />

using C1000 Thermal Cycler (BioRad). The PCR products were<br />

purified using QIAquick PCR Purification Kit (Qiagen) and their<br />

concentrations were determined spectrophotometrically using<br />

NanoDrop ND-1000 Spectrophotometer (NanoDrop Technologies,<br />

U.S.).<br />

Real Time PCR. Reactions were prepared in duplicates in 20<br />

µL handmade PCR mix consisting of 30 ng of pT77 template, p53for<br />

and p53-rev primers at concentrations 0.5 µM, 1 × DyNAzyme<br />

II buffer, 1.25 mM MgCl 2,20000× diluted SYBR Green I, each<br />

dNTP (or sum of dGTP+dG*TP) 125 µM,1UofDyNAzyme<br />

II DNA polymerase. PCR involved 20 cycles (denaturation<br />

94 °C/30 s, annealing 56 °C/30 s, polymerization 72 °C/30 s,<br />

fluorescence measurement in SYBR Green I channel 15 s/78 °C,<br />

wavelength of the source 470 nm, wavelength of the detection<br />

filter 585 nm; final extension 72 °C/3 min) and was run and<br />

evaluated using the RotorGene-3000 (Qiagen, Germany).<br />

Native PAGE. The PCR products (1 µL each) were mixed<br />

with loading buffer (0.1% SDS, 5% glycerol, 5 mg mL -1 bromophenol<br />

blue) and subjected to electrophoresis in 5% native gel<br />

containing 1 × TBE buffer (pH 8) at 150 V, 4 °C for 35 min.<br />

Gels stained with ethidium bromide or SYBR dyes were<br />

visualized using LAS-3000 (FUJIFILM Corporation), those<br />

stained with the Stains-All reagent were scanned.<br />

Electrochemical Analysis. The PEX and PCR products were<br />

analyzed by using ex situ (adsorptive transfer stripping, AdTS)<br />

square-wave voltammetry (SWV). The DNA was accumulated at<br />

the basal-plane pyrolytic graphite electrode (PGE; prepared and<br />

pretreated as described in ref 25) surface from 5 µL aliquots<br />

containing 0.3 M NaCl for 60 s. Then the electrode was rinsed by<br />

deionized water and was placed into the electrochemical cell. SWV<br />

settings: initial potential -1.0 V, final potential +1.5 V, pulse<br />

amplitude 25 mV, frequency 200 Hz, potential step 5 mV. The<br />

measurements were performed at ambient temperature in 0.2 M<br />

acetate buffer pH 5 by using CHI440 Electrochemical Workstation<br />

(CH Instruments, Inc., U.S.) in a three-electrode setup (with the<br />

PGE as working electrode, Ag/AgCl/3 M KCl as reference, and<br />

(25) Fojta, M.; Havran, L.; Kizek, R.; Billova, S. Talanta 2002, 56, 867–874.<br />

platinum wire as counter electrode). Baseline correction of the<br />

voltammograms was performed by means of a moving average<br />

algorithm (GPES 4 software, EcoChemie).<br />

RESULTS AND DISCUSSION<br />

Primer extension (PEX) has recently been used for the<br />

introduction of labeled nucleobases into oligodeoxynucleotides<br />

(ODNs). 7-13 Using nucleobase-modified deoxynucleotide triphosphates<br />

(dNTPs) and DNA polymerases, we prepared ODNs<br />

bearing various tags producing analytically useful electrochemical<br />

responses. Since 7-deazapurines are electrochemically oxidized<br />

at considerably less positive potentials, compared to the respective<br />

natural purine nucleobases, 9,21,22 it is in principle possible to apply<br />

these nucleobase analogues themselves (without introducing any<br />

extra label groups) as electroactive markers of the in vitro synthesized<br />

DNA.<br />

We performed PEX with primrnd primer and 5′-end-biotinylated<br />

temprnd16 template, the nucleotide sequence of which was<br />

designed to accommodate all four DNA bases (four-times each)<br />

in the synthesized DNA stretch (Figure 1A). The doublestranded<br />

PEX products were pulled-down from the reaction<br />

mixture using streptavidin coated magnetic beads (MBstv) and,<br />

after magnetic separation, the captured duplex ODN was<br />

thermally denatured to release the extended primer strand. The<br />

latter was finally analyzed using adsorptive transfer stripping<br />

square wave voltammetry (AdTS SWV). As shown in Figure<br />

1B,C, voltammetric responses of the pexrnd16 products reflected<br />

the composition of dNTP mix used. For a standard dNTP mix<br />

(dATP, dGTP, dCTP, and dTTP), we observed two anodic<br />

signals corresponding to electrochemical oxidation of guanine<br />

(peak Gox at 1.15 V) and adenine (peak Aox at 1.35 V, Figure<br />

1B,C). When dGTP was replaced by 7-deaza-dGTP (dG*TP), peak<br />

Gox was decreased and a new anodic peak appeared at 0.80 V<br />

(peak G* ox , Figure 1B,C) which has been ascribed to electrochemical<br />

oxidation of G* introduced into the extended ODN<br />

stretch. Lower intensity of the peak Gox corresponded to lower<br />

number of G residues in the G*-modified PEX product,<br />

compared to the PEX product composed of standard nucleotides<br />

(note that after PEX in the presence of dG*TP, there<br />

are still natural guanines present in the primer stretch, but not<br />

in the extended stretch, decreasing the total number of G<br />

residues from 12 to 8). When dATP was replaced by 7-deazadATP<br />

(dA*TP), we observed a decrease of the peak Aox intensity (due to lowering of the total number of adenines in<br />

the entire extended primer strand) and increase of the peak<br />

at 1.15 V. The potential of electrochemical oxidation of A* was<br />

shown 9 to coincide with the potential of peak Gox , and thus<br />

currents due to A* oxidation were responsible for the increase<br />

of the apparent signal intensity. To verify this hypothesis, we<br />

prepared another PEX product using primnoG primer and<br />

tempnoG template (Table 1). Neither of these ODNs contains G<br />

residues and no G is incorporated during PEX, which is reflected<br />

in the absence of peak Gox at the voltammogram of unmodified<br />

pexnoG product, showing only peak Aox (black curve in Figure<br />

1C, inset). After performing PEX with the same primer-template<br />

pair but with dATP replaced by dA*TP, the peak Aox was<br />

decreased and another signal appeared at 1.15 V, corresponding<br />

to the A* oxidation (peak A* ox ; green curve in Figure 1C, inset).<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6809


Figure 1. Electrochemical responses of oligonucleotides containing 7-deazapurines incorporated by primer extension (PEX). (A) Scheme of<br />

the experiment. The PEX was performed using a 5′-terminally biotinylated template. After extension of the primer using either standard dNTP<br />

mix, a mixture with dGTP replaced by dG*TP, or a mixture with dATP replaced by dA*TP, the duplex ODN was captured at magnetic beads<br />

covered with streptavidin, the beads were washed and the modified strand released by thermal denaturation, followed by AdTS SWV measurements<br />

at the PGE (shown for temp rnd16 and prim rnd sequences; blue and red letters are used to highlight A and G positions, respectively). (B) SWV<br />

voltammograms of pex rnd16 : standard nucleobases (black); A* instead of A (blue); G* instead of G (red); background electrolyte (dotted); (C)<br />

baseline-corrected curves taken from (B). Inset: PEX products obtained with temp noG and prim noG for dTTP+dATP mix (black) or dTTP+dA*TP<br />

mix (green).<br />

In next experiment, we prepared a 347-bp DNA fragment by<br />

the polymerase chain reaction (PCR) using the pT77 template and<br />

various dNTP mixtures. The PCR products were isolated from<br />

the reaction mixture using Qiagen columns and analyzed by AdTS<br />

SWV as above. During thermal cycling in the PCR, forward and<br />

backward primers (Table 1) are elongated on templates of both<br />

DNA strands, resulting in exponential amplification of the fragment<br />

delimited by the two primers (see Figure 2A). Thus, each<br />

strand of the double-stranded PCR product begins with the primer<br />

at its 5′-terminus, followed by newly synthesized stretch, the<br />

nucleotide composition of which (i.e., presence or absence of the<br />

7-deazapurines) is dictated by the composition of the dNTP mix<br />

(Figure 2A). Accordingly, when using mix of four standard dNTPs<br />

(in Figure 2B denoted as G+A), we observed peak G ox and peak<br />

A ox corresponding to natural purine bases. Replacement of<br />

dGTP with dG*TP resulted in appearance of an intense peak<br />

G* ox (due to G* incorporated into the polymerase-synthesized<br />

strands) and strong decrease of the intensity of peak G ox which,<br />

however, never reached zero. This peak G ox was due to G<br />

residues in the primer stretches of the G*-modified amplicons<br />

(Figure 2A). In addition, certain amount of unconsumed primers<br />

and primary templates remaining in the samples after the<br />

6810 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

purification step might contribute to the peak G ox intensity, as<br />

indicated by negative PCR control experiment (PCR mixture<br />

not subjected to thermal cycling, dotted curve in Figure 2B).<br />

(It should be however noted that contributions to the peak G ox<br />

from ODN primers in the negative PCR control, as apparent<br />

from Figure 2B, and those from primer stretches in the G*modified<br />

amplicon are not simply additive, because concentration<br />

of residual free primers after the PCR cycling is much lower than<br />

their initial concentration). When dATP was replaced with dA*TP,<br />

we observed increase of the signal close to 1.15 V (peak G ox +<br />

peak A* ox ) and decrease of peak A ox (again, the residual peak<br />

A ox was yielded by A residues in the primer stretches).<br />

Further, we focused our attention on analysis of DNA amplicons<br />

containing G* as an independently detectable marker and<br />

performed PCR experiments using dNTP mixes containing both<br />

dGTP and dG*TP at different ratios (while keeping the total dGTP<br />

+ dG*TP concentration constant). The resulting PCR products<br />

were analyzed electrochemically as above. As shown in Figure 3,<br />

changes in the relative abundances of G and G* in the PCR<br />

product (given by the molar fraction of the respective dNTP) were<br />

reflected in changes of corresponding electrochemical signals<br />

(peak G ox and peak G* ox ). Peak A ox due to adenine, the content


of which was constant for all amplicons analyzed in this<br />

experiment, did not show any significant trend, suggesting<br />

approximately constant yields of the PCR products obtained<br />

for any dGTP/dG*TP ratio after the 30 PCR cycles (which was<br />

also confirmed by UV-vis spectrophotometry, not shown).<br />

Dependences of the heights of peak Gox (yielded by the<br />

unmodified 347-bp PCR product) and peak G* ox (yielded by<br />

the same DNA fragment amplified with 100% of dG*TP) on<br />

concentrations of the amplicons followed similar trends (Figure<br />

4A), showing approximately linear regions below 10 µgmL-1and sublinear trends, suggesting saturation of the electrode surface,<br />

at higher DNA concentrations. The lowest detectable DNA<br />

concentrations were around 1 µg mL-1 in both cases. Electrochemical<br />

determination of the modified and unmodified 347bp<br />

amplicons was compared to other commonly used techniques,<br />

based on gel electrophoresis followed by DNA staining<br />

(Figure 4B). The amplicons were separated in native 5% polyacrylamide<br />

gel (1 µL of the reaction mixture, containing about 30<br />

µg mL-1 Figure 2. Electrochemical responses of a PCR-amplified 347-bp<br />

DNA fragment modified with 7-deazapurines. (A) Scheme of the PCR.<br />

Primers at the 5′-ends of both strands of the PCR product (black)<br />

contain only standard nucleobases, regardless of the dNTP composition.<br />

The synthesized stretches contain modified nucleobases depending<br />

on the dNTP mix. (B) Baseline-corrected voltammograms<br />

obtained for DNA fragment resulting from PCR in the presence of<br />

standard dNTP mix (black), for a mix with G* instead of G (red) and<br />

for a mix with A* instead of A (blue). The PCR was conducted in 30<br />

cycles and the products were purified using Qiagen PCR Purification<br />

Kit. Dotted curve corresponds to control PCR mixture (with G*+A)<br />

which was not subjected to thermal cycling.<br />

of the amplicon, and three binary dilutions). After<br />

electrophoresis, the gels were stained with ethidium bromide,<br />

Figure 3. Dependence of the heights of peak G* ox (red), peak G ox<br />

(black) and peak A ox (empty triangles) on the [dG*TP]/[dGTP]+[dG*TP]<br />

ratio in the PCR reaction used for amplification of the 347-bp DNA<br />

fragment. Other conditions as in Figure 2.<br />

SYBR Green I or Stains-All reagent (Figure 4B). Ethidiumstained<br />

bands of the G*-modified PCR products were considerably<br />

weaker than those of the unmodified amplicon (even when only<br />

50% of Gs were substituted by G*, see Figure 4B). Such<br />

observation was in agreement with literature data 26 showing that<br />

ethidium fluorescence is quenched when the dye is intercalated<br />

next to G*. Notably, we observed even stronger quenching effect<br />

of G* on the fluorescence of SYBR Green I (Figure 4B) and SYBR<br />

Gold (not shown) dyes. Thus, results of the fluorescent DNA<br />

staining might be misinterpreted, for the densely G*-modified<br />

DNA, in terms of (strongly) decreasing amount of the PCR<br />

products with increasing G/G* ratio. On the other hand, staining<br />

of the polyacrylamide gel with the Stains-All reagent (Figure 4B)<br />

did not reveal significant differences in the amounts of the<br />

unmodified and G*-substituted amplicons, in agreement with the<br />

electrochemical data.<br />

Despite the approximately same yields of the PCR products<br />

after 30 cycles, we were interested whether we are able to follow<br />

electrochemically differences in the kinetics of the PCR reactions<br />

through analysis of the amplicons after lower number of amplification<br />

cycles. Previous data 15 revealed less efficient DNA amplification<br />

by PCR in the presence of dG*TPs, compared to PCR with<br />

standard dNTPs only. We followed electrochemical signals of the<br />

unmodified and fully G*-substituted PCR products after 5, 10, 20,<br />

and 30 cycles (Figure 5A). For five cycles, peak G ox produced<br />

by the unmodified amplicon was considerably higher than peak<br />

G* ox of the G*-modified PCR product. Even after subtraction<br />

of signal intensity produced by the control PCR mix not<br />

subjected to the thermal cycling (containing initial concentrations<br />

of ODN primers and the primary template), the peak G ox<br />

was at least twice higher than peak G* ox produced by the G*modified<br />

amplicon after the same number of cycles. Large<br />

differences between the signal intensities were also observed<br />

after 10 amplification cycles, while after 20 and 30 cycles both<br />

peaks reached their limiting values. Hence, more cycles were<br />

required to reach the limiting amount of the G*-modified PCR<br />

(26) Latimer, L. J. P.; Lee, J. S. J. Biol. Chem. 1991, 266, 13849–13851.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6811


Figure 4. (A) Dependence of the intensities of peak G ox (measured<br />

for unmodified 347-bp amplicon, black) and peak G* ox (measured for<br />

the same DNA fragment amplified in the presence of dG*TP instead<br />

of dGTP, red) on DNA concentration. Other conditions as in Figure<br />

2. (B) Staining of the 347-bp amplicon in polyacryalmide gels with<br />

(from top to bottom) ethidium bromide, SYBR Green I and Stains-<br />

All. The PCR reaction contained either four standard dNTPs, G/G*<br />

) 1, or dGTP fully replaced with dG*TP as indicated on the top.<br />

Loading of the PCR products: 1 µL of the undiluted reaction mixture<br />

(about 30 ng of the amplicon), followed by binary dilutions as<br />

indicated.<br />

product, compared to PCR with standard dNTP mix. To support<br />

this conclusion, we performed a real-time PCR experiment<br />

using standard and dG*TP-containing dNTP mixes. Figure 5B<br />

shows increase of the fluorescence of SYBR Green I (which was<br />

used here as fluorescent dye indicating progression of DNA<br />

amplification) as a function of the number of PCR cycles for dNTP<br />

mixes containing 0, 10, 20, 30, 50, and 100% of dG*TP (of<br />

dGTP+dG*TP total). The steepest increase of the signal was<br />

observed for the unmodified amplicon and the amplification rate<br />

decreased with the G* content, showing clear difference even for<br />

10% G*. For 100% of G* the apparent amplification rate was<br />

remarkably depressed. However, it should be taken into consideration<br />

that SYBR Green I fluorescence is quenched by G*<br />

incorporated (see above), and thus the weak signal detected for<br />

the G*-substituted amplicon can have reflected the dense DNA<br />

modification rather than the amount of the PCR product. We<br />

therefore focused on the evaluation of the shapes of the amplification<br />

curves rather than the absolute signal magnitudes. RotorGene-<br />

3000 software enables to compare reactions parameters in the<br />

6812 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 5. (A) Effects of the number of PCR cycles on intensities of<br />

peak G ox (measured for unmodified 347-bp amplicon, gray columns)<br />

and peak G* ox (measured for the same DNA fragment amplified in<br />

the presence of dG*TP instead of dGTP, red columns). Other<br />

conditions as in Figure 2. (B) Quantitative (real time) PCR amplification<br />

of the 347-bp DNA fragment: standard dNTP mix (no G*) (1);<br />

10% G* (2); 20% G*(3); 30% G* (4); 50% G (5); G fully replaced with<br />

G* (6); negative “no template” control for standard dNTP mix (7). The<br />

graph shows fluorescence of SYBR Green I complexes with the DNA<br />

amplicons as a function of the number of PCR cycles. Inset, takeoff<br />

graphs derived from the real time PCR data (colors correspond to<br />

the raw data plot). All samples are plotted in duplicate.<br />

comparative quantification mode. The takeoff points for each<br />

reaction (sample) are calculated from the second derivatives of<br />

raw data (Figure 5B, inset). Generally, the takeoff value is not<br />

possible to determine exactly and it is taken as 80% below the<br />

peak of second derivative (i.e., below the point where the<br />

amplification curve is increasing most steeply). In the same mode,<br />

reaction efficiencies of particular reactions are reporting the<br />

amplificability of the template under the given conditions. For<br />

example, amplification factors of 1.72, 1.65, and 1.61 obtained for<br />

reactions with 100% G, G/G* ) 1 and 100% G*, respectively, clearly<br />

show less efficient amplification in the dG*TP-substituted reactions.<br />

Similar effects of G* on the PCR kinetics were obtained<br />

using Taqman hydrolytic probes which release fluorescent indicators<br />

into solution as the PCR progresses, and thus the fluorescence<br />

intensity was not affected by the incorporated G* as it was in the<br />

case of the intercalative dyes (not shown). Hence, results of the<br />

real time PCR experiment were in a qualitative accordance with<br />

the above electrochemical data indicating lower amplification rate<br />

for the reaction with dG*TP.<br />

CONCLUSIONS<br />

7-deazapurines G* and A* enzymatically incorporated into DNA<br />

are electrochemically oxidizable at carbon electrodes, producing


analytically useful signals at less positive potentials, compared to<br />

the corresponding natural purine nucleobases. Peak A* ox due to<br />

electrooxidation of A* occurs at the same potential as the peak<br />

G ox due to oxidation of natural G, preventing qualitative<br />

discrimination of A* incorporated into DNA containing guanine.<br />

On the other hand, total or partial substitution of G with G*<br />

results in appearance of a new anodic signal, peak G* ox ,at<br />

potential less positive than potentials of oxidation of any natural<br />

component of DNA, allowing independent determination of G*<br />

incorporated. G* can thus be utilized as an inexpensive,<br />

commercially available electroactive label for easy monitoring<br />

of primer extension or polymerase chain reactions via simple<br />

direct electrochemistry using cheap, widely accessible carbon<br />

electrodes. In turn, considering applications of 7-deazaguanine<br />

as DNA modifications preventing formation of multistranded<br />

alternative structures in PCR analysis of G-rich sequences 17<br />

and problems with quenching of fluorescence of ethidium 26<br />

or “SYBR” (to our knowledge, for the first time reported in<br />

this paper) dyes, electrochemical analysis appears an attractive<br />

complementary approach providing modification-specific signal<br />

suitable for quantitation of the fully modified amplified DNA<br />

as well as for the determination of the DNA modification extent.<br />

Specifically for the G*-modified DNA, the voltammetric analysis<br />

represents a simple and direct way to differentiate between the<br />

(27) Fojta, M.; Brazdilova, P.; Cahova, K.; Pecinka, P. Electroanalysis 2006, 18,<br />

141–151.<br />

(28) Fojta, M.; Havran, L.; Vojtísˇková, M.; Palecek, E. J. Am. Chem. Soc. 2004,<br />

126, 6532–6533.<br />

natural G and G* residues and to determine relative content<br />

of both. In contrast to using peak G ox or peak A ox , produced<br />

by natural purines, determination of DNA amplicons based on<br />

the measurement of peak G* ox is not affected by signals<br />

produced by residual ODN primers and/or the primary<br />

template. It has to be naturally taken into consideration that<br />

peak G* ox intensity must depend on the G + C content within<br />

the amplified region; alternatively, this feature may potentially<br />

be useful for the determination of the G + C content or<br />

estimation of the length of the amplified DNA fragment (e.g.,<br />

triplet repeat expansion 27,28 ) provided that a proper normalization<br />

of the signal intensity (such as fragment ends “counting” through<br />

the intensity of natural G in primers) is used. Besides the PCR<br />

applications, the G* electroactive is also potentially useful for taillabeling<br />

of DNA probes for electrochemical hybridization assays<br />

and other bioanalytical applications which are metter of our<br />

ongoing research and will be reported elsewhere.<br />

ACKNOWLEDGMENT<br />

This work was supported by Grant Agency of the ASCR (grant<br />

IAA400040901), by the ASCR (AV0Z50040507 and AV0Z50040702)<br />

and by the MEYS CR (LC06035, MSM0021622415). H.P. and P.H.<br />

contributed equally to this work.<br />

Received for review March 24, 2010. Accepted July 10,<br />

2010.<br />

AC100757V<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6813


Anal. Chem. 2010, 82, 6814–6820<br />

Screening Assay of Very Long Chain Fatty Acids in<br />

Human Plasma with Multiwalled Carbon<br />

Nanotube-Based Surface-Assisted Laser<br />

Desorption/Ionization Mass Spectrometry<br />

Wei-Yi Hsu, † Wei-De Lin, †,‡ Wuh-Liang Hwu, § Chien-Chen Lai,* ,‡,| and Fuu-Jen Tsai* ,†,|,&<br />

Department of Medical Research, China Medical University Hospital, Taichung, Taiwan, Institute of Molecular Biology,<br />

National Chung Hsing University, Taichung, Taiwan, Department of Medical Genetics, National Taiwan University<br />

Hospital and National Taiwan University College of Medicine, Taipei, Taiwan, Graduate Institute of Chinese Medical<br />

Science, China Medical University, Taichung, Taiwan, and Department of Health and Nutrition Biotechnology, Asia<br />

University, Taichung, Taiwan<br />

Peroxisomal disorders are characterized biochemically by<br />

elevated levels of very long chain fatty acids (VLCFAs) in<br />

serum. Herein, we describe a novel approach for quantification<br />

of VLCFAs in serum, namely, eicosanoic acid<br />

(C20:0), docosanoic acid (C22:0), tetracosanoic acid<br />

(C24:0), and hexacosanoic acid (C26:0). The methodology<br />

is based on (i) enrichment of VLCFA derivatives using<br />

multiwalled carbon nanotubes (MWCNTs); (ii) quantification<br />

using stable isotope-labeled internal standards; and<br />

(iii) direct detection using MWCNT-based surface-assisted<br />

laser desorption/ionization-time-of-flight mass spectrometry<br />

(SALDI-TOFMS). Four kinds of MWCNTs (Aldrich<br />

636843, 636495, 636509, and 636819) of different<br />

lengths and diameters were tested using the developed<br />

technique. The data show that 636843, the MWCNT with<br />

the largest outer diameter (o.d.), the widest wall thickness,<br />

and shortest length, had the best limit of detection<br />

(0.5-1 µg/mL) We also found that there was no significant<br />

difference in enrichment efficiency of VLCFAs between<br />

the four MWCNTs, which suggests that the size of<br />

the MWCNT may contribute to desorption/ionization<br />

efficiency. To our knowledge, this is the first study to test<br />

the enrichment of VLCFAs using MWCNTs of different<br />

sizes. We have shown that the VLCFAs adsorbed by<br />

MWCNTs can be analyzed by SALDI-TOFMS. In addition,<br />

this method does not require liquid/gas chromatography<br />

separation, thereby allowing for high-throughput screening<br />

of VLCFAs in peroxisomal disorders.<br />

Carbon nanotubes (CNTs) are basically elongated fullerenes<br />

consisting of rolled graphite sheets with diameters in the nanom-<br />

* To whom correspondence should be addressed. (C.-C.L.) Phone: (886)<br />

4-22840485ext. 235. Fax: (886) 4-22858163. E-Mail: lailai@dragon.nchu.edu.tw.<br />

(F.-J.T.) Phone: (886) 4-22062121ext. 7076. Fax: (886) 4-22033295. E-Mail:<br />

d0704@mail.cmuh.org.tw.<br />

† China Medical University Hospital.<br />

‡ National Chung Hsing University.<br />

§ National Taiwan University Hospital and National Taiwan University College<br />

of Medicine.<br />

| China Medical University.<br />

& Asia University.<br />

6814 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

eter range and lengths in the micrometer range. CNTs were first<br />

discovered in the cathode deposit of arc discharge between<br />

graphite rods. 1 CNTs are classified based on the number of carbon<br />

atom layers into multiwalled carbon nanotubes (MWCNTs) and<br />

single-walled carbon nanotubes (SWCNTs). 2 CNTs have unique<br />

chemical, physical, mechanical, and electrical properties, and<br />

therefore, they have potential applications in a variety of fields. 3,4<br />

The hydrophobic features and large adsorption surface areas of<br />

CNTs make them effective solid-phase extraction adsorbents of<br />

a wide range of chemical compounds including endocrine disruptors,<br />

5 chlorophenols, 6 polycyclic aromatic hydrocarbons, 7 benzodiazepine<br />

residues, 8 barbiturates, 9 herbicides, 10 tetracyclines, 11<br />

and polybrominated diphenyl ethers. 12 It has been shown that<br />

functional groups (i.e., hydroxyl or carboxyl groups) introduced<br />

onto the surface of CNTs can modulate their affinity for and<br />

selectivity of hydrophilic solutes or compounds with polar functional<br />

groups. 13 From a structural point of view, buckminster<br />

fullerene (C60) has also been shown to be an effective solid-phase<br />

extraction adsorbent. Munoz and Valcarcel showed that MWCNTs<br />

as well as C60 and C70 fullerenes were superior to graphitized<br />

(1) Iijima, S. Nature 1991, 354, 56–58.<br />

(2) Iijima, S.; Ichihashi, T. Nature 1993, 363, 603–605.<br />

(3) Saito, N.; Usui, Y.; Aoki, K.; Narita, N.; Shimizu, M.; Hara, K.; Ogiwara, N.;<br />

Nakamura, K.; Ishigaki, N.; Kato, H.; Taruta, S.; Endo, M. Chem. Soc. Rev.<br />

2009, 38, 1897–1903.<br />

(4) Valcarcel, M.; Cardenas, S.; Simonet, B. M. Anal. Chem. 2007, 79, 4788–<br />

4797.<br />

(5) Cai, Y.; Jiang, G.; Liu, J.; Zhou, Q. Anal. Chem. 2003, 75, 2517–2521.<br />

(6) Cai, Y. Q.; Cai, Y. E.; Mou, S. F.; Lu, Y. Q. J. Chromatogr., A 2005, 1081,<br />

245–247.<br />

(7) Wang, W. D.; Huang, Y. M.; Shu, W. Q.; Cao, J. J. Chromatogr., A 2007,<br />

1173, 27–36.<br />

(8) Wang, L.; Zhao, H.; Qiu, Y.; Zhou, Z. J. Chromatogr., A 2006, 1136, 99–<br />

105.<br />

(9) Zhao, H.; Wang, L.; Qiu, Y.; Zhou, Z.; Zhong, W.; Li, X. Anal. Chim. Acta<br />

2007, 586, 399–406.<br />

(10) Zhou, Q.; Xiao, J.; Wang, W.; Liu, G.; Shi, Q.; Wang, J. Talanta 2006, 68,<br />

1309–1315.<br />

(11) Suarez, B.; Santos, B.; Simonet, B. M.; Cardenas, S.; Valcarcel, M.<br />

J. Chromatogr., A 2007, 1175, 127–132.<br />

(12) Wang, J. X.; Jiang, D. Q.; Gu, Z. Y.; Yan, X. P. J. Chromatogr., A 2006,<br />

1137, 8–14.<br />

(13) Sae-Khow, O.; Mitra, S. J. Chromatogr., A 2009, 1216, 2270–2274.<br />

10.1021/ac100772j © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/22/2010


carbon black and RP-C18 for the extraction of the organometallic<br />

compounds. 14<br />

Surface-assisted laser desorption/ionization (SALDI), a matrixfree<br />

method for laser desorption/ionization mass spectrometry, 15<br />

has overcome the limitations of matrix-assisted laser desorption/<br />

ionization (MALDI). The lower matrix interference in the low mass<br />

spectrum (molecular mass < 500 Da) makes SALDI more<br />

appropriate for quantifying small molecules. There are four major<br />

types of SALDI substrates: silicon (DIOS), 16,17 sol-gels, 18 carbonbased<br />

substrates, and metal particle-based substrates. 19-22 Carbonbased<br />

SALDI, which utilizes graphite with sizes in the micrometer<br />

range, was first used as a desorption/ionization substrate for<br />

detection of peptides and proteins in 1995. 20 CNTs have also been<br />

used as SALDI substrates to analyze small peptides (


Scheme 1. (A) Procedure of Derivatizating VLCFAs to Quaternary Ammonium Salt and a Neutral Loss in the<br />

Postsource Decay (PSD) and (B) Multiple Functions of MWCNTs for Analyte Enrichment and as the SALDI Substrate<br />

was added to dissolve the residue at room temperature for 10<br />

min. The residue was removed in a stream of nitrogen at 40 °C<br />

and was reconstituted in 50% (v/v) methanol immediately before<br />

analysis.<br />

Enrichment of VLCFAs and Analysis with SALDI-TOFMS<br />

by MWCNTs. The procedure was performed according to the<br />

procedure reported by Pan et al. with minor modifications 32<br />

(Scheme 1B). MWCNTs (10 mg) were first suspended in 1 mL<br />

of 50% (v/v) methanol and then sonicated for 3 min.<br />

Then, 10 µL of the suspension was pipetted immediately into<br />

a centrifuge tube containing the analyte solution. The tube was<br />

sonicated for less than 5 s, after which the MWCNTs with analytes<br />

solution were homogeneously spread in the solution and the<br />

analytes were adsorbed from the liquid phase to the surface of<br />

MWCNTs within 10 min. The analyte-adsorbed MWCNTs were<br />

precipitated by centrifugation at 10 000 rpm for 10 min. Then, the<br />

supernatant was removed, and 4 µL of dispersant solution of 50%<br />

methanol (v/v) with 5% glycerol (v/v) and 1% sucrose (w/w) was<br />

added to the centrifuge tube to resuspend the MWCNTs. Finally,<br />

about 2 µL of the MWCNT solution was pipetted onto the SALDI-<br />

TOFMS sample plate. The sample plate was left at room temperature<br />

for 10-15 min to allow for the evaporation of the solvent<br />

before analysis by SALDI-TOFMS.<br />

(32) Pan, C.; Xu, S.; Zou, H.; Guo, Z.; Zhang, Y.; Guo, B. J. Am. Soc. Mass<br />

Spectrom. 2005, 16, 263–270.<br />

6816 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

SALDI-TOFMS analyses were performed on an AXIMA-CFR<br />

plus (Shimazu/KRATOS, Manchester, UK) instrument equipped<br />

with a nitrogen laser (wavelength at 337 nm) and delayed<br />

extraction. Ions were accelerated by energy of 20 kV before<br />

entering the TOF mass spectrometer. All spectra were acquired<br />

in the linear mode. In the SALDI-TOF/PSD MS experiments,<br />

spectra were acquired in the reflectron mode. The mass spectra<br />

of 50 different profiles of a sample spot were averaged. Each<br />

profile included five laser shots. The mass calibration was<br />

achieved using low-mass standards (MassPREP Calibration Mix-<br />

DIOS, Waters) in the mass range 20-1500 Da. Quantitative<br />

analysis of VLCFAs was carried out by measuring the ion peak<br />

intensity of the individual mass peaks.<br />

Limit of Detection (LOD) of Derivatized VLCFAs with<br />

MWCNT-Based SALDI-MS and MALDI-MS. Serial concentrations<br />

of the standard solutions of four VLCFAs (60 µL) were<br />

prepared: 50, 25, 10, 5, 1, and 0.5 µg/mL. Following the derivatization<br />

of VLCFAs to quaternary ammonium salt derivatives, the<br />

products were reconstituted in 60 µL of 50% (v/v) methanol. The<br />

derivatized products (10 µL) were enriched by MWCNTs according<br />

to the procedure mentioned above. The amounts of VLCFAs<br />

analyzed per spot were 250, 125, 50, 25, 5, and 2.5 ng.<br />

The sensitivity of MALDI-TOFMS to detect VLCFAs was<br />

tested using CHCA (10 mg/mL in 80% acetone) as the matrix.<br />

The reconstituted solution (10 µL) was mixed with CHCA


solution in an equal volume, and 1 µL of the mixture was<br />

pipetted onto the MALDI-MS sample plate without the enrichment<br />

procedure.<br />

Adsorption Efficiency of VLCFAs by MWCNTs in the<br />

Enrichment Procedure. Adsorption efficiency was determined<br />

using standard mixtures of four VLCFAs (40 µg/mL each).<br />

Standard mixtures were derivatized first and then subjected to<br />

the enrichment procedure with MWCNTs as mentioned above.<br />

After the standard mixture had adsorbed to the surface of the<br />

MWCNTs, the supernatants were analyzed by HPLC/ESI-MS.<br />

HPLC analysis was performed on a 3 µm C18 column (AtlantisdC18,<br />

2.1 mm i.d × 50 mm, Waters) with an isocratic program<br />

(mobile phases of acetonitrile/methanol/formic acid ) 80/20/<br />

0.1). The autosampler was a Finnigan Surveyor autosampler fitted<br />

with a 10 µL loop. The HPLC and autosampler systems were all<br />

synchronized via Xcalibur software (Xcalibur, Finnigan Corp.). A<br />

Finnigan LCQ DECA XP PLUS quadrupole ion trap mass spectrometer<br />

(Finnigan Corp., San Jose, CA) equipped with a<br />

pneumatically assisted electrospray ionization source was used.<br />

The mass spectrometer was operated in positive ion mode by<br />

applying a voltage of 4.5 kV to the ESI needle. The temperature<br />

of the heated capillary in the ESI source was set at 260 °C.<br />

The flow rate of the sheath gas (nitrogen) was set at 25<br />

(arbitrary units). Helium was used as the damping gas at a<br />

pressure of 10 -3 Torr. Voltages across the capillary and the<br />

octapole lenses were tuned by an automated procedure to<br />

maximize the signal of the ion of interest. We used the<br />

following SIM (selected ion monitoring) transitions for the<br />

analysis: C20:0 (center mass 398.2), C22:0 (center mass 426.2),<br />

C24:0 (center mass 454.2), and C26:0 (center mass 482.2). The<br />

isolation width was set as 1.5 Da. The experimental programs<br />

and data analyses were performed with the software package<br />

Xcalibur (Xcalibur, Finnigan Corp.).<br />

Preparation of Standard Solutions and Calibration Curves.<br />

Stock solutions of internal standard mixtures of three stable<br />

isotope-labeled VLCFAs (C20:0-d3, C22:0-d3, and C26:0-d4) and<br />

standard mixtures of four VLCFAs (C20:0, C22:0, C24:0, C26:<br />

0) were prepared at a concentration of 50 µg/mL in methanol<br />

and kept in the dark at -20 °C when not in use. For the<br />

calibration curve, the concentrations of the calibration solutions<br />

of standard mixture were 5, 10, 15, and 20 µg/mL, and the<br />

concentration of the calibration solution of the internal standard<br />

mixture was 10 µg/mL. The calibration solutions (50 µL) were<br />

evaporated in a stream of nitrogen prior to the derivatization<br />

procedure and reconstituted in 10 µL of 50% (v/v) methanol<br />

for the enrichment procedure.<br />

Hydrolysis of Lipids and Extraction of Total Fatty Acids<br />

from Plasma. For hydrolysis of lipids, a solution of acetonitrile<br />

(720 µL) and 5N hydrochloric acid (80 µL) was added to 50 µL of<br />

plasma in a1mLglass tube and heated at 80 °C for 1 h. After<br />

that, 50 µL of internal standard mixture (C20:0-d3, C22:0-d3, and<br />

C26:0-d4) at a concentration of 10 µg/mL was added before<br />

extraction of fatty acids with 2 mL of hexane. The hexane<br />

extraction step was repeated three times. The hexane extracts<br />

were combined and evaporated in a stream of nitrogen prior<br />

to the following derivatization procedure and reconstituted in<br />

10 µL of 50% (v/v) methanol for the enrichment procedure.<br />

Figure 1. Enrichment and SALDI-MS analysis of VLCFAs (1 µg/<br />

mL) derivatized to quaternary ammonium salt (TMAE-VLCFAs)<br />

(A-D); and MALDI-MS analysis of VLCFAs with CHCA as matrix (E).<br />

(* indicated the background peaks).<br />

RESULTS AND DISCUSSION<br />

Preparation of Quaternary Ammonium Salt Derivatives<br />

of VLCFAs. We found that VLCFAs (stock concentration 50 µg/<br />

mL) could not been detected in MALDI-MS or SALDI-MS ether<br />

in the positive or negative ion mode. Using ESI-MS, Johnson et<br />

al. 30 reported that trimethyaminoethyl-VLCFAs (TMAE-VLCFAs)<br />

afforded 8- to 12-fold greater signal intensity than the corresponding<br />

dimethylaminoethyl-VLCFAs (DMAE-VLCFAs). In order to<br />

enhance the detection in SALDI-MS, the same derivatization<br />

strategy was used in this study. The carboxylic acid groups of<br />

VLCFAs were first derivatized to DMAE-VLCFAs. The limit of<br />

detection (LOD) in MALDI-MS with CHCA as the matrix ranged<br />

from 10 to 50 µg/mL (Figure S-1B in the Supporting Information).<br />

Subsequently, DMAE-VLCFAs were reactived with methyl iodide<br />

to form TMAE-VLCFAs. The LOD of TMAE-VLCFAs in MALDI-<br />

MS (15 µg/mL) were about 10 times greater than the LOD of<br />

DMAE-VLCFAs (Figure S-1A in the Supporting Information). The<br />

DMAE-VLCFAs were not detected in the SALDI analysis. TMAE-<br />

VLCFAs are quaternary ammonium salts with permanent positive<br />

charges. The mode of ionization of quaternary ammonium salts<br />

under SALDI conditions is by dissociation of the salts in the ion<br />

source, leading to the detection of the positively charged moieties.<br />

So, the quaternary TMAE-VLCFAs iodide is expected to lead to<br />

higher SALDI-TOFMS sensitivity due to the dissociation of the<br />

iodide ion. Besides, in the process of SALDI, no additional acids<br />

were added. This ensured that the permanent positively charged<br />

VLCFAs have high priority in ionization over the uncharged or<br />

neutral compounds.<br />

Detection of Derivatized VLCFAs with Various MWCNTs<br />

for Enrichment and as SALDI Substrates (Scheme 1).<br />

MWCNTs come in a variety of diameters and lengths, depending<br />

on the growth process. In this study, MWCNTs of different<br />

diameters and lengths were chosen from commercial products<br />

with the same producing method (CVD method), carbon content,<br />

melting point, and density. The sizes of the four MWCNTs listed<br />

in Table 1 were provided by the manufacturer. The structural<br />

information was confirmed by SEM and TEM and is provided in<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6817


Figure 2. SALDI-PSD MS spectra of TMAE-VLCFAs.<br />

the Supporting Information (Figures S-2 and S-3, respectively).<br />

The SALDI-TOFMS spectra for the enrichment and SALDI<br />

analysis of TMAE-VLCFAs with various MWCNTs are depicted<br />

in Figure 1A-D and compared with the MALDI-MS with CHCA<br />

as the matrix (Figure 1E). The TMAE-VLCFAs (1 µg/mL) ions<br />

were only detected in MWCNT 636843 (Figure 1A). The TMAE-<br />

VLCFAs ions were [C20:0-TMAE] + at m/z 398.4, [C22:0-TMAE] +<br />

at m/z 426.4, [C24:0-TMAE] + at m/z 454.5, and [C26:0-TMAE] +<br />

at m/z 482.5. SALDI-MS spectra showed LODs of 0.5-1 µg/<br />

mL for MWCNT 636843, 1-5 µg/mL for MWCNT 636495 and<br />

MWCNT 636509, and 5-10 µg/mL for MWCNT 636819<br />

(Figure 1 and S-4 in the Supporting Information). The MALDI-<br />

MS analysis with CHCA as the matrix showed an LOD of 1-5<br />

µg/mL. In addition, more background peaks were observed in<br />

MALDI-MS spectra compared to MWCNT-based SALDI-TOFMS<br />

spectra. There were at least two factors that contributed to the<br />

analytical sensitivity in this study: the adsorption efficiency in the<br />

enrichment procedure and the desorption/ionization efficiency in<br />

SALDI. The adsorption efficiency of the MWCNTs was analyzed<br />

by HPLC/ESI-MS. The LC/MS chromatograms of TMAE-VLCFAs<br />

are shown in Figure S-5 in the Supporting Information. The<br />

calibration and validation data of the LC/MS method are given<br />

in Figure S-6 and Table S-1 in the Supporting Information. The<br />

adsorption efficiencies of the four MWCNTs for VLCFAs were<br />

>99.8%. There were no significant differences (p > 0.05) between<br />

the four MWCNTs. We attribute the varied desorption/ionization<br />

efficiency in SALDI to the different sizes of MWCNTs. The<br />

MWCNT 636843 had a much larger outer diameter (40-70 nm),<br />

a wider wall thickness (0-65 nm), and a shorter length (0.5-2<br />

µm) than the MWCNT 636819, which had a much smaller outer<br />

diameter (


Table 2. Concentrations and Ratios of VLCFAs in Plasma from Patients with ALD and from Normal Controls as<br />

Measured by SALDI-TOFMS<br />

Postsource Decay (PSD) of TMAE-VLCFAs. The PSD MS<br />

spectra of the derivatized VLCFAs in SALDI-TOFMS, which<br />

detected a neutral loss of trimethylamine (59 Da), were similar<br />

to those in LC-MS/MS. 30 SALDI-PSD MS spectra of TMAE-<br />

VLCFAs are recorded in Figure 2. Fragment ions were detected<br />

at m/z 339.3 for C20:0, m/z 367.4 for C22:0, m/z 395.4 for C24:0,<br />

and m/z 423.4 for C26:0. Those data demonstrate that derivatized<br />

VLCFAs not only increased the ionization efficiency in SALDI but<br />

also made the PSD easier in TOFMS. Furthermore, the neutral<br />

loss fragmentations helped confirm the presence of VLCFAs in<br />

the complex plasma matrix.<br />

Quantitative Analysis of the TMAE-VLCFAs. CNT-based<br />

SALDI has been shown to have excellent reproducibility of<br />

spectrum signals, making it an acceptable modality for quantitative<br />

analysis. 22,19,32 In addition, MALDI-MS with and without the stable<br />

isotope-labeled method has been shown to be suitable for<br />

quantifiation of small compounds in biologic samples. 38-40 In this<br />

study, we used stable isotope-labeled VLCFAs (C26:0-d4, C22:0d3,<br />

C20:0-d3) as the internal standards. The mass difference of<br />

3-4 Da between VLCFAs and isotope-labeled VLCFAs provided<br />

sufficient m/z separation space to avoid interference with<br />

isotopologues of the ions. Figure 3 shows the SALDI-TOF mass<br />

spectra of the four derivative VLCFAs. Three of the derivative<br />

isotope-labeled VLCFAs had a molar ratio of 2:1 and a peak<br />

intensity ratio of 2:1. The isotope-labeled C22:0-d3 and C26:0-d4<br />

were both evaluated in the quantification of C24:0 as the<br />

internal standards, with C26:0-d4 showing superior linearity (R 2<br />

regression coefficient varied from 0.9876 to 0.9969). The<br />

accuracy of the method was measured by determining the<br />

mean concentration at various concentrations of analyte and<br />

was calculated as percentage error of theoretical versus<br />

measured concentrations. Both the intra- and interday accuracy<br />

and precision of the method were determined by triplicate<br />

analysis of standard samples containing VLCFAs at the concentrations<br />

used to construct the calibration curves. Precision<br />

was estimated as the coefficient of variance (CV) of the<br />

analyses. The interday accuracy ranged from -5.41 to 3.23%,<br />

and the intraday accuracy ranged from -3.95 to 4.27% (Table<br />

C24:0/C22:0<br />

(ion peaks<br />

intensity ratio)<br />

C24:0/C22:0<br />

(conc ratio)<br />

C20:0 (µg/mL) C22:0 (µg/mL) C24:0 (µg/mL)<br />

1 16.23 15.21<br />

patients with ALD<br />

17.58 1.02 1.16<br />

2 17.25 11.81 22.21 1.15 1.88<br />

3 14.53 9.10 15.30 1.49 1.68<br />

4 17.89 11.47 14.70 1.19 1.28<br />

mean ± SD 16.47 ± 1.46 11.90 ± 2.5 17.45 ± 3.41 1.21 ± 0.20 1.50 ± 0.34<br />

GC/MS median a<br />

(5-90% range)<br />

20.8 (11.8-29.5) 13.7 (7.0-21.1) 1.50 (1.18-1.74)<br />

1 22.52 28.32<br />

control group<br />

19.52 0.64 0.69<br />

2 20.28 26.30 20.95 0.76 0.80<br />

3 18.86 22.13 18.25 0.76 0.82<br />

4 18.16 18.45 14.96 0.74 0.81<br />

mean ± SD 19.95 ± 1.92 23.80 ± 4.40 18.42 ± 2.55 0.72 ± 0.06 0.78 ± 0.06<br />

GC/MS median a<br />

(5-90% range)<br />

25.3 (10.5-51.0) 17.4 (8.5-35.7) 0.73 (0.48-0.89)<br />

a Value reported by Schutgens. 27<br />

S-2 in the Supporting Information). The interassay precision<br />

ranged from 1.05 to 7.52% and the intra-assay precision ranged<br />

from 0.89 to 8.85% (Table S-2 in the Supporting Information). It<br />

was found that the quantification of VLCFAs by MWCNT-based<br />

SALDI, using isotope-labeled VLCFAs as internal standards, was<br />

of acceptable accuracy and precision.<br />

Patient Plasma Analysis of VLCFAs. In order to test the<br />

clinical applicability of our assay, plasma samples collected from<br />

four subjects with X-linked adrenoleukodystrophy (ALD) and four<br />

normal controls were analyzed by MWCNT-based enrichment and<br />

SALDI-TOFMS. The SALDI-TOFMS spectra of VLCFAs in ALD<br />

patients and in normal controls are shown in Figure 4. The<br />

intensity ratio of C24:0/C22:0 was significantly higher in the ALD<br />

patients (1.02-1.49) than in the normal controls (0.64-0.76). The<br />

peaks at m/z 440.4 and 452.4 were assumed to be C23:0 and<br />

C24:1. 25,27,30,41 Table 2 summarizes the quantification data for<br />

VLCFAs concentrations and ratios in this study and compares<br />

them with the reference range reported by Schutgens. 27 In their<br />

study, samples from 109 healthy subjects and 41 patients with<br />

X-ALD were analyzed by GC-MS following derivatization by methyl<br />

ester and purification by thin-layer chromatography. The quantification<br />

results in our study were within the reported range.<br />

Usually, concentration ratios of C24:0/C22:0 are used to diagnose<br />

ALD. Stable-isotope quantification is important in the MALDI-<br />

TOFMS method because of the suppression effect of matrix<br />

interference peaks. In the MWCNT-based SALDI-TOFMS method<br />

described herein, the relative quantification can also be directly<br />

determined by measuring the ion peak intensities without the<br />

isotope-labeled internal standards calibration. Although the sample<br />

numbers are small, the concentration ratios of C24:0/C22:0<br />

correlated well with the ion peak intensity ratios of C24:0/C22:0<br />

(Pearson Correlation coefficient ) 0.88) (Figure S-8 in the<br />

(38) Koulman, A.; Petras, D.; Narayana, V. K.; Wang, L.; Volmer, D. A. Anal.<br />

Chem. 2009, 81, 7544–7551.<br />

(39) Sleno, L.; Volmer, D. A. Rapid Commun. Mass Spectrom. 2005, 19, 1928–<br />

1936.<br />

(40) Hsu, W. Y.; Lo, W. Y.; Lai, C. C.; Tsai, F. J.; Tsai, C. H.; Tsai, Y.; Lin, W. D.;<br />

Chao, M. C. Rapid Commun. Mass Spectrom. 2007, 21, 1915–1919.<br />

(41) Aveldano, M. I.; Donnari, D. Clin. Chem. 1996, 42, 454–461.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6819


Supporting Information). That finding suggests that ALD can be<br />

screened for by directly comparing the ion peak intensity ratios<br />

of C24:0/C22:0 when stable isotope-labeled internal standards are<br />

not available in the laboratory.<br />

Although GC-MS is a full quantitative method, quantification<br />

of VLCFAs in serum or plasma is typically time-consuming. 25,27<br />

For example, esterification of fatty acids with methylating reagent<br />

can take up to 16 h. Ultrasound-assisted transmethylation of fatty<br />

acids has been reported 28,29 to simplify the sample preparation<br />

procedure and improve the efficiency of transmethylation of fatty<br />

acids. However, the VLCFAs are not all detected in these reports<br />

(especially the specific markers C22:0 and C24:0 for the screening<br />

of peroxisomal disorder).<br />

CONCLUSION<br />

The described method has a number of advantages that satisfy<br />

the increased demand for selective and sensitive high-throughput<br />

techniques. Derivative VLCFAs make desorption/ionization in the<br />

SALDI process more efficient, and the neutral loss in PSD allows<br />

6820 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

for the specific identification of VLCFAs in plasma matrix.<br />

Enrichment of VLCFAs by MWCNT without eluting and evaporation<br />

procedures is an easy method for prepurification and<br />

concentration of samples. Use of VLCFAs-adsorbed MWCNTs as<br />

the SALDI substrate works well for detection in the low mass<br />

range without background peak interference and suppression.<br />

ACKNOWLEDGMENT<br />

The study was funded by a grant from the National Research<br />

Council of the Republic of China and the China Medical University<br />

Hospital (DMR 93-018).<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review March 25, 2010. Accepted July 2,<br />

2010.<br />

AC100772J


Anal. Chem. 2010, 82, 6821–6829<br />

MacroSEQUEST: Efficient Candidate-Centric<br />

Searching and High-Resolution Correlation<br />

Analysis for Large-Scale Proteomics Data Sets<br />

Brendan K. Faherty † and Scott A. Gerber* ,†,‡<br />

Department of Genetics, Dartmouth Medical School, Lebanon, New Hampshire 03756, and Norris Cotton Cancer<br />

Center, Lebanon, New Hampshire 03756<br />

Modern mass spectrometers are now capable of producing<br />

tens of thousands of tandem mass (MS/MS) spectra<br />

per hour of operation, resulting in an ever-increasing<br />

burden on the computational tools required to translate<br />

these raw MS/MS spectra into peptide sequences. In the<br />

present work, we describe our efforts to improve the<br />

performance of one of the earliest and most commonly<br />

used algorithms, SEQUEST, through a wholesale redesign<br />

of its processing architecture. We call this new program<br />

MacroSEQUEST, which exhibits a dramatic improvement<br />

in processing speed by transiently indexing the array of<br />

MS/MS spectra prior to searching FASTA databases. We<br />

demonstrate the performance of MacroSEQUEST relative<br />

to a suite of other programs commonly encountered in<br />

proteomics research. We also extend the capability of<br />

SEQUEST by implementing a parameter in MacroSE-<br />

QUEST that allows for scalable sparse arrays of experimental<br />

and theoretical spectra to be implemented for highresolution<br />

correlation analysis and demonstrate the<br />

advantages of high-resolution MS/MS searching to the<br />

sensitivity of large-scale proteomics data sets.<br />

Mass spectrometry (MS) coupled with computer-assisted<br />

database spectral matching has evolved into a cornerstone<br />

technology that drives research for the field of proteomics. 1 Recent<br />

developments in MS instrumentation have resulted in commercial<br />

mass spectrometers capable of generating tens of thousands of<br />

tandem mass spectra (MS/MS) per penultimate online reversephase<br />

liquid chromatography (LC) separation. 2-4 When coupled<br />

with biochemical prefractionation methods, a complete data set<br />

for a proteomics experiment (including technical and biological<br />

replicates) can consist of millions of MS/MS spectra per biological<br />

* To whom correspondence should be addressed. E-mail: scott.a.gerber@<br />

dartmouth.edu.<br />

† Dartmouth Medical School.<br />

‡ Norris Cotton Cancer Center.<br />

(1) Aebersold, R.; Mann, M. Nature 2003, 422, 198–207.<br />

(2) Makarov, A.; Denisov, E.; Kholomeev, A.; Balschun, W.; Lange, O.; Strupat,<br />

K.; Horning, S. Anal. Chem. 2006, 78, 2113–2120.<br />

(3) Olsen, J. V.; Schwartz, J. C.; Griep-Raming, J.; Nielsen, M. L.; Damoc, E.;<br />

Denisov, E.; Lange, O.; Remes, P.; Taylor, D.; Splendore, M.; Wouters, E. R.;<br />

Senko, M.; Makarov, A.; Mann, M.; Horning, S. Mol. Cell. Proteomics 2009,<br />

8, 2759–2769.<br />

(4) Second, T. P.; Blethrow, J. D.; Schwartz, J. C.; Merrihew, G. E.; MacCoss,<br />

M. J.; Swaney, D. L.; Russell, J. D.; Coon, J. J.; Zabrouskov, V. Anal. Chem.<br />

2009, 81, 7757–7765.<br />

sample. Importantly, although the absolute sensitivity of new<br />

instruments to detect a single peptide species has improved, it<br />

has been suggested that much of the credit for increased depth<br />

of peptide and protein coverage, and improved detection of<br />

substoichiometric species belongs to the increased rate of MS/<br />

MS spectral acquisition of these instruments, which allows them<br />

to penetrate transient rasters of precursor ions to greater ion peak<br />

depth per chromatographic unit time. 5 Indeed, the number of<br />

candidate precursor ions (MS1 features) is often at least an order<br />

of magnitude greater than the number of MS/MS sequencing<br />

events in a typical LC-MS analysis. 6 Given the current level of<br />

success with this strategy, it is only reasonable to predict that<br />

this trend of increasing MS/MS bandwidth to improve overall<br />

peptide identification rates per sample will continue, which places<br />

an additional burden on the computational tools required to search<br />

these larger data sets.<br />

A single raw MS/MS spectrum is translated into a peptide<br />

spectral match (PSM) through the use of algorithms that first<br />

search translated genomic databases from an organism of interest<br />

for candidate peptides by a defined enzyme specificity (based on<br />

the protease used for digestion) and precursor mass (based on<br />

the MS1 feature mass from which the MS/MS spectrum was<br />

derived and a desired mass precision). 7 Other parameters (fixed<br />

protein modifications, variable post-translational modifications,<br />

number of missed enzyme cleavage loci, maximum and/or<br />

minimum peptide size, etc.) may also be included. This search<br />

results in a list of candidate peptides that may contain the correct<br />

peptide sequence from which the MS/MS spectrum was derived.<br />

These candidate peptides are then evaluated for correctness by<br />

comparing the observed MS/MS spectrum with a dynamically<br />

generated theoretical spectrum for each candidate peptide and<br />

given a score that reflects the quality of their match. These scores<br />

are then ranked and, in some cases, further evaluated 8 before<br />

reporting the “best” candidate peptide match (PSM) for the MS/<br />

MS spectrum. In a collection of MS/MS spectra from a single<br />

LC-MS run, a series of LC-MS runs for a given sample or from<br />

(5) Haas, W.; Faherty, B. K.; Gerber, S. A.; Elias, J. E.; Beausoleil, S. A.;<br />

Bakalarski, C. E.; Li, X.; Villen, J.; Gygi, S. P. Mol. Cell. Proteomics 2006,<br />

5, 1326–1337.<br />

(6) Hoopmann, M. R.; Finney, G. L.; MacCoss, M. J. Anal. Chem. 2007, 79,<br />

5620–5632.<br />

(7) Nesvizhskii, A. I. Methods Mol. Biol. 2007, 367, 87–119.<br />

(8) Kall, L.; Canterbury, J. D.; Weston, J.; Noble, W. S.; MacCoss, M. J. Nat.<br />

Methods 2007, 4, 923–925.<br />

10.1021/ac100783x © 2010 American <strong>Chemical</strong> Society 6821<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/28/2010


a series of different samples, the general process is iterated<br />

thousands, if not millions, of times.<br />

Today, the modern proteomics researcher can choose from<br />

several computational tools to translate MS/MS spectra into<br />

PSMs, including SEQUEST, 9 Mascot, 10 X!Tandem, 11,12 and<br />

OMSSA, 13 among others. A number of studies have been<br />

performed that evaluate the analytical performance of these<br />

algorithms, with regards to precision and sensitivity of the PSM<br />

collections they generate. 14-16 While it appears that certain<br />

algorithms perform slightly better or worse than others based on<br />

the nature of the sample (e.g., phosphorylation, enzyme digest),<br />

the type of mass spectrometer used to generate these MS/MS<br />

data, and the mechanism of peptide fragmentation, etc., they are<br />

in general more similar than they are different, at least, in terms<br />

of the nature of the PSM collections they produce. However, given<br />

the recent trend toward larger and larger numbers of MS/MS<br />

spectra per experiment, the relative processing speed of these<br />

algorithms has become an important practical consideration, as<br />

some of these algorithms perform significantly slower than others.<br />

In particular, SEQUEST has lagged significantly behind, although<br />

limited efforts to improve performance by parallelization have been<br />

reported. 17 In general, the most basic solution to any large<br />

discrepancies in performance (or “productivity”, defined as the<br />

number of MS/MS spectra processed per unit computational time)<br />

has been to index the protein database by predigesting the protein<br />

sequences to peptides, then indexing each candidate peptide by<br />

precursor mass. While these peptide indices do result in significant<br />

performance gains that in general level the playing field<br />

among algorithms, there are drawbacks to using indexed databases.<br />

For example, the use of an indexed database as opposed<br />

to the use of a FASTA database results in a significant expansion<br />

of disk memory, requires separate indices for each enzyme used<br />

for digestion, and is problematic when considering post-translational<br />

modifications, all of which is repeated and exacerbated for<br />

each organism and/or release version of the organism-specific<br />

genome sequence. A recent report attempted to reconcile the<br />

issues of both performance and indexed database files by reading<br />

the target FASTA database once for the first MS/MS spectrum<br />

to be analyzed, indexing “on-the-fly” and storing this peptide index<br />

in fast CPU memory for use in finding candidates for subsequent<br />

MS/MS spectra. 18<br />

An alternative approach that addresses the performance and<br />

convenience issues associated with both indexed database files<br />

(9) Eng, J. K.; Mccormack, A. L.; Yates, J. R. J. Am. Soc. Mass Spectrom. 1994,<br />

5, 976–989.<br />

(10) Perkins, D. N.; Pappin, D. J.; Creasy, D. M.; Cottrell, J. S. Electrophoresis<br />

1999, 20, 3551–3567.<br />

(11) Craig, R.; Beavis, R. C. Bioinformatics 2004, 20, 1466–1467.<br />

(12) Craig, R.; Beavis, R. C. Rapid Commun. Mass Spectrom. 2003, 17, 2310–<br />

2316.<br />

(13) Geer, L. Y.; Markey, S. P.; Kowalak, J. A.; Wagner, L.; Xu, M.; Maynard,<br />

D. M.; Yang, X.; Shi, W.; Bryant, S. H. J. Proteome Res. 2004, 3, 958–964.<br />

(14) Bakalarski, C. E.; Haas, W.; Dephoure, N. E.; Gygi, S. P. Anal. Bioanal.<br />

Chem. 2007, 389, 1409–1419.<br />

(15) Elias, J. E.; Haas, W.; Faherty, B. K.; Gygi, S. P. Nat. Methods 2005, 2,<br />

667–675.<br />

(16) Kapp, E. A.; Schutz, F.; Connolly, L. M.; Chakel, J. A.; Meza, J. E.; Miller,<br />

C. A.; Fenyo, D.; Eng, J. K.; Adkins, J. N.; Omenn, G. S.; Simpson, R. J.<br />

Proteomics 2005, 5, 3475–3490.<br />

(17) Sadygov, R. G.; Eng, J.; Durr, E.; Saraf, A.; McDonald, H.; MacCoss, M. J.;<br />

Yates, J. R. J. Proteome Res. 2002, 1, 211–215.<br />

(18) Park, C. Y.; Klammer, A. A.; Kall, L.; MacCoss, M. J.; Noble, W. S. J.<br />

Proteome Res. 2008, 7, 3022–3027.<br />

6822 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

as well as indexing the protein database in CPU memory is to<br />

consider an entire set of MS/MS spectra to be searched (e.g., a<br />

complete LC-MS/MS run) as an array of precursor masses and<br />

spectra and to transiently index these MS/MS spectra by precursor<br />

mass at the beginning of a search. This allows for efficient<br />

matching of candidate peptides during a single in silico digestion<br />

pass through a FASTA database. Because the number of MS/<br />

MS spectra in an LC-MS/MS run is small relative to the number<br />

of candidate peptides in a typical organism database, this indexing<br />

step is extremely fast and can easily be implemented at the launch<br />

of a search. Furthermore, additional performance gains can be<br />

realized through elimination of redundant candidate peptide<br />

spectral processing steps that occur when single MS/MS spectra<br />

are considered in the absence of other MS/MS spectra that have<br />

overlapping candidate peptide search spaces. Indeed, the benefits<br />

of this “candidate-centric” (as opposed to “spectrum-centric”)<br />

searching approach have been described explicitly, first in<br />

principle by Edwards and Lippert 19 and later in practice by Tabb<br />

and co-workers. 20<br />

In the present work, we demonstrate the general utility of this<br />

“candidate-centric” approach by modifying the commercial version<br />

of SEQUEST to perform searches in a candidate-centric fashion,<br />

resulting in a program we call MacroSEQUEST. MacroSEQUEST<br />

reads FASTA-formatted protein databases and returns PSMs for a<br />

collection of MS/MS spectra in a fraction of the time it takes legacy<br />

SEQUEST. We compare the performance of MacroSEQUEST to a<br />

suite of other commonly used database search engines. Finally, we<br />

demonstrate a useful application for the performance gains associated<br />

with MacroSEQUEST by leveraging this increase in speed to perform<br />

high-resolution MS/MS correlation analysis and describe the benefits<br />

to classical SEQUEST scores associated with high-mass precision<br />

fragment ion searching.<br />

MATERIALS AND METHODS<br />

Sample Preparation. Peptide synthesis was performed by<br />

New England Peptide (Gardner, MA). Yeast cells were harvested<br />

during logarithmic growth, pelleted, and lysed by addition of SDS-<br />

PAGE pH 8.1 sample buffer (3× volume buffer/pellet weight) and<br />

bead beating at 4 °C. HeLa cell lysate was prepared by harvesting<br />

a confluent 15 cm dish of HeLa cells by trypsinization, washing<br />

2× in PBS, and addition of 4 mL of lysis buffer (0.5% Triton X-100,<br />

50 mM Tris pH 8.1, 150 mM NaCl, 1 mM MgCl2, Roche Minicomplete<br />

protease inhibitors), followed by sonication and<br />

clarification of the lysate in a centrifuge (14 000g) for 10 min<br />

at 4 °C. Both protein preps were reduced by addition of DTT<br />

to 5 mM and incubation in a water bath at 55 °C for 20 min,<br />

followed by alkylation of cysteines in 12.5 mM iodoacetamide<br />

at room temperature for 45 min. After the alkylation reaction<br />

was quenched (addition of 2.5 mM DTT), the lysates were<br />

aliquotted, snap frozen in liquid nitrogen, and stored at -80<br />

°C until use. To prepare the protein digests, an aliquot of cell<br />

lysate was warmed rapidly under warm water and mixed with<br />

SDS-PAGE sample buffer to a protein concentration of ∼0.25<br />

(19) Edwards, N.; Lippert, R. In Algorithms in Bioinformatics: Second International<br />

Workshop, WABI 2002, Rome, Italy, September 17-21, 2002 Proceedings;<br />

Guigo, R., Gusfield, D., Eds.; Springer: Berlin, Germany, 2002; Vol. 2452,<br />

pp 68-81.<br />

(20) Tabb, D. L.; Narasimhan, C.; Strader, M. B.; Hettich, R. L. Anal. Chem.<br />

2005, 77, 2464–2474.


mg/mL protein, followed by separation on 2-well, 4-12%<br />

NOVEX minigels (for 2D separations) with a protein marker<br />

in the narrow lane. Proteins were visualized with Coomassie<br />

blue, and the region between 80 and 125 kDa was excised,<br />

destained, and digested with trypsin. Peptide samples were<br />

analyzed on an LTQ Orbitrap (ThermoFisher Scientific, Bremen,<br />

Germany) per established procedures. 5 LTQ-Orbitrap<br />

.RAW files were converted to .mzXML files using ReAdW.exe<br />

(version 4.0, http://sourceforge.net/projects/sashimi/files/).<br />

Peptide precursor mass assignments were adjusted postacquisition<br />

with in-house software that (i) updates MS2 scan headers<br />

with high-mass accuracy precursor information from MS1 scans<br />

and (ii) averages multiple observations of these precursor<br />

values across each peptide chromatographic elution profile.<br />

MacroSEQUEST. MacroSEQUEST was written from scratch<br />

in ANSI C using standard libraries and compiled with GCC version<br />

4.1.2 on a Unix platform using the SEQUEST2.8 source code as<br />

a guide. Input is provided as command line arguments. Experimental<br />

data is read in .DTA file format, and candidate peptides<br />

are read from FASTA-formatted protein databases. Macro outputs<br />

SEQUEST-like .out files for each input spectrum.. To measure<br />

the efficiency of the “candidate-centric” method, we created a<br />

version of Macro (MacroParser) in which the Xcorr scoring<br />

function was replaced with a random number generator.<br />

Database Searching. Searches were performed using the<br />

latest database builds from the yeast proteome (Saccharomyces<br />

Genome Database, http://www.yeastgenome.org/) or the human<br />

proteome (UniProtKB, http://www.uniprot.org/). Target-decoy<br />

databases were generated using in-house scripts that reverse each<br />

protein sequence, label decoys proteins with a specific identifier,<br />

and append the decoy sequences to the end of a forward<br />

database. 21 Unless otherwise stated, all searches were performed<br />

with a 1.1 Da precursor mass tolerance and filtered to ±1.5 ppm<br />

precursor mass measurement accuracy (Figure S-1 in the Supporting<br />

Information); Xcorr and dCn (delta-correlation) cutoff<br />

values were adjusted to achieve a false discovery rate (FDR) of<br />

less than 1% (Table S-1 in the Supporting Information). Only the<br />

yeast searches were conducted with semitryptic enzyme specificity;<br />

all other searches were performed with full trypsin specificity.<br />

The maximum number of missed cleavages for all searches was<br />

set to three. All searches were also performed with acetamidemodified<br />

Cys as a static modification (+57.021 461 Da) and with<br />

oxidized Met as a variable modification (±15.991 915 Da); for<br />

phosphorylation searches, Ser, Thr, and Tyr were allowed to vary<br />

by ±79.966 331 Da. Variable modifications were limited to a<br />

maximum of 3 per peptide, where applicable. Refinement or<br />

iterative searches were not permitted when using X!Tandem and<br />

OMSSA. No multithreading was used for any search algorithm.<br />

RESULTS<br />

Candidate-Centric Database Spectral Matching. Historically,<br />

database spectral matching algorithms such as SEQUEST<br />

dealt with very small numbers of MS/MS spectra acquired per<br />

experiment, from as few as tens to at most one hundred spectra<br />

in a typical analysis. 9 Individual MS/MS spectra were then<br />

matched with candidate peptides by searching through a target<br />

database. In general, the focus for algorithm development has<br />

(21) Elias, J. E.; Gygi, S. P. Nat. Methods 2007, 4, 207–214.<br />

been on the steps involved in this single iteration and on improving<br />

the sensitivity, accuracy, and productivity of a single analysis, with<br />

the expectation that this “optimized” core process is iterated until<br />

PSMs have been generated for all MS/MS spectra in an analysis<br />

which, while computationally inefficient, was adequate given the<br />

historical context of the problem space. A generalized scheme<br />

describing this classical workflow for SEQUEST is depicted in<br />

Figure 1a. However, this workflow fails to recognize potential<br />

elements of relatedness between MS/MS spectra that are collected<br />

together. Clearly, all of these spectra require matching to<br />

candidate peptides in the same database; thus, presenting the<br />

entire collection of spectra to the database (or vice versa) as a<br />

“single analysis” has the potential to reduce areas of overhead<br />

associated with repeating a single process, such as digesting a<br />

database, thousands of times (Figure 1b). Our analysis of the work<br />

distribution for SEQUEST when searching a target-decoy human<br />

UniProtKB protein database (∼150 000 proteins) against 10 000<br />

spectra reveals that only a very small portion of the actual search<br />

time is spent on scoring candidate PSMs (∼1%), while the<br />

remainder is spent parsing the database (Figure 1c). Although<br />

the database must be read once in order to generate a PSM for<br />

a single spectrum, it does not need to be reread to search other<br />

MS/MS spectra under the same set of parameters if those<br />

additional spectra are considered simultaneously. In this way,<br />

legacy SEQUEST wastes a significant amount of search time when<br />

large numbers of MS/MS spectra are searched against large<br />

databases.<br />

We sought to address these issues through a wholesale<br />

rearchitecting of the SEQUEST scoring components in order to<br />

conduct them in a candidate-centric fashion and named the<br />

resulting program MacroSEQUEST or “Macro” for short. The<br />

Macro workflow is described in Figure 1b. Similar to the candidate-centric<br />

algorithm DBDigger, 20 a single Macro process at<br />

launch is fed a parameter set that includes a path to a collection<br />

of MS/MS spectra. This collection of spectra is first read into<br />

memory and indexed by precursor ion mass to create a “spectral<br />

data array” that includes preprocessing for each experimental<br />

spectrum and memory mapping for candidate peptide sequences,<br />

protein references, search scores, etc. After this data structure is<br />

created, Macro begins parsing the FASTA database of interest<br />

by digesting it based on a desired enzyme cleavage specificity.<br />

For each newly digested peptide, Macro calculates its mass and<br />

checks it against the spectral array, including a given mass<br />

tolerance; if a peptide falls outside these boundaries, Macro<br />

discards it and proceeds to the next logical peptide in the database.<br />

If, however, a peptide falls into one or more spectral “bins” in the<br />

spectral array, Macro enters a scoring loop in which a reimplementation<br />

of a fast SEQUEST cross correlation algorithm 22,23 is<br />

performed on the candidate peptide for each MS/MS spectrum<br />

that falls within the peptide’s desired precursor mass tolerance,<br />

and plugs these scores and associated peptide/protein information<br />

into the spectral array. The scoring loop then closes by returning<br />

to database parsing and to the next logical peptide in the database;<br />

Macro continues this cycle of peptide generation, spectral array<br />

scanning, and scoring until it reaches the end of the FASTA<br />

(22) Eng, J. K.; Fischer, B.; Grossmann, J.; Maccoss, M. J. J. Proteome Res. 2008,<br />

7, 4598–4602.<br />

(23) Venable, J. D.; Xu, T.; Cociorva, D.; Yates, J. R. Anal. Chem. 2006, 78,<br />

1921–1929.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6823


Figure 1. Comparison of legacy SEQUEST and MacroSEQUEST (Macro) workflows. (a) Scheme of the serial nature of SEQUEST. To analyze<br />

n number of spectra, SEQUEST must be executed n number of times, parsing the database with each instance and using both preliminary and<br />

XCorr scores. (b) Scheme of the Macro workflow. Macro considers all spectra in a single analysis and therefore requires only one pass through<br />

a target database during a search. (c) Analysis of the time distribution for SEQUEST and Macro searches of 10 000 spectra with a target-decoy<br />

human (UniProt) database. SEQUEST spends approximately 1% of the search time on scoring functions while Macro spends only 13% of the<br />

search time parsing the database. The inset shows an expanded view of the Macro time distribution for the first 20 min of the search time.<br />

database. Macro then concludes by calculating final score differences,<br />

etc. for the top-ranked candidate PSMs and writing this<br />

information to result files. Because Macro was developed from<br />

the original SEQUEST source code, the primary scoring metric<br />

Xcorr in Macro is identical to those created by legacy SEQUEST,<br />

a major difference between the two is that Macro no longer<br />

performs Sp scoring as a preliminary scoring step but instead<br />

calculates Xcorr for all candidate PSMs, owing to the computationally<br />

fast, non-FFT correlation algorithm. Macro spends significantly<br />

less time parsing the target-decoy human database than<br />

SEQUEST during a search of the same 10 000 spectra (Figure<br />

1c, inset), and almost 87% of the total run time performing scoring<br />

functions.<br />

Because such a small fraction of the actual SEQUEST process<br />

is spent scoring candidate PSMs and because this cycle is repeated<br />

in its entirety for each spectrum to be searched, SEQUEST’s<br />

productivity (number of spectra searched/unit time spent searching)<br />

is relatively flat as a function of the number of spectra<br />

searched (Figure 2). Macro, however, benefits from having more<br />

spectra to search by distributing the fixed time cost associated<br />

with a single digestion and parsing pass through the database<br />

across many spectra until the time spent parsing the database is<br />

small relative to the time spent scoring candidate PSMs, at which<br />

point Macro’s search productivity flattens out. This improvement<br />

in productivity as a function of number of spectra searched is<br />

clearly depicted in Figure 2, where searches of a target-decoy yeast<br />

(Figure 2a), human (Figure 2b), and human databases with<br />

dynamic phosphorylation on serine, threonine, and tyrosine<br />

(Figure 2c) using Macro all outperform legacy SEQUEST by 102×,<br />

107×, and 42×, respectively, when searching 10 000 MS/MS<br />

6824 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

spectra. Note that the increase in parsing logic associated with<br />

determining dynamic protein post-translational modifications, such<br />

as phosphorylation, substantially compresses and flattens the<br />

Macro productivity curve and extends the point in numbers of<br />

spectra at which Macro achieves maximum productivity to well<br />

beyond the 10 000 mark.<br />

In order to assess the level of spectral array processing<br />

overhead in Macro, we generated a “MacroParser” version of the<br />

program that replaces the XCorr scoring functions with a simple<br />

random number generator. This allowed us to subtract the time<br />

it takes Macro to parse the database and scan the array of spectra<br />

from the total search time, which generated an ideal power fit as<br />

a function of the number of spectra searched (Figure 2a-c).<br />

Comparison of MacroSEQUEST to Other Contemporary<br />

Tools. Although it was among the earliest algorithms to be<br />

adopted into widespread use, SEQUEST is now not the only<br />

program available to retrieve candidate PSMs from FASTA<br />

databases. Other commercial (Mascot) and noncommercial<br />

(OMSSA, X!Tandem) programs are also designed to execute<br />

database searches, although the core components and matching<br />

algorithms differ substantially. To establish Macro’s performance<br />

rank among programs that are commonly encountered in proteomics<br />

research laboratories and protein identification core<br />

facilities, we generated productivity curves for searches against<br />

our target-decoy human database for Macro, SEQUEST, Mascot,<br />

OMSSA, and X!Tandem. We also noted that all three of these<br />

programs have been written to take advantage of multiple physical<br />

and logical cores now increasingly common in modern CPU<br />

architectures, including the Intel Conroe chip used in this<br />

comparison. In order to cleanly reconcile differences in perfor-


Figure 2. Productivity of SEQUST and Macro under typical experimental conditions. We created three biological samples to use in testing the<br />

performance of Macro (green circles) versus legacy SEQUEST (orange circles) in common applications: (a) yeast 80-130 kDa protein digest,<br />

(b) human 80-130 kDa protein digest, and (c) human phosphorylation samples, all analyzed by a 90 min gradient LC-Orbitrap-MS/MS. Variable,<br />

random portions of each full collection of MS/MS spectra were searched to define the productivity (number of MS/MS spectra searched/minute<br />

of search time) of each search type as shown in the left panel. A version of Macro that does not produce scores and is useful as a measurement<br />

of database parsing work, MacroParser, was used to calculate the time spent on scoring by difference. In the right panel, the Macro, MacroParser,<br />

and calculated Macro Scoring productivity distributions are plotted as the log for the variable, random portion search sizes versus search time.<br />

A power function was fitted to the Macro Scoring distribution using a nonlinear least-squares method in R. Note that Macro is 102×, 107×, and<br />

42× faster than SEQUEST under each search condition, respectively.<br />

mance due to multithreading, we disabled one of the two cores<br />

on our Conroe prior to performing the comparison, the results of<br />

which are depicted in Figure 3. At peak productivity, Macro<br />

performs equivalent to OMSSA, about 20% faster than Mascot,<br />

and twice as fast X!Tandem, with legacy SEQUEST lagging far<br />

behind the others.<br />

High-Resolution MS/MS Correlation Analysis. In light of<br />

the significant performance improvement of Macro over legacy<br />

SEQUEST, we reasoned that this additional speed could be leveraged<br />

to also produce gains in the quality of spectral matches by creating<br />

sparse arrays of MS/MS fragment ions from high-resolution, highmass<br />

accuracy instruments such as the LTQ Orbitrap 2 and, in<br />

particular, the LTQ Orbitrap Velos 3 and by performing full correlation<br />

analyses on them to produce Xcorr scores, a feature not currently<br />

available in SEQUEST. Given the novelty of Xcorr and dCn values<br />

derived from high-resolution MS/MS spectra, we were also interested<br />

in evaluating the qualitative impact that variable search “resolution”<br />

might have on these scores.<br />

To do this, we created a parameter for Macro that defines the<br />

bin width of the arrays (in m/z) used for correlation analysis and<br />

used this factor to scale the number of bins that are generated<br />

within the range of observed fragment ions during MS/MS<br />

spectrum preprocessing steps. Macro then populates these bins<br />

with normalized ion intensity values using logic consistent with<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6825


Figure 3. Comparison of Macro performance with contemporary<br />

database searching algorithms. The same data set in Figure 2b was<br />

researched with other commonly used database searching algorithms,<br />

including Mascot, OMSSA, and X!Tandem. All algorithms were run<br />

in single-threaded mode to provide an accurate comparison.<br />

legacy SEQUEST and performs the fast Xcorr preprocessing math<br />

on the resultant sparse array. Given the mass precision and<br />

resolution of ion traps, MS/MS data from such instruments<br />

normally defines these bin widths as 1 m/z wide, and the spectral<br />

preprocessing logic of SEQUEST determines which ions fall into<br />

which m/z bin(s). Figure 4a depicts an MS/MS spectrum<br />

collected in an LTQ Orbitrap with a resolution setting of 15 000.<br />

Figure 4b describes this spectrum preprocessed for fast Xcorr<br />

analysis by Macro with a bin width of 1 m/z, while Figure 4c<br />

depicts an array from the same spectrum but with bin widths of<br />

0.01 m/z. Note that, in the insets, binning logic in Macro reduced<br />

the number of available ions for matching across a 2 m/z-wide<br />

window in the original spectrum from 6 to 2 when 1 m/z-wide<br />

bins were used versus 0.01 m/z-wide bins. Clearly, the higher<br />

“resolution” of the array in Figure 4c allows for more accurate<br />

ion-to-bin assignments and consequently a more robust discrimination<br />

between ion fragments of similar m/z. This is also apparent<br />

in the significant reduction in the average magnitude of negative<br />

values in the high-resolution preprocessed array (Figure 4c),<br />

which spreads these anticorrelations out over a much larger bin<br />

space. Similar to the preprocessing steps that were performed<br />

for experimental spectra, theoretical spectra for each candidate<br />

peptide were created using the same bin width scaling factor, and<br />

Xcorr values were then calculated from the dot product of the<br />

two arrays.<br />

We then tested the utility of these scalable MS/MS arrays by<br />

collecting data from our LTQ Orbitrap at relatively high resolution<br />

and mass accuracy (R ) 15 000). We analyzed our HeLa cell<br />

80-120 kDa protein lysate digest over a 90 min LTQ Orbitrap<br />

gradient, during which a total of 5 725 MS/MS spectra were<br />

collected. We then used Macro to iteratively search this run with<br />

fragment ion bin widths of 1.0, 0.75, 0.5, 0.25, 0.1, 0.075, 0.05, 0.025,<br />

and 0.01 m/z. Because the size of the arrays scales inversely with<br />

bin width, there is a modest but significant speed penalty<br />

6826 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

associated with high-resolution scoring. On our Intel Conroe test<br />

platform fitted with 8Gb of RAM, we observed a 3× increase in<br />

search time using bin widths of 0.025 m/z relative to unit<br />

resolution and a 6× increase in search time when using bin widths<br />

of 0.01 m/z. However, we also observed a significant increase in<br />

the total number of true positive (TP) PSMs at a 1% false discovery<br />

rate (FDR). Figure 5a depicts the total number of TP PSMs for<br />

this run as a function of search bin width. We observed a 24%<br />

increase in the number of TP PSMs when using bin widths of<br />

0.025 m/z versus 1.0 m/z bin widths. We considered the possibility<br />

that simply limiting the number of candidate PSMs by using a<br />

very narrow precursor ion tolerance may allow these “rescuable”<br />

PSMs to rise in rank. However, consistent with previous reports, 24,25<br />

we only observed a very slight increase in TP assignments in a<br />

search limited to ±10 ppm precursor ion mass measurement<br />

accuracy (Figure 5b). This suggests that the sensitivity gains we<br />

observe by high-resolution fragment ion searching represent<br />

PSMs that are not accessible by simply limiting the number of<br />

candidate PSMs in a search. Although these mediocre yet<br />

”rescuable” MS/MS spectra are being considered in the context<br />

of fewer competing candidates during a narrow window precursor<br />

ion search, their inherently poor correlation characteristics do not<br />

sufficiently enhance their differential Xcorr rank to allow for their<br />

rescue in this search mode.<br />

This increase in true positive assignments can be ascribed to<br />

a combination of features of merit associated with the higher<br />

resolution versions of Xcorr and dCn. Although we did not see an<br />

appreciable change in Xcorr for TP PSMs when searching at<br />

higher fragment ion resolution, we noted that false positive (FP)<br />

PSMs exhibited a decrease in their median Xcorr values (from<br />

1.6 to 1.2 for 1.0 and 0.025 m/z-wide bins, respectively; Figure<br />

5c). This allows for significantly lower Xcorr cutoff values to be<br />

used when establishing a 1% FDR. From Figure 5c, clearly many<br />

of the “rescued” TP PSMs come from low Xcorr scores that are<br />

otherwise inaccessible due to the extension of the FP PSM<br />

distribution into that score space. Equally dramatic was the change<br />

in TP PSM dCn scores, whose median values increased from 0.25<br />

to 0.42 across the entire data set (Figure 5d); for those “rescued”<br />

TP PSMs, the median dCn value jumped from 0.06 to 0.33. As<br />

dCn is a direct measure of the relative Xcorr score separation<br />

between the highest-ranked PSM and its nearest neighbor (de<br />

facto a FP PSM), our observations of overall reduced FP Xcorr<br />

scores and increased TP dCn scores are consistent with this<br />

relationship.<br />

To further describe the nature of some of these “rescued” TP<br />

PSMs, we examined their scores and individual PSM rankings.<br />

Figure 6a depicts the top two ranked PSMs for a particular MS/<br />

MS spectrum that yielded an excellent Xcorr score (4.3), but a<br />

very low dCn score of 0.005 when scored with a 1.0 m/z bin width.<br />

Note that these PSMs differ only by a K/Q (mass difference )<br />

0.036 Da) in the middle of the peptide sequences. Although<br />

distinguishable by Orbitrap precursor ion mass precision, for<br />

correlation analysis at unit resolution fragment ion correlation,<br />

this difference is transparent and results in almost identical Xcorr<br />

(24) Beausoleil, S. A.; Jedrychowski, M.; Schwartz, D.; Elias, J. E.; Villen, J.; Li,<br />

J.; Cohn, M. A.; Cantley, L. C.; Gygi, S. P. Proc. Natl. Acad. Sci. U.S.A.<br />

2004, 101, 12130–12135.<br />

(25) Hsieh, E. J.; Hoopmann, M. R.; Maclean, B.; Maccoss, M. J. J. Proteome<br />

Res. 2010, 9, 1138–1143.


Figure 4. High-resolution MS/MS spectra and Macro. (a) A typical, raw high-mass accuracy MS/MS spectrum generated using an LTQ-<br />

Orbitrap (R ) 15 000). (b) The raw spectrum in part a preprocessed for fast Xcorr using 1 m/z-wide bins to define the spectrum array. (c) The<br />

same spectrum preprocessed for fast Xcorr using 0.01 m/z-wide bins. Note in the inset that multiple ion features persist in a 2 m/z-wide space<br />

in the higher resolution array, while they are compressed into two features when searches are performed at unit resolution.<br />

scores for these two sequences. However, when searched by<br />

Macro using 0.025 m/z bin widths, this K/Q difference is readily<br />

distinguished by dCn, likely for all singly and doubly charged<br />

fragment ions from this quadruply charged precursor that contain<br />

the Lys residue. Figure 6b describes the general behavior of all<br />

candidate PSMs for this MS/MS spectrum at the two fragment<br />

ion tolerances. It is worth noting again in this plot that the<br />

discriminating power of these high-resolution correlations lies<br />

predominantly in positive overlap between the theoretical and<br />

experimental arrays and not in the “signal-to-noise” of this overlap<br />

relative to neighboring offsets (±75 m/z equivalents), an important<br />

aspect of the unit resolution Xcorr. This can be observed in the<br />

large decrease in the number of negative Xcorr values for lowranked<br />

PSMs between the high- and low-resolution searches.<br />

Indeed, of the 819 “rescued” PSMs, 26% of them were for<br />

peptide sequences that were correctly assigned (top-ranked) in<br />

the lower resolution correlation analysis but with extremely low<br />

dCn/Xcorr combinations such that they precluded a definitive<br />

assignment when score cutoffs were applied to achieve a 1% FDR.<br />

The remaining rescued peptide matches were reranked from<br />

lower ranks to the top-ranked candidate PSM. Figure 6c depicts<br />

the distribution across the entire run of original candidate PSM<br />

ranks when searches were performed with 1.0 m/z-wide fragment<br />

ion bins that are “rescued” when the search is done at 0.025 m/z<br />

bin widths. Of those PSMs that are reranked between the two<br />

searches, the median rank in the low-resolution search is 9; the<br />

lowest reranked candidate PSM moved from the 4 517th position<br />

to the top-ranked spot when searched at high resolution. Although<br />

this MS/MS spectrum is assigned a PSM (AASVHTVGEDTEET-<br />

PHR, M + 4H + , MMA ) 0.1 ppm) with an Xcorr of only 0.67,<br />

the dCn is 0.08 (dCn = 0.001 at unit fragment ion resolution).<br />

This peptide is from the protein hPOP1 with a molecular weight<br />

of 115 kDa, which is in the range of molecular weights (80-125<br />

kDa) of SDS-PAGE-fractionated human cell lysate from which<br />

the sample was derived. Figure 6d displays a reciprocal plot of<br />

the endogenous, rescued MS/MS spectrum from the lysate<br />

with the matched ions noted and a mirrored spectrum from a<br />

synthetic peptide that we analyzed under identical LC-MS/MS<br />

conditions, confirming this rescued sequence as being a correct<br />

match. Additionally, there were two other peptides from the same<br />

protein that were identified in both the high- and low-resolution<br />

correlation analyses with scores and mass measurement accuracy<br />

assignments that pass cutoffs (KTHQPSDEVGTSIEHPR, M +<br />

4H + , Xcorr ) 3.78, dCn ) 0.43, MMA ) -0.2 ppm and<br />

IPILLIQQPGK, M + 2H + , Xcorr ) 3.31, dCn ) 0.51, MMA )<br />

0.1 ppm), supporting the likelihood that this PSM is a valid<br />

match.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6827


Figure 5. Behavior of primary scoring metrics for high-resolution correlation analysis. (a) Histogram depicting the sensitivity of an LC-Orbitrap-<br />

MS/MS analysis as a function of fragment ion bin widths at FDR < 1%. (b) High-resolution correlation analysis rescues false negatives that are<br />

inaccessible by high-mass accuracy precursor ion searching alone. (c) ROC plot using XCorr filtering during 1.0, 0.25, and 0.025 m/z-wide bin<br />

searches. Insets show histograms of the behavior of Xcorr during 1.0 and 0.025 m/z-wide bin searches. The 1.0 m/z TP distribution is blue,<br />

while the 1.0 m/z FP distribution is red. The 0.025 m/z TP distribution is green, while the 0.025 m/z FP distribution is orange. Note that while<br />

the primary score for TPs remains largely unchanged, FP scores are significantly reduced. (d) ROC plot using dCn filtering during 1.0, 0.25, and<br />

0.025 m/z-wide bin searches. Insets show histograms of the behavior of dCn during 1.0 and 0.025 m/z-wide bin searches; the distributions are<br />

colored the same as in part c.<br />

DISCUSSION<br />

Soon, proteomics researchers will be swimming in a sea of<br />

mass spectra. During the later stages of development of Macro,<br />

we were asked to search a5hLTQVelos run with Macro and<br />

were surprised to find that it consisted of ∼170 000 tandem mass<br />

spectra. Clearly, the relative processing speed of computer-assisted<br />

spectral matching algorithms will be a critical feature for these<br />

algorithms in the coming years. Our intent with developing Macro<br />

was to refresh a scoring metric (Xcorr) that adds significant value<br />

to proteomics research by reorganizing the way it conducted<br />

searches. The fact that Macro is also very competitive with other<br />

search engines reinforces our view that candidate-centric searching<br />

represents a practical approach to search algorithm design in<br />

general. The core Macro process is also scalable: the MS/MS<br />

spectral array may be accessed in main CPU memory by multiple<br />

independent processing cores, each of which could digest a<br />

different portion of the FASTA database or perform scoring<br />

functions, or both; design modifications such as this are planned<br />

for future versions of Macro.<br />

Although the SEQUEST-specific version of Macro contains<br />

proprietary information that precludes its pubic release, we offer<br />

the source code for MacroParser to the proteomics community<br />

as a general framework for candidate-centric database parsing into<br />

which any peptide scorer could be inserted. Details regarding the<br />

Macro project can be found on our lab Web server at http://<br />

proteomics.dartmouth.edu.<br />

We feel that the performance gains associated with Macro can<br />

allow for a host of features and functions to be added to the<br />

classical scoring function Xcorr, including post-translational modi-<br />

6828 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

fication scanning and, as we show here, high-resolution correlation<br />

analysis. Although a parameter does exist in legacy SEQUEST<br />

that appears to allow users to scale fragment ion tolerances, closer<br />

inspection of the SEQUEST source code and search output when<br />

using this parameter reveals that it is not performing these<br />

searches as the user likely intends. Our inclusion of a variable<br />

fragment ion bin width parameter that accurately bins fragment<br />

ions into a scalable sparse array enables users with a mass<br />

spectrometer capable of better than unit resolution MS/MS<br />

spectra to search their data with the corresponding fragment ion<br />

mass tolerance. This has the added benefit of improved overall<br />

search accuracy relative to unit resolution searching. Although<br />

this does come as a speed penalty, we note that even when<br />

running at 0.025 m/z bin width, Macro is still 20× faster than<br />

SEQUEST on our platform. We anticipate that these highresolution<br />

correlation searches will result in a significant decrease<br />

in the minimum Xcorr threshold required to achieve a desired<br />

FDR: we observed the median Xcorr value of rescued PSMs,<br />

across all charge states, to be 1.52, with the lowest rescued PSM<br />

yielding an Xcorr of 0.67, albeit with a dCn of 0.08. Legacy<br />

SEQUEST users will likely need some adjusting to these unusually<br />

low score results, but with careful application of strategies to<br />

estimate data set false discovery rates 21 and calculate PSM<br />

posterior probability assignments, 26 we hope they will also<br />

ultimately appreciate the benefit of high mass precision to the<br />

improved accuracy of their data sets.<br />

(26) Kall, L.; Storey, J. D.; MacCoss, M. J.; Noble, W. S. J. Proteome Res. 2008,<br />

7, 29–34.


Figure 6. Characterization of PSMs “rescued” by high-resolution correlation analysis. (a) Top-ranked PSMs and scores for a given MS/MS<br />

search result using 1.0 and 0.025 m/z-wide bins. Note that these sequences are identical, except for a K/Q substitution in the middle of the<br />

peptide, marked by a red box. While the 1.0 m/z-wide bin search rank 1 PSM satisfies a precursor ion tolerance and Xcorr cutoff, the dCn<br />

(0.005) precludes unambiguous TP assignment. The top-ranked PSMs and scores from the 0.025 m/z-wide bin search are shown below. The<br />

resolving power of much narrower fragment ion bins clearly allows for discrimination of the K-containing ion fragments, resulting in a dCn of<br />

0.36 and an unambiguous TP PSM assignment. (b) Graphical representation of the distribution of candidate PSMs by Xcorr for the 1.0 (blue)<br />

and 0.025 (red) m/z-wide bin searches of the PSM from part a. (c) Histogram depicting the relative ranks of rescued PSMs in the lower resolution<br />

search. The lowest ranked PSM was rescued from the 4 715th position. (d) Reciprocal MS/MS plots, sequence, and matched fragment ions for<br />

the PSM rescued from the 4 715th rank position. The upper MS/MS spectrum is from the endogenous peptide from the HeLa cell lysate, and<br />

the lower MS/MS spectrum is from the synthetic peptide “AASVHTVGEDTEETPHR”.<br />

ACKNOWLEDGMENT<br />

We thank Mike Senko, Justin Blethrow, and colleagues at<br />

ThermoFisher Scientific for facilitating transfer of the SEQUST<br />

source code, additional platform testing, and general discussions.<br />

We thank Arminja Kettenbach for preparing the HeLa cell lysate<br />

digest and phosphopeptide-enriched sample and also thank Jason<br />

Gilmore for critical reading of the manuscript. This work was<br />

supported by the National Institutes of Health Grant P20-RR018787<br />

from the IDeA Program of the National Center for Research<br />

Resources (to S.A.G.) and a predoctoral fellowship from the<br />

National Institute of General Medical Sciences Grant T32-<br />

GM008704 (to B.K.F.).<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review March 26, 2010. Accepted July 8,<br />

2010.<br />

AC100783X<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6829


Anal. Chem. 2010, 82, 6830–6837<br />

Colorimetric Sensing of Silver(I) and Mercury(II)<br />

Ions Based on an Assembly of Tween 20-Stabilized<br />

Gold Nanoparticles<br />

Cheng-Yan Lin, † Cheng-Ju Yu, † Yen-Hsiu Lin, † and Wei-Lung Tseng* ,†,‡<br />

Department of <strong>Chemistry</strong>, National Sun Yat-sen University, Taiwan, and National Sun Yat-sen University-Kaohsiung<br />

Medical University Joint Research Center, Kaohsiung, Taiwan<br />

We have developed a rapid and homogeneous method for<br />

the highly selective detection of Hg 2+ and Ag + using<br />

Tween 20-modified gold nanoparticles (AuNPs). Citrate<br />

ions were found to still be adsorbed on the Au<br />

surface when citrate-capped AuNPs were modified with<br />

Tween 20, which stabilizes the citrate-capped AuNPs<br />

against conditions of high ionic strength. When citrate<br />

ions had reduced Hg 2+ and Ag + to form Hg-Au alloys<br />

and Ag on the surface of the AuNPs, Tween 20 was<br />

removed from the NP surface. As a result, the AuNPs<br />

were unstable under a high-ionic-strength solution,<br />

resulting in NP aggregation. The formation of Hg-Au<br />

alloys or Ag on the surface of the AuNPs was demonstrated<br />

by means of inductively coupled plasma mass<br />

spectroscopy and energy-dispersive X-ray spectroscopy.<br />

Tween 20-AuNPs could selectively detect Hg 2+ and<br />

Ag + at concentrations as low as 0.1 and 0.1 µM inthe<br />

presence of NaCl and EDTA, respectively. Moreover,<br />

the probe enables the analysis of AgNPs with a minimum<br />

detectable concentration that corresponds to 1<br />

pM. This probe was successfully applied to detect Hg 2+<br />

in drinking water and seawater, Ag + in drinking water,<br />

and AgNPs in drinking water.<br />

Interest in monitoring toxic metal ions in aquatic ecosystems<br />

continues because these contaminants adversely affect the environment<br />

and have serious medical effects. 1 Silver and mercury<br />

are two of the most hazardous metal pollutants, and they are<br />

widely distributed in ambient air, water, soil, and even food. 2,3<br />

For example, silver can inactivate sulfhydryl enzymes and accumulate<br />

in the body, 4 and mercury exposure can damage a<br />

variety of organs and the immune system. 5 Current approaches<br />

to detecting these two metal ions include inductively coupled<br />

* To whom correspondence should be addressed. Fax: 011-886-7-3684046.<br />

E-mail: tsengwl@mail.nsysu.edu.tw.<br />

† Department of <strong>Chemistry</strong>, National Sun Yat-sen University.<br />

‡ National Sun Yat-sen University-Kaohsiung Medical University Joint Research<br />

Center.<br />

(1) Campbell, L.; Dixon, D. G.; Hecky, R. E. J. Toxicol. Environ. Health, Part B<br />

2003, 6, 325–356.<br />

(2) Wood, C. M.; McDonald, M. D.; Walker, P.; Grosell, M.; Barimo, J. F.;<br />

Playle, R. C.; Walsh, P. J. Aquat. Toxicol. 2004, 70, 137–157.<br />

(3) Boening, D. W. Chemosphere. 2000, 40, 1335–1351.<br />

(4) Ratte, H. T. Environ. Toxicol. Chem. 1999, 18, 89–108.<br />

(5) Holmes, P.; James, K. A.; Levy, L. S. Sci. Total Environ. 2009, 408, 171–<br />

182.<br />

6830 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

plasma mass spectrometry (ICP-MS), 6,7 atomic absorption<br />

spectrometry, 8,9 and stripping voltammetry. 10,11 Although these<br />

methods offer excellent sensitivity and multielement analysis, they<br />

are rather costly, time-consuming, complex, and nonportable.<br />

In response to these shortcomings, various sensors using small<br />

organic molecules, 12,13 oligonucleotides, 14,15 DNAzymes, 16,17 and<br />

semiconductor quantum dots 18,19 have been investigated for<br />

the selective detection of Ag + or Hg 2+ in aqueous solutions.<br />

Unfortunately, most of these methods suffer from low water<br />

solubility, a complex synthesis procedure, and time-consuming<br />

DNA probe preparation. Recently, gold nanoparticles (AuNPs)<br />

have become another emerging material for sensing Hg 2+ or<br />

Ag + , because they have a high extinction coefficient in the<br />

visible region and behavior that depends on the interparticle<br />

distance. When the distances between the AuNPs become less<br />

than the average particle diameter, the color of the AuNPs<br />

changes from red to purple. Because of the coordination<br />

between the carboxyl groups of thiols and Hg 2+ , thiol-capped<br />

AuNPs have been used for colorimetric sensing of Hg 2+ . 20-22<br />

Also, Hg 2+ can selectively coordinate thymine (T) bases and<br />

forms stable T-Hg 2+ -T complexes. The melting temperature<br />

of cDNA containing T-Hg 2+ -T complexes is higher than that<br />

(6) Karunasagar, D.; Arunachalam, J.; Gangadharan, S. J. Anal. At. Spectrom.<br />

1998, 13, 679–682.<br />

(7) Barriada, J. L.; Tappin, A. D.; Evans, E. H.; Achterberg, E. P. TrAC, Trends<br />

Anal. Chem. 2007, 26, 809–817.<br />

(8) Li, Y.; Chen, C.; Li, B.; Sun, J.; Wang, J.; Gao, Y.; Zhao, Y.; Chai, Z. J. Anal.<br />

At. Spectrom. 2006, 21, 94–96.<br />

(9) Chamsaz, M.; Arbab-Zavar, M. H.; Akhondzadeh, J. Anal. Sci. 2008, 24,<br />

799–801.<br />

(10) Kim, H. J.; Park, D. S.; Hyun, M. H.; Shim, Y. B. Electroanalysis 1998, 10,<br />

303–306.<br />

(11) Mikelova, R.; Baloun, J.; Petrlova, J.; Adam, V.; Havel, L.; Petrek, J.; Horna,<br />

A.; Kizek, R. Bioelectrochemistry 2007, 70, 508–518.<br />

(12) Chatterjee, A.; Santra, M.; Won, N.; Kim, S.; Kim, J. K.; Kim, S. B.; Ahn,<br />

K. H. J. Am. Chem. Soc. 2009, 131, 2040–2041.<br />

(13) Zhan, X. Q.; Qian, Z. H.; Zheng, H.; Su, B. Y.; Lan, Z.; Xu, J. G. Chem.<br />

Commun. 2008, 1859–1861.<br />

(14) Lin, Y.-H.; Tseng, W.-L. Chem. Commun. 2009, 6619–6621.<br />

(15) Wang, J.; Liu, B. Chem. Commun. 2008, 4759–4761.<br />

(16) Li, T.; Shi, L.; Wang, E.; Dong, S. <strong>Chemistry</strong> 2009, 15, 3347–3350.<br />

(17) Hollenstein, M.; Hipolito, C.; Lam, C.; Dietrich, D.; Perrin, D. M. Angew.<br />

Chem., Int. Ed. 2008, 47, 4346–4350.<br />

(18) Koneswaran, M.; Narayanaswamy, R. Sens. Actuators, B 2009, 139, 91–<br />

96.<br />

(19) Chen, J.-L.; Zhu, C.-Q. Anal. Chim. Acta 2005, 546, 147–153.<br />

(20) Huang, C.-C.; Chang, H.-T. Anal. Chem. 2006, 78, 8332–8338.<br />

(21) Yu, C.-J.; Tseng, W.-L. Langmuir 2008, 24, 12717–12722.<br />

(22) Darbha, G. K.; Singh, A. K.; Rai, U. S.; Yu, E.; Yu, H.; Chandra Ray, P.<br />

J. Am. Chem. Soc. 2008, 130, 8038–8043.<br />

10.1021/ac1007909 © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/16/2010


containing T-T mismatches. 23 Based on this concept, two<br />

strands of DNA, which are designed to be complementary<br />

except for a single T-T mismatch, are used to modify the<br />

surface of the AuNPs. The resulting two types of DNAfunctionalized<br />

AuNPs are selectively aggregated in the presence<br />

of Hg 2+ based on T-Hg 2+ -T coordination and temperature<br />

control. 24 Similarly, Hg 2+ was selectively detected using two<br />

types of T-rich DNA-modified AuNPs and a T-rich DNA linker<br />

at room temperature. 25 Citrate-capped AuNPs interacting with<br />

single strands of T-rich oligonucleotide were found to be stable<br />

in a high-salt solution. 26-28 When Hg 2+ causes the conformation<br />

of T-rich DNA into a folded structure, this folded DNA cannot<br />

be adsorbed onto the AuNP surface. Salt-induced NP aggregation<br />

occurs because T-rich DNA is removed. In addition, it is<br />

well-known that a Au surface exhibits a strong affinity for<br />

Hg 2+ . 29-32 Thus, after the reduction of Hg 2+ with NaBH4, the<br />

Hg(0) thus generated is strongly bonded onto the surface of<br />

Au-based nanomaterials to form a solid amalgam-like structure.<br />

The surface plasmon resonance (SPR) band of Au nanorods<br />

and NPs in an excess of NaBH4 has been found to undergo a<br />

blue shift and a decrease in intensity after Hg 2+ was added. 29,31<br />

The only work on the detection of Ag + reported that AuNPs<br />

functionalized with cytosine-(C)-rich oligonucleotide selectively<br />

aggregated in the presence of Ag + based on the formation of<br />

C-Ag + -C complexes. 33 Although these methods all show good<br />

sensitivity and selectivity to Hg 2+ or Ag + , analysis of these two<br />

metal ions using a single type of AuNPs remains a challenge.<br />

In this study, we present a label-free, rapid, and homogeneous<br />

method for sensing both Hg 2+ and Ag + using Tween 20stabilized<br />

AuNPs (Tween 20-AuNPs). Because the surfaces of<br />

Tween 20-AuNPs still had citrate ions, the reduction of Hg 2+<br />

or Ag + with citrate resulted in the formation of Hg-Au alloy<br />

and Ag on the Au surface. When the Tween 20 was removed,<br />

NP aggregation occurred. We also investigated the effect of<br />

masking agents on the selectivity of this probe. To demonstrate<br />

its practicality, the present method was further applied to the<br />

determination of Hg 2+ ,Ag + , and AgNPs in complex matrices.<br />

EXPERIMENTAL SECTION<br />

<strong>Chemical</strong>s. Hydrogen tetrachloroaurate (III) dehydrate,<br />

Na2HPO4, and Na3PO4 were purchased from Alfa Aesar (Ward<br />

Hill, MA). Trisodium citrate, ethylenediaminetetraacetic acid<br />

(EDTA), Tween 20, Tween 40, Tween 60, Tween 80, NaBH4,<br />

(23) Tanaka, Y.; Oda, S.; Yamaguchi, H.; Kondo, Y.; Kojima, C.; Ono, A. J. Am.<br />

Chem. Soc. 2007, 129, 244–245.<br />

(24) Lee, J.-S.; Han, M.-S.; Mirkin, C. A. Angew. Chem., Int. Ed. 2007, 46, 4093–<br />

4096.<br />

(25) Xue, X.; Wang, F.; Liu, X. J. Am. Chem. Soc. 2008, 130, 3244–3245.<br />

(26) Yu, C.-J.; Cheng, T.-L.; Tseng, W.-L. Biosens. Bioelectron. 2009, 25, 204–<br />

210.<br />

(27) Li, D.; Wieckowska, A.; Willner, I. Angew. Chem., Int. Ed. 2008, 47, 3927–<br />

3931.<br />

(28) Liu, C.-W.; Hsieh, Y.-T.; Huang, C.-C.; Lin, Z.-H.; Chang, H.-T. Chem.<br />

Commun. 2008, 2242–2244.<br />

(29) Rex, M.; Hernandez, F. E.; Campiglia, A. D. Anal. Chem. 2006, 78, 445–<br />

451.<br />

(30) Leopold, K.; Foulkes, M.; Worsfold, P. J. Anal. Chem. 2009, 81, 3421–<br />

3428.<br />

(31) Lisha, K. P.; Anshup; Pradeep, T. Gold Bull. 2009, 42, 144–152.<br />

(32) Barrosse-Antle, L. E.; Xiao, L.; Wildgoose, G. G.; Baron, R.; Salter, C. J.;<br />

Crossley, A.; Compton, R. G. New J. Chem. 2007, 31, 2071–2075.<br />

(33) Li, B.; Du, Y.; Dong, S. Anal. Chim. Acta 2009, 644, 78–82.<br />

ascorbic acid, and NaCl were ordered from Sigma-Aldrich<br />

(Louis, MO). LiCl, KCl, MgCl2, CaCl2, SrCl2, BaCl2, CrCl3,<br />

MnCl2, FeCl2, FeCl3, CoCl2, NiCl2, CuCl2, ZnCl2, Cd(ClO4)2,<br />

AlCl3, Pb(NO3)2, HgCl2, and AgNO3 were purchased from Acros<br />

(Geel, Belgium). Water used in all experiments was doubly<br />

distilled and purified by a Milli-Q system (Millipore, Milford,<br />

MA).<br />

Characterization of the AuNPs. Extinction spectra of the<br />

AuNPs were measured using a double-beam UV-visible spectrophotometer<br />

(Cintra 10e; GBC, Victoria, Australia). High-resolution<br />

transmission electron microscopy (HRTEM, FEI Tecnai G2 F20<br />

S-Twin working at 200 kV) was used to collect HRTEM images<br />

of dispersed and aggregated AuNPs. Energy-dispersive X-ray<br />

(EDX) spectra were obtained using a HRTEM microscope. The<br />

zeta potential and size distribution of the AuNPs were measured<br />

using Delsa nano zeta potential and submicrometer particle size<br />

analyzer (Beckman Coulter Inc., U.S.). The hydrodynamic size<br />

of the AuNPs was measured using dynamic light scattering (DLS)<br />

(N5 Submicrometer Particle Size Analyzer, Beckman Coulter Inc.,<br />

U.S.).<br />

To understand the sensing mechanism, we equilibrated aliquots<br />

(1.0 mL) of 0.48 nM Tween 20-AuNPs in the presence of<br />

Hg 2+ (0-10 µM) or Ag + (0-10 µM) for 5 min at ambient<br />

temperature. The resulting mixture was subjected to centrifugation<br />

at 17 000 rpm for 10 min. Following removal of the<br />

supernatants, the precipitates were washed with water. After<br />

five centrifugation/washing cycles, the pellets were resuspended<br />

in water. A portion of the samples (∼200 µL) was<br />

diluted to 50-fold and then measured by ICP-MS (Perkin-Elmer-<br />

SCIEX, Thornhill, ON, Canada). Additionally, the composition<br />

of the obtained pellets was analyzed by EDX spectroscopy.<br />

For surface-assisted laser desorption/ionization time-of-flight<br />

ionization mass spectrometry (SALDI-TOF MS) (Autoflex, Bruker)<br />

measurements, citrate-capped AuNPs and Tween 20-AuNPs were<br />

separately pipetted into a stainless steel 384-well target (Bruker<br />

Daltonics) and dried under ambient temperature. Desorption/<br />

ionization was obtained by using a 337-nm-diameter nitrogen laser<br />

witha3nspulse width. MS experiments were performed in the<br />

positive-ion mode on a reflectron-type TOF MS equipped with a<br />

3 m flight tube. To obtain good resolution and signal-to-noise<br />

ratios, the laser power was adjusted to slightly above the threshold,<br />

and each mass spectrum was generated by averaging 500 laser<br />

pulses.<br />

Nanoparticle Synthesis. We prepared citrate-capped AuNPs<br />

by means of the chemical reduction of a metal salt precursor<br />

(hydrogen tetrachloroaurate, HAuCl4) in the liquid phase. To<br />

achieve this, we rapidly added HAuCl4 (0.35 M, 54 µL) to a<br />

solution of sodium citrate (2.55 mM, 60 mL) that was heated<br />

under reflux. This heating continued for an additional 15 min,<br />

during which time the color of the solution changed to a deep<br />

red. The size of citrate-capped AuNPs determined by TEM<br />

images was 13 ± 1 nm. The SPR wavelength of citrate-capped<br />

AuNPs located at 520 nm. The particle concentration of the<br />

AuNP solution was estimated to be 4.8 nM by Beer’s law; the<br />

extinction coefficient of 13 nm AuNPs at 520 nm is 2.7 × 10 8<br />

M -1 cm -1 . Tween 20-AuNPs were synthesized by adding Tween<br />

20 (10% v/v, 240 µL) to a solution of citrate-capped AuNPs (4.8<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6831


Scheme 1. Illustration of the Mechanism of Tween 20-AuNPs for Sensing Hg 2+ and Ag +<br />

nM, 60 mL). 34 To investigate the effect of the kind of the<br />

surfactant on the sensing of Hg 2+ and Ag + , Tween 20 was<br />

replaced by Tween 40, Tween 60, and Tween 80, once at a time.<br />

Additionally, to test the effect of the type of the AuNPs on the<br />

detection of Hg 2+ and Ag + , we replaced citrate-capped AuNPs<br />

with bare AuNPs in the synthesis of Tween 20-AuNPs. The<br />

preparation of bare AuNPs (3.7 nM; 12 nm) is described in<br />

the Supporting Information (SI).<br />

Sample Preparation. Tween 20-AuNPs were prepared in<br />

100-1000 mM sodium phosphate solution at pH 12.0. For Hg 2+<br />

sensing, metal ions (800 µL, 125-1250 nM) were added to a<br />

solution containing Tween 20-AuNPs (100 µL, 0.96-14.4 nM)<br />

and NaCl (100 µL, 1 M). For Ag + sensing, metal ions (800 µL,<br />

125-1250 nM) were added to a solution containing Tween 20-<br />

AuNPs (100 µL, 0.96-14.4 nM) and EDTA (100 µL, 0.1 M).<br />

We equilibrated the resulting solutions at ambient temperature<br />

for the optimum incubation time, and then recorded the<br />

extinction spectra of the solutions.<br />

Analysis of Real Samples. Samples of drinking water and<br />

seawater (pH 7.9) were collected from National Sun Yat-sen<br />

University campus. We then prepared a series of samples by<br />

“spiking” them with standard solutions of Hg 2+ (125-1250 nM)<br />

or Ag + (375-1250 nM). These spiked samples (800 µL) were<br />

added either to a solution containing 100 µL of 2.4 nM Tween<br />

20-AuNPs and 100 µL of 1 M NaCl or to a solution containing<br />

100 of 4.8 nM Tween 20-AuNPs and 100 µL of 0.1 M EDTA.<br />

We incubated the resulting solutions for 5 min before measuring<br />

their extinction spectra.<br />

On the other hand, this proposed method was utilized to detect<br />

10 nm AgNPs in drinking water. The preparation of citrate-capped<br />

AgNPs 35 is described in the SI. Different concentrations of AgNPs<br />

(1-10 pM) present in drinking water were oxidized to Ag + with<br />

(34) Huang, C.-C.; Tseng, W.-L. Analyst 2009, 134, 1699–1705.<br />

(35) Wei, H.; Chen, C.; Han, B.; Wang, E. Anal. Chem. 2008, 80, 7051–7055.<br />

6832 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

a solution of 1 µM H2O2 and 1 µM H3PO4. After 10 min, the<br />

resulting solution (800 µL) was added to a solution containing<br />

100 of 4.8 nM Tween 20-AuNPs and 100 µL of 0.1 M EDTA.<br />

The extinction spectra of the resulting solutions were recorded<br />

after 5 min incubation.<br />

RESULTS AND DISCUSSION<br />

Sensing Mechanism. SALDI-TOF MS technique is applied<br />

to the detection of small molecules that are adsorbed on the<br />

surface of the AuNPs when AuNPs are used as SALDI matrices. 36,37<br />

Thus, we utilized this technique to determine whether citrate ions<br />

were adsorbed on the surface of Tween 20-AuNPs. SI Figure S1A<br />

shows the SALDI spectrum of citrate-capped AuNPs. The peak<br />

detected at m/z 258.00 corresponded to [citrate +3Na] + . This<br />

peak was also observed in the SALDI spectrum of Tween 20-<br />

AuNPs (SI Figure S1B), indicating that citrate ions are capped<br />

on the surface of Tween 20-AuNPs. Moreover, the zeta potential<br />

of Tween 20-AuNPs was found to be -14.0 ± 0.8 mV in 10 mM<br />

phosphate at pH 12.0. These results suggest that a neutral Tween<br />

20 coating only shields the citrate ions (no displacement occurs).<br />

Accordingly, we reasoned that citrate ions adsorbed onto the NP<br />

surface can act as a reducing agent for Hg 2+ and Ag + when<br />

citrate-capped AuNPs have been modified with Tween 20.<br />

Scheme 1 shows the mechanism by which Tween 20-AuNPs<br />

sense Hg 2+ and Ag + . Because of the high affinity between Au<br />

and Hg, 29-32 the reduced Hg(0) is directly deposited onto the<br />

Au surface through the formation of Hg-Au alloys; meanwhile,<br />

Tween 20 molecules are desorbed from the AuNPs. The removal<br />

of the stabilizer (Tween 20) causes the AuNPs to aggregate under<br />

conditions of high ionic strength. Also, citrate ions can be used<br />

(36) Su, C.-L.; Tseng, W.-L. Anal. Chem. 2007, 79, 1626–1633.<br />

(37) Wu, H.-P.; Yu, C.-J.; Lin, C.-Y.; Lin, Y.-H.; Tseng, W.-L. J. Am. Soc. Mass<br />

Spectrom. 2009, 20, 875–882.


Figure 1. Extinction spectra of solutions of (A) 0.48 nM Tween 20-<br />

AuNPs and (B) 0.37 nM Tween 20-modified bare AuNPs (a) before<br />

and (b, c) after the addition of (b) 1 µMHg 2+ and (c) 1 µMAg + . Tween<br />

20-AuNPs are prepared in 20 mM phosphate at pH 12.0. The<br />

incubation time is 5 min.<br />

to reduce Ag + onto the Au surface. 38,39 The formation of Agcoated<br />

AuNPs enables Tween 20 to be removed from the NP<br />

surface, thereby inducing aggregation of the AuNPs in a highionic-strength<br />

solution.<br />

Evidence for the Formation of Hg-Au Alloys and Ag<br />

Shells. To test the hypothesis mentioned above, we monitored<br />

the extinction spectra of Tween 20-AuNPs in the absence and<br />

presence of Hg 2+ and Ag + . Curve a in Figure 1A shows that the<br />

SPR wavelength of Tween 20-AuNPs appears at 520 nm, indicating<br />

that they are well dispersed in 20 mM phosphate solution at pH<br />

12.0. In other words, Tween 20 molecules can indeed protect<br />

citrate-capped AuNPs against a high-ionic-strength solution. 40 After<br />

separately adding 1.0 µMHg 2+ (curve b) and 1.0 µMAg + (curve<br />

c), we observed a decrease in the strength of SPR band at 520<br />

nm and the formation of a new red-shift band. These changes<br />

were characteristic of AuNP aggregation. The Hg 2+ - and Ag + -<br />

(38) Xie, W.; Su, L.; Donfack, P.; Shen, A.; Zhou, X.; Sackmann, M.; Materny,<br />

A.; Hu, J. Chem. Commun. 2009, 5263–5265.<br />

(39) Xia, H.; Bai, S.; Hartmann, J.; Wang, D. Langmuir 2009.<br />

(40) Shen, C.-C.; Tseng, W.-L.; Hsieh, M.-M. J. Chromatogr., A 2009, 1216,<br />

288–293.<br />

induced NP aggregation was nearly complete after 5 min (SI<br />

Figure S2). Obviously, the deposition of Hg and Ag onto the Au<br />

surface enables Tween 20 to be removed, thereby driving NP<br />

aggregation. We ruled out the possibility that coordination<br />

between Tween 20 and metal ions induces the NP aggregation<br />

because Tween 20 does not contain any functional group to<br />

interact with Hg 2+ and Ag + . To ensure the role of citrate ions<br />

in the reduction of Hg 2+ and Ag + , the extinction spectra of bare<br />

AuNPs modified with Tween 20 were examined under the same<br />

conditions. The addition of Hg 2+ and Ag + to this type of AuNP<br />

resulted in a rare shift in the SPR wavelength (Figure 1B),<br />

clearly indicating that citrate ions are indispensable for detecting<br />

Hg 2+ and Ag + using Tween 20-AuNPs. On the other hand, we<br />

prepared Hg-Au alloy- and Ag-coated AuNPs and modified<br />

them with Tween 20. Compared to Tween 20-AuNPs, a blue<br />

shift in SPR and a decrease in SPR intensity were observed<br />

for Tween 20-modified Hg-Au alloy-coated AuNPs (SI Figure<br />

S3). This is attributed to the formation of Hg-Au alloy on the<br />

surface of the AuNPs. A similar phenomenon was reported when<br />

citrate-capped AuNPs were exposed in the presence of Hg vapor. 41<br />

Moreover, this kind of NPs was found to be unstable in a highionic-strength<br />

solution. This result reflects that Tween 20 molecules<br />

were not attached to the surface of Hg-Au alloy, thereby<br />

incapable of protecting NPs against a high-ionic strength solution.<br />

SI Figure S4 shows that the deposition of Ag on the surface of<br />

the AuNPs resulted in an increase SPR intensity relative to Tween<br />

20-AuNPs. 42,43 Similarly, Tween 20-modified Ag-coated AuNPs<br />

were aggregated in a high-ionic-strength solution because Tween<br />

20 molecules were not adsorbed on the surface of Ag shell.<br />

To provide further evidence for the formation of Hg-Au alloys<br />

and Ag shells, we used ICP-MS to quantitatively determine the<br />

composition of the precipitates, which were obtained by five cycles<br />

of centrifugation of a solution of Tween 20-AuNPs and metal ions.<br />

When a series of concentrations (0-1 µM) of Hg 2+ were present<br />

in a solution of 0.48 nM Tween 20-AuNPs, the molar ratio of<br />

Hg to Au in the precipitates gradually increased with increasing<br />

Hg 2+ concentration (Figure 2A). A similar phenomenon was seen<br />

in the case of Ag + (Figure 2B). If Hg-Au alloys and Ag shells<br />

did not form on the Au surface, the molar ratios of Hg to Au and<br />

Ag to Au should remain constant under these conditions. Under<br />

identical treatment conditions (five centrifugation/washing cycles),<br />

the composition of precipitates was also determined by EDX<br />

analysis. The Hg content in the precipitates increased with<br />

increasing Hg 2+ concentration (Figure 2C). We observed a<br />

similar phenomenon when different concentrations of Ag + were<br />

present in a solution of Tween 20-AuNPs (Figure 2D). These<br />

results are in agreement with those obtained by ICP-MS. These<br />

findings strongly support the idea that Hg 2+ - and Ag + -induced<br />

aggregation of Tween 20-AuNPs is indeed the result of the<br />

formation of Hg-Au alloys and Ag on the Au surface.<br />

Effect of Surfactant Chain Length, NP Concentration, and<br />

Ionic Strength. We next explored the effect of surfactant chain<br />

length on the metal-ion-induced aggregation of the AuNPs. The<br />

(41) Morris, T.; Copeland, H.; McLinden, E.; Wilson, S.; Szulczewski, G.<br />

Langmuir 2002, 18, 7261–7264.<br />

(42) Anandan, S.; Grieser, F.; Ashokkumar, M. J. Phys. Chem. C 2008, 112,<br />

15102–15105.<br />

(43) Guerrini, L.; Garcia-Ramos, J. V.; Domingo, C.; Sanchez-Cortes, S. J. Raman<br />

Spectrosc. 2010, 41, 508–515.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6833


Figure 2. (A, B) Effect of the concentration of (A) Hg 2+ and (B) Ag + on the composition ratio of (A) Au to Ag and (B) Au to Hg of the precipitates.<br />

A series of concentration of (A) 0-1000 nM Hg 2+ and (B) 0-1000 nM Ag + was added to 1.0 mL of 0.48 nM Tween 20-AuNPs. The precipitates<br />

were obtained by five cycles of centrifugation of the resulting solutions. (C, D) EDX spectra of the precipitates obtained after the addition of (a)<br />

0.1, (b) 1, and (c) 10 µM (C) Hg 2+ and (D) Ag + to a solution of 0.48 nM Tween 20-AuNPs. Tween 20-AuNPs are prepared in 20 mM phosphate<br />

at pH 12.0. The incubation time is 5 min.<br />

extinction values of the solution at 650 and 520 nm corresponded<br />

to the quantities of dispersed and aggregated AuNPs, respectively.<br />

Thus, the molar ratio of dispersed to aggregated AuNPs can be<br />

expressed by the ratio of the extinction value Ex at 650 nm to<br />

that at 520 nm (Ex650 nm/Ex520 nm). As shown in Figure 3A, the<br />

addition of both Hg2+ and Ag + to a solution of Tween 20-AuNPs<br />

resulted in a high value of Ex650 nm/Ex520 nm. However, when<br />

we replaced Tween 20 with Tween 40, the value of Ex650 nm/<br />

Ex520 nm became small. This suggests a relatively small amount<br />

of Hg-Au alloys or Ag on the surface of Tween 40-modified<br />

AuNPs, resulting in a small degree of NP aggregation. Similar<br />

phenomena were observed in the case of Tween 60- and Tween<br />

80-modifed AuNPs. To further confirm our hypothesis, ICP-<br />

MS was used to determine the composition of the NPs. After<br />

adding 1 µM Hg2+ to different kinds of AuNPs, the concentrations<br />

of Hg in Tween 20-, 40-, 60-, and 80-modified AuNPs were<br />

148, 64, 58, and 51 ppb, respectively. Similarly, upon the<br />

addition of 1 µM Ag + , the concentrations of Ag in Tween 20-,<br />

40-, 60-, and 80-modified AuNPs were 65, 19, 24, and 25 ppb,<br />

respectively. On the basis of these results, Tween 20 was<br />

selected for the following studies.<br />

Previous studies have shown that the sensitivity of a AuNPbased<br />

sensor is highly dependent on the concentration of<br />

6834 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

AuNPs. 44 At relatively high concentrations of Tween 20-AuNPs,<br />

the deposition of Hg2+ or Ag + on the surface of a single particle<br />

decreased, reducing the degree of NP aggregation. Moreover,<br />

the aggregation rate of NPs increased with increasing NP<br />

concentration. It can be seen that the optimum concentrations<br />

of Tween 20-AuNPs for sensing Hg2+ and Ag + were 0.24 and<br />

0.48 nM, respectively, when the incubation time was fixed at 5<br />

min (Figure 3B). Moreover, we investigated the effect of Na3PO4<br />

concentration on the colorimetric sensitivity of Tween 20-<br />

AuNPs to Hg2+ and Ag + .SI Figure S5 shows that the zeta<br />

potential of Tween 20-AuNPs reduced with increasing Na3PO4<br />

concentration, implying that electrostatic repulsion between<br />

Tween 20-AuNPs decreased with increasing ionic strength of<br />

the solution. Thus, under conditions of high ionic strength, the<br />

slight electrostatic repulsion between Tween 20-AuNPs provided<br />

a low barrier for metal-ion-induced NP aggregation. As<br />

expected, the difference in Ex650 nm/Ex520 nm for the cases with<br />

and without 1.0 µM Hg2+ gradually increased with increasing<br />

Na3PO4 concentration and reached a plateau at 80 mM Na3PO4<br />

(Figure 3C). A similar effect was found in the analysis of Ag +<br />

(Figure 3D). Consequently, 80 mM Na3PO4 was chosen for the<br />

following studies.<br />

(44) Huang, C.-C.; Tseng, W.-L. Anal. Chem. 2008, 80, 6345–6350.


Figure 3. (A) The value of Ex650 nm/Ex520 nm of Tween 20- 40-, 60-, and 80-modified AuNPs (0.48 nM) after the addition of (a) 1 µM Hg 2+ and<br />

(b) 1 µM Ag + . (B) Effect of the concentration of Tween 20-AuNPs on the ratio Ex650 nm/Ex520 nm in the presence of (a) 1 µM Hg 2+ and (b) 1 µM<br />

Ag + . (C) Effect of phosphate concentration on the ratio Ex650 nm/Ex520 nm of 0.24 nM Tween 20-AuNPs in the (a) absence and (b) presence of 1<br />

µM Hg 2+ . (D) Effect of phosphate concentration on the ratio Ex650 nm/Ex520 nm of 0.48 nM Tween 20-AuNPs in the (a) absence and (b) presence<br />

of 1 µM Ag + . (A, B) Tween 20-AuNPs are prepared in 20 mM phosphate at pH 12.0. (A-D) The incubation time is 5 min.<br />

Selectivity, Sensitivity, and Application. To realize the<br />

selectivity of 0.24 and 0.48 nM Tween 20-AuNPs toward Hg 2+ and<br />

Ag + , respectively, other metal ionssincluding Li + ,Na + ,K + ,<br />

Mg 2+ ,Ca 2+ ,Sr 2+ ,Ba 2+ ,Cr 3+ ,Mn 2+ ,Fe 2+ ,Fe 3+ ,Co 2+ ,Ni 2+ ,Cu 2+ ,<br />

Zn 2+ ,Cd 2+ ,Al 3+ ,Pb 2+ ,Hg 2+ , and Ag + swere examined under<br />

identical conditions. Only Hg 2+ and Ag + caused the aggregation<br />

of both 0.24 and 0.48 nM Tween 20-AuNPs (SI Figure S6),<br />

revealing that this probe is selective for Hg 2+ and Ag + . To<br />

circumvent this problem, we tested the effect of masking<br />

agents, including NaCl and EDTA. It is well-known that the<br />

solubility product of AgCl is 1.8 × 10 -10 , whereas HgCl2 is<br />

soluble in water (70 g/L). Accordingly, we tested the ability of<br />

NaCl to mask the interfering metal ions in our sensing system.<br />

In the presence of 0.1 M NaCl, the addition of 1 µM Hg 2+ to a<br />

solution of 0.24 nM Tween 20-AuNPs resulted in an apparent<br />

change in Ex650 nm/Ex520 nm, whereas the remaining metal ions<br />

(100 µM) had negligible effects on the same system (Figure<br />

4A). The selectivity of this probe is more than 100-fold for Hg 2+<br />

over all other tested metal ions. Additionally, to improve the<br />

selectivity of 0.48 nM Tween 20-AuNPs for Ag + , we chose<br />

EDTA as a masking agent, since it forms a more stable complex<br />

with Hg 2+ than with Ag + . As expected, the presence of 0.01 M<br />

EDTA masked Tween 20-AuNPs toward Hg 2+ (Figure 4B). As<br />

a result, Tween 20-AuNPs provided high selectivity (>100-fold)<br />

toward Ag + over all other tested metal ions.<br />

Under optimum conditions (for sensing Hg 2+ : 0.24 nM<br />

Tween 20-AuNPs, 80 mM Na3PO4, and 0.1 M NaCl; for<br />

sensing Ag + : 0.48 nM Tween 20-AuNPs, 80 mM Na3PO4, and<br />

0.01 M EDTA), we evaluated the sensitivity of this probe<br />

toward Hg 2+ and Ag + . When the concentrations of Hg 2+<br />

varied from 0 to 1000 nM, the extinction spectra of 0.24 nM<br />

Tween 20-AuNPs showed a gradual increase in extinction<br />

at 650 nm and their color gradually changed from red to<br />

purple (Figure 5A). A solution of 0.48 nM Tween 20-AuNPs<br />

showed a similar response to Ag + (Figure 5B). These spectra<br />

showed clear isosbestic points at 556 and 549 nm upon the<br />

addition of Hg 2+ and Ag + , respectively. This observation<br />

reveals that the aggregation of Tween 20-AuNPs is directly<br />

related to the concentration of Hg 2+ or Ag + . This result was<br />

further confirmed by HRTEM images and DLS measure-<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6835


Figure 4. The value of Ex650 nm/Ex520 nm of a solution of 80 mM Na3PO4 containing (A) 0.24 nM Tween 20-AuNPs and 0.1 M NaCl and (B) 0.48<br />

nM Tween 20-AuNPs and 0.01 M EDTA upon the addition of (A) 1 µM Hg 2+ and 100 µM other metal ions and (B) 1 µM Ag + and 100 µM other<br />

metal ions. The incubation time is 5 min.<br />

ments. SI Figure S7 shows the concentration-dependent TEM<br />

images of the aggregated AuNPs after adding different concentrations<br />

(0, 0.1, 1, and 10 µM) of Hg 2+ or Ag + to a solution<br />

of Tween 20-AuNPs. Moreover, the hydrodynamic size of<br />

the aggregated AuNPs increased with an increase in the<br />

concentration of Hg 2+ or Ag + (SI Figure S8). These results<br />

provide clear evidence that the aggregation degree of Tween<br />

20-AuNPs is highly dependent on the concentration of Hg 2+<br />

or Ag + . We observed that the ratio Ex650 nm/Ex520 nm<br />

increased linearly with increasing Hg 2+ and Ag + concentration<br />

over the range of 200 to 800 nM (R 2 ) 0.9943) and 400<br />

to 1000 nM (R 2 ) 0.9935), respectively (Figure 5C and D).<br />

A difference in linearity between two metal ions could be due<br />

to that the sensing mechanism of Tween 20-AuNPs for Hg 2+<br />

is different from that for Ag + . This probe could detect Hg 2+<br />

and Ag + at concentrations as low as 100 and 100 nM,<br />

respectively. The result is useful for detecting Ag + in<br />

drinking water, because the maximum level of silver in<br />

drinking water permitted by the United States Environmental<br />

Protection Agency (EPA) is 50 µg/L (∼460 nM).<br />

To test the practicality of the present approach, a solution of<br />

0.24 nM Tween 20-AuNPs was used to analyze Hg 2+ in drinking<br />

water and seawater. As shown in SI Figures S9 and S10,<br />

Ex650 nm/Ex520 nm increased linearly upon increasing the spiked<br />

concentration of Hg 2+ in drinking water over the range of<br />

200-600 nM (R 2 ) 0.9944) and in seawater over the range of<br />

300-1000 nM (R 2 ) 0.9977). Evidence of the Hg 2+ -induced<br />

aggregation of Tween 20-AuNPs in seawater can be seen in<br />

the HRTEM images (SI Figure S11). The lowest detectable<br />

concentrations of Hg 2+ in drinking water and seawater were<br />

6836 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

estimated to be 200 and 100 nM, respectively. Although the<br />

sensitivity of this probe is insufficient to detect the maximum<br />

level of mercury (2 ppb) in drinking water permitted by the<br />

U.S. EPA, we suggest that Tween 20-AuNPs can be used as<br />

probes for solid-phase preconcentration of mercury in complex<br />

matrices prior to ICP-MS analysis. 30<br />

We also evaluated the feasibility of this approach for sensing<br />

of Ag + and AgNPs in drinking water. We obtained a linear<br />

correlation (R 2 ) 0.9963) between the ratio Ex650 nm/Ex520 nm<br />

and the concentration of Ag + spiked into the drinking water<br />

over the range of 400-1000 nM (SI Figure S12), which includes<br />

the maximum permissible limit of silver in drinking water.<br />

Moreover, since hazardous AgNPs may pose threats to human<br />

health or the environment, 45 this approach was further used to<br />

monitor AgNPs in drinking water. Under acidic conditions (1.0<br />

µM H3PO4), AgNPs (10 ± 2 nm) are oxidized to Ag + ions with<br />

1.0 mM H2O2. 12 The oxidation of AgNPs to Ag + was complete<br />

after 10 min. The generated Ag + was directly detected by 0.48<br />

nM Tween 20-AuNPs. The degree of aggregation of Tween 20-<br />

AuNPs increased when the concentration of AgNPs was<br />

increased from 1 to 10 pM (SI Figure S13). The correlation<br />

coefficient (R 2 ) for the determination of AgNPs in the range<br />

1-6 pM was 0.9988. These results suggest that this probe will<br />

be suitable for routine assays of AgNPs in consumer products<br />

such as cosmetics and fabrics. 46 Table 1 shows the quantitative<br />

measurements of Hg 2+ ,Ag + , and AgNPs in different matrices<br />

based on the use of Tween 20-AuNPs.<br />

(45) Lubick, N. Environ. Sci. Technol. 2008, 42, 8617.<br />

(46) Benn, T. M.; Westerhoff, P. Environ. Sci. Technol. 2008, 42, 4133–4139.


Figure 5. Extinction spectra and color changes of solutions containing (A) 0.24 nM Tween 20-AuNPs and 0.1 M NaCl and (B) 0.48 nM Tween<br />

20-AuNPs and 0.01 M EDTA upon the addition of (A) 0-1000 nM Hg 2+ and (B) 0-1000 nM Ag + . The arrows indicate the signal changes with<br />

increases in analyte concentrations (A: 0, 100, 200, 300, 400, 500, 600, 700, 800, and 1000 nM; B: 0, 100, 200, 300, 400, 500, 600, 700, 800,<br />

900, and 1000 nM). A plot of Ex650 nm/Ex520 nm versus the concentration of (C) Hg 2+ and (D) Ag + . The incubation time is 5 min. The error bars<br />

represent standard deviations based on three independent measurements.<br />

Table 1. Quantification of Hg 2+ ,Ag + and AgNPs in the<br />

Different Sample Matrix Based on the Use of Tween<br />

20-AuNPs<br />

analyte matrix linear rang (M) R 2 MDC (M) a<br />

Hg 2+ deionized water 2 × 10 -7 to 8 × 10 -7 0.9943 1 × 10 -7<br />

Hg 2+ drinking water 2 × 10 -7 to 6 × 10 -7 0.9944 2 × 10 -7<br />

Hg 2+ seawater 3 × 10 -7 to 1 × 10 -6 0.9977 1 × 10 -7<br />

Ag + deionized water 4 × 10 -7 to 1 × 10 -6 0.9935 1 × 10 -7<br />

Ag + drinking water 4 × 10 -7 to 1 × 10 -6 0.9963 3 × 10 -7<br />

AgNPs drinking water 6 × 10 -12 to 1 × 10 -11 0.9988 1 × 10 -12<br />

a MDC, minimum detectable concentration.<br />

CONCLUSIONS<br />

This study reports a new assay for the selective detection of<br />

Hg 2+ and Ag + using Tween 20-AuNPs. This probe can be<br />

further applied to detecting hazardous AgNPs. We demonstrated<br />

that the aggregation of Tween 20-AuNPs results from<br />

the formation of Hg-Au alloy or Ag on the surface of the<br />

AuNPs. Thus, in the opinion of the authors, Tween 20-AuNPs<br />

should be used as a selective probe for extracting a large<br />

volume of Hg 2+ and Ag + prior to ICP-MS analysis. Moreover,<br />

we believe that the present approach holds great potential for<br />

monitoring Ag + and AgNPs in environmental samples.<br />

ACKNOWLEDGMENT<br />

We thank National Science Council (NSC 98-2113-M-110-009-<br />

MY3) and National Sun Yat-sen University-Kaohsiung Medical<br />

University Joint Research Center for the financial support of this<br />

study.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Experimental details, additional references, and Figures S1-13.<br />

This material is available free of charge via the Internet at<br />

http://pubs.acs.org.<br />

Received for review March 27, 2010. Accepted June 25,<br />

2010.<br />

AC1007909<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6837


Anal. Chem. 2010, 82, 6838–6846<br />

Analysis of Inorganic Polyphosphates by Capillary<br />

Gel Electrophoresis<br />

Andrew Lee and George M. Whitesides*<br />

Department of <strong>Chemistry</strong> and <strong>Chemical</strong> Biology, Harvard University, 12 Oxford Street, Cambridge, MA<br />

This paper describes the development of a method that<br />

uses capillary gel electrophoresis (CGE) to analyze mixtures<br />

of inorganic polyphosphate ((Pi)n). Resolution of<br />

(Pi)n on the basis of n, the number of residues of<br />

dehydrated phosphate, is accomplished by CGE using<br />

capillaries filled with solutions of poly(N,N-dimethylacrylamide)<br />

(PDMA) and indirect detection by the UV<br />

absorbance of a chromophore, terephthalate, added to<br />

the running buffer. The method is capable of resolving<br />

peaks representing (P i)n with n up to ∼70; preparation<br />

and use of authentic standards enables the identification<br />

of peaks for (Pi)n with n ) 1-10. The main<br />

advantages of this method over previously reported<br />

methods for analyzing mixtures of (Pi)n (e.g., gel<br />

electrophoresis, CGE using polyacrylamide-filled capillaries)<br />

are its resolution, convenience, and reproducibility;<br />

gel-filled capillaries are easily regenerated by<br />

pumping in fresh, low-viscosity solutions of PDMA. The<br />

resolution is comparable to that of ion-exchange chromatography<br />

and detection of (P i)n by suppressed<br />

conductivity. The method is useful for analyzing (Pi)n<br />

generated by the dehydration of Pi at low temperature<br />

(125-140 °C) with urea, in a reaction that may have<br />

been important in prebiotic chemistry. The method<br />

should also be useful for characterizing mixtures of<br />

other anionic, oligomeric, or polymeric species without<br />

an intrinsic chromophore (e.g., sulfated polysaccharides,<br />

oligomeric phospho-diesters).<br />

This paper describes a technique for analyzing mixtures of<br />

inorganic oligo- and polyphosphates, (Pi)n, by capillary gel<br />

electrophoresis. Samples of condensed inorganic phosphate are<br />

typically mixtures of (Pi)n with different chain length, n.<br />

Mixtures of (Pi)n can have a wide range in n; (Pi)n with<br />

estimated values of n as high as 1000 have been reported. 1<br />

The analytical method developed here characterizes samples<br />

of (Pi)n at high resolution by separating and detecting each<br />

species in a mixture. Our primary motivation to develop a new<br />

method of analysis was to explore the synthesis and reactivity<br />

of (Pi)n relevant to the chemical origins of life (i.e., the prebiotic<br />

chemistry leading to self-replicating systems in a “pre-RNA”<br />

or “RNA world”). 2-7<br />

* To whom correspondence should be addressed. E-mail: gwhitesides@<br />

gmwgroup.harvard.edu.<br />

(1) Clark, J. E.; Wood, H. G. Anal. Biochem. 1987, 161, 280–290.<br />

(2) Joyce, G. F. Nature 1989, 338, 217–224.<br />

6838 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Species of (Pi)n in aqueous solution are anionic and differ<br />

from each other in the number of residues of condensed<br />

phosphate and in net negative electrostatic charge. Capillary<br />

electrophoresis (CE), in its most straightforward mode of<br />

operation (capillary zone electrophoresis (CZE), that is electrophoresis<br />

of analytes through free solution, combined with<br />

optical detection of chromophoric analytes), cannot resolve and<br />

detect (Pi)n. We developed a method that addresses the two<br />

problems of (i) separating (Pi)n in mixtures and (ii) detecting<br />

and quantifying each (Pi)n. Capillary gel electrophoresis (CGE),<br />

using capillaries filled with aqueous solutions of poly(N,Ndimethylacrylamide)<br />

(PDMA), resolved cyclic and linear (Pi)n<br />

in order of their electrophoretic mobility. Addition of the<br />

chromophoric anion, terephthalate (1,4-(CO2 - )2C6H4, abbreviated<br />

as TP 2- ), to the running buffer enabled the detection of<br />

separated (Pi)n by indirect UV absorbance.<br />

We demonstrated the resolution of mixtures of (Pi)n (with n<br />

up to ∼70) and the identification of components in mixtures<br />

with the use of authentic standards. The areas of peaks in<br />

electropherograms, determined by indirect detection, allowed<br />

us to quantify the relative concentration of each species in a<br />

mixture of (Pi)n. In addition to analyzing commercially available<br />

samples of (Pi)n, prepared by the thermal dehydration of Pi<br />

(>220 °C), we analyzed (Pi)n generated by dehydration reactions<br />

that might have occurred on the prebiotic earth and, thus,<br />

might have been involved in the chemical origins of life. 5,6,8<br />

MOTIVATION<br />

(Pi)n and adenosine triphosphate (ATP) have an essential<br />

functional group in common: residues of dehydrated phosphate<br />

connected by phosphoanhydride bonds. (Pi)n is simpler in<br />

composition than ATP but, in principle, may provide the same<br />

chemical function (e.g., activation of -OH groups) and has led<br />

to the suggestion that (Pi)n is a molecular fossil, a species<br />

important in the origin of life, and a precursor to ATP. 9-13 We<br />

(3) Joyce, G. F. Nature 2002, 418, 214–221.<br />

(4) Gilbert, W. Nature 1986, 319, 618–618.<br />

(5) Keefe, A. D.; Miller, S. L. Origins Life Evol. Biospheres 1996, 26, 15–25.<br />

(6) Keefe, A. D.; Miller, S. L. J. Mol. Evol. 1995, 41, 693–702.<br />

(7) Pasek, M. A.; Kee, T. P.; Bryant, D. E.; Pavlov, A. A.; Lunine, J. I. Angew.<br />

Chem., Int. Ed. Engl. 2008, 47, 7918–7920.<br />

(8) Osterberg, R.; Orgel, L. E.; Lohrmann, R. J. Mol. Evol. 1973, 2, 231–234.<br />

(9) Rao, N. N.; Gomez-Garcia, M. R.; Kornberg, A. Annu. Rev. Biochem. 2009,<br />

78, 605–647.<br />

(10) Kornberg, A. J. Bacteriol. 1995, 177, 491–496.<br />

(11) Kornberg, A.; Rao, N. N.; Ault-Riche, D. Annu. Rev. Biochem. 1999, 68,<br />

89–125.<br />

(12) Orgel, L. E.; Lohrmann, R. Acc. Chem. Res. 1974, 7, 368–377.<br />

(13) Lohrmann, R.; Orgel, L. E. Nature 1973, 244, 418–420.<br />

10.1021/ac1008018 © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/27/2010


wished to explore the dehydration of Pi in detail and needed a<br />

method that was more convenient and reproducible than those<br />

reported so far for the resolution of mixtures of (Pi)n varying<br />

in chain length. The method we have developed should also<br />

be useful for investigating the biochemistry of (Pi)n 9-11 and for<br />

developing (Pi)n as a reagent in chemical synthesis. In addition,<br />

the method may be useful in applications for quality control:<br />

(Pi)n is a component of many commercial materials (e.g.,<br />

fertilizers, food products, detergent formulations, building<br />

materials). 14<br />

PREVIOUSLY REPORTED METHODS FOR<br />

SEPARATING MIXTURES OF (PI)N<br />

Polyacrylamide gel electrophoresis (PAGE) can resolve mixtures<br />

of (Pi)n, with n ∼ 2-450. Analysis by PAGE involves gel<br />

electrophoresis (typical runs require g3 h) and the detection<br />

of (Pi)n by staining gels with the cationic dye toluidine blue O<br />

(TBO); 1,15,16 autoradiography can detect species of (Pi)n synthesized<br />

from 32 P-ATP. 17 The disadvantages to PAGE are the<br />

difficulty of the experiments and the time required for analysis.<br />

We found it difficult to cast 20% gels that are homogeneous<br />

and provide reproducible resolution; to run a gel requires<br />

several hours (for separation, staining, and destaining).<br />

Anion-exchange chromatography is useful both for analyzing<br />

mixtures of (Pi)n and for preparing samples of purified (Pi)n.<br />

The best results for the chromatographic resolution of (Pi)n<br />

are chromatograms showing up to ∼50 peaks, in runs of less<br />

than 30 min; 18,19 this method requires HPLC instrumentation<br />

with a suppressed conductivity detector and an online KOH<br />

gradient generator (for minimizing the amount of CO2/CO3 2adsorbed<br />

from the atmosphere). Distinguishing samples containing<br />

(Pi)n with n > 45 is difficult by this method, and the<br />

resolution of species with n ∼ 100 has not been demonstrated.<br />

Other qualitative or semiquantitative methods used to analyze<br />

mixtures of (Pi)n include paper chromatography, 20,21 31 P<br />

NMR, 22-27 ESI-MS, 28 and the analysis of terminal phosphate<br />

groups with phosphoglucokinase. 20<br />

(14) Phosphoric Acid and Phosphates. Kirk-Othmer Encyclopedia of <strong>Chemical</strong><br />

Technology, 4th ed.; John Wiley & Sons: New York, 1991; Vol. 18, pp 669-<br />

718.<br />

(15) Ogawa, N.; DeRisi, J.; Brown, P. O. Mol. Biol. Cell 2000, 11, 4309–4321.<br />

(16) Robinson, N. A.; Wood, H. G. J. Biol. Chem. 1986, 261, 4481–4485.<br />

(17) Gomez-Garcia, M. R.; Kornberg, A. Proc. Natl. Acad. Sci. U.S.A. 2004,<br />

101, 15876–15880.<br />

(18) Baluyot, E. S.; Hartford, C. G. J. Chromatogr., A 1996, 739, 217–222.<br />

(19) Sekiguchi, Y.; Matsunaga, A.; Yamamoto, A.; Inoue, Y. J. Chromatogr., A<br />

2000, 881, 639–644.<br />

(20) Greenfield, S.; Clift, M. <strong>Analytical</strong> <strong>Chemistry</strong> of the Condensed Phosphates; 1<br />

ed.; Pergamon Press: New York, 1975.<br />

(21) Osterberg, R.; Orgel, L. E. J. Mol. Evol. 1972, 1, 241–250.<br />

(22) Rao, N. N.; Roberts, M. F.; Torriani, A. J. Bacteriol. 1985, 162, 242–247.<br />

(23) Gard, D. R.; Burquin, J. C.; Gard, J. K. Anal. Chem. 1992, 64, 557–561.<br />

(24) Crutchfield, M. M.; Callis, C. F.; Irani, R. R.; Roth, G. C. Inorg. Chem. 1962,<br />

1, 813–817.<br />

(25) Teleman, A.; Richard, P.; Toivari, M.; Penttilla, M. Anal. Biochem. 1999,<br />

272, 71–79.<br />

(26) Moreno, B.; Urbina, J. A.; Oldfield, E.; Bailey, B. N.; Rodrigues, C. O.;<br />

Docampo, R. J. Biol. Chem. 2000, 275, 28356–28362.<br />

(27) Glonek, T.; Lunde, M.; Mudgett, M.; Myers, T. C. Arch. Biochem. Biophys.<br />

1971, 142, 508–513.<br />

(28) Choi, B. K.; Hercules, D. M.; Houalla, M. Anal. Chem. 2000, 72, 5087–<br />

5091.<br />

CAPILLARY ELECTROPHORESIS OF (PI)N AND<br />

DNA<br />

Capillary electrophoresis separates and resolves analytes based<br />

on differences in electrophoretic mobility. Since both the amount<br />

of negative charge and hydrodynamic drag of (Pi)n increase with<br />

n, in a way similar to that of single-stranded DNA, we expected<br />

poor resolution in the analysis of (Pi)n in free-solution capillary<br />

electrophoresis (i.e., CZE). 29 Capillaries filled with a sieving<br />

matrix, however, are capable of resolving DNA in order of chain<br />

length. 30-32 The use of replaceable solutions of entangled polymer<br />

(such as linear acrylamide or PDMA) in CGE enabled the automated<br />

and massively parallel analysis of samples of DNA and was<br />

essential to the completion of the Human Genome Project. 33,34<br />

Previous reports of methods for analyzing (Pi)n by CGE used<br />

capillaries coated and filled with linear polyacrylamide (capillaries<br />

were prepared by filling capillaries with aqueous acrylamide<br />

and polymerizing in situ). 35,36 We, and others, found<br />

that capillaries prepared this way had short lifetimes; 37-40 these<br />

capillaries could not be reused, required time-consuming preparation<br />

of new capillaries, and limited the reproducibility of the<br />

method. Stover 36,41 and Wang 29,35 detected (Pi)n by indirect UV<br />

absorbance in CGE experiments (chromate or pyromelltic acid<br />

were the chromophores added to the running buffer). These<br />

demonstrations did not, however, identify peaks for (Pi)n<br />

beyond P3 nor did they quantify resolved (Pi)n.<br />

EXPERIMENTAL DESIGN<br />

CZE: Separation of (Pi)n in Free Solution. We used the<br />

results of CZE experiments to guide the development of a CGE<br />

technique. Although the resolution of mixtures of (Pi)n in free<br />

solution is poor, CZE experiments are easier to run than CGE<br />

experiments and allowed us to (i) test different chromophores<br />

for indirect detection and identify terephthalate (TP 2- )asan<br />

optimal choice; (ii) measure µ for analytical standards of (Pi)n<br />

in free solution and determine the influence of pH and net<br />

electrostatic charge on the mobility of (Pi)n (pH in the ranges<br />

of 6.8-7.1 and 8.4-8.7); (iii) optimize the composition of the<br />

running buffer. Table 1 gives literature values of pKa for (Pi)n.<br />

Analysis by CZE required the migration of (Pi)n to the anode<br />

and the use of capillaries with suppressed electro-osmotic flow.<br />

Capillaries covalently modified by reaction of the fused silica<br />

surface with a copolymer of N,N-dimethylacrylamide and<br />

3-methacryloxy-propyltrimethoxysilane 42 had an electro-osmotic<br />

flow of


Table 1. Values of pKa of Inorganic Polyphosphates<br />

abbreviation name pKa<br />

P1<br />

P2<br />

P3<br />

cyclo-P3<br />

(Pi)n<br />

orthophosphate a<br />

pyrophosphate a<br />

tri(poly)phosphate b<br />

trimetaphosphate a<br />

polyphosphate d<br />

2.15, 7.20, 12.35<br />

0.8, 2.2, 6.7, 9.4<br />

0.5, 1.0, 2.4, 6.5, 9.4<br />

2.05 c<br />

∼1-2, 7.2-8.2 d<br />

a Values at infinite dilution and 25 °C, taken from ref 57. b Values at<br />

infinite dilution and 25 °C, taken from ref 58. c Results of titration of<br />

cyclo-P3 cannot distinguish the pKa of the three ionizable groups of<br />

cyclo-P3; each of the three groups has a value of pKa of approximately<br />

2.05. d Values taken from ref 14; ranges are inferred from titration<br />

curves for long chain polyphosphates and describe the strongly acidic<br />

hydrogen at each residue of phosphate (pKa of 1 to 2) and two weakly<br />

acidic hydrogens at the ends of the chain (pKa of 7.2-8.2).<br />

CGE: Resolution of (Pi)n in Order of n. To achieve a<br />

combination of high resolution, speed, and convenience in the<br />

analysis of (Pi)n, we used CGE with solutions of PDMA (average<br />

molecular weight of 57 kDa, 9.1% w/v) as the sieving medium.<br />

A key advantage of the use of a solution of PDMA 39,43-47 is<br />

its low viscosity ( 6.8; composition of the buffer<br />

discussed below). TP 2- carries current in the capillary and<br />

migrates to the anode during electrophoresis. Migration of TP 2-<br />

(43) Heller, C. Electrophoresis 1999, 20, 1978–1986.<br />

(44) Wang, Y. M.; Liang, D. H.; Ying, O. C.; Chu, B. Electrophoresis 2005, 26,<br />

126–136.<br />

(45) Rosenblum, B. B.; Oaks, F.; Menchen, S.; Johnson, B. Nucleic Acids Res.<br />

1997, 25, 3925–3929.<br />

(46) He, H.; Buchholz, B. A.; Kotler, L.; Miller, A. W.; Barron, A. E.; Karger,<br />

B. L. Electrophoresis 2002, 23, 1421–1428.<br />

(47) Barbier, V.; Viovy, J. L. Curr. Opin. Biotechnol. 2003, 14, 51–57.<br />

(48) Zhou, H. H.; Miller, A. W.; Sosic, Z.; Buchholz, B.; Barron, A. E.; Kotler,<br />

L.; Karger, B. L. Anal. Chem. 2000, 72, 1045–1052.<br />

(49) Kotler, L.; He, H.; Miller, A. W.; Karger, B. L. Electrophoresis 2002, 23,<br />

3062–3070.<br />

(50) Gao, Q. F.; Yeung, E. S. Anal. Chem. 1998, 70, 1382–1388.<br />

(51) Giovannoli, C.; Anfossi, L.; Tozzi, C.; Giraudi, G.; Vanni, A. J. Sep. Sci. 2004,<br />

27, 1551–1556.<br />

(52) Fung, E. N.; Pang, H. M.; Yeung, E. S. J. Chromatogr., A 1998, 806, 157–<br />

164.<br />

(53) Huang, M. F.; Huang, C. C.; Chang, H. T. Electrophoresis 2003, 24, 2896–<br />

2902.<br />

(54) Shamsi, S. A.; Danielson, N. D. Anal. Chem. 1995, 67, 1845–1852.<br />

(55) Wang, P. G.; Giesel, R. W. Anal. Biochem. 1995, 230, 329–332.<br />

(56) Wang, P. G.; Giese, R. W. Anal. Chem. 1993, 65, 3518–3520.<br />

(57) Smith, R. M.; Martell, A. E. Critical Stability Constants; Plenum Press: New<br />

York, 1989.<br />

(58) Smith, R. M.; Martel, A. E. NIST Critically Selected Stability Constants of<br />

Metal Complexes; National Institute of Standards and Technology: Gaithersburg,<br />

MD, 1997.<br />

6840 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

by the detector produces a steady-state signal in UV absorbance,<br />

generating the baseline of electropherograms. Sample<br />

zones containing (Pi)n also migrate toward the anode. Within<br />

a sample zone, the current is carried in part by (Pi)n, rather<br />

than TP 2- . Zones containing (Pi)n are detected by a decrease<br />

in UV absorbance, resulting from a decrease in the concentration<br />

of TP 2- .<br />

The sensitivity of detection of (Pi)n depends primarily on the<br />

extinction coefficient of TP 2- and the mobilities of TP 2- and<br />

(Pi)n. Terephthalate, TP 2- , absorbs at λ ) 254 nm (ε ) 8.2 · 10 3<br />

M -1 cm -1 ) and has mobility (µ ∼ 28 cm 2 kV -1 min -1 ) close to<br />

that of (Pi)n (26-34 cm 2 kV -1 min -1 in free solution, pH ) 7.0).<br />

The similarity in µ for TP 2- and (Pi)n is important for limiting<br />

dispersion in electromigration and destacking, which result in<br />

the asymmetry and broadening of peaks. 59 The estimated limit<br />

of detection of residues of Pi is ∼0.2 µM.<br />

Ions in the Running Buffer. The running buffer, by design,<br />

contains only one type of anion: TP 2- . This characteristic avoids<br />

complications in the indirect detection of (Pi)n by the absorbance<br />

of TP 2- . Anions in addition to TP 2- would lead to system<br />

zones (that show up as negative peaks unrelated to (Pi)n),<br />

artifacts in peak shape, and complications in the analysis of<br />

peak areas. 59-63 We, therefore, used buffers prepared by adding<br />

terephthalic acid to solutions of 18.0-24.0 mM bis-tris (pKa )<br />

6.46) or tris (pKa ) 8.06). The resulting buffers contained TP 2-<br />

(3.0 mM), protonated amine (6.0 mM bis-tris-H + or tris-H + ),<br />

and free amine (12.0-18.0 mM of bis-tris or tris).<br />

pH. Running buffers made with bis-tris or tris allowed us to<br />

compare the resolution of (Pi)n at two ranges of pH: 6.8-7.2<br />

(bis-tris) and 8.4-8.7 (tris). These values of pH are well<br />

beyond the pKa of the first ionizable group of phosphate<br />

residues (∼2) but are near the pKa of the second ionizable<br />

groups of the terminal residues of phosphate of (Pi)n (∼6.3-7.2)<br />

(Table 1). Values of mobility for (Pi)n are sensitive to the pH of<br />

the running buffer, particularly for n < 5.<br />

Addition of Polyethylene Glycol (PEG) to Reservoirs of<br />

Running Buffer. In CGE, capillaries are filled with a solution of<br />

PDMA (density ∼1.06 g/mL); the open ends of the capillary are<br />

immersed into reservoirs of running buffer that do not contain<br />

PDMA (∼1.01 g/mL). As a result, solutions of PDMA leak out of<br />

the capillary into the buffer reservoirs, under gravity. The<br />

reproducibility of this experiment was, therefore, poor; the current<br />

decreased by >10% within 2 h, and retention times increased by<br />

>5% in each subsequent run.<br />

The solution to this problem was to add polyethylene glycol<br />

(PEG) to the reservoirs of running buffer, generating solutions<br />

isodense with the solution inside the capillary. PEG is water<br />

soluble and uncharged; it does not migrate during electrophoresis.<br />

CGE experiments using reservoirs of running buffer with 9.0%<br />

PEG (w/v) showed stable currents and improved run-to-run<br />

reproducibility. This procedure was essential for maintaining a<br />

(59) Poppe, H.; Xu, X. In High Performance Capillary Electrophoresis: Theory,<br />

Techniques, and Applications, 1 ed.; Khaledi, M., Ed.; Wiley & Sons: New<br />

York, 1998; Vol. 146.<br />

(60) Gas, B.; Kenndler, E. Electrophoresis 2004, 25, 3901–3912.<br />

(61) Beckers, J. L.; Bocek, P. Electrophoresis 2003, 24, 518–535.<br />

(62) Bruin, G. J. M.; Vanasten, A. C.; Xu, X. M.; Poppe, H. J. Chromatogr. 1992,<br />

608, 97–107.<br />

(63) Macka, M.; Haddad, P. R.; Gebauer, P.; Bocek, P. Electrophoresis 1997,<br />

18, 1998–2007.


constant medium for separation; we routinely collected data for<br />

120 min of electrophoresis for each preparation of a filled capillary<br />

(enough for 5 to 6 typical analyses of (Pi)n). 64<br />

Loading Samples by Electrokinetic Injection. During electrokinetic<br />

injection, an applied field forces anions in the sample<br />

to enter the capillary and migrate toward the anode. Rather than<br />

transferring plugs of solution into the capillary, electrokinetic<br />

injection only transfers anions. This procedure avoids the creation<br />

of discontinuities in the capillary and improves the reproducibility<br />

of separation medium.<br />

The total number of ions injected is determined by the<br />

electrical field, conductivity of the running buffer, duration of the<br />

injection, and to a smaller extent, the conductivity of the sample<br />

(further discussion in the Supporting Information). Typical injections<br />

(175 V cm -1 for 2.0 s) correspond to the loading of ∼0.1<br />

nmol of charge. The amount of each ion injected depends on<br />

both the concentration and mobility of the ion, as well as the<br />

concentration and mobility of all other ions in the sample. Our<br />

quantitative treatment of the data accounts for the bias in<br />

sampling caused by electrokinetic injection (discussed in the<br />

Results and Discussion section and Supporting Information). 65,66<br />

Sample Preparation. The procedure described above results<br />

in the injection and detection of all anions in a sample, not just<br />

(Pi)n. 67 Analysis of (Pi)n by indirect detection works best on<br />

samples that are free of salts besides (Pi)n. The following steps<br />

were useful for preparing samples free of additional anions,<br />

originating from either synthesis or preparative separation: (i)<br />

adsorption to anion-exchange resin; (ii) elution of anions other<br />

than (Pi)n (e.g., Cl - ) with 0.1 M Na2CO3; (iii) elution of (Pi)n<br />

with concentrated 2.0 M NH4HCO3; (iv) removal of NH4HCO3<br />

under vacuum. 68<br />

Analytes: Authentic Standards of (Pi)n (n ) 1-10).<br />

Commercially available oligophosphates of a single chain length<br />

are limited to P1,P2,P3, and cyclo-P3. Preparative-scale separation<br />

of a mixture of (Pi)n (117% polyphosphoric acid 69 ), by anionexchange<br />

chromatography, generated samples of purified (Pi)n,<br />

n ) 4-10, that served as analytical standards. By analyzing<br />

standards added to mixtures of (Pi)n, we identified peaks<br />

representing (Pi)n with n ) 1-10.<br />

Analytes: Mixtures of (Pi)n. We demonstrated the resolution<br />

of the method by analyzing commercially available samples of<br />

higher (Pi)n, covering a range of average length (n¯): 117%<br />

polyphosphoric acid, n¯ ) 17, n¯ ) 21, n¯ ) 48, and n¯ ) 65. In<br />

addition, we prepared samples of (Pi)n generated by the<br />

dehydration of NH4H2PO4 in mixtures with urea, using conditions<br />

reported by Orgel et.al., which were presumed to be<br />

plausible in prebiotic chemistry. 8,21<br />

Internal standard(s). We added an internal standard,<br />

K + CH3SO3 - , to each sample prior to analysis by CE. The peak<br />

(64) We used PEG instead of PDMA in buffer reservoirs because commercial<br />

PDMA (∼$100 per g) is much more expensive than PEG (∼$0.1 per g);<br />

each set of experiments requires ∼1 g of polymer. PDMA can be obtained<br />

inexpensively, however, by preparing it from N,N-dimethylacrylamide.<br />

(65) Huang, X. H.; Gordon, M. J.; Zare, R. N. Anal. Chem. 1988, 60, 375–377.<br />

(66) Rose, D. J.; Jorgenson, J. W. Anal. Chem. 1988, 60, 642–648.<br />

(67) Satow, T.; Machida, A.; Funakushi, K.; Palmieri, R. J. High Resolut.<br />

Chromatogr. 1991, 14, 276–279.<br />

(68) Cohn, W. E.; Bollum, F. J. Biochim. Biophys. Acta 1961, 48, 588–590.<br />

(69) The nomenclature (i.e., 117%) is based on comparing the ratio of phosphorus<br />

to oxygen in samples of dehydrated phosphate to the ratio of a standard<br />

solution of 85% phosphoric acid.<br />

observed for the internal standard allowed us to (i) monitor<br />

the reproducibility of the method; (ii) define the mobilities of<br />

(Pi)n relative to that of a standard (CH3SO3 - )(µP,rel in eq 1);<br />

(iii) determine the relative and absolute concentrations of resolved<br />

(Pi)n, by comparing the areas of peaks for CH3SO3 - and (Pi)n.<br />

In eq 1, tCH3SO3 - and tP are the retention times for CH3SO3 - and<br />

(Pi)n, V is the voltage, L1 is the length of the capillary, and L2<br />

is the distance between the inlet and detector.<br />

µ P,rel ) µ P - µ CH3SO - )<br />

3 L1L2 V (<br />

1<br />

t CH3 SO 3 -<br />

- 1<br />

t P)<br />

The electrophoretic mobility of CH3SO3 - in free solution was<br />

near that of (Pi)n, but peaks for CH3SO3 - and (Pi)n did not<br />

overlap (µCH3SO3 - ) 26.7 cm 2 kV -1 min -1 ). For experiments<br />

requiring additional internal standards, we also used Cl - ,<br />

CF3CO2 - , and CH3-C6H4-SO3 - .<br />

RESULTS AND DISCUSSION<br />

Resolution and Detection of Lower Oligophosphates by<br />

CZE. We analyzed mixtures of P1, P2, P3, and cyclo-P3 by CZE,<br />

using coated capillaries and running buffer composed of 3.0<br />

mM TP 2- and 18.0 mM bis-tris (pH ) 6.9) or 24.0 mM tris<br />

(pH ) 8.4); 70 data from these experiments are shown in the<br />

Supporting Information. The results demonstrated (i) the indirect<br />

detection of anions, enabled by the steady-state absorbance signal<br />

of TP 2- at 254 nm and (ii) mobilities in the order cyclo-P3 > P3<br />

> P2 > P1 (at pH ) 6.9) that correspond to the number of<br />

ionizable groups, values of pKa, and structures of these (Pi)n<br />

(i.e., the radius of cyclo-P3 is constrained in a way that P3 is<br />

not).<br />

Resolution of Mixtures of (Pi)n by CGE. The resolution of<br />

mixtures of (Pi)n with n > 5 required the use of a sieving gel.<br />

The best results were obtained with capillaries filled with<br />

solutions of 9.1% PDMA (w/v; average molecular weight 58.9<br />

kDa) in running buffer (24.0 mM tris, 3.0 mM terephthalic acid,<br />

pH ) 8.4). We filled capillaries (100 µm internal diameter and<br />

57 cm in length) by pumping in solutions of PDMA with<br />

positive pressure (30 psi of ultrahigh purity N2 applied to the<br />

inlet) for 10 min. During electrophoresis, the ends of the<br />

capillary were immersed into reservoirs of running buffer<br />

containing 9% PEG (w/v; average molecular weight of 1.5 kDa).<br />

The upper trace in Figure 1A shows the separation of (Pi)n in<br />

a commercially available mixture of sodium polyphosphate<br />

(reported chain length of n¯ ∼ 17). The sample had a concentration<br />

of 19.0 mM (in phosphate residues); P3 (2.0 mM) and<br />

CH3SO3 - (2.0 mM) were added to serve as internal standards.<br />

For comparison, the lower trace in Figure 1A shows the analysis<br />

of the same sample by CZE.<br />

We identified peaks for P1, P2, and P3 (marked with a dotted<br />

line in Figure 1B) by comparing the traces in Figure 1A to those<br />

collected for samples containing standards added to the mixture.<br />

In CGE experiments, mobilities of (Pi)n with n > 3 are lower than<br />

that of CH3SO3 - . In CZE, the mobilities of all (Pi)n are greater<br />

than that of CH3SO3 - . The contrast in the order of mobility<br />

(70) Capillaries modified by the method of Cretich 42 were used over the course<br />

of several months in more than 100 runs, without observable change in<br />

the retention times for analytical standards.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(1)<br />

6841


Figure 1. Resolution of (Pi)n by CGE. (A) Analysis of sodium polyphosphate (average chain length 17) in capillaries filled with 9.1% PDMA<br />

(w/v) gel (upper trace) or running buffer alone (24.0 mM tris, 3.0 mM terephthalic acid, pH ) 8.4, 25 °C) (lower trace). Electrophoresis was<br />

performed by applying 14.6 kV across capillaries having a length of 57 cm (50 cm from the inlet to detector). (B,C) Expanded views of the upper<br />

and lower traces in A, respectively.<br />

indicates size-sieving during CGE and the separation of (Pi)n<br />

in order of n. The series of peaks detected by CGE (shown in<br />

Figure 1B) suggests species as large as ∼P35 in the sample.<br />

6842 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Preparation of (Pi)n, n ) 4-10. Identification of peaks<br />

observed for broad distributions of (Pi)n required oligophosphate<br />

standards that are not commercially available. We separated a


Figure 2. Standards of (Pi)n, n ) 1-10. CZE traces analyzing a mixture of commercially available P1, P2, P3, cyclo-P3, and CH3SO3 - (top<br />

trace) and analytical standards P4-P10 purified by anion-exchange chromatography (no internal standard added). CZE was performed by applying<br />

14.6 kV across coated capillaries (57 cm in length, 50 cm from the inlet to detector) filled with running buffer composed of 18.0 mM bis-tris and<br />

3.0 mM terephthalic acid (pH ) 6.9, 25 °C).<br />

Figure 3. Identification of P3-P10 in mixtures of (Pi)n by CGE. Traces are for the analysis of mixtures of (Pi)n (19 mM in total phosphate, pH<br />

∼ 8) with the addition of analytical standards for (Pi)n (2 mM). Mixtures were resolved by CGE at 14.6 kV using capillaries filled with 9.1% PDMA<br />

(w/v) and running buffer (24.0 mM tris, 3.0 mM TP 2- ,pH) 8.4). Dotted lines mark peaks for species identified by added standards. Traces for<br />

samples with added P7 or P9 are available in the Supporting Information.<br />

mixture of (Pi)n (n ∼ 1-10, derived from 117% polyphosphoric<br />

acid 69 ) on an anion-exchange chromatography column (Cl -<br />

form) by elution with a gradient of KCl(aq). A second application<br />

of anion-exchange chromatography removed KCl from fractions<br />

containing (Pi)n (HCO3 - form; elution with NH4HCO3). Removal<br />

of NH4HCO3 by vacuum generated oligophosphate standards<br />

as the ammonium salt.<br />

CZE traces analyzing purified (Pi)n, n ) 4-10, show one<br />

major peak in each sample (Figure 2). Small peaks next to the<br />

major peak and a small peak for P1 (top trace in Figure 2 for<br />

reference) suggest small amounts of (Pi)(n+1) or (Pi)(n-1), and P1.<br />

The relatively clean traces suggest that the conditions used in<br />

the preparation of samples do not cause extensive hydrolysis<br />

of (Pi)n or equilibration in chain length. Characterization of<br />

oligophosphate standards by 31 P NMR is available in the<br />

Supporting Information.<br />

Identification of Peaks for n ) 3-10 in Mixtures of (Pi)n.<br />

Figure 3 shows CGE data analyzing mixtures of (Pi)n (19.0 mM<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6843


Figure 4. Resolution of commercially available mixtures of (Pi)n. Samples consisting of commercially available mixtures of polyphosphate<br />

glass (20 mM in total phosphate, Na + salt) and CH3SO3 - (2 mM) were analyzed by CGE. After electrokinetic injection (4.0 s at 10 kV), (Pi)n were<br />

separated by electrophoresis at 14.6 kV in coated capillaries (57 cm in length, 50 cm between inlet and detector) filled with solutions of 9.1%<br />

PDMA (w/v) and running buffer (24.0 mM tris, 3.0 mM TP 2- ,pH) 8.4). Peaks in trace C, for n ∼ 10-40, appear less sharp than the peaks in<br />

other runs. The reason for this difference is unclear; one possibility is that the peaks are affected by irregularities in pH within capillaries during<br />

separation, caused by a small mismatch in pH between samples (∼8) and the running buffer (pH ) 8.4). 72<br />

in phosphate residues; n¯ ∼ 17) and added P3, P4, P6, P8, orP10<br />

(2.0 mM). The x-axis of traces in Figure 3 are in units of mobility<br />

relative to CH3SO3 - µp,rel (eq 1). The traces show the alignment<br />

of peaks from run to run and allowed us to assign peaks with<br />

increased area to specific (Pi)n (dotted lines in Figure 3). The<br />

dotted line for n ) 20 in Figure 3B is based on the reasonable<br />

assumption that the peaks continue in order of n. 71<br />

Mixtures of (Pi)n with n up to ∼70. The CGE data in Figure<br />

4 characterizes commercially available mixtures of (Pi)n with<br />

different distributions in chain length. Samples analyzed in<br />

traces A-D are in order of increasing average chain length.<br />

The results show that peaks with n up to ∼70 can be resolved<br />

in a single run; longer (Pi)n may potentially be resolved using<br />

lower concentrations of PDMA.<br />

Shapes of Peaks. The asymmetric peaks observed in both<br />

CZE and CGE experiments are typical for electropherograms<br />

collected by indirect UV absorbance. (Pi)n with mobility greater<br />

than that of TP 2- shows up as peaks broadened to the left,<br />

while (Pi)n with mobility less than that of TP 2- show up as peaks<br />

broadened to the right. These shapes are the results of<br />

dispersion by electromigration, originating from (i) differences<br />

in mobility between analytes and TP 2- and (ii) nonuniform<br />

electric fields inside sample zones. Discussion of the origin of<br />

the shapes of peaks is available in the Supporting Information,<br />

as well as in ref 59.<br />

Quantitative Analysis of (Pi)n: Areas of Peaks. Three<br />

contributions determine the area of a peak: (i) the response of<br />

[TP 2- ]to(Pi)n; (ii) the amount of (Pi)n transferred from the<br />

sample to the capillary by electrokinetic injection; (iii) residence<br />

time of the analyte passing the detector. Analysis of the areas<br />

of peaks for (Pi)n and CH3SO3 - enables the quantification of<br />

resolved (Pi)n. To account for the effect of (iii), areas of peaks<br />

are adjusted by the factor (1/ti), where t is the retention time<br />

of analyte i. 73 In eq 2, the ratio of adjusted areas for analyte i<br />

and CH3SO3 - (Ai and ACH3SO3 -) is related to the concentrations<br />

[i]S and [CH3SO3 - ] of the sample: zi and zCH3SO3 - are the<br />

electrostatic charge of i and CH3SO3 - ; µCH3SO3 - and µi are the<br />

mobilities of i and CH3SO3 - ; µC + is the mobility of the cation<br />

in the sample zone (bis-tris-H + or tris-H + ).<br />

A i<br />

A CH3 SO 3 -<br />

[i] S<br />

)<br />

[CH3SO3 - ] · [<br />

z i<br />

z CH3 SO 3 -]<br />

· µ + µ C+ i [ µ + µ C+ CH3SO -] 3<br />

(2)<br />

The Supporting Information contains our derivation of eq 2 and a<br />

discussion of the quantitative aspects of electrokinetic injection<br />

and indirect UV absorbance.<br />

Equation 2 reveals the advantage of analyzing electropherograms<br />

by comparing peaks for (Pi)n and CH3SO3 - . The ratio (Ai/<br />

ACH3SO3 -) depends on the ratio ([i]S/[CH3SO3 - ]) but does not<br />

depend on the concentration of other ions in the sample or<br />

the voltage and duration of electrokinetic injection.<br />

Calibration of Peak Areas to Concentrations of P1, P2,<br />

P3, and cyclo-P3. Analysis of samples of containing P1, P2, P3,<br />

or cyclo-P3 and added CH3SO3 - by CZE demonstrated the<br />

relationship between peak areas and concentrations. Figure 5<br />

shows values of (Acyclo-P3 /ACH3SO3 -) determined from the analysis<br />

of samples containing cyclo-P3 and CH3SO3 - . Samples had ratios<br />

(71) The alignment is not perfect; small differences from run to run are possibly<br />

due to dispersion in electromigration, caused by differences between<br />

samples in pH or ionic strength.<br />

(72) Another possibility is that electrokinetic injections may result in concentrations<br />

of (Pi)n at the inlet that can lead to precipitation. (73) Hilser, V. J.; Freire, E. Anal. Biochem. 1995, 224, 465–485.<br />

6844 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010


Figure 5. Quantitative calibration of cyclo-P3. The plot shows values<br />

for Acyclo-P3 /ACH3SO3 - determined in the analysis of cyclo-P3 and<br />

CH3SO3 - by CZE, using coated capillaries (57 cm in length, 50 cm<br />

between inlet and detector) filled with running buffer (18.0 mM<br />

bis-tris, 3.0 mM TP2- ,pH) 6.9). Points for ([cyclo-P3]/[CH3SO3 - ])<br />

) 0.1, 1.0, or 10.0 are the average taken from eight experiments.<br />

Standard deviations for (Acyclo-P3<br />

/ACH3SO3 -) are 40 and distinguished the composition of<br />

mixtures generated at 125 and 140 °C. Cyclo-P3 is the most<br />

abundant species in mixtures prepared at 140 °C. In contrast,<br />

the mixture prepared at 125 °C does not contain cyclo-P3, despite<br />

containing linear (Pi)n with n > 40. The reason for the<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6845


preferential formation of cyclo-P3 over linear (Pi)n at 140 °C is<br />

not clear; the synthesis of cyclo-P3 is, however, potentially<br />

important for prebiotic chemistry. Reactions of (Pi)n with -OH<br />

groups, leading to polyphosphorylated compounds, likely<br />

depend on whether (Pi)n are cyclic or linear. Reactions of linear<br />

(Pi)n potentially transfer phosphate residues from either terminal<br />

or middle positions of the chain, while reactions of cyclo-<br />

P3 are ring-opening and can generate a triphosphate group<br />

similar to that of ATP. The resolution and convenience of CGE<br />

should enable a broad survey of conditions for the synthesis<br />

of (Pi)n and for reactions of (Pi)n with -OH groups. The results<br />

should be helpful in refining hypotheses for the importance of<br />

dehydrated phosphate in the chemical origins of life.<br />

CONCLUSION<br />

Previously reported methods using electrophoresis (slab gels<br />

or CGE) to analyze (Pi)n successfully resolved species on the<br />

basis of their size (n). These methods provided a way to<br />

qualitatively characterize mixtures of (Pi)n at high resolution<br />

but involved difficult and time-consuming experiments with<br />

limited reproducibility. The method we have demonstrated in<br />

this paper exploits the sieving properties of low-viscosity<br />

solutions of PDMA and is capable of high resolution and<br />

quantitative analysis. The advantages of CGE using solutions<br />

of PDMA as the separation medium are convenient preparation,<br />

rapid analysis, and reproducibility (enabled by refilling capillaries<br />

with fresh solutions of gel). Our identification of P 1-P10<br />

and resolution up to P70 validates a method that will be useful<br />

6846 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

in studies of (Pi)n relevant to prebioitic chemistry and<br />

biochemistry.<br />

In addition to characterizing the composition of samples of<br />

(Pi)n, the method we have developed is potentially useful for<br />

analyzing other anionic, oligomeric species without a sensitive<br />

chromophore. Examples relevant to prebiotic chemistry are<br />

oligomers of phosphate condensed with organic compounds<br />

(e.g., structures with formula (PO3 - -RO)n or (P2O6 2- -RO)n)).<br />

Examples of biological polymers include teichoic acid, hyaluronic<br />

acid, and sulfated polysaccharides such as heparin and<br />

chondroitin sulfate.<br />

ACKNOWLEDGMENT<br />

We acknowledge the Harvard University Origins of Life Initiative<br />

for research support. We thank Douglas B. Weibel, Katherine L.<br />

Gudiksen, Paul J. Bracher, and Dosil Pereira de Jesus for technical<br />

assistance and helpful discussion. A.L. acknowledges salary support<br />

from NIH GM051559.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in the text, as well as<br />

information about experimental procedures and sources of<br />

chemicals. This material is available free of charge via the<br />

Internet at http://pubs.acs.org.<br />

Received for review March 28, 2010. Accepted July 10,<br />

2010.<br />

AC1008018


Anal. Chem. 2010, 82, 6847–6853<br />

Method to Determine 226 Ra in Small Sediment<br />

Samples by Ultralow Background Liquid<br />

Scintillation<br />

Joan-Albert Sanchez-Cabeza,* ,†,‡ Laval Liong Wee Kwong, § and Maria Betti §<br />

Institute of Environmental Science and Technology, and Physics Department, Autonomous University of Barcelona,<br />

ES-08193 Bellaterra, Spain, Departamento de Medio Ambiente, CIEMAT, 28040 Madrid, Spain, and Environment<br />

Laboratories, International Atomic Energy Agency, MC-98000 Monaco<br />

210 Pb dating of sediment cores is a widely used tool to<br />

reconstruct ecosystem evolution and historical pollution<br />

during the last century. Although 226 Ra can be<br />

determined by γ spectrometry, this method shows<br />

severe limitations which are, among others, sample<br />

size requirements and counting times. In this work,<br />

we propose a new strategy based on the analysis of<br />

210 Pb through 210 Po in equilibrium by r spectrometry,<br />

followed by the determination of 226 Ra (base or supported<br />

210 Pb) without any further chemical purification<br />

by liquid scintillation and with a higher sample throughput.<br />

Although γ spectrometry might still be required<br />

to determine 137 Cs as an independent tracer, the effort<br />

can then be focused only on those sections dated<br />

around 1963, when maximum activities are expected.<br />

In this work, we optimized the counting conditions,<br />

calibrated the system for changing quenching, and<br />

described the new method to determine 226 Ra in small<br />

sediment samples, after 210 Po determination, allowing<br />

a more precise determination of excess 210 Pb ( 210 Pbex).<br />

The method was validated with reference materials<br />

IAEA-384, IAEA-385, and IAEA-313.<br />

The natural radionuclide 210 Pb is used as an environmental<br />

tracer of numerous biogeochemical processes in the aquatic,<br />

soil, and atmospheric sciences. Among these, possibly the most<br />

common application is its use in the reconstruction of recent<br />

environmental changes through 210 Pb dating, as described in<br />

some seminal papers. 1-5 210 Pb (half-life ) 22.23 ± 0.12 years) 6<br />

* To whom correspondence should be addressed. Phone: +34 93 581 1915.<br />

E-mail: joanalbert.sanchez@uab.cat.<br />

† Autonomous University of Barcelona.<br />

‡ CIEMAT.<br />

§ International Atomic Energy Agency.<br />

(1) Goldberg, E. D. In Radioactive Dating; International Atomic Energy Agency:<br />

Vienna, Austria, 1963; pp 121-131.<br />

(2) Crozaz, G.; Picciotto, E.; de Breuck, W. J. Geophys. Res. 1964, 69, 2597–<br />

2604.<br />

(3) Krishnaswamy, S.; Lal, D.; Martin, J.; Meybeck, M. Earth Planet. Sci. Lett.<br />

1971, 11, 407–414.<br />

(4) Robbins, J. A. In Biochemistry of Lead; Nriagu, J. O., Ed.; Elsevier:<br />

Amsterdam, The Netherlands, 1998; pp 285-393.<br />

(5) Appleby, P. G.; Oldfield, F. Catena 1978, 5, 1–8.<br />

(6) All half-lives used in this work are from Decay Data Evaluation Project,<br />

http://www.nucleide.org/DDEP.htm, updated by the “Laboratoire National<br />

Henri Becquerel” on January 22, 2010.<br />

is a natural radionuclide of the 238 U radioactive chain. In closed<br />

systems, 226 Ra (T1/2 ) 1600 ± 7 years) decays, through various<br />

daughter radionuclides, to 210 Pb, named base or supported<br />

210 Pb, which should be in equilibrium with 226 Ra (base 210 Pb )<br />

226 Ra). On the other hand, some 210 Pb can reach bottom aquatic<br />

sediments from either the atmosphere (after 222 Rn exhalation<br />

from soils) or the water column (in situ production). This is<br />

called the excess or unsupported fraction ( 210 Pbex) and is the<br />

basis of all 210 Pb dating models. 7 Therefore, bottom sediments<br />

contain a mixture of base and excess 210 Pb, and 210 Pbex is<br />

determined by the difference between the total 210 Pb and 226 Ra<br />

concentration for each sediment section ( 210 Pbex ) 210 Pb -<br />

226 Ra). This is commonly determined in sections (typically 1<br />

cm width) of undisturbed sediment cores collected from areas<br />

of interest.<br />

High-resolution γ spectrometry with HPGe (high-purity Ge)<br />

detectors is commonly used to simultaneously determine both<br />

210 Pb and 226 Ra in sediment samples. 8 Some of the reasons for<br />

its success are (i) 137 Cs is also determined, which may be used<br />

to validate the 210 Pb chronology, 9 and (ii) it is nondestructive,<br />

as it does not require chemical separation. However, it has<br />

important drawbacks for 210 Pb dating: (i) calibration in the low<br />

γ emission energy region is difficult and requires tedious selfabsorption<br />

corrections, thus affecting method accuracy, (ii)<br />

precision is limited by the unavoidable presence of a relevant<br />

background due to Compton scattered electrons in the detector,<br />

(iii) counting times are long, thus sample throughput is small,<br />

and (iv) most importantly for 210 Pb dating applications, sample<br />

size requirement is usually large, typically more than 5 g dry<br />

weight (dw), although some laboratories with HPGe well-type<br />

detectors and/or special low-level counting conditions might<br />

use a sample size as low as 1-2 g dw. Sample size is in many<br />

cases a limitation for 210 Pb dating as commonly a large number<br />

of other analyses need to be carried out. Although all samples<br />

could be measured by γ spectrometry and saved for further<br />

analysis, special care is needed during sample manipulation<br />

(e.g., if the sample cannot be ground) and in order to avoid<br />

(7) Appleby, P. G. Chronostratigraphic techniques in recent sediments. In<br />

Tracking Environmental Change Using Lake Sediments: Basin Analysis,<br />

Coring, and Chronological Techniques, Last, W. M., Smol, J. P., Eds.; Kluwer<br />

Academic, 2001; Vol. 1, pp 171-203.<br />

(8) Schelske, C. L.; Peplow, A.; Brenner, M.; Spencer, C. N. J. Paleolimnol.<br />

1994, 10, 115–128.<br />

(9) Smith, J. N. J. Environ. Radioact. 2001, 55, 121–123.<br />

10.1021/ac1008332 © 2010 American <strong>Chemical</strong> Society 6847<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/16/2010


contamination of the substances of interest (e.g., trace metals).<br />

When this strategy is used, the time needed to complete γ<br />

spectrometry before further analyses can be performed also<br />

needs to be taken into consideration.<br />

In many laboratories, the total 210 Pb concentration is determined<br />

by R spectrometry, through its daughter radionuclide<br />

210 Po in equilibrium. The technique is simple, rapid, reliable<br />

(recoveries >95%), and shows good precision. 10,11 Furthermore,<br />

although the technique is destructive, only ca. 250 mg of dry<br />

sediment is needed, and because laboratories usually have several<br />

R detectors (they are much cheaper than Ge detectors), sample<br />

throughput is much larger than with γ spectrometry. Typical<br />

counting times range from 1 to 7 days, depending on sample<br />

activity and desired uncertainty. An experimented analyst might<br />

produce a full total 210 Pb profile within a month after sample<br />

receipt (ca. 40 samples). When R spectrometry is used, the<br />

base 210 Pb can only be indirectly determined by assuming that<br />

the 226 Ra concentration is constant along the profile and<br />

therefore estimated as the average of concentrations in the<br />

profile bottom sections, as equilibrium should have been<br />

reached. 12 However, as 226 Ra may vary along the profile, this<br />

may lead to inaccurate 210 Pbex values. It is not uncommon that,<br />

once the total 210 Pb profile is known, γ spectrometry is carried<br />

out in selected sections, thus optimizing the use of this limiting<br />

resource, although because of sample size requirements this<br />

might be impracticable for many laboratories. As γ spectrometry<br />

is carried out on the untreated sample and 210 Po on a<br />

digestate, these techniques sometimes yield results which are<br />

not fully consistent.<br />

Liquid scintillation counting (LSC), mostly used for the<br />

determination of � emitters, 13,14 has also been extensively used<br />

to quantify R emitters, such as 226 Ra, in environmental samples<br />

(mainly waters). 15,16 Some of the reported methods include direct<br />

R LSC after water sample concentration, 17 226 Ra extraction from<br />

water samples with specific scintillation cocktails such as<br />

Radaex, 18 or generation of a radium-barium sulfate coprecipitate<br />

that is transformed into a soluble chloride or nitrate. 19,20<br />

Villa et al. 21 opted for this approach for sediment: the sediment<br />

is digested, and after the elimination of actinides as hydroxides,<br />

radium is recovered as Ra-Ba-SO4, dissolved in EDTA 0.2 M<br />

ammonia solution, and counted. However, most of these<br />

methods cannot be directly used with sediment digestates and/<br />

or are excessively resource-consuming.<br />

(10) Sanchez-Cabeza, J. A.; Masqué, P.; Ani-Ragolta, I. J. Radioanal. Nucl. Chem.<br />

1998, 227, 19–22.<br />

(11) Vesterbacka, P.; Ikaheimonen, T. K. Anal. Chim. Acta 2005, 545, 252–<br />

261.<br />

(12) Binford, M. W. J. Paleolimnol. 1990, 3, 253–268.<br />

(13) Pujol, L.; Sanchez-Cabeza, J. A. J. Radioanal. Nucl. Chem. 1999, 2, 391–<br />

398.<br />

(14) Liong Wee Kwong, L.; LaRosa, J. J.; Lee, S. H.; Povinec, P. P. J. Radioanal.<br />

Nucl. Chem. 2000, 248, 751–755.<br />

(15) Salonen, L. Sci. Total Environ. 1993, 130-131, 23–35.<br />

(16) Salonen, L.; Hukkanen, H. J. Radioanal. Nucl. Chem. 1997, 226, 67–74.<br />

(17) Sanchez-Cabeza, J. A.; Pujol, L. Analyst 1998, 123, 399–403.<br />

(18) Aupiais, J. Anal. Chim. Acta 2005, 532, 199–207.<br />

(19) Repinc, U.; Benedik, L. J. Radioanal. Nucl. Chem. 2002, 254, 181–185.<br />

(20) Galan-Lopez, M.; Martin-Sanchez, A.; Tosheva, Z.; Kies, A. In LSC 2005<br />

Advances in Liquid Scintillation Spectrometry; Chalupnik, S., Schoenhofer,<br />

F., Noakes, J., Eds.; Radiocarbon: Tucson, AZ, 2006; pp 165-170.<br />

(21) Villa, M.; Moreno, H. P.; Manjón, G. Radiat. Meas. 2005, 39, 543–550.<br />

6848 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

After 210 Po analysis for R spectrometry, 226 Ra remains in<br />

solution. In this work, we propose, develop, and validate a new<br />

method to determine 226 Ra by 222 Rn emanation to a scintillation<br />

cocktail, which eliminates the need to perform any further<br />

purification, and counting with an ultralow background liquid<br />

scintillation system.<br />

EXPERIMENTAL SECTION<br />

Equipment. An ultralow background liquid scintillation system,<br />

Quantulus 1220 TM (Wallac, Turku, Finland), was used to<br />

carry out this work. In this system, background is reduced by<br />

an optimized combination of active and passive shields. A pulseshape<br />

analysis (PSA) circuit permits the discrimination of<br />

pulses produced by R and � radiation by comparing the area<br />

of the pulse tail after 50 ns from the start with its total area.<br />

Pulse-shape discrimination is accomplished using a software<br />

adjustable parameter (PSA parameter) which can vary between<br />

1 and 256. 16,22 Quenching (sample extinction) was quantified with<br />

the standard quenching parameter (SQP(E)) which is used to<br />

determine the counting efficiency for each sample through<br />

calibration curves. 23,24 Counting was performed with Wallac<br />

OptiScint HiSafe III, a diisopropyl naphthalene based aqueous<br />

immiscible cocktail, and low-diffusion PE counting vials (Packard<br />

BioScience).<br />

Counting Solutions. The tracer solutions were prepared by<br />

gravimetrically spiking 226 Ra/2 M HNO3 (NIST, SRM4967,<br />

U.S.A.) into known amounts of deionized water contained in<br />

20 mL low-diffusion PE counting vials. OptiScint HiSafe was<br />

then added to reach a total admixture volume of 20 mL. These<br />

were stored for 3 weeks in a dark temperature-controlled area<br />

to allow in-growth and equilibrium of the radioactive progenies.<br />

The background solutions, used for calibration purposes, were<br />

prepared with 10 mL of deionized water, acidified to match<br />

the standard solutions, to which 10 mL of the scintillation<br />

cocktail was added. In all cases, quenching was changed by<br />

adding different amounts of CCl 4, ranging from 0 to 200 µL.<br />

All reagents used in the experiments were of analytical grade<br />

(Fisher Scientific).<br />

RESULTS<br />

When counting a 226 Ra aqueous solution with an immiscible<br />

scintillant (such as OptiScint Hisafe), the R emitter 226 Ra decays<br />

to the R emitter 222 Rn (T1/2 ) 3.8332 ± 0.0008 days). Radon is<br />

highly soluble in oil-based scintillators and is selectively<br />

extracted in the cocktail, suffering some decay while this<br />

process takes place. Once solubilized in the organic phase,<br />

222 Rn decays to the R emitter 218 Po (T1/2 ) 3.094 ± 0.006 min),<br />

this mainly decays to the � emitter 214 Pb (T1/2 ) 26.8 ± 9 min),<br />

which decays to the � emitter 214 Bi (T1/2 ) 19.9 ± 0.4 min),<br />

which mainly decays to the R emitter 214 Po (T1/2 ) 162.3 ± 1.2<br />

µs), and this one to 210 Pb (T1/2 ) 22.23 ± 0.12 years). Therefore,<br />

in the scintillant, and after an appropriate equilibration time<br />

(usually set to about 3 weeks), the R emitters 222 Rn, 218 Pb, and<br />

214 Pb are in secular equilibrium (Figure 1). As the probability<br />

of these R decays is in all cases close to one, the maximum<br />

(22) Kaihola, L. J. Radioanal. Nucl. Chem. 2000, 243, 313–317.<br />

(23) Villa, M.; Manjon, G.; Garcia-Leon, M. Nucl. Instrum. Methods Phys. Res.,<br />

Sect. A 2003, 496 (2-3), 413–424.<br />

(24) Sanchez-Cabeza, J. A.; Pujol, L. Health Phys. 1995, 68 (5), 674–82.


Figure 1. 226 Ra spectrum by liquid scintillation (only 6 × 10 -4 Bq).<br />

Vertical dashed lines indicate the approximate counting windows.<br />

efficiency reached when counting all R events should be 300%.<br />

This is another advantage of R versus γ spectrometry, where final<br />

counting efficiencies for 226 Ra, depending on the counting<br />

configuration, normally do not usually exceed a few percent.<br />

The magnitudes used for optimization were those related to<br />

counting precision.<br />

• Counting efficiency E: this is the ratio between the observed<br />

number of counts and the 226 Ra decays. The maximum<br />

observed efficiency is close to 300% (Figure 1) because with<br />

this method we simultaneously count R particles from 222 Rn,<br />

218 Pb, and 214 Pb in equilibrium.<br />

• Minimum detectable activity (MDA): 25 we used the following<br />

expression<br />

2.71 + 4.65√Bt<br />

MDA)<br />

tVE<br />

where B is the background count rate, t is the counting time,<br />

and V is the volume of solution used (in the case of sediment<br />

analysis, V was substituted by m, the mass of the aliquot<br />

analyzed).<br />

• Figure of merit: this is a magnitude commonly used to<br />

compare methods, which emphasizes counting efficiency (and<br />

therefore sample throughput). FM was calculated as FM )<br />

E 2 /B.<br />

In this section, we describe the results of the optimization<br />

process for the relevant counting parameters.<br />

Optimal Admixture Composition. Once the scintillation vial<br />

and total volume were fixed, the optimal “scintillator-to-water” ratio<br />

was optimized. We prepared and counted several composition<br />

mixtures with tracer solutions and background solutions, by using<br />

1, 5, 10, 15, and 19 mL of scintillant (Figure 2).<br />

Not unexpectedly, the highest efficiency was observed for the<br />

maximum volume of scintillant, as this maximizes radiation<br />

interaction with the detector (the scintillant). However, as the<br />

scintillant itself is an important source of background, this also<br />

increases the MDA value. This behavior is well-captured with the<br />

FM, which shows a maximum value for a 10:10 mL admixture.<br />

Therefore, we used this proportion in all experiments.<br />

Optimal PSA Parameter. The PSA circuit of Quantulus 1220<br />

sends the signals to one of the two multichannel analyzers<br />

Figure 2. Determination of the optimal admixture composition.<br />

Figure 3. Change of background and 226 Ra counting efficiency with<br />

PSA.<br />

depending on the result of the PSA analysis. However, this method<br />

is not error-free and shows interference as (i) R and � events can<br />

be wrongly assigned and (ii) the efficiency of this method depends<br />

on the energy of the incident radiation particles. 26 In order to<br />

minimize the interference, tracer solutions of 226 Ra activity 5.34<br />

Bq and background solutions, with a 10:10 mL admixture<br />

composition, were counted by LSC during 12 h and with the<br />

PSA parameter ranging from PSA ) 0 (no R-� discrimination)<br />

to PSA ) 256 (maximum R-� discrimination) with a step<br />

increment of 5 (Figure 3).<br />

The background signals assigned to the R spectrum have a<br />

varied nature and may include γ interactions with the scintillant,<br />

� events wrongly assigned to the R spectrum, R emission from<br />

impurities in the vial, and a variety of cosmic-ray originated signals.<br />

Therefore, the background spectrum does not show prominent<br />

peaks. The background count rate versus PSA plot (Figure 3)<br />

shows an almost constant value until PSA ∼ 90, as all events are<br />

recorded in the R spectrum, and then its value decays smoothly<br />

to a value close to 0 cpm, when all background events are recorded<br />

in the � spectrum.<br />

On the other hand, the 226 Ra spectrum shows well-resolved<br />

peaks within a quite narrow energy window (Figure 1), and<br />

the effect of the R particle energy on the interference is clearly<br />

shown in the efficiency versus PSA curve (Figure 3). Until PSA<br />

∼ 110, the total efficiency exceeds the maximum value of 300%<br />

because of the misclassification of background and � events in<br />

the R spectrum. In the PSA range of 120-160, most background<br />

(25) Currie, L. A. Anal. Chem. 1968, 40, 586–593. (26) Pujol, L.; Sanchez-Cabeza, J. A. Analyst 1997, 122, 383–385.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6849


Figure 4. Variation of MDA and FM with respect to PSA for 226 Ra<br />

standard solutions.<br />

and � events are correctly discriminated and a counting plateau,<br />

of almost constant efficiency (ca. 265%), is observed.<br />

In order to choose the optimum value within the efficiency<br />

plateau, both MDA and FM were plotted (Figure 4). In broad<br />

terms, MDA followed the efficiency curve, suggesting that, from<br />

a limit of detection perspective and within the plateau, best results<br />

would be achieved at higher PSA, as the background slowly<br />

decreases with the PSA parameter and stabilizes for the PSA<br />

ranging from 145 to 190. It is worth noticing that, as from PSA )<br />

135 the background rate is very low, the MDA is correspondingly<br />

low. Higher PSA values are discarded from the discussion as the<br />

efficiency there tends rapidly to zero. On the other hand, the FM<br />

shows a clear peak at PSA ) 145, and shows the optimum<br />

combination of efficiency and background, taking advantage of a<br />

high efficiency (and therefore sample throughput) and providing<br />

the lowest possible MDA.<br />

Counting Window. Due to the presence of many substances<br />

in the digestate, which may be partially soluble in the scintillator,<br />

the sample may show different levels of quenching, which also<br />

affect the position of the peaks. In order to compensate for the<br />

peak shift, we opted to define counting windows with a fixed width<br />

but a variable position in the spectrum, manually set to comprise<br />

the peaks of interest.<br />

The first strategy was to define a wide counting window of<br />

150 channels. This window allowed to include, irrespective of the<br />

quenching level (quantified through the SQP(E) parameter), the<br />

three R peaks (Figure 1). The main advantage of this strategy is<br />

that counting efficiency is large, up to the maximum theoretical<br />

level of 300%, and above 200% for the real samples commonly<br />

analyzed. However, as the spectrum resolution does not allow us<br />

to distinguish other R impurities present in the sample and some<br />

� interferences, this method might lead to slight activity<br />

overestimation.<br />

In order to minimize the effect of background from impurities<br />

and interferences, the second strategy was to define a 50 channel<br />

wide counting window to only comprise the higher energy R peak<br />

( 214 Po), which stands in an area of much smaller background<br />

and interference, 15 and is easily identified. However, in this<br />

case the maximum theoretical efficiency is only 100%.<br />

Calibration. Quenching is the only parameter affecting the<br />

counting configuration that is sample-dependent and cannot be<br />

easily controlled. Therefore, several 226 Ra standard solutions were<br />

quenched with CCl4 to values covering the expected SQP(E)<br />

6850 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 5. Effect of quenching (SQP(E)) on R counting efficiency.<br />

Figure 6. Time stability test for sample quenching (SQP(E)).<br />

value in environmental samples (Figure 5). For both counting<br />

windows, efficiency is close to zero for SQP(E) < 575. Therefore,<br />

the calibration curves, obtained by regression analysis between<br />

the counting efficiency (y) and the SQP(E) parameter (x), were<br />

150 channels:<br />

y ) (0.74 ( 0.06)x - (435 ( 51) and r ) 0.95<br />

50 channels:<br />

y ) (0.21 ( 0.02)x - (119 ( 19) and r ) 0.91<br />

Usually, LSC techniques involve the measurement of a batch<br />

of samples in cycles. As the typical activities expected in this type<br />

of work are low (of the order of 10 mBq), long-term counting, of<br />

the order of days for a single batch, is required. We successfully<br />

checked the stability of the scintillator through the monitoring of<br />

SQP(E) for a quenched tracer with an activity of 10 mBq during<br />

4 days (Figure 6).<br />

Detection Limits. Each counting configuration (including<br />

counting window and quenching) shows different detection limits.<br />

In Table 1, we show the counting properties for a standard solution<br />

quenched to SQP(E) ) 850, which is an average value of sediment<br />

samples analyzed in our laboratory. The combination of the<br />

indirect detection of 226 Ra by 222 Rn emanation, the high resulting<br />

efficiency, and the ultralow background of Quantulus 1220<br />

result in an extremely low MDA (0.29 Bq kg -1 ) for samples as<br />

low as 250 mg, a typical amount used when analyzing 210 Pb by<br />

R spectrometry. 10 Although both MDA are very low, and<br />

therefore any of the windows can be chosen, the larger<br />

efficiency (and therefore precision of measurements for the<br />

same counting time) favors the use of the largest window,<br />

which is the recommended value in this work. From a


Table 1. Counting Properties and Detection Limits for a 226 Ra Standard Solution a<br />

window (channels) efficiency (%) background (cpm) b<br />

spectrometric point of view, a smaller window and higher R<br />

energy minimize the possibility of interferences with other<br />

emissions and should be favored when expected sample activity<br />

or desired analytical throughput are not compromised.<br />

DISCUSSION<br />

Although this technique is well-suited to easily determine 226 Ra<br />

in small samples from most environmental matrixes, the system<br />

has been designed and optimized for small sediment samples,<br />

subsequent to 210 Pb determination through 210 Po in equilibrium.<br />

Proposed Method. We recommend operating in batches of<br />

about 12 samples, where both a reference material and a blank<br />

are processed and measured at the same time. The solution from<br />

which the Po source was deposited (about 15 mL) is recovered<br />

and, in order to destroy ascorbic acid and its degradation<br />

compounds, heated under reflux with 5 mL of concentrated HNO3<br />

for half an hour or until a clear solution is obtained. This<br />

solution is then evaporated to incipient dryness, and small<br />

amounts of 0.5 M HCl are added to dissolve the residue. This<br />

operation is carried out two more times to ensure the elimination<br />

of HNO3, and the total volume is then adjusted to 10 mL<br />

in a 20 mL low-diffusion PE counting vial. Then 10 mL of<br />

OptiScint HiSafe cocktail is carefully added by avoiding disturbing<br />

the interface formed by the two immiscible liquids in the<br />

vial (Figure 7).<br />

The mixture is kept for 3 weeks in a dark temperaturecontrolled<br />

area, in order to wait for radioactive equilibrium and<br />

to minimize chemiluminescence, and is counted by liquid scintillation<br />

by using the following conditions:<br />

• Counting time: all samples of the batch are sequentially<br />

counted for 1 h each, during a minimum of 30 cycles.<br />

• Counting windows are approximately set to channels 750-900<br />

and 850-900. However, best results are obtained by summing<br />

all spectra and visually adjusting the windows manually to their<br />

optimum value.<br />

• Counting mode is set to discriminate R and other pulses,<br />

with PSA ) 145.<br />

Counting conditions and calibration are equipment-sensitive,<br />

and therefore, the quenching calibration must be performed for<br />

each instrument. We also recommend carrying out some tests<br />

with 226 Ra tracer solutions to confirm that the chosen PSA value<br />

is correct for each instrument.<br />

The counting information needed to calculate the activity are<br />

sample count rate, background count rate, and quenching parameter<br />

(such as SQP(E) in Quantulus 1220). From the quenching<br />

value the efficiency can be obtained (Figure 5), and the sample<br />

activity finally calculated.<br />

Low-Activity Test. We tested the linearity and the low-activity<br />

response of the detector in the chosen configuration by measuring<br />

a set of 226 Ra unquenched tracer solutions (0.5 M HCl) with<br />

Ld (cpm) MDA (Bq kg -1 ) c<br />

710-860 173 ± 12 0.0070 ± 0.0020 0.0076 0.29<br />

810-860 54 ± 4 0.0024 ± 0.0012 0.26 0.59<br />

a Activity ) 0.01 Bq. Quenched to SQP(E) ) 850. b Counting time of the background was 55 h. c MDA was calculated by assuming a typical<br />

sediment mass of 250 mg.<br />

Figure 7. Proposed procedure to measure 210 Po and 226 Ra in<br />

sediment samples.<br />

Table 2. Count Rates When Measuring 226 Ra<br />

Unquenched Standard Solutions by LSC<br />

sample activity (mBq) net count rate (cpm)<br />

0.017<br />

0.12<br />


activities ranging over 2 orders of magnitude, all yielding satisfactory<br />

results (Table 3).<br />

The IAEA-384, a Fangataufa Lagoon sediment, is a reference<br />

material designed for the determination of anthropogenic and<br />

natural radionuclides in sediment. 27 This material is certified for<br />

10 radionuclides, including 210Pb ( 210Po in equilibrium), and<br />

information values are given for 10 radionuclides, including<br />

226 226 -1 Ra. The Ra activity ranges from 2.0 to 2.9 Bq kg (95%<br />

confidence interval) with a median value of 2.4 Bq kg-1 , which<br />

is the low range of 226Ra activities in sediments. Dry sediment<br />

aliquots of 500 mg were analyzed by using the proposed<br />

method (N ) 15). Spectra were analyzed by using both a wide<br />

and a narrow counting window, and results are shown in Figure<br />

9. The median of both distributions is 2.2 Bq kg-1 , and means<br />

were 2.3 Bq kg-1 (50 channels) and 2.5 Bq kg-1 (150 channels),<br />

very close to the reference material reported value and within<br />

its 95% confidence interval.<br />

The IAEA-385, a certified reference material for radionuclides<br />

in an Irish Sea sediment, is intended to be used, among other<br />

purposes, for the development and validation of radiometric and<br />

mass spectrometry analytical methods. 28 Its certified median value,<br />

22.7 Bq kg-1 , is typical of most soils and sediments. The median<br />

value obtained with the proposed method (N ) 15) is 23.5 Bq<br />

kg-1 Figure 8. Calibration curve of unquenched<br />

, which, although slightly higher than the upper limit of<br />

the certified confidence interval, represents a deviation of only<br />

7.3%.<br />

226Ra. (27) Povinec, P. P.; Pham, M. K.; Sanchez-Cabeza, J. A.; Barci-Funel, G.;<br />

Bojanawski, R.; Boshkova, T.; Burnett, W.; Carvalho, F.; Chapeyron, B.;<br />

Cunha, I. L.; Dahlgaard, H.; Galabov, N.; Fifield, L.; Gaustaud, J.; Geering,<br />

J.-J.; Gomez, I. F.; Green, N.; Hamilton, T.; Ibanez, F. L.; Ibn Majah, M.;<br />

John, M.; Kanisch, G.; Kenna, T. C.; Kloster, M.; Korun, M.; Liong Wee<br />

Kwong, L.; La Rosa, J.; Lee, S.-H.; Levy-Plaomo, I.; Malatova, M.; Maruo,<br />

Y.; Michell, P.; Murciano, I. V.; Nelson, R.; Nouredine, A.; Oh, J.-S.; Oregioni,<br />

B.; Petit, G.; Pettersson, H. B. L.; Reineking, A.; Smedley, P. A.; Suckow,<br />

A.; Struijs, T.; Voors, P. I.; Yoshimiza, K.; Wyse, E. J. Radioanal. Nucl. Chem.<br />

2007, 273, 383–393.<br />

(28) Pham, M. K.; Sanchez-Cabeza, J. A.; Povinec, P. P.; Andor, K.; Arnold, D.;<br />

Benmansour, M.; Bikit, I.; Carvalho, F. P.; Dimitrova, K.; Edrev, Z. H.;<br />

Engeler, C.; Fouche, F. J.; Garcia-Orellana, J.; Gascó, C.; Gastaud, J.; Gudelis,<br />

A.; Hancock, G.; Holm, E.; Legarda, F.; Ikäheimonen, T. K.; Ilchmann, C.;<br />

Jenkinson, A. V.; Kanisch, G.; Kis-Benedek, G.; Kleinschmidt, R.; Koukouliou,<br />

V.; Kuhar, B.; LaRosa, J.; Lee, S.-H.; LePetit, G.; Levy-Palomo, I.;<br />

Liong Wee Kwong, L.; Llauradó, M.; Maringer, F. J.; Meyer, M.; Michalik,<br />

B.; Michel, H.; Nies, H.; Nour, S.; Oh, J.-S.; Oregioni, B.; Palomares, J.;<br />

Pantelic, G.; Pfitzner, J.; Pilviok, R.; Puskeiler, L.; Satake, H.; Schikowski,<br />

J.; Vitorovic, G.; Woodhead, D.; Wyse, E. Appl. Radiat. Isot. 2008, 66, 1711–<br />

1717.<br />

6852 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

The IAEA-313, a stream sediment from Indonesia, 29 is a<br />

reference material designed for the analysis of 226Ra, U, and Th<br />

in geological samples. Its median 226Ra activity, 343 Bq kg-1 ,<br />

is about 1 order of magnitude higher than typical sediment<br />

activities. The median value (N ) 5) is within the reported 95%<br />

confidence interval.<br />

The observed deviations are typical of the precision of<br />

environmental materials, and in overall, we concluded that the<br />

proposed method is adequate to analyze 226Ra in sediment<br />

samples.<br />

Calibration Stability. With each sample batch, we have<br />

analyzed the IAEA-384 reference material for quality control<br />

purposes. Results of all analyses are shown in Figure 10. The mean<br />

value, 2.3 Bq kg-1 , lies within the reference material 95%<br />

confidence interval (Table 3), and the time stability of the activity<br />

shows that the system’s calibration is stable for times spanning<br />

at least during the measurement period, namely, one year.<br />

Although this needs to be checked for each instrument in a<br />

continuous manner, we do not anticipate the need of recalibration<br />

of Quantulus 1220 for periods shorter than a year.<br />

210 226 Pb and Ra in a Sediment Core. The method is<br />

designed to provide fast and reliable results of 226Ra when 210Pb dating sediment cores through the analysis of 210Po in equilibrium.<br />

The advantage of LSC over γ spectrometry is that<br />

results for a large number of samples, typical of 210Pb dating<br />

experiments, can be obtained within a month after sample<br />

digestion. We provide here an example of the proposed<br />

strategy.<br />

The DYFAMED station (Ligurian Sea, 43°25′ N; 7°52′ E) has<br />

been the subject of multidisciplinary research since 1987, including<br />

the study of atmospheric deposition of metals and their association<br />

with marine particles in the water column. Martín et al. 30 studied<br />

the concentrations and fluxes of trace metals in a sediment core<br />

collected from 2300 m water depth at the sea floor beneath the<br />

DYFAMED site. In Figure 11, we show the 210Pb and 226 Figure 9. Box-and-whisker plots of<br />

Ra<br />

profiles obtained by R spectrometry and LSC, showing that<br />

226Ra activities in the reference<br />

material IAEA-384 (N ) 15). The two dots in the 50 channels counting<br />

window were statistically identified as outliers.<br />

(29) Strachnov, V.; Valkovic, V.; Zeisler, R.; Dekner, R. Report on the Worldwide<br />

Intercomparison Exercise IAEA-314: 226 Ra, Th and U in Stream Sediment;<br />

International Atomic Energy Agency (IAEA/AL/038): Vienna, Austria, 1991.<br />

(30) Martin, J.; Sanchez-Cabeza, J. A.; Eriksson, M.; Miquel, J. C. Mar. Pollut.<br />

Bull. 2009, 59, 146–153.


Table 3. Analysis of 226 Ra in Sediment Reference Materials (RM)<br />

reference material no. of analyses method median (Bq kg-1 ) RM median (Bq kg-1 ) RM 95% confidence interval deviation<br />

IAEA-384 15 2.2 2.4 2.0–2.9 -8.3<br />

IAEA-385 15 23.5 21.9 21.6–22.4 +7.3<br />

IAEA-313 5 372 343 307–379 +8.5<br />

Figure 10. Time stability test for 226 Ra concentrations in the IAEA-<br />

384 reference material.<br />

226Ra reaches equilibrium in the core bottom sections. An<br />

important finding was that, although a constant 226Ra is<br />

commonly estimated from core bottom sections, where 210Pb and 226Ra are assumed to be in equilibrium, this is not the case<br />

in the DYFAMED core, where 226Ra ranged from 22.6 to 41.7<br />

Bq kg-1 . This reinforces the need for paired 210Pb/ 226 Figure 11.<br />

Ra<br />

210Pb and 226Ra profiles in a sediment core from the<br />

DYFAMED site (Ligurian Sea).<br />

measurements and methodologies to accomplish it accurately,<br />

as does the technique described in this paper. This may have<br />

important implications on the dating results for this particular<br />

sediment core, which will be discussed elsewhere.<br />

CONCLUSIONS<br />

The accurate 210 Pb dating of sediment cores requires the<br />

determination of 226 Ra in all sections. Although this can be done<br />

by γ spectrometry, the large sample size required (>5 g dry<br />

weight in well detectors, >20 g in coaxial detectors) might be<br />

impossible to obtain when sediments are used for the analysis<br />

of multiple magnitudes (such as grain size, elemental composition,<br />

trace metals, organic substances, biomarkers, ...) and<br />

counting times (typically >2 days per sample) might be a<br />

limiting factor for many laboratories. The proposed strategy<br />

in this work is the determination of 210 Pb through 210 Po in<br />

equilibrium by R spectrometry 10 followed by 226 Ra by LSC<br />

without any further radiochemical processing.<br />

We optimized the counting parameters for an ultralow background<br />

scintillation system with R-� separation capabilities<br />

(Quantulus 1220, Wallac) and propose the use of a PSA parameter<br />

of 145. The system was calibrated with a series of quenched 226 Ra<br />

standard solutions. For a typical sediment sample quenching<br />

(SQP(E) ) 850), the efficiency was (173 ± 12)% in the wide<br />

energy counting window (150 channels) due to the simultaneous<br />

counting of three radionuclides in equilibrium ( 222 Rn,<br />

218 Po, and 214 Po). When analyzing 250 mg dw sediment samples,<br />

the MDA was as low as 0.29 Bq kg -1 , which is about 2 orders<br />

of magnitude lower than typical sediment concentrations,<br />

showing the usefulness of the technique for many other<br />

environmental applications. The method was validated with<br />

three reference materials spanning 3 orders of magnitude of<br />

concentration. The proposed method can greatly improve the<br />

reliability of 210 Pb chronologies of sediment cores, and can also<br />

be tested for 226 Ra/ 210 Pb dating of carbonates such as corals<br />

and speleothems.<br />

ACKNOWLEDGMENT<br />

The authors thank Mr. Jacobo Martín (IAEA) for providing<br />

samples of a sediment core from the DYFAMED site and helpful<br />

comments. The IAEA is grateful for the support provided to its<br />

Marine Environment Laboratories by the Government of the<br />

Principality of Monaco.<br />

Received for review March 31, 2010. Accepted June 30,<br />

2010.<br />

AC1008332<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6853


Anal. Chem. 2010, 82, 6854–6861<br />

Improved Sensitivity of DNA Microarrays Using<br />

Photonic Crystal Enhanced Fluorescence<br />

Patrick C. Mathias, †,‡ Sarah I. Jones, § Hsin-Yu Wu, ‡,| Fuchyi Yang, ‡,| Nikhil Ganesh, ‡,⊥<br />

Delkin O. Gonzalez, § German Bollero, § Lila O. Vodkin, § and Brian T. Cunningham* ,†,‡,|<br />

Department of Bioengineering, 1304 W. Springfield Ave., Micro and Nanotechnology Laboratory, 208 N. Wright St.,<br />

Department of Crop Sciences, 1102 S. Goodwin Ave., Department of Electrical and Computer Engineering,<br />

1406 W. Green St., and Department of Materials Science and Engineering, 1304 W. Green St., University of Illinois at<br />

Urbana-Champaign, Urbana, Illinois 61801<br />

DNA microarrays are used to profile changes in gene<br />

expression between samples in a high-throughput manner,<br />

but measurements of genes with low expression<br />

levels can be problematic with standard microarray substrates.<br />

In this work, we expand the detection capabilities<br />

of a standard microarray experiment using a photonic<br />

crystal (PC) surface that enhances fluorescence observed<br />

from microarray spots. This PC is inexpensively and<br />

uniformly fabricated using a nanoreplica molding technique,<br />

with very little variation in its optical properties<br />

within- and between-devices. By using standard protocols<br />

to process glass microarray substrates in parallel with<br />

PCs, we evaluated the impact of this substrate on a onecolor<br />

microarray experiment comparing gene expression<br />

in two developmental stages of Glycine max. The PCs<br />

enhanced the signal-to-noise ratio observed from microarray<br />

spots by 1 order of magnitude, significantly increasing<br />

the number of genes detected above substrate fluorescence<br />

noise. PC substrates more than double the number<br />

of genes classified as differentially expressed, detecting<br />

changes in expression even for low expression genes. This<br />

approach increases the dynamic range of a surface-bound<br />

fluorescence-based assay to reliably quantify small quantities<br />

of DNA that would be impossible with standard<br />

substrates.<br />

The DNA microarray is a valuable tool for high-throughput<br />

quantification of gene expression, allowing a large number of<br />

candidate genes to be examined for differential expression<br />

simultaneously without extensive prior knowledge of gene<br />

functions. Eukaryotic gene expression is typically characterized<br />

by a large number of genes expressed at very low levels and a<br />

decreasing number of genes expressed at high levels. 1,2 Often<br />

the noise present in DNA microarray experiments is high<br />

enough that only a small fraction of genes assayed can be<br />

* To whom correspondence should be addressed. Phone: 217-265-6291.<br />

E-mail: bcunning@illinois.edu.<br />

† Department of Bioengineering.<br />

‡ Micro and Nanotechnology Laboratory.<br />

§ Department of Crop Sciences.<br />

| Department of Electrical and Computer Engineering.<br />

⊥ Department of Materials Science and Engineering.<br />

(1) Kuznetsov, V. A.; Knott, G. D.; Bonner, R. F. Genetics 2002, 161, 1321–<br />

1332.<br />

6854 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

detected by fluorescence measurements. While sample variation<br />

and nonspecific binding play an important role in this experimental<br />

noise, microarray substrate fluorescence can contribute<br />

to noise as well. This may explain the poor performance of<br />

microarrays in detecting genes with low expression levels<br />

relative to other methods. 3,4 To overcome the difficulties of<br />

quantifying the abundance of low expression genes, substrates<br />

that enhance the fluorescence observed from microarray spots<br />

can be used to achieve better assay performance.<br />

Researchers have utilized various nanostructured metal substratres<br />

to achieve increased intensity from common microarray<br />

dyes, with signal enhancements ranging from 1 to 2 orders of<br />

magnitude. These methods include the growth of metal island<br />

films on a substrate 5,6 or the deposition of nanoparticles fabricated<br />

by spray pyrolysis 7 to produce optical resonances to which<br />

microarray dye excitation and/or emission can couple. However,<br />

the practical impact of these enhancement methods is unclear,<br />

because previous work has not characterized the signal-to-noise<br />

ratio (SNR) enhancement for a large number of spots over many<br />

substrates in a conventional gene expression microarray experiment.<br />

One potential obstacle in achieving this result is the need<br />

for an inexpensive nanoscale fabrication method achieving highthroughput<br />

and good uniformity over large areas; previous reports<br />

of microarray dye enhancement have not utilized photolithography<br />

to generate consistent patterns and thus are subject to a random<br />

arrangement of structures. Another potential obstacle is integration<br />

of these substrates with the existing commercial equipment<br />

used to fabricate and scan DNA microarrays, as previous literature<br />

has not explicitly demonstrated nanostructures over areas as large<br />

as conventional microscope slides. We addressed these issues by<br />

designing a nanostructured photonic crystal (PC) substrate<br />

capable of enhancing Cyanine-5 (Cy-5) fluorescence in a com-<br />

(2) Ueda, H. R.; Hayashi, S.; Matsuyama, S.; Yomo, T.; Hashimoto, S.; Kay,<br />

S. A.; Hogenesch, J. B.; Iino, M. Proc. Natl. Acad. Sci. U.S.A. 2004, 101,<br />

3765–3769.<br />

(3) Kuo, W. P.; Liu, F.; Trimarchi, J.; Punzo, C.; Lombardi, M.; et al. Nat.<br />

Biotechnol. 2006, 24, 832–840.<br />

(4) t Hoen, P. A. C.; Ariyurek, Y.; Thygesen, H. H.; Vreugdenhil, E.; Vossen,<br />

R. H. A. M.; de Menezes, R. X.; Boer, J. M.; van Ommen, G.-J. B.; den<br />

Dunnen, J. T. Nucleic Acids Res. 2008, 36, e141.<br />

(5) Sabanyagam, C. R.; Lakowicz, J. R. Nucleic Acids Res. 2007, 35, e13.<br />

(6) Moal, E. L.; Leveque-Fort, S.; Potier, M.-C.; Fort, E. Nanotechnology 2009,<br />

20, 225502.<br />

(7) Guo, S.-H.; Tsai, S.-J.; Kan, H.-C.; Tsai, D.-H.; Zachariah, M. R.; Phaneuf,<br />

R. J. Adv. Mater. 2008, 20, 1424–1428.<br />

10.1021/ac100841d © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/16/2010


mercial microarray scanner and fabricating it by nanoreplica<br />

molding to fit standard microscope slides.<br />

The PC substrates are composed of a subwavelength, periodic<br />

SiO2 surface structure coated with a high refractive index<br />

dielectric layer of TiO2, creating a periodic modulation in<br />

refractive index along the device surface. The periodic modulation<br />

gives rise to optical resonances 8 that can be used to achieve<br />

fluorescence enhancement. These resonances can be used to<br />

generate strong optical near-fields at the device surface when<br />

spectrally aligned to the excitation wavelength 9 and to spatially<br />

alter the fluorescence emission pattern to maximize light<br />

collection. 10 The overall effect of these phenomena is to amplify<br />

the fluorescent signal from molecules within approximately 100<br />

nm of the PC surface. While the first demonstrations of PC<br />

enhanced fluorescence for microarrays required expensive<br />

lithographic procedures for each device and yielded modest<br />

enhancement factors of approximately 6× signal enhancement<br />

in a commercial scanner, 11,12 inexpensive and uniform fabrication<br />

over large areas in a nanoreplica molding process currently used<br />

to make commercial label-free biosensors 13 has since been<br />

employed to make these structures.<br />

Recently, PCs have been engineered by our group to enhance<br />

the common microarray dye Cyanine-5 (Cy-5) by more than 1<br />

order of magnitude when scanned in a commercial microarray<br />

scanner. 14 This work details the application of this PC design to<br />

a microarray experiment assessing differential expression between<br />

Glycine max cotyledons and trifoliates, which represent tissues<br />

from two distinct developmental stages in the soybean plant.<br />

Multiple PCs exhibiting highly uniform optical characteristics over<br />

the area of entire microscope slides were fabricated by nanoreplica<br />

molding. These PCs were processed using published protocols<br />

in parallel with commercial microarray substrates. By enhancing<br />

fluorescence, a larger number of genes can be detected above<br />

noise on the PC compared with glass substrates. This effect more<br />

than doubles the number of genes identified as differentially<br />

expressed between the trifoliate and cotyledon tissues, demonstrating<br />

that enhanced fluorescence offers practical benefit to a<br />

DNA microarray experiment.<br />

MATERIALS AND METHODS<br />

Photonic Crystal Fabrication and Characterization. The<br />

PCs used for this work were designed by simulation software<br />

employing Rigorous Coupled-Wave Analysis (DiffractMOD, RSoft<br />

Design Group, Inc.) to align optical resonances to the excitation<br />

(632.8 nm) and emission wavelengths (670-710 nm) of Cy-5. As<br />

described in previous work, 14 the period of the structure, grating<br />

depth, and thickness of the high refractive index TiO2 layer were<br />

(8) Rosenblatt, D.; Sharon, A.; Friesem, A. A. IEEE J. Quantum Electron. 1997,<br />

33, 2038–2059.<br />

(9) Ganesh, N.; Mathias, P. C.; Zhang, W.; Cunningham, B. T. J. Appl. Phys.<br />

2008, 103, 083104.<br />

(10) Ganesh, N.; Block, I. D.; Mathias, P. C.; Zhang, W.; Chow, E.; Malyarchuk,<br />

V.; Cunningham, B. T. Opt. Express 2008, 16, 21626–21640.<br />

(11) Neuschafer, D.; Budach, W.; Wanke, C.; Chibout, S.-D. Biosensors Bioelectron.<br />

2003, 18, 489–497.<br />

(12) Budach, W.; Neuschafer, D.; Wanke, C.; Chibout, S.-D. Anal. Chem. 2003,<br />

75, 2571–2577.<br />

(13) Cunningham, B.; Lin, B.; Qiu, J.; Li, P.; Pepper, J.; Hugh, B. Sensors Actuators<br />

B 2002, 85, 219–226.<br />

(14) Mathias, P. C.; Wu, H.-Y.; Cunningham, B. T. Appl. Phys. Lett. 2009, 95,<br />

021111.<br />

manipulated given the known refractive indices of PC materials<br />

to achieve optical resonances overlapping both excitation and<br />

emission wavelength ranges. These PCs were then fabricated<br />

by nanoreplica molding 13 to create six distinct devices for<br />

microarray experiments. The silicon “master” for the molding<br />

process consisted of a 360 nm period one-dimensional grating<br />

structure with a 60 nm grating depth and 50% duty cycle,<br />

patterned on an 8 in. silicon wafer by deep-UV lithography.<br />

After immersion in 2% dichlorodimethylsilane (PlusOne Repel-<br />

Silane ES, GE Healthcare) to promote clean release, a UVcurable<br />

liquid polymer (Gelest, Inc.) with index of refraction<br />

npolymer ) 1.46 was dispensed on a sheet of polyethylene<br />

terephthalate (PET), and the grating pattern was transferred<br />

with a roller. After curing the polymer under a high-intensity<br />

ultraviolet lamp (Xenon) for 90 s through the transparent PET<br />

sheet, 300 nm of SiO2 (nSiO2 ) 1.46) and 160 nm of TiO2 (nTiO2<br />

) 2.35) were added to the grating structure by sputter coating.<br />

The completed PCs were cut into 1 in. × 3 in. sections and<br />

adhered to glass microscope slides with an optically clear<br />

adhesive (3M).<br />

The PCs were initially profiled for surface characteristics by<br />

atomic force microscopy (Dimension 3000, Digital Instruments)<br />

to compare the actual dimensions with the PC design. The PCs<br />

were then optically characterized by passing broadband light in<br />

the visible spectrum from a tungsten halogen lamp (Ocean Optics)<br />

through a polarizer and collimator before transmission through<br />

the device, which was aligned on a rotational stage to be<br />

perpendicular to the direction of incident light. 15 PCs were<br />

illuminated with both transverse magnetic (incident electric field<br />

perpendicular to grating lines) and transverse electric (incident<br />

electric field parallel to grating lines) polarizations to characterize<br />

the two distinct resonances. Light transmitted through the PCs<br />

was collected by an optical fiber and measured using a UV-visible<br />

light spectrometer (Ocean Optics).<br />

Slide Preparation. PCs were cleaned by O2 plasma treatment<br />

and incubated with 3-glycidoxypropyltrimethoxysilane at 185<br />

mTorr overnight for surface functionalization. The control slides<br />

for microarray experiments were commercially available silanized<br />

glass slides (Corning GAPS II). Oligonucleotides were<br />

printed on the slides using a Genetix QArray 2 robot. A set of<br />

previously annotated 192 70-mer oligonucleotides derived from<br />

publicly available soybean EST and mRNA sequences was<br />

spotted on the slides with 40 repeats per sequence per slide.<br />

These 192 oligonucleotides are a subset of a larger 19 200 set<br />

of oligonucleotides detailed in previous work. 16 Each of six<br />

spotted PCs was matched with a spotted glass control slide to<br />

receive identical experimental treatments.<br />

Microarray Sample Preparation and Hybridization. Sample<br />

RNA was extracted using previously published protocols. 16<br />

Cotyledon RNA was extracted from freeze-dried soybean cultivar<br />

Williams seeds with fresh weight between 100-200 mg. Cotyledon<br />

RNA was purified using a Qiagen RNeasy kit. Trifoliate RNA was<br />

extracted from freeze-dried rolled-up trifoliates of soybean cultivar<br />

Williams from leaves between 0.5 and 1.5 in. in length. Sample<br />

RNA was labeled with Cy-5 by reverse transcription. Three<br />

replicate slides were hybridized for each of the two tissue samples,<br />

(15) Yang, F.; Yen, G.; Cunningham, B. T. Opt. Express 2010, 18, 11846–11858.<br />

(16) Gonzalez, D. O.; Vodkin, L. O. BMC Genomics 2007, 8, 468.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6855


with an identical number of glass slides processed in parallel with<br />

the PCs. Slides were blocked prior to hybridization with bovine<br />

serum albumin to prevent nonspecific binding, hybridized at 42<br />

°C overnight, and washed as described previously. 16<br />

Data Collection. Slides were scanned with a Tecan LS<br />

Reloaded scanner with a transverse magnetic polarized laser (λ<br />

) 632.8 nm) at normal incidence and an emission filter spanning<br />

670-710 nm. All slides were scanned at 10 µm resolution. Initial<br />

scans were at equal photomultiplier tube (PMT) gain to compare<br />

fluorescence intensities at equal measurement conditions, but<br />

afterward, gains were adjusted for each slide such that spots with<br />

the largest fluorescence intensities did not saturate the scanner<br />

PMT. Fluorescence images were analyzed using GenePix Pro 6.0<br />

to compute spot and local background intensities as well as their<br />

standard deviations for each spot.<br />

Signal-to-Noise Ratio Analysis. Fluorescence data was<br />

analyzed to calculate signal-to-noise ratios (SNRs) for each spot<br />

at identical scan conditions, where the SNR is the local backgroundsubtracted<br />

spot intensity divided by the standard deviation of the<br />

local background pixels. For each of the six PCs and six glass<br />

slides, within-slide repeats (40 per probe sequence) were averaged<br />

to generate a SNR value for each of the 192 sequences probed in<br />

the experiment. A SNR enhancement factor was calculated by<br />

dividing the PC SNR value for each gene by the glass SNR value<br />

for the same gene. The proportion of detected genes was<br />

determined by calculating the percentage of genes on each slide<br />

with a SNR > 3.<br />

Differential Expression Analysis. Expression data was<br />

analyzed using the Linear Models for Microarray Data (LIMMA)<br />

package in R. Data was background corrected using the normalized<br />

plus exponential convolution model with an offset of one. 17<br />

Quantile normalization was used to normalize between arrays. Logtransformed<br />

Cyanine-5 intensities were condensed by averaging<br />

within-slide repeats and then fit to a linear model. Empirical Bayes<br />

moderated t-statistics were calculated to assess differential expression<br />

between the trifoliate and cotyledon samples, with p-values<br />

adjusted by the Benjamini and Hochberg method to control the<br />

false discovery rate. 18 The significance level for testing was set<br />

to R)0.05.<br />

High-Throughput RNA Sequencing and Analysis. The<br />

mRNAs were also subjected to high-throughput sequencing (RNAseq)<br />

performed at the Keck Center of the University of Illinois<br />

using the Illumina Genome Analyzer II resulting in 10-18 million<br />

total reads for leaf trifoliate and the immature cotyledon mRNA<br />

samples, respectively. After processing, the RNA-seq reads are<br />

all 70 bases in length. The sequence reads were aligned using<br />

Bowtie 19 to the approximately 78 700 predicted Glyma gene<br />

models available at Phytozome 5.0 (http://www.phytozome.net)<br />

for the recently sequenced soybean genome. 20 The Bowtie<br />

parameters allowed three mismatches to each Glyma model and<br />

allowed up to 25 gene model matches to detect repetitive gene<br />

models. Normalization of RNA-seq data as RPKM (reads per<br />

(17) Ritchie, M. E.; Silver, J.; Oshlack, A.; Holmes, M.; Diyagama, D.; Holloway,<br />

A.; Smyth, G. K. Bioinformatics 2007, 23, 2700–2707.<br />

(18) Benjamini, Y.; Hochberg, Y. J. R. Stat. Soc. Series B (Methodological) 1995,<br />

57, 289–300.<br />

(19) Langmead, B.; Trapnell, C.; Pop, M.; Salzberg, S. L. Genome Biol. 2009,<br />

10, R25.<br />

(20) Schmutz, J.; Cannon, S. B.; Schlueter, J.; Ma, J.; Mitros, T.; et al. Nature<br />

2010, 463, 178–183.<br />

6856 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 1. (a) Schematic of PC design dimensions. (b) Atomic force<br />

micrograph of completed PC structure (after TiO2 deposition), with a<br />

measured period of 366 nm and height of 50 nm.<br />

kilobase of gene model per million mapped reads) was calculated<br />

as shown in previous work. 21 Assignment of the 70-mer oligos to<br />

Glyma gene models was also performed using Bowtie with the<br />

same parameters. Most of the 70-mer oligos on the arrays matched<br />

only one or a few Glyma models.<br />

RESULTS<br />

Characterization of Photonic Crystal Substrates. A schematic<br />

of the nanoreplica-molded one-dimensional PC design<br />

optimized in previous work 14 to enhance fluorescence from Cy-5<br />

(with period Λ ) 360 nm and grating step height h ) 60 nm)<br />

appears in Figure 1a. A representative atomic force micrograph<br />

detailing the surface structure appears in Figure 1b, with a<br />

measured period of Λ ) 366 nm and a measured height of h )<br />

50 nm showing good agreement with the expected dimensions.<br />

Enhancement of Cy-5 was achieved by aligning a narrow PC<br />

resonance (full-width at half-maximum, fwhm ) 4 nm) with the<br />

laser excitation wavelength of 632.8 nm and engineering a second<br />

broad resonance (fwhm ) 20 nm) to overlap the emission filter<br />

wavelengths of 670-710 nm (Figure 2). A more narrow excitation<br />

resonance increases the magnitude of the enhanced optical near<br />

(21) Mortzazvi, A.; Williams, B. A.; McCue, K.; Schaeffer, L.; Wold, B. Nat.<br />

Methods 2008, 5, 621–628.


Figure 2. Optical transmission measurements from all six PCs (each<br />

represented by a colored solid line) used in this study, obtained by<br />

illuminating the devices with polarized, collimated white light. Resonances<br />

with narrow spectral features are excited when the PCs are<br />

illuminated with transverse magnetic polarized light and overlap the<br />

excitation wavelength of 632.8 nm (dotted line). Resonances with<br />

broader spectral features are excited when the PCs are illuminated<br />

with transverse electric polarized light and overlap the emission filter<br />

wavelengths of 670-710 nm (dotted box).<br />

fields at the device surface, 22 while the broad extraction resonance<br />

maximizes the spectrum of emitted light redirected toward the<br />

detection optics. 10 Good spectral uniformity of the narrow excitation<br />

resonance over large areas of the PC is required to ensure<br />

precise overlap of the resonance with a narrowband excitation<br />

source regardless of the location of a microarray spot on the PC.<br />

The nanoreplica molding fabrication process achieves this uniformity<br />

throughout individual microscope slide-sized PCs, with a<br />

maximum observed resonance wavelength standard deviation of<br />

σwithin-PC ) 0.239 nm over 6 distinct PCs. Figure 2 illustrates<br />

that excellent between-device uniformity is achieved as well; the<br />

standard deviation in mean resonance wavelength between the 6<br />

PCs used in this work is σbetween-PC ) 0.691 nm, making both<br />

the within-device variation and between-device variation in<br />

resonance wavelength significantly smaller than the spectral<br />

width of the resonance.<br />

Signal-to-Noise Ratio Analysis. After pairing each PC with<br />

a control glass slide and printing a set of 192 oligonucleotides<br />

and negative controls on the slides, 16 a one-color DNA microarray<br />

protocol was simultaneously run on each glass-PC pair. Three<br />

pairs of slides were hybridized with Cy-5 labeled cotyledon RNA<br />

(extracted from seeds), and three pairs were hybridized with Cy-5<br />

labeled trifoliate RNA (extracted from leaves) from Glycine max<br />

cultivar Williams, representing two distinct tissues and stages of<br />

development. After averaging the 40 duplicates of the 192 genes<br />

on each slide, a ratio of averaged PC SNR to averaged glass SNR<br />

was generated for each PC-glass slide pair, resulting in a median<br />

SNR enhancement across all slides of 10.6×. The effect of this<br />

SNR enhancement on the raw fluorescence data is observed in<br />

Figure 3, which shows line profiles of identical probes for a<br />

microarray grid on both a PC and its control. Considerable<br />

enhancement is observed for spots of varying expression levels<br />

(Figure 3c, d), with low expression genes being much easier to<br />

discriminate from noise on the PC.<br />

To explore the relevance of this SNR enhancement on gene<br />

expression measurements, additional fluorescence scans of PCs<br />

and glass slides were performed after gain adjustment to utilize<br />

the full dynamic range of the scanner photomultiplier tube (PMT).<br />

Because the PCs demonstrate fluorescence enhancement, they<br />

were scanned at lower PMT gain values, resulting in lower noise<br />

levels. A detection threshold of SNR ) 3 was applied to determine<br />

the proportion of spots that could be detected on each slide.<br />

Across all cotyledon samples, 25.0% of spots could be detected<br />

on the glass slides, while 46.3% of the spots were detected on the<br />

Figure 3. Fluorescence images at identical gains of a single identical microarray grid on glass (a) and PC (b), with brightness and<br />

contrast adjustment to make the maximum number of spots visible on both images. For comparison of spot intensities, line profiles of<br />

identical locations on the grid for glass and PC are illustrated on the same plots. Lower expression genes appear in (c) and higher<br />

expression genes appear in (d).<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6857


Figure 4. Logarithmic plots of duplicate-averaged SNR values for the 192 probed genes on selected glass-PC slide pairs for each tissue.<br />

Genes are organized in decreasing expression order for each chip, and an SNR detection threshold of 3 appears as the cutoff line in each<br />

graph. SNR expression profiles for cotyledon RNA appear in (a) and (c) for a glass slide and its paired PC, respectively. Trifoliate RNA expression<br />

profiles are plotted in (b) and (d) for a glass slide and its paired PC, respectively. Negative control spots on all slides appeared below the<br />

detection threshold.<br />

PCs. A more dramatic increase in the number of detected spots<br />

was observed across all trifoliate samples, with 14.7% of spots on<br />

the glass slides being detected as compared to 49.0% of the spots<br />

on the PC. The number of genes that could be detected above<br />

noise thus almost doubled for the cotyledon sample and more<br />

than tripled for the trifoliate sample, as illustrated in plots of SNR<br />

for each gene for representative slides in Figure 4. SNR values<br />

(averaged across duplicate spots) are graphed for each gene in<br />

decreasing expression order for a single slide pair hybridized to<br />

a cotyledon sample (Figure 4a, c) and a single slide pair hybridized<br />

to a trifoliate sample (Figure 4b, d). As expected, negative control<br />

spots on both the PCs and the glass slides had SNRs below the<br />

detection threshold.<br />

Differential Expression Analysis. Background-corrected, 17<br />

normalized, log-transformed spot intensities were fit to a linear<br />

model using the LIMMA package in R, and empirical Bayes<br />

moderated t-statistics were calculated to assess differential expression<br />

in the trifoliate sample relative to the cotyledon sample. The<br />

analysis was carried out for glass slides and PCs separately to<br />

allow for comparison between the two substrates. Volcano plots<br />

(simultaneously illustrating the fold change and the adjusted<br />

p-value for each gene across glass slides or PCs) appear in Figure<br />

5. Ideally, the plot should be a v-shape, since the p-value should<br />

decrease as the fold-change increases. However, this relationship<br />

is distorted by variation, since variation in fold-change measurements<br />

smears the curve horizontally and variation between<br />

samples leads to lower p-values (smearing the curve vertically).<br />

Figure 5a and b plot all 7680 spots without averaging of withinslide<br />

repeat spots in order to illustrate more clearly the effect of<br />

the PC on the experimental data. Red circles denote genes that<br />

have an adjusted p-value of less than 0.05 and a greater than 2-fold<br />

change. 1431 spots on the PCs fulfill these criteria, compared with<br />

865 spots on the glass slides. The PC data more tightly conforms<br />

to the expected v-shape, suggesting there is less variation between<br />

within-slide repeats in fold-change values and p-values, particularly<br />

6858 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

for spots with low fold-change value, compared to the glass slide.<br />

This lowered variation is expected to allow for discrimination of<br />

smaller changes in expression during statistical testing.<br />

A similar analysis was carried out after averaging within-slide<br />

repeats, reducing the data set into 192 genes, with Figure 5c and<br />

d illustrating volcano plots for the averaged data set. On the glass<br />

slide, 27 genes fulfilled the criteria of statistically significant<br />

changes at an adjusted p-value less than 0.05 and a greater than<br />

2-fold change, while on the PC, 68 genes fulfilled these criteria.<br />

Importantly, all 27 of the genes fulfilling these criteria on the glass<br />

slide were also identified as differentially expressed on the PCs<br />

(Table S-1 in the Supporting Information). The average measurement<br />

for these 27 genes are similar on both substrates, with<br />

measurements of 2350 counts (of fluorescence intensity) on the<br />

glass slides and 2880 counts on the PCs. However, an additional<br />

41 genes were identified as differentially expressed on the PCs,<br />

suggesting that statistically significant changes in expression that<br />

were overwhelmed by noise on the glass slide could be identified<br />

on the PCs. Genes with a greater than 2-fold change as measured<br />

on the PCs (p < 0.05) but not the glass slides (Table S-2 in the<br />

Supporting Information) had an average expression level of 180<br />

counts on the PCs, demonstrating lower expression levels than<br />

those genes classified as differentially expressed on the glass<br />

slides.<br />

Validation of the microarray data by an independent method<br />

was obtained by high-throughput sequencing with the Illumina<br />

platform yielding 10-18 million total reads for leaf trifoliate and<br />

the immature cotyledon mRNA samples, respectively. High quality<br />

reads of 70 bases in length were mapped to 78 700 soybean gene<br />

models to obtain a quantitative view gene expression. Fold change<br />

values for 5 randomly selected genes from the 41 genes found to<br />

be differentially expressed on the PC but not the glass substrates<br />

(22) Mathias, P. C.; Ganesh, N.; Zhang, W.; Cunningham, B. T. J. Appl. Phys.<br />

2008, 103, 094320.


Figure 5. Volcano plots detailing the relationship between fold-change and inverse p-value to assess differential expression between the<br />

trifoliate and cotyledon samples, with positive fold changes indicating increased trifoliate expression and negative fold changes indicating increased<br />

cotyledon expression. Green vertical lines represent the 2-fold change cutoff, and the yellow horizontal line denotes a p-value cutoff of 0.05.<br />

Genes meeting both thresholds are indicated by red spots. Unaveraged data representing all 7680 spots across all experimental slides (3<br />

replicates per tissue) appear in (a) for the glass slides and (b) for the PCs, with 865 spots differentially expressed on glass slides and 1431<br />

spots on the PCs. Averaging within-slide repeats condensed the data to 192 distinct genes and controls, which appear in (c) for the glass slides<br />

and (d) for the PCs. Of the 192 genes probed, 27 were classified as differentially expressed on the glass slides, while 68 met this classification<br />

on the PCs.<br />

(Table S-2 in the Supporting Information) are plotted in Figure 6,<br />

with a comparison of glass microarray, PC microarray, and<br />

sequencing data. The direction of differential expression of the<br />

genes represented by a majority of probes on the arrays was<br />

confirmed by the transcriptome sequencing data, although the<br />

absolute fold changes vary due to the fundamental differences<br />

between techniques. For example, Table S-3 in the Supporting<br />

Information shows agreement for 22 of 26 genes (one sequence<br />

was highly repetitive in the genome and thus could not be<br />

quantified) found to be differentially expressed on both PC and<br />

glass microarrays and Table S-4 in the Supporting Information<br />

shows agreement for 39 of 41 genes found to be differentially<br />

expressed only on PC microarrays. It is also apparent that the<br />

PC arrays detect genes with lower average RPKM values (average<br />

RPKM of 65.0 for both samples in Table S-4 in the Supporting<br />

Information) than detected reliably on the glass slides alone<br />

(average RPKM of 2160 for both samples in Table S-3 in the<br />

Supporting Information).<br />

As shown in Table S-1 in the Supporting Information (which<br />

lists genes with significant differences in expression as detected<br />

on both glass and PCs), a number of the genes found to be<br />

overexpressed in the seed cotyledons (with negative fold changes)<br />

include those that encode well-known soybean storage proteins<br />

(i.e., glycinin, lectin, Kunitz trypsin inhibitor, and the Bowman-<br />

Birk proteinase inhibitor) whose mRNA transcripts are abundant<br />

during seed embryogenesis. 23,24 For example, RNA-seq transcriptomics<br />

data confirms some of these with very large RPKM values<br />

of up to 17 340 in the seed and no detectable transcripts in the<br />

leaves (Table S-3 in the Supporting Information). On the other<br />

hand, those genes encoding photosynthetic proteins as the<br />

Rubisco small chain precursor and chlorophyll a/b binding protein<br />

are overexpressed in the trifoliate leaves, as expected. As shown<br />

in Table S-2 in the Supporting Information, the additional genes<br />

detected as differentially expressed with significant p-values on<br />

the PCs represent various enzymes and transcription factors found<br />

to be expressed at lower levels by RNA transcriptome sequencing<br />

(Table S-4 in the Supporting Information), demonstrating the<br />

usefulness of the PCs to detect low expression transcripts.<br />

DISCUSSION<br />

By engineering PC resonances for compatibility with a commercial<br />

laser scanner, the benefits of enhanced fluorescence can<br />

be applied to a standard microarray experiment with no changes<br />

(23) Thibaud-Nissen, F.; Shealy, R. T.; Khanna, A.; Vodkin, L. O. Plant Physiol.<br />

2003, 132, 118–136.<br />

(24) Jones, S. I.; Gonzalez, D. O.; Vodkin, L. O. BMC Genomics 2010, 11, 136.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6859


Figure 6. Logarithmic plot of fold change comparisons between<br />

glass microarray, PC microarray, and sequencing results for five<br />

genes randomly selected from the list of genes found to be differentially<br />

expressed on the PCs but not on the glass slides. Fold change<br />

values for glass and PC microarrays were calculated by determining<br />

the ratio between average microarray expression level for trifoliate<br />

samples to average microarray expression level for cotyledon<br />

samples. A similar ratio was calculated using number of reads for<br />

sequencing data. The PC microarray results show similar directions<br />

and magnitudes of change compared to sequencing data for all 5<br />

genes, while glass microarray data for CHP089 and CHP004 does<br />

not agree with the sequencing data.<br />

to the experimental protocol. While the initial photolithography<br />

process needed to fabricate the silicon mold of the grating has<br />

high costs, a single round of photolithography on silicon can be<br />

translated into thousands of devices that are fabricated uniformly<br />

over large areas. By fabricating the mold on an 8 in. wafer, there<br />

is a large degree of flexibility in fitting PCs to preferred labware<br />

formats such as microscope slides and microtiter plates. Because<br />

PCs were cut to fit standard microscope slides, they could be<br />

processed with existing protocols and scanned with commercially<br />

available equipment, allowing for convenient adoption of PC<br />

substrates in a standardized experiment. Not only does the<br />

nanoreplica molding process provide a convenient form factor for<br />

the substrates, it also enables the excellent level of optical<br />

uniformity required to ensure that every spot on the microarray<br />

experiences the same level of enhancement. This is key to ensure<br />

that the data obtained from PCs does not have a higher level of<br />

variation than the data obtained from glass slides.<br />

The signal enhancement factor primarily used in previous work<br />

in this field is defined as spot intensity subtracted by the local<br />

background observed on the enhancement substrate divided by<br />

the same value observed on the glass slide or control substrate.<br />

The signal enhancement factor observed from Cy-5 spots with<br />

high expression genes in this microarray experiment was approximately<br />

60×, which is identical to the enhancement demonstrated<br />

in previous work with this substrate. 14 Thus, the PC signal<br />

enhancement compares favorably to experiments with metal island<br />

films that have yielded signal enhancement factors of 10-40×. 5,6<br />

However, the signal enhancement factor is not an ideal measurement<br />

to assess the practical utility of the substrate. This work<br />

has focused on SNR enhancement rather than signal enhancement<br />

because microarray data analysis programs use SNR values to<br />

classify spots as detected or not detected. It is possible to achieve<br />

6860 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

good signal enhancement without achieving similar SNR enhancement<br />

if a substrate enhances fluorescence but has a large noise<br />

value thus voiding any advantages of fluorescence enhancement.<br />

Without knowledge of a substrate’s impact on the SNR observed<br />

from spots, it is difficult to ascertain whether a substrate will<br />

benefit a target assay. The PC not only attains a large signal<br />

enhancement but it also achieves an SNR enhancement of<br />

approximately 10× (measured over all spots in the experiment),<br />

suggesting that the array can detect hybridization at concentrations<br />

10× lower than can be detected on glass substrates.<br />

The noise in a DNA microarray experiment arises primarily<br />

from the following sources: sample variation, nonspecific binding,<br />

instrumentation, and substrate fluorescence. Variation in the<br />

amount of nucleic acid sample captured is accounted for by<br />

hybridizing multiple arrays and figures prominently into tests of<br />

significance for differential expression experiments. Nonspecific<br />

binding is controlled largely by blocking and hybridization<br />

conditions and is assessed by evaluating negative control spots.<br />

The noise observed from instrumentation can be characterized<br />

by measurements of dark noise, but in this experiment, this<br />

represents only


data as well. The 41 genes in Tables S-4 and S-2 in the Supporting<br />

Information corresponding to genes detected as differentially<br />

expressed on the PC slides had an average RPKM value of 56 in<br />

the leaf sample and 67 in the seed sample, whereas the 26 genes<br />

detected as differentially expressed on both PC and glass slides<br />

had much higher average RPKMs in both the leaf (341 PRKM)<br />

and seed (3965 RPKM) samples, respectively. Thus, both microarray<br />

and sequencing data suggest the PC can reliably quantify<br />

genes with expression levels at least 1 order of magnitude lower<br />

than measured with conventional glass microarrays. By expanding<br />

the dynamic range of the microarray experiment, the number of<br />

genes for which statistically significant changes in expression<br />

could be observed improved from 27 to 68 genes, or from 13 to<br />

34% of the genes probed in the experiment. Because the gene<br />

expression follows a power law distribution, modest enhancements<br />

in the performance of the assay can dramatically increase the<br />

number of genes researchers are able to probe in this microarray<br />

format. This data thus suggests that the detection capabilities of<br />

microarray protocols currently used today can be greatly expanded<br />

by substitution of conventional substrates with enhanced fluorescence<br />

substrates such as PCs.<br />

The increased SNRs provided by PCs may allow researchers<br />

to perform experiments that are currently problematic on glass<br />

slides. Because lower amounts of bound sample can be detected<br />

with the PC, sample sizes may be reduced to volumes that would<br />

be difficult to probe using normal glass substrates. This may be<br />

particularly helpful for profiling gene expression in smaller tissue<br />

samples or small populations of rare cells such as stem cells.<br />

Alternately, the reduction in experimental variation afforded by<br />

this substrate may allow researchers to confidently identify<br />

differentially expressed genes with fewer replicates, which may<br />

also prove useful with small sample sizes or rare cells. This<br />

approach is not limited to conventional DNA microarray experiments.<br />

Any surface-bound biomolecular assay can be performed<br />

on these PCs for improved performance, as is illustrated in<br />

previous work with immunoassays. 25 This substrate can also<br />

potentially be adapted to improve reliability of novel technologies<br />

such as next-generation genomic sequencing platforms, since<br />

these instruments make extensive use of fluorescent molecules.<br />

CONCLUSION<br />

We have demonstrated that enhanced fluorescence is capable<br />

of significantly improving a DNA microarray that probes changes<br />

(25) Mathias, P. C.; Ganesh, N.; Cunningham, B. T. Anal. Chem. 2008, 80,<br />

9013–9020.<br />

in gene expression between samples. By using a PC substrate<br />

with uniform optical characteristics over microscope slide-sized<br />

areas, the SNR from microarray spots was increased by an order<br />

of magnitude compared to commercial glass substrates. This SNR<br />

enhancement translated into a greater number of genes detected<br />

above the noise level and allowed for the detection of statistically<br />

signficant changes in low expression genes. After evaluating<br />

differential expression in soybean trifoliate tissue versus cotyledon<br />

tissue, more than twice as many genes were characterized as<br />

differentially expressed on the PCs compared to the glass slides,<br />

and many of these were validated by high-throughput mRNA<br />

sequencing data. Using a PC substrate for microarray experiments<br />

thus opens the possibility to interrogate the roles of genes that<br />

previously could not be reliably quantified in a high-throughput<br />

fashion.<br />

ACKNOWLEDGMENT<br />

This work was supported by the National Institutes of Health<br />

(grant no. GM086382A), the National Science Foundation (grant<br />

no. CBET 07-54122), SRU Biosystems, and the Illinois Soybean<br />

Association. Any opinions, findings, conclusions, or recommendations<br />

expressed in this material are those of the authors and do<br />

not necessarily reflect the views of the National Institutes of Health<br />

or the National Science Foundation. The authors thank Stephen<br />

Schulz, Brenda Hugh, Frank Jackson, and Kurt Albertson at SRU<br />

Biosystems for attaching photonic crystal substrates to microscope<br />

slides. The authors thank the staff at the Micro and Nanotechnology<br />

Laboratory at the University of Illinois at Urbana-Champaign.<br />

The authors also thank Sean Bloomfield for bioinformatics<br />

assistance and the staff of the Keck Center at the Biotechnology<br />

Center at University of Illinois for Illumina sequencing.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Tables S-1 and S-2 list genes found to have statistically<br />

significant changes in expression on glass and photonic crystal<br />

slides and Tables S-3 and S-4 show corresponding RNA transcript<br />

levels. This material is available free of charge via the Internet at<br />

http://pubs.acs.org.<br />

Received for review March 31, 2010. Accepted July 2,<br />

2010.<br />

AC100841D<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6861


Anal. Chem. 2010, 82, 6862–6869<br />

Gas Chromatography-Combustion-Mass<br />

Spectrometry with Postcolumn Isotope Dilution for<br />

Compound-Independent Quantification:<br />

Its Potential to Assess HS-SPME Procedures<br />

Sergio Cueto Díaz, Jorge Ruiz Encinar,* Alfredo Sanz-Medel, and J. Ignacio García Alonso*<br />

Department of Physical and <strong>Analytical</strong> <strong>Chemistry</strong>, University of Oviedo, Julián Clavería 8, 33006 Oviedo, Spain<br />

A quadrupole GC-MS instrument with an electron ionization<br />

(EI) source has been modified to enable application<br />

of postcolumn isotope dilution analysis for the standardless<br />

quantification of organic compounds injected in the<br />

gas chromatograph. Instrumental modifications included<br />

the quantitative conversion of the separated compounds<br />

into CO 2, using a postcolumn combustion furnace, and<br />

the subsequent mixing of the gas with a constant flow<br />

of 13 CO2 diluted in helium. The online measurement<br />

of the 12 CO2/ 13 CO2 (44/45) ratio in the EI-MS allowed<br />

us to obtain quantitative data without resorting to<br />

compound-specific standards. Validation of the procedure<br />

involved the analysis of standard solutions<br />

containing different families of organic compounds<br />

(C 9-C20 linear hydrocarbons, BTEX and esters) obtaining<br />

satisfactory results in all cases in terms of<br />

absolute errors (


corresponding compound if its chemical formula is known. Until<br />

now such methodology has been exclusively applied to the<br />

quantitative determination of compounds containing ICPMS<br />

detectable elements for trace-elemental speciation of metals 9 or<br />

semimetals. 10,11 Unfortunately, the use of the ICPMS for the<br />

detection of carbon, hydrogen, nitrogen, or oxygen is seriously<br />

hampered by the low ionization yields of these elements in the<br />

ICP and the high carbon background under normal ICP operating<br />

conditions (atmospheric pressure). In fact, there are only two<br />

reports so far describing the use of HPLC and the postcolumn<br />

addition of 13 C-labeled species for the quantification of organic<br />

compounds by ICP-MS. 12,13 In both cases, the observed detection<br />

limits were not satisfactory.<br />

For the MS detection of organic compounds, previously<br />

separated by GC, electron ionization (EI) is the most common<br />

ionization source employed. It provides both structural and<br />

quantitative information of any volatile organic compound injected<br />

in the gas chromatograph. Unfortunately, as pointed out before,<br />

it requires specific analytical standards as its response is structurespecific<br />

for each single molecule subjected to analysis. 14 Recently,<br />

we have introduced a new quantitative detection concept in gas<br />

chromatography, based on the postcolumn addition of 13 CO2 for<br />

carbon isotope dilution analysis using EI. 15 This concept<br />

constitutes a patented procedure in which organic compounds<br />

separated by liquid or gas chromatography are converted<br />

quantitatively into carbon dioxide, by an oxidation or combustion<br />

reaction, and then are mixed with a postcolumn flow of<br />

enriched 13 CO2. 16 This procedure should provide quantitative<br />

information of every single compound previously separated in<br />

the chromatograph without the need for individual standards.<br />

We selected electron ionization because it operates under high<br />

vacuum conditions providing much better sensitivity and lower<br />

background for carbon detection than the above-mentioned ICP<br />

atmospheric source. For the correct application of isotope<br />

dilution analysis conversion of all carbon containing compounds<br />

to a unique chemical species is required (isotope equilibration).<br />

The isotopic equilibrium between the 13 C-containing species<br />

continuously added postcolumn ( 13 CO2) and the separated<br />

organic compounds is reached by their quantitative conversion<br />

into CO2, after the chromatographic separation, in a combustion<br />

furnace. 17 As the only species to be finally measured by EI is<br />

CO2, compound independent response and isotopic equilibration<br />

is finally obtained. This approach can become a universal<br />

quantitative detector in GC-MS analysis, without the classical<br />

need for specific standards and, what is more, it is compatible<br />

with the exceptional structural identification capabilities provided<br />

by current GC-EI-MS instruments. Thus, in this work<br />

(9) Sariego-Muñiz, C.; Marchante-Gayón, J. M.; García-Alonso, J. I.; Sanz-Medel,<br />

A. J. Anal. At. Spectrom. 2001, 16, 587–592.<br />

(10) Giusti, P.; Schaumlöffel, D.; Ruiz-Encinar, J.; Szpunar, J. J. Anal. At. Spectrom.<br />

2005, 20, 1101–1107.<br />

(11) Heilmann, J.; Heumann, K. G. Anal. Chem. 2008, 80, 1952–1961.<br />

(12) Vogl, J.; Heumann, K. G. Anal. Chem. 1998, 70, 2038–2043.<br />

(13) Smith, C.; Jensen, B. P.; Wilson, I. D.; Abou-Shakra, F.; Crowther, D. Rapid<br />

Commun. Mass Spectrom. 2004, 18, 1487–1492.<br />

(14) Mark, T. D. Int. J. Mass Spectrom. Ion Phys. 1982, 45, 125–145.<br />

(15) Cueto-Díaz, S.; Ruiz-Encinar, J.; Sanz-Medel, A.; García-Alonso, J. I. Angew.<br />

Chem., Int. Ed. 2009, 48, 2561–2564.<br />

(16) Ruiz-Encinar, J.; García-Alonso, J. I World Intellectual Property Organization,<br />

Spain; International Patent WO 2007/042597 A1, 2007.<br />

(17) Merritt, D. A.; Freeman, K. H.; Ricci, M. P.; Studley, S. A.; Hayes, J. M.<br />

Anal. Chem. 1995, 67, 2461–2473.<br />

we describe the instrumental developments and its analytical<br />

features for integral characterization and determination of<br />

organic compounds in detail.<br />

Furthermore, the compound-independent quantification provided<br />

by the developed approach can be an inestimable tool in<br />

the optimization and quantitative assessment of sample extraction<br />

and preconcentration procedures applied prior to GC-MS analysis.<br />

For instance, solid-phase microextraction (SPME) is today a<br />

widely used preconcentration technique in Gas Chromatography. 18<br />

Fiber absorption and recovery are common parameters used when<br />

validating a given SPME procedure. However, those two parameters<br />

are almost impossible to compute using conventional<br />

approaches. 19 In fact, it is necessary to inject specific liquid<br />

standards and to assume that the transfer efficiency in the GC<br />

injector for every compound under analysis is the same when<br />

using conventional injection and the thermal desorption. Thus,<br />

the important topic of absolute absorption yields will be evaluated<br />

in this work, as exemplified for the HS-SPME analysis of BTEX<br />

(Benzene, Toluene, Ethylbenzene, and o,m,p-Xylenes) in different<br />

water samples, using the proposed IDA-EI-MS quantification<br />

procedure.<br />

EXPERIMENTAL SECTION<br />

Reagents and Materials. Solid enriched Na2CO3 (99% 13 C<br />

enrichment) was obtained as a highly pure chemical reagent<br />

(purity >98%) from Cambridge Isotopes Laboratories (Andover,<br />

Massachusetts, USA). A standard mixture of n-alkanes (40 µg/<br />

mL each in n-hexane) was purchased from Fluka (Seelze,<br />

Germany). Phosphoric acid (99% puriss.) was purchased from<br />

Sigma-Aldrich (St. Louis, USA). Individual compounds (undecane,<br />

dodecane, tridecane, pentadecane, butyl butyrate, and<br />

hexyl butanoate) with certified purities ranging from 99 to 99,7%<br />

were used as analytical standards (Fluka, Seelze, Germany or<br />

Sigma-Aldrich, St. Louis, USA). A solution of 1,2,4-trimethylbenzene<br />

in methanol (5000 µg/mL) and a BTEX standard<br />

mixture (2000 µg/mL each in methanol) were both from<br />

Supelco (Bellefonte, USA). Ultrapure water (18.2 MΩcm) was<br />

obtained with a Milli-Q system (Millipore, Bedford, MA).<br />

n-Hexane for organic trace analysis grade was purchased from<br />

Merck (Darmstadt, Germany). SPME holder and fibers were<br />

purchased from Supelco (Bellefonte, USA).<br />

Instrumentation. GC-MS. A Konik-Tech (Sant Cugat del<br />

Vallés, Spain) 4000-B Gas Chromatograph coupled to a Konik-<br />

Tech MS-Q12 quadrupole mass spectrometer with an electron<br />

ionization source was used. This instrument uses a specially<br />

designed injector which keeps the septum at relatively low<br />

temperature, does not require a septum purge flow, and minimizes<br />

losses of organic compounds in the injector port. The analytical<br />

column was a 30 m long (0.32 mm internal diameter, 0.25 µm<br />

stationary phase) DB-XLB column (Agilent J&W Scientific, Santa<br />

Clara, USA). The sample volume injected in the GC was always 1<br />

µL using manual injection. Sample injection was performed in the<br />

splitless (1 min)/split mode. Carrier gas flow was set at 1 mL/<br />

min for all samples and conditions.<br />

Six-Way Valve. A manually actuated 0.25 mm bore stainlesssteel<br />

six-way two-position valve (VICI AG International, Schenkon,<br />

(18) Vas, G.; Vekey, K. J. Mass Spectrom. 2004, 39, 233–254.<br />

(19) Langelfeld, J. J.; Hawthorne, S. B.; Miller, D. J. Anal. Chem. 1996, 68,<br />

144–155.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6863


Switzerland), was placed inside the chromatographic oven to<br />

bypass the combustion furnace when required. The valve has an<br />

extension between the body and the actuator which allows the<br />

valve to be mounted in the heated zone while the handle remains<br />

outside at ambient temperature. This valve prevented the solvent<br />

from entering the combustion unit and therefore enlarged the<br />

catalytic activity of the Cu and Pt wires. Additionally, the valve<br />

allowed the direct connection of the column with the EI source<br />

for conventional GC-MS work in order to identify the different<br />

species under analysis by their fragmentation pattern. Moreover,<br />

such qualitative analysis is essential to assess peak purity for each<br />

compound. For postcolumn isotope dilution analysis the valve was<br />

initially in the “load” position (combustion oven bypassed) and<br />

the valve was manually switched to the “inject” position at the<br />

same time that the EI filament switched on. All connections<br />

between the valve and the rest of the components of the system<br />

were performed by means of 0.32 mm i.d. deactivated fused silica<br />

capillaries and fixed to the valve with appropriate polyimide coated<br />

fused silica adapters (VICI AG International, Schenkon, Switzerland).<br />

Combustion Furnace. The laboratory-made combustion furnace<br />

consisted of a 60 cm long ceramic tube (3 mm O.D., 0.5 mm I.D.)<br />

(Elemental microanalysis, Devon, U. K.) filled with copper and<br />

platinum wires and heated by a Nichrome wire. Proper thermal<br />

isolation of the combustion furnace was provided with glass wool.<br />

The combustion furnace was set vertically on top of the GC with<br />

the lower end of the ceramic tube inside the chromatographic<br />

oven to avoid cold spots after the separation. High temperatures<br />

were accurately controlled using a temperature sensor and an<br />

external controller, allowing temperature settings inside the tube<br />

to be within ±1 °C over a wide temperature range (50-1200 °C).<br />

The copper wires were previously oxidized by passing an oxygen<br />

flow (1 mL/min) at 450 °C during 4-5 h and this procedure was<br />

performed on a weekly basis to preserve its oxidizing capabilities.<br />

Fused silica capillaries (0,32 mm i.d.) were used to connect the<br />

furnace to the valve and to the ion source respectively. For this<br />

purpose a length (∼1 cm) at one end of each capillary was<br />

uncoated, to prevent the polyimide coating to be burnt, and<br />

introduced into the ceramic tube. Reducing unions 1/8′′ to 1/16′′<br />

(SGE, Victoria, Australia) and appropriate graphite ferrules (0.5<br />

mm i.d.) were used to hold the capillaries in place at both sides<br />

of the ceramic tube. For oxidation and combustion, the furnace<br />

was operated at 850 °C.<br />

Gas Cylinder and Mass Flow Controller. The 13 CO2 container<br />

was a dual inlet 5 L high pressure stainless steel gas cylinder<br />

(Iberfluid, Barcelona, Spain). The cylinder was equipped on<br />

one end with, in this order, an opening valve, a Swagelok tee<br />

where helium could be introduced from a high pressure<br />

cylinder, and a second opening valve connected to the other<br />

end of the tee. This second valve was connected to a septum<br />

for the manual injection of gases into the cylinder. Before the<br />

cylinder was pressurized, 13 CO2 was injected into the container<br />

by means of a gastight syringe (Hamilton, Reno, U. S. A.)<br />

through this valve. After the injection of the tracer, the valve<br />

was closed and the container was pressurized up to 6 bar with<br />

helium using the connection in the Swagelok tee. When the<br />

set pressure was reached the filling valve was also closed. At<br />

the other end of the cylinder a pressure gauge, an opening<br />

6864 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

valve and a mass-flow controller (Bronkhorst, Ruurlo, Netherlands)<br />

calibrated for He were coupled for the accurate control<br />

of the tracer flow. The opening valve was closed during the<br />

filling of the cylinder, remaining open the rest of the time. The<br />

flow rate for the postcolumn spike was set at 0.5 mL/min.<br />

Effluent and spike flows were mixed after combustion by means<br />

of a 0.25 mm bore stainless steel microvolume “Y” connector<br />

(VICI AG International, Schenkon, Switzerland).<br />

The whole instrumental setup is shown in Figure 1. As can be<br />

observed, the instrument can be operated in the standard GC-<br />

MS configuration (qualitative) or in the combustion-postcolumn<br />

configuration (quantitative) depending on the position of the<br />

switching valve.<br />

Procedures. Preparation of 13 CO2. The spike was prepared<br />

from 13 C enriched Na2CO3 (99%). An accurate weighed amount<br />

(∼200 mg) was placed in a 25 mL three-necked round-bottomed<br />

flask, previously purged with He to avoid natural abundances<br />

CO2 contamination from ambient air. A small quantity (300 µL)<br />

of concentrated H3PO4, was injected into the flask through a<br />

septum cap. After the acid-base reaction, 4 mL of the gaseous<br />

phase, containing 13 CO2 diluted in He, were removed using a<br />

gastight syringe and injected into the container shown in<br />

Figure 1.<br />

Calibration of the 13 CO2 Postcolumn Flow. The flow rate of<br />

the spike could be accurately controlled by the mass flow<br />

controller between 0.1 and 5 mL min -1 . In our experiments, it<br />

was set at 0.5 mL min -1 . The exact mass flow (ng of 13 CO2 per<br />

min) being mixed with the natural CO2 coming from the<br />

column and combustion furnace was determined by adding<br />

internal standards of known concentration spiked to the sample<br />

as described before. 11,15<br />

Quantification using Postcolumn Isotope Dilution. In our case,<br />

the isotope ratio 12 C/ 13 C was measured as the signal ratio at<br />

masses 44 and 45 (I44/I45) corresponding to the continuous<br />

blend of natural abundance 12 CO2 present in the chromatographic<br />

eluent and the enriched 13 CO2, added postcolumn.<br />

Selected Ion Monitoring (SIM) at masses 44.0 and 45.0 was<br />

performed for the duration of the chromatogram with 70 ms<br />

integration time per mass. The mass window was ∼0.1 mass<br />

units. Then, the isotope ratio in the blend, Rb ) I44/I45, was<br />

calculated to build the isotope ratio chromatogram (Rb vs time).<br />

The postcolumn isotope dilution equation, 7 shown as equation 1<br />

below, was then applied to every point in the chromatogram<br />

to obtain the mass-flow chromatogram (ng of C/min vs time).<br />

The integration of the mass flow chromatogram directly<br />

provided the amount of carbon (in ng) eluted in each chromatographic<br />

peak.<br />

MF n ) MF t<br />

AW n<br />

13<br />

At AWt An 12( Rb - Rt 1 - RbRn) In this equation MFn corresponds to the mass flow of carbon<br />

from the natural abundance sample injected, whereas MFt<br />

corresponds to the mass flow of carbon from the postcolumn<br />

spike or tracer. AWn and AWt correspond to the atomic weight<br />

of carbon in the sample and tracer, respectively. The isotope<br />

abundances At 13 and An 12 correspond to the isotopic composition<br />

of 13 C in the tracer and 12 C in the sample. Finally, Rt is the<br />

(1)


Figure 1. Schematic illustration of the setup showing the modification made in the GC-MS instrument. The two possible valve configurations<br />

are shown in the inset: Left, for structural identification by conventional GC-EI-MS analysis, and right, for absolute quantification using GC-<br />

Combustion-MS and postcolumn addition of 13 CO2. Adapted from reference. 15<br />

isotope ratio (12/13) in the tracer and Rn is the natural isotope<br />

ratio (13/12) in the sample, whereas Rb is the measured isotope<br />

ratio in the blend.<br />

Additionally, eq 1 can be used to quantify the mass flow of<br />

enriched carbon dioxide when an internal standard of known<br />

concentration is spiked to the sample.<br />

Head-Space SPME Procedure. Test solutions (∼10 ng/g) were<br />

prepared by diluting appropriately the BTEX standards in Milli<br />

Q water. Samples of ∼7 g were pipetted into 10 mL glass vials<br />

with open-top phenolic closures and PTFE/silicone septa (Supelco,<br />

Bellefonte, USA). Three fiber coatings, 100 µm polydimethylsiloxane<br />

(PDMS), 70/30 µm polydimethylsiloxane/divinylbenzene<br />

(PDMS/DVB), and 50/30 µm carboxen/polydimethylsiloxane/<br />

divinylbenzene (CAR/PDMS/DVB) were tested for comparative<br />

purposes. Samples were saturated in all cases with NaCl to<br />

improve extraction efficiency and stirred at 900 rpm at 30 °C<br />

during 30 min to allow equilibration between the liquid and gas<br />

phases. Then the fibers were inserted in the headspace and<br />

extracted during 15 min to ensure equilibrium conditions between<br />

the headspace and the fiber. After extraction, the fibers were<br />

removed and thermal desorption was carried out in the injection<br />

port of the GC at 250 °C.<br />

RESULTS AND DISCUSSION<br />

Online Continuous Measurement of the 12 C/ 13 C Isotope<br />

Ratios. The optimum conditions for the continuous measurement<br />

of 12 C/ 13 C isotope ratios were established by connecting directly<br />

the reservoir that contained the 13 CO2 spike to the mass<br />

spectrometer. 15 The mass spectrum obtained for the mass<br />

range 43-47 when such postcolumn flow was set at 1 mL/<br />

min is provided in the Supporting Information (Figure SI-1).<br />

The 13 C isotope enrichment in the postcolumn spike (∼97%)<br />

can be clearly seen from the spectrum. Resolution was<br />

optimized while still preserving a good sensitivity. The mass<br />

resolution finally obtained allowed baseline separation even for<br />

isotope ratios very far from 1. The experiments performed for<br />

the selection of the optimum integration time are given in the<br />

Supporting Information (Figure SI-2). An integration time of 70<br />

ms was finally selected considering both the precision in the ratio<br />

measurement and the number of points per chromatographic peak<br />

(∼15) to obtain well-defined peak profiles.<br />

Spike Enrichment and Stability of the 12 CO2/ 13 CO2<br />

Isotope Ratio. Again, to estimate the 13 C enrichment obtained<br />

after the preparation of the 13 CO2 spike, its reservoir was<br />

connected directly to the mass spectrometer. A constant flow<br />

of1mLmin -1 was set and the signals at m/z ratios 44 and 45<br />

were continuously monitored during 2 min for each independent<br />

isotope ratio measurement under the measurement<br />

conditions described in the procedures. The isotope ratio was<br />

calculated obtaining an average value of 0.034 ± 0.0003, which<br />

corresponds to a 13 C abundance of 96.70 ± 0.02% (n ) 5) (see<br />

the mass spectrum shown in the Figure SI-1). The slight<br />

difference between the experimentally obtained enrichment and<br />

that provided for the precursor Na2 13 CO3 by the supplier (99%),<br />

can be ascribed to small contaminations by natural abundance<br />

CO2 from the ambient air during the generation of the tracer<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6865


Figure 2. Intensity chromatograms for the n-alkanes mixture<br />

obtained in the SIM mode: (a) at m/z ) 71 operating the setup in the<br />

qualitative configuration; (b) at m/z 44 and 45 operating the setup in<br />

the quantitative configuration.<br />

and the filling of the reservoir, traces of this compound possibly<br />

present in the Helium gas used to dilute de 13 CO2 and traces<br />

of air still present in the EI source. The high enrichment<br />

obtained provides a very wide range of optimum analyte/spike<br />

ratios and then a sample containing compounds in a large range<br />

of concentrations could be quantified in the same analysis<br />

without affecting the accuracy of the results. 3<br />

The short-term (30 min) and long-term (9 h) stability of the<br />

12 C/ 13 C isotope ratio is given in the Supporting Information<br />

(Figures SI-3 and SI-4). In both cases the measured isotope ratios,<br />

0.0339 ± 0.0002 and 0.034 ± 0.001 (n ) 8), respectively, were<br />

remarkably stable. Similar results were obtained in different days<br />

indicating that the setup was robust enough for routine analysis.<br />

Evaluation of the GC-Combustion-IDMS System. The<br />

overall performance of the developed instrumentation was evaluated<br />

here by analyzing a standard solution of a mixture of<br />

n-alkanes (C9-C20). To study peak broadening, 1 µL of a<br />

solution containing approximately 5 µg/g of these compounds<br />

in n-hexane was injected in the chromatograph both in the<br />

qualitative (combustion oven bypassed) and quantitative (through<br />

the oven) modes. The Selected Ion Monitoring chromatogram<br />

obtained at m/z ) 71 (fragment characteristic of n-alkanes) is<br />

shown in Figure 2a, whereas the chromatogram obtained at<br />

masses 44 and 45 after combustion and postcolumn isotope<br />

dilution analysis is shown in Figure 2b. As can be observed, no<br />

significant peak broadening due to the combustion unit or to the<br />

different connections was observed, being the peak width measured<br />

at the half height equal to that found when operating the<br />

6866 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 3. Mass flow chromatograms obtained for the n-akanes<br />

mixture.<br />

GC-MS in the conventional way (0.02 min, Figure 2a). Only a small<br />

increase in retention times was observed (∼20 s) because of the<br />

combustion furnace. The mass flow chromatogram obtained after<br />

the application of the online isotope dilution equation (eq 1) is<br />

shown in Figure 3. As can be observed, the sensitivity is roughly<br />

constant for all compounds which is in clear contrast with the<br />

results shown in Figure 2a for mass 71 without combustion (SIM<br />

detection). The peak areas in the mass flow chromatogram were<br />

linear with the amount of carbon injected over 2 orders of<br />

magnitude (the maximum range assayed). The detection limit<br />

obtained for tetradecane was 9 ppb (ng/g) based on three times<br />

the standard deviation of the baseline. Taking into account that<br />

the injection volume was 0.5 µL, this value corresponds to an<br />

absolute limit of detection of 3 pg of tetradecane injected. If we<br />

take into account the peak width at the baseline (4s approximately<br />

in this particular case), then we can make our detection limit<br />

independent of the column employed and the carrier gas flowrate.<br />

The value found was 0.8 pg C s -1 , being in the same order<br />

that those provided by a mass spectrometer in full scan mode<br />

(1 pg C s -1 ) and clearly better than those of a flame ionization<br />

detector (10 pg C s -1 ). This detection limit is mainly limited<br />

by the background level observed at m/z ) 44, and explains<br />

the difference in comparison with the detection limit achievable<br />

using the mass spectrometer in SIM mode bypassing the oven<br />

(typically 0.1 pg C s -1 ) in spite of the fact that both sensitivities<br />

were very similar. In contrast, it should be mentioned that the<br />

EI detection limit observed is 3 orders of magnitude lower than<br />

that obtained for carbon using 13 C postcolumn isotope dilution<br />

and ICP-MS detection (0.7 ng C s -1 ). 12<br />

For the quantitative analysis of the standard mixture of<br />

n-alkanes (C9-C20) we needed first to quantify the mass flow<br />

of postcolumn 13 CO2. Out of the two possibilities of using<br />

internal standard (IS) or external standard, we selected an IS<br />

because it provided the advantage of compensation for small<br />

variations in the sample volume injected (we were using manual<br />

injection). Moreover, as recently pointed out by Heilmann and<br />

Heumann, 11 an additional advantage of using an internal<br />

standard is that the mass flow of enriched carbon dioxide does<br />

not need to be determined. The peak areas of both the analyte<br />

and the internal standard in the mass flow chromatogram are<br />

related to the unknown mass flow of spike (eq 1). However,<br />

the ratio of peak areas analyte/IS will be equal to the actual ratio<br />

of concentrations in the injected sample and independent of the<br />

mass flow of spike used in the calculations. Thus, quantification


Table 1. Absolute Quantification Results for the<br />

Alkane Model Mixture a<br />

compound added (µg/g) found (µg/g)<br />

nonane 8.0 8.3 ± 0.5<br />

decane 8.0 8.3 ± 0.2<br />

undecane 8.0 8.4 ± 0.1<br />

dodecane 8.0 8.3 ± 0.3<br />

tridecane 8.0 8.3 ± 0.3<br />

tetradecane 8.0 I.S.<br />

pentadecane 8.0 7.7 ± 0.1<br />

hexadecane 8.0 7.7 ± 0.1<br />

heptadecane 8.0 7.4 ± 0.3<br />

octadecane 8.0 7.9 ± 0.1<br />

nonadecane 8.0 7.8 ± 0.1<br />

eicosane 8.0 8.3 ± 0.2<br />

a Tetradecane was used as an internal standard. The uncertainty is<br />

expressed as the standard deviation for n ) 3 injections.<br />

of the amount of carbon under each peak can be determined just<br />

by calculating the areas of the internal standard and those<br />

corresponding to the other organic compounds. Therefore, the<br />

knowledge of the spike mass flow is not needed. The quantitative<br />

results obtained from the data shown in Figure 3, using tetradecane<br />

as internal standard, are given in Table 1. As can be observed,<br />

this methodology allowed us to quantify a series of organic<br />

compounds containing from 9 to 20 atoms of carbon and with a<br />

range of boiling points from 128 to 343 °C, with acceptable<br />

precision (


Figure 4. Mass flow chromatogram obtained for the mixture of BTEX<br />

and 1,2,4-trimethylnenzene in water using HS-SPME-GC-Combustion-MS<br />

and postcolumn addition of 13 CO2.<br />

nately, the CAR/PDMS/DVB fiber generated broader peaks and<br />

tailing, resulting in lower signal-to-noise ratio and poorer precision<br />

in comparison with the PDMS/DVB fiber, which was finally<br />

selected for further experiments. In addition, the fiber/sample<br />

distribution constants (Kfs) were calculated for a more reliable<br />

comparison between the extraction efficiency of the fibers as<br />

it is a parameter related to the phase volume 23 (see Table SI-2<br />

in the Supporting Information). The Kfs values obtained for the<br />

PDMS fiber were of the same order of those cited in the<br />

literature. 19 Interestingly, Kfs values obtained for the PDMS/<br />

DVB fiber were significantly higher than those obtained for<br />

the CAR/PDMS/DVB fiber, in spite that both fibers provided<br />

very similar absolute recoveries (see Table 2). These results<br />

seem to indicate that the PDMS/DVB stationary phase showed<br />

higher affinity for the BTEX compounds. Notably, only nine<br />

analysis (three replicates per fiber tested) were necessary to obtain<br />

this valuable information in fundamental studies, critical to<br />

optimize new procedures of SPME.<br />

As an example, Figure 4 shows the mass-flow chromatogram<br />

obtained using the PDMS/DVB fiber. In this case, together with<br />

the impressive preconcentration factor typically observed by<br />

SPME, the carbon background was significantly decreased (two<br />

or three times lower). The detection limits obtained using HS-<br />

SPME were in the low ng/L level (


ment. Therefore, the long-time dreamed of chromatographic<br />

detection allowing generic quantification and complete characterization<br />

of organic compounds in a single instrument is<br />

realized. In addition, the proposed method turned out to be<br />

very cost-effective because expenses associated with the use<br />

of isotopically labeled 13 CO2 are negligible and there is no need<br />

for external calibration, resulting in considerable savings in time<br />

and money (e.g., cost of analytical standards). Beyond its<br />

potential use for quantitative quality control in a wide range of<br />

standard laboratories (e.g., environmental ones), a powerful<br />

application of our approach can be foreseen in oil-spill fingerprinting<br />

and in pharmaceutical analysis, for which the number<br />

of target organic analytes is increasing exponentially (so it is<br />

virtually impossible to have standards for each compound, even<br />

if they exist).<br />

In addition, the unique compound-independent calibration<br />

capabilities of the approach proposed could be advantageously<br />

exploited for the assessment of sample introduction procedures<br />

in GC. In the present work, we have selected HS-SPME to<br />

illustrate this promising ability of our approach. Quantitative data<br />

obtained for BTEX compounds in different water samples have<br />

demonstrated that quantitative assessment of HS-SPME procedures<br />

in terms of absolute absorption yields is successful (the<br />

analytical features of different fiber coatings could be evaluated<br />

and critically compared in this way), opening its use to assess<br />

the performance of many other reported sample preparation and<br />

preconcentration methods.<br />

ACKNOWLEDGMENT<br />

Financial support was provided by the Spanish Ministry of<br />

Education (CTQ2006-05722) and FICYT (PC06-016) and technical<br />

support by KONIK-TECH. J.R.E. acknowledges the MEC (European<br />

Social Fund) for a Ramon y Cajal contract, and S.C.D.<br />

acknowledges the FICYT for a PhD grant.<br />

SUPPORTING INFORMATION AVAILABLE<br />

The mass spectrum obtained for the mass range 43-47 when<br />

1 mL/min of the spike flow was introduced directly to the ion<br />

source (Figure SI-1). The optimization of the integration time for<br />

the 44/45 isotope ratio measurement (Figure SI-2). The short term<br />

(30 min) and long term (9 h) stability of the 12 CO2/ 13 CO2 isotope<br />

ratio (Figures SI-3 and SI-4, respectively). The quantification<br />

results of a mixture of different families of compounds containing<br />

saturated, insaturated and functionalized compounds (Table SI-<br />

1). The Kfs values obtained for reliable comparison between the<br />

different fibers used (Table SI-2). The multiple headspace<br />

experiment carried out for internal validation of the approach,<br />

comparing the sum of the amount obtained in each single<br />

extraction to the total amount initially present in the spiked<br />

solution (Table SI-3). This material is available free of charge via<br />

the Internet at http://pubs.acs.org.<br />

Received for review April 12, 2010. Accepted July 2, 2010.<br />

AC101103N<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6869


Anal. Chem. 2010, 82, 6870–6876<br />

Autonomous Microfluidic Control by <strong>Chemical</strong>ly<br />

Actuated Micropumps and Its Application to<br />

<strong>Chemical</strong> Analyses<br />

Atsushi Takashima, Kenichi Kojima, and Hiroaki Suzuki*<br />

Graduate School of Pure and Applied Sciences, University of Tsukuba, 1-1-1 Tennodai, Tsukuba,<br />

Ibaraki 305-8573, Japan<br />

Autonomous control of microfluidic transport was realized<br />

through the use of chemically actuated diaphragm micropumps<br />

connected to a network of controlling flow channels.<br />

A hydrogen peroxide (H 2O2) solution was transported<br />

in the controlling flow channel by capillary<br />

action. Upon the solution’s arrival at the lower compartment<br />

of a micropump filled with manganese dioxide<br />

(MnO 2) powder, a volume change that accompanied<br />

the production of oxygen caused by the catalytic<br />

decomposition of H 2O2 induced inflation of the diaphragm.<br />

This in turn caused the movement of a<br />

solution in another network of flow channels formed<br />

in the upper layer. Micropumps that only exert pressure<br />

were also fabricated. By positioning the micropumps<br />

at appropriate locations in conjunction with<br />

additional flow-delaying components, the ejection of<br />

solutions from the reservoir of each micropump could<br />

be initiated at coordinated times. Furthermore, the<br />

solutions could be transported by the application of<br />

pressure from other micropumps. In other words, the<br />

information for switching from one micropump to<br />

another could be described on the chip in the form of<br />

a network of flow channels. This autonomous processing<br />

of solutions was demonstrated for enzymatic analyses<br />

of H 2O2, glucose, and lactate.<br />

With the progress now being made in microfluidic technologies,<br />

innovative devices that make possible the complicated<br />

manipulation of solutions have been proposed for various<br />

applications. 1-5 However, as far as microfluidic transport is<br />

concerned, many of these previous devices have relied on external<br />

instruments such as microsyringe pumps or power sources to<br />

produce pressure-driven flows or to generate electroosmotic flows.<br />

Such bulky instruments, however, are obstacles to the increased<br />

integration of components and miniaturization of the entire system.<br />

* To whom correspondence should be addressed. Phone: +81-29-853-5598.<br />

Fax: +81-29-853-4490. E-mail: hsuzuki@ims.tsukuba.ac.jp.<br />

(1) Thorsen, T.; Maerkl, S. J.; Quake, S. R. Science 2002, 298, 580–584.<br />

(2) Balagaddé, F. K.; You, L.; Hansen, C. L.; Arnold, F. H.; Quake, S. R. Science<br />

2005, 309, 137–140.<br />

(3) Shiu, J.-Y.; Chen, P. Adv. Mater. 2005, 17, 1866–1869.<br />

(4) Wang, C.-H.; Lee, G.-B. Biosens. Bioelectron. 2005, 21, 419–425.<br />

(5) Satoh, W.; Hosono, H.; Yokomaku, H.; Morimoto, K.; Upadhyay, S.; Suzuki,<br />

H. Sensors 2008, 8, 1111–1127.<br />

6870 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

To address this problem, trials have been performed to integrate<br />

active microfluidic components on a single chip. 6-10<br />

The long history of the development of micropumps and<br />

microvalves has produced a variety of devices that are based on<br />

various principles. 11-13 In the reported devices, actuation of the<br />

components has usually been based on switching by electrical<br />

signals that are programmed in a number of ways. In this<br />

approach, however, specially designed electronic circuits and<br />

software are needed to realize cooperative operation of the<br />

components. In addition, for disposable devices, it would be<br />

preferable for the microfluidic system to function autonomously.<br />

To resolve this problem, chemical actuators that are based upon<br />

the spontaneous volume change of a hydrogel have been<br />

reported. 14-16 Capillary action has also been used to realize a<br />

variety of devices for the autonomous transport of solutions and<br />

various other applications. 17-23 These previous approaches,<br />

particularly the latter one, suggest a direction for the realization<br />

of more sophisticated devices in the next generation. The<br />

manipulation of solutions in devices has been based on programmed<br />

instructions described on the chip as a structural<br />

(6) Choi, J.-W.; Oh, K. W.; Han, A.; Okulan, N.; Wijayawardhana, C. A.; Lannes,<br />

C.; Bhansali, S.; Schlueter, K. T.; Heineman, W. R.; Halsall, H. B.; Nevin,<br />

J. H.; Helmicki, A. J.; Henderson, H. T.; Ahn, C. H. Biomed. Microdevices<br />

2001, 3, 191–200.<br />

(7) Srinivasan, V.; Pamula, V. K.; Fair, R. B. Lab Chip 2004, 4, 310–315.<br />

(8) Satoh, W.; Hosono, H.; Suzuki, H. Anal. Chem. 2005, 77, 6857–6863.<br />

(9) Nashida, N.; Satoh, W.; Fukuda, J.; Suzuki, H. Biosens. Bioelectron. 2007,<br />

22, 3167–3173.<br />

(10) Abdelgawad, M.; Wheeler, A. Adv. Mater. 2009, 21, 920–925.<br />

(11) Gravesen, P.; Branebjerg, J.; Jensen, O. S. J. Micromech. Microeng. 1993,<br />

3, 168–182.<br />

(12) Shoji, S.; Esashi, M. J. Micromech. Microeng. 1994, 4, 157–171.<br />

(13) Laser, D. J.; Santiago, J. G. J. Micromech. Microeng. 2004, 14, R35–R64.<br />

(14) Beebe, D. J.; Moore, J. S.; Yu, Q.; Liu, R. H.; Kraft, M. L.; Jo, B.-H.; Devadoss,<br />

C. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 13488–13493.<br />

(15) Suzuki, H.; Kumagai, A.; Ogawa, K.; Kokufuta, E. Biomacromolecules 2004,<br />

5, 486–491.<br />

(16) Suzuki, H.; Tokuda, T.; Kobayashi, K. Sens. Actuators, B 2002, 83, 53–59.<br />

(17) Zhao, B.; Moore, J. S.; Beebe, D. J. Science 2001, 291, 1023–1026.<br />

(18) Ahn, C. H.; Choi, J.-W.; Beaucage, G.; Nevin, J. H.; Lee, J.-B.; Puntambekar,<br />

A.; Lee, J. Y. Proc. IEEE 2004, 92, 154–173.<br />

(19) Bouaidat, S.; Hansen, O.; Bruus, H.; Berendsen, C.; Bau-Madsen, N. K.;<br />

Thomsen, P.; Wolff, A.; Jonsmann, J. Lab Chip 2005, 5, 827–836.<br />

(20) Delamarche, E.; Juncker, D.; Schmid, H. Adv. Mater. 2005, 17, 2911–<br />

2933.<br />

(21) Chung, K. H.; Hong, J. W.; Lee, D.-S.; Yoon, H. C. Anal. Chim. Acta 2007,<br />

585, 1–10.<br />

(22) Zimmermann, M.; Hunziker, P.; Delamarche, E. Microfluid. Nanofluid.<br />

2008, 5, 395–402.<br />

(23) Swickrath, M. J.; Burns, S. D.; Wnek, G. E. Sens. Actuators, B 2009, 140,<br />

656–662.<br />

10.1021/ac1009657 © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/29/2010


Figure 1. <strong>Chemical</strong>ly actuated micropumps with flow channels. (A) Exploded view of the micropumps and the flow channels. (B) Operation of<br />

the micropump. Cross-sections are shown that include the flow channel for transport, the diaphragm, and the lower compartment for the H2O2<br />

solution. First, the reservoir of the micropump is filled with a solution to be transported (top). When a H2O2 solution is transported in the controlling<br />

flow channel and reaches the lower compartment of the micropump, bubbles are produced, the diaphragm inflates, and the solution in the upper<br />

reservoir is injected into the upper flow channel (bottom).<br />

arrangement of components, including the flow channel network. 18<br />

Although there have been limitations in manipulation that is<br />

performed only through capillary action, even the complicated<br />

manipulation of solutions may be realized by coupling a programmed<br />

microfluidic network with chemically actuated microfluidic<br />

components.<br />

In a number of previous studies, gas bubbles produced by the<br />

electrolysis of water were used to produce a volume change that<br />

would mobilize a solution in a microflow channel. 24-29 This<br />

principle of operation is attractive for the realization of a chemically<br />

actuated micropump, because gas production is accompanied by<br />

many chemical reactions. In creating our device, we used the<br />

volume change of oxygen bubbles produced by the catalytic<br />

decomposition of H2O2. 26 To trigger the pumping action, a H2O2<br />

solution was transported and supplied to the micropumps by<br />

capillary action in a controlling flow channel. A network of<br />

controlling flow channels described on a chip could be used<br />

as a program to operate many micropumps cooperatively. In<br />

other words, the timing of the switching among pumps could<br />

be adjusted by changing the relative positions of the micropumps<br />

and the length or other dimensional parameters of the<br />

flow channels. In this paper we present the basic concept for<br />

a chemically actuated micropump and its programming and<br />

characterize the performance of the device.<br />

(24) Böhm, S.; Timmer, B.; Olthuis, W.; Bergveld, P. J. Micromech. Microeng.<br />

2000, 10, 498–504.<br />

(25) Suzuki, H.; Yoneyama, R. Sens. Actuators, B 2003, 96, 38–45.<br />

(26) Choi, Y. H.; Son, S. U.; Lee, S. S. Sens. Actuators, A 2004, 111, 8–13.<br />

(27) Satoh, W.; Shimizu, Y.; Kaneto, T.; Suzuki, H. Sens. Actuators, B 2007,<br />

123, 1153–1160.<br />

(28) Shimizu, Y.; Takashima, A.; Satoh, W.; Sassa, F.; Fukuda, J.; Suzuki, H.<br />

Sens. Actuators, B 2009, 140, 649–655.<br />

(29) Blanco-Gomez, G.; Glidle, A.; Flendrig, L. M.; Cooper, J. M. Anal. Chem.<br />

2009, 81, 1365–1370.<br />

EXPERIMENTAL SECTION<br />

Materials and Reagents. A thick-film photoresist (SU-8) was<br />

purchased from MicroChem, Newton, MA. A precursor solution<br />

of poly(dimethylsiloxane) (PDMS) (KE-1300T) was purchased<br />

from Shin-Etsu <strong>Chemical</strong>, Tokyo, Japan. A precursor solution of<br />

PVA-SbQ, SPP-H-13, was purchased from Toyo Gosei Kogyo,<br />

Chiba, Japan. H2O2, manganese dioxide, and poly(oxyethylene)<br />

sorbitan monolaurate (Tween 20) were purchased from Wako<br />

Pure <strong>Chemical</strong> Industries, Osaka, Japan. The enzymes and<br />

related reagents were obtained from the following commercial<br />

sources: horseradish peroxidase (HRP; 100 U/mg), lactate<br />

oxidase (LOD; 38 U/mg), and bovine serum albumin (BSA)<br />

from Wako Pure <strong>Chemical</strong> Industries, Osaka, Japan; glucose<br />

oxidase (GOD; 151 U/mg) and 25% glutaraldehyde (GA)<br />

solution from Sigma-Aldrich, St. Louis, MO; N-acetyl-3,7dihydroxyphenoxazine<br />

(Amplex Red) from AnaSpec, San Jose,<br />

CA.<br />

Basic Structure and Fabrication of the Microfluidic Devices.<br />

The devices were constructed by stacking two PDMS<br />

substrates on a glass substrate (Figure 1). Flow channels were<br />

formed with PDMS using a template formed with a thick-film<br />

photoresist (SU-8). The compartments for the pumps and solutions<br />

to be transported were formed in the lower and upper PDMS<br />

layers by punching.<br />

A critical part of each micropump was a circular compartment<br />

(diameter 2.5 mm) with a diaphragm. The diaphragm was formed<br />

by intercalating a 50 µm thick PDMS sheet between the two<br />

PDMS substrates. The lower part of the compartment was<br />

connected to a controlling flow channel for the transport of a H2O2<br />

solution. To form a MnO2 layer in the vicinity of the diaphragm,<br />

a droplet of water containing a suspension of MnO2 powder<br />

was put into the compartment that was then placed upside<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6871


down until the water evaporated. The amount of the powder<br />

that remained was between 0.75 and 1.35 mg, depending on<br />

the size of the micropump. To prevent leakage of the powder<br />

and more effectively produce oxygen bubbles and exert<br />

pressure upon the diaphragm, the compartment beneath the<br />

diaphragm was stuffed with a plug of a PVA-SbQ gel, leaving<br />

a space below it that could be filled with the H2O2 solution<br />

introduced from the controlling flow channel. In forming the<br />

plug, that space was filled with a precursor solution of PVA-<br />

SbQ, which was then cured under a UV light. Air vents were<br />

formed at appropriate locations to release pressure and facilitate<br />

the transport and filling of the H2O2 solution. A circular<br />

reservoir (diameter 1.0 mm) for a solution to be transported<br />

was also formed with PDMS on the diaphragm layer and was<br />

connected to a flow channel. The flow channels extending from<br />

the reservoirs of several micropumps formed an appropriate<br />

network that reflected their different purposes. An inlet was<br />

formed on the reservoir to be filled with a transported solution.<br />

Micropumps that were used only to apply pressure were<br />

formed in a similar manner. In this case, an inlet for the<br />

transported solution was not formed. For all devices, the<br />

heights of the controlling flow channels and of the flow<br />

channels in which solutions would be transported were 75 and<br />

150 µm, respectively. The width of the flow channel for<br />

transportation was 500 µm. After the reservoirs in the upper<br />

PDMS substrate were filled with the necessary solutions, the<br />

entire structure was inserted between two poly(methyl methacrylate)<br />

(PMMA) plates and then fixed in place with bolts<br />

and nuts.<br />

Procedure and Principle of Operation. A change in the<br />

volume in the upper solution reservoir is caused by the deformation<br />

of the diaphragm that follows the production of oxygen<br />

bubbles produced by the catalytic decomposition of H2O2. First,<br />

aH2O2 solution is transported in the controlling flow channel<br />

by capillary action and is injected into the lower compartment<br />

of the micropump (Figure 1B, top). When the solution reaches<br />

the MnO2 powder below the diaphragm, oxygen bubbles are<br />

produced by the catalytic decomposition of H2O2:<br />

MnO2 2H2O298 2H 2 O + O 2<br />

The diaphragm then inflates and exerts pressure upon the solution<br />

in the reservoir. As a result, the solution is pushed forward in the<br />

flow channel and is transported to the lower stream (Figure 1B,<br />

bottom). The structure can also be modified so that the reservoir<br />

is filled with only air, instead of a solution to be transported. In<br />

this case, only a change in pressure is generated to move a liquid<br />

column that may be present in the lower stream of the extending<br />

flow channel. Several micropumps can be connected to the<br />

controlling flow channel and the flow channel for the transport of<br />

necessary solutions. Each micropump can be switched on individually<br />

and sequentially according to a predetermined schedule<br />

that is programmed in the shape of the network of controlling<br />

flow channels.<br />

Construction and Operation of Analysis Systems. To<br />

demonstrate the sequential manipulation of solutions for chemical<br />

6872 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

analyses, two devices were fabricated. In one of them (Figure 5),<br />

a solution containing 50 U/mL HRP and another solution containing<br />

H 2O2 as an analyte and Amplex Red (5 mM) were filled in<br />

the reservoir of two micropumps. The solutions were prepared<br />

with a 50 mM Tris-HCl buffer solution (pH 7.4). After the<br />

solutions were ejected from the pumps and merged at the<br />

T-junction according to the programmed pumping, their<br />

intensity of fluorescence was measured using a fluorescence<br />

microscope (VB-G25, Keyence, Tokyo, Japan) equipped with<br />

a CCD detection system (Keyence VB-7000/7010). In a more<br />

complicated device (Figure 6), enzymes (GOD and LOD) were<br />

immobilized in two of three injection ports. HRP was immobilized<br />

in a reaction chamber formed in the lower stream. In the<br />

immobilization of the enzymes, an enzyme solution, a 0.1 wt %<br />

BSA solution, and a 0.1 wt % GA solution were mixed in a 1:1:1<br />

ratio. The mixed solution was then dropped into the corresponding<br />

reservoirs or the reaction chamber (5 µL for the GOD and LOD<br />

solutions and 1 µL for the HRP solution), and a cross-linking<br />

reaction was allowed to proceed. Following this, the enzymeimmobilized<br />

layers were immersed in a 0.1 M glycine solution<br />

for 60 min. The activity of the immobilized enzymes was 1.3 U<br />

for GOD, 0.17 U for LOD, and 3.3 × 10 -2 U for HRP. After the<br />

injection ports were filled with solutions containing either<br />

glucose or lactate or both of them, along with the immobilized<br />

enzymes, H2O2 was produced by the enzymatic reactions. The<br />

solutions were then transported to the reaction chamber. The<br />

enzymatic reactions of HRP were accompanied by the generation<br />

of fluorescence, whose intensity was measured. Values for<br />

time, flow velocity, and fluorescence intensity were obtained<br />

in five measurements, whose averages are used in the<br />

following discussion.<br />

RESULTS AND DISCUSSION<br />

Movement of a Solution in the Controlling Flow Channel.<br />

The controlling flow channel consisted of three walls of PDMS<br />

and a bottom of glass. Although PDMS is hydrophobic (contact<br />

angle 110°), the hydrophilic glass bottom alone (contact angle<br />

15°) was able to generate a sufficient driving force to produce<br />

capillary action. The velocity of the column of H2O2 changed<br />

depending on device parameters such as the width, height, and<br />

wettability of the flow channel. The velocity of a liquid column,<br />

x˙, in a straight flow channel is expressed as follows: 30-33<br />

x˙ ) γLV 8ηx( hw<br />

h + w) 2<br />

[<br />

2 cos θPDMS w<br />

+ cos θPDMS + cos θglass h ]<br />

Here, γLV is the interfacial tension between the solution and<br />

the air, η is the viscosity of the solution, x is the distance from<br />

the inlet of the capillary to the meniscus of the moving liquid<br />

column, h and w are the height and width of the flow channel,<br />

and θPDMS and θglass are the contact angles on PDMS and glass,<br />

respectively.<br />

As could be anticipated from the equation, the movement of<br />

the column slowed as it moved forward in the fabricated flow<br />

(30) Satoh, W.; Yokomaku, H.; Hosono, H.; Ohnishi, N.; Suzuki, H. J. Appl. Phys.<br />

2008, 103, 034903.<br />

(31) Janshoff, A.; Künneke, S. Eur. Biophys. J. 2000, 29, 549–554.<br />

(32) Delamarche, E.; Bernard, A.; Schmid, H.; Bietsch, A.; Michel, B.; Biebuyck,<br />

H. J. Am. Chem. Soc. 1998, 120, 500–508.<br />

(33) Kim, E.; Xia, Y.; Whitesides, G. M. Science 1995, 376, 581–584.


Figure 2. Influence of a surfactant (Tween 20) on microfluidic<br />

transport. The distance indicates that of the meniscus of the liquid<br />

column from the reservoir for the H2O2 solution in a straight flow<br />

channel of 500 µm × 75 µm in cross-section. Concentration of the<br />

surfactant: [, 0wt%;×, 0.005 wt %; 2, 0.01 wt %; b, 0.05 wt %; 9,<br />

0.1 wt %.<br />

channels. In addition, the movement occasionally became irregular,<br />

possibly due to the morphological or chemical nonuniformity<br />

of the channel. In the worst case, the column stopped<br />

midway in its journey and did not reach the lower compartment<br />

of the micropump. This problem was solved by adding a surfactant<br />

to the solution, which facilitated smooth movement. Figure 2<br />

shows the dependence of the movement of the solution on the<br />

concentration of the surfactant (Tween 20) that was added. The<br />

influence of the surfactant was dramatic, and the solution’s<br />

movement became smoother and faster with increasing concentration<br />

of surfactant. At concentrations higher than 0.01 wt %, the<br />

flow velocity almost leveled off. In the following experiments, the<br />

concentration was therefore fixed at 0.01 wt %.<br />

In biochemical analyses in microsystems, the length of time<br />

required for a reaction is often on the order of seconds or minutes.<br />

In view of this, an additional requirement in such cases is the<br />

presence of structures that can slow the flow velocity of solutions.<br />

For this reason, we then examined how the velocity of the column<br />

changed with changes in the width of the flow channel. In Figure<br />

3A, flow channels 2-4 are straight and have widths of 250 µm,<br />

500 µm, and 1.0 mm, respectively. With a change made only in<br />

the width, a marked difference in flow velocity was observed that<br />

demonstrated accelerated movement of liquid plugs in wider flow<br />

channels. The presence of compartments positioned along the flow<br />

channel exerts an additional similar influence. 34,35 Therefore,<br />

rectangular compartments with dimensions of 1.5 mm × 880 µm<br />

and 6.0 mm × 3.5 mm were attached to the 250 µm wide and 1.0<br />

mm wide flow channels (flow channels 1 and 5, respectively). A<br />

portion of the solution, however, also penetrated into the extending<br />

controlling flow channel while the solution filled each compartment.<br />

As a result, the movement of the column was not<br />

significantly different from the case in which there were no<br />

compartments. This result indicated that branched compartments<br />

are not effective for this purpose.<br />

We then tried a sequential arrangement. Figure 3B shows 500<br />

µm wide flow channels. Flow channel 1 is straight, and flow<br />

channels 2 and 3 have compartments of different sizes (2.0 mm ×<br />

2.0 mm and 3.5 mm × 3.5 mm, respectively). By locating the exit<br />

at an appropriate position in the compartment, the transport in<br />

the flow channel was resumed after the compartment was filled<br />

completely, and the effect of the structures was more significant<br />

than that in Figure 3A. There was a tendency for bubbles to<br />

remain in the corners of the square compartments. Although these<br />

bubbles were small and had no adverse effect on the transport of<br />

solutions, circular or elliptic compartments might be better, both<br />

to avoid this potential problem and to realize a more accurate<br />

adjustment of timing.<br />

The movement of the column could also be delayed by the<br />

introduction of constrictions. The width of the flow channels in<br />

Figure 3C is 500 µm. For flow channels 2 and 3, the constrictions<br />

were positioned near the inlet and had widths of 300 and 200 µm,<br />

respectively. The constrictions also had an effect, and the movement<br />

of the column was slowed with narrower constrictions.<br />

Although we used only simple delaying structures because of the<br />

limited space, microfluidic transport can be delayed further<br />

through the use of a more complicated network of flow channels. 22<br />

The movement of a solution in a flow channel with compartments<br />

and/or constrictions can be understood using numerical<br />

Figure 3. Movement of solutions in the flow channels with various delaying structures. (A) Effect of changing the width of the flow channel and<br />

attaching rectangular compartments on the sides. (B) Effect of adding rectangular compartments in series. (C) Effect of creating constrictions.<br />

The images were taken 1, 3, and 6 s (from left to right) after the introduction of the solution from the left. Scale bars correspond to 2 mm.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6873


simulations, 34,35 which can be helpful in the design of controlling<br />

flow channels that punctually provide H2O2 solution to micropumps<br />

according to a predetermined schedule.<br />

Autonomous Sequential Switching of the Micropumps.<br />

We next studied the function of the micropump. After the lower<br />

compartment of the pump was filled with H2O2 solution, the<br />

diaphragm inflated and the solution in the upper reservoir was<br />

ejected and then mobilized in the extending flow channel.<br />

During this step, the H2O2 solution moved forward in the<br />

controlling flow channel and filled the lower compartment of<br />

the next pump. From that point forward, the same step was<br />

repeated. Needless to say, the concentration of H2O2 injected<br />

into the controlling flow channel affects the flow velocity of<br />

the solution in the upper flow channel. Although the flow<br />

velocity definitely depends on the H2O2 concentration, significant<br />

leakage and destruction of the gel layer was observed over<br />

a certain threshold, which also depended on the size of the<br />

pump. Considering these dynamics, 1.6 or 3.2 M H2O2 was used<br />

in the following experiments, depending on the size of the<br />

pump.<br />

For the handling of many solutions, several micropumps can<br />

be connected to the controlling flow channel and to the upper<br />

flow channels for solutions to be transported. To adjust the timing<br />

and order of the injection of solutions, the distance between pumps<br />

can be adjusted in a network of flow channels. Compartments and<br />

constrictions can also be introduced, as was discussed earlier. In<br />

the device shown in parts A and B of Figure 4, micropumps are<br />

connected with one another by rectangular compartments at their<br />

sides and centers, respectively. In the device shown in Figure<br />

4B, the compartments are filled with solution after the compartment<br />

of a neighboring micropump in the upper stream is filled<br />

with solution. With the small compartments in the crowded layout,<br />

however, the extension of the liquid columns from the upper<br />

reservoirs was not so significant compared with that of the device<br />

shown in Figure 4A. The device shown in Figure 4C has additional<br />

small compartments to delay the movement of the solution. Note<br />

the difference in the length of the liquid columns, which shows<br />

that the actuation of the micropump can be distinctly delayed in<br />

the lower stream of the controlling flow channel, unlike the device<br />

shown in Figure 4A and 4B. This result demonstrates that<br />

additional delaying structures can in fact be used to adjust the<br />

timing to trigger the actuation of the micropumps.<br />

Coordinated Operation of Micropumps for <strong>Chemical</strong><br />

Analyses. By properly designing a network of flow channels with<br />

micropumps located in appropriate positions, microfluidic devices<br />

can be constructed for various purposes, including chemical<br />

analyses that require the specific processing of solutions. In the<br />

device shown in Figure 5, there are two micropumps to eject<br />

solutions and two pumps to apply pressure. Here, H 2O2 was also<br />

used as an analyte. A solution containing H2O2 and Amplex<br />

Red (5 mM) and another solution containing 50 U/mL HRP,<br />

both prepared with a 50 mM Tris-HCl buffer solution (pH 7.4),<br />

were used to fill the reservoirs. After another H2O2 solution<br />

was injected into the controlling flow channel, it first filled the<br />

compartments of the micropumps. Following this, the analyte<br />

H2O2 solution in the reservoir was injected into the upper flow<br />

(34) Erickson, D.; Li, D.; Park, C. B. J. Colloid Interface Sci. 2002, 250, 422–<br />

430.<br />

(35) Young, W.-B. Colloids Surf., A 2004, 234, 123–128.<br />

6874 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 4. Arrays of micropumps with controlling flow channels with<br />

different structures. Rectangular compartments are connected with<br />

the controlling flow channels at the edges (A), at the center (B), and<br />

via smaller delaying compartments (C).<br />

channels for transport after 9 s (Figure 5B, panels 1 and 2).<br />

After a time delay (117 s), the larger pumps that apply pressure<br />

were switched on and the solutions in the upper flow channel<br />

were pinched off from the rest of the solutions, transported to<br />

the center, and merged in the mixing channel (Figure 5B, panels<br />

3 and 4). The average flow velocity in the mixing channel was 94<br />

µm/s. Solutions containing the enzyme and the substrates were<br />

transported and mixed in the flow channel. The enzymatic reaction<br />

by HRP produced highly fluorescent resorufin, which generated<br />

red fluorescence under a fluorescence microscope. Figure 5C<br />

shows the dependence of the fluorescence intensity on the<br />

concentration. The fluorescence intensity was measured 3 min<br />

after mixing. In the graph plotted on the semilog scale, the<br />

dependence of the fluorescence intensity on the concentration<br />

could be clearly observed.


Figure 5. Device to eject and mix solution plugs. (A) Layout of<br />

the controlling flow channels (dashed line, shaded) and the flow<br />

channels for transport (solid line). (B) Movement of a H2O2 solution<br />

in the controlling flow channels (dashed arrows) and an analyte<br />

H2O2 solution in flow channels for the transport of solutions (solid<br />

arrows). (1) The reservoirs of the micropumps were filled with<br />

solutions. (2) The solutions were injected into the main channel.<br />

(3, 4) The solutions were transported by the exertion of pressure<br />

from the leftmost and rightmost pumps and were merged at the<br />

T-junction. (C) Dependence of the fluorescence intensity on the<br />

concentration of H2O2 ejected from a micropump. Five runs were<br />

performed, and the averages and standard deviations are shown.<br />

The dashed lines show the average +3σ of the background<br />

fluorescence. The concentration of the H2O2 solution injected into<br />

the controlling flow channel was 1.6 M. The dimensions of the chip<br />

were 25 mm × 16 mm.<br />

More complicated manipulation of solutions could also be<br />

carried out. The device shown in Figure 6 has three injection<br />

pumps and three micropumps to exert pressure upon the ejected<br />

plugs. Three lines of controlling flow channels were used. In each<br />

channel, a pump to eject a necessary solution and a pump to apply<br />

pressure were connected. The lower compartment of each<br />

injection pump was used to delay the arrival of the H2O2 solution<br />

to the pumps to apply pressure located in the lower stream.<br />

GOD and LOD were immobilized at the bottom of the<br />

reservoirs of pumps A and C. First, we used solutions containing<br />

either glucose or lactate. The reservoir of pump A was filled<br />

with a phosphate buffer solution containing glucose and Amplex<br />

Red (5 mM), and the reservoir of pump C was filled with<br />

another phosphate buffer solution containing lactate and<br />

Amplex Red (5 mM). The reservoir of pump B was filled with<br />

a phosphate buffer solution to wash the reaction chamber. The<br />

enzymatic reactions by the oxidases produced H2O2. In view<br />

of the volume of the solutions and the activity of the immobilized<br />

enzymes, it could be assumed that the enzymatic<br />

conversion was virtually completed in the examined ranges of<br />

concentration of glucose and lactate during this preparatory<br />

period. Three minutes after the filling with solutions, the H2O2<br />

solution was introduced into the controlling flow channel.<br />

Following this, a row of plugs was injected into the main flow<br />

channel from the reservoirs of the pumps located in the lower<br />

stream. In accordance with the flow channel design, pumps<br />

A-F were switched on 28, 85, 174, 187, 259, and 336 s after<br />

the injection of the H2O2 solution into the controlling flow<br />

channel. The velocity of a plug passing through the reaction<br />

chamber was 35 µm/s, the same velocity at which the three<br />

pumps would apply pressure. When the first plug containing<br />

Amplex Red and H2O2 produced by GOD arrived at the reaction<br />

chamber, fluorescence was generated in the chamber, as it was<br />

in the previously mentioned device, and its intensity was<br />

measured 90 s after the solution reached the chamber. After<br />

the reaction chamber was rinsed with a rinsing plug that was<br />

ejected from pump B, the last plug containing Amplex Red and<br />

H2O2 produced by LOD was introduced into the reaction<br />

chamber, and the fluorescence intensity was measured as<br />

before. The same experiment was carried out using the same<br />

solution containing glucose, lactate, and Amplex Red to fill<br />

pumps A and C. Figure 7 shows the calibration plots obtained<br />

using solutions containing only glucose or lactate and the data<br />

points obtained for solutions containing both of them. For the<br />

limited ranges of concentration that were employed, the plots were<br />

apparently linear. The values obtained for solutions containing<br />

both glucose and lactate in different combinations of concentration<br />

came close to the calibration plots, which demonstrated that the<br />

device can be used for the analysis of different analytes that coexist<br />

in the same solution.<br />

We note again that no wires or tubes were deployed around<br />

the chip to apply electrical signals or pressure. The necessary<br />

reagent solutions can be stored in the injection ports beforehand.<br />

The only thing that must be done is to introduce a sample solution<br />

and initiate the flow of the H2O2 solution in the controlling flow<br />

channel. The H2O2 solution can also be stored in a separate<br />

reservoir and can then be injected by applying a small degree<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6875


Figure 6. Enzyme analysis that accompanies the manipulation of<br />

three plugs. (A) Layout of the controlling flow channels (dashed line,<br />

shaded) and flow channels for transport (solid line). For purposes of<br />

clarity, the two flow channel networks are drawn separately in the<br />

lower figure. (B) Fluorescence images showing the movement of the<br />

H2O2 solution in the network of controlling flow channels (dashed<br />

arrows) and solution plugs in the network of upper flow channels (solid<br />

arrows). (1) The first solution was transported to the reaction chamber.<br />

(2) After flushing of the first solution, the reaction chamber was<br />

washed with the second plug. (3) After flushing of the solution, the<br />

third solution was transported to the reaction chamber. Although the<br />

flows in the controlling flow channels are described separately, they<br />

began to flow in the flow channels simultaneously. The concentration<br />

of the H2O2 solution injected into the controlling flow channel was<br />

3.2 M. The dimensions of the chip were 33 mm × 27 mm.<br />

6876 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 7. On-chip detection of glucose and lactate using the device<br />

shown in Figure 6. Dependence of the fluorescence intensity on the<br />

concentration of glucose (O, 0) and lactate (b, 9). O and b indicate<br />

data obtained in experiments using solutions that contained only<br />

glucose or lactate. Five runs were performed, and the averages and<br />

standard deviations are shown. 0 and 9 indicate data obtained by<br />

filling the injection ports of pumps A and C with solutions containing<br />

both glucose and lactate. The inset shows the plot at a lower<br />

concentration range near the detection limits. The dashed lines show<br />

the average +3σ of the background fluorescence.<br />

of pressure that enables it to pass through a hydrophobic valve<br />

set at the entrance.<br />

CONCLUSIONS<br />

<strong>Chemical</strong>ly actuated micropumps can be realized by making<br />

use of the volume change produced by the catalytic decomposition<br />

of H2O2. The pumps are located along a controlling flow channel<br />

that transports a H2O2 solution via capillary action. The row of<br />

pumps can be switched on sequentially following the introduction<br />

of the H2O2 solution into the controlling flow channel. The<br />

structure of the pump itself can also be used to exert pressure<br />

upon a solution in a flow channel. The timing of the switching<br />

among pumps can be adjusted by locating them at appropriate<br />

positions in a network of flow channels or by employing<br />

additional structures such as compartments and/or constrictions.<br />

In other words, the information for switching among<br />

pumps is directly described on the chip as a program.<br />

As demonstrated, one potential application for our autonomous<br />

devices is that of portable analysis systems. Although on-chip<br />

biochemical analyses for molecules such as proteins have already<br />

been carried out using integrated microfluidic components, 6,27,28<br />

this technique will simplify the construction of the entire system.<br />

Moreover, such an autonomous device can be useful for a variety<br />

of purposes, such as micromixing, 23 tissue culturing, 19 and<br />

gas-liquid reactions, 17 since it can minimize the burden involved<br />

in the handling of solutions.<br />

Received for review April 12, 2010. Accepted July 8, 2010.<br />

AC1009657


Anal. Chem. 2010, 82, 6877–6886<br />

Immunoaffinity Purification Using Anti-PEG<br />

Antibody Followed by Two-Dimensional Liquid<br />

Chromatography/Tandem Mass Spectrometry for<br />

the Quantification of a PEGylated Therapeutic<br />

Peptide in Human Plasma<br />

Yang Xu,* John T. Mehl, † Ray Bakhtiar, † and Eric J. Woolf<br />

Department of Drug Metabolism and Pharmacokinetics, Regulated Bioanalysis, Merck Research Laboratories,<br />

WP75B-300, West Point, Pennsylvania 19486<br />

Quantification of a PEGylated peptide in human plasma<br />

using LC-MS/MS to support clinical studies presented<br />

challenges in terms of assay sensitivity, selectivity, and<br />

ruggedness. To ensure specific recognition of PEGylated<br />

species, an immunoaffinity purification method (IAP)<br />

using anti-PEG antibody followed by two-dimensional<br />

(2D) LC-MS/MS was developed for MK-2662, an investigational<br />

peptide containing 38 amino acids with a 40<br />

kDa branched PEG [poly(ethylene glycol)] at C-terminus.<br />

Biotinylated anti-PEG antibody, bound to streptavidincoated<br />

magnetic beads, was used to capture MK-2662<br />

and its stable-isotope-labeled internal standard from human<br />

plasma. After on-bead digestion with trypsin, the<br />

supernatant was injected on a 2D high-performance liquid<br />

chromatography (HPLC) system constructed with strong<br />

cation-exchange and reversed-phase columns, followed by<br />

MS/MS detection of the surrogate N 1-12-mer of MK-<br />

2662 on an API5000. The assay ruggedness was<br />

improved by optimizing the trypsin digestion and<br />

sample storage conditions. The intraday validation,<br />

conducted in parallel with protein precipitation (PPT)<br />

assay, demonstrated 94.8-105.8% accuracy with<br />


Figure 1. <strong>Chemical</strong> structures of MK-2662 and its internal standard,<br />

[ 13 C18, 15 N2] MK-2662 (ISTD). The underlined region indicates the<br />

N-terminal tryptic peptide used as surrogate for quantification.<br />

is time-consuming; therefore, reagent availability could become<br />

a rate-limiting step. In the case of MK-2662 analysis, ELISA was<br />

abandoned because a drug-specific antibody was not available at<br />

the time of assay development.<br />

Bioassay is a cell-based or enzyme-based assay that measures<br />

the biologic activity of a specific biological process. 7 Since bioassay<br />

measures the total activity of the sample irrespective of the<br />

chemical structures, the assay specificity is usually less than that<br />

of ELISA. It could provide a measure of pharmacodynamic (PD)<br />

properties 7 but may not be appropriate for pharmacokinetic (PK)<br />

assessment.<br />

Liquid chromatography-mass spectrometry (LC-MS) has<br />

been increasingly used to quantify peptides and proteins in<br />

biological matrixes 9-11 because of its selectivity. Despite commonly<br />

used sample cleanup techniques, such as protein precipitation<br />

(PPT), 9 solid-phase extraction (SPE), 12-14 two-dimensional<br />

(2D) SPE, 15 and 2D high-performance liquid chromatography<br />

(HPLC), 16,17 the immunoaffinity purification (IAP) coupled with<br />

LC-MS/MS represents an emerging strategy for peptide and<br />

protein bioanalysis. The IAP strategies include immunoaffinity<br />

depletion that can be used to remove abundant proteins from<br />

biological matrixes 18,19 and immunoaffinity capture that utilizes<br />

a single antibody to isolate and enrich the target peptides or<br />

(9) van den Broek, I.; Sparidans, R. W.; Schellens, J. H. M.; Beijnen, J. H.<br />

J. Chromatogr., B 2008, 872, 1–22.<br />

(10) Careri, M.; Mangia, A. J. Chromatogr., A 2003, 1000, 609–635.<br />

(11) John, H.; Walden, M.; Schafer, S.; Genz, S.; Forssanmm, W. G. Anal.<br />

Bioanal.Chem. 2004, 378, 883–897.<br />

(12) van den Broek, I.; Sparidans, R. W.; Huitema, A. D. R.; Schellens, J. H. M.;<br />

Beijnen, J. H. J. Chromatogr., B 2006, 837, 49–58.<br />

(13) van den Broek, I.; Sparidans, R. W.; Huitema, A. D. R.; Schellens, J. H. M.;<br />

Beijnen, J. H. J. Chromatogr., B 2007, 854, 245–259.<br />

(14) Heudi, O.; Barteau, S.; Zimmer, D.; Schmidt, J.; Bill, K.; Lehmann, N.; Bauer,<br />

C.; Kretz, O. Anal. Chem. 2008, 80, 4200–4207.<br />

(15) Yang, Z.; Hayes, M.; Fang, X.; Daley, M. P.; Ettenberg, S.; Francis, L. S. T.<br />

Anal. Chem. 2007, 79, 9294–9301.<br />

(16) Motoyama, A.; Xu, T.; Ruse, C. I.; Wohlschlegel, J. A.; Yates, J. R. Anal.<br />

Chem. 2007, 79, 3623–3634.<br />

(17) Linke, T.; Ross, A. C.; Harrison, E. H. J. Chromatogr A. 2006, 1123, 160–<br />

169.<br />

6878 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

proteins from biological samples. Although many applications<br />

using online immunoaffinity columns have been reported, 20-24<br />

offline affinity purification using different carriers, such as<br />

macroporous polymeric beads, agarose or sepharose beads, 25,26<br />

and magnetic beads, 27-29 etc., for immobilization of a variety of<br />

enzyme and/or capture antibodies enables more flexibility in<br />

terms of selection of carriers, antibodies, assay formats, and<br />

experimental conditions as compared to the online approach.<br />

This report presents a sensitive and reproducible bioanalytical<br />

approach for quantification of a PEGylated peptide in plasma. To<br />

support human clinical studies for MK-2662, a 2D HPLC-MS/<br />

MS method using a turbo ion spray (TIS) interface, monitoring<br />

the surrogate N-terminal peptide (HAibDGTFTSDYSK) of MK-<br />

2662 after tryptic digestion, has been developed and validated to<br />

quantify MK-2662 in human plasma. A stable-isotope-labeled<br />

internal standard, [ 13 C18, 15 N2] MK-2662, was employed. Two<br />

sample preparation methods, PPT versus IAP using an anti-<br />

PEG antibody, followed by trypsin digestion in a 96-well format<br />

using 0.2 mL plasma sample, have been evaluated and applied<br />

to clinical sample analysis. The potential caveats of the protein<br />

precipitation approach and the benefit of immunoaffinity<br />

purification in regard to assay specificity are discussed. To our<br />

best knowledge, this is the first report describing use of anti-<br />

PEG antibody to capture a PEGylated peptide in combination<br />

with LC-MS/MS analysis. This approach is potentially applicable<br />

to analysis of other PEGylated peptides or proteins.<br />

EXPERIMENTAL SECTION<br />

<strong>Chemical</strong>s and Reagents. MK-2662 and stable-isotope-labeled<br />

internal standard (ISTD, Figure 1) were synthesized at the Merck<br />

Research Laboratories, Merck & Co. (Rahway, NJ). HPLC grade<br />

acetonitrile, HPLC grade methanol, laboratory grade formic acid<br />

(88%), certified ammonium formate, and certified ammonium<br />

bicarbonate were obtained from Fisher Scientific (Pittsburgh, PA).<br />

Human control plasma (K2EDTA as anticoagulant) was purchased<br />

from Biological Specialty Co. (Colmar, PA). Water was<br />

purified by a Milli-Q ultrapure water system from Millipore<br />

(Bedford, MA). Bovine serum albumin (BSA) at 22% in 0.85%<br />

(18) Hagman, C.; Ricke, D.; Ewert, S.; Bek, S.; Falchetto, R.; Bitsch, F. Anal.<br />

Chem. 2008, 80, 1290–1296.<br />

(19) Dekker, L. J.; Bosman, J.; Burgers, P. C.; van Rijswijk, A.; Freije., R.; Luider,<br />

T.; Bischoff, R. J. Chromatogr., B 2007, 847, 65–69.<br />

(20) Radabaugh, M. R.; Nemirovskiy, O. V.; Misko, T. P.; Aggarwal, P.; Rodney<br />

Mathews, W. Anal. Biochem. 2008, 380, 68–76.<br />

(21) Li, W. W.; Nemirovskiy, O.; Fountain, S.; Rodney Mathews, W.; Szekely-<br />

Klepser, G. Anal. Biochem. 2007, 369, 41–53.<br />

(22) Berna, M.; Schmalz, C.; Duffin, K.; Mitchell, P.; Chambers, M.; Ackermann,<br />

B. Anal. Biochem. 2006, 356, 235–243.<br />

(23) Zhang, X.; Martens, D.; Kramer, P. M.; Kettrup, A. A.; Liang, X. J. Chromatogr.,<br />

A 2006, 1133, 112–118.<br />

(24) Edinboro, L. E.; Karnes, H. T. J. Chromatogr., A 2005, 1083, 127–132.<br />

(25) Huang, L.; Harvie, G.; Feitelson, J. S.; Gramatikoff, K.; Herold, D. A.; Allen,<br />

D. L.; Amunngama, R.; Hagler, R. A.; Pisano, M. R.; Zhang, W. W.; Fang,<br />

X. Proteomics 2005, 5, 3314–3328.<br />

(26) Hoofnagle., A. N.; Becker, J. O.; Wener, M. H.; Heinecke, J. W. Clin. Chem.<br />

2008, 54, 1796–1804.<br />

(27) Lee, Y. C.; Block, G.; Chen, H.; Folch-Puy, E.; Foronjy, R.; Jalili, R.;<br />

Jendresen, C. B.; Kimura, M.; Kraft, E.; Lindemose, S.; Lu, J.; McLain, T.;<br />

Nutt, L.; Ramon-Garcia, S.; Smith, J.; Spivak, A.; Wang, M. L.; Zanic, M.;<br />

Lin, S. H. Protein Expression Purif. 2008, 62, 223–229.<br />

(28) Dubois, M.; Becher, F.; Herbet, A.; Ezan, E. Rapid Commun. Mass Spectrom.<br />

2007, 21, 352–358.<br />

(29) Berna, M. J.; Zhen, Y.; Watson, D. E.; Hale, J. E.; Ackermann, B. L. Anal.<br />

Chem. 2007, 79, 4199–4205.


NaCl, phosphate buffer saline (PBS), phosphate buffer saline<br />

with 0.05% Tween-20 pH 7.4 (PBST), and Triton-X 100 were<br />

purchased from Sigma-Aldrich (Milwaukee, WI). Linco DPP-<br />

IV inhibitor was purchased from Millipore Corporation (St.<br />

Charles, MO). BD P700 blood collection tubes that contain<br />

proprietary DPP-IV inhibitors were purchased from BD Bioscience<br />

(Franklin Lakes, NJ). Lyophilized sequence grade<br />

modified trypsin (TRSEQZ at 1 mg/vial) was purchased from<br />

Worthington Biochemical Corporation (Lakewood, NJ). Dynabeads<br />

M-280-streptavidin was purchased from Invitrogen Corporation<br />

(Carlsbad, CA). Biotinylated anti-PEG rabbit monoclonal<br />

antibody, anti-PEG-biotin (or PEG-B-47-bio), was purchased<br />

from Epitomics, Inc. (Burlingame, CA).<br />

Instrumentation. Dynal MPC-L and MPC-96 (Invitrogen<br />

Corporation, Carlsbad, CA) were used to capture magnetic beads<br />

in the vial and 96-well plate, respectively. A TomTec Quadra 96<br />

workstation, model 320 (TomTec Corporation, Hamden, CT) was<br />

used to perform automated liquid transfer for protein precipitation.<br />

A MaxQ mini 4450 benchtop incubating orbital shaker (Thermo<br />

Scientific, Waltham, MA) and an SPE-Dry 96 micropate sample<br />

concentrator (Jones Chromatography, Lakewood, CO) were used<br />

for sample preparation. A Cohesive Aria 2300 system (Cohesive<br />

Technologies Inc., Franklin, MA), which included two quaternary<br />

Flux pumps, a valve module, and a CTC HTS autosampler, was<br />

used for 2D HPLC. A Sciex API 5000 triple-quadrupole mass<br />

spectrometer with a Sciex turbo ion spray interface (Sciex,<br />

Toronto, Canada) was used as a detector. The data were collected<br />

and processed through Analyst 1.4 software.<br />

Preparation of Standards and Quality Control (QC)<br />

Samples. MK-2662 stock solutions at about 40 µM were prepared<br />

from two separate weighings and dissolved in acetonitrile/water<br />

(30/70, v/v). One set of analyte stock solutions was used to<br />

prepare calibration standards, and the other set was used to make<br />

QC samples. Working standards, containing MK-2662 at the<br />

different concentration levels, were prepared by serial dilutions<br />

of analyte stock solution with 1% BSA, aliquoted to 1.5 mL<br />

polypropylene microcentrifuge tubes and stored at -70 °C. A 500<br />

nM ISTD working solution was obtained by dilution of an ∼40<br />

µM ISTD stock solution in 1% BSA. Plasma calibration standards<br />

were prepared daily by adding 40 µL of working standard and 40<br />

µL of 500 nM IS into 200 µL of control plasma to provide final<br />

concentrations of MK-2662 ranging from 1.0 to 1000 nM for the<br />

PPT process and 2-200 nM for the IAP process.<br />

The plasma QC samples were prepared in human control<br />

plasma that contained Linco DPP-IV inhibitor (20 µL inhibitor<br />

solution per milliliter of plasma) at 3, 50, 800, and 10 000 nM (for<br />

testing dilution integrity) MK-2662 for the PPT assay and 3, 6,<br />

50, and 150 nM for IAP. All QC aliquots were stored in a -70 °C<br />

freezer.<br />

Sample Preparation: Protein Precipitation Followed by<br />

Trypsin Digestion. Clinical samples and QCs were thawed at<br />

room temperature, mixed, and centrifuged at 4000 rpm (∼1300 g<br />

RCF), at 10 °C for 10 min. A 200 µL aliquot of sample was then<br />

mixed well with 40 µL of 1% BSA (to match the volume of<br />

standards), 40 µL of 500 nM ISTD, and 500 µL of 0.1% formic<br />

acid in acetonitrile/methanol (90/10, v/v) ina2mL96-well deepwell<br />

plate. After centrifuging at 2000 RCF for 5 min, 600 µL of<br />

supernatant was transferred into a clean 1.2 mL 96-well plate using<br />

a TomTec automated liquid handling system and dried on an SPE<br />

Dry-96 under a stream of nitrogen at 40 °C. After resuspending<br />

the residue in with 140 µL of 50 mM ammonium bicarbonate (pH<br />

8.0) containing 67 µg/mL trypsin, the samples were incubated at<br />

37 °C for3hinaMaxQmini shaker to allow trypsin digestion,<br />

and then the reaction was quenched by adding 10 µL of 44% formic<br />

acid to each well at the end of incubation. After mixing and<br />

centrifugation, 20 µL of sample was injected into the LC/LC-MS/<br />

MS for analysis.<br />

Sample Preparation: Immunoaffinity Capture Using Anti-<br />

PEG Antibody Followed by On-Bead Trypsin Digestion.<br />

DynaBeads M280-steptavidin (magnetic beads) were washed with<br />

PBST (pH 7.4) three times and resuspended into the same volume<br />

of PBST. Biotinylated anti-PEG rabbit monoclonal antibody, anti-<br />

PEG-biotin (or PEG-B-47-bio), was added to the bead suspension<br />

to reach the concentration of 20 µg/mL and then gently shaken<br />

at room temperature for 1htoallow binding of biotinylated<br />

antibody to the steptavidin-coated bead. The beads with antibody<br />

were washed and resuspended with PBST three times. A 50 µL<br />

final suspension was added into a sample mix that contained 200<br />

µL of plasma sample, 40 µL of 1% BSA (to match the volume for<br />

calibration standards), and 40 µL of 500 nM ISTD in a 96-well<br />

plate. The sample mix was shaken at 700 rpm at 30 °C for2hto<br />

allow antibody capture of MK-2662 from the plasma sample. After<br />

washing the beads with PBS (pH 7.4) that contained 0.1% Triton-X<br />

100 and then with ammonium bicarbonate (pH 8), an aliquot of<br />

140 µL of67µg/mL trypsin in 50 mM ammonium bicarbonate<br />

(pH 8) was added into each well, incubated at 37 °C with shaking<br />

for about 3 h. The supernatant was transferred into a 96-well plate<br />

with a 600 µL insert in each well, where 10 µL of 44% formic acid<br />

was added to quench the reaction. After mixing and centrifugation,<br />

20 µL of sample was injected on LC/LC-MS/MS for analysis.<br />

Two-Dimensional HPLC Conditions. The 2D HPLC was<br />

conducted with three basic steps: loading, eluting, and separating,<br />

using a dual-column quick-elution mode on a Cohesive Aria<br />

system. A cation-exchange column, Thermo BioBasic SCX (50 mm<br />

× 2.1 mm, 5 µm), was used as the first dimension with a salt<br />

gradient of mobile phases A and B [A, 50 mM ammonium formate<br />

(pH 3) in 10% acetonitrile; B, 200 mM ammonium formate (pH 3)<br />

in 10% acetonitrile]. A reversed-phase column, Thermo-Hypersil<br />

Gold PFP (50 mm × 2.1 mm, 5 µm), was used as the second<br />

dimension with an organic solvent gradient of mobile phases C<br />

and D [C, 10 mM ammonium formate in 0.1% formic acid; D, 10<br />

mM ammonium formate in 90% acetonitrile with 0.1% formic acid].<br />

The LC method on the Cohesive system is listed in the Supporting<br />

Information Table S1. The compartment of the autosampler was<br />

set at 5 °C, and the reversed-phase column was controlled at 40<br />

°C.<br />

Mass Spectrometry Detection and Calculation. A PE Sciex<br />

API 5000 triple-quadrupole mass spectrometer (MS/MS) with a<br />

turbo ion spray interface ionization source operated in a positive<br />

ion mode was used to quantitate MK-2662. The ion pairs (precursor<br />

ion f product ion) m/z 672.1 f 223.1 for MK-2662 and m/z<br />

676.9 f 223.1 for ISTD were selected for multiple reaction<br />

monitoring (MRM). The instrument setting was adjusted to<br />

maximize the response for the analyte and IS, respectively. The<br />

turbo gas temperature was 650 °C. The flow settings of nebulizing<br />

gas (nitrogen), collision gas (nitrogen), and curtain gas (nitrogen)<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6879


were 50, 6, and 30 L/min. The optimized declustering potential<br />

(DP), collision energy (CE), collision cell exit potential (CXP),<br />

and entrance potential (EP) were 140, 36, 15, and 10 V. The dwell<br />

time was 100 ms for MK-2662 and ISTD. Both Q1 and Q3<br />

quadrupoles were set at unit resolution. The total run time for<br />

each injection was 4.5 min, while the MS data acquisition window<br />

was started at 100 s after injection and kept open for 50 s. Peak<br />

area ratios were calculated using Analyst software version 1.4. A<br />

calibration curve was obtained by weighed (1/x 2 ) least-squares<br />

linear regression of the peak area ratio of the analyte to the IS<br />

versus the nominal concentration (x) of analyte.<br />

Method Validation. The selectivity of the assay was confirmed<br />

by processing control plasma from six different lots. Intraday<br />

precision and accuracy were determined by analyzing six sets of<br />

standard curve samples, each prepared in a different lot of control<br />

plasma. Assay accuracy was calculated from a least-squares<br />

regression curve constructed using all six replicate values at each<br />

concentration, and the intraday precision (% CV) was calculated<br />

from the peak area ratio of MK-2662 versus ISTD for each<br />

concentration used to construct the standard curve. QC samples<br />

were analyzed after first freezing and thawing, and the measured<br />

concentrations were considered as the initial values. Freeze-thaw<br />

stability was evaluated using QC samples that went through three<br />

cycles of freezing and thawing, with at least one day of storage at<br />

-70 °C between each thawing. Benchtop QC stability was tested<br />

following 5hatroom temperature and comparing the measured<br />

concentrations with their initial values. The stability of processed<br />

samples in the autosampler was assessed by comparing the results<br />

of QC samples analyzed at the end of a 14 h run with those<br />

analyzed at the beginning of the run. The reinjection stability was<br />

assessed by freezing the processed sample plate at -70 °C and<br />

reinjected after 8 days of storage. In order to examine the dilution<br />

integrity, five replicates of 10-fold of upper limit of quantification<br />

(ULOQ) were diluted by 20-fold with the corresponding control<br />

matrix during sample preparation and analyzed on LC/LC-MS/<br />

MS.<br />

RESULTS AND DISCUSSION<br />

Mass Spectrometry Detection of Surrogate Peptide of MK-<br />

2662. Tandem mass spectrometry (MS/MS) has been increasingly<br />

used for peptide quantification in biological matrixes because<br />

of its specificity. However, in the case of MK-2662, the PEGylated<br />

peptide exhibited a complex mass spectrum obtained using turbo<br />

ion spray (TIS), due to the large molecular weight (40 kDa) and<br />

oligomeric dispersity of the conjugated PEG polymer. DePEGylation<br />

of MK-2662 could be achieved by base hydrolysis and/or<br />

enzymatic hydrolysis. For MK-2662, base hydrolysis liberated the<br />

peptide via cleavage of the linker moiety. However, the MRM<br />

sensitivity of the dePEGylated peptide was low due to weak<br />

collision-induced dissociation (CID) of highly charged precursor<br />

ions. Additionally, base hydrolysis could add a risk of chemical<br />

degradation such as deamidation or hydrolysis to the peptide.<br />

Because of these limitations of base hydrolysis, enzymatic hydrolysis<br />

was selected to remove the PEG portion from MK-2662<br />

prior to MS/MS analysis.<br />

The N-terminus of MK-2662 is required for activity, and more<br />

importantly, it contains the non-native amino acid (Aib) substitution<br />

at the second amino acid in the peptide sequence, providing<br />

mass selectivity against possible low levels of endogenous oxyn-<br />

6880 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

tomodulin. Digestion of MK-2662 with trypsin produces a 12 amino<br />

acid N-terminal peptide (N 1-12-mer), HAibDGTFTSDYSK, and<br />

it was selected as a surrogate for quantification of MK-2662.<br />

In comparison to the other tryptic cleavage sites in the MK-<br />

2662 sequence, the single cleavage reaction at the 12th amino<br />

acid (to generate N1-12-mer) demonstrated more rapid and<br />

reproducible digestion. An internal standard was synthesized<br />

by preparing a stable-isotope-labeled MK-2662 (PEGylated) that<br />

contained 13 C/ 15 N-labeled phenylalanine at the sixth amino acid<br />

in the sequence. The use of a chemically matched IS resulted<br />

in improved assay precision by compensating for variation<br />

during the sample preparation and detection process.<br />

Precursor ions for the surrogate N1-12-mers of MK-2662 and<br />

its IS were monitored as the doubly charged molecular weight<br />

(MW) ions, [M + 2H] 2+ ,atm/z 672.1 (MW ) 1342) and 676.9<br />

(MW ) 1352), respectively. After optimizing CID, CE, and DP<br />

values, the product ion spectra of the N1-12-mers were collected,<br />

showing singly charged b- and y-ions (Figure 2). The b2 ion at<br />

223 m/z was selected based on its ion intensity and specificity,<br />

where a non-natural amino acid (Aib) in b2 could allow MS<br />

differentiation of MK-2662 from the endogenous OXM fragments.<br />

Therefore, the MRM at m/z 672 f 223 (MK-2662) and<br />

m/z 677 f 223 (ISTD) was used for quantification of MK-2662.<br />

Two-Dimensional Chromatographic Conditions. One major<br />

challenge for supporting clinical studies was the assay sensitivity<br />

requirement. Our initial effort on developing a traditional onedimensional<br />

(1D) reversed-phase HPLC-MS/MS assay allowed<br />

us to successfully achieve a lower limit of quantification (LLOQ)<br />

of 5 nM using protein precipitation followed by tryptic digestion<br />

(Figure 3A). Since the clinical assay required a LLOQ of 1 nM,<br />

addition assay development efforts were focused on reducing<br />

background and increasing signal-to-noise ratio.<br />

An attempt to use SPE to extract full length of MK-2662 from<br />

plasma was not successful (very low recovery), mostly likely due<br />

to the fact that the large PEG portion blocked or interfered with<br />

the interaction of the peptide with the SPE sorbent. Other<br />

approaches, such as optimizing the sample recovery and digestions<br />

conditions or evaluating different CID transitions, were<br />

unsuccessful toward obtaining an LLOQ below 5 nM, due to high<br />

background signal in the LC-MS/MS chromatogram.<br />

Multidimensional protein identification technology<br />

(MudPIT) 30-33 has been widely used in the proteomics area.<br />

Applying the same concept, a 2D LC-MS/MS was developed to<br />

separate a surrogate peptide of MK-2662 (after trypsin digestion)<br />

from the background using two different separating chromatographic<br />

mechanisms. The 2D HPLC was performed on a Cohesive<br />

Flux 2300 under quick elution mode with built-in switching valves.<br />

A strong cation-exchange column, BioBasic SCX (50 mm × 2.1<br />

mm, 5 µm), was used as the first dimension. An ∼60 s salt gradient<br />

from 50 to 200 mM ammonium formate was used at a flow rate of<br />

0.5 mL/mL to maintain a net positive charge on the N-terminal<br />

tryptic peptide and separate N1-12-mer from other charged<br />

components in the digest mix. A small percentage of organic<br />

(30) Lohrig, K.; Wolters, D. Methods Mol. Biol. 2009, 564, 143–153.<br />

(31) Florens, L.; Washburn, M. P. Methods Mol. Biol. 2006, 328, 159–175.<br />

(32) Cagney, G.; Park, S.; Chung, C.; Tong, B.; O’Dushlaine, C.; Shields, D. S.;<br />

Emili, A. J. Proteome Res. 2005, 4, 1757–1767.<br />

(33) Krapfenbauer, K.; Fountoulakis, M. Methods Mol. Biol. 2009, 566, 165–<br />

180.


Figure 2. Product ion mass spectra of tryptic peptides (A) HAibDGTFTSDYSK from MK-2662, [M + 2H] 2+ , and (B) HAibDGTF[ 13 C9, 15 N]TSDYSK<br />

from [ 13 C18, 15 N2] MK-2662 (ISTD), [M + 2H] 2+ .<br />

solvent (10% acetonitrile, v/v) was necessary to maintain good<br />

chromatographic peak shape and high recovery from the SCX<br />

column. Around the retention time of the N1-12-mer on the SCX<br />

column, a narrow peak window (about 0.3-0.4 min) was<br />

transferred from the SCX column to a reversed-phase (RP)<br />

columnsthe second dimensionsfollowed by continued organic<br />

gradient mobile phase to further separate the surrogate peptide<br />

from background noise, and therefore, to provide a clean<br />

chromatogram (Figure 3B). During method development, several<br />

reversed-phase columns were tested for the second dimension,<br />

and a Thermo Scientific Hyperil Gold, PFP column (50 mm × 2.1<br />

mm, 5 µm) was found to provide the best retention. The analyte<br />

focusing was achieved by introducing 10 mM ammonium formate<br />

(pH 3) into 20% acetonitrile as a mobile phase. The chromatographic<br />

conditions, including mobile phase selection, gradient,<br />

and the timing for column switching (SCX to RP and switching<br />

back), were optimized so as to provide a good separation within<br />

a reasonable run time. Figure S1 (see the Supporting Information)<br />

is a schematic of the column and switching valve arrangement<br />

for online 2D LC, and Table S1 (see the Supporting Information)<br />

details the chromatographic conditions including the gradients,<br />

flow rates, and event timing. A RP column heater was set at 40<br />

°C to produce a symmetric and well-resolved chromatographic<br />

peak. An SCX guard column was used to protect the SCX column<br />

which resulted in over 2000 injections per set of SCX/RP columns.<br />

The acquisition window on the MS was set for 50 s so that the<br />

MS ion source was kept clean and well-maintained. The overall<br />

run time was 4.5 min per sample.<br />

Pretreatment of Clinical Plasma Samples Using DPP-IV<br />

Inhibitors as Stabilizer. Clinical samples were collected in blood<br />

collection tubes (BD-P700) containing proprietary DPP-IV inhibitors,<br />

however the composition and quantity of the enzyme inhibitor<br />

was unknown. Since BD-supplied inhibitor can only be obtained<br />

by purchasing tubes from the vendor, it presented a practicality<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6881


Figure 3. Representative chromatograms of the N1-12-mer [m/z 672 f 223] from MK-2662 plasma sample analyzed using (A) 1D LC-MS/MS<br />

or (B) 2D LC-MS/MS. The left panel of part B is the chromatogram from the SCX column (the first dimension), where T18-30 and T19-30 were<br />

the peptides obtained from trypsin digestion of MK-2662. The right panel of part B is from the RP column (the second dimension). The total run<br />

time for 2D LC-MS/MS was 4.5 min per sample.<br />

problem for preparing control plasma in the analysis laboratories.<br />

Therefore, a test was conducted to determine equivalence of BD-<br />

P700 inhibitor tubes to Linco brand DPP-IV inhibitor that was<br />

available in easy to use solution form. The results demonstrated<br />

that the stabilizers from BD tubes (1.5 mL of plasma per tube)<br />

and Linco inhibitor (20 µL/mL plasma) provided comparable<br />

effects in terms of stability (including freeze-thaw, benchtop,<br />

autosampler, and reinjection stabilities), matrix effect, and recovery<br />

in human plasma.<br />

Optimization of Trypsin Digestion Conditions. A surrogate<br />

proteolytic peptide approach using trypsin digestion was selected<br />

for the MS/MS-based quantification of MK-2662 in human plasma.<br />

Following protein precipitation, the supernatant was brought to<br />

dryness and then reconstituted in NH 4HCO3 buffer that contained<br />

trypsin for digestion. This step needed to be optimized to<br />

ensure rapid, complete, and reproducible digest within a<br />

reasonable time frame.<br />

The digestion yield was tested by varying the amount of<br />

trypsin from 2 to 10 µg, the volume of plasma in the BD-P700<br />

tube, and different digestion time at 2, 3, 4, and 16 h in both<br />

Linco DPP-IV-containing plasma and the BD-P700-treated<br />

plasma, where the latter appeared to need more trypsin due to<br />

the inhibitory effect of proprietary enzyme inhibitors. The<br />

reason for testing plasma volume was that, at the clinical site,<br />

although 3 mL of blood per sample was required in BD tube,<br />

there was a chance that less volume was collected. Since the<br />

BD tubes contain a fixed amount of protease inhibitor, the lower<br />

the blood volume, the higher the inhibitor concentration in the<br />

tube, which may lead to the more difficult trypsin digestion.<br />

6882 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

The tested plasma volume at 1 and 1.5 mL covered a range of<br />

2-3 mL of blood collection.<br />

The test results are shown in Figure S2 (see the Supporting<br />

Information), where the experiment was conducted at 10 nM MK-<br />

2662 in all samples, and the yield was calculated assuming 100%<br />

digestion in a neat MK-2662 solution incubated with trypsin for<br />

3 h. The spiked plasma without inhibitor appeared to give higher<br />

yield compared to the neat sample, presumably due to surface<br />

absorption loss of the neat sample. The data indicated that a<br />

digestion with 7 µg of trypsin was very sensitive to the change of<br />

blood volumes1 mL of plasma (equivalent to 2 mL of blood<br />

collection) required much longer digestion time to achieve >80%<br />

digestion. In contrast, a digestion with 10 µg of trypsin demonstrated<br />

an improved tolerability, yielding nearly complete digestion<br />

within 3-4 h for both 1 and 1.5 mL of plasma (equivalent to 2-3<br />

mL of blood collected in the BD-P700 tube). Balancing the benefit<br />

and risk by considering digestion yield and sample processing<br />

time, the digestion conditions10 µg of trypsin for 3 hswas<br />

selected. The digestion outcome under these conditions was<br />

robust and highly reproducible. Consequently, the clinical site was<br />

instructed to collect 3 mL of blood (1.5 mL plasma equivalent)<br />

and to record significant deviation from this volume.<br />

Solvent for Working Standard Solutions. It is well-known<br />

that peptides are prone to absorption loss to container surfaces. 9<br />

Initially, 30% acetonitrile was used for preparation of both primary<br />

stock and working standard solutions. Following storage at 4 °C<br />

for 3 weeks, the primary stock was stable; however, more than<br />

30% of MK-2662 loss was observed in working stock solutions.<br />

Switching the working stock solvent to 1% BSA and storing at


-70 °C resolved the problem, giving


Figure 4. Representative chromatograms of MK-2662 in human plasma using LC/LC-MS/MS analysis: (A) single blank from protein purification,<br />

(B) LLOQ at 1 nM from protein purification, (C) single blank from immunoaffinity purification, and (D) LLOQ at 2 nM from immunoaffinity purification<br />

(left panel, MK-2662; right panel, ISTD).<br />

samples to the peak areas of neat standards prepared in the same<br />

solvent and injected directly. The results (Table S2 in the<br />

Supporting Information) show that recoveries ranged from 100.07%<br />

to 119.36%, and matrix effects ranged between 91.32% and 97.2%<br />

across the tested concentrations. On the basis of the intraday<br />

precision results obtained using six different lots of control plasma<br />

(Table 1), an absolute matrix effect should not have any impact<br />

on assay precision.<br />

Validation of the IAP 2D LC-MS/MS Method. The accuracy<br />

and precision of the IAP 2D LC-MS/MS was accessed<br />

and compared with the PPT 2D LC-MS/MS method. The IAP<br />

6884 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

intraday validation data obtained from three standard curves<br />

constructed in three lots of human plasma over the range of 2-200<br />

nM are presented in Table 1, and the QC precision and accuracy<br />

data from three sets of QCs at each of 3, 6, 50, and 150 nM MK-<br />

2662 are shown in Table 2. The chromatograms showed that there<br />

was no interference in the blank and an adequate single-to-noise<br />

ratio for the analyte peak at the LLOQ (Figure 4, parts C and D).<br />

Both methods demonstrated good precision and accuracy, and<br />

met the acceptance criteria for handling small molecules, specified<br />

in FDA guidelines. 34 In comparison to the PPT assay, the IAP<br />

method has a 2-fold higher LLOQ. This was due to differences in


Table 1. Intraday Precision and Accuracy for the<br />

Determination of MK-2662 in Different Lots of Human<br />

Plasma<br />

method<br />

nominal<br />

concn<br />

(nM)<br />

mean<br />

measured<br />

concn (nM)<br />

(n ) 6) a<br />

accuracy<br />

(%) b<br />

precision<br />

(%) c<br />

PPT (n ) 6) 1.00 1.01 101.00 2.97<br />

2.00 1.98 99.00 3.03<br />

10.00 9.93 99.30 2.92<br />

50.00 49.86 99.72 3.43<br />

200.00 201.30 100.65 2.93<br />

500.00 503.99 100.80 2.92<br />

800.00 801.50 100.19 3.05<br />

1000.00 998.47 99.85 2.93<br />

IAP (n ) 3) 2.00 1.97 98.6 9.76<br />

5.00 5.29 105.8 2.56<br />

10.00 9.48 94.8 3.44<br />

50.00 51.22 102.4 3.17<br />

100.00 96.73 96.7 4.24<br />

160.00 162.42 101.5 4.12<br />

200.00 200.14 100.1 1.98<br />

a Back-calculated by interpolation from the calibration curve, constructed<br />

by weighted (1/x 2 ) linear regression of peak area ratios<br />

(analyte over ISTD) vs concentrations of the calibration standards in<br />

plasma. b Expressed as [((mean observed concentration)/(nominal<br />

concentration)) × 100] (%). c Coefficient of variation of back-calculated<br />

concentration.<br />

the analytical recovery of the two methods, determined to be<br />

greater than 90% for the PPT method and approximately 30% for<br />

the IAP method (data not shown). The low recovery of IAP might<br />

be the result of relatively low binding affinity of anti-PEG antibody,<br />

which consequently leads to incomplete analyte capture during<br />

the binding step and partial analyte loss during the washing steps.<br />

However, since the stable-isotope-labeled PEGylated MK-2662 was<br />

used as the IS, the low recovery of IAP, as expected, did not have<br />

significant impact on the assay precision and accuracy.<br />

Application of PPT 2D LC-MS/MS and IAP 2D LC-MS/<br />

MS to Clinical Sample Analysis. The protein precipitation<br />

followed by 2D LC-MS/MS was initially applied to analysis of<br />

the clinical plasma samples to support the first in-human study<br />

after subcutaneous (sc) administration of MK-2662. The interday<br />

QC data collected during seven daily sample analysis runs showed<br />

accuracy ranging from 104.72% to 106.33% with


Table 3. MK-2662 QC (with DPP-IV Inhibitor) Stability in Human Plasma<br />

nominal concn (nM) 3 F/T [% CV] a<br />

room temp for 5h[%CV] a<br />

PEG antibody, it provides an alternative approach for PEGylated<br />

peptide analysis.<br />

CONCLUSIONS<br />

PPT 2D LC-MS/MS- and IAP 2D LC-MS/MS-based methods<br />

for the quantitative analysis of a PEGylated peptide, MK-2662, in<br />

human plasma were developed and validated in support of clinical<br />

studies. Challenges associated with surrogate peptide selection,<br />

assay sensitivity, selectivity, and ruggedness were addressed.<br />

Protein precipitation represented a simple sample preparation<br />

method. With the use of 2D HPLC-MS/MS, the background<br />

noise from PPT extract was significantly reduced, leading to<br />

increased assay sensitivity. The trypsin digestion conditions were<br />

optimized to overcome enzyme inhibition due to protease inhibitors<br />

used to stabilize clinical samples and, therefore, improved<br />

(35) Alley, S. C.; Benjamin, D. R.; Jeffery, S. C.; Okeley, N. M.; Meyer, D. L.;<br />

Sanderson, R. J.; Senter, P. D. Bioconjugate Chem. 2008, 19, 759–765.<br />

assay ruggedness. To address assay specificity, an IAP method<br />

was developed to selectively capture the PEGylated peptide from<br />

human plasma samples using anti-PEG antibody followed by<br />

surrogate peptide analysis. Comparison of two sample preparation<br />

methods (PPT and IAP) applied to clinical sample analysis showed<br />

that protein precipitation may overestimate the PEGylated peptide<br />

concentrations. This is consistent with the concern that the PPT<br />

2D LC-MS/MS method may detect intact surrogate peptide not<br />

only originating from the PEGylated but also from dePEGylated<br />

or other truncated metabolite species presented in the clinical<br />

samples.<br />

It is worth mentioning that the immunoaffinity purification<br />

coupled with MS/MS is a special assay formatssimilar to ELISA,<br />

the peptide is captured using a capture antibody; different from<br />

ELISA, the peptide-specific measurement uses tandem mass<br />

spectrometry rather than an antibody-based detection. LC-MS/<br />

MS detection combined with IAP not only provides high specificity<br />

but also overcomes a rate-limiting step of ELISAsproduction of<br />

peptide-specific antibody. Potentially, this new assay format could<br />

be generally applicable to other PEGylated peptides.<br />

ACKNOWLEDGMENT<br />

The authors gratefully acknowledge Maria Bednarek, Andrew<br />

Zhang, and Matt Braun for providing the stable-isotope-labeled<br />

internal standard to make this work possible. In addition, we<br />

express our gratitude to Tae Han, Vijay B. G. Reddy, and Carmen<br />

Fernandez-Metzler for project-related discussions and to Albert<br />

S. Yuan, Kuo-Chang Yin, and Thorsten Verch for antibody-related<br />

discussions.<br />

NOTE ADDED AFTER ASAP PUBLICATION<br />

This paper was published on July 16, 2010 with a minor text<br />

error. The revised version was published on July 21, 2010.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review April 14, 2010. Accepted June 27,<br />

2010.<br />

AC1009832<br />

autosampler for 14 h [% CV] a<br />

reinjection [% CV] a<br />

3 105.86 [1.17] 103.26 [2.55] 101.30 [1.53] 98.70 [1.82]<br />

50 103.20 [0.49] 101.71 [1.20] 99.75 [0.76] 97.00 [0.97]<br />

800 102.46 [0.29] 100.99 [0.31] 98.74 [1.21] 97.25 [0.53]<br />

a [((mean concentration after storage)/(initial mean concentration)) × 100] (%).<br />

Figure 5. Mean plasma concentration-time profiles following a<br />

single 6 mg subcutaneous administration of MK-2662 to healthy young<br />

men: (b) the measured mean peptide concentrations from individual<br />

samples (n ) 6) analyzed using protein precipitation (PPT) coupled<br />

with LC/LC-MS/MS; (9) the measured concentration from the pooled<br />

samples obtained from the same individual subjects using PPT LC/<br />

LC-MS/MS; (1) the measured concentration from the pooled<br />

samples obtained from the same individuals using immunoaffinity<br />

purification (IAP) coupled with LC/LC-MS/MS.<br />

6886 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010


Anal. Chem. 2010, 82, 6887–6894<br />

Ferrocene Bound Poly(vinyl chloride) as Ion to<br />

Electron Transducer in Electrochemical Ion<br />

Sensors<br />

Marcin Pawlak, Ewa Grygolowicz-Pawlak, and Eric Bakker* ,†<br />

Nanochemistry Research Institute, Department of <strong>Chemistry</strong>, Curtin University of Technology, Perth,<br />

Western Australia 6845, Australia<br />

We report here on the synthesis of poly(vinyl chloride)<br />

(PVC) covalently modified with ferrocene groups (FcPVC)<br />

and the electrochemical behavior of the resulting polymeric<br />

membranes in view of designing all solid state<br />

voltammetric ion sensors. The Huisgen cycloaddition<br />

(“click chemistry”) was found to be a simple and efficient<br />

method for ferrocene attachment. A degree of PVC modification<br />

with ferrocene groups between 1.9 and 6.1 mol<br />

% was achieved. The chemical modification of the PVC<br />

backbone does not significantly affect the ion-selective<br />

properties (selectivity, mobility, and solvent casting ability)<br />

of potentiometric sensing membranes applying this<br />

polymer. Importantly, the presence of such ferrocene<br />

groups may eliminate the need for an additional redoxactive<br />

layer between the membrane and the inner electric<br />

contact in all solid state sensor designs. Electrochemical<br />

doping of this system was studied in a symmetrical<br />

sandwich configuration: glassy carbon electrode |FcPVC|<br />

glassy carbon electrode. Prior electrochemical doping<br />

from aqueous solution, resulting in a partial oxidation of<br />

the ferrocene groups, was confirmed to be necessary for<br />

the sandwich configuration to pass current effectively. The<br />

results suggest that only ∼2.3 mol % of the ferrocene<br />

groups are electrochemically accessible, likely due to<br />

surface confined electrochemical behavior in the polymer.<br />

Indeed, cyclic voltammetry of aqueous hexacyanoferrate<br />

(III) remains featureless at cathodic potentials (down to<br />

-0.5 V). This indicates that the modified membrane is<br />

not responsive to redox-active species in the sample<br />

solution, making it possible to apply this polymer as a<br />

traditional, single membrane. Yet, the redox capacity of<br />

the electrode modified with this type of membrane was<br />

more than 520 µC considering a 20 mm 2 active electrode<br />

area, which appears to be sufficient for numerous<br />

practical ion voltammetric applications. The electrode<br />

was observed to operate reproducibly, with 1% standard<br />

deviation, when applying pulsed amperometric<br />

techniques.<br />

Poly(vinyl chloride) is a polymeric matrix commonly used for<br />

electrochemical electrode preparation, and is popular in classical<br />

* To whom correspondence should be addressed.<br />

† Current address: Department of Inorganic, <strong>Analytical</strong> and Applied <strong>Chemistry</strong>,<br />

University of Geneva, CH-1211 Geneva, Switzerland.<br />

aqueous inner contact 1 and solid state electrode (SSE) configurations<br />

2-10 because of its attractive mechanical properties and<br />

compatibility with electro-active components. With SSEs, an<br />

effective ion-to-electron transduction is necessary, 2,3 which is<br />

usually applied as a separate layer between the ion-selective<br />

membrane (ISM) and the inner metallic contact 4-8 or as an<br />

additive to the membrane. 9 The most popular redox-active materials<br />

used for this purpose are conductive polymers. 4-6,9 Many types<br />

of such polymers have been successfully applied in potentiometric<br />

electrodes, e.g., poly(pyrrole) (PPy), 4,5,11 poly(octyl thiophene)<br />

(POT), 9,10,12 poly(3,4-ethylenedioxythiophene) (PEDOT), 6,13,14 and<br />

poly(aniline) (PANI), 9 and the suppression of a water layer<br />

between the conducting polymer and underlying electrode was<br />

found to be a key characteristic to guarantee stable potentiometric<br />

behavior. 15-17 In recent years, ion-selective electrodes interrogated<br />

by voltammetric techniques have increased in importance. 18 This<br />

class of sensors borrows materials and techniques from the fields<br />

of ion-selective electrodes and the electrochemistry at the interface<br />

(1) Bakker, E.; Buhlmann, P.; Pretsch, E. Chem. Rev. 1997, 97, 3083–3132.<br />

(2) Cattrall, R. W.; Drew, D. M.; Hamilton, I. C. Anal. Chim. Acta 1975, 76,<br />

269–277.<br />

(3) Hulanicki, A.; Trojanowicz, M. Anal. Chim. Acta 1976, 87, 411–417.<br />

(4) Cadogan, A.; Gao, Z.; Lewenstam, A.; Ivaska, A. Anal. Chem. 1992, 64,<br />

2496–2501.<br />

(5) Michalska, A.; Hulanicki, A.; Lewenstam, A. Microchem. J. 1997, 57, 59–<br />

64.<br />

(6) Vazquez, M.; Bobacka, J.; Ivaska, A.; Lewenstam, A. Sens. Actuators, B<br />

2002, 82, 7–13.<br />

(7) Cosofret, V. V.; Erdosy, M.; Johnson, T. A.; Buck, R. P. Anal. Chem. 1995,<br />

67, 1647–1653.<br />

(8) Fibbioli, M.; Badyopadhyay, K.; Liu, S. G.; Echegoyen, L.; Enger, O.;<br />

Diederich, F.; Buhlmann, P.; Pretsch, E. Chem. Commun. 2000, 5, 339–<br />

340.<br />

(9) Bobacka, J.; Lindfors, T.; McCarrick, M.; Ivaska, A.; Lewenstam, A. Anal.<br />

Chem. 1995, 67, 3819–3823.<br />

(10) Song, F.; Ha, J.; Park, B.; Kwak, T. H.; Kim, I. T.; Nam, H.; Cha, G. S.<br />

Talanta 2002, 57, 263–270.<br />

(11) Sutter, J.; Lindner, E.; Gyurcsanyi, R. E.; Pretsch, E. Anal. Bioanal. Chem.<br />

2004, 380, 7–14.<br />

(12) Sutter, J.; Radu, A.; Peper, S.; Bakker, E.; Pretsch, E. Anal. Chim. Acta<br />

2004, 523, 53–59.<br />

(13) Bobacka, J. Anal. Chem. 1999, 71, 4932–4937.<br />

(14) Vazquez, M.; Danielson, P.; Bobacka, J.; Lewenstam, A.; Ivaska, A. Sens.<br />

Actuators, B 2004, 97, 182–189.<br />

(15) Fibbioli, M.; Morf, W. E.; Badertscher, M.; de Rooij, N. F.; Pretsch, E.<br />

Electroanalysis 2000, 12, 1286–1292.<br />

(16) Gyurcsanyi, R. E.; Rangisetty, N.; Clifton, S.; Pendley, B. D.; Lindner, E.<br />

Talanta 2004, 63, 89–99.<br />

(17) De Marco, R.; Veder, J.-P.; Clarke, G.; Nelson, A.; Prince, K.; Pretsch, E.;<br />

Bakker, E. Phys. Chem. Chem. Phys. 2008, 10, 73–76.<br />

(18) Samec, Z.; Samcova, E.; Girault, H. H. Talanta 2004, 63, 21–32.<br />

10.1021/ac1010662 © 2010 American <strong>Chemical</strong> Society 6887<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/28/2010


of two immiscible electrolyte solutions (ITIES) and has allowed<br />

researchers to design reversibly operating polyion sensors, 19-21<br />

flash chronopotentiometric electrodes for rapid determination of<br />

kinetically labile components, 22 ion channel mimetic systems, 23<br />

thin layer coulometric sensors, 24 thin layer electrodes for stripping<br />

voltammetry, 25 as well as controlled current coulometric sensors. 26<br />

The requirements for the realization of all solid state voltammetric<br />

ion sensors are more difficult to satisfy than the potentiometric<br />

sensors mentioned above, since the required redox capacity<br />

must be sufficiently high to sustain the ion transfer current.<br />

Shvarev showed that PEDOT is a promising conducting polymer<br />

underlayer for this purpose, 27 Amemiya 25,28 and our group 29<br />

studied conducting polymers for thin layer voltammetric applications.<br />

With these materials, the redox properties are typically<br />

conducive to anion extraction from the sample solution. To allow<br />

for cation extraction, the polymer may sometimes be doped to<br />

obtain its stable oxidized form. 29<br />

An alternative strategy was suggested by Samec et al. 30 by the<br />

incorporation of lipophilic ferrocene derivatives. This appears to<br />

be an attractive alternative to conducting polymer based systems,<br />

owing to the well-defined reversible redox couple and a potentially<br />

high redox capacity of the resulting membranes. Freely dissolving<br />

such redox active species may, however, result in eventual<br />

leaching from the membrane (especially in its oxidized, positively<br />

charged form) and in interferences from redox active species in<br />

the sample.<br />

To address these limitations and to introduce an alternative<br />

material, the chemical attachment of ferrocene groups to polymer<br />

chains is explored here. From a chemical point of view, poly(vinyl<br />

chloride) is easily modifiable due to the susceptibility for nucleophilic<br />

substitution of the chlorine atoms. 31,32 The most commonly<br />

used method for introducing new functional groups into the PVC<br />

chain is the reaction with sodium azide. 33,34 This modification was<br />

used for example in the cross-linking of the polymer. 35 Recently,<br />

the Huisgen cycloaddition, known also as “click chemistry”, has<br />

become a very popular method of transforming azide groups. 36<br />

This type of reaction is attractive because of its high selectivity,<br />

mild reaction conditions, very high yields, and lack of byproducts.<br />

It was successfully applied for modification of a wide range of<br />

(19) Shvarev, A.; Bakker, E. J. Am. Chem. Soc. 2003, 125, 11192–11193.<br />

(20) Samec, Z.; Trojanek, A.; Langmaier, J.; Samcova, E. Electrochem. Commun.<br />

2003, 5, 867–870.<br />

(21) Yuan, Y.; Amemiya, S. Anal. Chem. 2004, 76, 6877–6886.<br />

(22) Gemene, K. L.; Bakker, E. Anal. Chem. 2008, 80, 3743–3750.<br />

(23) Xu, Y.; Bakker, E. Langmuir 2009, 26, 568–573.<br />

(24) Yoshizumi, A.; Uehara, A.; Kasuno, M.; Kitatsuhi, Y.; Yoshida, Z.; Kihara,<br />

S. J. Electroanal. Chem. 2005, 581, 275.<br />

(25) Kim, Y.; Amemiya, S. Anal. Chem. 2008, 80, 6056–6065.<br />

(26) Bhakthavatsalam, V.; Shvarev, A.; Bakker, E. Analyst 2006, 131, 895–900.<br />

(27) Perera, H.; Fordyce, K.; Shvarev, A. Anal. Chem. 2007, 79, 4564–4573.<br />

(28) Guo, J.; Amemiya, S. Anal. Chem. 2006, 78, 6893–6902.<br />

(29) Si, P.; Bakker, E. Chem. Commun. 2009, 35, 5260–5262.<br />

(30) Langmaier, J.; Olsak, J.; Samcova, E.; Samec, Z.; Trojanek, A. Electroanalysis<br />

2006, 18, 1329–1338.<br />

(31) Kameda, T.; Ono, M.; Grause, G.; Mizoguchi, T.; Yoshioka, T. Polym.<br />

Degrad. Stab. 2009, 94, 107–112.<br />

(32) Kameda, T.; Fukuda, Y.; Grause, G.; Yoshioka, T. J. Appl. Polym. Sci. 2010,<br />

116, 36–44.<br />

(33) Sacristan, J.; Mijangos, C.; Reinecke, H.; Spells, S.; Yarwood, J. Macromolecules<br />

2000, 33, 6134–6139.<br />

(34) Martinez, G. J. Polym. Sci., Part A: Polym. Chem. 2006, 44, 2476–2486.<br />

(35) Jayakrishnan, A.; Sunny, M. C. Polymer 1996, 37, 5213–5218.<br />

(36) Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angew. Chem.,<br />

Int. Ed. 2002, 41, 2596–2599.<br />

6888 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

different polymers, 37 but it has not yet become popular for the<br />

modification of PVC. Considering that the modified polymer is<br />

intended to be applied for electrochemical electrode preparation,<br />

any redox activity is expected to be related to the ferrocene groups<br />

only. The redox active properties of a triazol linker obtained by<br />

applying the Huisgen cycloaddition has been studied by Kirrs and<br />

co-workers, 38 revealing that it is indeed electrochemically inert.<br />

In this paper we present studies on properties of membranes<br />

containing poly(vinyl chloride) modified with ferrocene groups<br />

via the click-chemistry reaction.<br />

EXPERIMENTAL SECTION<br />

General Methods and Materials. High molecular weight<br />

poly(vinyl chloride) (PVC), sodium azide, ethynylferrocene, Lascorbic<br />

acid, copper sulfate pentahydrate, tetradodecylammonium<br />

tetrakis(4-chlorophenyl)borate (ETH 500), bis(2-ethylhexyl) sebacate<br />

(DOS), potassium hexacyanoferrate (III), potassium chloride,<br />

sodium perchlorate, sodium nitrate, sodium thiocyanate,<br />

anhydrous tetrahydrofuran (THF), and dimethylformamide (DMF)<br />

were purchased from Sigma Aldrich and used without further<br />

purification. Aqueous solutions were prepared by dissolving the<br />

appropriate salts in Milli-Q-purified distilled water.<br />

IR spectra were taken using a Perkin-Elmer Spectrum 100 FT-<br />

IR spectrometer with attenuated total reflectance (ATR) sampling<br />

accessory and ATR corrected. UV-vis spectra were recorded in<br />

the 900-350 nm wavelength range on a GBC UV-vis 916<br />

spectrophotometer using 1 cm thick quartz cuvettes. The UV-vis<br />

calibration curve was obtained for three samples containing 2, 6,<br />

and 12 µmol/mL of ferrocene dissolved in a matrix containing 33<br />

mg of PVC per mL of THF. The first order derivative of the<br />

spectrum was used to determine the content of ferrocene attached<br />

to the PVC. Glassy carbon rods of 5 mm diameter were purchased<br />

from SPI Supplies/Structure Probe, Inc. (West Chester, PA) and<br />

assembled into an epoxy resin body (EpoFix Kit, Struers Australia,<br />

Milton QLD, Australia).<br />

Synthesis. The substitution reaction of PVC with NaN 3 was<br />

carried out using 1 g (16 mmol based on monomeric unit) of<br />

PVC and 1.04 g (16 mmol) of NaN3 in 42 mL of DMF-water<br />

mixture (5:1 volume ratio) and heated at 50 °C under nitrogen<br />

for various times from 24 to 168 h. When the desired degree<br />

of modification was achieved, the reaction mixture was cooled<br />

to room temperature, the product was filtered, washed with<br />

distilled water and methanol, and subsequently dried under<br />

reduced pressure. In the click-chemistry reaction between<br />

azide-modified PVC and ethynylferrocene, 300 mg of N3PVC,<br />

60 mg (0.29 mmol) of ethynylferrocene, 140 mg (0.56 mmol)<br />

of CuSO4 · 5H2O, and 500 mg (2.84 mmol) of ascorbic acid were<br />

added to 28 mL of a THF-water mixture (6:1 volume ratio).<br />

After 4hofreaction time, the product was precipitated with<br />

water. Ferrocene modified PVC was filtered, washed thoroughly<br />

with methanol, and then dissolved in 20 mL of THF.<br />

The THF solution was then filtered again to remove insoluble<br />

impurities, and the product was precipitated with methanol,<br />

filtered, and dried under reduced pressure.<br />

Composition of Membrane Cocktails. Membrane cocktail<br />

was prepared by dissolving 7.5 mg of FcPVC, 37.5 mg of DOS,<br />

(37) Lutz, J.-F. Angew. Chem., Int. Ed. 2007, 46, 1018–1025.<br />

(38) Verschoor-Kirss, M.; Kreisz, J.; Feighery, W.; Reiff, W. M.; Frommen, C. M.;<br />

Kirss, R. U. Organomet. Chem. 2009, 694, 3262–3269.


Figure 1. Infrared spectra of (A) unmodified PVC, (B) azide-modified PVC, and (C) ferrocene-modified PVC. Insets: structures of azide modified<br />

PVC and FcPVC.<br />

and 5 mg of ETH 500 in 1 mL of THF. DOS is a plasticizer<br />

commonly used as a solvent in PVC-based (liquid) membranes.<br />

The lipophilic electrolyte ETH 500 is added to provide counterions<br />

for the extracted analyte in the membrane and is also used to<br />

keep the electrical resistance of the membrane low. Electrodes<br />

were prepared by drop casting 40 µL of cocktail and letting the<br />

THF evaporate overnight under ambient conditions.<br />

Electrochemical Experiments. All cyclic voltammetry experiments<br />

were performed at a 50 mV s -1 scan rate. Normal pulse<br />

voltammograms were performed in a -0.6 to +0.8 V potential<br />

range using 10 mM solutions of NaCl, NaSCN, NaNO3, and<br />

NaClO4. Potential pulses (1 s) with a 20 mV potential increase<br />

were followed by a5sregeneration pulse (baseline potential)<br />

at open circuit potential (as measured before each voltammetric<br />

experiment). Membrane doping was performed with a single<br />

pulse of 10 min duration in a potential range of 400-800 mV<br />

in 10 mM NaClO4. A time of 10 min was chosen experimentally<br />

as the shortest time allowing the observed current to sufficiently<br />

decay. In the sandwiched electrode experiment, a three<br />

pulse procedure was applied. The first 30 s pulse at 0 mV was<br />

to confirm the steady state of the system before a 300 s long<br />

pulse at 200 mV was applied to induce electrochemical<br />

processes, while the last, carried out for 300 s at 0 mV, was to<br />

allow the system to return to its original state.<br />

Cyclic voltammetry in the presence of hexacyanoferrate(III)<br />

ion was performed in 1 mM K3Fe(CN)6 solution with 0.1 M KCl<br />

as a background, and the potential was scanned between a 0.8<br />

and -0.5 V potential range starting from 0.4 V and going to<br />

negative potentials first. The stability of the signal generated<br />

by the FcPVC modified electrode was examined by application<br />

ofa1spulse at 0.5 V and followed by 20 s regeneration at the<br />

applied open circuit potential, using a three electrode configuration<br />

and 10 mM sodium perchlorate. The procedure was<br />

repeated 40 times. The redox capacity of the FcPVC was tested<br />

in the same solution by applying a 0.8 V potential for 30, 60,<br />

120, 300, and 600 s, each time followed by a discharge pulse<br />

at 0 V for 200 s.<br />

Table 1. Reaction Time of Azidation and Resulting<br />

Degree of Ferrocene Modification of PVC, Estimated<br />

by Spectrophotometry<br />

reaction time [h] degree of modification [mol %]<br />

24 1.9<br />

60 3.2<br />

120 5.3<br />

168 6.1<br />

RESULTS AND DISCUSSION<br />

In an initial step, PVC was modified with azide groups using<br />

NaN3 according to Sacristan et al. 33 The efficiency of the<br />

reaction was confirmed by comparing the IR spectrum for<br />

commercial, not modified PVC, and the same polymer after<br />

modification with azide groups. As shown in Figure 1A,B, the<br />

distinctive azide band at 2100 cm -1 , not visible for unmodified<br />

PVC, gives a well developed peak after reaction, confirming<br />

the effective partial modification of the polymeric backbone<br />

with azide groups.<br />

Further modification of azide-modified PVC with ferrocene<br />

groups required the use of a large excess of catalyst (200 mol %<br />

relative to the ethynylferrocene) to allow a high yield of reaction.<br />

As seen in Figure 1C, the IR spectrum for the reaction product<br />

did not reveal any signal from azide groups, indicating full<br />

conversion of the substrate. Since the ferrocene groups absorb<br />

strongly at 440 nm, the content of the ferrocene groups in the<br />

obtained FcPVC could be conveniently estimated from a calibration<br />

curve prepared for mixtures of nonmodified PVC with various<br />

ethynylferrocene contents dissolved in THF. Since the baselines<br />

for a THF solution of the ferrocene-modified PVC were significantly<br />

different, the ferrocene content was calculated based on<br />

spectral derivatives. On the basis of the spectrophotometric results<br />

and considering complete conversion of azide to ferrocene in the<br />

polymer backbone (as indicated by FT-IR) it may be concluded<br />

that the content of the ferrocene groups in PVC can be easily<br />

controlled by adjusting the time of azidation reaction. As shown<br />

in Table 1 for the polymer that has been modified with azide<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6889


Figure 2. Chronoamperograms obtained for a pair of sandwiched electrodes modified with a FcPVC-DOS membrane containing 10 wt % of<br />

ETH 500, anodically doped with perchlorate from a 10 mM NaClO4 solution at the indicated potentials. Inset: charge calculated from the<br />

chronoamperograms as a function of the doping potential.<br />

groups in the first step reaction for only 24 h, the content of<br />

ferrocene groups attached during the second reaction step was<br />

1.9 mol %. The degree of modification significantly increased to<br />

3.2 mol % when the reaction time for the first step was increased<br />

to 60 h. After 120 h of the azidation, the reaction rate started to<br />

slow down while after 168 h the efficiency reached 6.1 mol % and<br />

its further increase was negligible.<br />

Unfortunately, with higher conversion rate the solubility of the<br />

modified PVC in THF drops significantly. For example, the highest<br />

concentration that could be obtained for a membrane cocktail<br />

containing FcPVC with 3.2 mol % of ferrocene groups is 50 mg/<br />

mL of THF (which is half of the typical cocktail concentration<br />

when using unmodified PVC), while only 17 mg of the cocktail<br />

containing polymer with 6 mol % of ferrocene groups could be<br />

dissolved in the same volume. This property is most likely related<br />

to the presence of aromatic linker between the ferrocene group<br />

and polymer backbone, which may result in a loose cross-linking<br />

due to π-stacking. Therefore, for practical reasons FcPVC containing<br />

3.2 mol % of ferrocene (461 mmol/kg) was chosen for further<br />

studies as a compromise between electrochemical activity and<br />

solubility.<br />

In initial studies, the redox properties of FcPVC in the ionselective<br />

membrane were studied independently of an aqueous<br />

electrolyte. The redox process at the inner side of the membrane<br />

is otherwise difficult to distinguish from the outer ion transfer<br />

process, which may be rate limiting owing to the low mobility of<br />

ion-selective membrane materials. 39,40 Removal of the aqueous<br />

electrolyte was achieved here in a fully symmetrical cell, sandwiching<br />

the ion-selective ferrocene functionalized membrane<br />

between two glassy carbon (GC) electrodes. Note that a fully<br />

reduced ferrocene based membrane is not expected to pass<br />

current effectively, since any oxidation at one electrode must be<br />

accompanied by a concurrent reduction at the second electrode.<br />

Ideally, therefore, the membrane must be composed of a reduced<br />

(39) Long, R.; Bakker, E. Electroanalysis 2003, 15, 1261–1269.<br />

(40) Long, R.; Bakker, E. Anal. Chim. Acta 2004, 511, 91–95.<br />

6890 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

and oxidized form in a one to one molar ratio for the following<br />

reaction to occur efficiently:<br />

GC anode| Fc0 Fc +| Fc+ Fc 0| GC cathode<br />

This was confirmed here by an electrochemical doping step from<br />

sodium perchlorate solution prior to assembling the sandwich<br />

configuration. Two identical glassy carbon electrodes modified<br />

with a FcPVC/DOS/ETH 500 membrane were placed in 10 mM<br />

NaClO4 solution and held at 400 mV for 600 s to partially oxidize<br />

ferrocene groups by incorporation of perchlorate ions into the<br />

membrane. Immediately after the doping process, the electrodes<br />

were sandwiched together. A potential of 200 mV was<br />

applied between the two electrodes for 300 s and the current<br />

recorded (see Figure 2). The observed current very soon decayed<br />

to near zero values, indicating a low ferrocenium concentration<br />

in the membrane. The total number of coulombs calculated from<br />

the integral of the chronoamperogram was 16.45 µC. After<br />

electrode separation, they were again immersed into the doping<br />

solution and held at 500 mV for the same time period. The<br />

sandwich procedure was repeated and higher currents were<br />

observed, resulting in a charge of 102.88 µC. This experimental<br />

sequence was repeated for increasing doping potentials. A plot of<br />

calculated charge as a function of doping potential is presented<br />

in the inset of Figure 2. Doping potentials higher than 600 mV<br />

gave a decreasing charge, consistent with ferrocene as a limiting<br />

reagent in reaction 1 above. Note that more positive potentials<br />

may also result in the oxidation of the tetraphenylborate anion, 28<br />

a component of the lipophilic electrolyte, and hence may explain<br />

the observed distortion of the expected bell-shaped curve (number<br />

of coulombs vs doping potential) at larger potentials.<br />

From the maximum charge found for the optimally doped<br />

electrodes it can be calculated that the amount of ferrocenium/<br />

ferrocene groups participating in the process was 3.16 nmol per<br />

20 mm 2 of electrode area, or about 10 mmol kg -1 (of which<br />

(1)


Figure 3. Cyclic voltammograms obtained for a pair of sandwiched electrodes modified with FcPVC membrane containing 461 mmol/kg of<br />

ferrocene groups and 10 wt % of ETH 500 anodically doped from 10 mM NaClO4 solution at 600 mV carried out for the indicated times after the<br />

doping.<br />

half can be oxidized and half reduced). This is just about 2.3<br />

mol % of what was calculated based on the spectrophotometric<br />

calibration curve. Considering that the mobility of ferrocene/<br />

ferrocenium groups attached to a long chain of the polymer is<br />

limited, only the molecules being at a close distance from the<br />

glassy carbon electrode may participate in the electrochemical<br />

processes within the experimental time frame. This, however,<br />

may be an advantage considering the method presented here,<br />

since it suggests that the material is insensitive to redox active<br />

species in the contacting aqueous solution (see below).<br />

When cyclic voltammetry was performed on a freshly doped<br />

sandwiched system, a redox couple was observed, with oxidation<br />

and reduction peaks comparable to the theoretical shape obtained<br />

for ferrocene dissolved in acetonitrile on a gold electrode. 41 The<br />

peak heights for the oxidation and reduction processes were<br />

almost identical indicating that the system is fully reversible. The<br />

peaks were separated by 320 mV which is more than in the case<br />

of ferrocene dissolved in acetonitrile (56 mV). This is not<br />

unexpected, given the fact that ferrocene is attached to the PVC<br />

backbone (and not diffusion limited as in the theoretical case)<br />

and the second electrode is acting as both the counter and<br />

reference electrode. However, for successive scans important<br />

changes of the maximum and minimum current values were<br />

observed. A series of 10 scans was repeated 40, 80, 120, and 160<br />

min after doping and showed that the maximum peak currents<br />

successively diminished with time (Figure 3). On the basis of our<br />

previous observations that within the time frame of the doping<br />

experiment (600 s) only about 2% of redox-active groups undergo<br />

oxidation, the decrease of observed currents may be explained<br />

by the gradual diffusion of oxidized ferrocene away from the<br />

electrode surface into the surrounding bulk membrane. Functionalities<br />

covalently attached to the PVC backbone are indeed<br />

not strictly immobile, as evidenced by earlier work demonstrating<br />

(41) Bard, A. J.; Faulkner, L. R. Electrochemical Methods; Wiley: New York, 2001.<br />

equilibration of solvent cast plasticized PVC films with covalently<br />

attached chromophore functionalities throughout the bulk of the<br />

material. 42<br />

Subsequent work focused on three electrode systems in order<br />

to characterize the solid state membrane in contact with aqueous<br />

electrolyte. The ion-selective properties of the new polymer were<br />

studied in 10 mM solutions with normal pulse voltammetry, which<br />

has been established as a robust and gentle method of ionselective<br />

membrane characterization. 43 The results are presented<br />

in Figure 4. The potentials at which the ion transfer of ClO4 - ,<br />

SCN - ,NO3 - , and Cl - ions occur at a current of 1 µA were<br />

found at about 420, 460, 580, and 700 mV, respectively. The<br />

order and the magnitude of the potential separation are<br />

consistent with the theoretical selectivity expected for lipophilicity-based<br />

extraction according to the Hofmeister series. 43<br />

This confirms that the ferrocene modification has a negligible<br />

effect on the anion extraction properties of the polymer. When<br />

scanning the electrode to a negative potential range, a cation<br />

extraction wave should normally occur with aqueous inner<br />

contact systems. 44 The inset of Figure 4 suggests for the sodium<br />

chloride sample that the voltammogram is very quiet at cathodic<br />

potentials, suggesting no cationic response. As mentioned above,<br />

the ferrocene groups present in the polymeric phase are predominately<br />

in their reduced state, suppressing cathodic response. It is<br />

expected that doping the membrane with lipophilic cationexchanger<br />

according to a procedure recently reported for poly-<br />

(octyl thiophene) (POT) 29 will allow one to yield cation transfer<br />

voltammetric characteristics, but this was not pursued here.<br />

Ferrocene groups are randomly distributed in the membrane<br />

phase and hence some of them are believed to be present on the<br />

membrane surface. Exposure of the conductive groups to the<br />

sample solution may cause undesired electrode sensitivity to<br />

redox-active solutes. Therefore, the studied electrodes were tested<br />

(42) Rosatzin, T.; Holy, P.; Seiler, K.; Rusterholz, B.; Simon, W. Anal. Chem.<br />

1992, 64, 2029–2035.<br />

(43) Jadhav, S.; Meir, A. J.; Bakker, E. Electroanalysis 2000, 12, 1251–1257.<br />

(44) Jadhav, S.; Bakker, E. Anal. Chem. 2001, 73, 80–90.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6891


Figure 4. Normal pulse voltammograms for an electrode modified with FcPVC-DOS with 10 wt % of ETH 500 performed in 10 mM solutions<br />

of the indicated anions. Inset: Response behavior to sodium chloride at more negative potentials, demonstrating the suppression of cation<br />

transfer processes.<br />

Figure 5. Cyclic voltammogram for an FcPVC modified electrode in 1 mM solution of K3Fe(CN)6 and 0.1 M KCl.<br />

for redox sensitivity in the presence of potassium hexacyanoferrate(III),<br />

which is known to undergo extraction into the organic<br />

phase at positive potentials, 45 giving anodic currents, while the<br />

undesired reduction of the Fe(III) species mediated by FcPVC<br />

should result in cathodic currents at negative potentials. To<br />

evaluate this, electrodes were scanned from -0.5 to 0.8 V in<br />

aqueous hexacyanoferrate(III) solutions (Figure 5). The cyclic<br />

voltammograms revealed indeed an anodic anion extraction wave<br />

at positive potentials, together with its subsequent stripping on<br />

the return scan. On the other hand, the cyclic voltammogram in<br />

the 0.4 to -0.5 V potential range remained featureless, suggesting<br />

that the Fe(III) cannot easily be reduced at the membrane surface.<br />

This is consistent with the observation above that only a small<br />

fraction of total membrane-bound ferrocene is electrochemically<br />

accessible. This suggests that the material may be used as a<br />

singular membrane matrix without the need for a more complex<br />

two-layer system that is necessary with the conducting polymer<br />

based ion to electron transducing layers.<br />

(45) Chen, S.-M.; Chzo, W.-Y.; Thangamuthu, R. J. Solid State Electrochem. 2008,<br />

12, 1487–1495.<br />

6892 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

One disadvantage of the conducting polymer layer POT as an<br />

ion-to-electron transducer appears to be its limited redox capacity.<br />

As shown by Kim et al., a layer of this polymer obtained during<br />

electro-deposition on a 5 mm diameter gold electrode and modified<br />

with a thin, spin-coated PVC membrane showed a capacity limit<br />

of about 3 µC. 25 This value was found for a stripping voltammetry<br />

analysis applying 30, 60, 120, 300, and 600 s doping times. An<br />

analogous experiment was performed here for FcPVC-based<br />

electrode applying the same doping times at 0.8 V. Since the<br />

stripping process was expected to be slower for a thicker<br />

membrane (about 100 µm), more suitable procedure for FcPVC<br />

redox-capacity study was to calculate the charge of a doping<br />

process applying chronoamperometry. The doping was performed<br />

at 0.8 V for 30, 60, 120, 300, and 600 s. The charge calculated<br />

based on integrals of obtained chronoamperograms is presented<br />

as a function of doping time in Figure 6. Contrary to results<br />

obtained by Kim et al. where the POT based electrode reached<br />

its maximum capacity after 2 min to a value of less than 3 µC, 25<br />

the doping charge was still growing for the FcPVC-based electrode<br />

after 600 s. It was found that after 600 s of FcPVC doping, the


Figure 6. Calculated charge from chronoamperometric perchlorate ion uptake (0.1 M NaClO4) into the ferrocene modified plasticized PVC<br />

membrane electrode, performed at 0.8 V vs Ag/AgCl and carried out for the indicated times.<br />

Figure 7. Normal pulse amperometry of a FcPVC-based membrane electrode immersed in 10 mM NaClO4 for 40 anodic perchlorate ion<br />

transfer pulses for 1sat0.5V,each followed by 20 s stripping pulses at 0.3 V (open circuit potential, found before the experiment). Shown here<br />

for clarity are the first and last two of the recorded pulses.<br />

calculated doping charge was more than 520 µC, which is 2 orders<br />

of magnitude larger than for POT based electrodes. The value is<br />

somewhat larger than for the sandwiched system described above<br />

(see Figure 2), likely because of the applied electrochemical<br />

doping times were not long enough to effect conversion of the<br />

entire membrane material. Moreover, more than 50 µC conversion<br />

was observed for FcPVC in the first 30 s, significantly more than<br />

the 2.5 µC for POT-based membranes in the same time period, 25<br />

which is advantageous for ion voltammetry sensor development.<br />

When operating the membrane by normal pulse amperometry<br />

or normal pulse voltammetry, the observed currents are often<br />

limited by ion transport processes in the aqueous or membrane<br />

phase. 43 If this holds true, changes in charge transfer kinetics of<br />

the underlying redox couple may not substantially influence the<br />

observed amperometric response if the potential of the redox<br />

couple remains constant in the course of the experiment. Figure<br />

7 demonstrates normal pulse amperometry in a 10 mM NaClO4<br />

solution, using a1sexcitation pulse at 0.5 V followed by a<br />

20 s baseline pulses at the open circuit potential. Indeed, the<br />

current response generated during the 1 s long pulse was very<br />

reproducible (1% relative SD) in the course of a 40 pulse<br />

sequence. This suggests that the ferrocene modified PVC<br />

introduced here is a promising material for voltammetric ion<br />

sensor design. While we have not tested the behavior of freely<br />

dissolved lipophilic ferrocene species in this work, we note that<br />

the group of Samec has reported on adequate stability with<br />

membranes containing freely dissolved 1,1′-dimethylferrocene<br />

as a redox active additive, although possible interference by<br />

redox active species in the sample were not yet studied there. 30<br />

CONCLUSIONS<br />

This paper presented a modification of PVC with ferrocene<br />

groups via an attractive “click chemistry” approach. The new<br />

material (FcPVC) was characterized in view of membrane electrode<br />

applications, specifically with voltammetric transduction<br />

principles. A new method applying two identical electrodes at the<br />

same doping state sandwiched together allowed one to confirm<br />

the redox properties of the FcPVC, independent of the processes<br />

occurring at the sample side of the membrane. This is important<br />

because the resulting sensors are often operationally limited by<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6893


ion transfer and diffusion processes, which are not the focus of<br />

this research. The redox capacity of the system was found to<br />

correspond to several nanomolar of oxidizable ferrocene groups<br />

in the membrane, which is only a fraction of nominal ferrocene<br />

concentration. This suggests that the ferrocene groups are<br />

sufficiently dilute and immobile in the membrane and that it is<br />

the surface confined functionalities that are predominantly electrochemically<br />

accessible. Nonetheless, the redox capacity appears<br />

to be adequate for most voltammetric ion transfer applications.<br />

The membranes were characterized in contact with aqueous<br />

electrolytes, and interrogation with normal pulse voltammetry<br />

suggests excellent reproducibility and an anion selectivity pattern<br />

that corresponds to the expected Hofmeister selectivity sequence.<br />

The cathodic potential window remained featureless, supporting<br />

the notion of a membrane containing predominantly ferrocene<br />

moieties that can be oxidized by the concomitant extraction of<br />

anions from the sample. It is expected that adequate chemical<br />

doping of the membrane with cation-exchanger species will make<br />

the cationic response region at cathodic potentials analytically<br />

6894 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

accessible, but this was not yet explored here. Experiments in<br />

the presence of potassium hexacyanoferrate(III) showed that<br />

possible presence of ferrocene groups at the surface of the ionselective<br />

membrane do not cause redox sensitivity of the electrode.<br />

Moreover, the high reproducibility of the sampled anion transfer<br />

currents (standard deviation below 1%) as well as the high redox<br />

capacity of the material (about 520 µC) makes ferrocene bound<br />

PVC an attractive material for electrochemical all solid state ionselective<br />

electrode fabrication.<br />

ACKNOWLEDGMENT<br />

The authors wish to thank the Australian Research Council<br />

(Grant DP0987851) and the National Institutes of Health (Grant<br />

EB002189) for financial support of this research.<br />

Received for review April 23, 2010. Accepted July 10,<br />

2010.<br />

AC1010662


Anal. Chem. 2010, 82, 6895–6903<br />

Electrophoretic Analysis of Biomarkers using<br />

Capillary Modification with Gold Nanoparticles<br />

Embedded in a Polycation and Boron Doped<br />

Diamond Electrode<br />

Lin Zhou, † Jeremy D. Glennon, † and John H. T. Luong* ,†,‡<br />

Innovative Chromatography Group, Irish Separation Science Cluster (ISSC), Department of <strong>Chemistry</strong> & the ABCRF,<br />

University College Cork, Cork, Ireland and Biotechnology Research Institute, National Research Council Canada,<br />

Montreal, Quebec, Canada H4P 2R2<br />

Field-amplified sample stacking using a fused silica<br />

capillary coated with gold nanoparticles (AuNPs) embedded<br />

in poly(diallyl dimethylammonium) chloride (PDDA)<br />

has been investigated for the electrophoretic separation<br />

of indoxyl sulfate, homovanillic acid (HVA), and vanillylmandelic<br />

acid (VMA). AuNPs (27 nm) exhibit ionic and<br />

hydrophobic interactions, as well as hydrogen bonding<br />

with the PDDA network to form a stable layer on the<br />

internal wall of the capillary. This approach reverses<br />

electro-osmotic flow allowing for fast migration of the<br />

analytes while retarding other endogenous compounds<br />

including ascorbic acid, uric acid, catecholamines, and<br />

indoleamines. Notably, the two closely related biomarkers<br />

of clinical significance, HVA and VMA, displayed differential<br />

interaction with PDDA-AuNPs which enabled the separation<br />

of this pair. The detection limit of the three analytes obtained<br />

by using a boron doped diamond electrode was ∼75 nM,<br />

which was significantly below their normal physiological<br />

levels in biological fluids. This combined separation and<br />

detection scheme was applied to the direct analysis of these<br />

analytes and other interfering chemicals including uric and<br />

ascorbic acids in urine samples without off-line sample<br />

treatment or preconcentration.<br />

Indoxyl sulfate (IXS), a metabolite of tryptophan (TRP) and a<br />

dietary protein, is derived from intestinal metabolism and liver<br />

conjugation 1 and excreted in urine in high concentration. It is also<br />

an endogenous compound in mammals, in mouse plasma and<br />

brain samples, as detected by liquid chromatography/tandem<br />

mass spectrometry. 2 This circulating protein-bound uremic toxin<br />

stimulates glomerular sclerosis, interstitial fibrosis, and the<br />

progression rate of renal failure. IXS induces endothelial dysfunction<br />

by inhibiting endothelial proliferation and migration in vitro<br />

* To whom correspondence should be addressed.<br />

† Innovative Chromatography Group, Irish Separation Science Cluster (ISSC),<br />

Department of <strong>Chemistry</strong> & the ABCRF, University College Cork.<br />

‡ Biotechnology Research Institute, National Research Council Canada.<br />

(1) Dealler, S. F.; Hawkey, P. M.; Millar, M. R. J. Clin. Microbiol. 1988, 26<br />

(10), 2152–2156.<br />

(2) Wang, G.-F.; Korfmacher, W. A. Rapid Commun. Mass Spectrom. 2008,<br />

23 (13), 2061–2069.<br />

and its role in oxidative stress is implicated. 3 Homovanillic acid<br />

(HVA) is a major catecholamine metabolite, associated with the<br />

brain dopamine level. As a biomarker of metabolic stress of<br />

2-deoxy-D-glucose, the HVA level in the brain and the cerebrospinal<br />

fluid is indicative of pheochromocytoma and neuoroblastoma. 4<br />

Metanephrine, one of the hormones produced by the adrenal<br />

glands, breaks down to normetanephrine and vanillylmandelic acid<br />

(VMA) via the intermediate 4-hydroxy-3-methoxy-phenylglycol.<br />

Thus, VMA is always detected in the urine together with HVA<br />

and other catecholamine metabolites from pheochromocytoma or<br />

catecholamine-secreting chromaffin tumor cells. 5,6 Catecholamine<br />

levels can be found in blood samples; however, the urine test<br />

reflects the production rate of the catecholamines over the<br />

collection period. Even at abnormal levels, the elevated quantities<br />

of these catecholamines are still very low, i.e., highly selective<br />

and sensitive analytical methods are needed for such important<br />

biomarkers. The urinary VMA and HVA values are most useful<br />

and can be correlated with the stage of disease, management, any<br />

maturation of tumor, and prognosis from children with neuroblastoma.<br />

4 The urine HVA/VMA ratio could be a screening tool<br />

to support earlier detection of Menkes disease, a disorder that<br />

affects copper levels in the body, leading to copper deficiency. 7a<br />

The mechanism that removes HVA from the brain is still poorly<br />

understood, however, the efflux transport of HVA from the brain<br />

plays an important role in controlling the HVA level in the brain.<br />

This HVA efflux transport system is inhibited by several organic<br />

anions including IXS, and metabolites of monoamine neurotransmitters<br />

but not neurotransmitters per se. 7b Thus, it is of clinical<br />

importance to develop a rapid and sensitive method for simultaneous<br />

analysis of HVA, VMA, and IXS in urine and other biological<br />

samples.<br />

(3) Tumur, Z.; Niwa, T. Am. J. Nephrol. 2009, 29, 551–557.<br />

(4) Liebner, E. J.; Rosenthal, I. M. Cancer 2006, 32 (3), 623–633.<br />

(5) Eisenhofer, G.; Lenders, J. W. M.; Linehan, W. M.; Walther, M. M.;<br />

Goldstein, D. S.; Keiser, H. R. New Engl. J. Med. 1999, 340, 1872–1879.<br />

(6) Lenders, J. W. M.; Keiser, H. R.; Goldstein, D. S.; Willemsen, J. J.; Friberg,<br />

P.; Jacobs, M.-C.; Kloppenborg, P. W. C.; Thien, T.; Eisenhofer, G. Ann.<br />

Intern. Med. 1995, 123, 101–109.<br />

(7) (a) Menkes, J. H.; Alter, M.; Steigleder, G. K.; Weakley, D. R.; Sung, J. H.<br />

Pediatrics 1962, 29, 764–779. (b) Mori, S.; Takanaga, H.; Ohtsuki, S.;<br />

Deguchi, T.; Kang, Y.-S.; Hosoya, K.-I.; Terasaki, T. J. Cereb. Blood Flow<br />

Metab. 2003, 23, 432–440.<br />

(8) (a) Issaq, H. J.; Delviks, K.; Janini, G. M.; Muschik, G. M. J. Liquid<br />

Chromatogr. Relat. Technol. 1992, 15/18, 3193–3201.<br />

10.1021/ac101105q © 2010 American <strong>Chemical</strong> Society 6895<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/16/2010


HVA and VMA in infant urine has been analyzed by capillary<br />

electrophoresis (CE) with UV detection with a detection limit<br />

(LOD) of 3.5 × 10 -4 M for HVA and 1.8 × 10 -4 M for VMA. 8a A<br />

simultaneous determination of VMA, HVA, creatinine, and uric<br />

acid uses capillary electrophoresis and a 30 mM phosphate<br />

buffer (pH 7.0) containing 150 mM sodium dodecyl sulfate<br />

(SDS). The detection is by UV absorbance at 245 nm and the<br />

run is rather lengthy (15 min). 9 Although the authors claimed<br />

lower LODs for both HVA and VMA (5.5 × 10 -5 M and 5.0 ×<br />

10 -5 M), such LODs are still significantly above the normal<br />

levels found in healthy subjects (8.2 to 41 µM for HVA and<br />

11.6 to 28.7 µM for VMA). 10 Besides high LOD, UV detection<br />

is also problematic as biological samples often consist of several<br />

compounds with strong absorption at low wavelengths. The<br />

neurotransmitters can be conjugated with a strong fluorophore<br />

and analyzed by MEKC with fluorescence detection to achieve<br />

high sensitivity. 11a This approach might be feasible for routine<br />

analysis of a few compounds but becomes impractical and timeconsuming<br />

for multiple analytes. The use of CE-MS for analysis<br />

of VMA, HVA, and other biomarkers for metabolic disorders<br />

in newborns is available elsewhere. 11b CE equipped electrochemical<br />

detection (ECD) using a bare silica capillary can be<br />

used to detect HVA and VMA after electrophoretic separation<br />

at pH 5. 12a,b However, HVA and VMA are not baseline separated<br />

and other catecholamines including IXS are not included. 12a HVA<br />

and VMA also become less electroactive at pH > 5, the minimal<br />

required pH for electrophoretic separation of HVA and VMA. In<br />

general, bare fused silica capillaries at high pH are used in such<br />

studies to resolve IXS, HVA, and VMA. The separation is lengthy<br />

since negatively charged IXS, HVA, and VMA migrate to the<br />

anode, i.e., opposite flow direction to the electroosmotic flow<br />

(EOF). In addition, HPLC with isocratic elution and spectrophotometric<br />

detection are often used for analysis of tryptophan<br />

metabolites including IXS. 12c HPLC is also coupled with mass<br />

spectrometry to detect compounds associated with purple urinary<br />

bag syndrome (PUBS) and IXS is one of these toxic<br />

compounds. 12d Micellar electrokinetic chromatography (MEKC)<br />

using a bare fused silica capillary and laser induced fluorescence<br />

detection (a KrF excimer laser, λ ) 248 nm) has been used to<br />

detect HVA, VMA, and IXS. Under the best condition with this<br />

expensive instrumentation, the detection limit for HVA and VMA<br />

is 170 nM and 150 nM, respectively. 12e<br />

This work describes a novel scheme for the analysis of IXS,<br />

HVA, and VMA in the presence of tryptophan and other important<br />

catecholamines and indoleamines (Table 1). The fused silica<br />

capillary is coated with a thin layer of poly(diallyl dimethylammonium)<br />

chloride (PDDA) or gold nanoparticles (AuNPs) embed-<br />

(9) Shirao, M. K.; Suzuki, S.; Kobayashi, J.; Nakazawa, H.; Mochizuki, E.<br />

J. Chromatogr. B 1997, 693, 463–467.<br />

(10) Garcia, A.; Heinanen, M.; Jimenez, L. M.; Barbas, C. J. Chromatogr. A 2000,<br />

871, 341–350.<br />

(11) (a) Caslavska, J.; Gassmann, E.; Thormann, W. J. Chromatogr. A 1995,<br />

709, 147–156. (b) Senk, P.; Kozak, L.; Foret, F. Electrophoresis 2004, 25,<br />

1447–1456.<br />

(12) (a) Li, X. J.; Jin, W. R. Chin. Chem. Lett. 2002, 13 (9), 874–876. (b) Li,<br />

X. J.; Jin, W. R.; Weng, Q. F. Anal. Chim. Acta 2002, 461 (1), 123–130. (c)<br />

Marklova, E.; Makovickova, I.; Krakorova, I. J. Chromatogr. A 2000, 870,<br />

289–293. (d) Bar-Or, D.; Rael, L. T.; Bar-Or, R.; Craun, M. L.; Statz, J.;<br />

Garrett, R. E. Clin. Chim. Acta 2007, 378, 216–218. (e) Paquette, D. M.;<br />

Sing, R.; Banks, P. R.; Waldron, K. C. J. Chromatogr. B: Biomed. Sci. Appl.<br />

1998, 714 (1), 47–57.<br />

6896 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

ded in PDDA to reverse the EOF, allowing fast migration of IXS,<br />

VMA, HVA, and tryptophan. In contrast, catecholamines and<br />

indoleamines migrate against the EOF and emerge very late in<br />

the electropherogram, i.e., they do not interfere with the analysis<br />

of IXS, VMA, HVA, and tryptophan. The presence of AuNPs plays<br />

an important role in baseline separation of several compounds<br />

and a mechanism is given to decipher the interaction between<br />

AuNPs and the analytes. Although the composite consisting of<br />

AuNPs embedded in PDDA has been reported by Chen et al., 13<br />

this is the first systematic application of an AuNP-PDDA coated<br />

capillary for simultaneous analysis of IXS, HMA, and VMA in urine<br />

samples. Together with sample stacking, the applicability of this<br />

approach for analysis of such important biomarkers in urine<br />

samples with improved detection sensitivity is also presented and<br />

discussed in detail.<br />

EXPERIMENTAL SECTION<br />

<strong>Chemical</strong>s. Poly(diallyl dimethylammonium) chloride (PDDA,<br />

MW ) 200 000-350 000, 20 wt % in water), hydrogen tetrachloroaurate<br />

tetrahydrate (HAuCl4 · 4H2O), Tris (hydroxymethyl)aminomethane,<br />

phosphoric acid (H3PO4), and other chemicals<br />

were purchased from Sigma (Dublin, Ireland). Unless otherwise<br />

stated, a 50 mM H3PO4 solution was adjusted to pH 3.0<br />

with 0.5 M Tris buffer and used as the separation buffer. The<br />

standard stock solutions (5.0 mM) of the analytes were<br />

prepared daily in deionized water. All solutions were prepared<br />

in Milli-Q ultrapure water and filtered through a 0.22 µm pore<br />

size membrane followed by sonication for 5 min prior to use.<br />

Synthesis of PDDA-Gold Nanoparticle (AuNPs) Composite.<br />

The PDDA-AuNP composite was prepared as described by<br />

Chen et al. 13 PDDA (250 µL, 4% wt. in H2O), 40 mL of H2O, 200<br />

µL of 0.5 M NaOH, and 300 µL of HAuCl4 (10 mg/mL) were<br />

thoroughly mixed in a beaker, covered with an inverted culture<br />

dish for 2 min. The mixture was then maintained at 100 °C for<br />

30 min, resulting in a ruby red solution. The UV-visible<br />

spectrum of the AuNP colloid was recorded on a HP 8453<br />

UV-visible spectrophotometer in a1cmoptical path quartz<br />

cuvette. The size distribution of AuNPs in PDDA was measured<br />

using a zetasizer Nano ZS system (Malvern Instruments, MA)<br />

which is based on a dynamic light scattering (DLS) technique.<br />

TEM micrographs were obtained by a Delong LVEM (Soquelec,<br />

Montreal, QC, Canada) low-voltage TEM at 5 kV. A<br />

small amount of PDDA-AuNPs was sonicated to disperse the<br />

material. A 20 µL sample of well dispersed suspension was then<br />

dried on a Formvar-carbon coated grid and analyzed.<br />

Preparation of Coated Capillaries. A fused-silica capillary<br />

(50 µm id and 365 µm od) purchased from Polymicro Technologies<br />

(Phoenix, AZ, USA) was cut to 45 cm as the effective capillary<br />

length. In order to expose the maximum number of silanol groups<br />

on the silica surface, the fused-silica capillary was rinsed with 1.0<br />

M NaOH and deionized water for 15 min each. The preconditioned<br />

capillary was then rinsed with the PDDA or the PDDA-AuNP<br />

(13) Chen, H.-J.; Wang, Y.-L.; Wang, Y.-H.; Dong, S.-J.; Wang, E. Polymer 2006,<br />

47 (2), 763–766.<br />

(14) (a) Luong, J. H. T.; Male, K. B.; Glennon, J. D. Analyst 2009, 134 (10),<br />

1965–1979. (b) Kraft, A. Int. J. Electrochem. Sci. 2007, 2, 355–385. (c) Swain,<br />

G. M.; Ramesham, R. Anal. Chem. 1993, 65, 345–351. (d) Xi, J.; Granger,<br />

M.; Chen, Q.; Strojek, K.; Lister, T.; Swain, G. Anal. Chem. 1997, 69, 591A–<br />

597A. (e) Tenne, R.; Patek, K.; Hashimoto, K.; Fujishima, A. J. Electroanal.<br />

Chem. 1993, 347, 409–415.


Table 1. <strong>Chemical</strong> Structure, pKa Values and Aqueous Solubilities (A, g/L) of Various Analytes<br />

solution for 15 min followed by an incubation period of 15 min.<br />

The coated capillary was gently rinsed with deionized water for 3<br />

min to flush out unadsorbed coating materials. Before the first<br />

run, the coated capillary was equilibrated with the running buffer<br />

for 15 min but only for 3 min between the runs. All of these<br />

procedures were performed at 25 °C. For overnight or prolonged<br />

storage, the capillary was rinsed with deionized water for 15 min<br />

and then stored with the capillary ends dipped in deionized water.<br />

Capillary Electrophoresis with Electrochemical Detection.<br />

The capillary outlet was epoxy sealed into a pipet tip so that only<br />

∼1 cm protruded. The pipet tip was firmly attached vertically into<br />

a micromanipulator (HS6, World Precision Instruments, Sarasota,<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6897


FL, USA) with three-dimensional adjustment capabilities. A<br />

cylindrical cathodic/detection reservoir (2 cm diameter ×1 cm<br />

height) contained Pt wires (1 mm in diameter, 99.9% purity),<br />

serving as the counter electrode for amperometric detection and<br />

the cathode for electrophoresis. An Ag/AgCl (3 M NaCl) reference<br />

electrode was placed vertically into the reservoir, whereas the<br />

BDD electrode was inserted upward from the reservoir’s bottom<br />

and sealed with epoxy (the working reservoir volume was ∼3 mL).<br />

The micromanipulator and a laboratory jack (to which the<br />

reservoir was solidly mounted) were attached to a solid breadboard<br />

to prevent movement during alignment. The capillary outlet<br />

was aligned to the detecting electrode using the micromanipulator<br />

with the aid of a surgical microscope (World Precision Instruments).<br />

The capillary outlet was adjusted until it touched the<br />

electrode surface (evident by a slight bend in the capillary<br />

observed by microscopic inspection) and it was then backed off<br />

25-30 µm using the micromanipulator’s z-control. The BDD<br />

electrode was connected to an electrochemical workstation<br />

(CHI660C, CH Instruments, Austin, TX, USA) consisting also of<br />

a platinum wire (1 mm in diameter) as counter electrode and an<br />

Ag/AgCl (3 M NaCl) electrode as reference electrode. The BDD<br />

electrode (3 mm in diameter, 0.1% doped diamond) was purchased<br />

from Windsor Scientific (Slough, Berkshire, U. K.). The analytes<br />

are detected by a boron doped diamond (BDD) electrode which<br />

is positioned close to the capillary outlet. The BDD electrode is<br />

advocated in this work due to its stable low background current<br />

and a wide applied potential window. BDD is also resistant to<br />

fouling due to the hydrogen surface termination and sp 3 carbon<br />

bonding (no extended pi-electron system). 14a,b BDD can be<br />

considered as general-purpose working electrodes with a broad<br />

array of applications for use with HPLC and CE. Pioneering work<br />

in the early 1990s was conducted by Swain and co-workers 14c,d<br />

and the Fujishima group. 14e<br />

The electrophoretic separation was conducted at -10 kV<br />

(reversed polarity) unless otherwise stated. A plastic cap with a<br />

central hole of ∼1 mm was firmly attached to the surface of the<br />

BDD electrode to reduce the active sensing area. The analyte<br />

sample was injected electrokinetically for 5sat-10 kV. Peak<br />

identification was based on the migration time of a single standard<br />

with that of unknown peaks. However, if the resolution between<br />

any peak pair was low, then peak identification was performed<br />

by spiking both solutes individually. The pKa values and aqueous<br />

solubility of the analytes were obtained using the ACD/<br />

Structure Designer software (Advanced <strong>Chemistry</strong> Development,<br />

Toronto, ON, Canada). The degree of ionization was<br />

estimated as pH ) pKa+ log [A - /AH] for acid and pH )<br />

pKb+log [BH + /B] for base (the Henderson-Hasselbalch<br />

equation).<br />

RESULTS AND DISCUSSION<br />

Performance of the PDDA Coated Capillary. PDDA was<br />

firmly adsorbed on the inner walls of the capillary via ionic<br />

interactions between the negatively charged SiO - of fused silica<br />

and the quaternary ammonium groups of the polymer. Indeed,<br />

immersion of a substrate (glass, quartz, silica wafer, gold, silver,<br />

and even Teflon) into an aqueous 1% solution of this positively<br />

charged polymer results in the strong adsorption of a mono-<br />

6898 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 1. Electropherograms obtained using a PDDA coating<br />

capillary (50 µm id and 45 cm effective length) for the separation of<br />

20 µM IXS, 25 µM VMA, 25 µM HVA, 25 µM TRP, 100 µM<br />

isoproterenol (ISP), 100 µM normetanephrine (NMN), 100 µM<br />

epinephrine (EP), 50 µM 5-hydroxytryptamine (5-HT), 125 µM 4-hydroxy-3-methoxybenzylamine<br />

(HMBA), and 75 µM tryptamine (TA).<br />

The running buffer consisted of 50 mM H3PO4-Tris, (a) pH 3, (b) pH<br />

4, and (c) pH 5. The separation voltage was applied at -10 kV with<br />

an injection time of 5s at -10 kV. BDD at +1.0 V vs Ag/AgCl, 3 M<br />

NaCl.<br />

layer (1.6 nm) of PDDA on the substrate. 15,16 The adsorption<br />

of a PDDA thin film on a glass substrate was also reported<br />

elsewhere. 17 Notice also that the charge of PDDA is not pH<br />

dependent, as reflected by constant EOF in the range pH 2-8<br />

provided the capillary is coated with high-molecular weight<br />

PDDA. 18 Thus, PDDA with MW of 200 000-350 000 was used in<br />

this study for coating the capillary.<br />

At -10 kV, except for the epinephrine (EP) and normetanephrine<br />

(NMN) pair, all analytes were baseline resolved when 50 mM<br />

H3PO4-Tris pH 3.0 was used as the running buffer (Figure 1,<br />

curve a). On the basis of the calculated pKa values for the analytes<br />

(Table 1), fully deprotonated and highly negatively charged IXS<br />

exhibited high electrophoretic mobility and migrated concomitantly<br />

with EOF as the first peak in the electropherogram. The<br />

carboxylate/carboxyl ratio, estimated as 10 (pH-pK a) ,is∼0.16 and<br />

0.04 for VMA and HVA, respectively. Thus, VMA should<br />

emerge before HVA and slightly ahead of EOF. At pH 3, the<br />

neutral form of TRP should be predominant (neutral TRP/TRP +<br />

) 4.2);therefore, it should migrate very closely to EOF and<br />

trail behind both VMA and HMA. The EP-NMN pair was<br />

slightly split further by conducting the separation at pH 4 with<br />

improved detection sensitivity but the running time was also<br />

slightly longer and VMA emerged very close to IXS (Figure 1,<br />

curve b). The run was lengthier at pH 5, with only 4 discernible<br />

peaks emerging in the electropherogram as HVA comigrated with<br />

VMA and TRP became more neutral at this pH and emerged far<br />

behind the HVA peak. The remaining analytes acquired more<br />

negative charges and interacted strongly with positively charged<br />

PDDA and were not eluted after 1200 s into the experiment. The<br />

detection sensitivity was also greatly compromised at this running<br />

(15) Kotov, N. A.; Harazsti, T.; Turi, L.; Zavala, G.; Geer, R. E.; Dekany, I.;<br />

Fendler, J. H. J. Am. Chem. Soc. 1997, 119, 6821–6832.<br />

(16) Moriguchi, I.; Teraoka, Y.; Kagawa, S.; Fendler, J. H. Chem. Mater. 1999,<br />

1, 1603–1608.<br />

(17) Hrapovic, S.; Liu, Y.; Enright, G.; Bensabaa, F.; Luong, J. H. T. Langmuir<br />

2003, 19, 3958–3965.<br />

(18) Wang, Y.; Dubin, P. L. Anal. Chem. 1999, 71, 3463–3468.


pH with a tilted baseline (Figure 1, curve c). It is of interest to<br />

compare the migration order of these peaks with MEKC performed<br />

at normal electrophoretic separation conditions by Caslavska<br />

et al. 11a In such a study, a buffer composed of 75 mM sodium<br />

dodecyl sulfate (SDS), 6 mM Na2B4O7, and 10 mM Na2HPO4<br />

(pH 9.2) is used with the separation potential set at +20 kV.<br />

The migration sequence is TRP, HVA, VMA, and IXS, which<br />

is opposite to the migration order shown in Figure 1. The<br />

remaining 7 catecholamines and indoleamines with a positive<br />

charge from the ammonium ion (NH3 + ) and/or the secondary<br />

amino group (-NH + -) were completely ionized at pH 3 and<br />

migrated to the cathode with high mobilities, i.e., countercurrent<br />

to the EOF. However, they were eventually driven out the<br />

capillary by EOF with higher mobility.<br />

Notice that a buffer composed of 200 mM boric acid, 100 mM<br />

potassium hydroxide, and 0.1% hydroxyethylcellulose, pH 9.2, has<br />

been used to separate VMA from various indoleamines under<br />

normal separation. 19 In this case, VMA trailed far behind other<br />

indoleamines, as expected from its negative charge at this pH.<br />

PDDA can also be added as a buffer additive as described by<br />

Tseng et al., 20 resulting in high and reversed EOF. The mobility<br />

of indolamines and catecholamines decreases as the PDDA<br />

concentration increases. The separation of 14 analytes including<br />

indolamines, catecholamines, and metanephrines is achieved<br />

within 33 min under optimal separation conditions (1.2% PDDA<br />

and 5 mM formic acid at pH 4.0). Indeed, PDDA has been used<br />

to coat fused silica capillary to form a thin film for the subsequent<br />

absorption of carbon nanotubes. 21 Besides adsorption, PDDA can<br />

be chemically bonded onto the interior capillary wall with an<br />

anodal EOF independent of pH ranging from 2.2 to 8.8. 22 The<br />

lifetimes of both the bonded and physically coated capillaries<br />

exceeded 40 h of continuous use at 240 V/cm at pH 4. Our<br />

experimental data confirmed that the PDDA could be reused for<br />

several repeated runs and the capillary was easily reconditioned<br />

as described earlier. The coated capillary also exhibited good<br />

tolerance to methanol and 0.1 HCl.<br />

Separation on Capillary Coated with PDDA-Gold Nanoparticles.<br />

Our next strategy was to form gold nanoparticles<br />

(AuNPs) on the capillary wall, since they have been known to<br />

interact with several compounds with amino, hydroxyl, and<br />

carboxylic groups. 23 In principle, AuNPs can be prepared separately<br />

and adsorbed on the PDDA layer. This concept has been<br />

used to coat the glass microchip channel with citrate-stabilized<br />

AuNPs for the separation of phenols. 24 Another approach is to<br />

form AuNPs by electroless plating via hydroxylamine-mediated<br />

reduction. 17 However, the charge of the PDDA-AuNP layer returns<br />

to negative and the separation must be performed under normal<br />

separation. Indeed, covalent attachment can be used to immobilize<br />

AuNPs onto the capillary wall, for instance, the preparation of<br />

dodecanethiol AuNPs on prederivatized 3-aminopropyl-trimethox-<br />

(19) Stocking, C. J.; Slater, J. M.; Simpson, C. F. Exp. Nephrol. 1998, 6, 415–<br />

420.<br />

(20) Tseng, W.-L.; Chen, S.-M.; Hsu, C.-Y.; Hsieh, M.-M. Anal. Chim. Acta 2008,<br />

613 (1), 108–115.<br />

(21) Luong, J. H.T.; Bouvrette, P.; Liu, Y.; Yang, D.-Q.; Sacher, E. J. Chromatogr.<br />

A 2005, 1 (2), 187–194.<br />

(22) Liu, Q.; Lin, F.; Hartwick, R. A. J. Chromatogr. Sci. 1997, 35 (3), 126–130.<br />

(23) Zhong, Z. Y.; Patskovskyy, S.; Bouvrette, P.; Luong, J. H. T.; Gendanken,<br />

A. J. Phys. Chem. B 2004, 108 (13), 4046–4052.<br />

(24) Pumera, M.; Wang, J.; Grushka, E.; Polsky, R. Anal. Chem. 2001, 73, 5625–<br />

5628.<br />

Figure 2. Electropherograms obtained using a PDDA-AuNPs coated<br />

capillary at pH 3, 4, and 5. Other conditions were same as Figure 1.<br />

ysilane or 3-mercaptopropyl-trimethoxysilane fused-silica capillaries.<br />

25 Again; several steps are required for the preparation of such<br />

modified capillaries.<br />

In order to maintain the positive charge for the coating layer,<br />

a one step procedure described by Chen et al. 13 was used to<br />

prepare the PDDA-AuNP composite since PDDA acts as both<br />

reducing and stabilizing agents for AuNPs. The PDDA-gold colloid<br />

exhibited a surface plasmon resonance (SPR) around 527-528<br />

nm, which could be related to AuNPs with the mean diameter of<br />

27-28 nm, in agreement with the report on the SPR peak obtained<br />

for AuNPs with an average diameter of 27 nm. 26 Dynamic light<br />

scattering confirmed that AuNPs should have a mean diameter<br />

of 28 nm with narrow particle size distribution. A TEM micrograph<br />

of PDDA-AuNPs indicated that AuNPs exhibited an average size<br />

of 25 nm (figure not shown). Notice that the excellent stability of<br />

AuNPs arises from the electrosteric effect of PDDA, in agreement<br />

with Mayer et al. 27 where PDDA is simply used as a protecting<br />

and stabilizing agent for AuNPs. Under the same electrophoretic<br />

and detection condition for the PDDA coated capillary, the<br />

electropherogram obtained for the 10 analytes using the PDDA-<br />

AuNP coated capillary is shown in Figure 2. At pH 3, EP was<br />

satisfactorily separated from NMN and the peaks were significantly<br />

sharper with low background current, resulting in significantly<br />

improved detection sensitivity. For the microchip channel<br />

with coated PDDA/AuNPs layer-by-layer, improved resolution and<br />

detection sensitivity has been reported for the three aminophenols,<br />

although the rationale behind such behavior is not known. 24 Our<br />

experimental data also confirmed that the run was longer at pH<br />

4 without any improvement in the separation of the NMN-EP pair<br />

with slightly reduced detection sensitivity. At pH 5, only four peaks<br />

(IXS, VMA, HVA and TRP) emerged in the electropherogram with<br />

the TRP peak far behind the HVA peak.<br />

The circumstances determining the effect of the PDDA-AuNP<br />

coated capillary are complex. The viscosity of the PDDA-AuNP<br />

solution is about 90% of the PDDA solution, i.e., it is easier to<br />

pump the PDDA-AuNP solution through the capillary to form a<br />

(25) (a) O’Mahony, T.; Owens, V. P.; Murrihy, J. P.; Guihen, E.; Holmes, J. D.;<br />

Glennon, J. D J. Chromatogr. A 2003, 1004 (1-2), 181–193. (b) Yang, L.;<br />

Guihen, E.; Holmes, J. D.; Loughran, M.; O’Sullivan, G. P.; Glennon, J. D.<br />

Anal. Chem. 2005, 77 (6), 1840–1846.<br />

(26) Nasir, S. M.; Nur, H. J. Fundam. Sci. 2008, 4, 245–252.<br />

(27) Mayer, A. B. R.; Hausner, S. H.; Mark, J. E. Polym. J. 2000, 32, 15–22.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6899


Figure 3. Electropherograms obtained using a PDDA-AuNP coated capillary with the BDD electrode for the study of (A) buffer concentration<br />

effect: 25, 50, 75, and 100 mM H3PO4-Tris, pH 3, (B) injection time effect (3, 5, 7, and 10 s), (C) effect of detection potential (+0.6, +0.8, +1.0,+<br />

and 1.2 V vs Ag/AgCl, 3 M NaCl), (D) effect of separation voltage (-5, -10, -15, and -20 kV). The sample mixture consisted of 20 µM IXS,<br />

25 µM (each) VMA, HVA, TRP, 100 µM ISP, 100 µM NMN, 100 µM EP, 50 µM 5-HT, 125 µM HMBA, 125 µM DHBA, and 75 µM TA.<br />

homogeneous layer. 28 The incorporation of AuNPs in the polymer<br />

network could be attributed to the structural change in pristine<br />

PDDA after the synthesis of AuNPs. It has been shown that CdC<br />

and C-N are produced after PDDA reacts with HAuCl4 by<br />

(28) Zhang, Z.-X.; Yan, B.; Liu, K.; Liao, Y.-P.; Liu, H.-W. Electrophoresis 2009,<br />

30, 379–387.<br />

6900 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

comparing the FTIR spectra of PDDA and PDDA-AuNPs. 13 For<br />

the latter, a new peak appears at 1578 cm -1 , which can be<br />

assigned for CdC and CdN in-plane vibration. In aqueous<br />

milieu, gold nanoparticles are negatively charged and the<br />

absolute value of the �-potential decreases as the pH decreases,<br />

whereas above pH 5.8, the absolute value of the �-potential


Table 2. Linearity and Detection Limit of the Analytes a,b<br />

analytes linear range (µM) calibration equation R2 detection limit (µM)<br />

indoxyl sulfate (IXS) 1.00-40 I ) 0.4839X+0.1490 0.9998 0.3<br />

vanillylmandelic acid (VMA) 1.25-50 I ) 0.3635X+0.3036 0.9998 0.4<br />

homovanillic acid (HVA) 1.25-50 I ) 0.3034X+0.5667 0.9998 0.4<br />

tryptophan (TRP) 1.25-50 I ) 0.3025X+1.0870 0.9998 0.4<br />

isoproterenol (ISP) 5.00-200 I ) 0.0871X+0.7481 0.9995 1.7<br />

normetanephrine (NMN) 5.00-200 I ) 0.0865X+0.6429 0.9995 1.7<br />

epinephrine (EP) 5.00-200 I ) 0.0945X+0.3850 0.9995 1.7<br />

5-hydroxytryptamine (5-HT) 2.50-100 I ) 0.1686X+0.9364 0.9995 0.8<br />

4-hydroxy-3-methoxybenzylamine (HMBA) 6.25-250 I ) 0.0822X+0.0446 0.9991 2.1<br />

3,4- dihydroxybenzylamine (DHBA) 6.25-250 I ) 0.0866X+0.3842 0.9991 2.1<br />

tryptamine (TA) 3.75-150 I ) 0.1476X+0.5307 0.9995 1.3<br />

a I ) peak current (nA), X ) analyte concentration (µM). b Electrophoretic separation was carried out using a PDDA-AuNP coated capillary (50<br />

µm id, 365 µm od) with an effective length of 45 cm, running buffer, 50 mM H3PO4-Tris, pH 3.0, separation voltage, -10 kV, injection for 5s at -10<br />

kV. Detection: BDD electrode at +1.0 V vs. Ag/AgCl, 3M NaCl.<br />

Figure 4. The comparison electropherograms for (a) nonsample<br />

stacking using a running buffer of 50 mM H3PO4-Tris, pH 4 with an<br />

injection time of 5 s (b) sample stacking, injection buffer: 10 mM<br />

H3PO4-Tris, pH 2, and the running buffer of 50 mM H3PO4-Tris, pH 4<br />

with an injection time of 10 s. (c) sample stacking, injection buffer:<br />

10 mM H3PO4-Tris, pH 2, and the running buffer was similar to the<br />

injection buffer with an injection time of 10 s. Concentration of<br />

analytes: 50 µM (each) IXS, VMA, AA, HVA, UA, TRP. Separation<br />

voltage: -10 kV. BDD at +1.0 V vs Ag/AgCl, 3 M NaCl.<br />

becomes almost constant (∼-45 to -50 mV). The negative<br />

charge of the gold surface may be explained by the fact that<br />

the Au-OH present on the gold surface can lose protons as<br />

the pH increases to form Au-O-groups on the nanoparticle<br />

surface. 29 AuNPs have a partially hydroxylated surface with a<br />

pKa value of 3.2 (Au-OH) in the presence of water. 29 At pH 3,<br />

around the pKa value of AuNPs, -OH groups are predominant<br />

(-OH groups/O - groups ) 1.6). The rest of the gold surface<br />

should be metallic, i.e., essentially hydrophobic. Therefore,<br />

AuNPs would display ionic and hydrophobic interaction, as well<br />

as hydrogen bonding with the PDDA network. The synthesis<br />

and incorporation of AuNPs in the polymer network would<br />

affect the PDDA structural change, which in turn increased<br />

the charge density and coverage efficiency of the coating. At<br />

pH4or5,Au-O - groups were predominant, resulting in a<br />

(29) Sylvestre, J.-P.; Kabashin, A. V.; Sacher, E.; Meunier, M.; Luong, J. H. T.<br />

J. Am. Chem. Soc. 2004, 126, 7176–7177.<br />

decrease of the �-potential of the PDDA-AuNPs composite, i.e.,<br />

the EOF is reduced. Notice that TRP also becomes more<br />

neutral at pH 5, whereas the charges of the catechoamines and<br />

indolamines were highly negative and such analytes were<br />

expected to interact strongly with PDDA. Indeed, all catecholamines<br />

and indoleamines were not eluted after 24 min into the<br />

experiment. In contrast, Zhang et al. 28 reported that the<br />

separation using the PDDA coated capillary is significantly<br />

longer than the one modified with PDDA-AuNPs for analysis<br />

of heroin and impurities. In such a study, the separation is<br />

performed at pH 5.2 with addition of 3% methanol into the<br />

running buffer consisting of 120 mM ammonium acetate.<br />

Therefore, it was somewhat difficult to compare the result<br />

obtained in this work with that of Zhang et al. 28<br />

Other Optimal Conditions and Sample Stacking. The 50<br />

mM phosphoric-Tris buffer pH 3 appeared to provide the highest<br />

detection sensitivity with good separation resolution. Indeed,<br />

4-hydroxy-3-methoxybenzylamine comigrated with 3,4-dihydroxybenzylamine<br />

when the buffer strength increased to 100 mM and<br />

the run was longer with lower detection sensitivity (Figure 3A).<br />

This pair was coeluted as one single peak if the separation was<br />

carried out with the PDDA coated capillary. Thus, a running buffer<br />

consisting of 50 mM phosphoric-Tris, pH 3 was used to optimize<br />

the separation potential and the detection potential of the BDD<br />

electrode. With respect also to detection sensitivity and separation<br />

efficiency (N/cm), the electrokinetic injection of the sample at<br />

-10 kV for 5 s was optimal compared to the results obtained at<br />

shorter (3 s) or longer (7 and 10 s) injection times (Figure 3B).<br />

Beyond 5 s, the resulting peaks were very broad with compromised<br />

detection sensitivity and separation efficiency. In contrast,<br />

3 s was not sufficient, as reflected by considerably lower peaks.<br />

The BDD detection potential, poised at +1 V vs 3 M Ag/AgCl<br />

provided highest detection sensitivity compared to the detection<br />

performed at lower or higher applied potentials (Figure 3C). Both<br />

resolution and detection sensitivity were severely deteriorated<br />

when the separation was conducted at -15 kV and -20 kV,<br />

compared to the run at -10 kV (Figure 3D). Linearity and LOD<br />

obtained for 11 analytes are summarized in Table 2. The four fast<br />

migrating analytes, IXS, VMA, HVA, and tryptophan exhibited a<br />

LOD of 0.3 µM, whereas a higher LOD, ∼1 µM was obtained for<br />

the slower migrating group.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6901


Figure 5. Analysis of urine sample using a running buffer consisting<br />

of 50 mM H3PO4-Tris, pH 4. Separation voltage: -10 kV with an<br />

injection time of 5sat-10 kV. Detection: BDD at +1.0 V vs Ag/<br />

AgCl, 3 M NaCl. The urine samples were filtered using a 0.22 µm<br />

Millipore filter and diluted in the injection buffer with different dilution.<br />

(A) 10-fold dilution, pristine and spiked urine samples were presented<br />

as the lower and upper curve, respectively. Peak identification: (1)<br />

IXS; (2) VMA; (3) AA; (4) HVA; (5) UA; and (6) TRP. The diluted<br />

urine sample was spiked with 20 µM IXS, 20 µM VMA, 100 µM AA,<br />

20 µM HVA, 100 µM UA, and 20 µM TRP. (B) 15-fold dilution, pristine<br />

and spiked urine samples were presented as the lower and upper<br />

curve, respectively. Peak identification: (1) IXS; (2) VMA; (3) AA; (4)<br />

HVA; (5) UA; and (6) TRP. The diluted urine sample was spiked with<br />

20 µM IXS, 20 µM VMA, 100 µM AA, 20 µM HVA, 100 µM UA, and<br />

20 µM TRP.<br />

For further LOD improvement, a series of experiments based<br />

on field-amplified sample stacking (FASS) 30 was performed by<br />

exploiting the conductivity difference between the sample zone<br />

and the running buffer to effect preconcentration. A standard<br />

solution of 6 analytes including ascorbic and uric acids, two<br />

endogenous urinary compounds, was prepared in 10 mM H3PO4<br />

pH 2 and injected at -10 kVfor 10 s with the separation<br />

performed at -10 kV using 50 mM H3PO4 pH 4 as the running<br />

buffer. It should be noted that at pH 3, ascorbic acid migrated<br />

very closely with HVA, whereas uric acid was not baseline<br />

separated from TRP. In principle, the amount of stacking is<br />

proportional to the conductivity difference between the running<br />

(30) Weng, Q.-F.; Xu, G.-W.; Yuan, K.-L; Tang, P J. Chromatogr. B 2006, 835,<br />

55–61.<br />

6902 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 6. Analysis of an urine sample using a running buffer<br />

consisting of 50 mM H3PO4-Tris, pH 4. Separation voltage: -10 kV<br />

with an injection time of 5sat-10 kV. Detection: BDD at +1.0Vvs<br />

Ag/AgCl, 3 M NaCl. The urine sample was filtered using a 0.22 µm<br />

Millipore filter and diluted 8-fold in the injection buffer. (A) with sample<br />

stacking and (B) without sample stacking.<br />

buffer and the sample solution. A preconcentration factor of 4<br />

was observed for IXS, AA, and HVA compared to 3 for VMA<br />

and UA and only 1.25 for TRP (Figure 4, curve b). With sample<br />

stacking, the detection limit of IXS, AA, and HVA was 75 nM<br />

compared with nonstacking (300 nM, Figure 4, curve a), significantly<br />

lower than the physiological levels of these biomarkers in<br />

urine as discussed later. Notice that peak-width was broadened<br />

when the injection time was greater than 10 s (data not shown).<br />

In another attempt, the capillary was first filled with 50 mM H3PO4<br />

pH 4 followed by a sample injection as described above. The<br />

cathodic running buffer was changed to 10 mM H3PO4, pH2<br />

instead of 50 mM H3PO4, pH 4. In this case, the separation<br />

window was narrower, resulting in the comigration of three<br />

pairs: IXS/VMA, AA/HVA, and UA/TRP (Figure 4, curve c).<br />

No further improvement in separation resolution was attained by<br />

adding several organic modifiers including 10% methanol, 10%<br />

acetonitrile, 5 mM methyl �-cyclodextrin, and 5 mM cetyl<br />

trimethylammonium bromide in the running buffer.<br />

Analysis of Biomarkers in Urine. The electropherogram of<br />

an authentic urine sample obtained from a healthy female shows<br />

5 low peaks and one very high peak. All biomarkers were detected<br />

even if the urine sample was diluted to 10- and 15-fold, respectively<br />

in the injection buffer (Figure 5). Standard IXS, HVA, VMA,


ascorbic acid (AA), uric acid (UA), and TRP were spiked into the<br />

urine sample for peak identification. Dilution of the urine sample<br />

also improved peak to peak separation and peak identification.<br />

All peaks were identified and the sequence was assigned as IXS,<br />

VMA, AA, HVA, UA, and TRP. The highest peak was attributed<br />

to UA as expected since its normal concentration in urine ranges<br />

from 1.23 to 3.7 mM (1.48-4.43 mmol/day with an average urine<br />

volume of 1200 mL). 31 The UA peak was confirmed by matching<br />

the migration time and spiking the urine sample with 30 µM uric<br />

acid. Peak deconvolution and peak area were obtained using<br />

WinPLOTR (Version: May, 2009, http://www.cdifx.univ-rennes1.fr/<br />

winplotr/winplotr.htm). The concentration for each analyte was<br />

estimated by comparing the peak areas of the authentic and spiked<br />

urine samples. The average concentration of IXS, VMA, HVA, AA,<br />

and UA for two different urine samples obtained from the same<br />

healthy female was determined to be 170, 55, 87, 142, and 1,075<br />

µM, respectively. Normal levels of such analytes in urine are<br />

100-1000 µM for IXS, 32 50-1000 µM for VMA, 10 14-125 µM for<br />

HVA, 10 and 150-200 µM for AA, 33 i.e., the values estimated by<br />

CE-ECD fall in the normal range for healthy subjects. Therefore,<br />

sample staking was not mandatory for the simultaneous analysis<br />

of IXS, HVA, and VMA in the presence of UA, AA, and TRP since<br />

the LOD of such analytes was significantly lower than their normal<br />

physiological levels.<br />

Sample stacking of the urine samples at low dilution (below<br />

5-fold) did not improve the detection limits (figure not shown),<br />

very likely due to a high level of salts in the urine sample. With<br />

an 8-fold dilution, sample stacking provided a significant improvement<br />

in the LOD with a sharper corresponding peak for each<br />

analyte as shown in Figure 6A. In particular, sample stacking also<br />

allowed for the detection of TRP, the last peak that migrated very<br />

(31) Tietz, N. W. Fundamentals of Clinical <strong>Chemistry</strong>: W.B. Saunders Co.:<br />

Philadelphia, 1987.<br />

(32) Harlit, H. J. Biol. Chem. 1933, 537–545.<br />

(33) Ridi, E.; Moubasher, M. S. R.; Hassan, Z. F. Biochem. J. 1951, 49, 246–<br />

251.<br />

(34) Matsuo, M.; Tasaki, R.; Kodama, H.; Hamasaki, Y. J. Inherit. Metab. Dis.<br />

2005, 28 (1), 89–93.<br />

(35) Luo, D.; Wu, L.; Zhi, J. ACS Nano 2009, 3 (8), 2121–2128.<br />

closely to the uric acid peak. In addition, both HVA and VMA<br />

were positively detected and identified compared with nonsample<br />

stacking (Figure 6B). Thus, the method became useful for<br />

estimation of the urine HVA/VMA ratio, a useful screening<br />

method for Menkes disease. This ratio could range from 4.1 to<br />

69.7 owing to impaired activity of dopamine �-hydroxylase, a<br />

copper-dependent enzyme. 34<br />

CONCLUSIONS<br />

In brief, a novel scheme was described for electrophoretic<br />

separation and detection of several important biomarkers in urine.<br />

A fused silica capillary with a coating layer of PDDA and AuNPs<br />

reversed the electroosmotic flow and served as a stable layer for<br />

resolving the analytes. No electrode fouling was observed during<br />

repeated analysis and even with the urine sample. The detection<br />

limit obtained for IXS, HVA, and VMA was considerably below<br />

their normal physiological levels in biological samples. The<br />

method was simple and capable of measuring several important<br />

biomarkers in urine samples without sample pretreatment with<br />

excellent selectivity and detection sensitivity. Sample stacking<br />

could be easily performed to improve detection limits of these<br />

analytes in urine samples provided such samples were diluted<br />

properly to reduce the level of uric acid and salts. Furthermore,<br />

many other basic neurotransmitters and their acidic metabolites<br />

could also be detected. Notice also that a BDD nanoforest<br />

electrode (BDDNF) can be fabricated by hot filament chemical<br />

vapor deposition. 35 This type of electrode exhibits improved<br />

detection sensitivity compared to conventional planar BDD<br />

electrodes. Integration of BDDSNF with capillary electrophoresis<br />

is a subject of future endeavor.<br />

ACKNOWLEDGMENT<br />

The authors thank the Science Foundation Ireland (SFI) for<br />

an SFI Walton Fellowship (JHTL), an IRCSET Embark Award<br />

(LZ), and an SFI-SRC Grant for the Irish Separation Science<br />

Cluster (ISSC).<br />

Received for review April 27, 2010. Accepted July 5, 2010.<br />

AC101105Q<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6903


Anal. Chem. 2010, 82, 6904–6910<br />

Human Plasma Copper Proteins Speciation by Size<br />

Exclusion Chromatography Coupled to Inductively<br />

Coupled Plasma Mass Spectrometry. Solutions for<br />

Columns Calibration by Sulfur Detection<br />

Souleiman El Balkhi, †,§,| Joël Poupon,* ,† Jean-Marc Trocello, ‡,§ France Massicot, |<br />

France Woimant, ‡,§ and Olivier Laprévote †,|<br />

Laboratoire de toxicologie biologique, Service de Neurologie, Centre de Référence Pour la Maladie De Wilson,<br />

AP-HP, Hôpital Lariboisière, 2, rue Ambroise Paré, 75475 Paris cedex 10, France, and ChimiesToxicologie<br />

Analytique et Celullaire, EA 4463, Faculté de Pharmacie, Université Paris Descartes, 4, avenue de l’Observatoire,<br />

75006 Paris, France<br />

Among the hyphenated techniques used to probe and<br />

identify metalloproteins, size exclusion chromatography<br />

coupled to inductively coupled plasma mass spectrometry<br />

(SEC-ICP-MS) has shown to have a central place. However,<br />

the calibration of SEC columns reveals to be tedious and<br />

always involves UV detection prior to ICP-MS. The presence<br />

of sulfur in 98% of proteins allows their detection by quadrupole<br />

ICP-MS, despite the isobaric interference ( 16 O 16 O) on<br />

S, by monitoring 32 S 16 O at mass to charge ratio (m/z) 48.<br />

The formation of SO occurs spontaneously in the argon<br />

plasma but can be optimized by the introduction of oxygen<br />

gas into a reaction cell (RC) to achieve nM levels. In this<br />

article, sulfur detection was discussed upon instrumental<br />

conditions and S detection was then optimized by applying<br />

O 2 as a reaction gas. SO formation was used to calibrate<br />

SEC columns without UV detection. This simple SEC-ICP-<br />

MS method was used for plasma copper proteins in plasma<br />

healthy subjects (HS) and an untreated Wilson disease<br />

(WD) patient. Copper proteins identified in healthy subjects<br />

were transcuprein, ceruloplasmin (Cp) and albumin. The<br />

method led to results in good agreement with other methods<br />

of determination. Copper bound to Cp in the WD<br />

patient was lowered with regard to the HS, and the exchangeable<br />

Cu was highly increased.<br />

In the postgenomics era, proteomics has become a central branch<br />

in life sciences. Information from proteomics (and peptidomics)<br />

studies may reveal alterations in gene expression due to a disease<br />

and facilitate the understanding of the metabolic pathways such as<br />

copper metabolism in Wilson and Menkes diseases. 1 Since one-third<br />

* Author for correspondence: Dr Joël Poupon, Laboratoire de Toxicologie<br />

biologique, Hôpital Lariboisière, 2 rue Ambroise Paré, 75475 Paris cedex 10,<br />

France, E-Mail: joel.poupon@lrb.aphp.fr, Phone: 33 1 49 95 66 00, Fax: 33 1 49<br />

95 65 71.<br />

† Laboratoire de toxicologie biologique.<br />

‡ Service de Neurologie.<br />

§ Centre de Référence Pour la Maladie De Wilson.<br />

| Université Paris Descartes.<br />

6904 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

of all proteins are metalloproteins, a new field recently defined by<br />

Haraguchi, namely “metallomics”, has been introduced. This topic<br />

attracted a fast growing interest during the past decade. It integrates<br />

different analytical approaches that focus on metalloproteins and<br />

metal-containing biomolecules. Metal ions in metalloproteins have a<br />

regulatory and structural role allowing them to fulfill their different<br />

functions. The biocatalysis of some specific enzymatic reactions<br />

including gene DNA synthesis, metabolism, antioxidation, and<br />

detoxification are examples of these functions. 2,3<br />

Since the separation, identification, and quantification of<br />

metalloproteins are mandatory in our understanding of their<br />

complex role in biological systems, a growing number of studies<br />

are covering these fields. Usually, authors employ hyphenated<br />

techniques combining a separation method followed by mass<br />

spectrometry analysis. Matrix-assisted laser desorption/ionization<br />

(MALDI) MS, electrospray ionization (ESI) MS, two-dimensional<br />

polyacrylamide gel electrophoresis (2D-PAGE) laser ablation (LA)<br />

ICP-MS and high performance liquid chromatography (HPLC)<br />

ICP-MS with different modes of separation have been proposed.<br />

More recently, capillary electrophoresis (CE) ICP-MS and Fourier<br />

transform ion cyclotron resonance mass spectrometry (FT-ICR-<br />

MS) have been used for this purpose. 3,4<br />

The overwhelming majority of recent applications concerning<br />

probing and quantification of metalloproteins were developed<br />

using ICP-MS coupled to size-exclusion chromatography (SEC). 5-9<br />

ICP-MS is the most widely used mode of detection because of its<br />

extremely low limits of detection (LOD), a wide dynamic range,<br />

(1) Wang, M.; Feng, W. Y.; Zhao, Y. L.; Chai, Z. F. Mass Spectrom. Rev. 2009,<br />

29, 326–348.<br />

(2) Haraguchi, H. J. Anal. At. Spectrom. 2004, 19, 5–14.<br />

(3) Shi, W.; Chance, M. R. Cell. Mol. Life Sci. 2008, 65, 3040–3048.<br />

(4) Szpunar, J. Analyst 2005, 130, 442–465.<br />

(5) Cabrera, A.; Alonzo, E.; Sauble, E.; Chu, Y. L.; Nguyen, D.; Linder, M. C.;<br />

Sato, D. S.; Mason, A. Z. Biometals 2008, 21, 525–543.<br />

(6) Franca Maltez, H.; Villanueva Tagle, M.; Fernandez De La Campa Maria<br />

Del, R.; Sanz Medel, A. Anal. Chim. Acta 2009, 650, 234–240.<br />

(7) Gonzalez-Fernandez, M.; Garcia-Barrera, T.; Arias-Borrego, A.; Bonilla-<br />

Valverde, D.; Lopez-Barea, J.; Pueyo, C.; Gomez-Ariza, J. L. Anal. Bioanal.<br />

Chem. 2008, 390, 17–28.<br />

(8) Lopez-Avila, V.; Sharpe, O.; Robinson, W. H. Anal. Bioanal. Chem. 2006,<br />

386, 180–187.<br />

(9) Sviridov, D.; Meilinger, B.; Drake, S. K.; Hoehn, G. T.; Hortin, G. L. Clin.<br />

Chem. 2006, 52, 389–397.<br />

10.1021/ac101128x © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/19/2010


the multielement capabilities, the continuous spectra signals<br />

survey and a high sample throughput. 10 In addition, SEC has the<br />

advantage, in theory, to preserve metalloproteins from solid<br />

phase/analytes interactions. Maintaining the integrity of the<br />

probed proteins is actually of a critical importance when metal<br />

ions simply coordinate to organic ligands.<br />

Nevertheless, the calibration of SEC columns presents some<br />

drawbacks rending SEC-ICP coupling laborious because of (i) a<br />

tedious installation of a UV detector in parallel to the ICP-MS,<br />

(ii) the use of an additional software for the integration of<br />

chromatographic peaks, and (iii) the increase of dead volume.<br />

This way of SEC calibration involves the injection of high<br />

concentrations of proteins. These are expensive and can irreversibly<br />

bound to columns leading to columns clogging. It is<br />

noteworthy that other detection techniques have been proposed<br />

to calibrate SEC columns such as IR, NMR, MS, evaporative light<br />

scattering, and viscometers 11 but their coupling to ICP has not<br />

yet been reported.<br />

The capability of ICP-MS to detect in a single run the metal<br />

ions of interest and hetero nonmetallic elements present in<br />

proteins such as S and P offers great opportunities. It allows a<br />

specific detection, molecular characterization, and even quantification<br />

of some metalloproteins. For instance, when the amino acid<br />

(AA) sequence is known for a given protein, and hence the<br />

number of AA containing S, the determination of S provides an<br />

accurate value of molar protein concentration. 12,13<br />

Unfortunately, these heteroelements (i.e., S and P) have high<br />

ionization energies resulting in low sensitivities in ICP-MS.<br />

Moreover, both S and P are extensively interfered by isobaric<br />

polyatomic ions generated in the atmospheric pressure argon<br />

plasma ( 16 O 16 O for 32 S and 15 N 16 O for 31 P). The use of high<br />

resolution mass spectrometry systems such as sector field ICP-<br />

MS with a resolution higher than 1 800 m/∆m 14-17 allows<br />

distinguishing of these isobaric species. Alternatively, the detection<br />

of S remains feasible with lower resolution (m/∆m 400) quadrupole<br />

ICP-MS by using a reaction cell. This device is a quadrupole<br />

filled with a reaction gas (oxygen or xenon) and placed between<br />

the ion lenses and the quadrupole mass filter of the ICP-MS<br />

instrument. Oxygen used as reacting gas in the reaction cell<br />

converts S ions into 32 S 16 O. Sulfur is then measured as SO at<br />

m/z 48 which shows less interferences. This technique was<br />

previously used for the determination of some sulfured amino<br />

acids. 18,19 In the same way, the measurement of phosphorylation<br />

degree of some peptides 20 and the determination of metal-sulfur<br />

ratios in some metalloproteins 12 were obtained by detecting S and<br />

(10) Lobinski, R.; Moulin, C.; Ortega, R. Biochimie 2006, 88, 1591–1604.<br />

(11) Striegel, A. M. Anal. Chem. 2005, 77, 104 A113 A.<br />

(12) Hann, S.; Koellensperger, G.; Obinger, C.; Furtmuller, P. G.; Stingeder, G.<br />

J. Anal. At. Spectrom. 2004, 19, 74–79.<br />

(13) Profrock, D.; Leonhard, P.; Prange, A. Anal. Bioanal. Chem. 2003, 377,<br />

132–139.<br />

(14) Bettmer, J.; Montes Bayon, M.; Encinar, J. R.; Fernandez Sanchez, M. L.;<br />

Fernandez de la Campa Mdel, R.; Sanz Medel, A. J. Proteomics 2009, 72,<br />

989–1005.<br />

(15) Rappel, C.; Schaumloffel, D. Anal. Bioanal. Chem. 2008, 390, 605–615.<br />

(16) Wind, M.; Wesch, H.; Lehmann, W. D. Anal. Chem. 2001, 73, 3006–3010.<br />

(17) Zinn, N.; Kruger, R.; Leonhard, P.; Bettmer, J. Anal Bioanal. Chem. 2008.<br />

(18) Schaumloffel, D.; Giusti, P.; Preud’Homme, H.; Szpunar, J.; Lobinski, R.<br />

Anal. Chem. 2007, 79, 2859–2868.<br />

(19) Yeh, C. F.; Jiang, S. J.; Hsi, T. S. Anal. Chim. Acta 2004, 502, 57–63.<br />

(20) Bandura, D. R.; Baranov, V. I.; Tanner, S. D. Anal. Chem. 2002, 74, 1497–<br />

1502.<br />

Pas 32 S 16 O + and 33 P 16 O + species (m/z 48 and 49, respectively)<br />

together with metals of interest. More recently, Wang used<br />

SEC-ICP-MS for a quantitative analysis of proteins via S<br />

determination. 21<br />

Interestingly, by using argon plasma, under atmospheric<br />

condition, oxides are formed spontaneously at percentages corresponding<br />

to ≈3% of measured elements. 22,23 These percentages<br />

are influenced by instrumental parameters of ICP-MS. 24,25 Consequently,<br />

in applications involving high concentrations of proteins,<br />

such as undiluted or weakly diluted plasma, the amount of<br />

produced SO could be sufficient for its detection. Moreover,<br />

recommended concentrations of proteins in SEC column calibration<br />

kits are around 1-10 g · L -1 (i.e., S concentration higher<br />

than 50 µM) which is enough to monitor SO without any<br />

addition of O2. In other words, the calibration of SEC columns<br />

should be possible by using SEC-ICP-MS in standard mode<br />

and without any prior UV detection.<br />

In this study, we present a simple SEC-ICP-MS method for<br />

plasma copper proteins speciation. This method was applied to<br />

healthy subjects (HS) and an untreated Wilson disease (WD)<br />

patient. The distribution of copper in the blood is still an important<br />

subject of research. The knowledge of copper distribution contributes<br />

to better understand copper metabolism and to better<br />

diagnose and monitor related diseases. In WD, a mutation in the<br />

gene ATP7B leads to a dysfunction of ceruloplasmin (Cp) which<br />

is the major protein binding Cu. This binding is mediated by the<br />

protein ATP7B that lacks in WD. Clinically, serum Cp concentration<br />

diminishes and the so-called “free Cu” increases and becomes<br />

toxic due to Cu deposits in target organs (liver, brain, kidney,<br />

and eyes). If not treated, irreversible damages can occur. 26-28<br />

For the speciation purpose, we demonstrate that SEC column<br />

calibration can be conducted with a simple SEC-ICP-MS at<br />

moderate m/z resolution and without any UV detection. ICP-MS<br />

operational parameters influence on oxides formations in standard<br />

mode (without O2) was studied. The method was then optimized<br />

by the application of O2 as reaction gas in the dynamic reaction<br />

cell (DRC). Excellent limits of detection (LODs) for S and Cu<br />

were obtained after optimization. Copper proteins identification,<br />

stability and quantification are explored.<br />

EXPERIMENTAL SECTION<br />

Reagents and <strong>Chemical</strong>s. Mobile phase (NH4NO3 200 mM)<br />

was prepared daily by dissolving 16 g of NH4NO3 (Sigma<br />

<strong>Chemical</strong> Co, St-Quentin Fallavier, France) in 1Lofultrapure<br />

water Milli-Q (Millipore, Molsheim, France) and degassed by<br />

vacuum filtration on a 0.22 µm Millipore filters.<br />

For SEC column calibration, two SEC marker kits were used.<br />

Kit 1 (obtained from Sigma) included: Blue dextran (2000 kDa),<br />

thyroglobulin (670 kDa), apoferritin (443 kDa), �-amylase (200<br />

(21) Wang, M.; Feng, W.; Lu, W.; Li, B.; Wang, B.; Zhu, M.; Wang, Y.; Yuan,<br />

H.; Zhao, Y.; Chai, Z. Anal. Chem. 2007, 79, 9128–9134.<br />

(22) Divjak, B.; Goessler, W. J. Chromatogr., A 1999, 844, 161–169.<br />

(23) Du, Z.; Houk, R. S. J. Anal. At. Spectrom. 2000, 15, 383–388.<br />

(24) Gray, A. L.; Williams, J. G. J. Anal. At. Spectrom. 1987, 2, 599–606.<br />

(25) Vaughan, M. A.; Horlick, G. Appl. Spectrosc. 1986, 40, 434–445.<br />

(26) Das, S. K.; Ray, K. Nat. Clin. Pract. Neurol. 2006, 2, 482–493.<br />

(27) El Balkhi, S.; Poupon, J.; Trocello, J. M.; Leyendecker, A.; Massicot, F.;<br />

Galliot-Guilley, M.; Woimant, F. Anal. Bioanal. Chem. 2009, 394, 1477–<br />

1484.<br />

(28) Linder, M. C.; Hazegh-Azam, M. Am. J. Clin. Nutr. 1996, 63, 797S–811S.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6905


Table 1. ICP-MS Parameters<br />

nebulizer<br />

glass SeaSpray concentric<br />

(nominal flow:1 mL · min-1 )<br />

spray chamber glass cyclonic baffled (50 mL)<br />

nebulizer gas flow 0.92 mL min-1 auxiliary gas flow 1.20 mL min-1 plasma gas flow 15 L min-1 ICP RF power 1125 W<br />

O2flow rate (DRC mode) 0.60 (0.1 - 1mL· min-1 )<br />

RPQ 0.25 (0.05-0.7)<br />

pulse stage voltage 800 V<br />

analog stage voltage -1612.5 V<br />

measured m/z<br />

32 16 34 16 63 65 70 45 151 S O, S O, Cu, Cu, Ga Rh, Eu<br />

scan mode peak hopping<br />

dwell time per isotope 250 ms<br />

kDa), alcohol dehydrogenase (150 kDa), bovine serum albumin<br />

(BSA) (66 kDa), carbonic anhydrase (29 kDa). Human serum<br />

albumin (HSA) (69 kDa) was needed for void volume (V0)<br />

determination (see results). A second ready-to-use marker kit<br />

(Kit 2) (Column Performance Check Std, Aqueous SEC 1,<br />

Phenomenex Inc., Le Pecq, France) was used to control SEC<br />

calibration containing bovine thyroglobulin (670 kDa), human<br />

gamma globulin (150 kDa), ovalbumin (44 kDa), and human<br />

myoglobin (17 kDa).<br />

D-penicillamine (D-pen) (purity >97%), chosen as a source of<br />

S, was obtained from Sigma. Cu, Ga, Eu, and Rh 1 g · L -1 standard<br />

solutions (Inorganic Ventures, distributed by Analab, Bischeim,<br />

France) were used to prepare working solutions by appropriate<br />

dilution in the mobile phase. Ga, Rh, and Eu were tested as<br />

internal standards (IS). All reagents were tested for copper<br />

contamination before use. Milli-Q water was used to make<br />

intermediate dilutions.<br />

Instrumentation. Separation of proteins was performed via<br />

SEC using a BioSep-SEC-S 2000 and Biosep-SEC-S 3000 (300 ×<br />

7.8 mm both) columns from Phenomenex. A Series 200 Perkin-<br />

Elmer chromatography system, composed of an injector and a<br />

high pressure pump equipped of PEEK tubing, (Perkin-Elmer,<br />

Courtaboeuf, France) was operating at a flow rate of 1 mL · min -1 .<br />

The injection volume was 100 µL. The column effluent was<br />

directly coupled to a1mL· min -1 glass SeaSpray concentric<br />

nebulizer (Perkin-Elmer) and a 50 mL cyclonic Baffled spray<br />

chamber from Perkin-Elmer. Oxygen (purity N48, Air Liquide<br />

Santé, Puteaux, France) was used as the reaction gas. Detection<br />

and quantification of elements was made by an ICP-MS Elan-<br />

DRCe (Perkin-Elmer) and the Chromera software (Perkin-<br />

Elmer) was used for the HPLC signal monitoring. Signal<br />

intensity was given as counts per seconde (Cps). The ICP-MS<br />

operation parameters are given in Table 1.<br />

The two isotopes of Cu were used for copper detection and<br />

quantification in all experiments because of the polyatomic<br />

interference 40 Ar 23 Na that was observed on 63 Cu. The ion peak<br />

at m/z 63 was attributed to pure Cu when ( 63 Cu/ 65 Cu) isotopes<br />

ratio was 0.45 ± 5%.<br />

DRC and Reaction Gas Flow Optimization. A10µg · L -1<br />

solution of D-pen (source of S) was used for the optimization<br />

of the oxygen reaction gas flow in the DRC. The optimal<br />

flow was obtained by varying the O2 flow (from 0.1 to 1.0<br />

mL · min -1 ) as so to attain the highest signal to background<br />

ratio (S/BKG). However, the effect of O2 on the Cu signal<br />

6906 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(Cps) had to be estimated. Therefore, a solution of D-pen<br />

and Cu at 200 and 0.5 µM, respectively, was used to measure<br />

SO and Cu signals by varying the oxygen gas flow in the<br />

DRC.<br />

SEC Column Calibration. Proteins were dissolved in the<br />

mobile phase and 100 µL of the proteins solutions of Kit 1 were<br />

injected on columns: apoferritin, �-amylase, alcohol dehydrogenase,<br />

BSA and carbonic anhydrase at 10, 5, 4, 5, and 10 g · L -1 ,<br />

respectively. These concentrations are recommended by the<br />

manufacturer and are used for a UV detection to give A280 nm<br />

≈ 1ona90× 1.6 cm column. In our study, the proteins were<br />

detected by their sulfur content (i.e., 32 S 16 O + referred to as 48 SO<br />

hereafter) by ICP-MS without any reaction gas in the DRC. A<br />

calibration curve was calculated between of the known molecular<br />

masses of proteins and their respective volume of elution<br />

divided by the dead volume (Ve/V0). To determine the void<br />

volume, a1g· L -1 solution of HSA was injected separately,<br />

vortex mixed with blue dextran (Bdx at 3 g · L -1 ) and injected<br />

again (see results for explanation).<br />

After a 100 fold dilution of the previous proteins solutions in the<br />

mobile phase, the same volume was injected while the optimal flow<br />

of oxygen gas in the DRC was applied. The calibration curve was<br />

calculated again with the new data. At last, to confirm the proteins<br />

separation, two mixtures were prepared and injected: (1) the readyto-use<br />

mixture and (2) a mixture of apoferritin, alcohol dehydrogenase,<br />

BSA and carbonic anhydrase at 1, 15, 6, 15 g · L -1 , respectively.<br />

Copper and Sulfur Calibration Curves. External calibrations<br />

for Cu and S were performed by using five solutions containing<br />

increasing amounts of Cu and D-pen. Concentrations were chosen<br />

on the basis of expected amounts of metalloproteins in the plasma<br />

samples (0.1-5.0 µM for Cu and 100-5000 µM for S). Ga was<br />

added as IS (100 µg · L -1 ) to all calibration solutions. External<br />

calibrations were, first, carried out by flow injection (FI) without<br />

any column connected to the HPLC system. The calibrations<br />

were operated in both standard (STD) and DRC modes. The<br />

peaks’ areas of measured elements reported to IS areas were<br />

tested for linearity. The same experiments were conducted<br />

under HPLC condition by injecting the calibration points on<br />

the Precolumn Biosep 2000.<br />

Species Unspecific Calibration. For Cu and S quantification<br />

in metalloproteins, the external calibration previously obtained by<br />

FI was used for the quantification of all Cu and S peaks in the<br />

chromatograms by using the peak areas. This is referred to as<br />

the “species unspecific” mode of calibration. 12,13<br />

Limits of Detection (LODs). LODs were calculated for each<br />

measured elements in both modes (STD and DRC) using FI first<br />

and the precolumn afterward. Sulfur and Cu LODs were calculated<br />

as 3-fold the standard deviation (SD) of the baseline variations<br />

for each elements on three blank injections.<br />

Sample Preparation. Blood samples were collected on a Liheparin<br />

Vacutainer tubes (ref. 365952, Becton- Dickinson, Le Pont<br />

de Claix, France) centrifuged for 10 min at 2000g, diluted (1:3) in<br />

a mobile phase solution containing 100 µg · L -1 of IS and injected<br />

on the column within 15 min. For exchangeable Cu study,<br />

determined and discussed in a previous work, 27 plasmas were<br />

diluted (1:1) in a 3 g· L -1 EDTA solution, kept at room<br />

temperature for one hour precisely and injected on the column<br />

after a final dilution (1:2) in the mobile phase.


RESULTS AND DISCUSSION<br />

Sulfur Detection by SO Signal in Standard Mode. Sulfur<br />

is present in two naturally occurring amino acids, cysteine, and<br />

methionine, together exhibiting a global abundance of about 5%<br />

of all amino acids in eukaryotic proteins. Therefore sulfur is<br />

present in the vast majority of proteins (>98%) and its direct<br />

detection by ICP-MS constitutes an almost general screening<br />

method for peptides and proteins. 14,15<br />

The high first ionization potential (10.36 eV) of S 15 makes the<br />

ionization efficiency in argon plasma to be less than 15% leading<br />

to low S sensitivity in ICP-MS. However, when Sulfur is present<br />

at concentrations sufficiently elevated its survey as 48 SO in<br />

standard mode is still possible due to a spontaneous oxide<br />

formation corresponding to ≈3% of the total element (i.e., SO<br />

formation in normal atmospheric pressure argon plasma).<br />

Attempts to increase SO signal in the standard mode by varying<br />

RF power or nebulizer gas flow (not shown) in order to facilitate<br />

oxides formation did not lead to satisfying results. This is in<br />

disagreement with Divjak et al. who’s results found that cold<br />

plasma increases CeO formation and expected that S would<br />

behave the same. 22 Indeed, cold plasma and high nebulizer gas<br />

flow were favorable for the formation of CeO in our experiments<br />

as well. However, increasing gas flow and decreasing RF was<br />

deleterious for both Cu and SO signals. The reaction between Ce<br />

and O2 is highly exothermic and was selected by some<br />

manufacturers as an example of a bimolecular reaction whose<br />

rate is inversely proportional to temperature. 29 Unfortunately,<br />

this is not the case of the reaction between S and O2 that seems<br />

to be endothermic under atmospheric pressure.<br />

SEC Column Calibration. Standard Mode. The separation of<br />

proteins on a SEC column and their detection by 48 SO offer the<br />

possibility to calibrate the column on the basis of their apparent<br />

molar mass (Mr) vs their volume of elution. This strategy<br />

permits to overcome the tedious installation of a UV detector<br />

in parallel to the ICP-MS and to use only one data integration<br />

software instead of two needed in general.<br />

The proteins of the marker kit 1 at the recommended<br />

concentrations were used to calibrate the Biosep 3000 column.<br />

Respective Ve of reference proteins are shown in Figure 1. To<br />

calculate SEC Calibration curve V0 is needed to plot molar<br />

masses vs Ve/V0. However, V0 could not be determined by Bdx<br />

alone because it does not have any measurable element by ICP-<br />

MS. This obstacle could be sidestepped as follows: giving that<br />

HSA easily binds to Bdx 30,31 we tried to produce what manufacturer<br />

advice against (i.e., mixing Bdx with other proteins of the marker<br />

kit). In fact, the injection of 100 µL ofBdxat2g· L -1 did not differ<br />

from baseline for all measured elements even for 12 C. On the other<br />

hand, HSA eluted at 11.3 min (namely Ve ) 11.3 mL as shown in<br />

Figure 2). HSA was then dissolved directly ina3g· L -1 Bdx solution<br />

(HSA ) 1g· L -1 ), incubated for one hour and injected on the<br />

column. We noticed that HSA 48 SO peak disappeared while a new<br />

peak at 7.31 min (namely Ve ) 7.31 mL) appeared. This peak<br />

results from the binding of HSA to Bdx and allows measuring<br />

the void volume V0. In such a way, all data needed to calculate<br />

the calibration curve are obtained (Figure 3).<br />

(29) Baranov, V. I.; Tanner, S. D. J. Anal. At. Spectrom. 1999, 14, 1133–1142.<br />

(30) Antoni, G.; Casagli, M. C.; Bigio, M.; Borri, G.; Neri, P. Ital. J. Biochem.<br />

1982, 31, 100–106.<br />

(31) Ponder, E.; Ponder, R. V. J. Gen. Physiol. 1960, 43, 753–758.<br />

Figure 1. 48 SO chromatograms of proteins injected separately.<br />

Elution volume of proteins: (1) apoferritin (10 g/L) Ve ) 10.45 mL, (2)<br />

�-amylase (4 g/L) Ve ) 11.32, (3) alcohol dehydrogenase (5 g/L) Ve<br />

) 11.64, (4) bovine albumin (10 g/L) 11.94 mL, (5) carbonic<br />

anhydrase (30 g/L) Ve ) 12.9 mL, (6) cytochrome C (2 g/L) Ve )<br />

13.71 mL. 1, 2, and 4 are on the left scale; 3, 5, and 6 are on the<br />

right scale.<br />

Figure 2. Void volume determination by interaction of Human<br />

albumin and Blue dextran. 48 SO signal in STD mode for human<br />

albumin (1 g/L) (blue line), and Human albumin bound to blue dextran<br />

(1 g/L and 3 g/L, respectively) (red line).<br />

The separation of proteins from Kits 1 and 2 was satisfying<br />

(not shown). However, It is noteworthy that myoglobin peak is<br />

hardly distinguishable from the baseline because of a poor<br />

sulfured AA content (only two methionines and one cysteine) and<br />

its coelution with low Mr sulfured residues present in all proteins.<br />

Signals Optimization for SO Detection in DRC Mode. The<br />

O2 flow offering the highest S/BKG for SO was found to be<br />

0.6 mL · min -1 . At this optimal flow, Cu signal for both isotopes<br />

decreases by about 20-30% compared to STD mode. Above<br />

0.6 mL · min -1 O2 all signals decreased.<br />

Sulfured Proteins in DRC Mode. By using O2 as a reacting<br />

gas in the DRC, S detection was markedly improved to achieve<br />

nanomolar levels (see after LOD). This implies the possibility<br />

(1) to inject lower concentrations of proteins on the column<br />

and so preserving it from irreversible proteins adsorption, (2)<br />

to quantify precisely proteins even those with a low sulfur<br />

content, (3) to use minor isotope 34 S by following 34 S 16 O( 50 SO)<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6907


Figure 3. SEC calibration curve between the known molar mass of<br />

each protein and their respective volume of elution divided by the<br />

dead volume (Ve/V0).<br />

and finally, (4) to use S/Cu ratios as a specific tool to Cu<br />

proteins detection since these ratios are theoretically known<br />

(Table 2 adapted from refs 32 and 33). Actually, when S/Cu ratio<br />

for a detected peak is higher than the theoretical value, proteins<br />

coelution should be suspected.<br />

Proteins injected separately could be identified by their<br />

calculated molar masses on the calibration curve confirming the<br />

validity of the column calibration. Figure 4 shows an example of<br />

chromatograms of cytochrom C (2 g · L -1 ) and B-amylase (4<br />

g · L -1 ). 48 SO signals are 1000 fold higher after O2 introduction<br />

and 50 SO becomes readily measurable offering a very wide<br />

dynamic range for S measurement.<br />

Copper and Sulfur Calibration Curves. By flow injection<br />

as well as under HPLC conditions, linearity was found for both<br />

Cu isotopes and in both modes (STD and DRC). 48 SO in STD<br />

mode was linear up to 10 000 µM and only to 2000 µM inDRC<br />

mode because of signal saturation. 50 SO linearity test was only<br />

possible in the DRC mode because it was not detectable in<br />

standard mode. The possibility to measure both isotopes<br />

permits to choose one or other depending on S concentrations<br />

in measured samples. All coefficients of variations (CV) of three<br />

injections of each of the calibration points in the same day were<br />

less than 3% and the coefficients of correlation (r 2 ) were higher<br />

than 0.99 for calibrated elements. In ICP techniques, internal<br />

standard is used to compensate all kind of instrumental drifts.<br />

The use of Ga as internal standard improves r 2 to reach 0.999<br />

and higher giving equal slopes for both Cu isotopes in STD<br />

and DRC modes. Rh was abandoned as internal standard<br />

because it binds to plasma proteins and Eu revealed to be less<br />

sensitive to the instrumental drift than Cu and S. None of the<br />

tested proteins could be used as source of S for calibration<br />

because of the sulfured residues of small molar mass detected<br />

in all protein samples. In addition, manufacturer did not<br />

mention the purity degree of the used proteins. This explains<br />

the lack of accuracy that we observed when we tried to quantify<br />

them by the obtained calibration curves.<br />

(32) Manley, S. A.; Byrns, S.; Lyon, A. W.; Brown, P.; Gailer, J. J Biol. Inorg.<br />

Chem. 2009, 14, 61–74.<br />

(33) Michalski, W. P. J Chromatogr., B 1996, 684, 59–75.<br />

6908 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Limits of Detection. The introduction of O2 in the DRC and<br />

the use of a chromatographic support are sensitive parameters<br />

influencing LODs of Cu and S and had to be estimated. LODs<br />

of 65 Cu, 63 Cu, 48 SO, and 50 SO are given in Table 3. 48 SO LOD<br />

turned out to be about 27 000 nM in standard mode and about<br />

264 nM in DRC mode. This confirms that the detection of<br />

48 SO in standard mode is still possible when S concentrations<br />

are sufficiently high. On the other hand, when 48 SO saturates<br />

in DRC mode (as it is the case of weakly diluted plasmas) 50 SO<br />

could be used to monitor and measure sulfur despite its high<br />

LOD.<br />

Plasma Copper Proteins Analysis. Detection. Three healthy<br />

subjects plasmas were injected pure or diluted (1:3) on the<br />

columns. Neither Biosep 2000 nor Biosep 3000 were able to<br />

separate ceruloplasmin copper (Cp-Cu) from albumin copper<br />

(Alb-Cu) peaks. Only three peaks were observed in these<br />

plasmas: the first one corresponded to transcuprein copper<br />

(TC-Cu) (270 kDa), the second to (Cp-Cu) (132 KDa) coeluted<br />

with (Alb-Cu) and the last one corresponded to low molar mass<br />

Cu containing molecules (LMr). Therefore, the two columns were<br />

connected in series to improve the separation. The connected<br />

columns were then calibrated as previously described (in DRC<br />

mode by monitoring 48 SO signal of different proteins at concentrations<br />

10-100 fold lower than recommended). Excellent<br />

separation between the Cp-Cu and the Alb-Cu peaks was<br />

obtained (Figure 5A). The following peaks were detected:<br />

TC-Cu, a major peak of Cu-Cp, Alb-Cu, and one LMr peak. It<br />

should be noted that the obtained molar masses are relative to<br />

the SEC column calibration and further analysis are needed to<br />

clearly identified the detected proteins. Transcuprein, for instance,<br />

was recently sequenced and its molar mass determined, 34 but one<br />

should be aware that the retention volume is not sufficient to<br />

identify this protein.<br />

Manley et al. 32 have been able to detect Cu coagulation factor<br />

Va (FVa-Cu), TC-Cu, Cp-Cu, Alb-Cu, and LMr-Cu in freshly<br />

sampled rabbit plasma by injecting a large quantity (500 µL) of<br />

undiluted plasma. The concentrations of FVa-Cu and Alb-Cu<br />

they found were surprisingly abnormal as noted by authors. In<br />

order to detect FVa-Cu in human plasma, we sampled blood on<br />

a Li heparin tube (LiH 65 UI) to which 50 UI of extra LiH were<br />

added (in order to prevent any possible coagulation to which FVa<br />

participates). One hundred microliters of pure plasma were then<br />

injected immediately after centrifugation (within 10 min). No peak<br />

corresponding to the FVa-Cu was detected despite the immediate<br />

injection. This is not surprising because of the very low theoretical<br />

concentration in human plasma (FVa-Cu ) 30 nM approximately).<br />

Conversely, in the majority of published studies working<br />

on Cu proteins speciation, tris-hydroxyl-methyl-aminomethane<br />

(Tris) or acetate buffers were used. We noted that these mobile<br />

phases either precipitate ionic Cu or prevent it to be eluted form<br />

SEC columns. For instance, injection of CuNO3 or CuCl2 solutions<br />

(1-5 µM) on the tested columns was not reproducible at all<br />

when Tris or acetate buffers were used. In addition, the<br />

injection of EDTA 3 g/L after the injection of plasma or a simple<br />

aqueous Cu solution provided an important and variable peak<br />

of Cu confirming that SEC columns are able to bound Cu when<br />

inappropriate mobile phases are used. This is in concordance<br />

with Inagaki et al. and Meng et al. observations. 21,35 To avoid


Table 2. Molecular Properties and Concentrations of Cu Proteins and Stoichiometric S/Cu Ratio (Adapted from Refs<br />

31 and 32<br />

protein<br />

molar mass<br />

(kDa)<br />

number of Cu bound<br />

per protein<br />

this, Inagaki choose to pretreat plasmas on a chelating resin before<br />

SEC analysis to be sure that loosely bound Cu is captured.<br />

Plasma from the untreated Wilson disease patient was also<br />

tested. Despite a retention time shift that was observed, due<br />

to columns loss of efficiency, three major peaks were detected:<br />

Cp-Cu, Alb-Cu, and low molar mass Cu complexes (Figure<br />

5B). The loss of efficiency, is a well-known phenomena that<br />

occurs when SEC column performance deteriorates after a large<br />

number of injections or after injection of highly concentrated<br />

samples.<br />

Copper Proteins Quantification. In ICP-MS, a simple<br />

inorganic compound of an element can calibrate the element in<br />

all kind of proteins because the response is independent of the<br />

element environment. Therefore, the conventional calibration<br />

methods (internal and external standardization and standard<br />

additions) can be used for protein quantification but only if the<br />

analytes have been identified. 1,4 This is referred to as the species<br />

unspecific mode of calibration. 12,13<br />

Koellensperger et al demonstrated that isotope dilution remains<br />

the method of choice for species unspecific quantification in<br />

hyphenated ICP-MS techniques although requiring mathematical<br />

approximations. 36 They explained, however, that flow injection<br />

yielded acceptable uncertainties and could be used for quantifica-<br />

(34) Liu, N.; Lo, L. S.; Askary, S. H.; Jones, L.; Kidane, T. Z.; Trang, T.; Nguyen,<br />

M.; Goforth, J.; Chu, Y. H.; Vivas, E.; Tsai, M.; Westbrook, T.; Linder, M. C.<br />

J. Nutr. Biochem. 2007, 18, 597–608.<br />

(35) Inagaki, K.; Mikuriya, N.; Morita, S.; Haraguchi, H.; Nakahara, Y.; Hattori,<br />

M.; Kinosita, T.; Saito, H. Analyst 2000, 125, 197–203.<br />

number of sulfur<br />

containing AA<br />

S/Cu<br />

ratio<br />

plasma proteins<br />

concentration<br />

maximum Cu<br />

concentration (nM)<br />

blood coagulation factor V 330 1 17 17 10 mg/L 30.3<br />

transcuprein 190 0.5 56 112 180 µg/L 0.47<br />

ceruloplasmin 132 6 40 6.7 0.2-0.6 g/L 9090-27 272<br />

albumin 66 0-1 42 42 36-53.6 g/L<br />

Cu, Zn-SOD 32 0-2 0.014-0.021 mg/L<br />

Figure 4. 48 SO and 50 SO chromatograms of separate injections of<br />

�-amylase (4 g/L) and cytochrome C (2 g/L) in DRC mode. (1) 48 SO<br />

�-amylase; (2) 50 SO �-amylase; (3) 48 SO cytochrome C; (4) 50 SO<br />

cytochrome C; (5) sulfured residues.<br />

tion. In our study, no difference was observed between the<br />

calibration curves obtained by flow injection and those obtained<br />

by injection on the precolumn, allowing the quantification of all<br />

species by using the daily flow injection calibrations.<br />

Ceruloplasmin coeluted with other sulfured proteins making<br />

its quantification impossible for healthy subjects and WD. Nevertheless,<br />

Alb was well separated and could be measured by its<br />

50 SO peak and found at 47.3 g · L -1 (716 µM) to be compared<br />

with 49.7 g · L -1 obtained by a routine immunochemical analysis.<br />

Calculated concentrations of Cu species in a healthy subject<br />

showed that Transcuprein, Ceruloplasmin, Alb and LMr bound<br />

0.4%, 87.5% 6.0%, and 6.0% of total Cu, respectively, (namely 0.08,<br />

14.7, 1.0, and 1.0 µM) giving a total Cu in good agreement with<br />

the concentrations found by atomic absorption spectrometry<br />

(AAS) (16.8 vs 15.5 µM). For the WD patient, total Cu was at 4.4<br />

µM by HPLC-ICP-MS vs 3.6 µM by AAS and included: 14.9%<br />

Cp-Cu, 12.4% Alb-Cu, and 72.7% of LMr-Cu (namely 0.66, 0.55,<br />

and 3.2 µM, respectively).<br />

Unexpectedly, Cu-LMr represents 6% or more of total Cu in<br />

the healthy subject plasma. In our previous study, 27 normal values<br />

of ultrafilterable Cu (CuUF) were established (obtained by the<br />

ultrafiltration of 44 healthy subjects plasmas on a 30 kDa cutoff<br />

filter) and we know that this fraction does not exceed 0.15 µM in<br />

healthy subjects (less than 1% of total Cu). This means that only<br />

0.15 µM of Cu do not bind proteins having molar mass greater<br />

than 30 kDa. We have described in this previous work that loosely<br />

bound Cu is the copper fraction complexed to Alb and/or other<br />

proteins that can be mobilized in the presence of high copper<br />

affinity chelators, such as EDTA. To quantify this exchangeable<br />

Cu (CuEXC) fraction the most useful ETDA concentration was 3<br />

g/L with incubation time of one hour (time found necessary to<br />

reach exchange equilibration). Quantification of CuEXC in healthy<br />

subjects offered normal values ranging from 0.6 to 1.1 µM.<br />

Interestingly, the LMr fraction of the healthy subject (1 µM Cu)<br />

lies within this reference range and is likely to correspond to<br />

CuEXC. Similarly, the Cu-LMr fraction found in the WD patient<br />

(3.2 µmol/L) is in good agreement with the CuEXC fraction<br />

determined by EDTA-ultrafiltration (2.8 µmol/L).<br />

To confirm this, we injected the plasmas after one hour<br />

incubation with EDTA 3 g/L (1:1) in order to probe the CuEXC.<br />

No difference was observed between the chromatograms before<br />

and after incubation confirming that loosely bound Cu is released<br />

on the column after injection and further elutes in the low Mr<br />

range. This loosely bound Cu is originally Alb copper and the<br />

(36) Koellensperger, G.; Hann, S.; Nurmi, J.; Prohaska, T.; Stingeder, G. J. Anal.<br />

At. Spectrom. 2003, 18, 1047–1055.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6909


Table 3. LODs in nM for Measured Elements in Different Operating Conditions (NM: Not Measurable)<br />

flow injection flow injection precolumn precolumn<br />

measured masses STD mode DRC mode SEC STD mode SEC DRC mode<br />

63Cu 4.8 5.8 13.0 20.0<br />

65Cu 6.8 7.6 17.0 27.0<br />

48SO 27 077 264 55 760 747<br />

50SO NM 1816 NM 5241<br />

Figure 5. (A) chromatogram of diluted (1:3) healthy subject plasma<br />

and (B) Wilson disease untreated patient plasma by using Biosep<br />

3000 and Biosep 2000 connected in series. Cu peak (blue line): (1)<br />

TC-Cu; (2) Cp-Cu; (3) Alb-Cu; (4) LMWM-Cu. 50 SO peak (red<br />

line): (5) unknown protein(s); (6) albumin.<br />

type 1 copper from ceruloplasmin reported to be as labile<br />

copper. 37<br />

(37) Musci, G.; Fraterrigo, T. Z.; Calabrese, L.; McMillin, D. R. J. Biol. Inorg.<br />

Chem. 1999, 4, 441–446.<br />

6910 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Even if SEC columns are known to be neutral toward the<br />

analytes/solid phase interactions, our results show that weakly<br />

bound Cu by unspecific coordination with proteins such as<br />

albumin and the type 1 Cu of ceruloplasmin can be partially<br />

released. SEC columns are neutral toward proteins but not toward<br />

metals of the proteins if it is not structurally protected. The copper<br />

atoms in our case only coordinate to amino acids of the protein<br />

and can readily be captured by remaining silanol groups in this<br />

silica based column. Upon the profile obtained from the WD we<br />

better understand the deposits of Cu in target organs which is<br />

unlikely to happen in healthy subjects because their loosely bound<br />

Cu is much lower. We believe that this fraction represents the<br />

copper that proteins can deliver and that its quantification is useful<br />

in WD.<br />

CONCLUSION<br />

This method is designated to be applied to a larger number of<br />

WD patients in order to use it for diagnostic purpose. Moreover,<br />

a possible correlation between the loosely bound Cu and the<br />

clinical state should be explored in WD. The same approach could<br />

be undertaken for other metalloproteins where metal ions are<br />

bound by coordination such as zinc proteins. At last, the sulfur<br />

monitoring improved tremendously our proteins analysis method.<br />

It allowed the SEC column calibration and offered a wide dynamic<br />

range for S determination, and hence proteins quantification, by<br />

offering the possibility to use both S isotopes.<br />

ACKNOWLEDGMENT<br />

We thank the Association Bernard PEPIN pour la Maladie de<br />

Wilson for financial support.<br />

Received for review April 29, 2010. Accepted July 7, 2010.<br />

AC101128X


Anal. Chem. 2010, 82, 6911–6918<br />

Direct Quantification of Single-Molecules of<br />

MicroRNA by Total Internal Reflection<br />

Fluorescence Microscopy<br />

Ho-Man Chan, † Lai-Sheung Chan, ‡ Ricky Ngok-Shun Wong, ‡ and Hung-Wing Li* ,†<br />

Department of <strong>Chemistry</strong>, Hong Kong Baptist University, Kowloon Tong, Hong Kong, P.R. China, and Department of<br />

Biology, Hong Kong Baptist University, Kowloon Tong, Hong Kong, P.R. China<br />

MicroRNAs (miRNAs) express differently in normal and<br />

cancerous tissues and thus are regarded as potent cancer<br />

biomarkers for early diagnosis. However, the short length<br />

and low abundance of miRNAs have brought challenges<br />

to the established detection assay in terms of sensitivity<br />

and selectivity. In this work, we present a novel miRNA<br />

detection assay in single-molecule level with total internal<br />

reflection fluorescence microscopy (TIRFM). It is a solution-based<br />

hybridization detection system that does not<br />

require pretreatment steps such as sample enrichment<br />

or signal amplification. The hsa-miR-21 (miR-21) is<br />

chosen as target miRNA for its significant elevated content<br />

in a variety of cancers as reported previously. Herein,<br />

probes of complementary single-stranded oligonucleotide<br />

were hybridized in solution to miR-21 and labeled with<br />

fluorescent dye YOYO-1. The fluorescent hybrids were<br />

imaged by an electron-multiplying charge-coupled device<br />

(EMCCD) coupled TIRFM system and quantified by<br />

single-molecule counting. This single molecule detection<br />

(SMD) assay shows a good correlation between the<br />

number of molecules detected and the factual concentration<br />

of miRNA. The detection assay is applied to quantify<br />

the miR-21 in extracted total RNA samples of cancerous<br />

MCF-7 cells, HepG2 cells, and normal HUVEC cells,<br />

respectively. The results agreed very well with those from<br />

the prevalent real-time polymerase chain reaction (qRT-<br />

PCR) analysis. This assay is of high potential for applications<br />

in miRNA expression profiling and early cancer<br />

diagnosis.<br />

MicroRNAs (miRNAs) are categorized as a class of small<br />

noncoding RNAs with approximately 19-23 nucleotides. It serves<br />

as the gene expression and cell development regulator in animals,<br />

plants, and viruses by interfering protein synthesis. 1,2 Evidences<br />

indicated that miRNAs play important roles in cell proliferation,<br />

differentiation, and apoptosis. 2-4 Recent researches have also<br />

explored the differential expression levels of miRNAs in cancerous<br />

* Corresponding author: (phone) +852-3411-7065; (fax) +852-3411-7348;<br />

(e-mail) hwli@hkbu.edu.hk.<br />

† Department of <strong>Chemistry</strong>, Hong Kong Baptist University.<br />

‡ Department of Biology, Hong Kong Baptist University.<br />

(1) Ambros, V. Cell 2003, 113, 673–676.<br />

(2) Ambros, V. Nature 2004, 431, 350–355.<br />

(3) Carthew, R. W. Curr. Opin. Genet. Dev. 2006, 16, 203–208.<br />

(4) Nakahara, K.; Carthew, R. W. Curr. Opin. Cell Biol. 2004, 16, 127–133.<br />

and noncancerous tissues and displayed its roles as tumor<br />

suppressors or oncogenes. 5-7 Tumor formation caused by miR-<br />

NAs is notable since it has been proven that a single sequence of<br />

miRNA can regulate multiple gene targets. In the meanwhile, a<br />

single gene target can also be regulated by multiple miRNAs.<br />

Therefore, on account of the connections between miRNA and<br />

cancer development, profiling of miRNA expression levels is<br />

proposed as a vital tool for the preliminary diagnosis and prognosis<br />

of cancer. 8<br />

A majority of the miRNA detection methods are based on the<br />

approach of hybridization, in which target molecules of interest<br />

arecapturedorhybridizedwiththecomplementaryoligonucleotides. 8,9<br />

Hereafter, the frequency of hybridization events are presented in<br />

form of measurable signals for the quantification of miRNA.<br />

However, the detection of the small size and trace amount of<br />

miRNAs is always challenging. Northern blotting is the most<br />

prevalent and accredited method for the quantification of miR-<br />

NAs. 10 The technique allows multiplex detection, but it is laborious<br />

and sample intensive. Another convincing technique for the<br />

determination of miRNA is the real-time polymerase chain reaction<br />

(qRT-PCR). 11-13 With PCR, copies of specific miRNAs and their<br />

corresponding signals are amplified and intensified, respectively.<br />

Nonetheless, the short length of target miRNA and intricate design<br />

of primer limited the reliability of amplification and labeling. 8<br />

Other amplification strategies have also been proposed in recent<br />

years aiming at higher detection sensitivity, specificity, and lower<br />

consumption of starting materials. For instance, nanoparticles such<br />

as OsO2, 14 quantum dots, 15 and gold 15,16 are chemically labeled<br />

on detecting probes as signal transducers and amplifiers. Besides,<br />

(5) Calin, G. A.; Croce, C. M. Nat. Rev. Cancer 2006, 6, 857–866.<br />

(6) Esquela-Kerscher, A.; Slack, F. J. Nat. Rev. Cancer 2006, 6, 259–269.<br />

(7) Volinia, S.; Calin, G. A.; Liu, C. G.; Ambs, S.; Cimmino, A.; Petrocca, F.;<br />

Visone, R.; Iorio, M.; Roldo, C.; Ferracin, M.; Prueitt, R. L.; Yanaihara, N.;<br />

Lanza, G.; Scarpa, A.; Vecchione, A.; Negrini, M.; Harris, C. C.; Croce, C. M.<br />

Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 2257–2261.<br />

(8) Wark, A. W.; Lee, H. J.; Corn, R. M. Angew. Chem., Int. Ed. 2008, 47,<br />

644–652.<br />

(9) Cissell, K. A.; Deo, S. K. Anal. Bioanal. Chem. 2009, 394, 1109–1116.<br />

(10) Lagos-Quintana, M.; Rauhut, R.; Lendeckel, W.; Tuschl, T. Science 2001,<br />

294, 853–858.<br />

(11) Chen, C. F.; Ridzon, D. A.; Broomer, A. J.; Zhou, Z. H.; Lee, D. H.; Nguyen,<br />

J. T.; Barbisin, M.; Xu, N. L.; Mahuvakar, V. R.; Andersen, M. R.; Lao, K. Q.;<br />

Livak, K. J.; Guegler, K. J. Nucleic Acids Res. 2005, 33, 9.<br />

(12) Raymond, C. K.; Roberts, B. S.; Garrett-Engele, P.; Lim, L. P.; Johnson,<br />

J. M. RNA 2005, 11, 1737–1744.<br />

(13) Schmittgen, T. D.; Jiang, J. M.; Liu, Q.; Yang, L. Q. Nucleic Acids Res. 2004,<br />

32, e43.<br />

(14) Gao, Z. Q.; Yang, Z. C. Anal. Chem. 2006, 78, 1470–1477.<br />

10.1021/ac101133x © 2010 American <strong>Chemical</strong> Society 6911<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/23/2010


enzymatic amplification techniques such as RAKE assay, 17 bioluminescence-enzyme<br />

labeling, 18 and rolling-circle amplification<br />

(RCA) 19 are also reported for sensitivity enhancement. These<br />

reported detection methods bring great benefit and improvement<br />

to the sensitivity of miRNA profiling to certain extends. Nevertheless,<br />

sample pretreatment and the modifications may result in loss<br />

of samples throughout the multiple pretreatment steps such as<br />

sample enrichment, labeling, and purification. The detection<br />

sensitivity and reliability are hence hindered. Since cellular miRNA<br />

concentration can be as low as 1000 molecules per cell, 20<br />

deficiencies in sensitivity may result in unsuccessful quantification<br />

of low-abundance miRNAs and, thus, false diagnostic result.<br />

Therefore, an ultrasensitive miRNA-profiling assay without the<br />

need of pretreatment is demanded.<br />

Single-molecule detection (SMD) technique has been widely<br />

applied in the behavioral study of individual biomolecules. 21,22<br />

Scientific issues such as monitoring enzymatic kinetics, 23,24 DNA<br />

adsorption/desorption behavior, 25,26 verification of nucleic acid<br />

hybridization, 27 DNA mismatch discrimination, 28 protein and DNA<br />

conformation dynamics, 29,30 and DNA mapping 31,32 have successfully<br />

been accomplished with SMD. Besides, several groups have<br />

also demonstrated the competence of SMD in the quantitation of<br />

biomolecules such as proteins 33 and viral DNA 34,35 by singlemolecule<br />

counting. Recently, quantitation of miRNA in singlemolecule<br />

level was displayed by Neely and co-workers. 36 The<br />

novel miRNA detection assay is free of sample enrichment and<br />

(15) Liang, R. Q.; Li, W.; Li, Y.; Tan, C. Y.; Li, J. X.; Jin, Y. X.; Ruan, K. C. Nucleic<br />

Acids Res. 2005, 33.<br />

(16) Yang, W. J.; Li, X. B.; Li, Y. Y.; Zhao, L. F.; He, W. L.; Gao, Y. Q.; Wan,<br />

Y. J.; Xia, W.; Chen, T.; Zheng, H.; Li, M.; Xu, S. Q. Anal. Biochem. 2008,<br />

376, 183–188.<br />

(17) Nelson, P. T.; Baldwin, D. A.; Scearce, L. M.; Oberholtzer, J. C.; Tobias,<br />

J. W.; Mourelatos, Z. Nat. Methods 2004, 1, 155–161.<br />

(18) Cissell, K. A.; Rahimi, Y.; Shrestha, S.; Hunt, E. A.; Deo, S. K. Anal. Chem.<br />

2008, 80, 2319–2325.<br />

(19) Cheng, Y.; Zhang, X.; Li, Z.; Jiao, X.; Wang, Y.; Zhang, Y. Angew. Chem.,<br />

Int. Ed. 2009, 48, 3268–3272.<br />

(20) Lim, L. P.; Lau, N. C.; Weinstein, E. G.; Abdelhakim, A.; Yekta, S.; Rhoades,<br />

M. W.; Burge, C. B.; Bartel, D. P. Genes Dev. 2003, 17, 991–1008.<br />

(21) Joo, C.; Balci, H.; Ishitsuka, Y.; Buranachai, C.; Ha, T. Annu. Rev. Biochem.<br />

2008, 77, 51–76.<br />

(22) Weiss, S. Science 1999, 283, 1676–1683.<br />

(23) Li, H. W.; Yeung, E. S. Anal. Chem. 2005, 77, 4374–4377.<br />

(24) Li, J. W.; Yeung, E. S. Anal. Chem. 2008, 80, 8509–8513.<br />

(25) Kang, S. H.; Shortreed, M. R.; Yeung, E. S. Anal. Chem. 2001, 73, 1091–<br />

1099.<br />

(26) Li, H. W.; Park, H. Y.; Porter, M. D.; Yeung, E. S. Anal. Chem. 2005, 77,<br />

3256–3260.<br />

(27) Kang, S. H.; Kim, Y. J.; Yeung, E. S. Anal. Bioanal. Chem. 2007, 387,<br />

2663–2671.<br />

(28) Gunnarsson, A.; Jonsson, P.; Marie, R.; Tegenfeldt, J. O.; Hook, F. Nano<br />

Lett. 2008, 8, 183–188.<br />

(29) Cohen, A. E.; Moerner, W. E. Proc. Natl. Acad. Sci. U.S.A. 2007, 104,<br />

12622–12627.<br />

(30) Funatsu, T.; Harada, Y.; Tokunaga, M.; Saito, K.; Yanagida, T. Nature 1995,<br />

374, 555–559.<br />

(31) Chan, E. Y.; Goncalves, N. M.; Haeusler, R. A.; Hatch, A. J.; Larson, J. W.;<br />

Maletta, A. M.; Yantz, G. R.; Carstea, E. D.; Fuchs, M.; Wong, G. G.; Gullans,<br />

S. R.; Gilmanshin, R. Genome Res. 2004, 14, 1137–1146.<br />

(32) Xiao, M.; Phong, A.; Ha, C.; Chan, T. F.; Cai, D. M.; Leung, L.; Wan, E.;<br />

Kistler, A. L.; DeRisi, J. L.; Selvin, P. R.; Kwok, P. Y. Nucleic Acids Res.<br />

2007, 35, e16.<br />

(33) Tessler, L. A.; Reifenberger, J. G.; Mitra, R. D. Anal. Chem. 2009, 81, 7141–<br />

7148.<br />

(34) Lee, J. Y.; Li, J. W.; Yeung, E. S. Anal. Chem. 2007, 79, 8083–8089.<br />

(35) Li, J. W.; Lee, J. Y.; Yeung, E. S. Anal. Chem. 2006, 78, 6490–6496.<br />

(36) Neely, L. A.; Patel, S.; Garver, J.; Gallo, M.; Hackett, M.; McLaughlin, S.;<br />

Nadel, M.; Harris, J.; Gullans, S.; Rooke, J. Nat. Methods 2006, 3, 41–46.<br />

6912 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

amplification, but continuous sample flow is needed for improvement<br />

in sensitivity. This microfluidic-assisted fluorescence correlation<br />

spectroscopy platform is also comparatively sophisticated.<br />

In this article, we present a quantitative single-molecule<br />

detection of miRNAs using total internal reflection fluorescence<br />

microscopy (TIRFM). For the proof of concept, we have chosen<br />

hsa-miR-21 (miR-21) as our detection target. The miR-21 is known<br />

as one of the most significant miRNAs elevated in at least six types<br />

of cancers including breast, colon, lung, pancreas, prostate, and<br />

stomach cancers. 7 Studies have indicated the miR-21 mediated<br />

tumor growth by serving as an oncogene 37 and targeting the<br />

tumor suppressor genes such as TPM1 and PDCD 4 in invasion<br />

and metastasis. 38-40 Compared to the conventional methods, the<br />

developed assay here is straightforward because no pretreatment<br />

steps are involved. Both the probe and target oligonucleotides<br />

are free of chemical modifications. The YOYO-1 labeled miRNA<br />

hybrids diffuse freely on unmodified coverslips and are monitored<br />

by electron-multiplying charge-coupled device (EMCCD) under<br />

TIRFM. TIRFM is a highly sensitive microscopic technique that<br />

has been used for SMD in solution. The total internal reflection<br />

(TIR) generates evanescent field layer that has a penetration depth<br />

of about 100-300 nm depending on the incident angle of the<br />

excitation laser beam. The excitation of fluorophores is confined<br />

within the evanescent field layer such that background signal from<br />

the bulk is greatly suppressed. Herein, the diffusing hybrids are<br />

observed as single fluorescent spots when they enter the excitation<br />

volume and are excited. Image of fluorescent molecules are<br />

acquired for single-molecule counting. The counted number is<br />

found to be proportional to the quantity of miRNAs in bulk<br />

solution. The developed assay was also employed for the determination<br />

of miR-21 in normal and cancerous cell lines and the<br />

results were validated with that of qRT-PCR detection.<br />

EXPERIMENTAL SECTION<br />

Slide Pretreatment. All coverslips were prewashed prior to<br />

experiments. Briefly, No. 1 22-mm square cover glasses (Gold<br />

Seal, Electron Microscopy System, Hatfield, PA) were sequentially<br />

sonicated for 30 min in household detergent, 30 min in acetone<br />

(AR grade, Labscan), and 30 min in absolute ethanol. The slides<br />

were then successively soaked for 30 min in Piranha solution<br />

(H2SO4/30% H2O2) (v/v 1:1), rinsed with distilled water<br />

extensively, sonicated for 30 min in HCl/30% H2O2/H2O (v/<br />

v/v 1:1:1) solution, sonicated in distilled water for 15 min,<br />

further sonicated for 30 min in Piranha solution, and finally<br />

sonicated for 15 min in distilled water twice. The slides were<br />

stored in distilled water and blow-dried with nitrogen before<br />

use.<br />

Preparation of Hybridization Buffers. A1× Tris-NaCl-EDTA<br />

(TNE) buffer containing 20 mM pH 8.0 Tris-HCl (Invitrogen,<br />

Carlsbad, CA), 1 mM EDTA, and various concentration (0, 50,<br />

150, 250, and 500 mM) of sodium chloride was prepared with<br />

DEPC-treated water (Ambion, Austin, TX) accordingly as the<br />

(37) Si, M. L.; Zhu, S.; Wu, H.; Lu, Z.; Wu, F.; Mo, Y. Y. Oncogene 2007, 26,<br />

2799–2803.<br />

(38) Lu, Z.; Liu, M.; Stribinskis, V.; Klinge, C. M.; Ramos, K. S.; Colburn, N. H.;<br />

Li, Y. Oncogene 2008, 27, 4373–4379.<br />

(39) Zhu, S. M.; Si, M. L.; Wu, H. L.; Mo, Y. Y. J. Biol. Chem. 2007, 282, 14328–<br />

14336.<br />

(40) Zhu, S. M.; Wu, H. L.; Wu, F. T.; Nie, D. T.; Sheng, S. J.; Mo, Y. Y. Cell<br />

Res. 2008, 18, 350–359.


dilution and hybridization buffer. The pH of the TNE buffers was<br />

adjusted to 7.4 by addition of 1 M HCl dropwisely. The buffer<br />

solution was then filtered through a 0.22 µm nylon membrane<br />

filter, autoclaved, and photobleached with UV-C lamp overnight<br />

prior to use.<br />

Preparation of Probe and Target MicroRNA Oligonucleotides.<br />

Commercial available LNA-modified oligonucleotide probe<br />

(miRCURYLNAmicroRNADetectionProbe,Productno:38102-00)<br />

(3′-TCAACATCAGTCTGATAAGCTA-5′) specific to hsa-miR-21<br />

was purchased from Exiqon (Denmark). Two HPLC-purified<br />

synthetic DNA oligonucleotides having the complementary sequence<br />

to hsa-miR-21 (3′-TCAACATCAGTCTGATAAGCTA-5′)<br />

and the mature sequence of hsa-miR-214 (3′-ACAGCAGGCACA-<br />

GACAGGCAGU-5′) were custom-designed and obtained from<br />

Invitrogen (Hong Kong) as the DNA probe and the negative<br />

control of the experiments. Two Anti-miR miRNA inhibitors of<br />

miR-21 (AM17000, Product ID: AM10206, 3′-CAACACCAGUC-<br />

GAUGGGCUGU-5′), and anti-miR-21 (AM17000, Product ID:<br />

AM12979, 3′-UAGCUUAUCAGACUGAUGUUGA-5′), were purchased<br />

from Ambion, acting as the RNA probe and target miR-21<br />

strands, respectively. All oligonucleotides were suspended in 1<br />

µL of DEPC-treated water (Ambion) and further diluted into<br />

appropriate concentration with 1× TNE buffer. The melting<br />

temperatures (Tm) of the oligonucleotides were predicted using<br />

the Probe Tm Predictor accessible online (www.exiqon.com)<br />

based on the thermodynamic nearest neighbor model.<br />

Optimization of Hybridization Conditions. Ionic Strength.<br />

For the determination of optimum hybridization ionic strength,<br />

sodium chloride concentrations from 0 to 500 mM in the TNE<br />

buffer were used throughout the oligonucleotides dilution and<br />

hybridization mixture preparation.<br />

Selection of Probe. For the selection of optimal probe with the<br />

best hybridization affinity toward miRNA, probes of DNA, RNA<br />

and LNA were prepared and hybridized with target miR-21 as<br />

described below. In addition, two control experiments including<br />

probes only and negative control miR-214 were also performed.<br />

Hybridization Time. For the determination of optimum hybridization<br />

time, hybrids of same concentration and ionic strength<br />

(250 mM NaCl) were incubated for 15 min, 30 min, 1 and 3 h<br />

respectively.<br />

Hybridization and Labeling of MicroRNA. The hybridization<br />

and fluorescence labeling of miRNAs were performed according<br />

to the following procedures. LNA probe, DNA probe, RNA<br />

probe, miRNA target, and DNA negative control oligonucleotides<br />

were diluted with 1× TNE buffers (pH 7.4) to 300 pM in<br />

concentration, respectively. The hybridization cocktail contained<br />

15 µL of 300 pM probe strand, 15 µL of appropriate concentration<br />

of target miR-21 strand, and 14 µL of TNE buffer. The cocktail<br />

was incubated in form of free-diffusing solutions in dry bath<br />

(AccuBlock Digital Dry Bath D1100, Labnet, NJ) for 1 h. The<br />

hybridization temperature was set to be 20 °C below the T m of<br />

the probe, that is, 52, 47, and 36 °C for LNA, DNA, and RNA<br />

probe, respectively. After incubation, 1 µL of 100 nM YOYO-1<br />

Iodide (YOYO) (Invitrogen) was added subsequently to label<br />

the hybrids. YOYO labels the hybrids in a dye to base pair<br />

(dye/bp) ratio of 1:1. The mixture was kept for 5 min in order<br />

to achieve equilibrium and 10 µL of solution was pipetted to<br />

the precleaned coverslips for TIRF imaging.<br />

External Calibration of MicroRNA. A calibration curve was<br />

established to correlate number of detected single fluorescent<br />

spots and concentration of miRNAs. Synthetic miRNA with final<br />

concentration of 100, 75, 50, 25, 10, 5, and 1 pM target miR-21<br />

strand was hybridized with probe strand of final concentration of<br />

100 pM under optimal conditions as mentioned above.<br />

Cell Culture. Human umbilical vein endothelial cells (HU-<br />

VEC) were purchased from Lonza (Walkersville, MD). M199<br />

medium, heparin, gelatin, and endothelial cell growth supplement<br />

(ECGS) were purchased from Sigma. Fetal bovine serum (FBS)<br />

and penicillin-streptomycin (PS) were purchased from Invitrogen.<br />

Dulbecco’s modified Eagle medium (DMEM) and Roswell Park<br />

Memorial Institute medium (RPMI) were purchased from Gibco<br />

(Grand Island, NY). HUVEC were grown in M199 medium<br />

supplemented with 20% heat-inactivated FBS, 20 µg/mL ECGS,<br />

90 µg/mL heparin, 1% PS in 75 cm 2 culture flasks coated with<br />

0.1% gelatin. HUVEC from passage 2-6 were used in this study.<br />

In addition, human hepatocellular carcinoma (HepG2) was<br />

cultured in DMEM supplemented with 10% FBS and 1% PS,<br />

and invasive breast ductal carcinoma (MCF-7) was cultured in<br />

RPMI supplemented with 10% FBS and 0.5% PS. All cells were<br />

maintained in a humidified incubator at 37 °C with 5% CO2 and<br />

80% relative humidity.<br />

Total RNA Isolation. Total RNA of HUVEC, HepG2, and<br />

MCF-7 were extracted using TRIzol Reagent (Invitrogen) according<br />

to the manufacturer’s protocol. Briefly, the sample was lysed<br />

and homogenized with TRIzol Reagent. Phase separation was<br />

followed by addition of chloroform and centrifugation. RNA in the<br />

aqueous phase was then recovered by precipitation with isopropyl<br />

alcohol. The RNA pellet was washed with 75% ethanol and finally<br />

redissolved in RNase-free water. The RNA quantity was determined<br />

by measuring optical density at 260 nm using the NanoDrop<br />

ND-1000 Spectrophotometer (NanoDrop Technologies, Wilmington,<br />

DE), and the RNA quality was assessed by performing<br />

agarose gel electrophoresis.<br />

Quantification of miR-21 in Cells by Quantitative Real-<br />

Time PCR (qRT-PCR). Complementary DNA (cDNA) was<br />

generated from 10 ng of total RNA per 5 µL of gene specific<br />

reverse transcription (RT) reaction by using reagents from<br />

TaqMan MicroRNA Reverse Transcription Kit (Applied Biosystems,<br />

Foster City, CA) and mature hsa-miR-21-specific RT primer<br />

(5×) from TaqMan MicroRNA Assays (P/N: 4373090, Applied<br />

Biosystems). Each RT reaction contained 1 mM dNTPs, 50 U<br />

MultiScribe reverse transcriptase, 1× reverse transcription buffer,<br />

and 3.8 U RNase inhibitor as well as 1× specific RT primer. The<br />

reaction mixture was incubated in PTC-100 Programmable Thermal<br />

Controller (MJ Research, Inc., Waltham, MA) for 30 min at<br />

16 °C, 30 min at 40 °C, 5 min at 85 °C, and then held at 4 °C. The<br />

cDNA sample was then amplified by PCR using TaqMan 2×<br />

Universal PCR Master Mix (No AmpErase UNG) (Applied<br />

Biosystems) and TaqMan Assay (20×) from TaqMan MicroRNA<br />

Assays (P/N: 4373090, Applied Biosystems). Each PCR reaction<br />

included 0.67 µL of RT product, 5 µL of TaqMan 2× Universal<br />

PCR Master Mix, No AmpErase UNG, and 0.5 µLof20× TaqMan<br />

Assay (a mix preformulated miRNA-specific forward PCR primer,<br />

specific reverse PCR primer and miRNA-specific TaqMan MGB<br />

probe). The reaction mixture was then run at 95 °C for 10 min,<br />

followed by 40 cycles of 95 °C for 15 s and 60 °C for 1 min in<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6913


Figure 1. Schematic illustration of the hybridization-based TIRFM assay for the detection of single miRNA molecules in solution.<br />

Bio-Rad iCycler, Version 4.006 (Bio-Rad Laboratories, Inc., Hercules,<br />

CA). The PCR reactions were run along with no-template<br />

control and RT-minus control. Data were analyzed by iCycler iQ<br />

Optical System Software, Version 3.0a (Bio-Rad Laboratories), and<br />

the miRNA expression level was measured using threshold cycle<br />

(Ct) which is the cycle number at which the fluorescence<br />

generated within a reaction crosses the threshold. The Ct values<br />

were then converted to absolute amount using a standard curve<br />

of mature miR-21. Six independent experiments were performed,<br />

and each experiment was run in duplicate.<br />

Establishment of miR-21 Calibration by qRT-PCR. A<br />

mixture of synthetic RNA oligonucleotides from mirVana miRNA<br />

Reference Panel v9.1 (Applied Biosystems) was used to generate<br />

a standard curve for miR-21. The RNA oligonucleotides representing<br />

mature miR-21 ranging from 10 -6 to 1 fmol were reverse<br />

transcribed, also using TaqMan MicroRNA Reverse Transcription<br />

Kit (Applied Biosystems) and mature miR-21-specific<br />

TaqMan MicroRNA Assays (P/N: 4373090, Applied Biosystems).<br />

The PCR amplification of the cDNA was then performed<br />

using same materials as mentioned above. Three independent<br />

experiments were performed, and each experiment was run<br />

in duplicate. Calibration curve was shown in Supporting<br />

Information.<br />

Quantification of miR-21 in Cells by TIRFM. Total RNA<br />

of HUVEC, HepG2, and MCF-7 was diluted to 150 ng/µL with<br />

TNE buffer, respectively. Consequently, 7.5 µL of diluted total<br />

RNAs were spiked into a mixture standard solution of miR-21 of<br />

a final concentration of 0, 1, 2, 5, 10, and 15 pM and a LNA probe<br />

with a final concentration of 100 pM, and finally diluted with TNE<br />

buffer to a volume of 44 µL. The solution was incubated for 1hat<br />

52 °C and followed by addition of 1 µL of YOYO dye as mentioned<br />

previously. After equilibrium for 5 min, 10 µL (with 250 ng of total<br />

RNA) of solution was pipetted to coverslips for TIRF imaging and<br />

quantitation. The contents of miR-21 in the three cell lines were<br />

estimated by standard addition method and the value of miR-21<br />

quantified by the TIRFM platform was compared with the outcome<br />

of the qRT-PCR method.<br />

Imaging System. An inverted Olympus IX-71 microscope<br />

(Olympus, Tokyo, Japan) was equipped with a high-numerical<br />

aperture 60× oil-immersion objective (1.45 NA, PlanApo, Olympus)<br />

as shown in Figure S1A. The sample coverslip was located under<br />

the fused-silica Isosceles Brewster Prism (CVI Melles Griot,<br />

Carksvadm, CA) and above the 60× objectives with immersion<br />

oil (η ) 1.52, Nonfluorescence, Olympus) in between. A 488 nm<br />

cyan laser (50 mW, CMA1-01983, Newport, NJ) was used as the<br />

excitation source. The laser beam was first filtered with neutral<br />

6914 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

density filter (FSR-OD 60, Newport), focused by cylindrical lens<br />

(focal length, 150 mm; CVI Melles Griot), and eliminated with<br />

the aid of pinholes before its entry to the prism with an incident<br />

angle of approximately 66°. Evanescent field was generated by<br />

the TIR of laser beam occurring at prism and was used to excite<br />

the fluorescently labeled hybrid molecules. A band-pass emission<br />

filter (D535/40 m Chroma Technology) was placed between the<br />

objective lens and the EMCCD to collect emitted photons. A<br />

uniphase mechanical shutter (model LS272, Vincent Associates,<br />

Rochester, NY) and a driver (model VMM-T1, Vincent Associates)<br />

were synchronized with the PhotonMax: 512B electron-multiplied<br />

CCD camera (EMCCD, Princeton Instruments, Princeton, NJ) in<br />

external synchronization mode and frame-transfer mode. The<br />

mechanical shutter blocked laser beam when the camera was off<br />

in order to reduce photobleaching. The ADC rate of the camera<br />

was 10 Hz, exposure time was 50 ms, and the multiplication gain<br />

was set at 4000, with the shutter driver set to 100 ms exposure<br />

and 100 ms delay. Typically, an image series of 10 sequential<br />

frames on 10 locations were acquired from a single slide. Images<br />

were obtained with the WinSpec/32 software (Version 2.5.22.0,<br />

Downingtown, PA) provided by Princeton Instruments.<br />

Data Analysis. All captured images were analyzed with a<br />

public-domain image-processing program Image J (version 1.43i,<br />

NIH, Bethesda, MD). A region of interest (ROI) with 200 pixelssquare<br />

at the center of the light spot with relatively even laser<br />

intensity was selected for single molecule counting. Intensity of<br />

images may also suffer while the shutter was triggered on or off.<br />

Five subframes (frame 3 to frame 7) of the image series were<br />

selected for analysis. The threshold for image acquirement was<br />

chosen at a value of three times the standard deviation of the mean<br />

intensity of the image. The image was then further processed with<br />

the Analyze Particles function in Image J to determine the number<br />

of single fluorescence particles computationally. The size of<br />

particles was set at 2-10 pixels to reduce false positive signals<br />

generated from noises. Number of spots in five frames was<br />

counted separately and summed up (i.e., accumulation of spots<br />

imaged within 250 ms). Consequently, the sum of spots from 10<br />

image series of a single slide was averaged. All experiments were<br />

done in triplicate and the error bars of charts shown in the Results<br />

and Discussion refer to the standard error of mean of the triplicate<br />

experiments unless specified.<br />

RESULTS AND DISCUSSION<br />

In this study, the miRNA detection is based on the hybridization<br />

approach as illustrated in Figure 1. Complementary probes<br />

of oligonucleotides were hybridized with the target miR-21 in


solution under appropriate incubation conditions. Compared with<br />

surface-based hybridization that involves washing of excess<br />

reagents, solution-based hybridization offers the advantage of<br />

higher hybridization efficiency. The hybridized duplexes were<br />

labeled with fluorescent dye YOYO-1 iodide (YOYO) for the singlemolecule<br />

fluorescence detection. YOYO is an intercalating fluorescent<br />

dye which electrostatically binds to the backbone of<br />

oligonucleotide. 41-43 The binding affinity of YOYO dyes on<br />

double-stranded oligonucleotide is very high (∼6 × 10 8 M -1 ) and<br />

there is a fluorescence intensity enhancement of approximately<br />

400-fold upon binding to DNA. 42 Since the binding mechanism<br />

is based on geometrical insertion, neither dye molecules nor<br />

the oligonucleotides have to be chemically modified during the<br />

labeling process. The detection of miRNA is, thus, straightforward.<br />

To improve the fluorescence signal intensity, the hybrids<br />

(∼20 bp) were labeled with YOYO in the ratio of 1 dye<br />

molecule/1 bp. After direct labeling of hybrids with YOYO,<br />

microliters of sample solution were sandwiched in a pair of<br />

precleaned coverslips with a solution depth of approximately<br />

20 µm, and observed under EMCCD-TIRF microscope. The<br />

cleanness of the coverslips was found to be very significant in<br />

the assay as scattering and autofluorescence of any dirt and<br />

stains on glass surface may result in false positive signals. It is<br />

crucial to clean coverslips extensively before use.<br />

YOYO-labeled miRNA hybrids were visualized by a singlemolecule<br />

TIRFM imaging system (Figure S1A). A 488 nm laser<br />

was used to excite the bound YOYO dyes. The laser beam with<br />

an incident angle of approximately 66° was total-internal-reflected<br />

at the glass/solution interface. The thickness of evanescent field<br />

layer (EFL) generated by total internal reflection was calculated<br />

to be ∼190 nm by d ) λ/(2π(η2 2 sin 2 θ - η1 2 ) 1/2 ), where d is the<br />

penetration depth of the field, λ is the wavelength of the<br />

excitation light in vacuum, η1 and η2 are the refraction indices<br />

of the solution and glass slides, and θ is the angle of incidence.<br />

For more homogeneous exciting laser intensity, a central<br />

region of 200 × 200 pixels (53 × 53 µm 2 ) of the EMCCD image<br />

was selected as the sampling area, and thus, the probe volume<br />

is estimated to be 0.54 pL. When 100 pM of miRNA hybrids<br />

was loaded on the coverslip, the theoretical number of observed<br />

hybrid molecules existing in the sampling region was 100 pM<br />

× (53 µm × 53 µm) × 190 nm × 6.02 × 10 23 molecules/mol )<br />

33. 44 Figure S1B shows a typical TIRFM image of miRNA hybrids<br />

acquired in single-molecule level. It is noted that molecules<br />

undergo random diffusional motion in a nonimmobilized system. 45<br />

The diffusion coefficient of the miRNA hybrids was calculated as<br />

76 µm 2 s -1 in bulk solution. 46 However, it was showed that<br />

molecular diffusion rate is much slower at the glass/solution<br />

interface compared to the bulk because of the electrostatic<br />

interaction between molecules and macroscopic glass surface<br />

in microsized domain. 45,47-49 The fluorescence signal generated<br />

(41) Cosa, G.; Focsaneanu, K. S.; McLean, J. R. N.; McNamee, J. P.; Scaiano,<br />

J. C. Photochem. Photobiol. 2001, 73, 585–599.<br />

(42) Gurrieri, S.; Wells, K. S.; Johnson, I. D.; Bustamante, C. Anal. Biochem.<br />

1997, 249, 44–53.<br />

(43) Rye, H. S.; Yue, S.; Wemmer, D. E.; Quesada, M. A.; Haugland, R. P.;<br />

Mathies, R. A.; Glazer, A. N. 1992, 20, 2803–2812.<br />

(44) He, Y.; Li, H. W.; Yeung, E. S. J. Phys. Chem. B 2005, 109, 8820–8832.<br />

(45) Xu, X. H.; Yeung, E. S. Science 1997, 275, 1106–1109.<br />

(46) Zhdanov, V. P. Mol. BioSyst. 2009, 5, 638–643.<br />

(47) Xu, X. H. N.; Yeung, E. S. Science 1998, 281, 1650–1653.<br />

Figure 2. Correlation between the number of observed miR-21<br />

molecules and expected number of miR-21 in the sampling volume<br />

(0.54 pL). The number of observed miR-21 was corrected as the net<br />

number of hybrids (number of observed molecules - number of<br />

observed in blank). The slope of the correlation curve is 0.80, which<br />

indicates the assay has a hybridization efficiency of approximately<br />

80%.<br />

from the single molecules is spreaded as they diffuse and the size<br />

of the fluorescence spots in the image is larger than the physical<br />

size of molecules of interest (see also movie file in Supporting<br />

Information). 45 Although molecules interact with the glass surface,<br />

nonspecific adsorption of hybrid molecules on the surface of glass<br />

slide was insignificant as both the oligonucleotides and the glass<br />

surface are highly negative-charged at pH 7.4. Figure S2 shows<br />

the histogram of residence time for each miRNA hybrids in 8<br />

consecutive frames. Among the 114 molecules detected in the 8<br />

frames, ∼ 80% of the molecules appeared and then disappeared<br />

in a single frame; 11%, 4%, and 4% of the molecules stayed at the<br />

same position for 2, 3, and 4 frames, respectively, and less than<br />

1% retained for 7 frames and eventually desorbed from the surface.<br />

In the TIRFM image, each fluorescent spot was regarded as a<br />

single molecule as a linear correlation on the number of counted<br />

fluorescence spots and the expected number of miR-21 calculated<br />

from the corresponding concentration was established (R 2 )<br />

0.991). Herein, hybridization is the main factor attributed to<br />

the differences in observed and expected number of molecules<br />

and its efficiency was approximated to be 80% from the slope<br />

of the plot as shown in Figure 2.<br />

Optimization of Hybridization Conditions. The stringency<br />

of hybridization governs the detection sensitivity and selectivity.<br />

It is crucial to optimize the hybridization conditions before<br />

performing further detection. The effects of ionic strength,<br />

selection of probes, and incubation time on the hybridization<br />

efficiency were evaluated.<br />

First, the effect of buffer ionic strength on hybridization<br />

efficiency was studied (Figure 3A). In general, hybridization<br />

affinity is improved at higher ionic strength as the electrostatic<br />

repulsions between the negatively charged oligonucleotides can<br />

be effectively shielded in the presence of salt. However, high ionic<br />

strength may also result in aggregation of molecules. The<br />

aggregates will be misinterpreted as an individual in SMD, and<br />

(48) Lyon, W. A.; Nie, S. M. Anal. Chem. 1997, 69, 3400–3405.<br />

(49) Isailovic, S.; Li, H. W.; Yeung, E. S. J. Chromatogr., A 2007, 1150, 259–<br />

266.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6915


Figure 3. Optimization of miRNA hybridization conditions. (A) Effect<br />

of ionic strength adjusted by NaCl on hybridization efficiency (number<br />

of hybrid counts). (B) Performance of LNA, DNA, and RNA probe on<br />

hybridization with complementary miR-21 and negative control of miR-<br />

214. (C) Effect of incubation time on hybridization efficiency (number<br />

of hybrid counts). The data depict the averages of three experiments,<br />

and the error bars are the standard error of mean of the three trials.<br />

as a consequence, the number of counts drops. Here, we adjusted<br />

the ionic strength by varying the concentration of NaCl in 1× Tris-<br />

NaCl-EDTA (TNE) buffer. Figure 3A shows the molecule counts<br />

as a function of the concentration of NaCl in the hybridization<br />

buffers for probes of locked nucleic acid (LNA), DNA, and RNA,<br />

6916 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

respectively, with the same concentration of the target miR-21.<br />

The experimental condition was considered optimum when the<br />

highest number of molecule counts was obtained; indicating that<br />

the largest number of individual hybrids was formed. A gradual<br />

increase in the number of hybrids was observed from buffer<br />

containing 0-250 mM NaCl and a decrease after the salt<br />

concentration exceeds. We analyzed the pixel size of each<br />

fluorescent spots of LNA/miRNA hybrids prepared in TE buffer<br />

with 250 and 500 mM NaCl, respectively, as shown in Figure S3.<br />

It is obvious that higher percentages of molecules with larger pixel<br />

sizes (size >4 pixels) were found in the series of 500 mM NaCl<br />

compared to that of 250 mM. This implies that increase in ionic<br />

strength would induce aggregation among hybrids and cause<br />

underestimation on the count of molecules in single molecule<br />

level. Thus, it is concluded that the optimal NaCl concentration<br />

was 250, 150, and 150 mM with respect to probes of LNA, DNA,<br />

and RNA.<br />

Oligonucleotide probe is commonly used as a reporter in<br />

hybridization-based nucleic acid detection. Conventional miRNA<br />

detection assays use DNA oligonucleotides as the capturing probe,<br />

although the stability of DNA-miRNA duplex is not high.<br />

Recently, several groups demonstrated the potential of LNAs as<br />

an alternative to DNA probes. 8,9,50,51 LNA is a nucleic acid<br />

analogue known as a mimic of RNA that has high binding affinity<br />

to RNA molecules. The thermostability of the LNA-miRNA<br />

duplex is significantly higher than that of unmodified DNA<br />

probes. 52-55 Consequently, both the detection sensitivity and<br />

hybridization discrimination efficiency are enhanced due to the<br />

increase in binding affinity and stability of LNA-miRNA complex.<br />

To assess the performance of different nucleic acid analogues of<br />

probes in the detection of miRNA, the hybridization affinities of<br />

LNA, DNA, and RNA probes toward (i) complementary miR-21<br />

and (ii) negative control miR-214 (sequence showed in Experimental<br />

Section) in free solution were investigated. As shown in<br />

Figure 3B, LNA probe-based hybridization yields a 1.5-fold<br />

enhancement in the molecule counts compared with those of DNA<br />

and RNA probes, while the binding of LNA probes to negative<br />

control of miR-214 was maintained at low level. This indicates that<br />

LNA probe not only promotes higher capturing efficiency, but also<br />

displays a good mismatch discrimination capability.<br />

Incubation time also plays a role on the overall hybridization<br />

efficiency. Counts of hybrids obtained after incubation for 15, 30,<br />

60, and 180 min were shown in Figure 3C. It was observed that<br />

the number of counts increased and reached maximum in the<br />

first 60 min incubation and then reduced after 3hofincubation.<br />

The reduction in counts after prolonged incubation may be due<br />

to the structural conformation change of LNA strands as proven<br />

by MALDI-TOF mass spectrometry analysis (see Supporting<br />

(50) Castoldi, M.; Schmidt, S.; Benes, V.; Noerholm, M.; Kulozik, A. E.; Hentze,<br />

M. W.; Muckenthaler, M. U. RNA 2006, 12, 913–920.<br />

(51) Kloosterman, W. P.; Wienholds, E.; de Bruijn, E.; Kauppinen, S.; Plasterk,<br />

R. H. A. Nat. Methods 2006, 3, 27–29.<br />

(52) Bondensgaard, K.; Petersen, M.; Singh, S. K.; Rajwanshi, V. K.; Kumar, R.;<br />

Wengel, J.; Jacobsen, J. P. Chem.sEur. J. 2000, 6, 2687–2695.<br />

(53) Braasch, D. A.; Corey, D. R. Chem. Biol. 2001, 8, 1–7.<br />

(54) Nielsen, K. E.; Rasmussen, J.; Kumar, R.; Wengel, J.; Jacobsen, J. P.;<br />

Petersen, M. Bioconjugate Chem. 2004, 15, 449–457.<br />

(55) Petersen, M.; Nielsen, C. B.; Nielsen, K. E.; Jensen, G. A.; Bondensgaard,<br />

K.; Singh, S. K.; Rajwanshi, V. K.; Koshkin, A. A.; Dahl, B. M.; Wengel, J.;<br />

Jacobsen, J. P. J. Mol. Recognit. 2000, 13, 44–53.


Figure 4. Standardization curve for the quantification of miR-21.<br />

Different concentrations of miR-21 were hybridized with 100 pM LNA<br />

probe in solution at 52 °C for 1 h. The data depict the averages of<br />

three experiments, and the error bars are the standard error of mean<br />

of the three trials.<br />

Information, Figure S4). As a result, the hybridization time was<br />

set to 60 min for later experiments in this study.<br />

Quantification of Synthetic miR-21. Under the optimal<br />

hybridization conditions of miRNAs as discussed above, a calibration<br />

plot of molecules counts as a function of the target miR-21<br />

concentration was constructed. Synthetic miRNA of 0-100 pM<br />

was hybridized with 100 pM LNA probes in solution and labeled<br />

with YOYO. By single-molecule counting on the series of images<br />

acquired, a plot of the number of molecule counts as a function<br />

of the concentration of synthetic miR-21 was obtained with a<br />

coefficient of determination of 0.991 (Figure 4). The detection limit<br />

of the assay was estimated to be 5 pM (i.e., 50 amol in 10 µL of<br />

sample). The ultimate theoretic limit of single molecule detection<br />

is to detect a sole molecule in the bulk sample solution sandwiched<br />

between the glass slides. Ideally, the theoretic limit is calculated<br />

to be 170 zM, with a single target molecule in 10-µL sample<br />

solution. To achieve such an ultimate theoretic limit requires<br />

sufficiently long sampling time until the single target molecule<br />

enters the probe volume (0.54 pL) by chance and gets excited<br />

and detected. Otherwise, regardless of the sampling time, the<br />

concentration of sample solution should be ∼3 pM, so that a single<br />

molecule always locates in the probe volume. To improve the<br />

detection limit, several factors including the probe volume,<br />

viscosity of sample solution, sampling time, photostability, and<br />

Figure 5. Quantification of miR-21 contents in total RNA of (A) MCF-7, (B) HepG2, and (C) HUVEC cells by standard addition methods with<br />

TIRFM. Synthetic miR-21 was spiked into matrix of total RNA and LNA probe. The data depict the averages of three experiments, and the error<br />

bars are the standard error of mean of the three trials. (D) Comparison between the miR-21 contents in total RNA of HUVEC, HepG2, and<br />

MCF-7 cells determined by qRT-PCR assay and single-molecule TIRFM assay.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6917


quantum efficiency of fluorescence dyes may be optimized as<br />

reported by Nie and co-workers. 56<br />

Quantification of miR-21 in Normal and Cancerous Cell<br />

Lines. MCF-7 (Invasive breast ductal carcinoma cell line) and<br />

HepG2 (human hepatocarcinoma cell line) are regarded as a<br />

common breast and liver cancer cell line model, respectively, while<br />

HUVEC (human umbilical vein endothelial cells) is regarded as<br />

normal cell line. Herein, we employed the single-molecule detection<br />

assay to quantify differential-expressing miRNAs in normal<br />

and cancerous cell lines. Briefly, total RNA of each cell line was<br />

extracted prior to the direct miRNA detection. The YOYO dye<br />

has no selectivity toward oligonucleotides, mRNAs, tRNAs, and<br />

other small RNAs in such environment. To eliminate the signals<br />

drawn from the complex matrixes without supplementary pretreatments<br />

on the cell samples, standard addition method is<br />

adopted in the manner that synthetic target miR-21 is spiked to<br />

the mixture of total RNA sample and probes. The original<br />

concentration of miR-21 in the sample of total RNA of each cell<br />

lines can eventually be obtained by extrapolation of the calibration<br />

curve. Three independent standardization curves were prepared<br />

for the three cell lines and the amounts of miR-21 in each of the<br />

cell line were determined. All three calibration curves showed<br />

strong correlation between the numbers of counted molecules and<br />

the miRNA concentration with the coefficients of determination<br />

for all three of them greater than 0.997 (Figure 5A-C). Since the<br />

hybridization efficiency was ∼80% as mentioned previously, the<br />

concentration of miR-21 in each cell line estimated by standard<br />

addition method was multiplied by the efficiency conversion factor<br />

of 1.25 such that the actual miR-21 contents in total RNA can be<br />

obtained. The contents of miR-21 in total RNA of each cell line<br />

were found to be 0.92, 0.43, and 0.32 amol/ng for MCF-7, HepG2,<br />

and HUVEC, respectively.<br />

For the purpose of result validation, we quantified the content<br />

of miR-21 in the three cell lines using the same batch of cells by<br />

qRT-PCR method. Quantitative RT-PCR is a technique commonly<br />

adopted as a standard method for miRNA profiling. It is superior<br />

to other detection assays for its high specificity and minute<br />

amounts of starting materials used in the detection. Amplification<br />

steps however are involved and it takes approximately 5 h for the<br />

whole process. The output of qRT-PCR is usually expressed in<br />

terms of fold-change and so the raw data is semiquantitative.<br />

Additional calibration curve has to be established for quantitation<br />

purpose. To compare, our SMD detection assay is relatively rapid<br />

as it takes only 1hofsample incubation and promptly followed<br />

by microscopic detection. The result is quantitative and obtained<br />

by applying standard addition methods. The standardization curve<br />

(56) Nie, S. M.; Chiu, D. T.; Zare, R. N. Anal. Chem. 1995, 67, 2849–2857.<br />

6918 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

by qRT-PCR is shown in the Supporting Information (Figure S5).<br />

Figure 5D displays the contents of miR-21 in MCF-7, HepG2, and<br />

HUVEC cells determined by the SMD assays and those by qRT-<br />

PCR with calibration standardization. The SMD result agrees very<br />

well with the outcome of qRT-PCR. The high correlation with the<br />

accredited qRT-PCR methods demonstrated that the pretreatmentfree<br />

SMD system developed here is of high potential in profiling<br />

expression of miRNAs in different cell lines and thus applicable<br />

in early cancer diagnosis.<br />

CONCLUSIONS<br />

We developed a direct and amplification-free quantitative assay<br />

of single miRNA molecules using solution-based hybridization<br />

approach and fluorescence-based detection with TIRFM. This<br />

assay is straightforward, rapid, and highly sensitive. The fluorescent<br />

hybrids diffuse randomly in the refined detection volume.<br />

Improvement on the limit of detection could be achieved by<br />

increasing sampling time or immobilizing target fluorescent<br />

hybrids onto the coverslips. As a proof of concept, the content of<br />

miR-21 were determined in cancerous MCF-7 and HepG2 and<br />

noncancerous HUVEC cell lines and the result agreed very well<br />

with that of conventional qRT-PCR. Both the success in discriminating<br />

differentially expressing miR-21 in different cell lines and<br />

the high correlation with the conventional detection method<br />

justified the potential of the system in application for future early<br />

cancer diagnosis.<br />

ACKNOWLEDGMENT<br />

This work was fully supported by the Faculty Research Grant<br />

of Hong Kong Baptist University (FRG/07-08/II-68) and grant<br />

from the University Grants Council of the Hong Kong Special<br />

Administrative Region, China (HKBU I/06C). We thank Dr.<br />

C. K. C. Wong from the Department of Biology of HKBU for<br />

providing the MCF-7 cells.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information includes video of miRNA diffusing in<br />

probe volume, adsorption time analysis of single hybrid molecules,<br />

pixel size analysis of hybrids prepared in TE buffer containing<br />

250 and 500 mM NaCl, MALDI-TOF mass spectrum of LNA<br />

strands, calibration curve of qRT-PCR. This material is available<br />

free of charge via the Internet at http://pubs.acs.org.<br />

Received for review April 30, 2010. Accepted July 14,<br />

2010.<br />

AC101133X


Anal. Chem. 2010, 82, 6919–6925<br />

Electrochemical Modulation for Signal<br />

Discrimination in Surface Enhanced Raman<br />

Scattering (SERS)<br />

Emiliano Cortés,* ,† Pablo G. Etchegoin,* ,‡ Eric C. Le Ru, ‡ Alejandro Fainstein, § María E. Vela, †<br />

and Roberto C. Salvarezza †<br />

Instituto de Investigaciones Fisicoquímicas Teóricas y Aplicadas (INIFTA), Universidad Nacional de La<br />

Plata-CONICET, Sucursal 4 Casilla de Correo 16 (1900), La Plata, Argentina, The MacDiarmid Institute for Advanced<br />

Materials and Nanotechnology, School of <strong>Chemical</strong> and Physical Sciences, Victoria University of Wellington,<br />

PO Box 600, Wellington, New Zealand, Centro Atómico Bariloche and Instituto Balseiro, Comisión Nacional de<br />

Energía Atómica and Universidad Nacional de Cuyo, (8400) San Carlos de Bariloche, Río Negro, Argentina<br />

Electrochemical modulation to induce controlled fluctuations<br />

in SERS signals is introduced as a method to<br />

discriminate and isolate different contributions to the<br />

spectra. The modulationswhich can be changed in potential<br />

range, amplitude, and frequencysacts as a controllable<br />

“switch” to turn on, off, or change specific Raman<br />

signals which can then be correlated within the spectra<br />

by different fluctuation analysis techniques. Principal<br />

component analysis (PCA), either by itself or assisted by<br />

fast fourier transform (FFT) prefiltering, are shown to<br />

provide viable tools to isolate the different components<br />

of the spectra. Electrochemical modulation provides,<br />

therefore, a technique to study complex cases of coadsorption,<br />

and resolve problems of spectral congestion in<br />

SERS signals.<br />

The links between surface-enhanced raman scattering (SERS) 1,2<br />

and electrochemistry go all the way back to the discovery of the<br />

effect in the 1970s. 3-5 Since then, electrochemical studies have<br />

provided invaluable information on the details of the electronic<br />

interaction of molecules with the underlying metal substrate<br />

responsible for the SERS enhancement. Particularly important<br />

have been studies where the origin of the so-called “chemical<br />

enhancement” 6,7 in SERS can be discerned. From a more modern<br />

point of view, SERS and electrochemistry have diversified into a<br />

myriad of different areas that include the studies of coadsorption<br />

* To whom correspondence should be addressed. E-mail: emilianocll@gmail.com<br />

(E.C.), pablo.etchegoin@vuw.ac.nz (P.G.E.).<br />

† Universidad Nacional de La Plata-CONICET.<br />

‡ Victoria University of Wellington.<br />

§ Comisión Nacional de Energía Atómica and Universidad Nacional de Cuyo.<br />

(1) Le Ru, E. C.; Etchegoin, P. G. Principles of Surface Enhanced Raman<br />

Spectroscopy and Related Plasmonic Effects; Elsevier: Amsterdam, 2009.<br />

(2) Aroca, R. F. Surface-Enhanced Vibrational Spectroscopy; John Wiley & Sons:<br />

Chichester, 2006.<br />

(3) Fleischmann, M.; Hendra, P. J.; McQuillan, A. J. Chem. Phys. Lett. 1974,<br />

26, 163–166.<br />

(4) Jeanmaire, D. L.; Van Duyne, R. P. J. Electroanal. Chem. 1977, 84, 1–20.<br />

(5) Albrecht, M. G.; Creighton, J. A. J. Am. Chem. Soc. 1977, 99, 5215–5217.<br />

(6) Otto, A. Surface-Enhanced Raman Scattering: Classical and <strong>Chemical</strong> Origins;<br />

Springer-Verlag: Berlin, 1984.<br />

(7) Tian, Z. Q. Faraday Discuss. 2006, 132, 309.<br />

(multiple species), biological redox system, 8,9 corrosion, surface<br />

science, 10 fuel cells, electrocatalysis, electronic models for chargetransfer<br />

processes on surfaces, 11-16 and single-molecule or single<br />

hot-spot SERS, 17,18 to name only a few. 19,20 The simultaneous<br />

presence of the laser field with electric potentials and/or currents<br />

in the many different experimental configurations of substrates<br />

with simple, multiple, or tip-like electrodes, together with the<br />

variety of environmental variables (like pH or the exact nature of<br />

the electrolyte) produce the vast diversity of experimental situations<br />

found in the modern applications of SERS to electrochemistry.<br />

Tian and co-workers 20 provide a lucid summary of recent<br />

advances in SERS for electrochemical applications. It is argued<br />

in ref 20 in fact, that electrochemical systems are among the most<br />

complex one can study in SERS. The combination of electrochemical<br />

processes at interfaces with SERS make these systems<br />

particularly challenging, but undoubtedly very relevant to understand<br />

fundamental aspects in real analytical and bioanalytical<br />

applications. Last, but not least, there are several techniques under<br />

development in SERS that combine the basic elements of<br />

electrochemical experiments, even though they would not be<br />

strictly speaking considered as such. Examples of the latter are<br />

recent attempts to combine SERS and molecule transport proper-<br />

(8) Murgida, D. H.; Hildebrandt, P. Chem. Soc. Rev. 2008, 37, 937–945.<br />

(9) Hildebrandt, P.; Murgida, D. H. Bioelectrochemistry 2002, 55, 139.<br />

(10) Cai, W. B.; Ren, B.; Li, X. Q.; She, C. X.; Liu, F. M.; Cai, X. W.; Tian, Z. Q.<br />

Surf. Sci. 1998, 406, 9–22.<br />

(11) Gersten, J. I.; Birke, R. L.; Lombardi, J. R. Phys. Rev. Lett. 1979, 43, 147–<br />

150.<br />

(12) Lombardi, J. R.; Birke, R. L.; Lu, T.; Xu, J. J. Chem. Phys. 1986, 84, 4174–<br />

4180.<br />

(13) Xie, Y.; Wu, D. Y.; Liu, G. K.; Huang, Z. F.; Ren, B.; Yan, J. W.; Yang, Z. L.;<br />

Tian, Z. Q. J. Electroanal. Chem. 2003, 554, 417–425.<br />

(14) Lombardi, J. R.; Birke, R. L. Acc. Chem. Res. 2009, 42, 734–742.<br />

(15) Tognalli, N.; Fainstein, A.; Bonazzola, C.; Calvo, E. J. J. Chem. Phys. 2004,<br />

120, 1905.<br />

(16) Tognalli, N.; Scodeller, P.; Flexer, V.; Szamocki, R.; Ricci, A.; Tagliazucchi,<br />

M.; Calvo, E. J.; Fainstein, A. Phys. Chem. Chem. Phys. 2009, 11, 7412.<br />

(17) dos Santos, D. P.; Andrade, G. F. S.; Temperini, M. L. A.; Brolo, A. G. J.<br />

Phys. Chem. C 2009, 113, 17737–17744.<br />

(18) Shegai, T.; Vaskevich, A.; Rubinstein, I.; Haran, G. J. Am. Chem. Soc. 2009,<br />

131, 14390.<br />

(19) Tian, Z. Q.; Ren, B. Annu. Rev. Phys. Chem. 2004, 55, 197–229.<br />

(20) Wu, D. Y.; Li, J. F.; Ren, B.; Tian, Z. Q. Chem. Soc. Rev. 2008, 37, 1025–<br />

1041.<br />

10.1021/ac101152t © 2010 American <strong>Chemical</strong> Society 6919<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/26/2010


ties, 21 or the use of electrostatic fields to control the adsorption/<br />

desorption of molecules on SERS substrates according to their<br />

cationic/anionic nature in solution. 17,22<br />

It is perhaps due to the complexity of the systems under study<br />

that despite all the accumulated work over the last decades the<br />

combination of SERS and electrochemistry is still undergoing<br />

methodological and analytical advances as a technique. There is<br />

a genuine need to develop both new and more advanced methods<br />

and analysis tools to attack the ever increasing complexity of the<br />

problems that are being investigated (in particular in bioelectrochemical<br />

areas). An early example of the latter is the technique<br />

of potential averaged SERS developed by Tian and co-workers 23<br />

to address the problem of fast electrochemical processes and<br />

SERS monitoring of unstable species on the electrodes. Detection<br />

of intermediate species in electrochemical reactions by timeresolved<br />

SERS has also been tried by Lombardi et al. 24 A more<br />

recent example, on the other hand, is the introduction of local<br />

imaging of the electrochemical current by surface-plasmon<br />

resonances in ref 25. Many of these new techniques are not<br />

directly linked or applied to SERS, 26,27 even though they share<br />

the same basic elements; that is, the combination of electrochemistry<br />

with a detection technique based on surface plasmon<br />

resonances. 1<br />

This paper aims at a methodological development in the<br />

combination of SERS and electrochemistry. As such we shall show<br />

some basic examples of the technique we propose. To this end,<br />

we borrow ideas from well established techniques of optical<br />

modulation 28 by using the ability of electrochemistry to turn “on”<br />

and “off” Raman signals (by changing the resonance conditions<br />

of the reduced or oxidized states), as well as to introduce other<br />

(more subtle) spectral changes as a function of the applied<br />

potential. Variations in SERS signals with potential have been<br />

extensively studied in the literature. A common trait is the<br />

observation of changes in the overall intensity of the spectra (as<br />

the main consequence of the applied potential) when switching<br />

from oxidized to reduced species. As a result, changes in the<br />

intensity of the SERS spectra due to the oxidation state have been<br />

used as a detection parameter in nanobiosensing, 29 to check the<br />

permeability of phospholipid membranes 30 or thiol self-assembled<br />

monolayers, 31 and to evaluate integrity and redox behavior in<br />

proteins; 32 among others.<br />

SERS provides a tool to monitor electrochemical phenomena<br />

at concentration levels (∼nanomolars, and below) that cannot be<br />

followed either in the voltamperogram or with other techniques.<br />

(21) Ward, D. R.; Halas, N. J.; Ciszek, J. W.; Tour, J. M.; Wu, Y.; Nordlander,<br />

P.; Natelson, D. Nano Lett. 2008, 8, 919–924.<br />

(22) Lacharmoise, P. D.; Etchegoin, P. G.; Le Ru, E. C. ACS Nano 2009, 3,<br />

66–72.<br />

(23) Tian, Z. Q.; Li, W. H.; Mao, B. W.; Zou, S. Z.; Gao, J. S. Appl. Spectrosc.<br />

1996, 50, 1569.<br />

(24) Shi, C.; Zhang, W.; Birke, R. L.; Lombardi, J. R. J. Phys. Chem. 1990, 94,<br />

4766–4769.<br />

(25) Shan, X.; Patel, U.; Wang, S.; Iglesias, R.; Tao, N. Science 2010, 327, 1363–<br />

1366.<br />

(26) Huang, B.; Yu, F.; Zare, R. N. Anal. Chem. 2007, 79, 2979–2983.<br />

(27) Wang, S.; Huang, X.; Shan, X.; Foley, K. J.; Tao, N. Anal. Chem. 2010, 82,<br />

935–941.<br />

(28) Yu, P. Y.; Cardona, M. Fundamentals of Semiconductors: Physics and<br />

Materials Properties; Springer: Berlin, 2004.<br />

(29) Scodeller, P.; Flexer, V.; Szamocki, R.; Calvo, E. J.; Tognalli, N.; Troiani,<br />

H.; Fainstein, A. J. Am. Chem. Soc. 2008, 130, 12690–12697.<br />

(30) Daza Millone, M. A.; Vela, M. E.; Salvarezza, R. C.; Creczynski-Pasa, T. B.;<br />

Tognalli, N. G.; Fainstein, A. Chem. Phys. Chem. 2009, 10, 1927–1933.<br />

6920 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Its sensitivity can, in fact, go routinely all the way down to single<br />

molecule levels. 33 But spectral congestion is very often cited as a<br />

common problem for its use 34,35 as was recently reported (for<br />

example) to separate resonances of a target molecule from a<br />

surrounding lipid matrix. 36 In the technique developed here,<br />

electrochemistry provides an external “switch” wherefrom variations<br />

in specific SERS signals can be introduced at will (and with<br />

a given periodicity). Fluctuation analysis with methods like<br />

principal component analysis (PCA)seither by itself or assisted<br />

by Fourier “lock-in” prefiltering (vide infra)scan easily follow from<br />

here, thus providing a systematic method to isolate SERS signals<br />

from different species along the electrochemical cycle (in the<br />

background of many possible, sometimes undesirable, contributions).<br />

Electrochemical modulation and signal discrimination could<br />

become a very important tool in areas like bioelectrochemistry,<br />

to isolate the signals from different species in redox active sites,<br />

in samples where we cannot choose at will the purity or spectral<br />

characteristics of all the components. The situation of overlapping<br />

peakssor much larger signals from spurious moleculesscan<br />

prevent the isolation of the interesting spectra displaying an<br />

oxidation/reduction cycle. Therefore, the technique is aimed at<br />

applications in this latter case, and it applies particularly well to<br />

address cases of coadsorption; which is a classic case applicable<br />

to the electrochemistry of complex multicomponent systems (like<br />

biological systems).<br />

EXPERIMENTAL SECTION<br />

A three-electrode cell with a Ag/AgCl (1 M Cl - ) electrode<br />

and a high-area platinum foil as reference and counter electrodes,<br />

respectively, was used. The working electrode (see<br />

below for further details) is immersed in the electrolyte solution<br />

(phosphate buffer, pH 6) and this is placed inside an open<br />

electrochemical cell that allows focusing on the substrate with<br />

long working distance objectives (×10, ×20, ×50, or ×100)<br />

through a water/air interface (see the Supporting Information<br />

(SI) for further details). All the potentials reported here are<br />

referenced to the Ag/AgCl (1 M Cl - ) electrode. Cyclic voltammetry<br />

is performed with a potentiostat with digital data<br />

acquisition. The whole assembly is placed on top of a motorized<br />

x-y stage for microscopy (to allow exploration and mapping<br />

of the electrode) and is used on a BX41 Olympus microscope<br />

attached to a Jobin-Yvon LabRam spectrometer. All the experiments<br />

performed in this paper are done with the 633 nm line<br />

of a HeNe laser with 3 mW at the sample. For samples with<br />

“high” concentrations of dyes (for SERS standards) like<br />

∼20-40 nM, we are mainly interested in average signals over<br />

the substrate. Therefore, we typically use for these experiments<br />

(31) Tognalli, N. G.; Fainstein, A.; Vericat, C.; Vela, M. E.; Salvarezza, R. C. J.<br />

Phys. Chem. C 2008, 112, 3741–3746.<br />

(32) Kranich, A.; Naumann, H.; Molina-Heredia, F. P.; Moore, H. J.; Lee, T. R.;<br />

Lecomte, S.; de la Rosa, M. A.; Hildebrandt, P.; Murgida, D. H. Phys. Chem.<br />

Chem. Phys. 2009, 11, 7390–7397.<br />

(33) Etchegoin, P. G.; Van Duyne, R. P. Phys. Chem. Chem. Phys. 2008, 10,<br />

6079.<br />

(34) Casadio, F.; Leona, M.; Lombardi, J. R.; Van Duyne, R. P. Acc. Chem. Res.<br />

2010, 43, 782–791.<br />

(35) Golightly, R. S.; Doering, W. E.; Natan, M. J. ACS nano 2009, 3, 2859–<br />

2869.<br />

(36) Nguyen, T. T.; Rembert, K.; Conboy, J. C. J. Am. Chem. Soc. 2009, 131,<br />

1401–1403.


the ×10 objective, to have a large spot diameter ∼10 µm, and<br />

minimize photobleaching effects as much as possible.<br />

Rhodamine 6G (RH6G), nile blue (NB), and crystal violet (CV)<br />

were obtained from commercial sources (Aldrich) and mixed at<br />

the appropriate concentrations with borohydride-reduced Ag<br />

colloids 37 (to avoid a citrate capping layer) and with 20 mM KCl;<br />

to induce a slight destabilization and the formation of clusters. 38<br />

For each case reported here, the details about the specific<br />

concentrations being used are specified in the captions. RH6G,<br />

NB, and CV adsorb to the negatively-charged colloids in this case<br />

through electrostatic interactions. The colloidal solution is subsequently<br />

drop-casted and dried under a mild heat on a clean Ag<br />

foil. Once dried, the colloids stick to the Ag foil by van der Waals<br />

forces, and remain attached to it upon reimmersion in the<br />

phosphate buffer solution. Several regions with multiple dried<br />

clusters of colloids can be easily distinguished in the microscope<br />

image on the Ag surface after drying. Hence, the Ag foil with the<br />

colloids and the dyes is our working electrode, and also provides<br />

the ideal means whereupon SERS enhancements (to observe low<br />

dye concentrations) can coexists with the ability to perform<br />

electrochemistry on the attached molecules. We checked at much<br />

higher concentrations (∼1 mM) that, as far the voltamperogram<br />

is concerned, the electrochemistry of the dyes on the colloids is<br />

indistinguishable from the one where the dyes are directly<br />

attached to the Ag foil itself. Note also that, in all of the<br />

experiments performed in this work, the dyes were adsorbed<br />

beforehand on the colloids (i.e., the dyes are not dissolved in the<br />

electrolyte solution itself).<br />

ELECTROCHEMICAL MODULATION AND SIGNAL<br />

ISOLATION<br />

Mixture of NB and RH6G. To fix ideas, we develop the<br />

method through an example. Consider the case of a mixture of<br />

two classic SERS dyes: RH6G and NB in concentrations of 40 and<br />

20 nM, respectively. We perform cyclic voltammetric runs varying<br />

the scan rate (i.e., the period), typically in the range ∼20-100 s<br />

per cycle, while monitoring the SERS signal on the electrode<br />

simultaneously with a much smaller integration time to follow the<br />

dynamics. This gives us the possibility to choose the best scan<br />

rate that allows us to follow the dynamics of the system. According<br />

also to the choice of the potential window, the different coadsorbed<br />

species will be modulated differently. This adds then a second<br />

degree of freedom (besides the period of the modulation) to be<br />

chosen by the experimentalist: that is, the potential range of the<br />

modulation. In the case of Figure 1 we modulated the potential<br />

in the range between -50 and -550 mV at a fix scan rate of 50<br />

mV · sec -1 . NB will experience oxidation/reduction cycles<br />

between these two values, while RH6G will only start to be<br />

reduced at a potential of ∼-400 mV and lower. 17 Another<br />

important detail is that the electrochemical modulation does<br />

not cross at any point the potential of zero charge (pzc )-0.9<br />

V) of the electrode, 39 meaning that the dyes (which are<br />

positively charged in solution) remain all the time on a<br />

(37) Creighton, J. A.; Blatchford, C. G.; Albrecht, M. G. J. Chem. Soc., Faraday<br />

Trans. 2 1979, 75, 790–798.<br />

(38) Meyer, M.; Le Ru, E. C.; Etchegoin, P. G. J. Phys. Chem. B 2006, 110,<br />

6040.<br />

(39) Niaura, G.; Gaigalas, A. K.; Vilker, V. L. J. Phys. Chem. B 1997, 101, 9250–<br />

9262.<br />

Figure 1. (a) Average spectrum for a sample with 20 nM of NB and<br />

40 nM of RH6G which is modulated electrochemically in the potential<br />

range -50 to -550 mV at a scan rate of 50 mV · sec -1 . Both signals<br />

of NB and RH6G are clearly visible, and both dyes are modulated in<br />

different electrochemical potential ranges. The “fingerprint” region<br />

containing the 590 cm -1 mode of NB and the 610 cm -1 one of RH6G<br />

(dashed box in (a)) is analyzed by PCA, resulting in the two dominant<br />

eigenvectors shown in (b). The spectra in (b) contains a linear<br />

combination of the two independent spectra of NB and RH6G, which<br />

can be obtained by a suitable change of basis (described in full detail<br />

in ref 40). The new eigenvectors (obtained from a linear combination<br />

of the PCA ones) that describe the two physically independent<br />

components (NB and RH6G) are shown in (c). With these eigenvectors<br />

we can make a decomposition of the spectra to obtain the two<br />

coefficients that describe (for each event) the intensity as a linear<br />

combination of v1 and v2. These coefficients are shown in (d), with a<br />

blown-out region in (e) where the dephasing of the electrochemical<br />

modulation of both signals can be observed.<br />

positively charged electrode in the entire modulation range.<br />

None of the phenomena reported in this paper can be ascribed<br />

to adsorption/desorption produced by a change in polarity of<br />

the electrode across the pzc point, but rather to oxidation/<br />

reduction of the species. Note also that the scan rates used in<br />

our experiences are fast enough compared to those reported<br />

for “electrostatic guiding”. 17,22<br />

In order to optimize the presentation of the material here<br />

without dwelling too much into collateral issues, we present the<br />

main details of the basic electrochemistry of NB and RH6G in<br />

the SI, as well as the basic aspects of the analysis with principal<br />

component analysis (PCA), and PCA with FFT prefiltering. 41,42<br />

As explained in the SI, both analysis techniques (PCA and FFT)<br />

are quite widespread, and we shall therefore assume that the<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6921


eader has a basic understanding of them. Hereafter, we shall go<br />

directly to the presentation of the main results.<br />

Figure 1(a) shows the average signal over 400 spectra taken<br />

with integration time of 1 s, with an electrochemical modulation<br />

period of 20 s. The average has clear fingerprint peaks of both<br />

dyes; for example the 590 and 610 cm -1 modes of NB and RH6G,<br />

respectively. The intensity of these peaks is modulated differently<br />

in the potential window selected, that is, their redox<br />

potentials are different. This is a case nevertheless where both<br />

signals of the coadsorbed species are clearly visible, and show<br />

in some cases some degree of overlap (like the 1645 and 1650<br />

cm -1 modes of NB and RH6G, respectively). We first show an<br />

analysis in Figure 1 based on a spectral region that contains<br />

fingerprint modes of both dyes, like the region around ∼600 cm -1<br />

shown in a dashed box in Figure 1(a). A PCA analysis of this<br />

region 40 immediately identifies the presence of two main modulated<br />

components in the spectra. As explained in the SI though,<br />

being a linear decomposition technique, PCA always provides a<br />

linear combination of the right answer in the form of principal<br />

“eigenvector” spectra (in this case the first two of them, shown<br />

in Figure 1(b)). In order to go from this two PCA eigenvectors to<br />

the real spectra that represent the individual contributions of NB<br />

and RH6G, a linear transformation (which we shall call “rotation”)<br />

of these eigenvectors is required. An entirely equivalent situation<br />

occurs with the application of PCA to the analysis of fluctuations<br />

from single molecule spectra; as explained in full length in ref<br />

40. The “rotated” eigenvectors obtained in this case are shown in<br />

Figure 1(c), and represent the isolated contributions of NB and<br />

RH6G to each individual spectrum. Once these two eigenvector<br />

spectra (v 1 and v2 in Figure 1(c)) are known, a standard linear<br />

decomposition can be performed on each spectra Ii in the time<br />

series (i ) 1, 2, ..., 400), to represent it as a linear combination<br />

(with two coefficients) of the rotated PCA eigenvectors; that<br />

is, Ii ) c1v1 + c2v2. The values of the coefficients, therefore,<br />

represent the particular contribution of NB and RH6G to an<br />

individual spectrum, and this is shown explicitly in Figure 1(d).<br />

Figure 1(e), in addition, shows a blown-out region of the time<br />

evolution of the coefficients where it can be appreciated that the<br />

two signals are being modulated differently at different times,<br />

accounting for the different intrinsic electrochemical properties<br />

of both dyes as a function of the sweeping potential. Note that<br />

time evolution can be converted to potential, since we know<br />

exactly the period of electrochemical modulation applied externally.<br />

An example of this is shown in the SI. The coefficients in<br />

Figure 1(d) and (e) are obviously in arbitrary units, since the<br />

Raman intensity itself is in arbitrary units too. What matters is<br />

not the absolute units, but rather the relative values of the<br />

coefficients representing the particular contributions to the total<br />

coming from the different electrochemical species.<br />

Therefore, Figure 1 demonstrates clearly that the combination<br />

of electrochemical modulation with PCA analysis can (i) resolve<br />

the problem of spectral congestion, and (ii) identify in addition<br />

the ranges of electrochemical potentials where the different<br />

species experience their modulation. This is emphasized in Figure<br />

(40) Etchegoin, P. G.; Meyer, M.; Blackie, E.; Le Ru, E. C. Anal. Chem. 2007,<br />

79, 8411.<br />

(41) Jolliffe, I. T. Principal Component Analysis; Springer Verlag:Berlin, 2002.<br />

(42) Hyvrinen, A.; Karhunen, J.; Oja, E. Independent Component Analysis; John<br />

Wiley & Sons: New York, 2001.<br />

6922 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

1(e), where the “dephasing” of the coefficients c 1 and c2 representing<br />

the relative contributions of NB and RH6G, respectively,<br />

can be easily seen. The fact that RH6G only starts to be<br />

reduced toward the lowest negative range of the potential (E<br />

< -400 mV) appears as a “dephasing” of the coefficients in<br />

time, while the successive electrochemical cycles evolve. Figure<br />

1(d) also shows a slight reduction in the amplitude of the<br />

coefficients produced by photobleaching at the observation point.<br />

It may also have partial contribution from natural long-term<br />

desorption of the dyes into the buffer. Photobleaching, in general,<br />

tends to be one of the major limiting factors to obtain unlimited<br />

sampling of the signal.<br />

Moving on from the simplest case in Figure 1, if we want to<br />

carry out the analysis over the entire spectral range (rather than<br />

a small window with fingerprint modes, like in Figure 1), special<br />

care must be taken with background contributions. Backgrounds<br />

are in general problematic in SERS 43 and, unlike the Raman peaks<br />

themselves, they can have a variety of origins (including molecules<br />

that are not in contact with the electrode). Furthermore, there<br />

will be in general a correlation between the background and the<br />

Raman signal created by the dispersion of the plasmon resonances<br />

producing SERS. 43-45 Here, we want to concentrate in a first<br />

approximation only on the Raman signals themselves and,<br />

therefore, the background will be removed 46 to avoid its presence<br />

in the PCA analysis. This should be done with care and the results<br />

should be assessed on a case-by-case basis. Backgrounds in SERS<br />

can be particularly problematic in the single molecule limit, where<br />

the dispersion of an individual plasmon resonance at a hot-spot<br />

can be revealed. 43,47 Measuring an average over many molecules<br />

is in a way an advantage, for differing types of backgrounds tend<br />

to average out and we are left with a case of background<br />

contribution that is easier to subtract. The background of each<br />

individual spectrum is subtracted with a wavelet transform that<br />

is fully described in ref 46 (including the program which is freely<br />

available 46 ). The success of the background subtraction procedure<br />

to deconvolute the spectra is judged here by the ability of the<br />

PCA analysis to distinguish the two main contributions to the<br />

signal in the full spectral range (which are known in this case).<br />

Effectively, Figure 2 shows the equivalent analysis of Figure<br />

1 in the full spectral range. Once the backgrounds are subtracted,<br />

and the two main PCA eigenvector spectra rotated, we recover<br />

the independent spectra of NB and RH6G from the mixture.<br />

Except for a few minor imperfections, the separation of the spectral<br />

components is almost perfect, and can be also distinguished in<br />

terms of the respective electrochemical modulation ranges for the<br />

potential. As in Figures 1(d) and (e), the dephasing of the<br />

coefficients from the contributions of NB and RH6G, according<br />

to the different values of the redox potentials over the cyclic<br />

voltammetry runs, can be clearly observed again in Figure 2(d).<br />

It is particularly worth emphasizing that the spectral deconvolution<br />

(43) Buchanan, S.; Le Ru, E. C.; Etchegoin, P. G. Phys. Chem. Chem. Phys. 2009,<br />

11, 7406.<br />

(44) Le Ru, E. C.; Etchegoin, P. G.; Grand, J.; Felidj, N.; Aubard, J.; Levi, G.;<br />

Hohenau, A.; Krenn, J. R. Curr. Appl. Phys. 2008, 8, 467.<br />

(45) Le Ru, E. C.; Grand, J.; Felidj, N.; Aubard, J.; Levi, G.; Hohenau, A.; Krenn,<br />

J. R.; Blackie, E.; Etchegoin, P. G. J. Phys. Chem. C 2008, 112, 8117.<br />

(46) Galloway, C.; Le Ru, E. C.; Etchegoin, P. G. Appl. Spectrosc. 2009, 63,<br />

1371.<br />

(47) Itoh, T.; Yoshida, K.; Biju, V.; Kikkawa, Y.; Ishikawa, M.; Ozaki, Y. Phys.<br />

Rev. B 2007, 76, 085405.


Figure 2. With a suitable removal of the background 46 for all events,<br />

the full spectra of NB and RH6G can be deconvoluted again from<br />

the modulated time series through PCA. The region analyzed now<br />

is the full spectral window shown in (a) (dashed line). The two main<br />

eigenvectors and the “rotated” ones 40 (representing the independent<br />

contributions of NB and RH6G) are shown in (b) and (c), respectively.<br />

The spectra of two compounds can be deconvoluted, and their time<br />

(or potential) dependence followed through the time evolution of the<br />

two coefficients in (d), that describe each event as a linear combination<br />

of v1 and v2 in (c). The dephasing of the electrochemical<br />

modulation of the two compounds can again be appreciated in (d).<br />

through the electrochemical modulation works very well even in<br />

regions where a substantial overlap of peaks occurs. An example<br />

of the latter is shown in Figure 3, for the region around ∼1650<br />

cm -1 , which has contributions from breathing modes of both<br />

NB and RH6G (but at slightly different frequencies). As can<br />

be appreciated from Figure 3, the different modulation induced<br />

through the electrochemical cycle on both compounds is good<br />

enough to “deconvolute” closely laying peaks within their natural<br />

widths.<br />

Mixture of NB and CV. All in all, the case presented in the<br />

previous section represents a relative “easy” case of spectral<br />

deconvolution for PCA with electrochemical modulation; starting<br />

from the fact that both compounds have comparable contributions<br />

to the intensity of the spectra and can be easily identified in terms<br />

of their contributions to fluctuations (which is the whole basis of<br />

PCA 41,42 ). A substantially more challenging case is to try recover<br />

a small signal from a background of other contributions to the<br />

spectra (that might be unavoidable in many real cases). We treat<br />

this case here also with an example.<br />

Figure 3. The deconvolution of the spectra works even in regions<br />

where there is a substantial overlap of peak intensities for the different<br />

compounds. Here we show explicitly the region ∼1650 cm -1 , where<br />

the RH6G peak is deconvoluted from the equivalent ring-breathing<br />

mode in NB, which is at a slightly lower frequency.<br />

Consider the case of a mixture of 80 and 20 nM of CV and<br />

NB, respectively, as shown in Figure 4(a). The mixture is prepared<br />

purposely so that the signal of NB is barely visible in the average<br />

spectrum (Figure 1(a)). This is shown explicitly, for example, with<br />

the arrows pointing at the ∼590 and ∼1650 cm -1 fingerprint<br />

peaks of NB in Figure 4(a), which are almost buried in the<br />

immediate surrounding of much larger peaks from CV. Both NB<br />

and CV experience (different) electrochemical modulations in the<br />

potential range between -50 and -550 mV (the electrochemistry<br />

of CV is also summarized in the SI). We can basically follow in a<br />

first approximation the analysis performed in the previous section.<br />

With an unlimited amount of sampling, PCA has all what it<br />

needs to distinguish the different components in the spectra. This<br />

is because the eigenvalues and eigenvectors of the covariance<br />

matrix (which is the basis of PCA 41,42 ) will eventually distinguish<br />

what is correlated and what is random in the signal. For the same<br />

reason that unlimited integration time should in principle always<br />

lift a signal above the noise level, sufficient sampling will always<br />

provide PCA with all the information required to distinguish the<br />

different contributions to the spectra. Nevertheless, when big<br />

signals are mixed with very small ones from other independent<br />

contributions, it is the comparative (with respect to other larger<br />

signals in the spectrum) size of the small signals that fix the<br />

amount of sampling we need to obtain the principal components<br />

from the spectra. This amount of sampling can be, however,<br />

prohibitively large. In the case of SERS, for example, photobleaching<br />

and other long-term stability problems of the signal puts<br />

serious limitation to how much sampling we can obtain. We are<br />

left here basically with a conundrum: random fluctuations in large<br />

signals with limited sampling might be comparablesas far as<br />

contributions to the covariance matrix of PCA is concernedsto<br />

the small signals we are trying to modulate and detect on purpose.<br />

An example of this is explicitly shown in Figure 4; the region<br />

encircled in a box in Figure 4(a) contains a contribution from NB<br />

and a much larger peak of CV in the same range. This region is<br />

PCA-analyzed and the first four eigenvectors are shown in Figure<br />

4(b). We can see that there is no clear signature of the NB peak<br />

in the PCA analysis. The first eigenvector is clearly dominated<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6923


Figure 4. With limited sampling, PCA can have problems to<br />

distinguish small components in the spectra. An example is shown<br />

here with a mixture of 80 and 20 nM of CV and NB, respectively. In<br />

(a) we show the average (over 400 spectra), with arrows showing<br />

the positions where the fingerprint modes of NB at 590 cm -1 and<br />

1645 cm -1 should appear. The signal is dominated by CV throughout<br />

the entire spectral range, with many overlap regions. In (b) we show<br />

a PCA analysis of the ∼1630 cm -1 region, where both the 1620 and<br />

1645 cm -1 peaks of CV and NB are expected, respectively. But with<br />

a difference of factor of ∼20 between the intensities of CV and NB in<br />

this range, fluctuations larger than ∼5% in the CV signal have a larger<br />

weight for the covariance matrix than the induced (electrochemical)<br />

variations of NB. With limited sampling, PCA cannot clearly distinguish<br />

the NB signal. We show here the first four PCA eigenvectors as an<br />

example. The limited sampling problem can be partially compensated<br />

by a prefiltering of the data with a FFT-transform, filtered at the<br />

appropriate electrochemical modulation frequency (see Figure 5).<br />

by the intensity fluctuations of the much larger CV peak, whereas<br />

the second one has the typical characteristic of an eigenvector<br />

accounting for frequency shifts of the CV peak. 40 From the third<br />

eigenvector onward, the covariance matrix in PCA is trying to<br />

find “correlations in the noise”. With enough sampling, these<br />

correlations will be zero and the peak of NB will appear as a third<br />

eigenvector. But in this case, the limited sampling establishes a<br />

competition between the random fluctuations of the much larger<br />

peak and the signal we are trying to observe (from NB).<br />

It is in situations like this one that a prefiltering of the data<br />

can help. The basic idea of Fourier prefiltering is explained in<br />

full in the SI with further examples, but here we shall give a brief<br />

version of it. From a FFT-transform of the total integrated intensity<br />

as a function of time we can easily reveal the main modulation<br />

frequency in the FFT-spectrum, as can be seen in Figure 5(a).<br />

This comes primarily from the modulation of the CV signal, but<br />

that is irrelevant at this stage: the FFT-transform of the total<br />

integrated intensity is used to identify the modulation in FFTfrequency<br />

units. 48-50 We can now perform a wavelength-bywavelength<br />

(i.e., pixel-by-pixel) FFT filtering of the data at this<br />

6924 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 5. (a) A FFT of the total integrated intensity of the spectra<br />

(which varies mainly here due to the electrochemical modulation of<br />

CV), reveals clearly the main frequency of the modulation (i.e., scan<br />

rate of 50 mV. sec -1 ). We can filter each pixel (wavelength) at this<br />

frequency and then transformed back the spectra into the time<br />

domain. The average of the filtered data (at the electrochemical<br />

modulation frequency) is shown in (b). Now the signal of NB is more<br />

clearly visible in the average and its relative intensity with respect to<br />

the CV peak is larger (because only the part of the CV signal that is<br />

modulated survives the filtering). A PCA analysis of the prefiltered<br />

signal shows a first eigenvector representing the CV intensity, and a<br />

second eigenvector which now contains the NB peak at ∼1645 cm -1 ,<br />

and a small fraction of frequency shifts of the CV peak. The vertical<br />

intensity scale for both curves in (c) is the same. Green lines are<br />

guides to the eye.<br />

particular frequency. This is simply done by an FFT-transform to<br />

the time evolution of each pixel, multiplication by an appropriate<br />

filter centered at the spikes of Figure 5(a), and followed by an<br />

inverse FFT-transform to go back to the time domain. In these<br />

data, therefore, we have enhanced the importance of only those<br />

fluctuations that happen at the known electrochemical modulation<br />

frequency, with respect to any other random fluctuation.<br />

Figure 5(b) and (c) show the resulting PCA analysis after FFTfiltering.<br />

It is worth noting that the average of the filtered signal<br />

has now a much better ratio of intensities between the CV and<br />

NB signals; this is because only the fraction of the CV signal that<br />

is being modulated survives the filtering. Accordingly, this puts<br />

(48) Press, W. H.; Flannery, B. P.; Teukolsky, S. A.; Vetterling, W. T. Numerical<br />

Recipes in FORTRAN: The Art of Scientific Computing; Cambridge University<br />

Press: Cambridge, 1989.<br />

(49) Kreyszig, E. Advanced Engineering Mathematics, 9th ed.; John Wiley & Sons:<br />

New York, 2006.<br />

(50) Kauppinen, J.; Partanen, J. Fourier Transforms in Spectroscopy; Wiley-VCH<br />

Verlag: Berlin, 2001.


the NB signal in a much better footing to be identified among<br />

the first PCA eigenvectors. Effectively, we can see now in Figure<br />

5(c) that the first eigenvector is still dominated by the CV signal,<br />

but the second one contains both the NB signal and some<br />

contributions to frequency shifts of the CV peak. What we have<br />

achieved by FFT-prefiltering is to put the NB signal at the same<br />

level of importance of frequency shifts of the CV peak (according<br />

to the covariance matrix) and this therefore “lifts” a very small<br />

signal from accidental correlations in the noise. As explained<br />

further in the SI, this is nothing but the well established principle<br />

of “lock-in” amplification; applied here as a prefiltering for PCA.<br />

In general, in any other experimental situation, if we have access<br />

to unlimited sampling then time lock-in “amplification” might not<br />

be needed; for the signal will always increase linearly with time<br />

while the noise increases like the square root of it. Therefore,<br />

the signal-to-noise ratio will always improve with integration time<br />

and eventually a small signal can be resolved. But we can achieve<br />

a much faster result in a shorter time (with less sampling) with<br />

a lock-in. Like in the case at hand here, unlimited amount of<br />

sampling is sometimes not possible in other experimental situations<br />

because the long-term stability of the experiment is compromised<br />

or it is simply too long. The situation depicted here is<br />

the exact analog of this latter situation but for the problem of<br />

spectral deconvolution with PCA.<br />

The combination of FFT-prefiltering and PCA might not be<br />

necessary in many situations and whenever possible, the simplest<br />

analysis should prevail. But the covariance matrix of PCA is<br />

basically “blind” to the time sequence of the electrochemical cycle.<br />

If we scrambled all the spectra (in time) we would still have the<br />

same covariance matrix and the same PCA eigenvectors. However,<br />

we do know the frequency (or period) with which we are inducing<br />

the electrochemical modulation. What we are doing with FFTprefiltering<br />

is, accordingly, to “pick” the appropriate fluctuations<br />

that happen at the frequency where we know the real physical<br />

effect is happening. In that manner, we bias the PCA analysis<br />

toward a physically relevant set of fluctuations, and this helps to<br />

improve the ability of PCA to distinguish what is correlated and<br />

what is not, in a situation where limited sampling is unavoidable.<br />

In the language of lock-in amplification, we have improved the<br />

signal-to-noise ratio by discarding all fluctuations except those that<br />

happen at the right frequency. The example chosen here is tailormade<br />

to show the concept and, as such, we have different options<br />

at hand. We could have done, for example, a PCA analysis is a<br />

much narrower window around the NB peak (to try to minimize<br />

the influence of the CV peak nearby). However, in real applications,<br />

we might not even know where to expect the peaks, and<br />

considerable spectral overlap is always possible. FFT-prefiltering<br />

is one additional tool to consider in these situation, and it could<br />

-in many cases- be the difference between being able to isolate a<br />

particular spectral feature of the weaker signal or not.<br />

CONCLUSIONS<br />

We have shown a combination of electrochemical modulation<br />

with fluctuation analysis to discriminate and isolate different<br />

species in cases where there is spectral congestion and coadsorption<br />

of molecules. Needless to say, the technique is not limited<br />

to two species, and can be applied to multicomponent systems in<br />

general. PCA contains, in principle, all the elements it needs to<br />

distinguish all spectral components given enough sampling. The<br />

restriction of limited sampling to “lift” small signals from other<br />

contributions has also been shown through the addition of FFTprefiltering.<br />

Prefiltering is not strange in PCA analysis. The most<br />

common type of prefiltering used in the PCA literature is<br />

“pre-whitening”. 41,42 In this latter case, this is done for a different<br />

purpose: to ensure that the samples are as unbiased as possible<br />

before the PCA analysis is carried out. Here we use prefiltering<br />

in a different way: to bias the study of correlations in the data<br />

that happen at a prescribed frequency, imposed externally by the<br />

electrochemical modulation. We believe that techniques to solve<br />

spectral congestion of coadsorbed species will be of great interest,<br />

as SERS moves into areas (like bioelectrochemistry) where the<br />

purity and characteristics of the sample cannot be chosen at will.<br />

We hope the concepts developed here in our paper will contribute<br />

to that endeavor.<br />

ACKNOWLEDGMENT<br />

E.C. acknowledges the financial support of UNLP, ANPCyT<br />

(Argentina) and the MacDiarmid Institute (New Zealand) for a<br />

research/exchange program between Argentina and New Zealand.<br />

Thanks are also given to the host institution, Victoria University<br />

of Wellington, where the work was carried out. We acknowledge<br />

financial support from ANPCyT (Argentina, PICT06-621, PAE<br />

22711, PICT06-01061, PICT-CNPQ 08-019). E.C., A.F., and RCS<br />

are also at CONICET. MEV is a member of the research career<br />

of CIC BsAs. R.C.S and A.F. are Guggenheim Foundation Fellows.<br />

PGE and ECLR are indebted to the Royal Society of New Zealand<br />

for additional financial support under a Marsden Grant.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information and figures. This material is available<br />

free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 3, 2010. Accepted July 10, 2010.<br />

AC101152T<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6925


Anal. Chem. 2010, 82, 6926–6932<br />

Study of Highly Selective and Efficient Thiol<br />

Derivatization Using Selenium Reagents by<br />

Mass Spectrometry<br />

Kehua Xu, †,‡ Yun Zhang, † Bo Tang,* ,‡ Julia Laskin, § Patrick J. Roach, § and Hao Chen* ,†<br />

Center for Intelligent <strong>Chemical</strong> Instrumentation, Department of <strong>Chemistry</strong> and Biochemistry, Ohio University, Athens,<br />

Ohio 45701, Key Laboratory of Molecular and Nano Probes, Ministry of Education, College of <strong>Chemistry</strong>, <strong>Chemical</strong><br />

Engineering and Materials Science, Shandong Normal University, Jinan, China, 250014, and <strong>Chemical</strong> and Materials<br />

Sciences Division, Pacific Northwest National Laboratory, P.O. Box 999, MSIN K8-88, Richland, Washington 99352<br />

This paper reports a systemic mass spectrometry (MS)<br />

investigation of a novel strategy for labeling biological<br />

thiols, involving the cleavage of the Se-N bond by thiol<br />

to form a new Se-S bond. Our data show that the reaction<br />

is highly selective, rapid, reversible, and efficient. Among<br />

20 amino acids, only cysteine is reactive toward Se-N<br />

containing reagents and the reaction occurs in seconds.<br />

With the addition of dithiothreitol, peptides derivatized<br />

by selenium reagents can be recovered. The high reaction<br />

selectivity and reversibility provide potential in both<br />

selective identification and isolation of thiols from mixtures.<br />

Also, with dependence on the selenium reagent<br />

used, derivatized peptide ions exhibit tunable dissociation<br />

behaviors (either facile cleavage or preservation of the<br />

formed Se-S bond upon collision-induced dissociation),<br />

a feature that is useful in proteomics studies. Equally<br />

importantly, the thiol derivatization yield is striking, as<br />

reflected by 100% conversion of protein �-lactoglobulin<br />

A using ebselen within 30 s. In addition, preliminary<br />

applications such as rapid screening of thiol peptides from<br />

mixtures and identification of the number of protein free<br />

and bound thiols have been demonstrated. The unique<br />

selenium chemistry uncovered in this study would be<br />

valuable in the MS analysis of thiols and disulfide bonds<br />

of proteins/peptides.<br />

Biological thiols, such as glutathione (GSH) and thiol proteins,<br />

are critical physiological components found in animal tissues and<br />

fluids and involved in a plethora of important cellular functions. 1,2<br />

In the past decades chemical characterization of biological thiols<br />

has attracted significant attention. Various measurement<br />

methods have been developed including UV, 3 fluorescence (FL)<br />

* To whom correspondence should be addressed. Hao Chen: phone, 740-<br />

593-0719; fax, 740-597-3157; e-mail, chenh2@ohio.edu. Bo Tang: phone, (86)531-<br />

86180010; e-mail, tangb@sdnu.edu.cn.<br />

† Ohio University.<br />

‡ Shandong Normal University.<br />

§ Pacific Northwest National Laboratory.<br />

(1) Basford, R. E.; Huennekens, F. M. J. Am. Chem. Soc. 1955, 77, 3873–<br />

3877.<br />

(2) Rahman, I.; MacNee, W. Free Radical Biol. Med. 2000, 28, 1405–1420.<br />

(3) Kusmierek, K.; Chwatko, G.; Glowacki, R.; Bald, E. J. Chromatogr., B 2009,<br />

877, 3300–3308.<br />

6926 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

spectroscopy, 4-6 and mass spectrometry (MS), 7-12 in which the<br />

derivatization of thiol groups with a suitable chemical reagent is<br />

necessary for increasing thiol stability and improving detection<br />

selectivity.<br />

While spectroscopic methods such as FL are very useful for<br />

detection and imaging of thiol compounds, MS can provide<br />

molecular weight and structural information. The inherent sensitivity<br />

and chemical specificity offered by MS 13-16 is essential in<br />

further identifying biological thiols and investigating their physiological<br />

functions on a molecular level. Also, a number of excellent<br />

MS studies of thiols and disulfides of proteins/peptides based on<br />

the novel ion chemistry have been reported. 10,11,17-20 At present,<br />

MS labeling strategies are mainly based on three types of chemical<br />

reactions. Nucleophilic substitution such as using heptafluorobutyl<br />

chloroformate 21 and iodoacetamide is commonly used for the<br />

analysis of protein tryptic digest; 22 however, these reagents are<br />

not specific for thiol compounds, that is, they can also couple with<br />

amino/hydroxyl groups. The second approach involves Michael-<br />

(4) Lee, J. H.; Lim, C. S.; Tian, Y. S.; Han, J. H.; Cho, B. R. J. Am. Chem. Soc.<br />

2010, 132, 1216–1217.<br />

(5) Pullela, P. K.; Chiku, T.; Carvan, M. J.; Sem, D. S. Anal. Biochem. 2006,<br />

352, 265–273.<br />

(6) Tang, B.; Yin, L.; Wang, X.; Chen, Z.; Tong, L.; Xu, K. Chem. Commun.<br />

2009, 5293–5295.<br />

(7) Gorman, J. J.; Wallis, T. P.; Pitt, J. J. Mass Spectrom. Rev. 2002, 21, 183–<br />

216.<br />

(8) Bilusich, D.; Bowie, J. H. Mass Spectrom. Rev. 2009, 28, 20–34.<br />

(9) Chrisman, P. A.; Pitteri, S. J.; Hogan, J. M.; McLuckey, S. A. J. Am. Soc.<br />

Mass Spectrom. 2005, 16, 1020–1030.<br />

(10) Diedrich, J. K.; Julian, R. R. J. Am. Chem. Soc. 2008, 130, 12212–12213.<br />

(11) Gunawardena, H. P.; O’Hair, R. A. J.; McLuckey, S. A. J. Proteome Res. 2006,<br />

5, 2087–2092.<br />

(12) Seiwert, B.; Karst, U. Anal. Chem. 2007, 79, 7131–7138.<br />

(13) Wang, Y.; Vivekananda, S.; Zhang, K. Anal. Chem. 2002, 74, 4505–4512.<br />

(14) Pingitore, F.; Wesdemiotis, C. Anal. Chem. 2005, 77, 1796–1806.<br />

(15) Blanksby, S. J.; Bierbaum, V. M.; Ellison, B. G.; Kato, S. Angew. Chem.,<br />

Int. Ed. 2007, 46, 4948–4950.<br />

(16) Gronert, S. Chem. Rev. 2001, 101, 329–360.<br />

(17) Kim, H. I.; Beauchamp, J. L. J. Am. Soc. Mass Spectrom. 2009, 20, 157–<br />

166.<br />

(18) Chrisman, P. A.; McLuckey, S. A. J. Proteome Res. 2002, 1, 549–557.<br />

(19) Li, J.; Shefcheck, K.; Callahan, J.; Fenselau, C. Int. J. Mass Spectrom. 2008,<br />

278, 109–133.<br />

(20) Qiao, L.; Bi, H.; Busnel, J.-M.; Liu, B.; Girault, H. H. Chem. Commun. 2008,<br />

47, 6357–6359.<br />

(21) Simek, P.; Husek, P.; Zahradnickova, H. Anal. Chem. 2008, 80, 5776–<br />

5782.<br />

(22) Williams, D. K., Jr.; Meadows, C. W.; Bori, I. D.; Hawkridge, A. M.; Comins,<br />

D. L.; Muddiman, D. C. J. Am. Chem. Soc. 2008, 130, 2122–2123.<br />

10.1021/ac1011602 © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/16/2010


Scheme 1. Reactions of Selenium Reagents 1 and 2 with Thiols of Proteins/Peptides<br />

addition of thiols onto unsaturated CdC bonds such as using<br />

acrylate 23 or maleimide derivatives as common labeling agents. 24<br />

These reagents have selectivity toward thiols and have been often<br />

used in practical thiol analysis. However, the Michael-addition<br />

product is irreversible, which cannot allow enriching and purifying<br />

analyte compounds from complex matrices. 25 The third approach<br />

involves thiol exchange reaction such as using 5,5′-dithiobis(2nitrobenzoic<br />

acid) known as Ellman’s reagent; 26,27 a large excess<br />

amount of the disulfide reagent is necessary to ensure complete<br />

derivatization of all thiols in the sample. As a result, the newly<br />

formed disulfides in the reaction can possibly further react with<br />

residual thiols of the target molecule to form undesirable disulfides.<br />

For example, Udgaonkar et al. used Ellman’s reagent in a<br />

100-fold molar excess to label the thiol protein based on the<br />

exchange reaction in the study of the cooperativity of a fast protein<br />

folding reaction by MS. 27 Besides the problems mentioned above,<br />

these derivatization reactions have other limitations, including a<br />

long reaction time, low conversion yield, or more than one possible<br />

site for tagging. Therefore, new derivatization chemistry suitable<br />

for MS detection of thiol, particularly with high thiol selectivity,<br />

fast reaction speed, good reaction reversibility, and high conversion<br />

yield, is still needed for the structural analysis of peptides<br />

and proteins in biological samples.<br />

Because selenium is an essential element in vivo 28 and a key<br />

component of selenoproteins known as antioxidant enzymes,<br />

selenium chemistry has recently attracted increasing attention.<br />

Ebselen, 2-phenyl-1,2-benzisoselenazol-3(2H)-one (reagent 1,<br />

Scheme 1), has been considered as a mimetic of glutathione<br />

peroxidase (GPx) 29-32 and an anti-inflammatory drug since being<br />

found. 33,34 The mechanism for ebselen-catalyzed thiol oxidation<br />

(23) Wang, G.; Hsieh, Y.; Wang, L.; Prelusky, D.; Korfmacher, W. A.; Morrison,<br />

R. Anal. Chim. Acta 2003, 492, 215–221.<br />

(24) Seiwert, B.; Hayen, H.; Karsta, U. J. Am. Soc. Mass Spectrom. 2008, 19,<br />

1–7.<br />

(25) Jalili, P. R.; Ball, H. L. J. Am. Soc. Mass Spectrom. 2008, 19, 741–750.<br />

(26) Sevcikova, P.; Glatz, Z.; Tomandl, J. J. Chromatogr., A 2003, 990, 197–<br />

204.<br />

(27) Kumar Jha, S.; Udgaonkar, J. B. J. Biol. Chem. 2007, 282, 37479–37491.<br />

(28) Yoneda, S. J.; Kazuo, T. S. Toxicol. Appl. Pharmacol. 1997, 143, 274–280.<br />

(29) Forstrom, J. W.; Zakowski, J. J.; Tappel, A. L. Biochemistry 1978, 17, 2639–<br />

2644.<br />

(30) Birringer, M.; Pilawa, S.; Flohé, L. Nat. Prod. Rep. 2002, 19, 693–718.<br />

(31) Jacob, C.; Giles, G. I.; Giles, N. M.; Sies, H. Angew. Chem, Int. Ed. 2003,<br />

42, 4742–4758.<br />

(32) Haenen, G. R.; De Rooij, B. M.; Vermeulen, N. P.; Bast, A. Am. Soc.<br />

Pharmacol. Exp. Therapeut. 1990, 37, 412–422.<br />

has been proposed: 35,36 the Se-N bond of ebselen is cleaved by<br />

thiol RSH to produce the corresponding selenenyl sulfide Se-S,<br />

which further reacts with excess thiol RSH to produce selenol<br />

and disulfide compound RSSR. However, Mugesh et al. recently<br />

reported that ebselen as a GPx-like redox catalyst is indeed<br />

inactive, and the reaction of ebselen with PhSH only produces<br />

the Se-S product, not disulfide PhSSPh, even with an excess<br />

amount of PhSH. 36 Very recently, based on this novel selenium<br />

chemistry, we synthesized fluorescent probes carrying Se-N<br />

bonds and used them for detecting and imaging thiols in living<br />

cells 6,37 by FL. It was shown that the FL probes capture cellular<br />

thiols selectively in the presence of diverse species such as<br />

inorganic metal ions, dopamine, histamine, L-adrenaline, etc.<br />

Following that, Zhang et al. used piazselenole containing Se-N<br />

bond to probe physiological thiols based on electrochemical<br />

reactions. 38 Highly specific derivatization of thiols using compounds<br />

containing Se-N bonds makes them promising candidates<br />

to derivatize peptides and proteins for subsequent MS characterization.<br />

Although the analysis of protein thiols and disulfide bonds<br />

is important in proteomics applications, the selenium chemistry<br />

received surprisingly limited attention. 39 Here we present a<br />

systematic study of derivatization of amino acids, peptides, and<br />

proteins using Se-N containing reagents and examine the utility<br />

of this chemistry for analytical applications involving mass<br />

spectrometry.<br />

In this study, we carried out a series of reactions of two Se-N<br />

containing reagents, ebselen and N-(phenylseleno)phthalimide<br />

(reagent 2, Scheme 1), with various biological thiol compounds<br />

such as cysteine, reduced peptide glutathione GSH, and �-lactoglobulin<br />

A protein. As shown in Scheme 1, the Se-N bonds in<br />

reagents 1 and 2 are cleaved by thiols to form new Se-S bonds<br />

as shown in products 3 and 4, respectively (Scheme 1). Our<br />

experimental results showed that this thiol derivatization reaction<br />

is highly selective, rapid, reversible, and efficient (quantitative in<br />

(33) Müller, A.; Cadenas, E.; Graf, P.; Sies, H. Biochem. Pharmacol. 1984, 33,<br />

3235–3239.<br />

(34) Sies, H.; Masumoto, H. Adv. Pharmacol. 1997, 38, 229–246.<br />

(35) Mugesh, G.; du Mont, W.-W.; Sies, H. Chem. Rev. 2001, 101, 2125–2179.<br />

(36) Sarma, B. K.; Mugesh, G. J. Am. Chem. Soc. 2005, 127, 11477–11485.<br />

(37) Tang, B.; Xing, Y.; Li, P.; Zhang, N.; Yu, F.; Yang, G. J. Am. Chem. Soc.<br />

2007, 129, 11666–11667.<br />

(38) Wang, W.; Li, L.; Liu, S.; Ma, C.; Zhang, S. J. Am. Chem. Soc. 2008, 130,<br />

10846–10847.<br />

(39) Sakurai, T.; Kanayama, M.; Shibata, T.; Itoh, K.; Kobayashi, A.; Yamamoto,<br />

M.; Uchida, K. Chem. Res. Toxicol. 2006, 19, 1196–1204.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6927


Figure 1. ESI-MS spectrum showing the selective reaction of ebselen with cysteine in the presence of methionine, tryptophan, and valine. The<br />

insets show the agreement of the simulated product peak isotopic distribution (top) with the recorded reaction product ion of m/z 397 (bottom).<br />

some cases) at room temperature. On the basis of these revealed<br />

features of the reaction, associated analytical applications were<br />

explored, such as the selective identification of thiol-containing<br />

peptides from mixtures and measurement of the number of<br />

cysteine residues of proteins. In addition, the dissociation behaviors<br />

of the derivatized protein/peptide ions were also examined.<br />

EXPERIMENTAL SECTION<br />

Electrospray ionization (ESI) of samples was performed using<br />

a Thermo Finnigan LCQ DECA Mass Spectrometer (San Jose,<br />

CA) or a hybrid triple-quadrupole-linear ion trap mass spectrometer<br />

(Q-trap 2000; Applied Biosystems/MDS SCIEX, Concord,<br />

Canada). The sample injection flow rate was 10 µL/min. A high<br />

voltage of +5 kV was applied to the spray probe for the positive<br />

ion mode. For the Thermo Finnigan LCQ DECA mass spectrometer,<br />

the optimized heated transfer capillary tube temperature was<br />

150 °C. Collision-induced dissociation (CID) was used for further<br />

structural confirmation of the product ions. Data acquisition was<br />

performed using Xcalibur (rev. 2.0.7, Thermo Scientific, San Jose,<br />

CA). For the Q-trap 2000 mass spectrometer, the mass spectrometer<br />

curtain gas (N 2) was kept as 20 (manufacturer’s units) and<br />

the declustering potential was set at 10 V. Collision induced<br />

dissociation (CID) was carried out to provide ion structural<br />

information using enhanced product ion scan mode. Precursor<br />

ion scanning (PIS) was used to select reaction products by<br />

monitoring the characteristic fragment ions, and N2 was used<br />

as the collision gas. Data acquisition was performed using the<br />

Analyst software (version 1.4.2, Applied Biosystems/MDS<br />

SCIEX, Concord, Canada). Deconvolution of mass spectra was<br />

carried out using Mag-Tran 1.03b2 software (Amgen Inc.,<br />

Thousand Oaks, CA) written based on the ZScore algorithm. 40<br />

High-Resolution MS. High-resolution LTQ-Orbitrap mass<br />

spectrometer (Thermo Electron, Bremen, Germany) with a<br />

(40) Zhang, Z.; Marshall, A. J. Am. Soc. Mass Spectrom. 1998, 9, 225–233.<br />

6928 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

modified nanoelectrospray ionization (nano-ESI) was used for<br />

collecting high-resolution data. The system was operated in the<br />

positive ion mode with a resolving power of 60 000 at m/z 400.<br />

The Molecular Weight Calculator (http://ncrr.pnl.gov/software/)<br />

was used to simulate the isotope distribution for thiol-derivatized<br />

products.<br />

<strong>Chemical</strong>s. The 20 amino acids, angiotensin II human (FW<br />

1046 Da), bradykinin acetate (FW 1060 Da), MRFA (Met-Arg-<br />

Phe-Ala) acetate salt (FW 523 Da), �-lactoglobulin A from bovine<br />

milk, tris(2-carboxyethyl)phos-phine hydrochloride (TCEP), DLdithiothreitol<br />

(DTT), TPCK-treated trypsin from bovine pancreas<br />

(MW ∼23.8 KDa), ammonium bicarbonate, 1,4-benzoquinone, and<br />

N-(phenylseleno)phthalimide were purchased from Sigma-Aldrich<br />

(St. Louis, MO). Glutathione (GSH) (reduced form, MW 307 Da)<br />

and ebselen were obtained from TCI America (Tokyo, Japan) and<br />

Calbiochem (Cincinnati, OH), respectively. HPLC-grade methanol<br />

and acetonitrile from GFS <strong>Chemical</strong>s (Columbus, OH) and Sigma-<br />

Aldrich (St. Louis, MO) were used, and acetic acid was purchased<br />

from Fisher Scientific (Pittsburgh, PA). The deionized water used<br />

for sample preparation was obtained using a Nanopure Diamond<br />

Barnstead purification system (Barnstead International, Dubuque,<br />

IA).<br />

Protein Reduction. A volume of 175 µL of 0.1 mM �-lactoglobulin<br />

A in methanol/water (1:1 by volume) containing 2% acetic<br />

acid and 17.5 µL of 50 mM TCEP in 20 mM ammonium<br />

bicarbonate aqueous solution were mixed resulting in the molar<br />

ratio of 1:50 (protein/TCEP). The protein was reduced by TCEP<br />

for 3.5 h at room temperature. Then Millipore-ZipTip Pipette Tips<br />

were used to remove TCEP and ammonium bicarbonate via<br />

desalting.<br />

RESULTS AND DISCUSSION<br />

Reaction Selectivity. In this experiment, the selectivity of the<br />

derivatization reaction was examined first using amino acids as


Figure 2. (a) ESI-MS spectrum showing the reaction of GSH with ebselen; (b) CID MS 2 spectrum of ebselen derivatized GSH product ion (m/z<br />

583); (c) ESI-MS spectrum showing the reaction of GSH with reagent 2; (d) CID MS 2 spectrum of the product ion of GSH derivatized by reagent<br />

2 (m/z 464); (e) ESI-MS spectrum showing the selective reaction of ebselen with GSH in the presence of angiotensin II, bradykinin, and MRFA;<br />

(f) precursor ion scanning based on the monitoring of the characteristic fragment ion of m/z 276 showing the selective detection of GSH.<br />

reaction substrates and using ebselen as a reagent. We found that<br />

ebselen reacts exclusively with side chains of cysteine residues<br />

that contain a free thiol. As illustrated in Figure 1, the ESI-MS<br />

spectrum displays the reaction of ebselen (10 µM) with cysteine<br />

(5 µM) in the presence of methionine, tryptophan, and valine (5<br />

µM for each) in methanol/water (1:1 by volume) containing 0.5%<br />

acetic acid. A dominant peak at m/z 397 corresponds to the<br />

reaction product between ebselen and cysteine. The inset shows<br />

the zoom-in area for the reaction product ion C 16H17O3N2SSe (m/z<br />

397) (bottom) in comparison with the calculated product<br />

isotope distribution (top). Exact match is observed, confirming<br />

the peak assignment. The characteristic isotope distribution<br />

helps to identify products of the derivatization reaction. Upon<br />

CID, the ion at m/z 397 fragments into m/z 379, 353, and 276<br />

by losses of one water molecule, one carbon dioxide, and one<br />

cysteine, respectively (data not shown), confirming the product<br />

ion structure and the covalent nature of the newly formed bond.<br />

In addition, peaks at m/z 118, 122, 150, 205, and 276 are seen<br />

in the Figure 1, corresponding to the protonated valine, cysteine,<br />

methionine, tryptophan, and ebselen, respectively. We also tested<br />

the reaction of ebselen with cysteine in the presence of 19 other<br />

natural amino acids (Figure 1-S in the Supporting Information).<br />

Again, only the reaction product of ebselen and cysteine (m/z<br />

397) were observed, demonstrating that ebselen has exclusive<br />

specificity toward cysteine derivatization. Similarly, it was also<br />

found that another selenium reagent 2 has similar selectivity for<br />

cysteine in the presence of other amino acids. Furthermore, in<br />

our observation, the derivatization reaction takes place rapidly and<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6929


Figure 3. ESI-MS spectrum showing the recovery of unmodified<br />

GSH 30 min after adding DTT to the reaction mixture of GSH and<br />

ebselen.<br />

is completed in seconds at room temperature. 41,42 This is an<br />

advantage for shortening analysis time in real-world applications,<br />

particularly for high-throughput proteomics analysis.<br />

Reactions with Peptide Thiols. In addition to reacting with<br />

amino acid cysteine, both ebselen and the selenium reagent 2<br />

were also tested with thiol peptides. Figure 2a displays the ESI-<br />

MS spectrum showing the reaction of GSH (0.1 mM, the structure<br />

of this tripeptide is shown in the figure inset) with ebselen (0.2<br />

mM) in acetonitrile/water (1:1 by volume) containing 1% acetic<br />

acid. As expected, the reaction product resulting from the reaction<br />

of ebselen with GSH (m/z 583) was observed. As shown in the<br />

CID spectrum of m/z 583 (Figure 2b), water loss (the formation<br />

of m/z 565), backbone cleavages (the formation of m/z 508 and<br />

m/z 454), and the protonated GSH (m/z 308) and ebselen (m/z<br />

276) were seen, confirming the product ion structure and the<br />

addition of peptide thiol onto ebselen (step a of eq 1 in Scheme<br />

1). It can be seen that, besides backbone cleavages, Se-S bond<br />

cleavage was also observed leading to the formation of the<br />

fragment ion of m/z 276 and 308, which may be driven by the<br />

reformation of the five-membered ring of ebselen as assisted via<br />

the nucleophilic attack of selenium by the adjacent amide nitrogen<br />

of the ebselen tag (shown in the inset of Figure 2b). On the basis<br />

of this hypothesis, the CID behavior of the derivatized peptide<br />

ions could be tuned, simply by removing the adjacent amide group<br />

of the derivatizing selenium reagent. We thus tested the reagent<br />

2. Figure 2c illustrates ESI-MS spectrum showing the reaction of<br />

GSH (0.1 mM) with reagent 2 (0.5 mM) in acetonitrile/water (1:1<br />

by volume) containing 1% acetic acid, in which the reaction product<br />

(m/z 464) is seen. As shown in eq 2 in Scheme 1, the derivatized<br />

thiol tag does not contain amide group as the reagent 2 is split<br />

during the reaction. Indeed, the CID of the product ion at m/z<br />

464 (Figure 2d) does not lead to the cleavage of the Se-S bond.<br />

Instead, it shows the loss of one water molecule (the formation<br />

of m/z 446), the backbone cleavages (the formation of m/z 389,<br />

335, and 318), and the cleavage of the S-C bond (the formation<br />

of m/z 189). In this case, the selenium tag is stable, surviving the<br />

CID process, which would be valuable in top-down proteomics<br />

(41) Cotgreave, I. A.; Mogenstern, R.; Engman, L.; Ahokas, J. Chem.-Biol.<br />

Interact. 1992, 84, 69–76.<br />

(42) Haenen, G. A. M. M.; Rooij, B. M. D.; Vermeulen, N. P. E.; Bast, A. Mol.<br />

Pharm. 1990, 37, 412–422.<br />

6930 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

studies (e.g., for pinpointing thiol location). In Figure 2c, one peak<br />

at m/z 191 corresponding to the protonated PhSe(dO)OH<br />

probably arose from the impurity of the reagent 2 used, which<br />

was technical grade.<br />

In addition, it appears that ebselen is more reactive than<br />

reagent 2, probably due to the tension of its five-membered ring.<br />

The calculated Se-N bond length of ebselen of 1.916 Å is longer<br />

than that of reagent 2 (1.855 Å, Figure 2-S in the Supporting<br />

Information), so the Se-N bond of ebselen is easier to break by<br />

thiol than that of reagent 2. Further support for this conclusion<br />

comes from the fact that nearly all of the GSH was reacted with<br />

ebselen in the mixing ratio of 1:2 (GSH to ebselen, Figure 2a),<br />

while a small amount of GSH remained when the ratio of GSH to<br />

reagent 2 used was 1:5 (Figure 2c).<br />

In the case of labeling by ebselen, the facile cleavage of Se-S<br />

could be useful in selective detection of cysteine containing<br />

peptides from mixtures, based on the resulting characteristic<br />

fragment ion of m/z 276. In this experiment, selective reaction of<br />

ebselen (0.1 mM) with GSH (50 µM) in the presence of other<br />

peptides of angiotensin II, bradykinin, and MRFA (50 µM for each)<br />

in methanol/water (1:1 by volume) containing 1% acetic acid was<br />

carried out. As illustrated in Figure 2e, the ESI-MS spectrum<br />

displays that the reaction of ebselen with GSH is highly selective,<br />

only a product (m/z 583) resulting from the reaction of ebselen<br />

with GSH was detected. However, the spectrum has multiple<br />

peaks and appears complicated. Precursor ion scanning (PIS) was<br />

applied to ions generated by ESI of the same reaction mixture<br />

and a much clearer spectrum containing only m/z 276 (the<br />

protonated ebselen) and 583 (the protonated ion of ebselenderivatized<br />

GSH) are seen (Figure 2f), in which cysteinecontaining<br />

peptide GSH can be rapidly identified.<br />

Reversibility of a derivatization reaction is very important as it<br />

allows enrichment and purification of analyte compounds from<br />

complex matrices. 43 However, as already mentioned above,<br />

commonly used thiol derivatization reactions such as nucleophilic<br />

substitution and Michael-addition are irreversible, limiting their<br />

analytical utilities. In this study, the reversibility of the thiol<br />

derivatization reaction was investigated by using dithiothreitol<br />

(DTT) as a reductant. As shown in Figure 3, the ESI-MS spectrum<br />

reveals the recovery of GSH (m/z 308) 30 min after adding DTT<br />

(0.5 mM) to the reaction mixture containing GSH (5 µM) and<br />

ebselen (10 µM). In the spectrum, one can see that m/z 583, the<br />

protonated molecule of the ebselen-derivatized GSH, disappears<br />

upon reaction with DTT. In addition, m/z 553 arose, probably<br />

indicating that the resulting reduction product, ebselen selenol<br />

(compound 5, Scheme 1), was air sensitive and readily oxidized<br />

by O 2 in the ambient environment into a more stable diselenide<br />

Se-Se product. 41,44 The CID spectrum of m/z 553 shows the<br />

loss of PhNH2 and the formation of the protonated ebselen (m/z<br />

276, Figure 3-Sa in the Supporting Information), in agreement<br />

with the peak assignment. The m/z 575 corresponds to the<br />

sodiated molecule of the diselenide (see its CID in Figure 3-Sb in<br />

the Supporting Information). These results clearly demonstrate<br />

the excellent reversibility of the thiol derivatization. In combination<br />

with the extraordinary selectivity of the reaction, this reactivity<br />

(43) Smith, M. E. B.; Schumacher, F. F.; Ryan, C. P.; Tedaldi, L. M.; Papaioannou,<br />

D.; Waksman, G.; Caddick, S.; Baker, J. R. J. Am. Chem. Soc. 2010, 132,<br />

1960–1965.<br />

(44) Engman, L.; Hallberg, A. J. Org. Chem. 1989, 54, 2964–2966.


Figure 4. (a) ESI-MS spectrum showing �-lactoglobulin A; (b) ESI-MS spectrum showing the �-lactoglobulin A fully derivatized by ebselen<br />

(charge numbers are labeled with primes). The insets show the corresponding deconvoluted spectra; (c) high-resolution Orbitrap MS spectrum<br />

showing the exact match of the isotopic peak distribution between the simulated (shown in blue discrete line) and detected peak (shown in black<br />

solid line) of 16+ derivatized �-lactoglobulin A ions; (d) ESI-MS spectrum showing the partially derivatized �-lactoglobulin A (charge numbers<br />

are labeled with double primes) 30 s after mixing the protein (5 µM) with 1,4-benzoquinone (10 µM); in this spectrum, the major peaks correspond<br />

to intact protein ions.<br />

can be used in a variety of applications focused on enrichment<br />

and purification of biological thiols from cells and tissues.<br />

Reaction with Protein Thiols. Selenium derivatizing reagents<br />

were used for identification of free cysteine residues in proteins<br />

due to the high selectivity for thiol groups. �-Lactoglobulin A used<br />

as a model system in this study (162 amino acid residues) contains<br />

two disulfide bridges (Cys 66 -Cys 160 and Cys 106 -Cys 119 ) and one<br />

free Cys 121 . 45 Figure 4a shows the ESI-MS spectrum of �-lactoglobulin<br />

A (5 µM) in methanol/water (1:1 by volume) containing<br />

1% acetic acid. Multiply charged ions of �-lactoglobulin A are<br />

detected, and deconvolution of the mass spectrum provides the<br />

protein mass of 18 364 Da (shown in the inset of Figure 4a). In<br />

Figure 4b, the ESI-MS spectrum shows the reaction of �-lactoglobulin<br />

A (5 µM) and ebselen (10 µM) in methanol/water (1:1<br />

by volume) containing 1% acetic acid. Deconvolution of the mass<br />

spectrum of ebselen-derivatized �-lactoglobulin A provides a mass<br />

of 18 639 Da (shown in the inset of Figure 4b). After derivatization<br />

of �-lactoglobulin A by ebselen, a mass gain of 275 Da was<br />

observed, indicating addition of one ebselen (MW 275 Da) to the<br />

sole free cysteine residue of the protein. Furthermore, only the<br />

ions of the ebselen-derivatized �-lactoglobulin A were seen in the<br />

spectrum shown in Figure 4b, suggesting that all of the protein<br />

was reacted with ebselen and a quantitative conversion yield was<br />

obtained. High-resolution mass analysis using Orbitrap MS was<br />

used for unambiguous identification of the derivatization product.<br />

(45) Surroca, Y.; Haverkamp, J.; Heck, A. J. R. J. Chromatogr., A 2002, 970,<br />

275–285.<br />

Figure 4c displays the Orbitrap MS spectrum showing the exact<br />

match between the simulated (shown in blue discrete line) and<br />

experimentally observed (shown in black solid line) isotope<br />

distribution of the 16+ charge state of the derivatized �-lactoglobulin<br />

A ions. Furthermore, we performed tandem MS analysis<br />

using the LTQ-Orbitrap. High-resolution Orbitrap CID MS 2<br />

spectrum of the 16+ derivatized �-lactoglobulin A ions (m/z<br />

1166, in Figure 4-S in the Supporting Information) contains<br />

fragment ions y′139 12+ and y′130 11+ that carry the selenium tags.<br />

This suggests that the derivatization site is located on the last<br />

130 amino acid residues of the protein, in agreement with the<br />

position of the free cysteine residue in �-lactoglobulin A<br />

(Cys 121 ). These results indicate the potential use of the selenium<br />

derivatization strategy in top-down proteomics studies.<br />

In the case of �-lactoglobulin A derivatization, ebselen was<br />

compared with one commonly used thiol tagging reagent, 1,4benzoquinone.<br />

It was found by ESI-MS that the protein can be<br />

fully reacted with ebselen 30 s after mixing (The spectrum is the<br />

same as shown in Figure 4b). By contrast, under the same<br />

conditions (e.g., concentrations and solvents used for the reaction<br />

were kept the same), only ∼30% protein was derivatized with 1,4benzoquinone<br />

30 s after mixing the protein and the reagent. This<br />

indicates that the labeling using selenium reagents is much faster<br />

and more efficient than using the Michael-addition reagents such<br />

as 1,4-benzoquinone. The result suggests that the selenium<br />

chemistry would be quite useful in the high-throughput analysis<br />

of thiol containing proteins.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6931


Figure 5. ESI spectrum showing the derivatized protein resulting<br />

from reacting the reduced �-lactoglobulin A with ebselen. The<br />

superscript in the charge number shows the number of selenium tags<br />

added to the protein after derivatization.<br />

High selectivity and efficiency of the selenium chemistry<br />

investigated in this study makes it useful in identification of the<br />

number of free and bound thiol groups in proteins, which is of<br />

importance in the protein structural analysis. The reaction of intact<br />

protein �-lactoglobulin A with ebselen as described above shows<br />

that the protein has only one free cysteine residue. We further<br />

examined the derivatization reaction with reduced protein. In the<br />

experiment, the �-lactoglobulin A was first reduced by TCEP,<br />

which is known to be more stable and effective to reduce disulfide<br />

bonds than DTT. 46 After reduction and removal of the excess<br />

amount of TECP, the reagent ebselen was added to the protein<br />

solution for thiol derivatization. As shown in Figure 5, the reduced<br />

�-lactoglobulin A containing three selenium tags has a high<br />

relative abundance. It is likely that the reduction of disulfide bond<br />

of Cys 66 -Cys 160 is easier than the other Cys 106 -Cys 119 bond<br />

leading to reduced protein mainly with three free thiols. 47<br />

Another possible reason is that the reduced protein is partially<br />

folded so that it is not easy for the relatively large ebselen<br />

reagent to access all of the free thiols. Nevertheless, in Figure<br />

5, the multiply charged ions of fully modified proteins with five<br />

selenium tags were clearly observed, suggesting that the reduced<br />

protein has a maximum of five free cysteine residues. This<br />

indicates that the four additional free cysteine residues result from<br />

reduction and the protein has two disulfide bonds prior to<br />

reduction, which is exactly in agreement with the known structure<br />

(46) Han, J. C.; Han, G. Y. Anal. Biochem. 1994, 220, 5–10.<br />

(47) Sakai, K.; Sakurai, K.; Sakai, M.; Hoshino, M.; Goto, Y. Protein Sci. 2000,<br />

9, 1719–1729.<br />

6932 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

of �-lactoglobulin A as mentioned before. Thus, it is possible to<br />

identify the number of free cysteine, total cysteine residues, and<br />

disulfide linkages in proteins using a simple approach based on<br />

the selenium chemistry.<br />

CONCLUSIONS<br />

In summary, ebselen and N-(phenylseleno)phthalimide, as<br />

Se-N bond containing compounds, are excellent labeling reagents<br />

for characterization of thiol-containing compounds by mass<br />

spectrometry. In this study we examined a series of reactions of<br />

these two selenium reagents with amino acids, peptides, and<br />

proteins. Our study reveals that the thiol derivatization reaction<br />

is highly selective, rapid, reversible, and efficient (quantitative in<br />

the case of �-lactoglobulin A derivatization by ebselen), which is<br />

of high value in proteomics research. In comparison to the wellknown<br />

Michael-addition reactions used for thiol tagging, selenium<br />

reagents appear to be more efficient and faster. <strong>Analytical</strong><br />

applications stemming from this investigation include (i) fast<br />

screening of peptides/proteins containing free cysteine residues<br />

from complex mixtures and (ii) identification of the number of<br />

free and bound thiols of proteins and their locations using MS/<br />

MS experiments. Given the significance of thiols in life and the<br />

important reaction features uncovered, it is expected that there<br />

will be many novel MS applications based on the powerful<br />

selenium chemistry reported in this study.<br />

ACKNOWLEDGMENT<br />

This work was supported by U.S. NSF (Grant CHE-0911160),<br />

National Basic Research Program of China (973 Program, Grant<br />

2007CB936000), National Natural Science Funds for Distinguished<br />

Young Scholar (Grant No. 20725518), National Natural Science<br />

Foundation of China (Grant No. 20875057), and Natural Science<br />

Foundation of Shandong Province in China (Grant No. Y2007B02).<br />

Part of the research described in this manuscript was performed<br />

at the W. R. Wiley Environmental Molecular Sciences Laboratory<br />

(EMSL), a national scientific user facility sponsored by the U.S.<br />

Department of Energy’s Office of Biological and Environmental<br />

Research and located at Pacific Northwest National Laboratory<br />

(PNNL). PNNL is operated by Battelle for the U.S. Department<br />

of Energy. We also thank Mr. Xiaoyong Lu for his help.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional supporting mass spectra. This material is available<br />

free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 4, 2010. Accepted July 6, 2010.<br />

AC1011602


Anal. Chem. 2010, 82, 6933–6939<br />

Difference between Ultramicroelectrodes and<br />

Microelectrodes: Influence of Natural Convection<br />

Christian Amatore,* Cécile Pebay, Laurent Thouin,* Aifang Wang, and J-S. Warkocz<br />

Ecole Normale Supérieure, Département de Chimie, UMR CNRS-ENS-UPMC 8640 “Pasteur”, 24 rue Lhomond,<br />

F-75231 Paris Cedex 05, France<br />

Natural convection in macroscopically immobile solutions<br />

may still alter electrochemical experiments performed<br />

with electrodes of micrometric dimensions. A model<br />

accounting for the influence of natural convection allowed<br />

delineating conditions under which it interferes with mass<br />

transport. Several electrochemical behaviors may be<br />

observed according to the time scale of the experiment,<br />

electrode dimensions, and intensity of natural convection.<br />

The range of parameters in which ultramicrelectrodes<br />

behave under a true diffusional steady state was identified.<br />

Mapping of concentration profiles was performed experimentally<br />

by scanning electrochemical microscopy in the<br />

vicinity of microelectrodes of various radii. The results<br />

validated remarkably the predictions of the model, evidencing<br />

in particular the alteration of the diffusional<br />

steady state by natural convection.<br />

Microelectrodes are versatile tools for the study of electrochemical<br />

processes of mechanistic and/or analytical interest. Their<br />

advantageous properties stem from their small size. Microelectrodes<br />

may be used in highly resistive environments and in very<br />

small sample volumes. They enable the detection of very small<br />

amounts of material and allow short time responses. 1-9 However,<br />

the definition of a microelectrode is still nowadays ambiguous.<br />

Actually, the notion of a microelectrode differs greatly according<br />

to the particular origin of electrochemists, i.e., electroanalytical<br />

chemists or molecular electrochemists. The term microelectrode<br />

may then encompass electrodes of either millimetric or micrometric<br />

dimensions. Electrodes of smaller sizes are referred to as<br />

ultramicroelectrodes. Such definitions, based mainly on historical<br />

* To whom correspondence should be addressed. E-mail: christian.amatore@<br />

ens.fr (C.A.); laurent.thouin@ens.fr.<br />

(1) Fleischmann, M.; Pons, S.; Rolison, D. R. Ultramicroelectrodes; Datatech<br />

Systems, Inc.: Morgantown, NC, 1987.<br />

(2) Bond, A. M.; Oldham, K. B.; Zoski, C. G. Anal. Chim. Acta 1989, 216,<br />

177–230.<br />

(3) Wightman, R. M.; Wipf, D. O. Electroanalytical <strong>Chemistry</strong>; Marcel Dekker:<br />

New York, 1989; Vol. 15, pp 267-353.<br />

(4) Montenegro, M. I.; Queiros, M. A.; Daschbach, J. L. Microelectrodes: Theory<br />

and Applications; Kluwer Academic Press: Dordrecht, The Netherlands,<br />

1991; Vol. 197.<br />

(5) Aoki, K. Electroanalysis 1993, 5, 627–639.<br />

(6) Heinze, J. Angew. Chem., Int. Ed. 1993, 32, 1268–1288.<br />

(7) Amatore, C. Electrochemistry at ultramicroelectrodes. In Physical Electrochemistry;<br />

Rubinstein, I., Ed.; Marcel Dekker: New York, 1995.<br />

(8) Stulik, K.; Amatore, C.; Holub, K.; Marecek, V.; Kutner, W. Pure Appl. Chem.<br />

2000, 72, 1483–1492.<br />

(9) Forster, R. J. Encyclopedia of Electrochemistry; John Wiley & Sons: New<br />

York, 2003; Vol. 3, pp 160-195.<br />

criteria, may appear useless since they better define the origin of<br />

the users than the object itself. A better classification of these<br />

electrodes would be obtained if it were based on their particular<br />

properties. Since electrochemical reactions are interfacial reactions,<br />

mass transport is one of the key processes to consider. 10<br />

In a liquid, elementary contributions in the mass transport are<br />

diffusion, migration, and convection. Under most circumstances,<br />

migration is suppressed by adding a large excess of dissociated<br />

inert salt or supporting electrolyte. Convection is often neglected<br />

at electrodes of micrometric dimensions in macroscopically still<br />

solutions. Indeed, convection originates from movement of fluid<br />

packets of micrometric size. 11 It necessarily vanishes close to the<br />

electrode interface over distances where concentrations differ<br />

significantly from their bulk values. 12,13 In such a case, only<br />

diffusion is assumed to govern the final approach of an electroactive<br />

molecule toward the electrode. However, according to the<br />

size of these electrodes and time scale of the experiments,<br />

convective fluxes due to natural convection may still compete with<br />

diffusional fluxes in motionless solutions. This may occur even<br />

in the absence of any density gradients 14 or effects induced by a<br />

magnetic field. 15 These situations arise as soon as the thickness<br />

of the diffusion layer becomes comparable to the thickness of the<br />

convection-free domain. 7 Under such conditions, the responses<br />

do not follow the classical relationships given for currents in<br />

dynamic and steady-state regimes. Therefore, under given experimental<br />

conditions, it is of importance to decide the largest<br />

size of an electrode for eliminating any influence of natural<br />

convection. 16,17 Such a criterion may then allow distinguishing<br />

properties of ultramicroelectrodes from those of other electrodes<br />

of micrometric sizes.<br />

To assess the conditions of convection-free regimes at electrodes<br />

of micrometric dimensions, we investigated in some<br />

previous studies the current responses of micrometric disk<br />

(10) Bard, A. J.; Faulkner, L. R. Electrochemical Methods, 2nd ed.; John Wiley &<br />

Sons: New York, 2001.<br />

(11) Moreau, M.; Turq, P. <strong>Chemical</strong> Reactivity in Liquids: Fundamental Aspects;<br />

Kluwer Academic/Plenum Press: New York, 1988; pp 561-606.<br />

(12) Levich, V. G. Physicochemical Hydrodynamics; Prentice Hall: Englewoods<br />

Cliffs, NJ, 1962.<br />

(13) Davies, J. E. Turbulence Phenomena; Academic Press: New York, 1972.<br />

(14) Li, Q. G.; White, H. S. Anal. Chem. 1995, 67, 561–569.<br />

(15) Grant, K. M.; Hemmert, J. W.; White, H. S. J. Electroanal. Chem. 2001,<br />

500, 95–99.<br />

(16) Hapiot, P.; Lagrost, C. Chem. Rev. 2008, 108, 2238–2264.<br />

(17) Molina, A.; Gonzalez, J.; Martinez-Ortiz, F.; Compton, R. G. J. Phys. Chem.<br />

C 2010, 114, 4093–4099.<br />

10.1021/ac101210r © 2010 American <strong>Chemical</strong> Society 6933<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/26/2010


electrodes in various conditions, in chronoamperometry 18 and<br />

cyclic voltammetry. 19 The mapping of concentration profiles was<br />

alsoperformedintheirvicinityusingamethodalreadydescribed. 18-23<br />

At the same time, we proposed a theoretical model to evaluate<br />

the influence of natural convection on mass transport in still<br />

media. 18 The good agreement observed between theory and<br />

experiments demonstrated the validity of this model over a wide<br />

range of experimental conditions. 18,22-26 The purpose of this study<br />

is then to take the benefit of this model to delineate the<br />

experimental conditions that allow convection-free regimes to be<br />

observed in dynamic and steady-state regimes. These conditions<br />

are better presented in a zone diagram showing the influence of<br />

all the parameters: time scale of the experiment, electrode radius,<br />

and thickness of the convection-free domain. Comparison with<br />

experimental data will also serve to illustrate the relative contributions<br />

of convection and diffusion at electrodes of different sizes<br />

performing under the steady-state regime.<br />

MODEL OF NATURAL CONVECTION<br />

The influence of convection on the electrode responses can<br />

be quantified from deviations of their diffusive currents or from<br />

alteration of their diffusion layers. Under pure diffusional conditions,<br />

the concentration profile of a species at a disk electrode is<br />

given by integration of Fick’s second law: 10<br />

∂c(r, z, t)<br />

∂t<br />

) D( ∂2c(r, z, t)<br />

∂r 2<br />

+ ∂2c(r, z, t)<br />

∂z 2<br />

+ 1 ∂c(r, z, t)<br />

r ∂r )<br />

(1)<br />

where r describes the radial position normal to the axis of<br />

symmetry at r ) 0 and z describes the linear displacement normal<br />

to the plane of the electrode at z ) 0. D is the diffusion coefficient.<br />

For a chronoamperometric experiment, the pertinent boundary<br />

conditions are<br />

t < 0; r, z g 0; c(r, z, t) ) c° (2)<br />

t g 0; r e r 0 ; c(r,0,t) ) 0 (3)<br />

r, z f ∞; c(r, z, t) ) c° (4)<br />

The current is readily obtained from integration of the concentration<br />

gradients at the electrode surface with<br />

(18) Amatore, C.; Szunerits, S.; Thouin, L.; Warkocz, J. S. J. Electroanal. Chem.<br />

2001, 500, 62–70.<br />

(19) Amatore, C.; Pebay, C.; Thouin, L.; Wang, A. F. Electrochem. Commun.<br />

2009, 11, 1269–1272.<br />

(20) Amatore, C.; Pebay, C.; Scialdone, O.; Szunerits, S.; Thouin, L. Chem.sEur.<br />

J. 2001, 7, 2933–2939.<br />

(21) Amatore, C.; Szunerits, S.; Thouin, L.; Warkocz, J. S. Electroanalysis 2001,<br />

13, 646–652.<br />

(22) Amatore, C.; Knobloch, K.; Thouin, L. Electrochem. Commun. 2004, 6, 887–<br />

891.<br />

(23) Baltes, N.; Thouin, L.; Amatore, C.; Heinze, J. Angew. Chem., Int. Ed. 2004,<br />

43, 1431–1435.<br />

(24) Rudd, N. C.; Cannan, S.; Bitziou, E.; Ciani, L.; Whitworth, A. L.; Unwin,<br />

P. R. Anal. Chem. 2005, 77, 6205–6217.<br />

(25) Amatore, C.; Sella, C.; Thouin, L. J. Electroanal. Chem. 2006, 593, 194–<br />

202.<br />

(26) Amatore, C.; Knobloch, K.; Thouin, L. J. Electroanal. Chem. 2007, 601,<br />

17–28.<br />

6934 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

r0 ∂c(r, z, t)<br />

i )(2πnFD∫ r ∂r (5)<br />

0 ∂z<br />

In still solutions, natural convection operates perpendicularly<br />

to the electrode surface. It is based on microscopic motions of<br />

the solution except in the very near vicinity of the electrodes,<br />

where it vanishes. Experimentally, the resulting velocity field is<br />

extremely difficult to estimate. Moreover, it is almost impossible<br />

to master mathematically since it depends on many parameters<br />

which are not easy to control (vibrations, temperature gradients,<br />

movement of the cell atmosphere, etc.). However, beyond these<br />

difficulties, we demonstrated successfully that, for electroactive<br />

species, the influence of natural convection can be assimilated to<br />

that of an apparent diffusion coefficient depending on the<br />

orthogonal distance z from the electrode plane. 18 Moreover, since<br />

the electrochemical perturbation affects only the viscous sublayer<br />

adjacent to the electrode, we showed that Dapp could be evaluated<br />

by<br />

z<br />

Dapp ) D( 1 + 1.522( δconv) 4<br />

)<br />

where δconv is the thickness of the convection-free layer. This<br />

is the only parameter introduced into the model to account for<br />

the effects of natural convection. It is possible to evaluate its<br />

influence on the electrode response by replacing D by Dapp in<br />

eq 1 and solving numerically the new mass transport equation in<br />

association with the same boundary conditions (eqs 2-4).<br />

EXPERIMENTAL SECTION<br />

All the solutions were prepared in purified water (Milli-Q,<br />

Millipore). A 10 mM concentration of K4Fe(CN)6 (Acros) was<br />

dissolved in 1 M KCl (Aldrich), which was used as the<br />

supporting electrolyte. Reciprocally 2 mM FcCH2OH (Acros)<br />

was prepared in 0.1 M KNO3 (Fluka). The diffusion coefficients<br />

were DFe ) (6.0 ± 0.5) × 10 -6 cm 2 s -1 27 for Fe(CN)6 4- /<br />

Fe(CN)6 3- and DFc ) (7.6 ± 0.5) × 10 -6 cm 2 s -1 for FcCH2OH/<br />

FcCH2OH + .<br />

The working electrodes were Pt disk electrodes of 12.5, 25,<br />

62.5, 125, 250, and 500 µm radii. They were obtained from the<br />

cross section of Pt wires (Goodfellow) sealed into soft glass. The<br />

reference electrode was a Ag/AgCl electrode, and the counter<br />

electrode was a platinum coil. A scanning electrochemical microscope<br />

(910B CH Instruments) was used to establish the concentration<br />

profiles. The amperometric probe was a Pt disk electrode<br />

of submicrometric dimension (∼500 nm radius). Its fabrication<br />

and the related procedure to map the concentrations have already<br />

been reported. 23 For Fe(CN)6 4- /Fe(CN)6 3- experiments, the<br />

working electrode was biased at +0.6 V/ref on the oxidation<br />

plateau of Fe(CN)6 4- . The probe was biased at +0.6 V/ref to<br />

collect Fe(CN)6 4- or -0.1 V/ref to collect Fe(CN)6 3- . For<br />

FcCH2OH/FcCH2OH + experiments, the working electrode was<br />

biased at +0.25 V/ref. In this case, the probe was biased at<br />

+0.25 V/ref to collect FcCH2OH and -0.1 V/ref to collect<br />

FcCH2OH + .<br />

The mass transport equation was solved numerically in the<br />

conformal space adapted to the geometry of a microdisk elec-<br />

(27) Amatore, C.; Szunerits, S.; Thouin, L.; Warkocz, J.-S. Electrochem. Commun.<br />

2000, 2, 353–358.<br />

(6)


trode 28 by a finite element using Comsol Multiphysics software.<br />

The thickness of the convection-free layer, δconv, was determined<br />

before each experiment by chronoamperometry as described<br />

previously. 18<br />

RESULTS AND DISCUSSION<br />

An important property of disk ultramicroelectrodes is that their<br />

diffusion layers develop with time until reaching a steady-state<br />

limit imposed by hemispherical-type diffusion. However, natural<br />

convection may interfere with the mass transport as soon as the<br />

thickness of the expanding diffusion layer becomes comparable<br />

to δconv. In such a case, a steady-state regime is still achieved<br />

but is then controlled by the respective contributions of<br />

diffusion and natural convection. Depending on the electrode<br />

radius, r0, and the thickness of the convection-free layer, δconv,<br />

two situations may be encountered, whether the diffusion at<br />

the electrode surface is planar or not. Indeed, at short time<br />

scales, the thickness of the diffusion layer is considerably<br />

smaller than the electrode radius. The electrodes then behave<br />

as electrodes of infinite dimensions, and planar diffusion<br />

operates. In that particular situation, the Cottrell equation<br />

applies with<br />

i planar )( nFADc°<br />

√πDt<br />

for electrodes of surface area A. Using the Nernst formulation,<br />

eq 7 is similar to that given in hydrodynamic electrochemical<br />

methods: 10<br />

i )( nFADc°<br />

δ<br />

where δ ) (πDt) 1/2 . Therefore, natural convection interferes<br />

significantly with the mass transport as soon as δconv ≈ (πDt) 1/2 .<br />

Whenever this condition is not met, the diffusion layer may<br />

develop, possibly reaching a hemispherical behavior until being<br />

eventually limited by δconv.<br />

However, since diffusion and natural convection occur together<br />

in steady-state regimes, their contributions in mass transport<br />

remain difficult to comprehend. To solve this problem, one needs<br />

first to investigate the concentration profiles established under<br />

the pure diffusional steady-state regime, i.e., without any influence<br />

of natural convection. In such a case, solution of eq 1 shows that<br />

most of the concentration gradients operate over a distance<br />

comparable to the electrode radius (Figure 1A). Concentration<br />

along the z axis (Figure 1B) varies according to<br />

c 2 z<br />

) arctan( c° π r0) The diffusion layer presents a hemispherical shape, and the<br />

steady-state current is given by:<br />

(7)<br />

(8)<br />

(9)<br />

i hemisph )(4nFr 0 Dc° (10)<br />

(28) Amatore, C.; Fosset, B. J. Electroanal. Chem. 1992, 328, 21.<br />

Figure 1. Steady-state concentration profile simulated at a disk<br />

electrode without considering the influence of natural convection: (A)<br />

2D concentration profile with isoconcentration lines ranging from c/c°<br />

) 0.1 to c/c° ) 0.9. (B) Concentration profile along the vertical axis<br />

of symmetry.<br />

Using the Nernst formulation again, comparison between eqs 8<br />

and 10 leads to an equivalent diffusion layer thickness, δ ) πr0/<br />

4. This thickness differs slightly from δz, which may be obtained<br />

by extrapolating the concentration gradient at z ) 0, r ) 0 along<br />

the z axis. Indeed, when z f 0, the concentration profile at r<br />

) 0 tends to (Figure 1B)<br />

c 2 z<br />

)<br />

c° π r0 (11)<br />

which gives δz ) πr0/2. The difference in the δ and δz values<br />

results from the nonradial distribution of diffusion fields at disk<br />

electrodes. Concentration gradients are higher at the electrode<br />

edges than at the center (Figure 1A). Since δ is evaluated from<br />

the integration of concentration gradients over the whole electrode<br />

surface (eq 5), it is necessarily smaller than δz. In the following,<br />

the variation of these two parameters will provide an accurate<br />

estimation of the influence of natural convection, either from<br />

the current (i.e., through δ) or from the alteration of concentration<br />

profiles in the z direction (i.e., through δz), where natural<br />

convection prevails.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6935


Figure 2. Steady-state concentration profiles and steady-state<br />

currents simulated at disk electrodes of different radii under the<br />

influence of natural convection. (A) Isoconcentration lines c/c° ) 0.9<br />

for disk electrodes of various radii. From left to right, r0/δconv ) 0.05,<br />

0.5, 1.0, 1.5, 2.0, 4.2, and 6.0. (B, C) Variations of the diffusion layer<br />

thicknesses δ and δz as a function of the electrode radius, with (solid<br />

curves) and without (dashed curves) the influence of natural convection.<br />

(D) Error on steady-state currents due to the influence of natural<br />

convection as a function of r0/δconv.<br />

Figure 2A displays isoconcentration lines c/c° ) 0.9 calculated<br />

under the influence of natural convection for electrodes of various<br />

radii. They were obtained from combination of eqs 1-6 under<br />

the steady-state regime. To provide a more general representation<br />

of the problem, the spatial coordinates r and z were normalized<br />

by δconv. When the electrode radii are small enough (i.e., r0/<br />

δconv < 0.5), one observes that the diffusion layers retain their<br />

quasi-hemispherical shapes, like those previously simulated in<br />

the absence of natural convection (Figure 1A). For larger<br />

electrode radii, the concentration profiles become flattened, their<br />

expansion along the z axis being restricted by the boundary at<br />

δconv. Therefore, when convection operates, it has two major<br />

effects depending on the ratio r0/δconv. On one hand, concentration<br />

gradients (∂c/∂z)z)0 become more uniform over the central<br />

area of the electrodes than when natural convection is absent.<br />

On the other hand, the development of the diffusion layers still<br />

operates laterally along the r axis. This can be easily observed<br />

in parts B and C of Figure 2, where the variations of δ/δconv and<br />

δz/δconv, respectively, are reported as a function of r0/δconv. In<br />

particular, one observes that when r0/δconv ) 4, δz has reached<br />

6936 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 3. Zone diagrams describing the influence of natural convection<br />

on planar and hemispherical diffusion at disk electrodes. (A) Zone<br />

diagram established from two boundary conditions, δ ) δconv (solid curve)<br />

and r0/(πDt) 1/2 ) 4/π (vertical straight line). Note that this diagram is<br />

independent of the model of convection. (B) Zone diagram established<br />

on the basis of the present model of natural convection with four<br />

boundary conditions: |δ - δdiff|/δ or |i - idiff|/|idiff| ) 0.1 (curve 1), |δ -<br />

δconv|/δ or |i - iconv|/|iconv| ) 0.1 (curve 2), |δ - δplanar|/δ or |i - iplanar|/<br />

|iplanar| ) 0.1 (curve 3), and |δ - δhemisph|/δ or |i - ihemisph|/|ihemisph| ) 0.1<br />

(curve 4). The black symbols correspond to the experimental conditions<br />

considered in Figure 4: from left to right, r0 ) 12.5, 25, 62.5, 125, 250,<br />

and 500 µm, with δconv ) 200-250 µm.<br />

its limit δconv whereas δ is only equal to 0.8δconv. Accordingly,<br />

the convolution of these two effects on the development of<br />

diffusion layers leads to a deviation of the currents from eq<br />

10. This alteration may be drastic since the relative error |i -<br />

ihemisph|/|ihemisph| increases almost linearly with r0/δconv (Figure<br />

2D). In particular, |i - ihemisph|/|ihemisph| ≈ 0.65 when δconv ≈ r0.<br />

A first attempt to summarize all these situations is to establish<br />

a zone diagram describing the boundary condition imposed by<br />

δconv on δ, whether the diffusion is planar or hemispherical.<br />

As previously mentioned, two other parameters have to be<br />

considered: the electrode radius, r0, and the diffusion length,<br />

(πDt) 1/2 . The transition between planar and quasi-hemispherical<br />

diffusion depends on the ratio r0/(πDt) 1/2 , whereas the influence<br />

of convection is fixed by r0/δconv. Therefore, a diagram<br />

with two coordinates, r0/(πDt) 1/2 and r0/δconv, allows plotting<br />

the limit, which differentiates the domains where convection<br />

or diffusion prevails independently. This limit is then δ ) δconv.


Figure 4. Comparison between simulated (curves) and experimental (symbols) steady-state concentration profiles along the vertical axis of<br />

symmetry at disk electrodes of different radii when the electrode potential is poised on the oxidation plateau of Fe(CN)6 4- . Concentration profiles<br />

simulated without (dashed curves) or with (solid curves) natural convection (δconv ) 200-250 µm). Experimental concentration profiles of the<br />

substrate Fe(CN)6 4- (0) and product Fe(CN)6 3- (O). t ) 60 s. [Fe(CN)6 4- ] ) 10 mM in 1 M KCl.<br />

The diagram is reported in Figure 3A, where the zones above<br />

and below this limit correspond to the control of convection and<br />

diffusion, respectively. In the lower zone, the equality between<br />

eqs 7 and 10 discriminates by a vertical line located at r0/(πDt) 1/2<br />

) 4/π two other domains where planar diffusion (i.e., r0/<br />

(πDt) 1/2 > 4/π) and quasi-hemispherical diffusion (i.e., r0/<br />

(πDt) 1/2 < 4/π) dominate.<br />

One must note that this diagram is independent of the model<br />

of natural convection since δ was calculated without considering<br />

eq 6 and by only assuming δ ) δconv. In this context, the model<br />

enables the transitions between the three regimes of the<br />

diagram to be determined. For this purpose, the model is used<br />

to evaluate δ and to compare its value to a given reference<br />

thickness, δref, predicted by considering only one specific<br />

regime: (1) δref ) δdiff for pure diffusion control without any<br />

influence of convection, (2) δref ) δconv, (3) δref ) δplanar for<br />

planar diffusion with δplanar ) (πDt) 1/2 , and (4) δref ) δhemisph<br />

for hemispherical diffusion with δhemisph ) πr0/4. The transition<br />

from one of these specific regimes to the others may then be<br />

estimated by setting a relative threshold on δ such as<br />

Note that eq 12 is equivalent to<br />

|δ - δref |<br />

) 0.1 (12)<br />

δ<br />

|<br />

i - iref i | ) 0.1 (13)<br />

ref<br />

where iref is the reference current obtained from eq 8 with<br />

i ref )( nFADc°<br />

δ ref<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(14)<br />

6937


The zone diagram built from eq 12 or 13 is reported in Figure<br />

3B. Thus, three curves (curves 2-4) delineate a new domain<br />

corresponding to a mixed regime between the limiting ones<br />

previously identified (Figure 3A). In particular, the transitions from<br />

hemispherical diffusion to convection (i.e., vertical displacement<br />

on the diagram) and from hemispherical diffusion to planar<br />

diffusion (i.e., horizontal displacement) are relatively broad since<br />

they occur approximately over 2 orders of magnitude on the r0/<br />

δconv and r0/(πDt) 1/2 scales, respectively. Only the transition<br />

from planar diffusion to convection is very sharp. Indeed, as<br />

soon as δplanar ≈ δconv, the diffusion layer reaches its steadystate<br />

limit and mass transport is fully controlled by convection.<br />

In contrast, when the condition δz ≈ δconv is met for hemispherical-type<br />

diffusion, the layer may still expand laterally<br />

along the r axis until reaching its steady-state limit (see Figure<br />

2A-C). This latter condition corresponds to curve 1 in Figure<br />

3B when |δ - δdiff|/δ ) 0.1 or |i - idiff|/|idiff| ) 0.1. It allows<br />

delineating the upper zone of the diagram where convection<br />

starts to interfere in the mass transport.<br />

A chronoamperometric experiment can be represented on the<br />

diagram by a horizontal straight line described from the right to<br />

the left when the time duration increases. According to the size<br />

of the electrode, r0, and thickness, δconv, the nature of the steadystate<br />

regime reached at longer time may be different. On the<br />

one hand, if log(r0/δconv) > 0.95, a sharp transition from planar<br />

diffusion to convection occurs. On the other hand, if log(r0/<br />

δconv) < -0.7, a broad transition with a mixed regime from<br />

planar diffusion to quasi-hemispherical diffusion operates<br />

without any influence of natural convection. When log(r0/<br />

(πDt) 1/2 ) < -0.75, a steady-state regime is always observed<br />

though its nature (diffusional or convective) only depends on<br />

the ratio r0/δconv.<br />

Under given experimental conditions (i.e., the same position<br />

of the electrode in the cell, temperature, viscosity of the electrolyte,<br />

environment, etc.), δconv is approximately constant so that the<br />

mass transport regime under steady state depends only on the<br />

electrode dimension. This was checked experimentally by<br />

mapping diffusion layers in the vicinity of electrodes of various<br />

radii. Figure 4 shows the concentration profiles along the vertical<br />

axis of symmetry of the electrodes for both the reactant and<br />

product. δconv was evaluated independently by chronoamperometry<br />

at a large electrode 18 and was found to range from 200<br />

to 250 µm. It was thus possible to compare the experimental<br />

data with concentration profiles predicted with or without<br />

natural convection. A very good agreement was observed in<br />

Figure 4 whatever the size of the electrodes between experimental<br />

data and predictions issued from the model when natural convection<br />

was taken into account. Alterations on the concentration<br />

profiles due to convection were apparent as soon as r0 ) 25 µm.<br />

The experimental conditions pertaining to each concentration<br />

profile in Figure 4 are reported as symbols in the zone diagram<br />

of Figure 3B. According to the threshold previously defined with<br />

|δ - δhemisph|/δ ) 0.1 or |i - ihemisph|/|ihemisph| ) 0.1, the results<br />

show that a hemispherical diffusion regime was reached for r0<br />

) 12.5 and 25 µm while a mixed regime was achieved for the<br />

other radii (r0 ) 62.5-500 µm).<br />

These experimental data validate the predictions of the present<br />

model, yet they involved only the effect of natural convection along<br />

6938 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 5. Comparison between simulated (curves) and experimental<br />

(symbols) steady-state concentration profiles at a disk electrode of<br />

radius r0 ) 25 µm when the electrode potential is poised onto the<br />

oxidation plateau of FeCH2OH. (A) Experimental concentration profiles<br />

of the product FcCH2OH + along the vertical axis of symmetry (circles).<br />

(B) Experimental concentration profiles of FcCH2OH + along the r axis<br />

at various vertical distances z: z ) 6(O), 16 (0), 26 (]), 36 (×), 46<br />

(+), and 56 µm (∆). The black area indicates the extent of the<br />

electrode coordinates along the r axis. The concentration profiles are<br />

simulated without (dashed curves) and with (solid curves) consideration<br />

of the influence of natural convection (δconv ) 200 µm).<br />

[FeCH2OH] ) 2 mM in 0.1 KNO3.<br />

the axis of symmetry of the electrodes. Conversely, we showed<br />

above (see Figure 2) that this effect is also effective along lateral<br />

directions due to the compensation of transport between vertical<br />

and lateral fluxes. In the following, we investigated this latter issue<br />

experimentally by performing 2D imaging. Figure 5 reports the<br />

mapping of concentration profiles established in the steady-state<br />

regime along the z axis and r axis when r0 ) 25 µm. As in Figure<br />

4, the concentration profiles were compared to the predictions<br />

established with and without the influence of convection. Apart<br />

from the good agreement obtained between the data and predictions,<br />

these results clearly illustrate the fact that convection may<br />

still alter the diffusion layers even when quasi-hemispherical<br />

diffusion is expected to prevail (Figure 3B). In the present case,<br />

the concentration profiles are distorted over distances z equivalent<br />

to 10 times the electrode radius, r0. Simultaneously, the relative


Figure 6. Comparison between simulated (curves) and experimental<br />

(symbols) thicknesses of the diffusion layer at disk electrodes of<br />

various radii: δz/δconf (dashed lines, O) and δ/δconf (solid curve, 0).<br />

error in the current obtained by the model is |i - ihemisph|/<br />

|ihemisph| ) 0.07.<br />

Finally, variation of δz issued from the mapping of concentration<br />

profiles in Figure 4 is reported in Figure 6 as a function of<br />

the electrode size and then compared to the predicted one. The<br />

diffusion layer thicknesses, δ, estimated from the experimental<br />

steady-state currents (through eq 8) are also plotted. As observed,<br />

all these data show that the model applies satisfactorily under the<br />

steady-state regime to predict the influence of natural convection<br />

on current responses or concentration profiles.<br />

CONCLUSION<br />

The model elaborated in this work predicts within a very<br />

good accuracy the relative contributions of diffusion and natural<br />

convection to the mass transport at disk electrodes. The<br />

electrochemical behaviors of the electrodes not only are related<br />

to their dimensions but also depend on the time scale of the<br />

experiment and thickness of the convection-free layer (i.e.,<br />

δconv). These results stress once more the futility of trying<br />

to propose an absolute definition of ultramicroelectrodes<br />

based on the objects themselves. Indeed, the same electrode<br />

may behave as a microelectrode or an ultramicroelectrode,<br />

depending on these parameters. Our model allowed us to<br />

clearly delineate the situations where natural convection<br />

alters both the dynamic and steady-state regimes at disk<br />

electrodes. The properties of ultramicroelectrodes are mainly<br />

achieved when r0/δconv < 0.2. This condition has practical<br />

consequences if one needs, for example, to exploit the<br />

characteristics of ultramicroelectrodes to detect or measure<br />

concentrations in restricted volumes, without any alteration<br />

of natural convection on the measurements.<br />

ACKNOWLEDGMENT<br />

This work has been supported in part by the CNRS (Grant<br />

UMR8640), Ecole Normale Superieure, UPMC, and French<br />

Ministry of Research.<br />

Received for review May 7, 2010. Accepted July 9, 2010.<br />

AC101210R<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6939


Anal. Chem. 2010, 82, 6940–6946<br />

Determination of Double Bond Location in Fatty<br />

Acids by Manganese Adduction and Electron<br />

Induced Dissociation<br />

Hyun Ju Yoo and Kristina Håkansson*<br />

Department of <strong>Chemistry</strong>, University of Michigan, 930 North University Avenue, Ann Arbor, Michigan 48109-1055<br />

Double bond locations in fatty acids can be determined<br />

from characteristic charge-remote fragmentation patterns<br />

of alkali metal-adducted fatty acids following high energy<br />

collision activated dissociation (CAD). With low energy<br />

CAD, several chemical derivatization methods, including<br />

ozonization, epoxidation, and hydroxylation, have been<br />

used to generate characteristic fragments. However, high<br />

energy CAD is not universally available and involves a high<br />

degree of scattering, causing product ion loss. Further,<br />

derivatization reactions involve side reactions and sample<br />

loss. Here, we analyzed metal-adducted fatty acids to<br />

investigate the utility of electron induced dissociation<br />

(EID) for determining double bond location. EID has been<br />

proposed to involve both electronic excitation, similar to<br />

high energy CAD, and vibrational exciation. Various<br />

metals (Li, Zn, Co, Ni, Mg, Ca, Fe, and Mn) were<br />

investigated to fix one charge at the carboxylate end of<br />

fatty acids to promote charge-remote fragmentation. EID<br />

of Mn(II)-adducted fatty acids allowed determination of<br />

all double bond locations of arachidonic acid, linolenic<br />

acid, oleic acid, and stearic acid. For Mn(II)-adducted<br />

fatty acids, reduced characteristic charge-remote product<br />

ion abundances at the double bond positions are indicative<br />

of double bond locations. However, other metal<br />

adducts did not generally provide characteristic product<br />

ion abundances at all double bond locations.<br />

Polyunsaturated fatty acids are essential for cell membrane<br />

functioning because many membrane properties, such as<br />

fluidity and permeability, are closely related to the level of<br />

unsaturation. 1,2 Lipid peroxidation results in loss of membrane<br />

polyunsaturated fatty acids. 1 Also, when an oil or fat becomes<br />

oxidized, health concern is due to the potential production of<br />

free radicals, which can be highly carcinogenic. 3-6 Double bond<br />

sites in unsaturated fatty acids and lipids are plausibly oxidized<br />

* To whom correspondence should be addressed. E-mail: kicki@umich.edu.<br />

Tel: (734) 615-0570. Fax: (734) 647 4865.<br />

(1) Farooqui, A. A.; Horrocks, L. A. Cell. Mol. Neurobiol. 1998, 18, 599–608.<br />

(2) Mitchell, T. W.; Pham, H.; Thomas, M. C.; Blanksby, S. J. J. Chromatogr.,<br />

B 2009, 877, 2722–2735.<br />

(3) Yao, D.; Shi, W.; Gou, Y.; Zhou, X.; Yee Aw, T.; Zhou, Y.; Liu, Z. Free Radical<br />

Biol. Med. 2005, 39, 1385–1398.<br />

(4) North, J. A.; Spector, A. A.; Buettner, G. R. Am. J. Physiol. 1994, 267, C177–<br />

188.<br />

(5) Pandey, M.; Sharma, L. B.; Singh, S.; Shukla, V. K. World J. Surg. Oncol.<br />

2003, 1, 5.<br />

(6) Black, H. S. Integr. Cancer Ther. 2004, 3, 279–293.<br />

6940 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

and form free radicals, which can cause tissue damage and<br />

alterations in cell membranes. 1,3,4,7,8 Thus, the identification<br />

of double bond locations in fatty acids can be beneficial for<br />

understanding lipid biology and also its related disease states. 1,2<br />

Double bond locations in aliphatic compounds, including fatty<br />

acids, can be obtained from mass spectrometry (MS). 9-14 Such<br />

information can be determined by charge-remote fragmentation<br />

processes of alkali metal-adducted fatty acids in high energy collisional<br />

activation with fast atom bombardment (FAB) ionization. 9,15<br />

However, FAB desorption of fatty acid mixtures can result in<br />

preferential desorption of some ions, chemical noise, and low<br />

sensitivity. 17,19 In addition, sector-type instruments for high energy<br />

collision activated dissociation (CAD) are not universally available,<br />

and high energy CAD involves a high degree of scattering, causing<br />

product ion loss. 20 Recently, low energy CAD of Cu(II)-adducted<br />

fatty acids in an ion trap instrument was used to provide diagnostic<br />

product ions to aid the identification of double bond locations in<br />

unsaturated fatty acids; 32 however, double bond localization<br />

remains challenging for monounsaturated fatty acids.<br />

Charge remote fragmentation occurs remote from a charge<br />

site and appears to readily occur in high energy CAD of fatty acids,<br />

lipids, steroids, and other compounds containing long alkyl<br />

chains. 15,17,33 Charge remote fragmentation has been used to<br />

provide double bond positions of fatty acids from FAB-MS/MS<br />

(high energy CAD) of lithiated fatty acids using sector-type mass<br />

spectrometry by Gross and others. 17,33,34 More recently, tandem<br />

time-of-flight (TOF/TOF) MS was applied to determine double<br />

bond locations of lithiated fatty acids by McEwen and co-workers<br />

using solvent-free matrix-assisted laser desorption ionization<br />

(MALDI). 10 Charge-remote bond cleavages in fatty acids are<br />

(7) de Kok, T. M.; ten Vaarwerk, F.; Zwingman, I.; van Maanen, J. M.; Kleinjans,<br />

J. C. Carcinogenesis 1994, 15, 1399–1404.<br />

(8) Montine, T. J.; Morrow, J. D. Am. J. Pathol. 2005, 166, 1283–1289.<br />

(9) Adams, J.; Gross, M. L. Anal. Chem. 1987, 59, 1576–1582.<br />

(10) Trimpin, S.; Clemmer, D. E.; McEwen, C. N. J. Am. Soc. Mass Spectrom.<br />

2007, 18, 1967–1972.<br />

(11) Van Pelt, C. K.; Brenna, J. T. Anal. Chem. 1999, 71, 1981–1989.<br />

(12) Buser, H. R.; Arn, H.; Cuerin, P.; Rauscher, S. Anal. Chem. 1983, 55, 818–<br />

822.<br />

(13) Schneider, B.; Budzikiewicz, H. Rapid Commun. Mass Spectrom. 1990, 4,<br />

550–551.<br />

(14) Malosse, C.; Kerhoas, L.; Einhorn, J. J. Chromatogr. 1998, 803, 203–209.<br />

(15) Cheng, C.; Gross, M. L. Mass Spectrom. Rev. 2000, 19, 398–420.<br />

(16) Hayes, R. N.; Gross, M. L. Methods Enzymol. 1990, 193, 237–263.<br />

(17) Adams, J. Mass Spectrom. Rev. 1990, 9, 141–186.<br />

(18) Jensen, N. J.; Tomer, K. B.; Gross, M. L. J. Am. Chem. Soc. 1985, 107,<br />

1863–1868.<br />

(19) Adams, J.; Gross, M. L. Org. Mass Spectrom. 1988, 23, 307–316.<br />

(20) Fallick, A. E. Int. J. Mass Spectrom. Ion Phys. 1983, 46, 59–62.<br />

10.1021/ac101217x © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/27/2010


educed in the vicinity of double bonds, thus providing information<br />

on double bond locations. 16 Fixed charge sites are important for<br />

predominant charge-remote fragmentation processes because<br />

charge-driven processes compete with charge-remote processes.<br />

In charge-driven fragmentation, charge migration results in<br />

rearrangement of chemical structure, making identification of<br />

double bond positions impossible. 15 High energy CAD is known<br />

to involve electronic excitation as a dominant process for generating<br />

charge-remote product ions, and vibrational/rotational excitation<br />

is considerably less efficient. 16<br />

Electron induced dissociation (EID) involves interactions<br />

between singly charged analyte ions and free electrons. This<br />

concept was first shown with 3-9 eV electrons in 1979 by Cody<br />

and Freiser for radical cations. 21 EID does not require multiply<br />

charged precursor ions, thereby differing from other ion-electron<br />

interactions such as electron capture dissociation (ECD) 22-25 and<br />

electron detachment dissociation (EDD). 26-28 Consequently, EID<br />

is compatible with smaller biomolecules (such as fatty acids) for<br />

which formation of gas-phase multiply charged ions is energetically<br />

unfavorable. Zubarev and co-workers applied EID (10-13<br />

eV electron irradiation) to singly charged oligosaccharide cations,<br />

29 and we have applied EID for structural characterization of<br />

phosphate-containing metabolites. 30 EID of phosphate-containing<br />

metabolites provided complementary structural information compared<br />

to CAD and infrared multiphoton dissociation (IRMPD) and<br />

generally generated more extensive fragmentation than the latter<br />

two techniques. 44 O’Hair and co-workers have proposed that<br />

EID occurs via electronic and vibrational excitation, based on<br />

similarities between the types of product ions observed from<br />

EID, ultraviolet photodissociation, and electron ionization (EI)<br />

mass spectra. 31 Thus, EID may be an alternative technique to<br />

high-energy CAD for revealing information on double bond<br />

locations in fatty acids. To our knowledge, EID has not<br />

previously been applied toward fatty acid analysis. We used<br />

several metals (Li, Zn, Co, Ni, Mg, Ca, Fe, and Mn) to fix a<br />

positive charge at the end of fatty acids. Mn(II) adduction<br />

consistently generated charge-remote fragmentation in EID.<br />

The charge-remote product ion abundances at each carbon<br />

location were compared to deduce double bond positions. EID<br />

of Mn(II)-adducted arachidonic acid was compared to IRMPD<br />

of the same species to illustrate that EID involves both<br />

electronic and vibrational excitation in contrast to IRMPD<br />

which only involves the latter.<br />

(21) Cody, R. B.; Freiser, B. S. Anal. Chem. 1979, 51, 541–551.<br />

(22) Zubarev, R. A.; Horn, D. M.; Fridriksson, E. K.; Kelleher, N. L.; Kruger,<br />

N. A.; Lewis, M. A.; Carpenter, B. K.; McLafferty, F. W. Anal. Chem. 2000,<br />

72, 563–573.<br />

(23) Zubarev, R. A.; Kelleher, N. L.; McLafferty, F. W. J. Am. Chem. Soc. 1998,<br />

120, 3265–3266.<br />

(24) Zubarev, R. A. Curr. Opin. Biotechnol. 2004, 15, 12–16.<br />

(25) Cooper, H. J.; Hakansson, K.; Marshall, A. G. Mass Spectrom. Rev. 2005,<br />

24, 201–222.<br />

(26) Budnik, B. A.; Haselmann, K. F.; Zubarev, R. A. Chem. Phys. Lett. 2001,<br />

299–302.<br />

(27) Yang, J.; Mo, J.; Adamson, J. T.; Hakansson, K. Anal. Chem. 2005, 77,<br />

1876–1882.<br />

(28) Wolff, J. J.; Chi, L. L.; Linhardt, R. J.; Amster, I. J. Anal. Chem. 2007, 79,<br />

2015–2022.<br />

(29) Budnik, B. A.; Haselmann, K. F.; Elkin, Y. N.; Gorbach, V. I.; Zubarev, R. A.<br />

Anal. Chem. 2003, 75, 5994–6001.<br />

(30) Yoo, H. J.; Liu, H.; Hakansson, K. Anal. Chem. 2007, 20, 7858–7866.<br />

(31) Lioe, H.; O’Hair, R. A. Anal. Bioanal. Chem. 2007, 389, 1429–1437.<br />

EXPERIMENTAL SECTION<br />

Sample Preparation. Fatty acids used in this work include<br />

stearic acid, oleic acid, linolenic acid, and arachidonic acid. Fatty acids<br />

and metal salts, including MnCl2, CoBr2, and NiBr2, were purchased<br />

from Sigma-Aldrich (St. Louis, MO). Fatty acid (70-200 µM) was<br />

mixed with 200-600 µM metal salt in methanol/water (80/20<br />

v/v). Sample solutions of metal (Met)-adducted fatty acids were<br />

freshly made 10-30 min prior to MS analysis.<br />

Fourier Transform Ion Cyclotron Resonance Mass Spectrometry.<br />

Singly charged metal-adducted fatty acids, [M + Met<br />

- H] + ([M + 2Met - H] + for Li), were generated by external<br />

electrospray ionization (ESI) at 70 µL/h (Apollo II dual stage<br />

ion funnel ion source, Bruker Daltonics, Billerica, MA). All<br />

experiments were performed with a 7 T quadrupole (Q)-FTICR<br />

mass spectrometer (APEX-Q, Bruker Daltonics) as previously<br />

described. 28 All data were obtained in positive ion mode.<br />

Briefly, ions produced by ESI were mass-selectively externally<br />

accumulated 32,33 in a hexapole for 0.1-2 s, transferred via high<br />

voltage ion optics, and captured in the ICR cell by dynamic<br />

trapping. This accumulation sequence was looped three times to<br />

improve precursor ion abundance. In MS/MS experiments, mass<br />

selective external accumulation of [M + Met - H] + ([M + 2Met<br />

- H] + for Li) was employed. In some cases, mass selective<br />

external accumulation was followed by further isolation via<br />

correlated harmonic excitation fields (CHEF) 34 inside the ICR<br />

cell to eliminate unwanted peaks caused by impurities and<br />

byproducts from adduct forming reactions. An indirectly heated<br />

hollow dispenser cathode was used for electron generation. 35<br />

A heating current of 1.8 A was applied to a heater element<br />

located behind the cathode. For EID, performed inside the ICR<br />

cell, the cathode bias voltage was pulsed to 25-50 eV for<br />

50-500 ms. IRMPD was performed inside the ICR cell with a<br />

25 W, 10.6 µm, CO2 laser (Synrad, Mukilteo, WA). The laser<br />

beam was deflected by two mirrors for alignment through the<br />

hollow dispenser cathode to the center of the ICR cell. The<br />

beam entered the vacuum system through a BaF2 window.<br />

Photon irradiation was performed for 300-600 ms at 8.75-10<br />

W laser power. All mass spectra were acquired with XMASS<br />

software (version 6.1, Bruker Daltonics) in broadband mode<br />

from m/z 21 to 1000 with 256 K data points and summed over<br />

10-30 scans. Data processing was performed with the MIDAS<br />

analysis software. 36 Calculated masses of precursor ions, [M<br />

+ Met - H] + ([M + 2Met - H] + for Li), and one of the most<br />

abundant product ions were used for internal calibration.<br />

RESULTS<br />

Charge-Remote Fragmentation in EID of Mn(II)-Adducted<br />

Fatty Acids. Figure 1 shows EID spectra of Mn(II)-adducted<br />

arachidonic acid, linolenic acid, oleic acid, and stearic acid, where<br />

(32) Belov, M. E.; Nikolaev, E. N.; Anderson, G. A.; Udseth, H. R.; Conrads,<br />

T. P.; Veenstra, T. D.; Masselon, C. D.; Gorshkov, M. V.; Smith, R. D. Anal.<br />

Chem. 2001, 73, 253–261.<br />

(33) Hendrickson, C. L.; Quinn, J. P.; Emmett, M. R.; Marshall, A. G. 49th ASMS<br />

Conference on Mass Spectrometry and Allied Topics, Chicago, IL, 2001; CD-<br />

ROM.<br />

(34) de Koning, L. J.; Nibbering, N. M. M.; van Orden, S. L.; Laukien, F. H. Int.<br />

J. Mass Spectrom. 1997, 165, 209–219.<br />

(35) Tsybin, Y. O.; Witt, M.; Baykut, G.; Kjeldsen, F.; Hakansson, P. Rapid<br />

Commun. Mass Spectrom. 2003, 17, 1759–1768.<br />

(36) Senko, M. W.; Canterbury, J. D.; Guan, S.; Marshall, A. G. Rapid Commun.<br />

Mass Spectrom. 1996, 10, 1839–1844.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6941


Figure 1. EID of Mn(II)-adducted fatty acids: (a) arachidonic acid (d ) 4), (b) linolenic acid (d ) 3), (c) oleic acid (d ) 1), and (d) stearic acid (d )<br />

0), where d indicates the number of double bonds in each fatty acid. Only charge-remote product ion peaks, [CxHyO2 + Mn] + , are labeled. The<br />

precursor ion peaks, [M + Mn - H] + , are outside the displayed m/z range. The Y-axis is zoomed 50 or 200 times as indicated by “×50” or “×200”.<br />

6942 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010


Figure 2. Normalized product ion abundances of [CxHyO2 + Met] + vs Cn (n denotes carbon position from the carboxylate end of a fatty acid)<br />

in EID of Mn(II)-adducted fatty acids: (a) arachidonic acid (d ) 4), (b) linolenic acid (d ) 3), (c) oleic acid (d ) 1), and (d) stearic acid (d ) 0),<br />

where d indicates the number of double bonds in each fatty acid. Three observations were made for each Mn(II)-adducted fatty acid.<br />

each fatty acid has 4, 3, 1, and 0 double bonds, respectively. In<br />

this figure, only characteristic charge-remote fragments, [CxHyO2<br />

+ Mn] + , are labeled whereas unlabeled peaks are mostly due<br />

to charge-driven fragmentation, including [CxHy + Mn] + .<br />

Charge-remote product ion abundances at each carbon-carbon<br />

cleavage site were calculated by adding all of the [CxHyO2 +<br />

Met] + -type ion abundances at each carbon position and<br />

normalizing to the total [CxHyO2 + Met] + -type ion abundances<br />

in the EID spectrum. In EID of Mn(II)-adducted fatty acids,<br />

even and odd electron species were observed, generated<br />

by heterolytic (resulting in product ions of, e.g., the types<br />

[CnH2n-1O2 + Mn] + , [CnH2n-3O2 + Mn] + , and [CnH2n-5O2 +<br />

Mn] + ) and homolytic bond cleavages (producing, e.g., the<br />

product ion types [CnH2n-2O2 + Mn] + ,[CnH2n-4O2 + Mn] + , and<br />

[CnH2n-6O2 + Mn] + ), respectively. Similar behavior was noted<br />

in high energy CAD of ESI-generated precursor ions, while<br />

mostly heterolytic bond cleavages were observed in FAB-high<br />

energy CAD. 37<br />

For clarity, Supplementary Table 1 (Supporting Information)<br />

shows peak assignments for charge-remote product ions from EID<br />

of Mn(II)-adducted ararchidonic acid as well as an example (C5<br />

position) of how normalized charge-remote product ion abundances<br />

were calculated. Double bond locations can be identified<br />

from normalized product ion abundances at each carbon site,<br />

as shown in Figure 2. EID experiments for each fatty acid were<br />

(37) Cheng, C.; Pittenauer, E.; Gross, M. L. J. Am. Soc. Mass Spectrom. 1998,<br />

9, 840–844.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6943


Figure 3. Normalized product ion abundances of [CxHyO2 + Met] + vs Cn (n denotes carbon position from the carboxylate end of a fatty acid)<br />

in EID of Ni(II)- and Mg(II)-adducted arachidonic acid. The structure of arachidonic acid is shown with double bond locations.<br />

repeated three times on three different days to verify the reliability<br />

of EID as a method for double bond localization in fatty acids. As<br />

shown in Figure 2, EID spectra of Mn(II)-adducted fatty acids<br />

provided highly reproducible structural information regarding<br />

double bond positions for each fatty acid. Lower product ion<br />

abundances at the C4 position, observed in EID of Mn(II)adducted<br />

linolenic acid, oleic acid, and stearic acid, indicated<br />

that those fatty acids do not have double bonds between C5<br />

and C6. In contrast, the lower product ion abundance at C5<br />

compared to C4 was indicative of the existence of a double bond<br />

between C5 and C6 in arachidonic acid. Other double bond<br />

locations in each fatty acid could be obtained from valley<br />

positions in the graph of normalized charge-remote product<br />

ion abundances vs carbon position (Cn), as shown in Figure 2.<br />

Gas-phase ion fragmentation reactions are charge remote when<br />

there is no important interaction between the charge and the<br />

cleavage sites. However, hydrocarbon chains are flexible, and such<br />

interactions can, therefore, occur even in the presence of a<br />

terminal fixed charge, particularly at less remote sites. 15 Gross<br />

and co-workers have reported that some product ions formed by<br />

cleavage near the charge are stabilized by a cyclic conformation. 15<br />

6944 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

For example, C4-C5 bond cleavage in deprotonated palmitic acid<br />

was enhanced due to formation of a cyclic structure. 18 Similarly,<br />

enhanced cleavage at the C3 position occurred in EID of Mn(II)adducted<br />

linolenic, oleic, and stearic acid (Figure 2b-d),<br />

possibly due to ring formation.<br />

Mn(II) is expected to bind tightly with the carboxylate anion<br />

end of fatty acids because both Mn(II) and the carboxylate anion<br />

have hard Lewis acid and base properties, respectively. Binding<br />

energies of divalent metal ions and H2O/OH - increase sharply<br />

at the transition between d 0 (Ca(II)) and d n (Sc(II)) due to<br />

electron occupation in d orbitals. 38 These chemical properties<br />

may explain why Mn(II) appears to be more efficient than other<br />

divalent metals examined (see below) for “fixing” a charge at<br />

the carboxylate end of fatty acids.<br />

Charge-Remote Fragmentation in EID of Arachidonic Acid<br />

Adducted with Metals Other than Mn(II). In addition to Mn(II),<br />

which yielded successful charge remote fragmentation in EID (see<br />

above), Li(I) and other divalent metals (Zn(II), Co(II), Ni(II),<br />

Mg(II), Ca(II), and Fe(II)) were examined as adducts in EID of<br />

(38) Magnusson, E.; Moriarty, N. W. Inorg. Chem. 1996, 35, 5711–5719.


Figure 4. EID (a) and IRMPD (b) of Mn(II)-adducted arachidonic acid. ν2 and ν3 denote the second and third harmonic peaks. [CxHyO2 + Mn] +<br />

product ion peaks are labeled with “b”, and [CxHy] + -type peaks are labeled with “4”. Figure 4a shows the same spectrum as in Figure 1a, but<br />

here, all peaks are labeled for comparison with the IRMPD spectrum. Y-axis is zoomed 200 or 5 times, indicated by “×200” or “×5”.<br />

arachidonic acid. Lithium was our first metal of choice because<br />

doubly Li(I)-adducted fatty acids, [M + 2Li - H] + , have been<br />

shown to yield charge-remote fragmentation in high energy<br />

CAD. 9,10,16,18 Electron irradiation (∼26 eV electrons) of the [M<br />

+ 2Li - H] + form of arachidonic acid did not yield any product<br />

ions (Supplementary Figure 1a, Supporting Information). One<br />

explanation for this discrepancy may be the different means of<br />

electronic excitation in EID vs high energy CAD: the latter<br />

involves interactions between ions and neutrals whereas the<br />

former involves ion-electron interactions. Thus, physical parameters<br />

such as polarizibility (which is low for lithium) should be<br />

more important in EID compared to high energy CAD.<br />

EID (∼20 eV electrons) of Zn(II)- and Co(II)-adducted arachidonic<br />

acid yielded very limited charge-remote fragmentation<br />

(Supplementary Figure 1b,c, Supporting Information). In contrast,<br />

EID of Ni(II)-, Mg(II)-, Ca(II)-, and Fe(II)-adducted arachidonic<br />

acid provided extensive charge-remote product ions of the type<br />

[C xHyO2 + Met] + . (As an example, EID of Ni(II)-adducted<br />

arachidonic acid is shown in Supplementary Figure 1d, Supporting<br />

Information.)<br />

Figure 3 shows charge-remote product ion abundances as a<br />

function of alkyl chain carbon position from EID spectra of Ni(II)and<br />

Mg(II)-adducted arachidonic acid. Similar to the Mn(II) data in<br />

Figure 2, [CxHyO2 + Met] + -type product ions were used to<br />

generate this plot. EID of Ca(II)- and Fe(II)-adducted arachidonic<br />

acid provided very similar results; thus, the corresponding data<br />

are not shown. EID of these metal-adducted arachidonic acid did<br />

not provide characteristic ion abundance patterns at all double<br />

bond positions. The double bond (C14-C15) far from the carboxylate<br />

end could not be identified from EID of Ni(II)-adducted<br />

arachidonic acid (Figure 3a). Further, the double bond (C5-C6)<br />

close to the carboxylate end could not be determined from EID<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6945


of Ni(II)- nor Mg(II)-adducted arachidonic acid (Figure 3a,b). We<br />

hypothesize that Ni(II) or Mg(II) adduction may yield a “less fixed”<br />

charge compared to Mn(II) adduction, thereby failing to provide all<br />

double bond locations in EID.<br />

EID of Mn(II)-adducted arachidonic acid was compared to<br />

IRMPD of the same species. In contrast to EID (believed to occur<br />

via both electronic and vibrational activation), dissociation in<br />

IRMPD occurs solely via vibrational excitation. The mixture of<br />

mostly [C xHy + Mn] + and [CxHyO2 + Mn] + species observed<br />

in EID (Figure 4a) implies competition between charge-driven<br />

and charge-remote processes. However, IRMPD of the same<br />

precursor ions mainly yielded product ions from charge-driven<br />

fragmentation, [CxHy + Mn] + , as expected from vibrational<br />

activation (Figure 4b). In contrast, high energy CAD (known to<br />

involve electronic excitation) provides mainly charge-remote<br />

product ions. 9 The internal energy required for charge-remote<br />

fragmentation is estimated to be ∼1.4-2.9 eV for protonated fatty<br />

acids. 39 In 70 eV EI, molecular ions of small alkenes were found<br />

to be isomerized to a mixture of interconverting structures. 40 We<br />

propose that the 25-50 eV electron energies used in EID are<br />

sufficient to promote charge-remote fragmentation but not high<br />

enough to cause isomerization of Mn(II)-adducted fatty acids,<br />

which would result in double bond migration.<br />

CONCLUSION<br />

Mn(II)-adducted fatty acids were analyzed to investigate the<br />

utility of EID for determining double bond locations. Charge-<br />

(39) Deterding, L.; Gross, M. L. Org. Mass. Spectrom. 1988, 23, 169–177.<br />

(40) Borchers, F.; Levsen, K.; Schwartz, H.; Wesdemiotis, C.; Winkler, H. C.<br />

J. Am. Chem. Soc. 1977, 99, 6359–6365.<br />

6946 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

remote product ion abundances of [CxHyO2 + Mn] + -type fragments<br />

generated by EID are significantly reduced at double<br />

bond positions. Analysis of [CxHyO2 + Mn] + -type product ion<br />

abundances from EID of Mn(II)-adducted fatty acids allowed<br />

determination of all double bond positions. However, other<br />

metal adducts did not generally provide characteristic product<br />

ion abundances at all double bond locations. The resulting<br />

structural information on double bond locations for Mn(II)adducted<br />

fatty acids may be explained by dominant electronic<br />

excitation processes in EID and efficient generation of a fixed<br />

charge at the carboxylate end due to strong interaction between<br />

Mn(II) cation and carboxylate anion. EID of Mn(II)-adducted<br />

arachidonic acid was compared with IRMPD of the same<br />

species. As expected, mostly charge-driven fragmentation was<br />

observed in IRMPD whereas both charge-remote and chargedriven<br />

product ions were observed in EID. In contrast, high<br />

energy CAD is known to occur mainly via electronic excitation<br />

and results in dominant charge-remote product ions.<br />

ACKNOWLEDGMENT<br />

This work was supported by the University of Michigan and<br />

an NSF CAREER award to K.H. (CHE-05-47699).<br />

SUPPORTING INFORMATION AVAILABLE<br />

Peak assignments and calculations for Figures 1a and 2a. EID<br />

spectra of Li-, Zn-, Co-, and Ni-adducted arachidonic acid. This material<br />

is available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 8, 2010. Accepted July 14, 2010.<br />

AC101217X


Anal. Chem. 2010, 82, 6947–6957<br />

Identification of Metallothionein Subisoforms in<br />

HPLC Using Accurate Mass and Online Sequencing<br />

by Electrospray Hybrid Linear Ion Trap-Orbital Ion<br />

Trap Mass Spectrometry<br />

Sandra Mounicou,* ,† Laurent Ouerdane, † BéatriceL’Azou, ‡ Isabelle Passagne, ‡<br />

CélineOhayon-Courtès, ‡ Joanna Szpunar, † and Ryszard Lobinski †<br />

CNRS/UPPA, Laboratoire de Chimie Analytique Bio-Inorganique et Environnement, UMR 5254, 2, av. Pr. Angot,<br />

64053 Pau, France, and EA 3672 Santé-Travail-Environnement, Université Victor Segalen, 146, rue Léo Saignat,<br />

33076 Bordeaux<br />

A comprehensive approach to the characterization of<br />

metallothionein (MT) isoforms based on microbore HPLC<br />

with multimodal detection was developed. MTs were<br />

separated as Cd7 complexes, detected by ICP MS and<br />

tentatively identified by molecular mass measured with<br />

1-2 ppm accuracy using Orbital ion trap mass spectrometry.<br />

The identification was validated by accurate<br />

mass of the corresponding apo-MTs after postcolumn<br />

acidification and by their sequences acquired online<br />

by higher-energy collision dissociation MS/MS. The<br />

detection limits down to 10 fmol and 45 fmol could<br />

be obtained by ESI MS for apo- and Cd7-isoforms,<br />

respectively, and were lower than those obtained by<br />

ICP MS (100 fmol). The individual MT isoforms could<br />

be sequenced at levels as low as 200 fmol with the<br />

sequence coverage exceeding 90%. The approach was<br />

successfully applied to the identification of MT isoforms<br />

induced in a pig kidney cell line (LLC-PK1)<br />

exposed to CdS nanoparticles.<br />

Mammalian metallothioneins (MT) are low molecular weight<br />

proteins (60-62 amino acid with molecular mass of ca. 6000-7000<br />

Da) characterized by high cysteine content (up to 30% residues)<br />

enabling to bind a wide range of transition and heavy metal<br />

ions. 1-3 They are involved in a variety of biochemical processes<br />

essential for life, such as cellular growth, stress response, copper<br />

and zinc homeostasis, and detoxification of heavy metals. Hence,<br />

MTs can be valuable biomarkers of stress conditions and several<br />

pathologies which require the development of methods for the<br />

determination and identification of the individual MT isoforms and<br />

products of their post-translational modifications. 4-7 The primary<br />

* To whom correspondence should be addressed.<br />

† Laboratoire de Chimie Analytique Bio-Inorganique et Environnement.<br />

‡ Université Victor Segalen.<br />

(1) Hamer, D. H. Annu. Rev. Biochem. 1986, 55, 913–951.<br />

(2) Kagi, J. H. R. Methods Enzymol. 1991, 205, 613–626.<br />

(3) Metallothioneins; Stillman, M. J.; Shaw, C. F. I.; Suzuki, K. T., Eds.; VCh:<br />

New York, 1992.<br />

(4) Kagi, J. H. R.; Schaffer, A. Biochemistry 1988, 27, 8509–8515.<br />

(5) Cousins, R. J. Physiol. Rev. 1985, 65, 238–309.<br />

(6) Klaassen, C. D.; Liu, J.; Choudhuri, S. Ann. Rev. Pharmacol. Toxicol. 1999,<br />

39, 267–294.<br />

structure of metallothioneins exhibits modifications up to 15 amino<br />

acids leading to the occurrence of several subisoforms. 1-3<br />

Electrospray ionization (ESI) MS was first proposed by<br />

Fenselau group for in vitro studies of the complexation of metalions<br />

by MTs. 8,9 These seminal works have been followed by a<br />

large number of studies using ESI MS to study the stoichiometry<br />

of metal binding to MTs and its domains, 10-14 reactivity and<br />

kinetics of metal exchange, 15-18 and reaction of MT with metallodrugs<br />

19 which have been competently reviewed. 15,20,21 Recombinant<br />

and highly purified MTs were used in these studies because<br />

of the very low purity of the commercial MT which are mixtures<br />

of different isoforms.<br />

Electrospray MS, coupled with a high-resolution separation<br />

techniques, such as reversed-phase chromatography 22 or capillary<br />

electrophoresis, 23-25 was shown to be an attractive technique to<br />

(7) Coyle, P.; Philcox, J. C.; Carey, L. C.; Rofe, A. M. Cell. Mol. Life Sci. 2002,<br />

59, 627–647.<br />

(8) Yu, X.; Wojciechowski, M.; Fenselau, C. Anal. Chem. 1993, 65, 1355–<br />

1359.<br />

(9) Afonso, C.; Hathout, Y.; Fenselau, C. J. Mass Spectrom. 2002, 37, 755–<br />

759.<br />

(10) Gehrig, P. M.; You, C.; Dallinger, R.; Gruber, C.; Brouwer, M.; Kôgi,<br />

J. H. R.; Hunziker, P. E. Protein Sci. 2000, 9, 395–402.<br />

(11) Merrifield, M. E.; Huang, Z.; Kille, P.; Stillman, M. J. J. Inorg. Biochem.<br />

2002, 88, 153–172.<br />

(12) Ngu, T. T.; Krecisz, S.; Stillman, M. J. Biochem. Biophys. Res. Commun.<br />

2010, 396, 206–212.<br />

(13) Orihuela, R.; Domènech, J.; Bofill, R.; You, C.; Mackay, E. A.; Kägi, J. H. R.;<br />

Capdevila, M.; Atrian, S. J. Biol. Inorg. Chem. 2008, 13, 801–812.<br />

(14) Palumaa, P.; Eriste, E.; Kruusel, K.; Kangur, L.; Jörnvall, H.; Sillard, R. Cell.<br />

Mol. Biol. (Noisy-le-Grand, Fr.) 2003, 49, 763–768.<br />

(15) Duncan, K. E. R.; Ngu, T. T.; Chan, J.; Salgado, M. T.; Merrifield, M. E.;<br />

Stillman, M. J. Exper. Biol. Med. 2006, 231, 1488–1499.<br />

(16) Palumaa, P.; Eriste, E.; Njunkova, O.; Pokras, L.; Jörnvall, H.; Sillard, R.<br />

Biochemistry 2002, 41, 6158–6163.<br />

(17) Vaher, M.; Romero-Isart, N.; Vasak, M.; Palumaa, P. J. Inorg. Biochem.<br />

2001, 83, 1–6.<br />

(18) Zeitoun-Ghandour, S.; Charnock, J. M.; Hodson, M. E.; Leszczyszyn, O. I.;<br />

Blindauer, C. A.; Stürzenbaum, S. R. FEBS J. 2010, 277, 2531–2542.<br />

(19) Karotki, A. V.; Vasak, M. J. Biol. Inorg. Chem. 2009, 14, 1129–1138.<br />

(20) Chan, J.; Huang, Z.; Merrifield, M. E.; Salgado, M. T.; Stillman, M. J. Coord.<br />

Chem. Rev. 2002, 233-234, 319–339.<br />

(21) Ngu, T. T.; Stillman, M. J. Dalton Trans. 2009, 5425–5433.<br />

(22) Chassaigne, H.; Lobinski, R. J. Chromatogr., A 1998, 829, 127–136.<br />

(23) Mounicou, S.; Polec, K.; Chassaigne, H.; Potin-Gautier, M.; Lobinski, R. J.<br />

Anal. At. Spectrom. 2000, 15, 635–642.<br />

(24) Andon, B.; Barbosa, J.; Sanz-Nebot, V. Electrophoresis 2006, 27, 3661–<br />

3670.<br />

10.1021/ac101245h © 2010 American <strong>Chemical</strong> Society 6947<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/29/2010


separate and identify the individual isoforms but the hitherto<br />

studies have been plagues by poor sensitivity preventing the direct<br />

analysis of biological cytosols. Indeed, most of the reports referred<br />

to the analysis of a1mg· mL -1 solution by capillary electrophoresis<br />

and 100 µg · mL -1 by HPLC. Such high MT concentrations<br />

are largely superior to those encountered in real-world<br />

biological systems and can be achieved only after a large-scale<br />

purification. Moreover, the ESI MS sensitivity is severely<br />

degraded by the coelution of other biomolecules, especially in<br />

ESI TOF MS. Consequently, the most widely applied MS<br />

technique for the screening of biological cytosols for MTs has<br />

been inductively coupled plasma mass spectrometry (ICP MS)<br />

used in combination with HPLC or capillary electrophoresis. 26,27<br />

ICP MS is a convenient technique for the detection of the<br />

individual isoforms and the determination of the stoichiometry of<br />

the metal-complexes formed, indeed, but does not allow their<br />

identification.<br />

The few successful examples of the identification of MTs in<br />

biological cytosols by ESI MS were carried out on the basis of<br />

the molecular mass determined using quadrupole 28,29 and TOF<br />

mass spectrometers 30 with limited mass accuracy. The acquisition<br />

of sequence information by conventional bottom-up proteomics<br />

approaches is hampered by resistance of MT to tryptic digestion<br />

and frequent miscleavages in the presence of residual metals. 31<br />

Also, as MT isoforms exhibit a significant sequence homology<br />

(70-90%), 32 many tryptic peptides are common for many isoforms<br />

thus preventing de novo identification. Indeed, examples of<br />

successful identification of MTs on the basis of MS/MS of tryptic<br />

peptides have been scarce. 30,33<br />

The above drawbacks and the tediousness of off-line bottomup<br />

identification procedures can be alleviated by top-down MS<br />

allowing one to obtain structural information from intact proteins. 34<br />

Although most of data have been obtained with high magnetic<br />

field strength FT ICR MS using infrared multiple photon dissociation<br />

(IRMPD) or electron capture detection (ECD), 35 the efficiency<br />

of direct fragmentation of intact protein in quadrupole collision<br />

cells is increasingly explored. 36,37 Particularly interesting is the<br />

combination of hybrid linear quadrupole ion trap with an Orbitrap<br />

mass spectrometer which offers resolution exceeding 60 000 and<br />

often sub-ppm mass accuracy. 38-42<br />

(25) Benavente, F.; Andon, B.; Gimenez, E.; Olivieri, A. C.; Barbosa, J.; Sanz-<br />

Nebot, V. Electrophoresis 2008, 29, 4355–4367.<br />

(26) Prange, A.; Schaumloffel, D. Anal. Bioanal. Chem. 2002, 373, 441–453.<br />

(27) Szpunar, J. Analyst 2005, 130, 442–465.<br />

(28) Infante, H. G.; Cuyckens, F.; Van Campenhout, K.; Blust, R.; Claeys, M.;<br />

Van Vaeck, L.; Adams, F. C. J. Anal. At. Spectrom. 2004, 19, 159–166.<br />

(29) Polec, K.; Perez-Calvo, M.; Garcia-Arribas, O.; Szpunar, J.; Ribas-Ozonas,<br />

B.; Lobinski, R. J. Inorg. Biochem. 2002, 88, 197–206.<br />

(30) Wang, R.; Sens, D. A.; Albrecht, A.; Garrett, S.; Somji, S.; Sens, M. A.; Lu,<br />

X. Anal. Chem. 2007, 79, 4433–4441.<br />

(31) Wang, R.; Sens, D. A.; Garrett, S.; Somjii, S.; Sens, M. A.; Lu, X.<br />

Electrophoresis 2007, 28, 2942–2952.<br />

(32) Sewell, A. K.; Yokoya, F.; Yu, W.; Miyagawa, T.; Murayama, T.; Winge,<br />

D. R. J. Biol. Chem. 1995, 270, 25079–25086.<br />

(33) Feng, W.; Benz, F. W.; Cai, J.; Pierce, W. M.; Kang, Y. J. J. Biol. Chem.<br />

2006, 281, 681–687.<br />

(34) Kelleher, N. L.; Lin, H. Y.; Valaskovic, G. A.; Aaserud, D. J.; Fridriksson,<br />

E. K.; McLafferty, F. W. J. Am. Chem. Soc. 1999, 121, 806–812.<br />

(35) Tolmachev, A. V.; Robinson, E. W.; Wu, S.; Pasa-Tolic, L.; Smith, R. D. Int.<br />

J. Mass Spectrom. 2009, 287, 32–38.<br />

(36) Mandal, R.; Li, X. F. Rapid Commun. Mass Spectrom. 2006, 20, 48–52.<br />

(37) Moreno-Gordaliza, E.; Canas, B.; Palacios, M. A.; Gomez-Gomez, M. M.<br />

Anal. Chem. 2009, 81, 3507–3516.<br />

6948 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

The goal of this study was the description on the molecular<br />

level of the chemical response of a cell line exposed to the stress<br />

of CdS nanoparticles, increasingly used in diverse areas from<br />

electronics to targeted drug delivery, in view of the understanding<br />

of the mechanisms of their toxicity. As the preliminary experiments<br />

indicated MT induction at picomole levels, analytical<br />

development focused on the detection of traces of metal-complexes<br />

with individual MT isoforms and the unambiguous identification<br />

of the latter using both the accurate mass and topdown sequencing<br />

information, directly in the MT fraction, without extensive<br />

purification.<br />

EXPERIMENTAL SECTION<br />

Reagent and Standards. <strong>Analytical</strong> reagent grade chemicals<br />

purchased from Sigma-Aldrich (Saint-Quentin Fallavier, France)<br />

and water (18 MΩ cm) obtained from a Milli-Q system (Millipore,<br />

Bedford, MA) were used throughout unless stated otherwise. The<br />

rabbit liver metallothionein-2 isoform standard (purity: 95%) was<br />

purchased from Enzo Life Sciences (Villeurbanne, France). It was<br />

reported by the manufacturer to be a mixture of isoforms (major:<br />

MT-2a, minor: MT-2b and MT-2c) and to contain 67% of protein<br />

and 9% of metals (Cd and Zn). The stock solution (1 mg · mL -1 )<br />

was prepared by dissolving 1 mg of metallothionein in 1 mL of<br />

water, subdivided in 10 µL aliquots to avoid multiple thawing<br />

and freezing, and frozen at -20 °C. Working solutions were<br />

prepared daily by dilution with water at 4 °C.<br />

Instrumentation. Microbore reversed-phase HPLC (µRP<br />

HPLC) separations were carried out using an Agilent 1100<br />

capillary HPLC system (Agilent, Tokyo, Japan) equipped with a<br />

100 µL · min -1 splitter module. ICP MS detection was achieved<br />

using a model 7500cs instrument (Agilent) fitted with platinum<br />

cones, 1 mm i.d injector torch and a T-connector allowing the<br />

introduction of 5% O2. The µRP HPLC-ICP MS coupling was<br />

done via an Isomist interface (Glass Expansion, Melbourne,<br />

Vic, Australia) consisting of a 20 mL model Cinnabar spray<br />

chamber cooled at 2 °C and fitted with a 50 µL · min -1<br />

Micromist nebulizer.<br />

For µRP HPLC-ESI MS experiments, the µRP HPLC system<br />

was connected to a LTQ Orbitrap Velos mass spectrometer<br />

(ThermoFisher Scientific, Bremen, Germany). The coupling was<br />

achieved via a heated electrospray ionization source (H-ESI II)<br />

(ThermoFisher Scientific). The postcolumn acidification manifold,<br />

described in detail elsewhere, 43 consisted of a zero dead volume<br />

PEEK T-piece allowing the mixing of the chromatographic effluent<br />

with a formic acid:MeOH (30/70%, v/v) solution delivered by<br />

means of a syringe pump (Pump 33 model, Harvard Apparatus,<br />

South Natick, MA) and a mixing PEEK tubing coil (250 µm i.d. ×<br />

250 mm) connected to the inlet of the electrospray ion source.<br />

Only PEEK tubing and connectors were used to avoid metal<br />

contamination.<br />

(38) Bondarenko, P. V.; Second, T. P.; Zabrouskov, V.; Makarov, A. A.; Zhang,<br />

Z. J. Am. Soc. Mass Spectrom. 2009, 20, 1415–1424.<br />

(39) Scigelova, M.; Makarov, A. Proteomics 2006, 1, 16–21.<br />

(40) Macek, B.; Waanders, L. F.; Olsen, J. V.; Mann, M. Mol. Cell. Proteomics<br />

2006, 5, 949–958.<br />

(41) Wynne, C.; Fenselau, C.; Demirev, P. A.; Edwards, N. Anal. Chem. 2009,<br />

81, 9633–9642.<br />

(42) Erales, J.; Gontero, B.; Whitelegge, J.; Halgand, F. Biochem. J. 2009, 419,<br />

75–82.<br />

(43) Chassaigne, H.; ?obin?ski, R. J. Chromatogr., A 1998, 829, 127–136.


Figure 1. Schematic flowchart showing the type of information obtained in the different modes of the multimodal approach developed.<br />

The cell-line cytosol was separated by ultracentrifugation using<br />

a HimaCs 120GX model (Hitachi, Tokyo, Japan) and fractionated<br />

by size-exclusion chromatography using an Agilent 1100 HPLC<br />

system (Agilent, Wilmington, DE). The metal elution was monitored<br />

by splitting part of the effluent to an Agilent 7500ce ICP<br />

MS (Agilent, Tokyo, Japan).<br />

Procedures. The purpose of the different detection modes<br />

used in this study was schematically illustrated in Figure 1.<br />

µRP HPLC. The column was a C8 Vydac (250 mm ×1 mm i.d.,<br />

5 µm) presented in passivated 316 copper-free stainless steel<br />

housing (Alltech/Grace, Templemars, France). The injection<br />

volume was 5 µL. The elution solvents were eluent A: 5 mM<br />

ammonium acetate pH 6 and eluent B: 5 mM ammonium acetate<br />

(pH 6) in 50% (v/v) acetonitrile. Elution was carried out at 40<br />

µL · min -1 using the following program: 0-50 min: 2-20% B;<br />

50-52 min: 2% B; 52-60 min: 2% B.<br />

µRP HPLC-ICP MS. ICP MS was optimized daily for the<br />

maximum sensitivity by introducing directly a solution containing<br />

1 µg · L -1 89 Y, 7 Li, 205 Tl and 140 Ce. Isotopes monitored in µRP<br />

HPLC-ICP MS were 114 Cd, 112 Cd 63 Cu, 65 Cu, 64 Zn, 66 Zn. The<br />

data were processed using Excel Microsoft software.<br />

µRP HPLC-ESI MS. Chromatographic separation conditions<br />

were identical as given above. For the postcolumn acidification<br />

experiments, the solution added was formic acid:methanol (30/<br />

70%, v/v) delivered at 4 µL · min -1 .<br />

Initial calibration of the mass spectrometer was performed<br />

using a mixture consisting of cafein (195.08765), MRFA peptide<br />

(524.26499) and Ultramark polymer (m/z 1221.99063). The ion<br />

source was operated in the positive ion mode at 3.2 kV. The<br />

vaporizer temperature of the source was set to 120 °C and the<br />

capillary temperature to 280 °C. Nitrogen sheath gas was set to<br />

20, the auxiliary gas to 5 and sweep gas to 0 (arbitrary unit). The<br />

ion lenses were automatically optimized using a MT solution (1<br />

µg · mL -1 in 0.01% formic acid in 50% (v/v) methanol) introduced<br />

by infusion at 4 µL · min -1 and monitored at m/z 1225.843 (z )<br />

5). In all experiments, the most abundant mass of [M + 5H] 5+<br />

ion was systematically monitored for better detection limit.<br />

Throughout this manuscript, the most abundant mass is always<br />

given for multicharged ions or uncharged molecules.<br />

In full scan mode, an m/z 1200-1500 range was scanned for<br />

the detection of apo- and metalated MTs (the resolution was set<br />

at 100 000 (m/∆m, fwhm at m/z 400)). Injection time was 400<br />

ms. In single ion monitoring (SIM) mode, the scan mode was<br />

centered on m/z 1230 ± 50 and m/z 1377.5 ± 65 for the detection<br />

of apo- and metalated rabbit liver MTs, respectively. Pig kidney<br />

apo-MTs were measured at m/z 1250 ± 60 and then at m/z 1170<br />

± 70 whereas the metalated forms at m/z 1367.5 ± 87. The<br />

resolution was set at 100 000 (m/∆m, fwhm at m/z 400) and<br />

injection time was 3 s. The mass accuracy was expressed in ppm<br />

and calculated as the ratio (M theoretical - Mdetermined)/Mtheoretical<br />

multiplied by 10 6 .<br />

µRP HPLC-ESI MS/MS. The [M+5H] 5+ of apo-MT was<br />

selected for top-down sequencing experiments. High energy<br />

collision fragmentation (HCD) was carried out at 50% energy,<br />

using an isolation width of 2 in full scan mode over m/z<br />

100-2000 mass range at a resolution of 100 000. The acquisition<br />

time (60 min) was split into 6 time segments. Segment 1:<br />

0-30.75 min, SIM scan mode (m/z 1100-1300) at 100 000<br />

resolution (m/∆m, fwhm at m/z 400); segment 2: 30.75-36<br />

min, HCD at m/z 1189.4: segment 3: 36-39 min, HCD at m/z<br />

1197.63, and m/z 1194.63; segment 4: 39-44 min, HCD at m/z<br />

1203.24; segment 5: 44-47 min, HCD at m/z 1271.96; segment<br />

6: 47-60 min, SIM scan mode centered at m/z 1200 ± 100 at<br />

100 000 resolution. Data were processed using Xcalibur 2.1<br />

software (Thermo Fisher Scientific) and masses were manually<br />

corrected using the traces of peaks of the Ultramark polymer<br />

calibrant as internal standard. Analyst Q.S. 1.1 software (Applied<br />

Biosystems MDS Sciex, Foster City, CA) was used to obtain<br />

the theoretical y ions lists of the amino acids sequence of MT<br />

subisoforms.<br />

Cell Growth and CdS Nanoparticles Exposure. Kidney pig cell<br />

line (LLC-PK1) were grown in 100 mm cell culture Petri dish<br />

with EMEM (Eagle’s minimal essential medium) media containing<br />

1% antibiotics (penicillin, streptomycin), 2 mM Lglutamin,<br />

1 mM 4-(2-hydroxyethyl)-1-piperazine-ethanesulfonic<br />

acid (HEPES), non-essential amino acids, and 5% of fetal calf<br />

serum. Petri dishes were incubated at 37 °C ina5%CO2<br />

incubator. At subconfluence, cells were exposed during 24 h<br />

with5mLof120µM solution of 10 nm-CdS nanoparticles. Cells<br />

from 11 Petri dishes were pooled. Cells not submitted to CdS<br />

were considered as control.<br />

Recovery of the MT Fraction. After cell growth, the supernatant<br />

was discarded and cells were rinsed twice with 1 mL phosphate<br />

buffered saline (PBS). A 1-mL aliquot of 25 mM Tris/HCl buffer<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6949


(pH 7.2) containing 1 mM dithiothreitol (DTT) and 0.1 mM<br />

phenylmethyl-sulfonylfluoride (PMSF) was added. Cells were<br />

scrapping off and recovered into an eppendorf tube. Another 500<br />

µL of buffer was added to recover the maximum number of cells.<br />

Cellular lysis was carried out by a successive freeze-thaw<br />

procedure by immersing the tube into liquid nitrogen for 3 min<br />

followed by 3 min in a water bath at 37 °C (all steps repeated<br />

three times). An aliquot of 50 µL was kept aside to quantify<br />

proteins amount by the Bradford protein assay. A 1.45 µL PMSF<br />

aliquot was added to the remaining 1.45 mL sample to reach a<br />

final concentration of 0.2 mM. Then, the sample was ultracentrifuged<br />

(4 °C, 120 000 g, 20 min). The recovered supernatant was<br />

heated at 95 °C for 5 min and ultracentrifuged (4 °C, 120 000 g,<br />

20 min). A 100 µL aliquot of the supernatant was fractionated using<br />

a Superdex peptide HR 10/30 (GE Healthcare, Uppsala, Sweden)<br />

size exclusion column (10 × 300 mm × 5 µm) and isocratic elution<br />

at 0.7 mL · min -1 50 mM TRIS HCl pH 7.4. The SEC column<br />

was cleaned beforehand by flushing with mobile phase containing5mM�-mercaptoethanol<br />

and 2 mM EDTA in order to<br />

remove adsorbed metal ions. The main Cd-containing fraction<br />

(8.75 - 10.73 mL) was collected from nine injections, freezedried,<br />

resolubilized in 900 µL of water and desalted using a 5<br />

mL HiTrap column (GE Healthcare, Uppsala, Sweden) by<br />

elution with NH3aq (pH 8.0) at 1.5 mL · min -1 . The Cd-containing<br />

fraction was heartcut, freeze-dried, and resolubilized in 25 µL<br />

prior to µHPLC analyses. No special precautions, for example,<br />

by Ar or N2 saturation during the sample preparation, were<br />

taken to reduce oxidation of MTs.<br />

RESULTS AND DISCUSSION<br />

ESI MS Analysis of Rabbit Liver MT-2 Standard. As poor<br />

detection limits are the principal drawback of the existing methods<br />

for MT detection, effort has been focus on the optimization of<br />

the signal-to-noise ratio. The choice of the scan mode between<br />

full scan and single ion monitoring (SIM) was first investigated.<br />

In the infusion conditions proper for the ionization of the Cd7-<br />

MT form (5 mM AcNH4, pH6in50%(v/v) MeOH) the S/N<br />

ratio of the most intense peak at m/z 1380.691 (δ 0.3 ppm,<br />

[M+H + ] 5+ ) was 10 times higher in the SIM mode (centered<br />

at m/z 1367.5 ± 85) than in the full scan (m/z 1325-1410)<br />

mode. Similar results were observed for the corresponding apo-<br />

MT signal (m/z 1225.848 (δ 0.3 ppm) in 0.01% FA in 50% (v/v)<br />

MeOH, pH 1.6) for which the S/N ratio was 7 times higher in<br />

the SIM mode (centered at m/z 1230 ± 13) than in the full<br />

scan mode (m/z 1215-1250). Note that in full scan the injection<br />

time was decreased from 3000 to 400 ms in order to limit the<br />

number of concomitant ions entering the ion trap but the MT<br />

peak S/N ratio was not affected.<br />

In order to evaluate the detection limits in the chromatographic<br />

mode taking into account the abundance of the individual<br />

isoforms, the purity of the MT standard available was investigated<br />

by µHPLC-ICP MS. The chromatogram (Figure 2a) obtained in<br />

the conditions of full stoichiometric metalation (seven Cd atoms)<br />

shows two major peaks ( 114 Cd). The contribution of Zn and Cu<br />

is insignificant (


difference,<br />

ppm compound<br />

Table 1. Identification of Rabbit Liver MT2 Subisoforms<br />

apo MTs complexed MTs<br />

form time, m/z, z ) 5, most abundant most abundant difference,<br />

amino acid<br />

m/z, z ) 5, most abundant most abundant<br />

N° min measured mass, measured mass, theoretical ppm composition compound compound measured mass, measured mass, theoretical<br />

1 34.3 1230.049 6145.209 6145.212 0.5 Ac-M1D3P2N1C20S9A6T4G4K8E1Q1I1 N Ac-MT-2b 1384.894 6919.435 6919.427 -1.2 Cd7 N Ac-MT2b<br />

1375.300 6871.464 6871.450 -2.0 Cd6Zn N Ac-MT2b<br />

1375.100 6870.464 6870.452 -1.7 Cd6Cu N Ac-MT2b<br />

1365.506 6822.494 6822.476 -2.6 Cd5ZnCu N Ac-MT2b<br />

1365.906 6824.494 6824.474 -2.9 Cd5Zn2 N Ac-MT2b<br />

1365.306 6821.494 6821.478 -2.3 Cd5Cu2 N Ac-MT2b<br />

2 36.45 1217.4455 6082.1911 6082.1909 -0.04 M1D3P3N2C20S8A9T3G4K7Q1I1 MT-2a 1372.290 6856.414 6856.406 -1.1 Cd 7 -MT2a<br />

1362.697 6808.449 6808.429 -2.9 Cd6Zn MT2a<br />

3 36.7 1223.446 6112.194 6112.201 1.2 M1D3P3N2C20S8A8T4G4K7Q1I1 MT-2c / / / / /<br />

4 38.5 1243.862 6214.274 6214.281 1.2 Ac-M1D3P2N1C20S9A7T3G3K8E1Q1I1R1 N Ac-MT-2d 1398.707 6988.499 6988.486 -1.8 Cd7 N Ac-MT2d<br />

1389.113 6940.529 6940.520 -1.2 Cd6Zn N Ac-MT2d<br />

5 38.9 1231.849 6154.209 6154.212 0.5 Ac-M1D3P3N2C20S8A8T4G4K7Q1I1 N Ac-MT-2c 1386.694 6928.434 6928.427 -1.0 Cd7 N Ac-MT2c<br />

1376.900 6879.464 6879.452 -1.7 Cd6Cu N Ac-MT2c<br />

1377.101 6880.469 6880.451 -2.6 Cd6Zn N Ac-MT2c<br />

6 39.3 1225.848 6124.204 6124.201 -0.4 Ac-M1D3P3N2C20S8A9T3G4K7Q1I1 N Ac-MT-2a 1380.693 6898.429 6898.416 -1.8 Cd 7 N Ac-MT2a<br />

1371.299 6851.459 6851.440 -2.7 Cd6Zn N Ac-MT2a<br />

1361.305 6801.489 6801.466 -3.3 Cd5ZnCu N Ac-MT2a<br />

1361.705 6803.489 6803.464 -3.6 Cd5Zn2 N Ac-MT2a<br />

1361.104 6800.484 6800.467 -2.4 Cd5Cu2 N Ac-MT2a<br />

7 42.3 1249.072 6240.324 6240.333 1.5 Ac-M 1D 3P 2N 1C 20S 8A 7T 3G 3K 8Q 1I 1E 1R 1L 1 N Ac-MT-2e 1403.917 7014.549 7014.548 -0.1 Cd 7 N Ac-MT2e<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6951


Figure 3. Comparison of CID and HCD fragmentation modes for the acquisition of sequence information for rabbit liver MT-2a isoform (6.7<br />

µg · mL -1 in 0.01% formic acid, MeOH 50%) directly introduced at 4 µL · min -1 . CID: 40% energy; HCD: 50% energy. Mass range scanned: m/z<br />

335-2000; activation time: 30 ms; injection time: 4000 ms; resolution: 100 000.<br />

smaller than that of 1 order of magnitude observed elsewhere 44<br />

possibly because of the use of a more efficient heated ionization<br />

source.<br />

Mass spectra of the individual MT isoforms [M+H + ] 5+<br />

contributing to the TICs are given in Figure 1-ESI (Supporting<br />

Information). The data are summarized in Table 1. Five MT-2<br />

subisoforms (a-e) could be tentatively identified as N-acetylated<br />

species. Non-acetylated subisoforms MT-2a and MT-2c could also<br />

be detected as minor species. The identification was based on<br />

the accurate mass determination which could be achieved for the<br />

apo-forms with sub-ppm accuracy. The exception was two isoforms<br />

(4 and 7) which were not detected in the ICP MS chromatogram<br />

and for which the measured mass accuracy was 1.5 and 1.2 ppm,<br />

respectively. The achieved mass accuracy allowed the online<br />

determination of the amino acid composition using only few pmole<br />

of MT. The subtraction of the molecular masses of the apo forms<br />

from the corresponding metalated isoforms, the latter determined<br />

with the mass accuracies between 0.1 and 3.7 ppm, allowed the<br />

identification of 20 metal complexes with different rabbit liver<br />

isoforms (Table 1). The overall data confirms the low purity of<br />

the commercial rabbit liver MT-2 preventing its use for most<br />

metalation studies. Note that all the MT-complexes detected could<br />

be efficiently converted to the apo-form by postcolumn acidification<br />

which demonstrates the efficiency of the manifold employed.<br />

(44) Polec, K.; Mounicou, S.; Chassaigne, H.; Lobinski, R. Cell. Mol. Biol. 2000,<br />

46, 221–235.<br />

6952 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

ESI MS Detection Limits. In order to assess the detection<br />

limit of the method an amount of 4.5 ng (ca. 750 fmol, 9-times<br />

less than in Figure 1) of MT-2 standard was analyzed. The five<br />

predominant MT subisoforms could be detected with S/N ratios<br />

comprised between 3 and 66 depending on the MT-isoform<br />

abundance. No significant degradation of the mass accuracy was<br />

observed as metalated isoforms could be measured with mass<br />

accuracies below 3 ppm. Taking the N-Ac-MT2b as example (the<br />

only pure isoform in Figure 1a accounting for ca. 300 fmol of MT)<br />

which was detected with a S/N ratio of 100, the detection limit<br />

(MT concentration producing the signal superior to 3 times<br />

standard deviation of the number of counts on a given mass in<br />

the blank chromatogram) of the apo-isoform could be estimated<br />

as 9 fmol and that of the metalated isoforms for 45 fmol. These<br />

detection limits are by far the lowest ever reported for MT analysis<br />

(500-fold lower than with a former-generation triple quad 44 or with<br />

a recent TOF MS. 28 The HPLC- ICP MS detection limit could be<br />

estimated, for the N-Ac-MT2b subisoform, as 100 fmol of protein<br />

(S/N ) 120 for 4.5 pmol) and was higher than that of ESI MS.<br />

Note that the ESI MS data discussed above were obtained in<br />

the SIM scan mode. When repeated in the full scan mode, the<br />

S/N ratio was found to decrease 2-7 times depending on the<br />

isoform. Mass accuracies obtained for both scan modes were<br />

essentially similar.<br />

Online Sequencing of MT Isoforms. Although the accurate<br />

mass is a valuable parameter for the identification of the MT


isoforms, its successful use, in the absence of other information,<br />

is limited to well-described systems, such as, for example, rabbit<br />

liver. For less characterized systems, especially if genetic information<br />

is not available, the accurate molecular mass information<br />

needs to be completed by sequence information. Consequently,<br />

it was further attempted to investigate the possibility of top-down<br />

sequencing of MT subisoforms eluted in conditions of Figure 2.<br />

Preliminary experiments were carried out in the infusion mode<br />

and consisted in the study of collision induced dissociation (CID)<br />

and higher energy collision dissociation (HCD) modes for the<br />

optimization of the fragmentation of MT-isoforms. The m/z<br />

1225.84 ion corresponding to the +5 charged state of apo-MT2a<br />

was selected as an example.<br />

Using CID, 40% of maximum available energy was sufficient<br />

to fragment the parent ion; an increase in energy did not change<br />

the pattern of the product ion mass spectra. When HCD was<br />

investigated, the energy had to be set at 50% to obtain significant<br />

fragmentation which remained virtually unchanged with a further<br />

increase in energy. In both modes product ions with charge states<br />

from +1 to+5 were generated. Figure 3 shows that a larger<br />

proportion of smaller m/z fragments was obtained in the HCD<br />

mode than in CID mode. More y ions with +3 and +5 charge<br />

states were observed in CID than in HCD; the opposite was true<br />

for y ions with +2 and +1 charge states. y ions with +3 charge<br />

states generated by CID covered the largest part of the sequence<br />

(42%), followed by +4 charge states (34%), +2 charge states (27%),<br />

+1 charge states (21%) and +5 charge states (12.9%). In HCD<br />

fragmentation, the proportion of +3 and +2 charged ions was<br />

inversed. Indeed, +2 charge states covered the largest part of<br />

the sequence (40%), followed by +4 ions (35.5%), +3 ions (30.6%),<br />

+1 ions (27.4%) and +5 charged ions (3.2%). Fragments obtained<br />

using CID and HCD covered 90.3% and 93.5% of the protein<br />

sequence, respectively, with mass accuracies in the 1-2 ppm<br />

range regardless of the charge state. Some y ions (three in each<br />

fragmentation mode) suffered from >3 ppm mass accuracy due<br />

to peak overlap, but in such cases a non-overlapped differently<br />

charged ion could always be found. High energy collision<br />

dissociation (HCD), providing slightly better protein sequence<br />

coverage, was chosen for further studies; in case of a > 90%<br />

sequence coverage was insufficient, it could be complemented<br />

by the CID mode. Using this procedure, the identity assignment<br />

of the rabbit liver MTs in Table 1 could be confirmed. Taking<br />

into account the MT2 concentration (ca. 6.7 µg · mL -1 ) used for<br />

the acquisition of the HCD mass spectrum and the fragments<br />

of lowest abundance, the minimum concentration required to<br />

obtain a sequence coverage superior to 90% could be assessed<br />

to as 2-3 µg · mL -1 .<br />

µHPLC-ICP MS and ESI MS Detection of MT Subisoforms<br />

in Pig Kidney Cell Line Exposed to CdS Nanoparticles. The<br />

developed approach was applied to the characterization of MTs<br />

expressed in a pig kidney cell line exposed to CdS nanoparticles.<br />

A reversed-phase chromatogram with ICP MS 114 Cd detection<br />

of the MT fraction (ca. 10 pM of MT) showed three fairly<br />

intense peaks accompanied by a number of minor ones (Figure<br />

4a). Many of these peaks had their equivalents in the chromatogram<br />

obtained with 63 Cu detection (Figure 4b) which indicates<br />

the existence of mixed metal complexes. The Cu/Cd ratio was<br />

below 10% in all the peaks except peak 8. The presence of Cu-<br />

Figure 4. µHPLC analysis of pig kidney cell line by (a) ICP MS<br />

detection of 114 Cd, (b) ICP MS detection of 66 Zn and 63 Cu, (c) TIC of<br />

ESI MS detection in conditions preserving metalation, (d) TIC of ESI<br />

MS detection with postcolumn acidification. The peak numbers<br />

correspond to data in Table 2.<br />

binding MTs isoforms is not surprising as Cu was present in the<br />

medium used for cell culture and exposure to CdS nanoparticles.<br />

No peaks were observed on the 64 Zn channel. The TIC ESI MS<br />

chromatograms of the metalated MTs (Figure 4c) and apo-MTs<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6953


Table 2. Amino Acid and Elemental Composition of Pig Kidney Metallothioneins<br />

MTs (under demetalation conditions) complexed MTs (under metalation conditions)<br />

mass difference<br />

(demetalatedcomplexed)<br />

time, m/z, z ) 5, most abundant most abundant difference,<br />

amino acid<br />

m/z, z ) 5, most abundant most abundant difference,<br />

min measured mass, measured mass, theoretical ppm<br />

composition<br />

measured mass, measured mass, theoretical ppm<br />

1 33.8 1189.032 5940.124 5940.128 0.7 M1D2P3N1C20S9A7T3G6K6R1Q1I1 1343.877 6714.349 6714.342 -1.0 7Cd - 14H<br />

2 35 1189.837 5944.149 5944.159 1.7 M1D2P2N1C20S10A6T3G6K7R1V2 1344.685 6718.389 6718.374 -2.2 7Cd - 14H<br />

3 37.4 1197.437 5982.149 5982.138 -1.8 Ac-M1D2P3N1C20S9A7T3G6K6R1Q1I1 1352.280 6756.364 6756.353 -1.6 7Cd - 14H<br />

4 38.1 1194.840 5969.164 5969.163a -0.1 M1D2P3N1C20S9A7T3G6K5R2Q1I1 1349.482 6742.374 6742.377 0.5 7Cd - 14H<br />

M1D3P2 N1C20S8A7T2G6K8Q1I1V1 5 38.5 1198.243 5986.179 5986.170 -1.4 Ac-M1D2P2N1C20S10A6T3G6K7R1V2 1353.086 6760.394 6760.384 -1.4 7Cd - 14H<br />

6 40.5 1203.246 6011.194 6011.190 -0.6 Ac-M1D3P2N1C20S8A7T2G6K8Q1I1V1 1358.090 6785.414 6785.405 -1.3 7Cd - 14H<br />

7 41 1203.036 6010.1436 6010.1440 0.1 Ac-M1D2P3N1C20S9A7T3G6K5R2Q1I1 1357.884 6784.384 6784.359 -3.6 7Cd - 14H<br />

8 44.3 1272.167 6355.799 6355.792b -1.0 Ac-M1D3P2N1C20S8A7T2G6K8Q1I1V1 1360.478 6797.354 6797.344 -1.4 4Cd - 8H<br />

form<br />

N°<br />

6954 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

a Theoretical mass for a mixture of MT1a/MT2a (1/1.65). b Mass of N-AcMT2a after thiols oxidation (formation of one disulfide bridge) + 2Cd(II) + 2Cu(I) - 6H.<br />

(Figure 4d) showed a number of fairly intense quite well resolved<br />

peaks in the 33-45 min range which corresponded to the peaks<br />

in the ICP MS chromatograms.<br />

The mass spectrometric data allowed spectra of MT-like ions<br />

([M+5H] 5+ ) to be detected at retention times corresponding<br />

to the ICP MS peaks as demonstrated in Figure 5. In seven<br />

cases a mass spectrum of the apo-isoform and a corresponding<br />

mass spectrum of the Cd7-metalated isoform could be found<br />

resulting in a XIC with a single intense peak. In the case of<br />

peak 8 (Figure 4b), no apo-isoform could be detected under acidic<br />

conditions; the isotopic pattern of the peak revealed the presence<br />

of metals. The elemental composition calculated on the basis of<br />

the accurate molecular mass indicated the presence of two atoms<br />

of Cd and two atoms of Cu. The mass spectrometric data are<br />

shown in Table 2. As it can be seen the mass accuracies fitting a<br />

tentative amino acid composition are usually


Figure 5. Extracted ion current ESI MS chromatograms of MTs subisoforms of pig kidney cell line with mass spectra in insets (upper: metalated<br />

subisoform; lower: subisoform under acidic conditions) at retention time (a) 33.8 min, (b) 35 min, (c) 37.4 min, (d) 38.1 min, (e) 38.5 min, (f) 40.5<br />

min, (g) 41 min, (h) 44.3 min.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6955


Figure 6. List of ions used for the determination of pig kidney MT sequences. (a) MT-1g, (b) MT-1h, (c) N-Ac-MT-1g, (d) MT-1a and MT-2a,<br />

(e) N-Ac-MT-1h, (f) N-Ac-MT-2a, (g) N-Ac-MT-1a, (h) N-Ac-MT-2a. Bold characters highlight residues where modifications are observed.<br />

no mass fragment specific to MT-1e in MS/MS spectra, its<br />

presence or absence cannot be unambiguously confirmed. Even<br />

if present, its contribution to the MTs pool would be very low. By<br />

selecting appropriate and narrow retention time ranges in the MS/<br />

MS chromatogram, y ions detected covered 74% and 87% of the<br />

MT-1a and MT-2a amino acid sequence (Figure 6d) confirming<br />

the identification of Cd7MT-1a and Cd7MT-2a.<br />

An amino acid sequence of MT-2a without the N-terminal was<br />

observed for the peak eluting at 40.5 min (Figure 6f). The mass<br />

difference with regard to MT-2a isoforms indicated N-acetylation<br />

on methionine residue leading to the identification of N-Ac-MT-<br />

2a. This was also confirmed by two peaks of the MT2a specific<br />

fragments (acetylated and non acetylated) in Figure 3-ESI e. The<br />

sequence of the MT eluting at 41 min suggests strongly its<br />

6956 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

identification as N-Ac-MT-1a (δ -0.3 ppm). Indeed, the presence<br />

of the sequence fragment 39 Ala -Gly-Cys-Ala 42 and the overall<br />

sequence coverage of 95% supported this identification.<br />

The molecular masses of the remaining four apo-MTs in Table<br />

2 did not correspond to any of the reported pig MTs subisoforms.<br />

The 42-Da difference between the forms eluting at 33.8 and 37.4<br />

min (peaks 1 and 3) and between the forms eluting at 35 and<br />

38.5 min (peaks 2 and 5) indicated two pairs of acetylated and<br />

non-acetylated isoforms. The amino acid sequence of the MT<br />

subisoform eluting at 33.8 min (5940.124 Da, Figure 6a) was found<br />

to be as a mixture of MT-1a and MT-1e as m/z 1091.924 and m/z<br />

940.374 (charge +2) ions were observed for, respectively, Ala39<br />

and Lys43 residues and matched the theoretical y ions (m/z<br />

1091.929 and m/z 940.376 (charge +2). Furthermore, a y ion at


m/z 1172.935 was observed for Ser14 residue (theoretical m/z<br />

1172.939, z ) 4, δ 3.1 ppm) which differentiated this sequence<br />

from the one of MT-1f which contains a Thr14 residue. As the<br />

sequence matched 98% of pig MT-1 sequence and has never been<br />

reported before, it was denoted here as MT-1g. The sequence<br />

coverage with y ions was 88.5%. Consequently, the subisoform<br />

eluting at 37.4 min (5982.149 Da, Figure 6c) was referred to as<br />

N-Ac-MT-1g.<br />

Regarding the identification of the remaining pair of acetylated<br />

and non-acetylated isoforms eluting at 35.0 min (5944.149<br />

Da) and 38.5 min (5986.179 Da), the HCD spectrum of the<br />

N-acetylated form (m/z 1198.243, charge +5) gives the sequence<br />

information. It was found identical to MT1c until Pro38<br />

and then residues Ala39, Arg43, Gln46, Ile49 of MT1c were,<br />

respectively, replaced by Val39, Lys43, Lys46, Val49 as m/z<br />

1098.4547 (theoretical m/z 1098.4550, δ 0.2 ppm), m/z 933.3862<br />

(theoretical m/z 933.3869, δ 0.7 ppm), m/z 781.813 (theoretical<br />

m/z 781.816, δ 3.2 ppm) and m/z 1274.494 (theoretical m/z<br />

1274.499, δ 4.1 ppm) were detected at +2 charge state and the<br />

last one at +1 charge state. An ExPASy database search<br />

indicated that these modifications on 46 and 49 amino acids<br />

residues were reported for MT of others organisms (pig (in<br />

MT3), sheep, bovine and human). The sequence coverage was<br />

92%. As this MT has not been reported before, this pair of<br />

isoforms was denoted as MT-1h and N-Ac-MT-1h, respectively.<br />

The composition of the complexes formed indicates a release<br />

of Cd ions from the nanoparticles and their complexation by<br />

MTs formed.<br />

The identification of the last peak in the µHPLC-ICP MS<br />

chromatogram (peak 8 at 44.3 min (cf. Figure 4b) was less<br />

straightforward as no apo-isoform could be obtained. The presence<br />

of the fragment sequence 7 Cys-Ala-Ala 9 among +4 charge state<br />

fragments is characteristic for an MT-2 class isoforms. Another<br />

part of an MT sequence could be identified among +2 charge<br />

state fragments as 37 Cys-Pro-Val-Gly-Cys-Ala-Lys-Cys-Ala 45 (δ<br />

3.1 ppm), confirming the identity of the MT subisoform as<br />

MT2a. The end of the MT sequence 46 Gln-Gly-Cys-Ile-Cys-Lys-<br />

Gly-Ala-Ser-Asp 55 is observed as +1 charge state (δ 2.3 ppm).<br />

The metallothionein is obviously metalated and the mass<br />

difference in comparison with the N-Ac-MT2a theoretical mass<br />

(6011.190 Da) corresponds to the addition of Cu2Cd2 (351.665<br />

Da) with a loss of 8H + . This state of metalation is corroborated<br />

by the observation of three y ions corresponding to the 7 Cys-<br />

(47) Meloni, G.; Faller, P.; Vasak, M. J. Biol. Chem. 2007, 282, 16068–16078.<br />

(48) Salgado, M. T.; Stillman, M. J. Biochem. Biophys. Res. Commun. 2004, 318,<br />

73–80.<br />

Ala-Ala 9 fragment (m/z 1417.393, m/z 1391.644 and m/z<br />

1373.883 charge +4) differing by 85.89 mass unit from the<br />

corresponding apo subisoforms. The isotopic pattern was<br />

similar to the theoretical isotopic pattern of fragments with<br />

Cu2Cd2 adduct. Therefore peak 8 (Figure 4b) was identified as<br />

N-terminal acetylated Cu2Cd2-MT2a (6355.799 Da, theoretical<br />

6355.792 Da, δ -1.0 ppm) under acidic conditions and as<br />

N-terminal acetylated Cu2Cd6-MT2a (6797.354 Da, theoretical<br />

6797.344 Da, δ -1.4 ppm) under neutral conditions. The late<br />

elution of a similar complex from a rat kidney extract was<br />

reported elsewhere. 23 The comparison of the molecular mass<br />

of the complex with that of the apoMT indicates the presence<br />

of a disulfide bridge formed by oxidation of Cys by Cu(II)<br />

leading to the loss of additional two protons. 17,47 Indeed, the<br />

product ion mass spectra of Cd2Cu2MT2-a (cf. Figure 6), show<br />

a non-fragmented region (from Gly10 until Cys37) of the sequence<br />

suggesting the binding of four metal ions to Cys. This binding<br />

site is shared between the R and � domains of the MT2-a which<br />

is in agreement with the hypothesis of Vaher et al. 17 We can also<br />

notice that the unfragmented central region binding metals<br />

contains 11 Cys like the R domain of MT2a which would explain<br />

the stabilization of Cd2Cu2-MT2a complex. 48<br />

CONCLUSIONS<br />

µHPLC with the combined elemental detection by ICP MS, subppm<br />

molecular mass determination by electrospray orbital trap MS<br />

in conditions favoring metalation and demetalation, and HCD<br />

fragmentation allowing online protein sequencing was demonstrated<br />

as the, to date, most comprehensive approach to the characterization<br />

of metallothioneins in a biological cytosol. The detection limits down<br />

to the low femtomole levels allowed the characterization of metallothionein<br />

subisoforms induced in cell cultures exposed to CdS<br />

nanoparticles confirming the presence of known and the identification<br />

of new MT subisoforms.<br />

ACKNOWLEDGMENT<br />

The contribution of the Region of Aquitaine and the FEDER<br />

funds via CPER A2E (31486/08011464) project is acknowledged.<br />

We thank ICMCB, Bordeaux, for providing the 10 nm CdS<br />

nanoparticles used in this study.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Figure 1-ESI, Figure 2-ESI, and Figure 3-ESI. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 12, 2010. Accepted July 8, 2010.<br />

AC101245H<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6957


Anal. Chem. 2010, 82, 6958–6968<br />

Cleavable Cross-Linker for Protein Structure<br />

Analysis: Reliable Identification of Cross-Linking<br />

Products by Tandem MS<br />

Mathias Q. Müller, † Frank Dreiocker, ‡ Christian H. Ihling, † Mathias Schäfer,* ,‡ and Andrea Sinz* ,†<br />

Department of Pharmaceutical <strong>Chemistry</strong> and Bioanalytics, Institute of Pharmacy, Martin-Luther-Universität<br />

Halle-Wittenberg, Wolfgang-Langenbeck-Strasse 4, D-06120 Halle (Saale), and Institute for Organic <strong>Chemistry</strong>,<br />

Department of <strong>Chemistry</strong>, Universität zu Köln, Greinstrasse 4, D-50939 Cologne, Germany<br />

<strong>Chemical</strong> cross-linking combined with a subsequent<br />

enzymatic cleavage of the created cross-linked complex<br />

and a mass spectrometric analysis of the resulting crosslinked<br />

peptide mixture presents an alternative approach<br />

to high-resolution analysis, such as NMR spectroscopy or<br />

X-ray crystallography, to obtain low-resolution protein<br />

structures and to gain insight into protein interfaces.<br />

Here, we describe a novel urea-based cross-linker, which<br />

allows distinguishing different cross-linking products by<br />

collision-induced dissociation (CID) tandem MS experiments<br />

based on characteristic product ions and constant<br />

neutral losses. The novel cross-linker is part of our<br />

ongoing efforts in developing collision-induced dissociative<br />

reagents that allow an efficient analysis of cross-linked<br />

proteins and protein complexes. Our innovative analytical<br />

concept is exemplified for the Munc13-1 peptide and the<br />

recombinantly expressed ligand binding domain of the<br />

peroxisome proliferator-activated receptor r, for which<br />

cross-linking reaction mixtures were analyzed both by<br />

offline nano-HPLC/MALDI-TOF/TOF mass spectrometry<br />

and by online nano-HPLC/nano-ESI-LTQ-orbitrap mass<br />

spectrometry. The characteristic fragment ion patterns of<br />

the novel cross-linker greatly simplify the identification<br />

of different cross-linked species, namely, modified peptides<br />

as well as intrapeptide and interpeptide cross-links,<br />

from complex mixtures and drastically reduce the potential<br />

of identifying false-positive cross-links. Our novel ureabased<br />

CID cleavable cross-linker is expected to be highly<br />

advantageous for analyzing protein 3D structures and<br />

protein-protein complexes in an automated manner.<br />

<strong>Chemical</strong> cross-linking 1 combined with mass spectrometry<br />

presents an alternative approach to study the tertiary and<br />

quaternary structure of proteins and protein complexes. 2-4<br />

However, an unambiguous, sensitive, reliable, and fast identifica-<br />

* To whom correspondence should be addressed. (A.S.) Phone: +49-345-<br />

5525170. Fax: +49-345-5527026. E-mail: andrea.sinz@pharmazie.uni-halle.de.<br />

(M.S.) Phone: +49-221-4703086. Fax: +49-221-4703064. E-mail: mathias.schaefer@<br />

uni-koeln.de.<br />

† Martin-Luther-Universität Halle-Wittenberg.<br />

‡ Universität zu Köln.<br />

(1) Sinz, A. Angew. Chem., Int. Ed. 2007, 46, 660–662.<br />

(2) Sinz, A. J. Mass Spectrom. 2003, 38, 1225–1237.<br />

(3) Trakselis, M. A.; Alley, S. C.; Ishmael, F. T. Bioconjugate Chem. 2005, 16,<br />

741–750.<br />

6958 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

tion of the amino acids involved in the covalent derivatization by<br />

chemical cross-linking remains challenging. A mass spectrometric<br />

identification of the chemically modified amino acids in the<br />

respective protein is performed after proteolytic digestion of<br />

the cross-linking reaction mixtures. This is often hampered by<br />

the complexity of the created peptide mixtures, wherein only a<br />

relatively small percentage of cross-linked products are present<br />

besides a majority of unmodified peptides. To safeguard for a<br />

selective identification of cross-linked peptides using electrospray<br />

ionization (ESI) 5-7 combined with collision-induced dissociation<br />

(CID) 8,9 tandem MS, 10 several approaches have been suggested.<br />

11-14 They include cross-linking reagents that incorporate<br />

the use of marker ions resulting from low-energy CID 15,16 or<br />

metastable decay, 17 isotope-coding strategies, such as proteolytic<br />

digestion in 18 O water, 18 isotope coding of the cross-linking<br />

reagents 19-22 or of the proteins, 23 and an enrichment of cross-<br />

(4) Sinz, A. Mass Spectrom. Rev. 2006, 25, 663–682.<br />

(5) Fenn, J. B.; Mann, M.; Meng, C. K.; Wong, S. F.; Whitehouse, C. M. Science<br />

1989, 246, 64–71.<br />

(6) Cole, R. Electrospray Ionization Mass Spectrometry: Fundamentals, Instrumentation<br />

and Applications; 1997.<br />

(7) Kebarle, P.; Verkerk, U. H. Mass Spectrom. Rev. 2009, 28, 898–917.<br />

(8) Wells, J. M.; McLuckey, S. A. The Encyclopedia of Mass Spectrometry; 2003;<br />

p1.<br />

(9) Shukla, A. K.; Futrell, J. H. J. Mass Spectrom. 2000, 35, 1069–1090.<br />

(10) Hunt, D. F.; Yates, J. R., III; Shabanowitz, J.; Winston, S.; Hauer, C. R. Proc.<br />

Natl. Acad. Sci. U.S.A. 1986, 83, 6233–6237.<br />

(11) Soderblom, E. J.; Goshe, M. B. Anal. Chem. 2006, 78, 8059–8068.<br />

(12) Soderblom, E. J.; Bobay, B. G.; Cavanagh, J.; Goshe, M. B. Rapid Commun.<br />

Mass Spectrom. 2007, 21, 3395–3408.<br />

(13) Dreiocker, F.; Müller, M. Q.; Sinz, A.; Schäfer, M. J. Mass Spectrom. 2010,<br />

45, 178–189.<br />

(14) Müller, M. Q.; Dreiocker, F.; Ihling, C. H.; Sinz, A.; Schäfer, M. J. Mass<br />

Spectrom., in press.<br />

(15) Back, J. W.; Hartog, A. F.; Dekker, H. L.; Muijsers, A. O.; de Koning, L. J.;<br />

de Jong, L. J. Am. Soc. Mass Spectrom. 2001, 12, 222–227.<br />

(16) Tang, X.; Munske, G. R.; Siems, W. F.; Bruce, J. E. Anal. Chem. 2005, 77,<br />

311–318.<br />

(17) Young, M. M.; Tang, N.; Hempel, J. C.; Oshiro, C. M.; Taylor, E. W.; Kuntz,<br />

I. D.; Gibson, B. W.; Dollinger, G. Proc. Natl. Acad. Sci. U.S.A. 2000, 97,<br />

5802–5806.<br />

(18) Back, J. W.; Notenboom, V.; de Koning, L. J.; Muijsers, A. O.; Sixma, T. K.;<br />

de Koster, C. G.; de Jong, L. Anal. Chem. 2002, 74, 4417–4422.<br />

(19) Petrotchenko, E. V.; Olkhovik, V. K.; Borchers, C. H. Mol. Cell. Proteomics<br />

2005, 4, 1167–1179.<br />

(20) Schulz, D. M.; Kalkhof, S.; Schmidt, A.; Ihling, C.; Stingl, C.; Mechtler, K.;<br />

Zschörnig, O.; Sinz, A. Proteins 2007, 69, 254–269.<br />

(21) Ihling, C.; Schmidt, A.; Kalkhof, S.; Schulz, D. M.; Stingl, C.; Mechtler, K.;<br />

Haack, M.; Beck-Sickinger, A. G.; Cooper, D. M.; Sinz, A. J. Am. Soc. Mass<br />

Spectrom. 2006, 17, 1100–1113.<br />

10.1021/ac101241t © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/22/2010


linked products via affinity tags. 1,24,25 Identification of cross-linked<br />

peptide ions by CID tandem MS or MS n is not trivial as product<br />

ion spectra of these contain a number of product ions, i.e., band<br />

y-type ions, 26,27 originating from both peptides involved in<br />

the cross-linking product. To enable a more effective analysis of<br />

cross-linked peptides by CID tandem MS analysis, we have<br />

developed a novel concept for collision-induced dissociative crosslinking<br />

reagents. As such, the novel cross-linker contains a labile<br />

covalent bond located within the linker region, which is selective<br />

and preferably cleaved by collision activation in the gas phase.<br />

Highly efficient and reproducible cleavage of our novel urea-based<br />

cross-linker is mediated by a nucleophilic attack of a carbonyl<br />

oxygen in the cross-linker’s spacer chain at the urea carbonyl<br />

carbon. The synthesis and application of the predecessor of this<br />

cross-linker, a thiourea derivative, has been studied in detail<br />

previously. 14 However, that thiourea-based cross-linker exhibited<br />

a number of properties which were unfavorable for analyzing<br />

cross-linked products by CID in an automated fashion. Most<br />

importantly, we were unable to differentiate between the different<br />

types of cross-linked species, namely, peptides that are modified<br />

by a partially hydrolyzed cross-linker or intramolecular type 1<br />

cross-linked products.<br />

The urea-based cross-linker presented herein overcomes the<br />

deficiencies of the thiourea-based reagent. It leads to the formation<br />

of indicative fingerprint mass shifted product ions and neutral<br />

losses in the product ion mass spectra, which allow unambiguous<br />

identification of the different types of cross-linking products. The<br />

properties of the novel cross-linker were evaluated for the Munc<br />

13-1 peptide and the 28 kDa ligand binding domain of the<br />

peroxisome proliferator-activated receptor R (PPARR), impressively<br />

indicating its power for efficiently analyzing cross-linked<br />

products in a highly automated fashion.<br />

EXPERIMENTAL SECTION<br />

Materials. All chemicals and solvents used for synthesis of<br />

the cross-linker were used as purchased without further purification<br />

(Acros Organics, Geel, Belgium; ABCR, Karlsruhe, Germany).<br />

Dichloromethane and pyridine were dried by distillation from<br />

calcium hydride. All other solvents were distilled over a column<br />

prior to use.<br />

Buffer reagents and chemicals were obtained from Sigma<br />

(Taufkirchen, Germany). Proteases (trypsin and GluC, sequencing<br />

grade) were obtained from Roche Diagnostics (Mannheim,<br />

Germany). Nano-HPLC solvents were spectroscopic grade (Uvasol,<br />

VWR, Darmstadt, Germany). MALDI matrixes and calibration<br />

standards were purchased from Bruker Daltonik (Bremen,<br />

Germany). Water was purified with a Direct-Q5 water purification<br />

system (Millipore, Eschborn, Germany). The ligand binding<br />

domain of PPARR was expressed in Escherichia coli and purified<br />

(22) Schmidt, A.; Kalkhof, S.; Ihling, C.; Cooper, D. M.; Sinz, A. Eur. J. Mass<br />

Spectrom. 2005, 11, 525–534.<br />

(23) Taverner, T.; Hall, N. E.; O’Hair, R. A.; Simpson, R. J. J. Biol. Chem. 2002,<br />

277, 46487–46492.<br />

(24) Hurst, G. B.; Lankford, T. K.; Kennel, S. J. J. Am. Soc. Mass Spectrom. 2004,<br />

15, 832–839.<br />

(25) Sinz, A.; Kalkhof, S.; Ihling, C. J. Am. Soc. Mass Spectrom. 2005, 16, 1921–<br />

1931.<br />

(26) Roepstorff, P.; Fohlman, J. Biomed. Mass Spectrom. 1984, 11, 601.<br />

(27) Papayannopoulos, I. A.; Biemann, K. Protein Sci. 1992, 1, 278–288.<br />

Scheme 1. Structure of the Symmetric<br />

NHS-BuUrBu-NHS Compound (1) for <strong>Chemical</strong><br />

Cross-Linking<br />

according to a previously published protocol. 28 The Munc13-1<br />

peptide was synthesized by Dr. O. Jahn (MPI for Experimental<br />

Medicine, Göttingen, Germany).<br />

Synthesis of the Cross-Linker. Aminobutanoic acid was<br />

dissolved in 4 M NaOH, and 0.175 equiv of triphosgen in dioxane<br />

was added at 0 °C (see Scheme S1 of the Supporting Information).<br />

The reaction mixture was brought to room temperature overnight,<br />

and the solid was removed by filtration. The solvent was removed<br />

under reduced pressure, and the residue was recrystallized from<br />

6 M hydrochloric acid. The product was isolated by filtration,<br />

washed with acetone, and dried in vacuum. HO-BuUrBu-OH was<br />

isolated as a colorless solid in a yield of 26%. 29 An abbreviated<br />

notation, BuUrBu, is used throughout this paper for the backbone<br />

of cross-linker 1, which is derived from urea (Ur) and aminobutyric<br />

acid (Bu) (Scheme 1).<br />

HO-BuUrBu-OH was dissolved in pyridine and treated with<br />

N-(trifluoroacetoxy)succinimidesprepared from N-hydroxysuccinimide<br />

(NHS) and trifluoroacetic anhydridesat 0 °C. 30 The<br />

reaction mixture was brought to room temperature within 2 h.<br />

After addition of ethyl acetate the raw product was isolated by<br />

filtration and dissolved in a mixture of dichloromethane and<br />

methanol. Insoluble components were removed by filtration, and<br />

the filtrate was evaporated. NHS-BuUrBu-NHS (1) was isolated<br />

as a colorless solid in a yield of 82%.<br />

Cross-Linking Reactions. For cross-linking experiments,<br />

aqueous stock solutions of Munc13-1 peptide (100 µg/mL) or<br />

PPARR ligand binding domain (2 mg/mL) were diluted with 20<br />

mM HEPES, 150 mM NaCl, 1 mM TCEP, 10% (v/v) glycerol, pH<br />

8.0, to a volume of 1 mL, giving a final protein/peptide concentration<br />

of 10 µM. The cross-linker (200 mM stock solution in DMSO)<br />

was added in 50, 100, and 200 M excess to the protein/peptide<br />

solution, and the reactions were allowed to proceed for 5, 15, 30,<br />

60, and 120 min. Reactions were quenched with ammonium<br />

bicarbonate (20 mM final concentration). One 200 µL aliquot was<br />

taken from each sample and stored at -20 °C before MS analysis.<br />

Cross-linked Munc13-1 peptide was digested with trypsin, whereas<br />

PPARR was digested either with trypsin or with trypsin/GluC<br />

(each 1:100 (w/w) enzyme to substrate ratio) overnight at 25 °C<br />

according to an existing protocol. 20 The resulting digests were<br />

stored at -20 °C before MS analysis.<br />

Linear Mode MALDI-TOF Mass Spectrometry. MALDI-<br />

TOF mass spectrometry of intact cross-linked PPARR was<br />

(28) Müller, M. Q.; Roth, C.; Sträter, N.; Sinz, A. Protein Expression Purif. 2008,<br />

62, 185–189.<br />

(29) Coe, S.; Kane, J. J.; Nguyen, T. L.; Toledo, L. M.; Wininger, E.; Fowler,<br />

F. W.; Lauher, J. W. J. Am. Chem. Soc. 1997, 119, 86–93.<br />

(30) Rao, T. S.; Nampalli, S.; Sekher, P.; Kumar, S. Tetrahedron Lett. 2002, 43,<br />

7793–7795.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6959


performed in linear and positive ionization mode on an Ultraflex<br />

III instrument (Bruker Daltonik) equipped with a Smartbeam laser<br />

using sinapinic acid (SA) as the matrix. A 5 pmol portion of protein<br />

was applied to a ground steel target (Bruker Daltonik) using the<br />

double layer method as described in the manufacturer’s manual<br />

of the Ultraflex III mass spectrometer.<br />

Nano-HPLC/MALDI-TOF/TOF Mass Spectrometry. Peptide<br />

mixtures derived from enzymatic digestions of Munc13-1 and<br />

PPARR were analyzed by offline coupling of a nano-HPLC system<br />

(Ultimate 3000, Dionex Corp., Idstein, Germany) to a MALDI-<br />

TOF/TOF mass spectrometer (Ultraflex III, Bruker Daltonik).<br />

Samples were injected via the autosampler to a precolumn<br />

(Acclaim PepMap, C18, 300 µm × 5 mm, 5 µm, 100 Å, Dionex)<br />

and desalted by washing the precolumn for 15 min with 0.1% TFA<br />

before the peptides were eluted onto the separation column<br />

(Acclaim PepMap, C18, 75 µm × 250 mm, 5 µm, 100 Å, Dionex),<br />

which had been equilibrated with 95% solvent A (5% ACN, 0.05%<br />

TFA). For PPARR, peptides were separated with a 60 min gradient<br />

(0-60 min, 5-50% B; 60-62 min, 50-100% B; 62-67 min, 100%<br />

B; 67-72 min, 5% B; solvent B ) 80% ACN, 0.04% TFA) at a flow<br />

rate of 300 nL/min with UV detection at 214 and 280 nm. Eluates<br />

were fractionated into 18 s fractions using the fraction collector<br />

Proteineer fc (Bruker Daltonik), mixed with 1.1 µL of matrix<br />

solution (0.8 mg/mL HCCA in 90% ACN, 0.1% TFA, 1 mM<br />

NH4H2PO4) and directly prepared on a 384 MTP 800 µm<br />

AnchorChip target (Bruker Daltonik). For separation of<br />

Munc13-1 peptides, slightly different nano-HPLC conditions<br />

were employed. Tryptic peptide mixtures were separated by<br />

nano-HPLC (30 min, 5-50% B), and fractions were collected<br />

on the AnchorChip 384/800 target in 30 s steps.<br />

MALDI-TOF/TOF-MS/MS analyses were conducted in the<br />

positive ionization and reflectron mode by accumulating 2000 laser<br />

shots in the range m/z 700-5000 to one mass spectrum. Mass<br />

spectra were externally calibrated using Peptide Calibration<br />

Standard II (Bruker Daltonik). Signals with a signal-to-noise ratio<br />

(S/N) > 15 were selected and subjected to laser-induced fragmentation<br />

(LID). In-source decay (ISD) was performed manually<br />

on isolated cross-linking product candidates by increasing the laser<br />

energy by ca. 10% relative to the usual laser energy in LID-MS/<br />

MS. Data acquisition was done automatically by the WarpLC 1.1<br />

software (Bruker Daltonik) coordinating MS data acquisition<br />

(FlexControl 1.3) and data processing (FlexAnalysis 3.0).<br />

Nano-HPLC/Nano-ESI-LTQ-Orbitrap Mass Spectrometry.<br />

Fractionation of tryptic peptide mixtures (Munc13-1 and PPARR)<br />

was carried out on an Ultimate nano-HPLC system (Dionex) using<br />

reversed-phase C18 columns (precolumn, Acclaim PepMap, 300<br />

µm × 5 mm, 5 µm, 100 Å; separation column, Acclaim PepMap,<br />

75 µm × 250 mm, 5 µm, 100 Å; Dionex). After being washed on<br />

the precolumn for 15 min with water containing 0.1% TFA, the<br />

peptides were eluted and separated using gradients from 0% to<br />

50% B (90 min) and from 50% to 100% B (1 min) and 100% B (5<br />

min), with solvent A being 5% ACN containing 0.1% FA and solvent<br />

B being 80% ACN containing 0.1% FA. The nano-HPLC system<br />

was directly coupled to the nano-ESI source (Proxeon, Odense,<br />

Denmark) of an LTQ-orbitrap XL hybrid mass spectrometer<br />

(Thermo Fisher Scientific, Bremen, Germany). 31 MS data were<br />

(31) Hu, Q.; Noll, R. J.; Li, H.; Makarov, A.; Hardman, M.; Graham Cooks, R. J.<br />

Mass Spectrom. 2005, 40, 430–443.<br />

6960 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

acquired over 122 min in data-dependent MS 2 mode: Each highresolution<br />

full scan (m/z 300-2000, R ) 60 000) in the orbitrap<br />

was followed by three product ion scans in the orbitrap (R )<br />

7500) on the three most intense signals in the full-scan mass<br />

spectrum (isolation window 1.5 u). MS 3 was performed on the<br />

three most intense signals in CID tandem mass spectra, and<br />

fragment ions were either analyzed in the linear quadrupole<br />

ion trap (LTQ) or in the orbitrap (R ) 7500). Dynamic<br />

exclusion (exclusion duration 180 s, exclusion window -1 to<br />

+1 Th) was enabled to allow detection of less abundant ions.<br />

Data acquisition was controlled via XCalibur 2.0.7 (Thermo<br />

Fisher) in combination with DCMS link 2.0 (Dionex).<br />

Analysis of Cross-Linked Products. Cross-linked products<br />

were identified by analyzing the MS data using General Protein<br />

Mass Analysis for Windows (GPMAW), 32 version 8.10 (Lighthouse<br />

Data, Odense, Denmark, http://www.gpmaw.com) and the Cool-<br />

ToolBox software program, an upgrade of the VIRTUALMSLAB<br />

software program. 33 The length of the cross-linker was calculated<br />

using the ViewerLight 5.0 software (Accelrys). 34<br />

RESULTS AND DISCUSSION<br />

Cross-Linker Design. The dissociative cross-linker specifically<br />

designed for CID tandem MS is a simple and symmetric<br />

molecule with a central urea moiety (Scheme 1), similar to the<br />

previously introduced thiourea derivative. 14 The practical strategy<br />

of inducing a preferred and defined dissociation of the chemical<br />

cross-linker upon CID was adapted from the Edman degradation<br />

mechanism, which was exemplified for a cross-linker containing<br />

a thiourea moiety that is connected to proline via a glycine<br />

residue. 13 The refined analytical concept of the herein presented<br />

novel cleavable cross-linker relies on an intramolecular nucleophilic<br />

attack at the central urea carbonyl, resulting in an efficient<br />

cleavage of the cross-linker and giving rise to characteristic<br />

fragment ions and constant neutral losses (CNLs). The symmetric<br />

compound is simple as it consists of a central urea moiety that is<br />

flanked by two aminobutyric acids connected to activated NHS<br />

esters, which react with lysines or free N-termini in proteins. 4<br />

An abbreviated notation, BuUrBu, is used throughout this<br />

paper for the novel cross-linker 1, which is derived from urea<br />

(Ur) and aminobutyric acid (Bu) (Scheme 1). The protein<br />

derivatization reactions of 1 are illustrated in Scheme 2S of the<br />

Supporting Information. After one reactive site has reacted with<br />

a primary amine in a protein, the second reactive site has at least<br />

two options to react with proteins: (a) it might form an intermolecular<br />

cross-link (2) with another protein (type 2 cross-link 35 )<br />

or (b) it might react with another amine group of the same protein,<br />

leading to an intramolecular cross-link (type 1, structure 3).<br />

Alternatively, the second NHS functionality is hydrolyzedsin case<br />

no amine group is within the correct distance to be crosslinkedsor<br />

it is aminolyzed during reaction quenching with<br />

ammonium salts (type 0 cross-link linker, 35 structures 4 and 5,<br />

respectively).<br />

(32) Peri, S.; Steen, H.; Pandey, A. Trends Biochem. Sci. 2001, 26, 687–689.<br />

(33) de Koning, L. J.; Kasper, P. T.; Back, J. W.; Nessen, M. A.; Vanrobaeys, F.;<br />

Van Beeumen, J.; Gherardi, E.; de Koster, C. G.; de Jong, L. FEBS J. 2006,<br />

273, 281–291.<br />

(34) ViewerLight 5.0, Accelrys, San Diego, CA.<br />

(35) Schilling, B.; Row, R. H.; Gibson, B. W.; Guo, X.; Young, M. M. J. Am. Soc.<br />

Mass Spectrom. 2003, 14, 834–850.


Scheme 2. Fragmentation Mechanism of Protonated 4 (Scheme S2, Supporting Information) upon CID, Delivering a<br />

Doublet of 26 u Mass Shifted Product Ions [M + H + Bu] + (6) and [M + H + BuUr] + (7) by CNLs of 129 and 103 u<br />

Reactivity and CID Fragmentation. The novel symmetric<br />

urea-based cross-linker is unique in that it allows a discrimination<br />

of different types of cross-linked products, namely, type 0 (a<br />

peptide that is modified by a partially hydrolyzed cross-linker 35 ),<br />

type 1 (intrapeptide cross-link), and type 2 (interpeptide crosslink),<br />

based on characteristic fragmentation reactions. As such,<br />

the cross-linker allows a screening for cross-linked species in a<br />

highly automated fashion. The CID fragmentation mechanism of<br />

peptides modified with hydrolyzed linkers (the type 0 cross-link<br />

linker 4 35 is depicted in Scheme 2). A nucleophilic attack at the<br />

carbonyl carbon of the urea moiety leads to a cleavage of the crosslinker<br />

at either urea amide bond. Depending on which oxygen<br />

attacks the urea carbonyl carbon, either the peptide modified with<br />

4-aminobutyric acid (6) ([M + H + 85 u] + ) corresponding to a<br />

CNL of 129 u (1,3-oxazepan-2-one) or the peptide that is<br />

decorated with 1,3-oxazepan-2-one (7) ([M + H + 111 u] + )<br />

corresponding to a CNL of 103 u (4-aminobutyric acid) is<br />

observed. Intriguingly, the two product ions 6 and 7 formed<br />

are mass shifted by 26 u, which is a unique indicator of a “deadend”<br />

(type 0) cross-link, i.e., a peptide that is modified by a<br />

partially hydrolyzed cross-linker (Scheme 2).<br />

In analogy to the fragmentation behavior of the type 0 crosslink<br />

(Scheme 2), an interpeptide (type 2) cross-link (2) results in<br />

the cleavage of the urea amide bond, yielding a pair of complementary<br />

product ions that originate from the symmetric structure<br />

of the cross-linker (Scheme 3). Either peptide ions are generated<br />

that are modified with 4-aminobutyric acid (ions 6a and 6b) or,<br />

alternatively, peptide product ions are formed carrying a 1,3oxazepan-2-one<br />

(7a and 7b) modification. This specific fragmentation<br />

behavior of an intermolecular cross-linked precursor ion (2)<br />

gives rise to two doublets of peptide ions, which exhibit the<br />

characteristic mass difference of 26 u. This unique pattern of a<br />

“doublet of 26 u doublets” in CID mass spectra is highly indicative<br />

in evidencing the presence of an interpeptide cross-link (type 2),<br />

while a dead-end (type 0) cross-link only generates a single 26 u<br />

doublet as discussed above. Strikingly different from the characteristic<br />

fragment ions created for type 0 and type 2 cross-links is<br />

the fragmentation pattern of a type 1 intramolecular cross-link (3).<br />

As shown in Scheme 4, CID of intramolecular cross-links leads<br />

to the effective formation of pyrolidinone, detectable as a CNL<br />

for 85 u. A summary of the characteristic fragment ions and CNLs<br />

created for the different cross-linked products that allow for a rapid<br />

screening is shown in Table 1. We also observed mixed species,<br />

i.e., peptides containing more than one cross-linker molecule,<br />

which were not considered in this study.<br />

Nano-HPLC/Nano-ESI-LTQ-Orbitrap Mass Spectrometry.<br />

To assess the behavior of the NHS-BuUrBu-NHS cross-linker (1)<br />

in CID experiments, two model substances, the 22-amino acid<br />

Munc13-1 peptide (CRAKANWLRAFNKVRMQLQEAR) and the<br />

28 kDa ligand binding domain of PPARR, were selected. Both were<br />

cross-linked with 1, enzymatically digested and subjected to nano-<br />

HPLC/nano-ESI-LTQ-orbitrap-MS n . The Munc13-1 peptide is<br />

currently under investigation in the Sinz lab 36 and was chosen<br />

as a model peptide for evaluating the properties of 1 as it<br />

contains three primary amines (N-terminus, Lys-4 and Lys-13)<br />

that are separated by four arginines, allowing a tryptic cleavage.<br />

Peptides Modified with a Partially Hydrolyzed Cross-<br />

Linker (Dead-End Cross-Links, Type 0). In Figure 1A, the ESI-<br />

CID-MS/MS product ion spectrum of a modified tryptic peptide<br />

(AKANWLR) of Munc13-1 is presented. The abundant precursor<br />

ion at m/z 1072.60 was selected and completely dissociated by<br />

effective collision activation in the linear quadrupole ion trap. The<br />

ion at m/z 1072.60 was assigned as a type 0 cross-linked ion [M<br />

+ H + BuUrBu-OH] + that is modified with a partially hydrolyzed<br />

cross-linker (4) at the lysine residue. This assignment<br />

is consistent with the fragmentation scheme shown in Scheme<br />

2. The expected CNLs of 129 and 103 u consequently lead to the<br />

observation of specific fragment ions with an indicative mass<br />

difference of 26 u, in this case found at m/z 943.55 and at m/z<br />

969.53 (Figure 1A). Structure identification of these ions was<br />

achieved by interpretation of the LTQ-MS 3 product ion spectra<br />

shown in Figure 1B,C. The former spectrum shows the respective<br />

product ion spectra of [M + H + Bu] + at m/z 943.55 (6), while<br />

(36) Dimova, K.; Kalkhof, S.; Pottratz, I.; Ihling, C.; Rodriguez-Castaneda, F.;<br />

Liepold, T.; Griesinger, C.; Brose, N.; Sinz, A.; Jahn, O. Biochemistry 2009,<br />

48, 5908–5921.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6961


Scheme 3. Fragmentation Mechanism of Protonated 2 upon CID, Delivering Two Complementary Doublets of 26 u<br />

Mass Shifted Product Ions a<br />

a Product ions of peptide 1 are 6a and 7a, and product ions of peptide 2 are 6b and 7b.<br />

Scheme 4. Fragmentation Mechanism of a Protonated Type 1 Modified Peptide (3) upon CID, Delivering a Product<br />

Ion That Is Modified with BuUr [M + H + BuUr] + (7) by a CNL of Pyrolidinone (85 u)<br />

Table 1. Summary of the Characteristic CNLs and Fragment Ion Mass Shifts Used for Discrimination of the<br />

Different Cross-Linking Types (ESI-Ion Trap-MS/MS and MALDI-TOF/TOF-MS/MS)<br />

cross-link type hydrolyzed, type 0 aminolyzed, type 0 intrapeptide, type 1 interpeptide, type 2<br />

CNL 103 and 129 u 102 and 128 u 85 u<br />

no. of 26 u doublets 1 1 2<br />

the latter presents the respective product ion spectrum of [M<br />

+ H + BuUr] + at m/z 969.53 (7). In both experiments the<br />

product ion analysis was conducted in the orbitrap at a mass<br />

accuracy of 1-2 ppm employing a resolving power of 15 000.<br />

In the MS 3 product ion spectra presented in Figure 1B,C, the<br />

precursor ions exhibit frequent losses of water and ammonia, and<br />

more importantly, extensive backbone cleavages allow an efficient<br />

sequencing of the peptide. In the present case, b-type ions<br />

containing the cross-linker dominate the spectra. From the CID<br />

fragmentation pattern (Figure 1A) it is immediately obvious that<br />

fragmentation of the urea cross-linker is strongly preferred, while<br />

the peptide backbone remains virtually intact. The effective<br />

fragmentation of the partially hydrolyzed cross-linker 4 leads to<br />

the formation of prominent “26 u doublet” product ions 6 and 7<br />

(Table 1). As highlighted above, these product ions are indicative<br />

of a selective and sensitive detection of the respective cross-linked<br />

6962 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

species as they cannot be confused with b- and corresponding<br />

a-type ions giving a mass difference of 28 u (CO). Peptides that<br />

are modified by an amidated cross-linker (5), resulting from the<br />

quenching reaction with ammonium bicarbonate, are also observed<br />

in the product ion mass spectra, 37 accordingly exhibiting<br />

CNLs of 128 and 102 u (Scheme 2, Table 1). It should be noted<br />

that peptides modified by hydrolyzed (4) and amidated (5) linkers<br />

were readily separated by our nano-HPLC method with ca. 30 s<br />

elution time difference for a 1000 Da peptide.<br />

An analogous fragmentation behavior was observed for a<br />

modified tryptic peptide of PPARR comprising amino acids<br />

259-272. ESI-CID-MS 2 experiments resulted in characteristic<br />

CNLs of 129 and 103 u with the indicative 26 u pattern due to<br />

(37) Kalkhof, S.; Sinz, A. Anal. Bioanal. Chem. 2008, 392, 305–312.


Figure 1. (A) ESI-LTQ-CID-MS 2 product ion spectrum of the type 0 modified tryptic peptide AKANWLR ion [M + H + BuUrBu-OH] + (structure<br />

4 (Scheme S2, Supporting Information) at m/z 1072.60) of Munc13-1. ESI-LTQ-MS 3 product ion spectra of [M + H + Bu] + at m/z 943.55 (B)<br />

and of [M + H + BuUr] + at m/z 969.53 (C).<br />

cleavage of 4 (Figure S1 of the Supporting Information). In<br />

subsequent MS 3 product ion experiments, the peptide backbone<br />

was sequenced, clearly evidencing the modification of Lys-266<br />

in the flexible ω loop of PPARR (Figure S1B,C).<br />

Intrapeptide Cross-Linking Products (Type 1). As the next<br />

step, we aimed to evaluate the behavior of compound 1 in<br />

intrapeptide cross-linked products (3, Scheme 2S, Supporting<br />

Information). In Figure 2, ESI-CID-MS 2 and -MS 3 data of a<br />

derivatized PPARR peptide (aa 196-210) are exemplarily<br />

presented. In contrast to peptides that were modified by a<br />

partially hydrolyzed cross-linker (type 0), intrapeptide crosslinks<br />

(type 1, structure 3, Scheme 2S) exhibit a characteristic<br />

neutral loss of 85 u upon CID (Table 1), resulting from the<br />

elimination of a pyrolidinone as illustrated in Scheme 3. This<br />

property renders our novel cross-linker highly advantageous for<br />

distinguishing different types of cross-links from each other and<br />

thus makes it an extremely valuable tool for a rapid and automated<br />

screening of cross-linking products. ESI-MS 2 and -MS 3 product<br />

ion experiments enabled us to readily identify Lys-204 to be<br />

connected with Lys-208 in the N-terminal helix of PPARR. The<br />

efficient fragmentation of cross-linked species compared to<br />

conventional NHS ester cross-linkers, such as bis(sulfosuccin-<br />

imyl) suberate (BS 3 ), 38 makes it clearly superior to these<br />

noncleavable derivatives.<br />

Interpeptide Cross-linking Product (Type 2). Finally, we<br />

examined the fragmentation behavior of 1 for the analysis of<br />

interpeptide cross-links (type 2, structure 2), which present the<br />

most informative distance constraints with respect to analyzing<br />

3D structures of proteins or to mapping protein-protein interaction<br />

sites. Modified Munc13-1 (Figure 3) and PPARR (273 aa;<br />

Figure S2, Supporting Information) peptides deliver the two<br />

indicative 26 u doublets of fragment ions 6a, 7a and 6b, 7b<br />

(Scheme 3, Table 1), which allowed an unambiguous identification<br />

of cross-linked lysines. For the Munc13-1 peptide, amino acids<br />

3-9 were found to be connected with amino acids 10-15, clearly<br />

pointing to a cross-link between Lys-4 and Lys-13 (Figure 3). For<br />

the PPARR peptide, Lys-208 was demonstrated to be cross-linked<br />

to Lys-216 (Figure S2), a cross-linking product that had not been<br />

identified in previous cross-linking studies of PPARR with noncleavable<br />

cross-linkers. 39 Interestingly, the most abundant signals<br />

in the MS 3 product ion spectra (Figure S2B,C) correspond to<br />

both unmodified peptides comprising the cross-linking product<br />

(38) Dihazi, G. H.; Sinz, A. Rapid Commun. Mass Spectrom. 2003, 17, 2005–<br />

2014.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6963


Figure 2. (A) ESI-LTQ-CID-MS 2 product ion spectrum of the type 1 modified PPARR peptide (aa 195-209) [M + 2H + BuUrBu] 2+ at m/z<br />

945.47. (B) ESI-LTQ-MS 3 product ion spectrum of [M + 2H + BuUr] 2+ at m/z 902.94. $: signal not assigned.<br />

and, as such, originate from cleavage of the amide bond<br />

between the lysine ε-amine group and the cross-linker.<br />

Offline Nano-HPLC/MALDI-TOF/TOF Mass Spectrometry.<br />

To ensure and to probe the general applicability of 1 for<br />

both ESI- and MALDI-MS, we additionally examined the frag-<br />

(39) Müller, M. Q.; de Koning, L. J.; Schmidt, A.; Ihling, C.; Syha, Y.; Rau, O.;<br />

Mechtler, K.; Schubert-Zsilavecz, M.; Sinz, A. J. Med. Chem. 2009, 52,<br />

2875–2879.<br />

6964 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

mentation behavior of the cross-linker in the higher kinetic energy<br />

regime of collision activation (kinetic energy of the precursor ion<br />

in TOF/TOF-CID Ekin ≈ 1 keV) 40-42 in MALDI-TOF/TOF (ISD<br />

and CID) experiments. In the same fashion as for low-energy ESI-<br />

LTQ-orbitrap-MS n product experiments (kinetic energy of the<br />

precursor ion in LTQ-CID Ekin ≈ 1 eV) 43,44 Munc13-1 and PPARR<br />

peptides were created by tryptic digestion and separated by nano-<br />

HPLC before MS analysis.


Figure 3. (A) ESI-LTQ-CID-MS 2 product ion spectrum of the triply protonated type 2 modified Munc13-1 peptide (amino acids 3-9 connected<br />

with amino acids 10-15) at m/z 597.01. (B, C) ESI-LTQ-MS 3 product ion spectra of the modified R peptide [R +2H + Bu] 2+ at m/z 463.76 (B)<br />

and of the � peptide [� + H + Bu] + at m/z 819.48 (C).<br />

Peptides Modified with Hydrolyzed Linkers (Type 0). As<br />

with collision activation in the low-energy regime in an LTQ-CID-<br />

MS 2 product ion experiment of the type 0 modified Munc13-1<br />

peptide (Figure 1A), the MALDI-TOF/TOF-CID spectrum exhibits<br />

the 26 u doublet of product ions 6 and 7 as base peaks<br />

resulting from the highly preferential cleavage of the urea crosslinker<br />

(Figure 4, Table 1). However, due to the overall increased<br />

(40) Pittenauer, E.; Allmaier, G. Comb. Chem. High Throughput Screening 2009,<br />

12, 137–155.<br />

(41) Medzihradszky, K. F.; Campbell, J. M.; Baldwin, M. A.; Falick, A. M.; Juhasz,<br />

P.; Vestal, M. L.; Burlingame, A. L. Anal. Chem. 2000, 72, 552–558.<br />

(42) Shenar, N.; Sommerer, N.; Martinez, J.; Enjalbal, C. J. Mass Spectrom. 2009,<br />

44, 621–632.<br />

(43) Schwartz, J. C.; Senko, M. W.; Syka, J. E. J. Am. Soc. Mass Spectrom. 2002,<br />

13, 659–669.<br />

(44) Douglas, D. J. Mass Spectrom. Rev. 2009, 28, 937–960.<br />

collision activation in the TOF/TOF-CID experiment (Figure 4)<br />

compared to the LTQ-CID experiment (Figure 1A), a significant<br />

series of b and y ion signals resulting from the cleavage of peptide<br />

backbone amide bonds were additionally observed in the former.<br />

ISD-MS/MS data of primary product ions observed in Figure 4 for<br />

signals at m/z 943.6 (corresponding to a CNL of 129 u) and m/z<br />

969.6 (corresponding to a CNL of 103 u) are comparable to those of<br />

the fragmentation product of a Munc13-1 cross-link, in which Lys-4<br />

is connected with Lys-13 (Figure 3). Also, backbone fragmentation<br />

of the peptide was observed in MALDI-ISD experiments.<br />

MALDI-TOF/TOF data of a PPARR peptide comprising amino<br />

acids 259-272 exhibit a fragmentation behavior analogous to that<br />

in an ESI-CID-MS 2 experiment (Supporting Information, Figure<br />

S3).<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6965


Figure 4. MALDI-TOF/TOF product ion spectrum of the type 0 modified Munc13-1 peptide (peptide amino acids 3-9) [M + H + BuUrBu-OH] +<br />

at m/z 1072.6.<br />

Intrapeptide Cross-Linking Products (Type 1). Also for<br />

intrapeptide cross-links, MALDI-TOF/TOF data were comparable<br />

to ESI-CID-MS 2 data for Munc13-1 (data not shown) and a<br />

PPARR peptide (Supporting Information, Figure S4).<br />

Interpeptide Cross-Linking Products (Type 2). For interpeptide<br />

cross-links 2 within the Munc13-1 peptide, MALDI-TOF/<br />

TOF experiments impressively demonstrated the fragility of the<br />

urea cross-linker upon high-energy CID, which is the reason for<br />

the increased abundance of the two 26 u doublets of product ions<br />

dominating the product ion spectrum of the type 2 modified<br />

precursor ion at m/z 1789.2 (Figure 5, Table 1). According to this<br />

spectrum, the fragmentation characteristics of 1 seem to be<br />

perfectly suited for MALDI-TOF/TOF applications, being even<br />

better than for low-energy MS 2 experiments in a quadrupole ion<br />

trap (Figure 3). The two pairs of 26 u mass shifted product ions<br />

reveal that the R-peptide and �-peptide comprising the cross-linked<br />

product are modified either with 4-aminobutyric acid or with 1,3oxazepan-2-one.<br />

Additionally conducted MALDI-ISD-MS/MS experiments<br />

confirmed the amino acid sequences of both peptides<br />

connected via the BuUrBu linker. This fingerprint feature of two<br />

26 u doublets evidencing the interpeptide cross-link was also found<br />

in the MALDI-TOF/TOF fragment ion spectrum of an interpeptide<br />

(type 2) modified PPARR peptide (Lys-208 connected with Lys-<br />

216) as the most abundant signals in the mass spectrum (Figure<br />

S5, Supporting Information; Table 1). A full list of cross-linking<br />

products from PPARR, which were identified by nano-HPLC/<br />

MALDI-TOF/TOF-MS/MS, is presented in the Supporting Information<br />

(Table S1).<br />

CONCLUSIONS<br />

In this paper, we describe the synthesis and application of a<br />

novel dissociative amine-reactive cross-linker that shows enhanced<br />

6966 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

fragmentation capabilities in both ESI-CID-ion trap-MS n and<br />

MALDI-TOF/TOF-CID-MS/MS and ISD-MS/MS and as such<br />

possesses a versatile applicability. The reactivity and the<br />

kinetics of our novel cross-linker were found to be comparable<br />

with that of other NHS esters, such as BS 3 . The cross-linker 1<br />

contains a urea moiety allowing for a facilitated cleavage upon<br />

CID. The key step in the fragmentation mechanism of the<br />

cross-linker is a nucleophilic attack at the urea carbonyl group,<br />

enabling a highly effective cleavage of the urea moiety. The<br />

synthesis of the cross-linker is simple compared to that of the<br />

previously published “Edman” cross-linker 13 and is achieved<br />

in two steps with inexpensive starting materials. The BuUrBu<br />

structure is symmetric, and hence, assignment of cross-linked<br />

product ions is facilitated compared to the rather complicated<br />

product ion spectra obtained with the previously published<br />

unsymmetric tandem MS cleavable linker. 13 The effective<br />

fragmentation properties of the CID cleavable cross-linker<br />

safeguarding for a selective identification of modified peptides<br />

by tandem MS improves the sensitivity of chemical crosslinking<br />

for protein structure analysis. This is highly advantageous<br />

for discriminating cross-linked species compared to<br />

isotope-labeled, i.e., deuterated linkers, which are employed<br />

as 1:1 mixtures of nondeuterated and deuterated derivatives,<br />

and in constrast to those, no decrease in MS signal intensity<br />

is observed for CID labile cross-linkers.<br />

Especially during MALDI-TOF/TOF experiments, higher quality<br />

MS/MS data were obtained compared to those for conventional<br />

noncleavable NHS esters, such as BS 3 . From CID MS 2 product<br />

ion mass spectra, the type of cross-linking present is readily<br />

visible: Peptides that are modified with a hydrolyzed linker<br />

(dead-end or type 0 cross-links) deliver a single pair of product<br />

ions mass shifted by 26 u due to the loss of 103 and 129 u


Figure 5. (A) MALDI-TOF/TOF product ion spectrum of the type 2 modified Munc13-1 peptide at m/z 1789.2. (B) Upper panel: ISD-MS 2<br />

product ion spectrum of the �-peptide [� + H + Bu] + at m/z 819.4. Lower panel: ISD-MS/MS product ion spectrum of the �-peptide [� + H +<br />

BuUr] + at m/z 845.5. (C) Upper panel: ISD-MS 2 product ion spectrum of the R-peptide [R +H + Bu] + at m/z 943.5. (C) Lower panel: ISD-MS/MS<br />

product ion spectrum of the R-peptide [R +H + BuUr] + at m/z 969.5.<br />

(Table 1). The resulting 26 u product ion pair (“doublet”) greatly<br />

simplifies identification of the respective cross-linking type and<br />

therefore allows for an automated analysis of tandem MS data. A<br />

CNL of 85 u (pyrolidinone) indicates an intrapeptide cross-link<br />

(type 1), whereas an interpeptide (type 2) cross-link delivers two<br />

sets of 26 u doublets originating from both R- and �-peptides, each<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6967


eing modified by the cross-linker fragments Bu and BuUr. A<br />

summary of the characteristic fingerprint product ions and neutral<br />

losses that serve for a discrimination of the different cross-linking<br />

types is presented in Table 1. Additionally conducted ISD-MS/<br />

MS and ion trap-MS 3 experiments clearly confirm the amino<br />

acid sequences of cross-linked peptides and allow reliable<br />

structure identification. Our cleavable cross-linker 1 allows an<br />

automated identification of cross-links combined with a database<br />

search based on the characteristic constant neutral losses<br />

as well as the abundant 26 u mass shifted ion pairs. Thus, highabundance<br />

unmodified peptides are excluded from further<br />

analysis and allow a highly efficient analysis of cross-linked<br />

products. Our novel cross-linker is expected to have a bright<br />

future for protein structure analysis.<br />

ACKNOWLEDGMENT<br />

M.Q.M. is supported by the DFG Graduiertenkolleg 1026<br />

“Conformational Transitions in Macromolecular Interactions” at<br />

the Martin-Luther-Universität Halle-Wittenberg. Dr. O. Jahn kindly<br />

provided the Munc13-1 peptide. A.S. gratefully acknowledges<br />

financial support by the DFG and the BMBF (ProNet-T 3 project).<br />

M.S. and F.D. gratefully acknowledge financial support by the<br />

DFG.<br />

APPENDIX<br />

Glossary<br />

aa amino acid<br />

ACN acetonitrile<br />

6968 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

BS 3 bis(sulfosuccinimidyl) suberate<br />

CID collision-induced dissociation<br />

CNL constant neutral loss<br />

Da Dalton<br />

DCC N,N′-dicyclohexylcarbodiimide<br />

DMSO dimethyl sulfoxide<br />

ESI electrospray ionization<br />

FA formic acid<br />

HCCA R-cyano-4-hydroxycinnamic acid<br />

HEPES 4-(2-hydroxyethyl)piperazine-1-ethanesulfonic acid<br />

HPLC high-performance liquid chromatography<br />

ISD in-source decay<br />

LID laser-induced dissociation<br />

LTQ linear quadrupole ion trap (Thermo Fisher)<br />

MALDI matrix-assisted laser desorption/ionization<br />

MS mass spectrometry<br />

NHS N-hydroxysuccinimide<br />

RT room temperature<br />

TCEP tris(2-carboxyethyl)phosphine<br />

TFA trifluoroacetic acid<br />

Th Thomson (m/z)<br />

TOF time-of-flight<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in the text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 13, 2010. Accepted July 8, 2010.<br />

AC101241T


Anal. Chem. 2010, 82, 6969–6975<br />

Analyte Discrimination from Chemiresistor<br />

Response Kinetics<br />

Douglas H. Read* and James E. Martin<br />

Sandia National Laboratories, P.O. Box 5800, Albuquerque, New Mexico 87185-1415<br />

Chemiresistors are polymer-based sensors that transduce<br />

the sorption of a volatile organic compound into a resistance<br />

change. Like other polymer-based gas sensors that<br />

function through sorption, chemiresistors can be selective<br />

for analytes on the basis of the affinity of the analyte for<br />

the polymer. However, a single sensor cannot, in and of<br />

itself, discriminate between analytes, since a small concentration<br />

of an analyte that has a high affinity for the<br />

polymer might give the same response as a high concentration<br />

of another analyte with a low affinity. In this paper<br />

we use a field-structured chemiresistor to demonstrate<br />

that its response kinetics can be used to discriminate<br />

between analytes, even between those that have identical<br />

chemical affinities for the polymer phase of the sensor.<br />

The response kinetics is shown to be independent of the<br />

analyte concentration, and thus the magnitude of the<br />

sensor response, but is found to vary inversely with the<br />

analyte’s saturation vapor pressure. Saturation vapor<br />

pressures often vary greatly from analyte to analyte, so<br />

analysis of the response kinetics offers a powerful method<br />

for obtaining analyte discrimination from a single sensor.<br />

Polymer sorption is the basis of most simple methods of<br />

sensing vapors of volatile organic compounds. Such devices<br />

include quartz crystal microbalances and surface acoustic wave<br />

sensors that transduce mass sorption into a frequency change 1-4<br />

and methods that transduce mass uptake into a resistance or<br />

capacitance change, such as chemiresistors, chemicapacitors, and<br />

CHEMFETs. 5-13 Regardless of the sensing mechanism, each<br />

* To whom correspondence should be addressed. E-mail: dhread@sandia.gov.<br />

Phone: (505) 844-5338. Fax: (505) 844-4045.<br />

(1) Hierlemann, A.; Ricco, J.; Bodenho, K.; Dominik, A.; Golpel, W. Anal. Chem.<br />

2000, 72, 3696–3708.<br />

(2) Wibawa, G.; Takahashi, M.; Sato, Y.; Takishima, S.; Masuoka, H. J. Chem.<br />

Eng. Data 2002, 47, 518–524.<br />

(3) Ho, C. K.; Lindgren, E. R.; Rawlinson, K. S.; McGrath, L. K.; Wright, J. L.<br />

Sensors 2003, 3, 236–247.<br />

(4) Fang, M.; Vetelino, K.; Rothery, M.; Hines, J.; Frye, G. C. Sens. Actuators,<br />

B 1999, 56, 155–157.<br />

(5) Janata, J. Electroanalysis 2004, 16 (22), 1831–1835.<br />

(6) Wilson, M. W.; Hoyt, S.; Janata, J.; Booksh, K.; Obando, L. IEEE Sens. J.<br />

2001, 1 (4), 256–274.<br />

(7) Patel, S. V.; Mlsna, T. E.; Fruhberger, B.; Klaassen, E.; Cemalovic, S.; Baselt,<br />

D. R. Sens. Actuators, B 2003, 96, 541–553.<br />

(8) Donaghey, L. F. Resistive Hydrocarbon Leak Detector. U.S. Patent<br />

4,631,952, Dec 30, 1986.<br />

(9) Lewis, N. S.; Doleman, B. J.; Briglin, S.; Severin, E. J. Colloidal Particles<br />

Used in Sensing Arrays. U.S. Patent 6,537,498, March 25, 2003.<br />

(10) Lundberg, B.; Sundqvist, B. J. Appl. Phys. 1986, 60 (3), 1074–1079.<br />

(11) Lei, H.; Pitt, W. G.; McGrath, L. K.; Ho, C. K. Sens. Actuators, B 2004,<br />

101, 122–132.<br />

individual sensor, consisting of a single polymer, cannot discriminate<br />

between analytes if only the equilibrium mass uptake is used,<br />

unless somehow the partial pressure of the analyte is either known<br />

or measured. For these polymer-based sensors, analyte discrimination<br />

is currently based on the artificial nose concept, wherein<br />

arrays of sensors having differentiating chemical affinities are<br />

exposed to the vapor. 13-16 Any analyte will then give a more-orless<br />

unique relative equilibrium mass uptake to the array elements,<br />

generating a response fingerprint. This equilibrium approach can<br />

enable the discrimination of analytes having disparate chemical<br />

affinities, but will not be as useful for distinguishing homologous<br />

analytes, such as octane from decane or xylene from mesitylene.<br />

The ability to distinguish between homologous analytes requires<br />

nonequilibrium information.<br />

In this study we use magnetic field-structured chemiresistors<br />

to show that polymer sorption kinetics enables discrimination<br />

between even homologous analytes. The basis for this discrimination<br />

derives in part from Flory-Huggins theory, which shows that,<br />

for analytes having the same chemical affinity for a particular<br />

polymer, the analyte’s equilibrium mass sorption is determined<br />

by the analyte activity alone. 17 This chemical affinity is quantified<br />

by the Flory parameter, �, and activity is defined as the ratio of<br />

the analyte vapor pressure to its saturation vapor pressure, or a<br />

=P/P*. For linear alkanes the saturation vapor pressure decreases<br />

by about a factor of 3 for every additional carbon, so at the same<br />

activity, octane vapor will have roughly 10 times the number<br />

density of molecules as decane, yet will lead to about the same<br />

equilibrium polymer swelling. The flux of the analyte into the<br />

polymer, and therefore the mass transport into the chemiresistor,<br />

is proportional to its diffusivity times the analyte number density,<br />

and the latter can be obtained from equilibrium thermodynamics.<br />

Therefore, we expect swelling will be roughly 10 times faster for<br />

octane than for decane, provided their diffusivities are similar. If<br />

the analyte diffusivities are similar, then to first-order approximation,<br />

the characteristic swelling rate should simply be proportional<br />

to the analyte’s saturation vapor pressure. This swelling time is<br />

expected to be independent of the analyte’s concentration. This<br />

is because when in the linear swelling regime (a valid assumption<br />

(12) Ho, C. K.; Hughes, R. C. Sensors 2002, 2, 23–34.<br />

(13) Doleman, B. J.; Lonergan, M. C.; Severin, E. J.; Vaid, T. P.; Lewis, N. S.<br />

Anal. Chem. 1998, 70 (19), 4177–4190.<br />

(14) Grate, J. W. Chem. Rev. 2008, 108, 726–745.<br />

(15) Kim, Y. S.; Ha, S. C.; Yang, Y.; Kim, Y. J.; Cho, S. M.; Yang, H.; Kim, Y. T.<br />

Sens. Actuators, B 2005, 108, 285–291.<br />

(16) Eastman, M. P.; Hughes, R. C.; Yelton, G.; Ricco, A. J.; Patel, S. V.; Jenkins,<br />

M. W. J. Electrochem Soc. 1999, 146 (10), 3907–3913.<br />

(17) Flory, P. J. Principles of Polymer <strong>Chemistry</strong>; Cornell University: Ithaca, NY,<br />

and London, 1953; pp 495-514.<br />

10.1021/ac101259w © 2010 American <strong>Chemical</strong> Society 6969<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/21/2010


Figure 1. Response of an FSCR to a 2000 s exposure of p-xylene<br />

at an activity of 0.019. The conductance drops exponentially and then<br />

recovers on the same time scale during the nitrogen purge.<br />

for most vapor detection applications), twice the analyte vapor<br />

pressure gives twice the swelling, yet also gives twice the number<br />

density of analyte molecules, and thus twice the diffusive flux. A<br />

swelling time that is independent of the analyte concentration is<br />

highly desirable, since it would simplify the interpretation of<br />

kinetics data. In fact, the raw data we collect are sensor response<br />

versus time, so the response must be converted to polymer<br />

swelling before the swelling time can be computed. This is<br />

accomplished using the transduction curve of the particular<br />

chemiresistor, which is the relation between its change in<br />

conductance and equilibrium swelling. In a previous paper we have<br />

shown that this transduction curve is independent of the analyte,<br />

and so can be determined from testing the chemiresistor with<br />

any particular analyte, regardless of its affinity for the polymer. 18,19<br />

The kinetics data we present were taken with the fieldstructured<br />

chemiresistors (FSCRs) we have developed over the<br />

past several years. 18-23 These sensors differ from traditional<br />

carbon black chemiresistors in that the particle phase is composed<br />

of Au-coated Ni particles that are structured into conducting<br />

pathways using magnetic fields. Field-structuring consistently<br />

brings the particle phase to a conducting threshold over a wide<br />

range of particle volume fractions, and the Au-coated particles do<br />

not adsorb typical VOCs. As a result, FSCRs have significantly<br />

increased baseline stability, sensitivity, reversibility, and sensorto-sensor<br />

reproducibility. 18<br />

BACKGROUND<br />

Response Curve. The response of a chemiresistor exposed<br />

to a step increase in analyte concentration is given in Figure 1.<br />

The conductance drops to its equilibrium value as the polymer<br />

swells and then recovers on the same time scale as the polymer<br />

deswells. FSCR response is defined as the conductance ratio G/G0,<br />

where G and G0 are the conductances in the presence and<br />

absence of analyte, respectively. Plotting the equilibrium value<br />

of G/G0 as a function of analyte concentration gives a sigmoidal<br />

(18) Read, D. H.; Martin, J. E. Adv. Funct. Mater. 2010, 20 (11), 1–8.<br />

(19) Read, D. H.; Martin, J. E. Anal. Chem. 2010, 82 (12), 5373–5379.<br />

(20) Martin, J. E.; Anderson, R. A.; Odinek, J.; Adolf, D.; Williamson, J. Phys.<br />

Rev. B 2003, 67, 094207.<br />

(21) Martin, J. E.; Hughes, R. C.; Anderson, R. A. Sensor Devices Comprising<br />

Field Structured Composites. U.S. Patent 6,194,769, Feb 27, 2001.<br />

(22) Martin, J. E.; Hughes, R. C.; Anderson, R. A. Field-Structured Material Media<br />

and Methods for Synthesis Thereof. U.S. Patent 6,290,868, Sept 18, 2001.<br />

(23) Read, D. H.; Martin, J. E. Anal. Chem. 2010, 82, 2150–2154.<br />

6970 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 2. The FSCR equilibrium conductance versus analyte activity<br />

plot forms a sigmoidal curve, which is parametrized by Γ and the<br />

response midpoint, a1/2. For example, the acetone curve has a<br />

response midpoint of ∼0.15, and the detection range is from an<br />

activity of ∼0.01 to an activity of 0.36. Response curves for this single<br />

FSCR exposed to various analytes illustrate selectivity for more<br />

hydrophobic analytes. As evident, these equilibrium data do not<br />

discriminate between chemically similar analytes.<br />

equilibrium response curve, as in Figure 2. Because the relative<br />

resistance change, ∆R/R0, is exponential with the analyte<br />

concentration, the relative conductance can be fit by 18<br />

G<br />

)<br />

G [ 0<br />

1 + eΓa/a1/2 - 1<br />

e Γ -1<br />

- 1 ]<br />

Here a is the analyte vapor activity, Γ is a fit parameter related to<br />

the abruptness of the conductor-insulator transition (typically<br />

∼4), and a1/2 is the response midpoint, defined as the analyte<br />

activity that reduces G0 by half.<br />

Figure 2 shows that a poly(dimethylsiloxane) chemiresistor<br />

is much more sensitive to the hydrophobic analytes toluene,<br />

p-xylene, mesitylene, and undecane than to the hydrophilic analyte<br />

acetone. In fact, the response midpoint for propanol is nearly 10×<br />

that of toluene or p-xylene. Although this polymer is selective for<br />

hydrophobic analytes, the equilibrium response cannot be used<br />

to discriminate between different analytes unless the analyte<br />

activity is known. More typically, these data would be used to<br />

determine the activity of a known analyte.<br />

Transduction Curve. We have previously shown that the<br />

response of an FSCR is a universal function of polymer swelling,<br />

regardless of the chemical nature of the analyte. 18,19 The response<br />

curve in Figure 2 can thus be thought of as a combination of a<br />

solely device-dependent transduction curve (conductance as a<br />

function of swelling) and the solely analyte-dependent mass<br />

sorption isotherm that relates polymer swelling to the vapor<br />

concentration. The transduction curve for a typical sensor is shown<br />

in Figure 3, with the volume fraction of absorbed analyte<br />

determined gravimetrically. This curve is given by<br />

G<br />

)<br />

G [ 0<br />

1 + eΓφ/φ1/2 - 1<br />

e Γ -1<br />

- 1 ]<br />

where φ is the volume fraction of absorbed analyte and φ1/2 is<br />

the response midpoint. The transduction curve is parametrized<br />

by φ1/2 and Γ, and these are strongly dependent on the<br />

fabrication process used to make the sensor. 18,19,23 To a good<br />

(1)<br />

(2)


Figure 3. Equilibrium response as a function of the volume fraction<br />

of absorbed analyte, giving a master transduction curve that is<br />

dependent only on the characteristics of the particular sensor. (The<br />

data for pentane, octane, decane, TMP, TBT, and undecane are<br />

reprinted from Figure 8 of ref 19. Copyright 2010 American <strong>Chemical</strong><br />

Society.) Here the response midpoint, φ1/2, is 5.64 × 10 -3 , and Γ is<br />

2.75. a1/2 and � values for each analyte are reported in ref 19.<br />

approximation, the equilibrium sorption is proportional to the<br />

analyte activity, a, and is given by the linearized Flory-Huggins<br />

equation 17,19<br />

φ ) ae -(�+1)<br />

Here � is the Flory interaction parameter, which is a measure of<br />

an analyte’s affinity for the polymer. This transduction curve is of<br />

great importance in this paper, as we will use it to convert<br />

nonequilibrium response curves, such as that in Figure 1, into<br />

the nonequilibrium sorption curves from which we obtain kinetics<br />

data.<br />

EXPERIMENTAL SECTION<br />

Chemiresistor Fabrication. The chemiresistors used in this<br />

research consist of five identically fabricated FSCRs. The FSCR<br />

composite is composed of 15 vol % 3-7 µm gold-plated carbonylnickel<br />

particles (Goodfellow Inc., product no. NI06021) and a<br />

two-part, addition-cure poly(dimethylsiloxane) (PDMS) (Gelest<br />

Inc. optical encapsulant 41, PP2-OE41). The particles are immersion-gold-plated<br />

using Enthone Inc. Lectroless Prep (PCN 210004-<br />

001). The particles are mixed in the PDMS precursors, and a<br />

volume of hexane equal to the volume of PDMS is then added to<br />

thin the viscous composite precursor. A ∼5 µL volume of solventcast<br />

composite precursor is then deposited onto a glass substrate,<br />

forming a ∼3 mm diameter film, which spansa1mmgapbetween<br />

the two gold electrode pads. The sensors are cured at 40 °C for<br />

24hina∼750 G uniaxial magnetic field. The composite is placed<br />

in the magnetic field such that a conductive chainlike particle<br />

network is formed across the two electrodes. 18,22<br />

Analytes. The model analytes used in this research are<br />

acetone, toluene, p-xylene, mesitylene, and undecane. All of the<br />

analytes are from either Fisher <strong>Chemical</strong>s (ACS certified reagent<br />

grade) or Aldrich <strong>Chemical</strong>s (ReagentPlus grade). Pertinent<br />

physical data for the analytes (saturation vapor pressures and �<br />

(3)<br />

Table 1. Room Temperature Saturation Vapor<br />

Pressures 24 and � Parameters for the Analytes 19<br />

analyte P*(25 °C) (Torr) �<br />

toluene 28.97 1.268<br />

p-xylene 8.80 1.258<br />

mesitylene 2.55 1.424<br />

acetone 228.19 2.226<br />

undecane 0.39 1.126<br />

parameters) are included in Table 1. 19,24 The aromatic compounds<br />

toluene, xylene, and mesitylene and undecane were chosen for<br />

use as analytes due to their similar � parameters for with PDMS<br />

and dissimilar vapor pressures.<br />

Sensor Testing. The sensors are enclosed in a shielded flow<br />

cell with gas inlet and outlet ports and electrical throughputs.<br />

Known concentrations of analyte vapors are produced by mixing<br />

a controlled flow rate of analyte-saturated nitrogen from a<br />

temperature-controlled bubbler system with a controlled flow of<br />

pure nitrogen. 19 Conductance measurements are made by applying<br />

a constant 10 mV dc voltage across the electrodes with a power<br />

supply (Hewlett-Packard model 6552A) and measuring the current<br />

through the chemiresistors with picoammeters (Keithley Instruments<br />

Inc. model 6485).<br />

Sorption Kinetics. To study diffusion, it is necessary to<br />

determine analyte mass sorption as a function of time. This is<br />

accomplished by using the transduction curve in eq 2 to transform<br />

the FSCR response curves, such as that in Figure 1, into the<br />

volume fraction of absorbed analyte. Solving eq 2 for φ(t) gives<br />

φ(t) ) φ1/2 Γ ln[ 1 + (eΓ - 1) G0 - G(t)<br />

G(t) ] (4)<br />

Using the equilibrium transduction curve to relate the nonequilibrium<br />

response to the nonequilibrium sorption is an approximation,<br />

since it is an unproven assumption that the sensor<br />

conductance depends only on the total mass sorption and is<br />

insensitive to the swelling gradients that accompany diffusion. For<br />

direct mass uptake experiments gradients pose no problems, but<br />

gradients could potentially foil our approach, yet as we show in<br />

the following section, we obtain a response time that is clearly<br />

independent of the analyte activity. Perhaps this assumption is<br />

reasonable because gradients would have no effect except to<br />

second order: If the equilibrium sensor conductance is locally<br />

linear in the analyte activity, then only the curvature of the gradient<br />

would lead to a conductance response that is different from the<br />

zero gradient response at the same analyte uptake. The utility of<br />

this approximation can only be judged by the quality of the final<br />

experimental results shown in the following section.<br />

Figure 4 illustrates a time-dependent mass sorption curve<br />

obtained from the sensor response. In this figure the equilibrium<br />

swelling, calculated from applying eq 4 to the raw conductance<br />

data, differs from the prediction of the linearized Flory-Huggins<br />

equation (eq 3) by ∼10%. This experimental error does not cause<br />

an error in the computed sorption time. Diffusion into a finite slab<br />

(24) Yaws, C. L.; Narasimhan, P. K.; Gabbula, C. Yaws’ Handbook of Antoine<br />

Coefficients for Vapor Pressure [Online], 1st electronic ed.; Knovel: New York,<br />

2005. URL: http://www.knovel.com/web/portal/browse/display?_EXT_<br />

KNOVEL_DISPLAY_bookid)1183, accessed 3/5/2009.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6971


Figure 4. The volume fraction of absorbed analyte, φ, approaches<br />

its asymptotic value exponentially upon exposure to an analyte, in<br />

this case p-xylene vapors, at an activity of 0.054.<br />

Figure 5. The measured characteristic response time, τmeas, isthe<br />

y intercept of the line obtained by plotting A ) ∫0 t ∆φ(s)/φ∞ ds as a<br />

function of B ) ∆φ(t) ) φ∞ - φ(t).<br />

is a mathematically complex problem, but for practical purposes<br />

the kinetic data can be fit by the simple exponential expression<br />

φ(t) ) φ ∞ (1 - e -t/τmeas ) (5)<br />

Here t is time and τmeas is the measured response time of the<br />

chemiresistor to the analyte in question. Fitting to this form<br />

will prove useful in correcting the observed kinetic data for<br />

the time it takes for the analyte concentration to reach the<br />

steady state in the flow cell. Despite this, we can actually obtain<br />

the response time in a model-independent fashion. To do so,<br />

we simply plot the integral A ) ∫0 t ∆φ(s)/φ∞ ds against B )<br />

∆φ(t)/φ∞, where ∆φ(t) ≡ φ∞ - φ(t). We can operationally define<br />

τmeas as the y intercept of this curve in the limit as ∆φ(t) f 0,<br />

but if eq 5 is a reasonable fit, the result will be a straight line<br />

whose y intercept is easily obtained. The data in Figures 5 are<br />

indeed linear, so eq 5 is actually quite a good description of the<br />

raw kinetic data.<br />

RESULTS<br />

In the following we give the theoretical expression for the true<br />

sorption kinetics, which is shown to be independent of the analyte<br />

activity. We then develop an expression for the measured sorption<br />

kinetics, which is a convolution of the true sorption kinetics and<br />

the kinetics of filling the flow cell. We then show how the flow<br />

cell fill kinetics can be determined from the measured sorption<br />

kinetics and how the fill time can be used to extract the true<br />

6972 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

sorption time from the measured time. Both the measured and<br />

true sorption kinetics are shown to be independent of the activity,<br />

but strongly dependent on the analyte saturation vapor pressure.<br />

Predicted Sorption Time. From Raoult’s law, the partial<br />

pressure of a chemical species, P, in a gas is equal to its volume<br />

fraction, φ vap, times the total pressure, or φvap ≡ V/Vtot ) P/Ptot.<br />

Recall that the analyte activity is a ) P/P*; therefore, a ) φvap/<br />

φvap*, where φvap* is the volume fraction of analyte vapor at<br />

saturation. From the linearized form of the Flory-Huggins<br />

equation (eq 3), φ ) a/e 1+� , the partition coefficient, K, is then<br />

K ≡ φ 1<br />

)<br />

φvap φvap *e 1+�<br />

At high inlet stream fluxes the response time of an FSCR is<br />

diffusion limited, and from Fick’s second law of diffusion for the<br />

case of a semi-infinite slab, the sensor’s response time can be<br />

expressed as<br />

τd ∝ K d2 1<br />

)<br />

Dt Dtφvap *e 1+�d2<br />

(6)<br />

(7a)<br />

where d is the thickness of the composite and Dt is the diffusion<br />

coefficient of the analyte into the silicone. Note that even for<br />

analytes of identical chemical affinity there is discrimination<br />

based on sorption kinetics, due to variations in their diffusivity<br />

and saturation volume fraction. From the ideal gas law, the<br />

saturation vapor pressure is given by P* )Fφvap*RT/Mw, so<br />

the sorption time can also be written as<br />

τ d ∝ RT<br />

P*<br />

F<br />

M w D t e 1+�d2<br />

(7b)<br />

The saturation vapor pressure varies over a wide range, so the<br />

sorption kinetics should be a very useful method of discrimination.<br />

Measured Sorption Time. The goal is to determine the true<br />

sorption time, including any dependence this time might have on<br />

the analyte activity. Unfortunately, the measured sorption time is<br />

a convolution of two factors: the time it takes for the analyte<br />

activity in the flow cell to reach its steady-state value (flow cell<br />

fill time) and the true mass sorption time. In a constant-flow-rate<br />

apparatus, and for any particular analyte, neither of these factors<br />

should be dependent on the analyte activity, so the measured<br />

sorption time should be independent of the analyte activity. The<br />

data in Figure 6 show that this is indeed the case and also show<br />

significant differences in the sorption kinetics for different analytes,<br />

as expected. At low concentrations (where the slope of the<br />

response curve is low) and for long response times, sensor drift<br />

becomes an issue and response times are difficult to reliably<br />

measure; such is the case for low concentrations of undecane.<br />

This issue, however, can be addressed by using a sensor with<br />

high sensitivity at low analyte concentrations. 23<br />

The data in Figure 6 are an average for five sensors, each<br />

tested simultaneously in the same flow cell. Because of sensorto-sensor<br />

variations in polymer thickness, each chemiresistor had<br />

a somewhat different response time. For example, the five sensors<br />

exposed to mesitylene at an activity of 0.058 have an average<br />

response time, τmeas,of48± 13 s. To account for these thickness


Figure 6. The average measured response time, τmeas, is independent<br />

of the volume fraction of absorbed analyte, φ, and thus the<br />

analyte activity. This response time, however, is strongly analyte<br />

dependent, enabling discrimination between analytes having nearly<br />

identical chemical affinities such as toluene and undecane. At low<br />

concentrations (where the slope of the response curve is low) and<br />

for long response times, sensor drift becomes an issue and response<br />

times are difficult to reliably measure; such is the case for low<br />

concentrations of undecane.<br />

variations, we normalized the response time of each sensor to<br />

agree with that of an arbitrarily chosen reference sensor, by<br />

exposing the sensors to a mesitylene activity of 0.058, which<br />

resulted in a swelling of 0.005. These renormalized time scales<br />

were then used to determine both the average response time<br />

to an analyte and the associated measurement error in terms<br />

of the standard deviation. The dimensionless correction factors<br />

ranged from 0.6 to 1.3.<br />

To determine if it is a reasonable assumption that the variation<br />

in the measured response time is caused by variations in sensor<br />

thickness, we made 10 chemiresistors on a glass slide using<br />

methods and materials identical to those for the originals. Sensor<br />

thickness measurements, made with a Nikon mm-800 measuring<br />

microscope with a quadra-chek 200 advanced digital readout<br />

system, give a sensor thickness of 216 ± 26 µm. The average<br />

response time, τ meas, of the original five sensors to mesitylene<br />

at an activity of 0.058 was 48 ± 13 s. Using this average response<br />

time and the average composite thickness of the newly made<br />

sensors (i.e., 216 µm), we calculate a measured diffusion<br />

coefficient of 4.5 × 10 -2 cm 2 /s. From this diffusion coefficient<br />

we computed the predicted response times for the 10 new<br />

sensors, yielding 48 ± 8 s. Therefore, the variation in response<br />

times that we measured for the original sensors is similar to<br />

the variation in response time that we predict from the newly<br />

made sensors. It should be noted that this measured diffusion<br />

coefficient is not a true diffusion coefficient because it is<br />

calculated from τmeas, which is a convolution of the actual<br />

diffusion time scale and the time scale from the flow cell fill<br />

time as we will discuss in the following section.<br />

An accurate determination of the sorption kinetics ideally<br />

requires that the flow cell fill time is fast compared to the sorption<br />

kinetics. In the following we will determine this fill time from the<br />

measured sorption kinetics and use this value to extract the true<br />

sorption kinetics from the measured values. The true sorption<br />

time can then be compared to the flow cell fill time to determine<br />

whether this fast fill time condition is met in our experiments.<br />

Flow Cell Fill Time. After the inlet stream starts to deliver<br />

an analyte of fixed activity to the flow cell, the analyte activity<br />

rises continuously, due to the finite volume of pure nitrogen that<br />

must be displaced. The transient analyte activity can be obtained<br />

by modeling the flow cell as a continuously stirred tank and is of<br />

the form<br />

a(t) ) a ∞ (1 - e -t/τf ) (8a)<br />

In terms of the volume fraction of the vapor this is<br />

φ vap (t) ) φ vap (∞)(1 - e -t/τf ) (8b)<br />

The characteristic time for the flow cell to reach a steady-state<br />

concentration is given by τf ) Vf/F, where Vf is the volume of<br />

the flow cell and F is the volumetric flow rate of nitrogen from<br />

the inlet stream.<br />

Convolution of Time Scales. The measured composite<br />

swelling in Figure 4 is a convolution of the increase in the analyte<br />

vapor activity in the flow cell and the diffusive kinetics, which for<br />

simplicity we take to be of the form φ(t) ) φ∞(1 - e -t/τ d) for a<br />

step increase in the analyte activity at t ) 0. Using eq 5, the<br />

expression is<br />

t dφvap (t)<br />

φ(t) ) K∫ 0<br />

t)s dt [1 - e-(t-s)/τd ]ds (9)<br />

Using eq 8a to evaluate the derivative gives<br />

φ(t) ) Kφvap (∞) t<br />

τ ∫ e<br />

0<br />

f<br />

-s/τf -(t-s)/τd [1 - e ]ds (10)<br />

A straightforward integration leads to the final expression for the<br />

sorption kinetics”<br />

φ(t) ) Aφvap (∞)[ 1 - τd e<br />

τd - τf -t/τd +<br />

τ f<br />

e<br />

τd - τf -t/τf]<br />

(11)<br />

The measured lifetime can then be computed from this equation,<br />

and the surprisingly simple result is<br />

τmeas ) 1 ∞<br />

φ ∫ [1 - φ(t)] dt ) τ<br />

0<br />

d + τf ∞<br />

(12)<br />

The true diffusion time can thus be obtained from the measured<br />

time by τd ) τmeas s τf. (The case where τd = τf leads to division<br />

by zero in eq 11. This division looks troublesome, but can be<br />

handled by defining τd ) τf(1 + ε) and carefully taking the limit<br />

as ε f 0.)<br />

Corrected Sorption Times. The correction of the measured<br />

sorption times for the flow cell fill time requires a determination<br />

of the fill time. The fill time can be extracted from the measured<br />

sorption times themselves, through a limiting process, as we will<br />

now describe. To obtain accurate measured sorption times, we<br />

first average these measured times over all analyte activities to<br />

obtain a mean response time we call τjmeas. This averaging is valid,<br />

because the measured sorption time is independent of the<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6973


Figure 7. The measured sorption time, τmeas, is proportional to the<br />

inverse saturation vapor pressure. The nonzero y intercept of 13.8 s<br />

is the flow cell fill time, which when subtracted from the measured<br />

sorption time gives the true sorption time, τd. The sorption time of<br />

undecane is nearly 2 orders of magnitude larger than that of toluene<br />

despite their equilibrium response curves being nearly identical as in<br />

Figure 2.<br />

analyte activity as demonstrated in Figure 6. A plot of this<br />

average time versus the inverse analyte saturation vapor pressure<br />

results in a straight line with a nonzero y intercept, as seen in<br />

Figure 7. Of course, at infinite saturation vapor pressure the<br />

sorption time τd must be zero since the diffusive flux is infinite.<br />

Therefore, this finite intercept of 13.8 s corresponds to the fill<br />

time of the flow cell and can be used to produce the corrected<br />

sorption times, τd, in Figure 7. This measured flow cell fill time<br />

is in good agreement with the theoretical prediction, using Vf/F.<br />

Figure 7 shows a nearly perfect proportionality between the<br />

sorption time, τd, and the analyte saturation vapor pressure, P*.<br />

Equation 7b shows that other analyte parameters, such as the<br />

diffusivity and the � parameter, also determine this sorption time,<br />

but apparently these combined factors are more-or-less constant<br />

for the analytes we tested. Also, the saturation vapor pressure<br />

varies strongly from analyte to analyte. If finer resolution was<br />

desired for differentiation between similarly volatile analytes, these<br />

other analyte parameters could be taken into account. The<br />

homologous series of aromatic analytes, all of which have nearly<br />

identical Flory parameters, and thus identical equilibrium chemiresistor<br />

responses as a function of activity, are now easily distinguished.<br />

For toluene, p-xylene, and mesitylene the sorption times<br />

are 3.7, 10.0, and 37.0 s, respectively, which are in good agreement<br />

with their relative reciprocal saturation vapor pressures, 3.45, 11.4,<br />

and 39.2 (×10 -2 Torr -1 ). These sorption time differences are<br />

large compared to the errors associated with our measurements,<br />

even though the sorption time of toluene is much faster<br />

than the fill time of 13.8 s. The measurement of faster sorption<br />

times is possible, but would require either higher flow rates<br />

and a smaller flow cell volume, so that the fill time is faster, or<br />

thicker sensors, to increase the sorption time. Of course,<br />

increased sorption time comes at the cost of increased sensor<br />

response time, so this would only be desireable if high-volatility<br />

analytes are to be measured.<br />

DISCUSSION<br />

In the pulsed flow experiments the determination of the<br />

sorption time is straightforward, but would require a test unit with<br />

an engineered flow system. This mechanical requirement is<br />

antithetical to the inexpensive and simple chemiresistor concept.<br />

6974 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Other approaches are possible, but are yet untested in the context<br />

of vapor sorption, though analogous approaches have been used<br />

to determine photoluminescence lifetimes. In fact, luminescence<br />

is a useful language in which to describe the other possibilities.<br />

First, a luminescence lifetime can be measured by suddenly<br />

subjecting a material such as a phosphor to steady illumination.<br />

The rise time of the photoemission is the luminescent lifetime.<br />

This approach is analogous to the sorption kinetics reported above.<br />

If instead the steady illumination is suddenly shut off, the emission<br />

decay gives the lifetime. This is analogous to analyzing the<br />

desorption kinetics shown in Figure 1. A third approach is to excite<br />

a phosphor with amplitude-modulated light. The photoemission<br />

will also be amplitude modulated, but shifted in phase, and the<br />

lifetime can be determined from the phase shift and the modulation<br />

frequency. This is analogous to modulating the activity of<br />

the analyte, which is not really practical. The fourth approach is<br />

to excite a phosphor with light that randomly fluctuates in<br />

intensity. The photoemission then also fluctuates in intensity, but<br />

these fluctuations are time correlated. The associated correlation<br />

time can be determined from the photoemission time autocorrelation<br />

function, which is easily computed with a simple shift<br />

register device. This fourth method is applicable to a sensor in<br />

an environment where there is nonstationary air movement that<br />

causes fluctuations in the delivered analyte activity. An engineered<br />

mechanical system needed to deliver a flow pulse can potentially<br />

be supplanted by an electronic chip that computes the correlation<br />

time of the sensor response.<br />

Of course, we cannot expect the activity fluctuations delivered<br />

to the sensor to be purely random. Instead, these activity<br />

fluctuations will themselves have a correlation time (which is an<br />

interesting issue in and of itself). Therefore, the measured<br />

response fluctuations will in general be a convolution of both the<br />

activity fluctuations and the sorption kinetics. Extracting the true<br />

sorption kinetics will require two sensors that have widely<br />

disparate response kinetics (e.g., a thick and a thin polymer film).<br />

The thin sensor will respond rapidly, relative to the correlation<br />

time of the activity fluctuations, to changes in analyte activity. This<br />

sensor will thus measure the correlation time of the activity<br />

fluctuations. The thick sensor will have a sorption time long<br />

compared to this correlation time, and its response will be largely<br />

determined by the sorption kinetics. The thin sensor can be used<br />

to correct this measured sorption time to obtain the true sorption<br />

time. Therefore, by standard time correlation techniques, discrimination<br />

based on sorption kinetics can be developed without<br />

an engineered flow system. This approach will be the subject of<br />

our future research.<br />

CONCLUSIONS<br />

We have used a field-structured chemiresistor to demonstrate<br />

that response kinetics can be used to discriminate between<br />

analytes, even between those that have identical chemical affinities<br />

for the polymer phase of the sensor. To do this, we have used<br />

the analyte-independent transduction curve (conductance versus<br />

polymer swelling) to transform the time-dependent sensor conductance<br />

into a time-dependent polymer swelling. From these<br />

swelling data we can determine the measured sorption time, which<br />

is a combination of the true sorption kinetics and the fill time of<br />

the flow cell. The true sorption kinetics is obtained by correcting


for the fill time, and we find that the sorption kinetics is<br />

independent of the analyte activity, but inversely proportional to<br />

the saturation vapor pressure of the analyte. Saturation vapor<br />

pressures vary greatly from analyte to analyte, making response<br />

kinetics a powerful method of analyte discrimination, even with a<br />

single sensor. Finally, we suggest that when the sensing environment<br />

presents fluctuations in the analyte activity, an analysis of<br />

the fluctuations in the sensor response fluctuations can be used<br />

to extract the sorption kinetics. This approach would obviate the<br />

need for an engineered flow system that can deliver an analyte<br />

pulse and will be the focus of our future work.<br />

ACKNOWLEDGMENT<br />

This work was supported by the Division of Materials Sciences<br />

and Engineering, Office of Basic Energy Sciences, U.S. Department<br />

of Energy. Sandia National Laboratories is a multiprogram<br />

laboratory operated by Sandia Corp., a Lockheed Martin Co., for<br />

the Department of Energy’s National Nuclear Security Administration<br />

under Contract DE-AC04-94AL85000.<br />

Received for review May 13, 2010. Accepted July 1, 2010.<br />

AC101259W<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6975


Anal. Chem. 2010, 82, 6976–6982<br />

Improvement of Sensitivity and Dynamic Range in<br />

Proximity Ligation Assays by Asymmetric<br />

Connector Hybridization<br />

Joonyul Kim, Jiaming Hu, Rebecca S. Sollie, and Christopher J. Easley*<br />

Department of <strong>Chemistry</strong> and Biochemistry, 179 <strong>Chemistry</strong> Building, Auburn University, Auburn, Alabama 36849<br />

The proximity ligation assay (PLA) is one of the most<br />

sensitive and simple protein assays developed to date, yet<br />

a major limitation is the relatively narrow dynamic range<br />

compared to other assays such as enzyme-linked immunosorbent<br />

assays. In this work, the dynamic range of PLA<br />

was improved by 2 orders of magnitude and the sensitivity<br />

was improved by a factor of 1.57. To accomplish this,<br />

asymmetric DNA hybridization was used to reduce the<br />

probability of target-independent, background ligation. An<br />

experimental model of the aptamer-target-connector<br />

complex (apt A-T-aptB-C20,PLA) in PLA was developed<br />

to study the effects of asymmetry in aptamer-connector<br />

hybridization. Connector base pairing was varied from<br />

the PLA standard of 20 total bases (C 20) to an asymmetric<br />

combination with 15 total bases (C15). The<br />

results of this model suggested that weakening the<br />

affinity of one side of the connector to one aptamer<br />

would significantly reduce target-independent ligation<br />

(background) without greatly affecting target-dependent<br />

ligation (signal). These predictions were confirmed<br />

using PLA with asymmetric connectors for detection<br />

of human thrombin. This novel, asymmetric PLA<br />

approach should impact any previously developed PLA<br />

method (using aptamers or antibodies) by reducing<br />

target-independent ligation events, thus generally improving<br />

the sensitivity and dynamic range of the assay.<br />

The proximity ligation assay 1 (PLA) is among the most<br />

sensitive protein assays developed to date. PLA can be considered<br />

a descendent of immuno-PCR methods, 2 in which antibodies are<br />

attached to oligonucleotides to allow the coupling of target binding<br />

with exponential amplification by polymerase chain reaction<br />

(PCR). PLA is a homogeneous assay, requiring no washing steps,<br />

with protein detection limits as low as tens of zeptomoles 1 (fM in<br />

1 µL samples). As shown in Figure 1A, the mode of action of PLA<br />

results in a four-part complex composed of the targeted protein<br />

analyte (T), two proximity probes (DNA-conjugated antibodies or<br />

oligonucleotide aptamers; aptA and aptB), and a connector<br />

oligonucleotide (C20,PLA). The three separate DNA molecules<br />

* To whom correspondence should be addressed. Phone: +1 334-844-6967.<br />

Fax: +1 334-844-6959. E-mail: chris.easley@auburn.edu.<br />

(1) Fredriksson, S.; Gullberg, M.; Jarvius, J.; Olsson, C.; Pietras, K.; Gústafsdóttir,<br />

S. M.; Ostman, A.; Landegren, U. Nat. Biotechnol. 2002, 20, 473–<br />

477.<br />

(2) Adler, M.; Wacker, R.; Niemeyer, C. M. Analyst 2008, 133, 702–718.<br />

6976 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 1. Schematics of PLA and model systems described in the<br />

text. (A) Standard PLA, in which a connector length of 20 bases<br />

(C20,PLA, orb ) 20) is used. Ligated signal or background complexes<br />

are not differentiated by qPCR. (B) Asymmetric PLA, in which<br />

connector lengths of 15-19 bases (Cb,PLA, where 15 e b < 20) are<br />

tested. (C) Schematic of the experimental model system used to<br />

predict signal and background in PLA methods. The “signal” complex<br />

is defined as Loop-Cb, whereas the “background” complex is defined<br />

as FreeA-Cb-FreeB. Model connectors were varied from C15 to C20<br />

in this study.<br />

(or moieties) are designed to hybridize into a structure that<br />

promotes a DNA ligase enzyme to covalently couple the two<br />

proximity probes. In the presence of target, the four-part<br />

complex is preferred over target-independent hybridization<br />

(three-part complex) due to a cooperative “proximity effect”<br />

that essentially increases the local concentration of the oligo-<br />

10.1021/ac101762m © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/23/2010


nucleotide tails. 3-5 Thus, the amount of ligation products is<br />

proportional to the amount of protein analyte, and these products<br />

can be made recognizable by PCR for exponential amplification<br />

and highly sensitive quantitation by qPCR. Due to its simplicity<br />

and sensitivity, PLA has found much interest of late and has been<br />

successfully utilized in a variety of applications, such as sensitive<br />

protein detection, 1,6 in situ analysis of protein localization, 7<br />

DNA-protein interactions, 8 clinical diagnostics, 9 and proteinprotein<br />

interaction assays. 10<br />

In the beginning stages of PLA method development, Landegren<br />

and co-workers determined the optimal concentrations of<br />

each proximity probe and the connector oligonucleotide that<br />

maximize the cooperative effect induced by target binding, 1 and<br />

they further proved that this cooperative effect was negatively<br />

related to the Kd of the proximity probes. 6 Major advantages of<br />

PLA are sensitivity, selectivity, and ease of use. Use of at least<br />

two proximity probes, with discrete binding sites on the same<br />

target, allows unprecedented limits of target detection (LOD)<br />

and better specificity compared to other assays with one probe.<br />

Furthermore, the assay requires no washing steps due to its<br />

homogeneous nature. However, PLA has the limitation of a<br />

relatively narrow dynamic range. The reason for this flaw is<br />

the necessity to use very low amounts of proximity probes<br />

(∼10 -16 mol) to minimize a target-independent hybridization<br />

and ligation (background). Such a low amount of proximity<br />

probes allows PLA to assay only up to ∼10 -15 mol (10 times<br />

more analytes than proximity probes) if the Kd of the proximity<br />

probe is less than 0.4 nM. 6<br />

Our goal in this study was to improve the dynamic range of<br />

PLA without sacrificing LOD. To address this issue, we made two<br />

hypotheses: (1) target-independent hybridization (background;<br />

Figure 1A, bottom) is inversely proportional to the Kd of the<br />

connector oligonucleotide at a given concentration of proximity<br />

probes, and (2) the width of the dynamic range is proportional<br />

to the amount of a proximity probe to be used. If confirmed,<br />

these hypotheses suggest that dynamic range and signal to<br />

background could be improved by simply reducing the number<br />

of bases in the connector. To test these hypotheses, first we<br />

developed and tested an experimental model of the proximity<br />

effect. This model suggested that asymmetric connectors with<br />

higher K d values could significantly decrease target-independent<br />

hybridization (background; Figure 1A, bottom). In turn,<br />

this would lead to higher signal-to-background ratios and allow<br />

the use of more proximity probes, which should increase the<br />

dynamic range of PLA. As described below, we successfully<br />

(3) Zhou, H. X. Biochemistry 2001, 40, 15069–15073.<br />

(4) Zhou, H. X. J. Mol. Biol. 2003, 329, 1–8.<br />

(5) Tian, L.; Heyduk, T. Biochemistry 2009, 48, 264–275.<br />

(6) Gullberg, M.; Gústafsdóttir, S. M.; Schallmeiner, E.; Jarvius, J.; Bjarnegård,<br />

M.; Betsholtz, C.; Landegren, U.; Fredriksson, S. Proc. Natl. Acad. Sci.<br />

U.S.A. 2004, 101, 8420–8424.<br />

(7) Söderberg, O.; Leuchowius, K. J.; Gullberg, M.; Jarvius, M.; Weibrecht, I.;<br />

Larsson, L. G.; Landegren, U. Methods 2008, 45, 227–232.<br />

(8) Gustafsdottir, S. M.; Schlingemann, J.; Rada-Iglesias, A.; Schallmeiner, E.;<br />

Kamali-Moghaddam, M.; Wadelius, C.; Landegren, U. Proc. Natl. Acad. Sci.<br />

U.S.A. 2007, 104, 3067–3072.<br />

(9) Nordengrahn, A.; Gustafsdottir, S. M.; Ebert, K.; Reid, S. M.; King, D. P.;<br />

Ferris, N. P.; Brocchi, E.; Grazioli, S.; Landegren, U.; Merza, M. Vet.<br />

Microbiol. 2008, 127, 227–236.<br />

(10) Gustafsdottir, S. M.; Wennström, S.; Fredriksson, S.; Schallmeiner, E.;<br />

Hamilton, A. D.; Sebti, S. M.; Landegren, U. Clin. Chem. 2008, 54, 1218–<br />

1225.<br />

Table 1. Oligonucleotide Sequences Used in the<br />

Experimental Model of Proximity Hybridization a<br />

name sequence b<br />

C20 5′-TAATACTTGCTGAGGAATAA-3′ (similar to that<br />

used in standard PLA)<br />

C19 5′-TAATACTTGCTGAGGAATA-3′<br />

C18 5′-TAATACTTGCTGAGGAAT-3′<br />

C17 5′-TAATACTTGCTGAGGAA-3′<br />

C16 5′-TAATACTTGCTGAGGA-3′<br />

C15 5′-TAATACTTGCTGAGG-3′<br />

FreeA 5′-/IABkFQ/GCA AGT ATT ATT TTT TTT TTT TTT<br />

TTT TTT TTT TTT TTT TTC TCT TTT TC-3′<br />

FreeB 5′-/Phos/TCT TCT CTC TCT CTC TTT TTT TTT TTT<br />

TTT TTT TTT TTT TTT ATT CCT CA/6-FAM/-3′<br />

linker 5′-GAG AGA GAG AGA AGA GAA AAA GAG AAA<br />

AAA-3′<br />

Loop 5′-/IAbFQ/GCA AGT ATT ATT TTT TTT TTT TTT<br />

TTT TTT TTT TTT TTT TTC TCT TTT TC TCT TCT<br />

CTC TCT CTC TTT TTT TTT TTT TTT TTT TTT TTT<br />

TTT ATT CCT CA/6-FAM/-3′<br />

a Bases in bold depict the variable segments of the connectors used<br />

to alter connector affinity (see Table 2 for affinities). b Notes: IABkFQ,<br />

Iowa black FQ; Phos, phosphorylated end; 6-FAM, 5(6)-carboxyfluorescein<br />

(mixed isomer).<br />

improved the sensitivity and substantially widened the dynamic<br />

range for thrombin detection compared to standard PLA. This<br />

novel, asymmetric PLA approach should be generally applicable<br />

to any of the various PLA methods developed to date, using either<br />

aptamer- or antibody-based probes. 1,6-10<br />

EXPERIMENTAL SECTION<br />

Reagents and Materials. All solutions were prepared with<br />

deionized, ultrafiltered water (Fisher Scientific). T4 DNA ligase<br />

was purchased from New England BioLabs. Human thrombin was<br />

obtained from Sigma-Aldrich. All oligonucleotides except the<br />

Taqman probe were obtained from Integrated DNA Technologies<br />

(IDT; Coralville, Iowa), with purity and yield confirmed by mass<br />

spectrometry and HPLC, respectively. Sequences of ssDNA<br />

strands used in the experimental model are given in Table 1.<br />

Sequence Free A was labeled at its 5′-end with Iowa black FQ<br />

quencher (IABkFQ; absorbance maximum at 531 nm). Sequence<br />

FreeB was labeled at its 5′-end with a phosphate group<br />

(Phos) and at its 3′-end with 5(6)-carboxyfluorescein (6-FAM,<br />

mixed isomer; emission maximum at 520 nm). Loop strands<br />

(Table 1) were synthesized by ligation of FreeA and FreeB using<br />

T4 DNA ligase; further details on Loop synthesis are included<br />

in the Supporting Information text and Figure S-2. Sequences<br />

(listed 5′ to 3′) for PLA or asymmetric PLA were as follows.<br />

Thrombin aptamer A, aptA, CAG TCC GTG GTA GGG CAG GTT<br />

GGG GTG ACT TCG TGG AAC TAT CTA GCG GTG TAC<br />

GTG AGT GGG CAT GTA GCA AGA GG; thrombin aptamer<br />

B, aptB, /5Phos/GT CAT CAT TCG AAT CGT ACT GCA ATC<br />

GGG TAT TAG GCT AGT GAC TAC TGG TTG GTG AGG<br />

TTG GGT AGT CAC AAA; symmetric connector, C20,PLA, AAG<br />

AAT GAT GAC CCT CTT GCT AAA A; asymmetric connector,<br />

C18,PLA, TTA TGA TGA CCC TCT TGC TAA AA; asymmetric<br />

connector, C16,PLA, TAG ATG ACC CTC TTG CTA AAA; forward<br />

PCR primer, GTG ACT TCG TGG AAC TAT CTA GCG; reverse<br />

PCR primer, AAT ACC CGA TTG CAG TAC GAT TC; Taqman<br />

probe (Applied Biosystems, ABI), 4-chlorofluorescein-TGT<br />

ACG TGA GTG GGC ATG TAG CAA GAG G-carboxytetramethylrhodamine.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6977


Fluorescence Measurements for Model Titrations. Fluorescence<br />

measurements were conducted using a Nanodrop 3300<br />

fluorospectrometer (Thermo Scientific). The intensity of FAM<br />

emission was measured at 520 nm with LED excitation at 485<br />

nm. The buffer used for titrations was 50 mM Tris-HCl (pH 7.5),<br />

100 mM NaCl, 1 mM MgCl2. Titrations sets (14-16 points) were<br />

carried out in duplicate, and each fluorescence measurement was<br />

repeated six times. Fixed amounts of Loop or FreeA and FreeB<br />

strands were incubated with variable amounts of Cb overnight at<br />

room temperature prior to fluorescence measurements.<br />

Proximity Ligation Assays. Temperature control was achieved<br />

using a Mastercycler-EP gradient thermal cycler (Eppendorf).<br />

Thrombin and oligonucleotide solutions were made as described. 1<br />

An amount of 4 µL of a reaction mixture containing human<br />

thrombin, aptA, and aptB was incubated at 37 °C for 15 min<br />

before adding 1 µL of connector (C16,PLA, C18,PLA, orC20,PLA).<br />

An additional 30 min of incubation at 22 °C was performed to<br />

stabilize complexes. These 5 µL samples were then brought<br />

to a total volume of 55 µL for the ligation and amplification<br />

mixture, which contained 50 mM KCl, 10 mM Tris-HCl, pH<br />

8.3, 1.9 mM MgCl2, 0.4 Weiss unit T4 DNA ligase, 73 µM ATP,<br />

0.18 mM dNTPs, 0.45 µM forward and reverse PCR primers,<br />

45 nM Taqman probe for the 5′ nuclease assay, and 1.5 units<br />

of AmpliTaq Gold polymerase (ABI). The reactions were<br />

transferred to a real-time PCR instrument (CFX96, Bio-Rad)<br />

for temperature cycling: 5 min at 22 °C for ligation, 10 min at<br />

95 °C, and then cycling for 15 s at 95 °C and60sat60°C,<br />

repeated 45 times. Signal-to-background (S/BG) values were<br />

calculated from the relative number of ligation products<br />

detected in a sample with thrombin to that in a sample without<br />

thrombin. The relative quantity of the ligated products was<br />

calculated by using the comparative threshold cycle (Ct) method. 11 For measurements of PLA background, 15 pM aptA<br />

and 20 pM aptB were used, with different connector concentrations<br />

as follows: C16,PLA ) 178 nM; C18,PLA ) 112 nM; C20,PLA )<br />

45 nM. For asymmetric PLA, 900 pM aptA and 1200 pM aptB<br />

were used with 480 nM C16,PLA; and for symmetric PLA, 15 pM<br />

aptA and 20 pM aptB were used with 400 nM C16,PLA (same<br />

conditions as previously reported 1 ).<br />

Connector Kd Calculations. Base-pairing dissociation constants<br />

for the invariable segment (Kd,A) and for each variable<br />

segment (Kd,B) of the connectors were calculated using the wellestablished,<br />

thermodynamic nearest-neighbor model. 12 Dissociation<br />

constants were derived from thermodynamic parameters<br />

calculated via the BioPHP melting temperature calculator<br />

(http://insilico.ehu.es/tm.php), a free online service that uses<br />

SantaLucia’s method. 12 Source code is freely downloadable at<br />

http://www.biophp.org/. Kd,B values were also experimentally<br />

confirmed; in Table 2, the column labeled “Kd,B” shows the<br />

BG calculated values, and the column labeled “Kd,eff” shows the<br />

experimentally confirmed values. These values are in good<br />

agreement, as further shown in the Supporting Information<br />

(Figures S-3 and S-4).<br />

(11) Livak, K. J. User Bulletin No. 2: ABI PRISM 7700 Sequence Detection System;<br />

PE Applied Biosystems: Foster City, CA, 1997; pp 11-15.<br />

(12) SantaLucia, J. Proc. Natl. Acad. Sci. U.S.A. 1998, 95, 1460–1465.<br />

6978 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Table 2. Analysis of Connector Sequences Used in the<br />

Experimental Model of Proximity Hybridization a<br />

connector,<br />

C b<br />

length<br />

(bases)<br />

Kd,B BG S (nM) Kd,eff (nM) Kd,eff (nM) S/BG [Cn] opt<br />

(nM)<br />

C20 20 16 18.6 ± 2.7 3.25 ± 0.48 5.72 45<br />

C19 19 135 149 ± 11 5.54 ± 0.93 26.9 80<br />

C 18 18 539 446 ± 59 9.52 ± 1.33 46.8 110<br />

C 17 17 3370 2550 ± 220 7.12 ± 1.67 358 170<br />

C16 16 28 600 7100 ± 740 8.47 ± 0.38 838 480<br />

C 15 15 297 000 17.5 ± 6.3 20 000<br />

a<br />

Notes: Kd,B, calculated dissociation constant for variable segment<br />

of connector in 100 mM salt and 1 mM Mg2+ BG S ; Kd,eff and Kd,eff,<br />

measured<br />

effective dissociation constants of background and signal, respectively;<br />

[Cn]opt, optimal connector concentration for PLA; unvaried Kd,A ) 14.8<br />

nM.<br />

RESULTS AND DISCUSSION<br />

Signal and Background in Proximity Ligation. A schematic<br />

of the original, aptamer-based PLA 1,6 is shown in Figure<br />

1A, in the context of protein detection with aptamer probes.<br />

The quantitative capability of PLA is most heavily dependent<br />

upon two steps: the cooperative hybridization of the four-part<br />

complex, aptamer A-target-aptamer B-connector (aptA-TaptB-C20,PLA),<br />

followed by the proximity-dependent ligation of<br />

the two aptamers using a DNA ligase. This process results in<br />

an amount of ligated product that is proportional to the original<br />

amount of protein analyte (“signal” ligations), albeit with some<br />

analyte-independent ligation products present (“background”<br />

ligations). Although qPCR does not differentiate signal from<br />

background ligations, under optimized conditions 1,6 the signal<br />

will be dominant over background to allow highly sensitive,<br />

indirect detection of the protein analyte. Signal enhancement over<br />

background in PLA is based on the proximity effect. The effective<br />

concentrations of two aptamers are significantly increased due to<br />

their proximity, afforded by simultaneous binding to the same<br />

protein. 3-5<br />

The chief interference in PLA is target-independent, background<br />

ligation (Figure 1A, bottom). This background ligation<br />

has been reported to result from two sources. 1,6,13 First, based<br />

simply on binding equilibria, a fraction of aptamers will always<br />

hybridize with the 20-base connector sequences (C20,PLA), even<br />

in the absence of protein analyte, resulting in target-independent<br />

ligations that increase the background levels in the<br />

assay. 1,6 Second, the T4 DNA ligase is capable of ligating a small<br />

fraction of ssDNA sequences, even in the absence of a connector<br />

sequence. 13,14 This will also increase assay background (not<br />

shown in the figure). Importantly, the presence of this background<br />

limits the amount of usable aptamer (or antibody) probes, which<br />

affects the sensitivity and partially defines the upper limit of the<br />

dynamic range. 6 Methods to widen dynamic range would address<br />

a major weakness of PLA and significantly improve the applicability<br />

of the assay.<br />

Design of the Experimental Model of Proximity Hybridization.<br />

We hypothesized that by decreasing connector binding<br />

affinities for aptamers using an asymmetric connector (Figure 1B),<br />

the amount of background ligation products would be greatly<br />

reduced compared to signal ligation products. In turn, this should<br />

(13) Leslie, D. C.; Sohrabi, A.; Ikonomi, P.; McKee, M. L.; Landers, J. P.<br />

Electrophoresis 2010, 31, 1–8.<br />

(14) Kuhn, H.; Frank-Kamenetskii, M. D. FEBS J. 2005, 272, 5991–6000.


allow the use of higher aptamer concentrations, which can<br />

increase the upper limit of dynamic range and improve sensitivity.<br />

This hypothesis was based on the inherent signal enhancement<br />

given by cooperative binding in the proximity effect, 3-5 which does<br />

not apply to background. To understand the impact of connector<br />

binding affinity on S/BG ratio in PLA, we developed a simple<br />

experimental model using DNA hybridization and fluorescence<br />

quenching (Figure 1C). First, we define an “asymmetric connector”<br />

in the model system as a connector with one side shortened,<br />

compared to the original PLA connector, C20,PLA. Herein, model<br />

connector sequences are represented by Cb, where b is equal<br />

to the total number of bases in the connector that are<br />

complementary to the Loop sequence. For example, the<br />

symmetric connector used in the model is represented by C20,<br />

similar to the 20-base connector used in the original PLA<br />

work, 1,6 whereas an asymmetric connector with only six bases<br />

in the variable region is represented by C16. Sequences are noted<br />

in Table 1, with variable regions in bold. (Note that connectors<br />

used later for PLA or asymmetric PLA are differentiated from<br />

model connectors using the subscript “PLA,” such as C20,PLA or<br />

C16,PLA.) In our model system, signal and background were<br />

treated separately. The equilibria shown in Figure 1C represent<br />

signal complex formation with a sequential binding mode and<br />

background complex formation with an independent binding<br />

mode. These are not blind assumptions but are based upon Hill<br />

coefficients which were determined by nonlinear least-squares<br />

fitting of our experimental data (discussed in detail below).<br />

To model target-dependent signal in PLA, a single DNA loop<br />

was hybridized with connector sequences (Figure 1C, top). The<br />

aptA-T-aptB complex in PLA was modeled using a 100nucleotide<br />

DNA loop (Loop) that is capable of hybridizing with<br />

a connector sequence at both ends, representing the ideal case<br />

in which aptamers possess infinite affinities for their protein<br />

analyte (Kd’s ≈ 0). To measure the extent of hybridization, the<br />

Loop was covalently labeled with IABkFQ quencher at the 5′end<br />

(black circle) and with FAM at the 3′-end (gray circle).<br />

Quenching of FAM fluorescence ensued when both the 5′- and<br />

3′-ends of the Loop were hybridized to a connector sequence,<br />

mimicking the proximity hybridization complex in PLA (i.e.,<br />

Loop-Cb represents aptA-T-aptB-C20,PLA). Thus, a decrease<br />

in Loop fluorescence was referred to as “signal” in this case.<br />

The model of background ligations in PLA (Figure 1C, bottom)<br />

follows from the model of signal. The FAM- and IABkFQ-labeled<br />

Loop sequence was essentially split into two free strands (FreeA<br />

and FreeB), representing aptamers unbound by protein analytes.<br />

Upon combination of both free strands and a connector<br />

sequence, the extent of fluorescence quenching represented<br />

target-independent hybridization in PLA and was proportional<br />

to the amount of background ligation products that would be<br />

expected (i.e., FreeA-Cb-FreeB represents aptA-C20,PLA-aptB).<br />

Thus, a decrease in fluorescence was referred to as “background”<br />

in this case.<br />

Complex Formation with Asymmetric Connectors. Moving<br />

toward our goal of reduced background in PLA, the experimental<br />

model (Figure 1C) was used to determine effective dissociation<br />

constants (Kd,eff) of signal and background complexes at<br />

different connector lengths (Cb, where 15 e b e 20). The<br />

assumption was that the Kd,eff values could represent the relative<br />

amounts of signal and background. Separate titrations of the<br />

Loop and Free strands with the Cb strand gave the corresponding<br />

Kd,eff values of the complexes. The S/BG was then<br />

calculated as<br />

S/BG ) K S<br />

a,eff<br />

BG<br />

Ka,eff ) K BG<br />

d,eff<br />

S<br />

Kd,eff This S/BG represents the ideal case in which aptamers have<br />

infinite affinities for the target (Kd’s ≈ 0).<br />

Connectors’ affinities for the FreeB strand (or for one arm of<br />

the Loop) were reduced by decreasing the number of complementary<br />

bases to FreeB or the Loop arm (gray segments in<br />

Figure 1B; bolded text in Table 1). Base-pairing dissociation<br />

constants for the invariable segment (Kd,A) and for each variable<br />

segment (Kd,B) of the connectors were calculated (Table 2)<br />

using the well-established, thermodynamic nearest-neighbor<br />

model. 12 Starting with the 20-base connector (C20) similar to that<br />

used in standard PLA, the variable segment was decreased in<br />

length, resulting in five different asymmetric connectors (C15<br />

through C19). Titrations with these asymmetric connectors and<br />

the symmetric control connector (C20) were carried out on both<br />

the signal and background models (Figure 1C), and the<br />

decrease in fluorescence was used to assay stability.<br />

Background Reduction in the Experimental Model. Modeling<br />

of cooperative complex formation has proven difficult,<br />

particularly in the realm of predicting the affinity of bivalent<br />

ligands based on known affinities of each individual ligand. 3-5<br />

Rather than attempting to develop an extensive model of the<br />

system shown in Figure 1C, titration data was analyzed using the<br />

Hill equation, 15 a simplified model that has found wide use in<br />

biochemistry, physiology, and pharmacology to analyze binding<br />

equilibria and ligand-receptor interactions. 16 According to<br />

Weiss, 16 the Hill coefficient is best described as an “interaction<br />

coefficient” that reflects the extent of cooperativity among multiple<br />

ligand binding sites. Since it is expected that cooperativity is<br />

inherent to the formation of the aptA-T-aptB-C20 complex in<br />

PLA, or the Loop-Cb complexes in our model, we employed<br />

nonlinear least-squares (NLLS) fitting of our data to the Hill<br />

equation (eq 2) to assay this cooperativity and to determine Kd,eff<br />

values.<br />

S ) S i + (S f - S i )<br />

[C b ] n<br />

(K d,eff ) n + [C b ] n<br />

where S is the fluorescence signal measured as a function of [Cb],<br />

the connector concentration, Si and Sf are the initial and final<br />

signals, respectively, and Kd,eff represents the effective dissocia-<br />

S tion constant of the Loop-Cb complex (Kd,eff) or the FreeA-<br />

BG Cb-FreeB complex (Kd,eff). In this way, fluorescence measurements<br />

during titration with increasingly higher [Cb] could be<br />

fit to the Hill equation (eq 2) and the Kd,eff values could be<br />

extracted. Furthermore, each extracted Hill coefficient, n,<br />

represented the extent of cooperativity in that particular system.<br />

Our expectation was that the Loop-Cb complexes (signal)<br />

(15) Hill, A. V. J. Physiol. 1910, 40, iv–vii.<br />

(16) Weiss, J. N. FASEB J. 1997, 11, 835–841.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(1)<br />

(2)<br />

6979


Figure 2. Results of the experimental model. Fluorescence response curves for the signal (Loop, black circles) and background (Free, gray<br />

open squares) complexes when titrated with (A) C20 or (B) C16. NLLS fits to the Hill equation (eq 2) are shown as solid curves. Other curves<br />

(titration with C15,C17,C18, and C19) are given in Supporting Information Figure S-1. Also shown are (C) relative amounts of signal and background<br />

using each connector and (D) signal-to-background ratios as functions of the dissociation constants (Kd,B) of the variable segments of each<br />

connector, Cb.<br />

would have significantly increased cooperativity (higher n)<br />

compared to the FreeA-Cb-FreeB complexes (background).<br />

Example data sets from these titrations are shown in Figure<br />

2, with error bars representing standard deviations of the triplicate<br />

measurements. Figure 2A shows that the Loop-C20 complex<br />

(filled black circles) is more thermodynamically favored<br />

compared to the FreeA-C20-FreeB complex (open gray squares).<br />

This result was expected, owing to the success of various<br />

PLAs 1,6-10 using 20-base connectors such as C20. This data, and<br />

all titrations to follow, were fit to eq 2 via NLLS fitting to<br />

determine Kd,eff values and S/BG (via eq 1) for each asymmetric<br />

connector (Cb); measured and calculated values are shown in<br />

Table 2. With the use of the symmetric connector, C20, the S/BG<br />

value was determined to be 5.72. Figure 2B shows the<br />

fluorescence measurements from a similar experiment using C16.<br />

This figure clearly shows that the relative stability of the signal<br />

complex (Loop-C16, filled black circles) over the background<br />

complex (FreeA-C16-FreeB, open gray squares) is much<br />

greater than that measured with C20, without a large decrease<br />

in stability of the signal complex. In fact, the S/BG value using<br />

C16 was determined to be 838. Thus, a 146-fold increase in<br />

S/BG was achieved by simply decreasing the connector length<br />

in an asymmetric manner. Titration curves for C15, C17, C18,<br />

and C19 (Supporting Information Figure S-1) follow this trend of<br />

increasing S/BG with decreasing connector length (Figure 2, parts<br />

6980 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

C and D). As shown in Figure 2C, the background levels with<br />

C20 were 382-fold larger than the background with C16, whereas<br />

the signal differed by only 3.16-fold. These large decreases in<br />

background and only modest decreases in signal gave the<br />

expected increase in S/BG with decreasing connector affinity<br />

BG (Figure 2D). The agreement between measured Kd,eff and<br />

calculated Kd,B values was also encouraging (Figures S-2 and<br />

S-3 in the Supporting Information). It must be noted that the<br />

background complex titration experiment for C15 was impractical,<br />

due to the expense involved in working with millimolar<br />

concentrations of DNA. Although, using nearest-neighbor<br />

calculations, 12 it is predicted that the S/BG could be as high<br />

as 1.7 × 104 for C15. These results provided a first step toward<br />

confirmation of our hypothesis.<br />

Also shown in Figure 2D is a curve (dashed line) calculated<br />

from data presented by Tian and Heyduk, 5 who developed a<br />

similar experimental model using bivalent ligands of various<br />

affinities. It is clear that the general trend observed in our data<br />

matches with that seen by Tian and Heyduk with an obvious offset<br />

in magnitude. This offset is likely due to the model differences,<br />

in which Tian and Heyduk used a6nmspacer between their<br />

binding sites, whereas our model has essentially no spacer<br />

between the binding sites. Inclusion of an equivalent spacer in<br />

our system would be unreasonable, since PLA requires subse-


quent ligation of the aptamers after hybridization of the signal<br />

complex. Nonetheless, the similarities were encouraging.<br />

Figure 2D presents the relationship between S/BG and the<br />

affinities of the variable segments of the connectors, which are<br />

represented by the calculated Kd,B values 12 (Table 2). This<br />

analysis is useful, since Kd,B will ultimately become the independent<br />

variable in the asymmetric PLA approach. As shown<br />

in the figure, S/BG increases proportionally with Kd,B. In<br />

theory, one could increase Kd,B even further. In the extreme<br />

example, one could use C11 (only one base pair in the variable<br />

segment). However, the S/BG parameter is not useful if there<br />

is insufficient signal present with respect to the noise of<br />

measurement, and it is likely that the amount of either signal<br />

or background (or both) complexes using C11 would be<br />

negligible. The connector length, b, could be considered an<br />

alternative for the independent variable (shown as upper axis<br />

in Figure 2, parts C and D), but care must be taken, since b<br />

remains dependent upon base composition, making it less precise<br />

than Kd,B.<br />

Hill coefficients (n) were also determined by NLLS fitting of<br />

the data to eq 2. The mean n values for formation of all of the<br />

background (FreeA-Cn-FreeB) and signal (Loop-Cn) complexes<br />

were 1.42 ± 0.19 and 1.76 ± 0.11, respectively. These<br />

values were shown to be statistically different (p < 0.01).<br />

Cooperativity enhancement in the signal complex was expected,<br />

since this effect gives PLA the ability to discriminate signal<br />

from background. Furthermore, as reported by Weiss, 16 these<br />

Hill coefficients suggest that the signal complex forms in a<br />

sequential manner, whereas the background complex forms<br />

in an independent manner, as reflected in Figure 1C.<br />

In summary, the improvements in S/BG in the experimental<br />

model is derived from the more than 2 orders-of-magnitude<br />

BG increase in Kd,eff (reduced affinity) that is achieved using<br />

S asymmetric connectors, whereas Kd,eff is increased by less than<br />

1 order of magnitude (Table 2, Figure 2C); thus, by eq 1, S/BG<br />

is increased significantly (Figure 2D). However, application of<br />

asymmetric connectors to PLA requires experimental confirmation,<br />

as given below, since we began our model with the<br />

assumption of infinite aptamer affinities.<br />

Design of Asymmetric PLA. A schematic of asymmetric PLA<br />

is shown in Figure 1B. Since decreased background was achieved<br />

in the experimental model (Figure 2), the next step was to confirm<br />

that this approach was useful for PLA. The real assay is inherently<br />

more complicated than the experimental model, requiring not only<br />

a four-part complex formation but also ligation of aptA to aptB,<br />

followed by qPCR. 1,6 Furthermore, aptamer affinities are finite<br />

in the real system, compared to the assumption of infinite affinities<br />

in the model. On the basis of these factors, we expected that<br />

the observations from the model would be less pronounced in<br />

the real system. Asymmetric PLA (Figure 1B) is based on the<br />

formation of the complex aptA-T-aptB-Cb,PLA, where b < 10 is<br />

achieved by decreasing connector lengths on the side complementary<br />

to aptB. Shorter connectors, therefore, have higher<br />

Kd,B (Table 2), or lower affinity.<br />

For the standard PLA system (C20,PLA, b ) 20), the optimal<br />

connector concentration was determined to be >40 nM. 1 This<br />

correlates well with the model results shown in Figure 2A. At<br />

C20,PLA ) 40 nM, the stability of the signal complex is near<br />

maximal, whereas the stability of the background complex is<br />

minimized as much as possible. Working on this principle, the<br />

results from the titrations in the experimental model (Figure<br />

2, parts A and B and Supporting Information Figure S-1) were<br />

used to estimate the optimal connector concentrations, [Cn]opt,<br />

for each model connector of different length. These values are<br />

reported in the rightmost column of Table 2 and were used for<br />

all asymmetric PLA experiments to follow. This approach allowed<br />

the interrogation of each asymmetric PLA system at its optimal<br />

values of signal and background.<br />

Background Reduction Using Asymmetric PLA. Starting<br />

with the identical reagents used by Landegren and co-workers 1,6<br />

for sensitive detection of human thrombin, connector lengths were<br />

systematically decreased from the initial C20,PLA sequence. Background<br />

ligation levels were measured by carrying out the assay<br />

without the target protein (human thrombin) present. As shown<br />

in Figure 3A, the background levels in asymmetric PLA were<br />

significantly reduced by increasing Kd,B (or decreasing length,<br />

b), matching well with the experimental model (Figure 2C).<br />

Background levels in asymmetric PLA with C16,PLA were 46.7fold<br />

lower than with standard, symmetric PLA (C20,PLA), thus<br />

confirming our hypothesis. It should, therefore, be possible to<br />

use higher concentrations of aptamer probes with asymmetric<br />

PLA to improve the dynamic range and sensitivity.<br />

Dynamic Range and Sensitivity Enhancement with Asymmetric<br />

PLA. A major limitation of standard PLA is the relatively<br />

narrow dynamic range. This disadvantage is directly related to<br />

the generation of background ligations. The original work in PLA<br />

for human thrombin detection determined the optimal concentrations<br />

of aptamer probes to be 15 and 20 pM for aptA and aptB in<br />

the 5 µL incubation mixture, respectively. 1 At higher probe<br />

concentrations, increased background ligations were shown to<br />

be detrimental to the S/BG of the assay. Since the assay relies<br />

on the simultaneous binding of two aptamers to the same<br />

target, simple probability theory shows that the population of<br />

signal complexes (aptA-T-aptB-Cb,PLA) will first increase then<br />

decrease as the concentration of target is increased. The upper<br />

limit of the dynamic range is thus positively related to both<br />

the concentrations and the Kd’s of the aptamer probes. Gullberg<br />

et al. 6 showed that target at 10 times higher concentration than<br />

probes could not be assayed with probe Kd < 0.4 nM but could<br />

be assayed with a probe Kd ) 2.5 nM, albeit with relatively<br />

low S/BG. This implies that in standard PLA, a lower LOD is<br />

achievable with higher affinity probes at the expense of narrow<br />

dynamic range.<br />

Previous publications using standard, aptamer-based PLA for<br />

thrombin detection did not report the LOD or the dynamic<br />

range. 1,6 However, by analyzing the presented data, the reported<br />

LOD was determined to be 16 amol (significantly different from<br />

background, p < 0.05), with a dynamic range from 16 to 400 amol.<br />

In our laboratory, with the standard PLA system using C20,PLA,<br />

we obtained a detection limit about 3-fold higher than 16 amol,<br />

but we were able to achieve a significantly wider dynamic range<br />

compared to the previous reports. As shown in Figure 3B (open<br />

gray squares), our standard PLA detection limit was 50 amol, and<br />

our dynamic range was from 50 to 3000 amol (7.68-fold wider than<br />

previous reports).<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6981


Figure 3. Asymmetric PLA. (A) Amount of background ligations vs<br />

dissociation constant (Kd,B) of the variable segments of each connector,<br />

Cb,PLA, in the absence of target protein. (B) Titration with human<br />

thrombin. Asymmetric connector C16,PLA (filled black circles) and<br />

symmetric connector C20,PLA (open gray squares) were used for<br />

comparison. Error bars represent the standard error of quadruplicate<br />

qPCR measurements. Horizontal bars define the assay dynamic<br />

ranges. (C) Linear scale comparisons of dynamic range, sensitivity,<br />

and limit of detection (LOD) improvements using asymmetric connector<br />

C16,PLA (white bars) compared to the standard connector C20,PLA<br />

(gray bars). Error bars depict standard deviations of linear sensitivities.<br />

The cross-hatched bar represents the LOD obtained in the original<br />

reports of PLA (refs 1 and 6).<br />

With significant reductions in background using asymmetric<br />

PLA (Figure 3A), we hypothesized that the dynamic range could<br />

be improved using higher aptamer concentrations with connector<br />

C16,PLA. This hypothesis was confirmed by titration of the system<br />

with human thrombin (Figure 3B). Asymmetric (closed black<br />

circles) and standard (open gray squares) PLA were carried out<br />

under the following optimal conditions in the 5 µL incubation<br />

mixture: asymmetric, [C16,PLA] ) 480 nM, [aptA] ) 0.9 nM, and<br />

[aptB] ) 1.2 nM; standard, [C20,PLA] ) 45 nM, [aptA] ) 15 pM,<br />

and [aptB] ) 20 pM. Using asymmetric PLA, the dynamic range<br />

6982 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

was increased by 2 orders of magnitude, and the sensitivity<br />

was improved by a factor 1.57 compared to the standard system<br />

with C20,PLA (Figure 3, parts B and C). In addition, the maximum<br />

achievable S/BG was larger by a factor of 3.15. These improvements<br />

were accomplished with an additional 3-fold improvement<br />

in LOD to 5.0 amol of thrombin. When compared to previously<br />

reported data, 1,6 our dynamic range was improved by an even<br />

larger factor of 323. In Figure 3B, horizontal bars mark the<br />

dynamic ranges, with the upper bounds defined as the highest<br />

concentration with a statistically higher (p < 0.05) value of signalto-background<br />

ratio compared to the adjacent lower concentration.<br />

Error bars represent standard errors of quadruplicate runs of<br />

qPCR.<br />

Since logarithmic scales were used in Figure 3B, linear scale<br />

comparisons of dynamic ranges, sensitivities, and limits of detection<br />

are shown in Figure 3C. Sensitivities were determined as the<br />

slopes of the linear regressions of both data sets in the range<br />

from the LOD to 3000 amol. This figure helps to clarify the<br />

significant advantages provided by asymmetric PLA, in comparison<br />

to standard PLA. Most notably, our approach of using<br />

asymmetric connectors has allowed a significant improvement in<br />

one of the major limitations of the assay, dynamic range. This<br />

improvement should be generally applicable to any of the<br />

previously developed PLA methods. 1,6-10<br />

CONCLUSIONS<br />

This report describes a method of utilizing asymmetric connectors<br />

to significantly improve the dynamic range of PLA, thereby<br />

addressing a major limitation of the assay. Sensitivity, LOD, and<br />

maximum S/BG were also improved. By enhancing its already<br />

successful predecessor, this asymmetric PLA method approaches<br />

an ideal assay system, providing a homogeneous and highly<br />

specific protein assay with the capability to detect very small<br />

quantities of protein in free solution, at high throughput, and with<br />

good dynamic range.<br />

Furthermore, an experimental model of proximity hybridization<br />

was successfully developed and tested. This model system<br />

ultimately improved our understanding of the proximity effect and<br />

its inherent cooperative assembly in PLA, highlighting the<br />

importance of the interplay between connector-to-aptamer affinities<br />

and aptamer-to-target affinities in maximizing S/BG and dynamic<br />

range. Generally speaking, low connector affinities can be paired<br />

with high probe affinities (aptamers or antibodies) to achieve this<br />

goal.<br />

ACKNOWLEDGMENT<br />

Support for this work was provided by the Auburn University<br />

Department of <strong>Chemistry</strong> and Biochemistry. The authors thank<br />

Professor Douglas C. Goodwin for use of his PCR and gel<br />

preparation equipment, as well as Daniel C. Leslie for helpful<br />

discussions about proximity ligation and ligation background.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 14, 2010. Accepted July 8, 2010.<br />

AC101762M


Anal. Chem. 2010, 82, 6983–6990<br />

Improving Pressure Robustness, Reliability, and<br />

Versatility of Solenoid-Pump Flow Systems Using a<br />

Miniature Economic Control Unit Including Two<br />

Simple Pressure Pulse Mathematical Models<br />

Burkhard Horstkotte, † Erich Ledesma, ‡ Carlos M. Duarte, † and Víctor Cerdà* ,‡<br />

Department of Global Change Research, IMEDEA (CSIC-UIB) Institut Mediterráni d’Estudis Avançats, Miquel<br />

Marques 21, 07190 Esporles, Spain, and University of the Balearic Islands, Department of <strong>Chemistry</strong>, Carreterra de<br />

Valldemossa km 7, 5, 07011 Palma de Mallorca, Spain<br />

In this work we have systematically studied the behavior<br />

of solenoid pumps (SMP) as a function of flow rate and<br />

flow resistance. Using a new, economic, and miniature<br />

control unit, we achieved improvements of the systems<br />

versatility, transportability, and pressure robustness. A<br />

further important improvement with respect to pressure<br />

resistance was achieved when a flexible pumping tube was<br />

inserted between the solenoid pump and the flow resistance<br />

acting as a pressure reservoir and pulsation damper.<br />

The experimental data were compared with two pressure<br />

pulse models for SMP, which were developed during this<br />

work and which were well-suited to describe the SMP<br />

operation.<br />

Solenoid micropumps (SMP) have gained considerable importance<br />

as liquid drivers in analytical flow techniques (FT) since<br />

their first application in flow systems reported by Weeks and<br />

Johnson. 1 The initial motivation for the use of SMP was miniaturization<br />

of field applicable FT systems.<br />

SMP present an economic alterative to (multi)syringe and<br />

peristaltic pumps, typically used for flow injection analysis (FIA) 2<br />

and sequential injection analysis (SIA) 3 systems, respectively.<br />

They provide a semicontinuous flow with highly pronounced<br />

pulsation. It has been considered that this pulsation causes<br />

intermediate turbulent conditions in the flow manifold improving<br />

mixing of the sample and reagents in comparison with other<br />

unsegmented FT, where laminar flow conditions are typical.<br />

While in the first works, this feature was considered as a<br />

disadvantage due to the oscillations of the detector signal, Lapa et<br />

al. proved first that mixing of reagent and sample is enhanced and<br />

sensitivity improvements of 50% are feasible. 4 The authors further<br />

established the denomination “multipumping flow systems” (MPFS).<br />

In contrast to syringe pumps, SMP operate continuously, which<br />

bears the potential to shorten the time of analysis or to perform<br />

* To whom correspondence should be addressed. Phone: +34 971 173 261.<br />

Fax: +34 971 173 426. E-mail: Victor.Cerda@uib.es.<br />

† IMEDEA (CSIC-UIB) Institut Mediterráni d’Estudis Avançats.<br />

‡ University of the Balearic Islands.<br />

(1) Weeks, D. A.; Johnson, K. S. Anal. Chem. 1996, 68, 2717–2719.<br />

(2) Ruzicka, J.; Hansen, E. H. Anal. Chim. Acta 1975, 78, 145–157.<br />

(3) Ruzicka, J.; Marshall, G. Anal. Chim. Acta 1990, 237, 329–343.<br />

(4) Lapa, R. A. S.; Lima, J. L. F. C.; Reis, B. F.; Santos, J. L. M.; Zagatto, E. A. G.<br />

Anal. Chim. Acta 2002, 466, 125–132.<br />

analyte from large sample volumes. 5,6 In contrast to and in<br />

advantage over peristaltic pumps, in MPFS, each flow channel<br />

can be controlled individually. Both FIA and SIA manifold<br />

configurations have been successfully applied so far using SMP,<br />

i.e., confluent, codirectional flows, and contra-directional (aspiration<br />

and dispense), respectively. 4,7 <strong>Analytical</strong> applications of MPFS<br />

have been reviewed in detail elsewhere. 8-10<br />

However, SMP show two considerable shortcomings: they do<br />

not work reliable in the presence of gas bubbles and particulate<br />

matter affecting both the actuation of the build-in check valves<br />

and they are notably liable to flow backpressure leading to a<br />

decrease of the effective flow with increasing flow resistance.<br />

Although there is an obvious interest in the improvement of the<br />

pressure robustness and flow rate reliability, up to date, there have<br />

been hardly any efforts to complete this objective. In this work,<br />

we studied the pressure robustness of the SMP as a function of<br />

backpressure including two strategies to improve of the pressure<br />

robustness. These have been the adaptation of the activation time,<br />

made possible by the use a versatile and highly economic relay<br />

card and software control, and the use of inflatable pumping tubes<br />

to decrease the peak backpressure. We further present for the<br />

first time simple models of the pressure and flow pulse behavior<br />

as a function of pump dimensions and manifold characteristics,<br />

which were capable of explaining the observations made from the<br />

experiments undertake .<br />

MATERIALS AND METHODS<br />

Distilled water was used throughout. For the evaluation of flow<br />

rates, pumped volumes where quantified by weighing on an<br />

analytical balance, including compensation of the water density<br />

at ambient temperature.<br />

The self-priming SMP from Bio-Chem Fluidics (Boston, NJ)<br />

of nominal 8 µL, 40 µL, and two of 25 µL volumes were used (types<br />

(5) Pons, C.; Santos, J. L. M.; Lima, J. L. F. C.; Forteza, R.; Cerdà, V. Microchim.<br />

Acta 2008, 161, 73–79.<br />

(6) Pons, C.; Forteza, R.; Cerdà, V. Anal. Chim. Acta 2005, 550, 33–39.<br />

(7) Pinto, P. C. A. G.; Saraiva, M.L.M.F.S.; Santos, J. L. M.; Lima, J. L. F. C.<br />

Anal. Chim. Acta 2005, 539, 173–179.<br />

(8) Lima, J. L. F. C.; Santos, J. L. M.; Dias, A. C. B.; Ribeiro, M. F. T.; Zagatto,<br />

E. A. G. Talanta 2004, 64, 1091–1098.<br />

(9) Rocha, F. R. P.; Reis, B. F.; Zagatto, E. A. G.; Lima, J. L. F. C.; Lapa, R. A. S.;<br />

Santos, J. L. M. Anal. Chim. Acta 2002, 468, 119–131.<br />

(10) Santos, J. L. M.; Ribeiro, M. F. T.; Dias, A. C. B.; Lima, J. L. F. C.; Zagatto,<br />

E. E. A. Anal. Chim. Acta 2007, 600, 21–28.<br />

10.1021/ac101250h © 2010 American <strong>Chemical</strong> Society 6983<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/20/2010


Figure 1. Scheme of the operation of the used solenoid micropumps in (A) activation (aspiration) and (B) deactivation (expulsion). Elements:<br />

1, pretension spring; 2, solenoid; 3, spring; 4, inlet check valve; 5, outlet check valve; 6, metal core of solenoid; 7, membrane; 8, inner volume<br />

of pump; 9, inlet; 10, outlet.<br />

P/N090SP-12-8, P/N110TP-12-40, and P/N120SP-12-25) and are<br />

denoted as pumps 1, 4, 2, and 3, respectively. In the activation<br />

state, the SMP operate in suction whereas at deactivation the pulse<br />

volume is expelled in the forward direction by the pressure<br />

exposed by a metal spring as shown schematically in Figure 1.<br />

The product specifications indicate a pressure height of 1.3 m for<br />

aspiration and 3.5 m for dispensing. Detailed specifications cited<br />

in this article can be downloaded at the producer’s Web site. 11<br />

For control and powering, an eight channel relay card<br />

(SERDIO8R) from EasyDAC (Glasgow, United Kingdom) was<br />

used. It allowed remote and independent control of up to eight<br />

solenoids (SMP or valves) via an RS232 serial interface. Connection<br />

of the solenoid and powering is accomplished via screw<br />

terminals, and no further equipment was required for operation.<br />

For reliable operation, parallel connection of a diode to the<br />

connected solenoid devices in reverse direction is required to<br />

deduct the induced counter-voltage at the solenoid release. The<br />

relay card is highly appropriate for miniaturization of MPFS as<br />

the dimensions are about 11 cm × 10 cm × 3 cm. A thorough<br />

description can be downloaded from the manufacture’s Web site.<br />

Three PTFE tubes of different lengths (L) and inner diameters<br />

(i.d.) were used as model flow resistances (tube 1, 49 cm, 0.2<br />

mm i.d.; tube 2, 30 cm, 0.5 mm i.d.; and tube 3, 175 cm, 0.8 mm<br />

i.d.). For calibration of the effective pressure at a given flow rate,<br />

a Bu4S syringe pump module 12,13 (16.000 steps, 24-1024 s for<br />

total dispense) from Crison Instruments S.A. (Alella, Barcelona,<br />

Spain) was used, equipped with a glass syringe of 5 mL from<br />

Hamilton Bonaduz AG (Bonaduz, Switzerland). Pressure measurements<br />

were performed using a Bourdon pressure gauge of 6 bar<br />

working range connected by a peek tube (20 cm, 0.8 mm i.d.)<br />

and a three-way connector between the solenoid pump and the<br />

pressure resistance. For testing the potential of pressure pulse<br />

dampers, two pieces of purple/black marked peristaltic pumping<br />

tube (2.05 mm i.d.) of 30 and 6.5 cm effective length were used.<br />

(11) http://www.biochemfluidics.com/pdf/micro-pumps_brochure_dn_ibpmp-<br />

01_r1.pdf, accessed February 20, 2010.<br />

(12) Cerda, V.; Estela, J. M.; Forteza, R.; Cladera, A.; Becerra, E.; Altimira, P.;<br />

Sitjar, P. Talanta 1999, 50, 695–705.<br />

(13) Horstkotte, B.; Elsholz, O.; Cerdá, V. J. Flow Injection Anal. 2005, 22,<br />

99–109.<br />

6984 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Remote software control of the relay card as well as of the<br />

syringe pump was done via an RS232C serial interface using the<br />

AutoAnalysis 5.0 platform 14 from Sciware (Palma de Mallorca,<br />

Spain) including specific dynamic link libraries (DLL) for each<br />

communication port and connected instrument.<br />

The DLL made for the used relay card was adapted for both<br />

SMP and three-way solenoid valves. It included a calibration mode<br />

for connected SMP, definition of the mean pulse volume and<br />

minimal activation time for each relay, and the selection and<br />

definition of two of the three operation parameters (flow rate,<br />

dispense volume, and operation time) for each step and for each<br />

individual SMP. The control panels corresponding to configuration<br />

and operation instructions are shown in Figure 2, respectively.<br />

The DLL is commercially available from Sciware SL.<br />

RESULTS AND DISCUSSION<br />

Calibration of the Flow Resistances. For calibration of flow<br />

resistances, a smooth, nonpulsed flow was required. Therefore,<br />

a syringe pump was used. The obtained data are given in Table<br />

1. The relation between flow rate and pressure was linear, and<br />

effective flow resistances of 0.66, 0.062, and 0.017 bar min mL -1<br />

were found corresponding to hydrodynamic diameters of tubes<br />

of 0.27, 0.42, and 0.91 mm i.d.<br />

3.2. Calibration of the Solenoid Micropumps. For calibration,<br />

the volume of water resulting from 240 pump pulses was<br />

weighted in triplicate using an activation frequency of 2 Hz and,<br />

for both the in- and outflow of the SMP, PTFE tubes of 20 cm<br />

length and 1.5 mm i.d., respectively, which corresponded to<br />

neglectible flow resistances.<br />

The activation time was varied between 50 and 400 ms. Longer<br />

activation times were not feasible due to low operation reliability<br />

and excessive heating of the solenoid. The results and mean pulse<br />

volumes are represented in Figure 3.<br />

It was found that the mean pulse volume was approximately<br />

stable for activation times between 150 and 300 ms and exceeded<br />

in every case the nominal volume given by the producer: by 12%<br />

(pump 1), 21% (pump 2), 40% (pump 2), and 120% (pump 4) (P/<br />

(14) Becerra, E.; Cladera, A.; Cerda, V. Lab. Robotics Autom. 1999, 58, 131–<br />

140.


Figure 2. Configuration editor window (above) and instruction command window (below) for the software control of the used relay card using<br />

the program AutoAnalysis.<br />

Table 1. Data from Calibration of Model Flow<br />

Resistances<br />

flow rate<br />

[mL/min]<br />

tube 1<br />

(49 cm,<br />

0.2 mm i.d.)<br />

pressure [bar]<br />

tube 2<br />

(30 cm,<br />

0.5 mm i.d.)<br />

tube 3<br />

(175 cm,<br />

0.8 mm i.d.)<br />

0.50 0.30<br />

0.60 0.35<br />

0.75 0.47<br />

1.00 0.70<br />

1.50 1.00<br />

2.50 0.15<br />

3.75 0.22<br />

5.00 0.32 0.08<br />

6.00 0.40 0.10<br />

7.50 0.50 0.14<br />

pressure/flow rate 0.659 0.0651 0.0175<br />

[bar min mL -1 ]<br />

hydrodynamic diameter<br />

[mm]<br />

0.27 0.42 0.91<br />

N110TP-12-40). For pumps 2 and 3 (type P/N120SP-12-25), the<br />

same behavior of decreasing pulse volume with increasing<br />

activation time was found, while for pumps 1 and 4, the expulsion<br />

volume increased with the activation time. This was reduced to<br />

different responses of the rubber two check valves (see Figure<br />

1). Because of slight differences in fabrication and abrasion, the<br />

characteristics of the pump are assumed to be determined by the<br />

weaker check valve.<br />

For pump 4, activation times below 50 ms or above 400 ms<br />

disabled pumping operation. Activation times above 250 ms lead<br />

to notable pump heating, leading to higher pulse volumes.<br />

3.3. Pressure Pulse Model 1. To simulate the pressure<br />

pulse, a numerical flow model was created. The inner volume of<br />

the SMP (see Figure 1) can be described best by a cone, where<br />

the base is the cross section area of the inner cavity of the pump<br />

and the surface shell is the membrane elevated in its central point<br />

by attraction of the solenoid. The inner volume is therefore<br />

calculated by eq 1, where h i is the lift of the membrane at time<br />

i.<br />

Vin,i ) 1<br />

3 r 2<br />

pump πhi (1)<br />

At the release of the solenoid, the pressure is equal to the spring<br />

force F applied on the inner cross section area A of the pump.<br />

The spring force is a product of the spring constant D and the<br />

sum of the initial lift h0 and the pretension h′ according eq 2.<br />

The initial lift was calculated from the calibrated pulse volume of<br />

the pump by transformation of eq 1.<br />

p 0 ) F spring<br />

A pump<br />

) D(h0 + h′)<br />

)<br />

2<br />

π<br />

r pump<br />

D( 3Vpump + h′)<br />

2<br />

rpump π<br />

2<br />

rpump π<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(2)<br />

6985


Figure 3. Dependency of mean pulse volume on the activation time (240 pulses at 2 Hz, 3-fold repetition).<br />

The initial pressure leads to an outflow Q through the connected<br />

flow resistance characterized by tube length L and inner radius r<br />

and liquid viscosity η at a given temperature given by eq<br />

3(Hagen-Poiseuille equation). 15<br />

Q i ) p i r4 π<br />

8ηL<br />

The outflow leads to a spring and pressure release. The full release<br />

is reached when hi equals 0. The membrane position and<br />

pressure after one time interval ∆t are described by eqs 4 and<br />

2, respectively. Here, we applied a time interval ∆t of 10 ms for<br />

modeling.<br />

hi )hi-1 - 1 Qi∆t 3 2<br />

rpump π<br />

To describe the flow conditions, the Reynold number was<br />

calculated further being the product of flow rate, radius of the<br />

flow resistance tube, liquid density, and dynamic viscosity. 15<br />

Re i ) Q i r2F<br />

η<br />

3.4. Application of Solenoid Valves to Model Flow Resistances.<br />

The model flow resistances (tubes 1-3) were tested<br />

on the SMP for different activation times. The found mean pulse<br />

volume after expulsion of a theoretical volume of 2 mL was set in<br />

relation to the maximal found pulse volume during the former<br />

calibration, denoted further as operation efficiency (OE). The<br />

results are shown in Figure 4A-D.<br />

(15) Glück, B. Hydrodynamische und gasdynamische Rohrströmung; Druckverluste;<br />

Bausteine der Heizungstechnik; VEB Verlag für Bauwesen: Berlin, Germany,<br />

1998 (ISBN: 3-345-00222-1).<br />

6986 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(3)<br />

(4)<br />

(5)<br />

With the use of an activation time of 200 ms, it was found that<br />

even for the lowest flow resistance (tube 3) the OE were<br />

significantly reduced and that efficiency loss increased with the<br />

pump size or pulse volume, respectively. In fact, the found pulse<br />

volumes fitted approximately with the nominal pulse volume of<br />

the SMP given by the producer.<br />

For the lowest resistance, OE was about 90% (same range as<br />

the former reported 1 ) for pumps 1, 2, and 3 and about 40% for<br />

pump 4. For tube 3, OE values further showed to be independent<br />

from the flow rate, while for higher flow resistances, more<br />

pronounced for tube 1 than for tube 2, the OE values decreased<br />

considerably with higher flow rate. The behavior of pumps 2 and<br />

3 were found to be similar when the pulsation frequency was used<br />

as an independent variable of data representation.<br />

Decreasing the activation time from 200 to 50 ms but maintaining<br />

the same operation frequency improved the OE for pumps 2<br />

and 3 for flow resistance tube 1 up to 15%. A similar behavior was<br />

found for pump 1 at operation frequencies >1.5 Hz. This improvement<br />

was reduced to the prolongation of the deactivation time<br />

and in consequence of outflow progress.<br />

In contrast, the OE of pump 4 decreased mainly with the flow<br />

resistance with minimal effect on the flow rate. This was explained<br />

by the high volume of the SMP requiring deactivation times of<br />

several seconds for full expulsion. Leakage of the inlet SMP check<br />

valve has to be considered further. Such leakage would lead to a<br />

continuous loss over the entire pressure exhibition time.<br />

The observations indicate that the mean flow rate is not an<br />

appropriate parameter to describe or calculate the backpressure<br />

exposed by the flow resistances but the pulse volume and<br />

pulsation frequency are the main parameters affecting the flow<br />

rate. Pons et al. 6 recommended “periodic recalibration of the<br />

micropumps”. From the results presented here, it becomes<br />

clear that reliable off-system calibration of the SMP is not<br />

possible but has to be done within the flow system and at the


Figure 4. Dependencies between volumetric efficiency and mean flow<br />

rate for pump 1 (A), 2 (B), 3 (C), and 4 (D) for different activation times<br />

and flow resistances. In part D, the effect of peristaltic pumping tubes<br />

used as a pulsation damper on pressure robustness is shown further.<br />

aimed operation frequency leading otherwise to considerable<br />

volumetric errors. Following this rule, we did not observe any<br />

significant alterations of the pulse volume during this work.<br />

The pressure pulse of pump 3 was simulated. The constant of<br />

the integrated spring was determined by the force exerted on the<br />

plate of an analytical balance on compression. The spring tension<br />

length was measured simultaneously with the depth probe of a<br />

vernier caliber. The spring constant was determined to by about<br />

2500 N/m. Further geometric dimensions measured by the<br />

vernier caliber were a prestressing of 4 mm and an inner diameter<br />

of 14 mm. The initial lift h0 was then calculated from the pulse<br />

volume of 40 µL using eq 1.<br />

Simulations of the pressure pulse release made with model 1<br />

are shown in Figure 5 for pump 3 neglecting any elasticity of the<br />

flow resistance tubes. As expected, the pressure release follows<br />

a first-order lag (PT1) behavior. Since the lion share of the spring<br />

force is the result of the pretension, the pressure decrease over<br />

the pulse duration is low from 0.8 to 0.67 bar. This fact is<br />

fundamental for reliable SMP operation, and a simple way to<br />

improve the SMP pressure stability is to enhance the pretension<br />

and, if required, the operation voltage. However, break-through<br />

of the check valves and excessive solenoid heating probably limits<br />

this modification.<br />

From Figure 5C it becomes clear why the OE of pump 3<br />

decreases at frequencies above 0.5 Hz for tube 1. As a matter of<br />

fact, the deactivation time is not sufficiently long for the flow out<br />

of the entire pulse volume from the SMP. With the use of shorter<br />

activation times, i.e., longer deactivation times while keeping the<br />

operation frequency constant led therefore to improved operation<br />

efficiencies.<br />

As stated, the initial pressure of pump 3 was about 0.8 bar<br />

while the producer guarantees only about 0.33 bar. To evaluate<br />

the maximal pressure applicable by the SMP, they were directly<br />

connected to the aneroid barometer, obtaining stable levels of 1.35<br />

bar for pump 1, 1.55 bar for pumps 2 and 3, and 1.15 bar for pump<br />

4, overcoming the stated pressure stability up to 4-fold but at the<br />

expense of an OE of zero (dispense volume ) 0). The theoretical<br />

pressures at the tested flow rates considering pulse-less flow are<br />

given in Table 2. The OE started to decrease for theoretical<br />

pressures higher than 0.5 bar. It should be pointed out that with<br />

model 1 with pump 3, Reynold numbers higher than 2300 were<br />

not obtained, indicating that the pulsated flow is still laminar,<br />

which contradicts the generally stated turbulent flow conditions<br />

for MPFS.<br />

3.5. Pressure Pulse Model 2. Elasticity of the connected<br />

tubing was neglected in pressure pulse model 1. In reality, tubing<br />

elasticity leads to a faster pulse volume expulsion, faster pressure<br />

decrease, and further delay behavior of both pressure and flow<br />

rate in the downstream tube.<br />

In section 3.6, experiments made with an elastic pumping tube<br />

inserted between the SMP and the flow resistance will be<br />

discussed. To describe the fluid mechanics, we included the<br />

characteristics of an elastic pumping tube in a second pressure<br />

pulse model.<br />

If an elastic cavity is inflated, the wall tension will cause the<br />

increase of the fluid pressure within. The ratio of volume increase<br />

per pressure unit is denoted compliance and is, if simplified as a<br />

linear behavior, given by eq 6.<br />

C ) ∆V<br />

∆p ) V i - V 0<br />

p i - p 0<br />

In consequence, the pressure in the tube ptube could be calculated<br />

according to eq 7.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(6)<br />

6987


ptube,i ) ∆V<br />

C ) Vi - V0 C<br />

Figure 5. Modeling of the pressure and flow pulse of pump 3 for the used<br />

flow resistances (A) tube 1 (49 cm, 0.27 mm i.d.), (B) tube 2 (30 cm, 0.42<br />

mm i.d.), (C) tube 3 (175 cm, 0.91 mm i.d.) with no tube elasticity.<br />

6988 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(7)<br />

The two flow rates out of the pump into the elastic tube QIn and<br />

out of the elastic tube QOut through the flow resistance were<br />

defined. QOut was calculated from the pressure in the elastic<br />

tube and the connected flow resistance analogous to eq 3. QIn<br />

was calculated according eq 8 distinguishing three cases. (1)<br />

The SMP is already empty and consequently both ppump and QIn<br />

are zero.<br />

(2) The pressure in the tubing and the SMP are equal so that<br />

the QIn is equal to QOut. (3) The ppump overcomes ptube so that<br />

QIn is limited by the pressure difference and the flow resistance<br />

of the outlet check valve defined by its passage radius rValve<br />

and its length LValve according to eq 3 and its time response T<br />

and QOut.<br />

Qin,i )<br />

ppump,i ) 0|) 0<br />

{if ppump,i > ptube,i | ) (ppump,i - ptube,i )(1 - e t/T ) r }<br />

4<br />

Valve π<br />

+ Qout,i 8ηLValve ppump,i ) ptube,i | ) Qtube,i (8)<br />

The accumulating volumes of expulsion from the SMP into the<br />

elastic tube VIn and from the compliance elastic tube through<br />

the flow resistance VOut were calculated according to eq 9.<br />

V i )Q i-1 ∆t + V i-1<br />

The inner volume of the elastic tube was calculated by eq 10 being<br />

a simple balance of the initial volume, the flow out QOut, and the<br />

flow in QIn.<br />

(9)<br />

V tube,i )V tube,i-1 + Q In,i-1 ∆t - Q Out,i-1 ∆t (10)<br />

3.6. Improvement of Pressure Resistance. Increasing the<br />

elasticity of the flow resistance connected to the SMP allows a<br />

fast flow out of the pulse volume. This decreases the pulsation<br />

character of the flow and leads to a stabilized but lower pressure<br />

in the system. Since the backpressure is proportional to the 2nd<br />

power of the flow rate, smoothing the flow pulsations by an<br />

integral element such as an elastic pumping tube could decrease<br />

the peak flow rate through the flow resistance and improve the<br />

OE.<br />

Table 2. Theoretical Pressure for the Tested Mean<br />

Flow Rates Using the Model Flow Resistances and<br />

Found Dynamic Inner Diameters<br />

flow rate<br />

[mL/min]<br />

tube 2<br />

(0.27 mm<br />

i.d. 49 cm)<br />

pressure [bar]<br />

tube 2<br />

(0.42 mm<br />

i.d. 30 cm)<br />

tube 3<br />

(0.91 mm<br />

i.d. 175 cm)<br />

0.2 0.13<br />

0.3 0.19 0.02 0.01<br />

0.5 0.31 0.03 0.01<br />

0.8 0.50 0.05 0.01<br />

1.0 0.63 0.07 0.02<br />

1.5 0.94 0.10 0.03<br />

2.0 1.25 0.13 0.03<br />

3.0 1.88 0.20 0.05<br />

5.0 0.33 0.09


Since the time for liquid expulsion increases with the pulse<br />

volume of the used pump, the contribution of any existing leakages<br />

through the inlet check valve of the SMP would increase as well.<br />

Allowing a fast flow-out against an apparently lower flow resistance<br />

can therefore increase the OE because after expulsion the<br />

resistance of two check valves would have to be overcome to<br />

enable a counter-directional flow. By this, the solenoid membrane<br />

can reach its final position of the deactivated status in less time<br />

and can aspirate a full pulse volume again. On the other hand,<br />

another producer of equivalent solenoid micropumps stated that<br />

“it is recommended to use hard tubing for piping” 16 since soft<br />

tubing might absorb the pressure pulse and affect the OE of the<br />

SMP.<br />

For testing the potential of pressure pulse dampers, two pieces<br />

of purple/black marked peristaltic pumping tube (2.05 mm i.d.)<br />

of 30 and 6.5 cm effective length were used and inserted between<br />

solenoid pump 4 and the model flow resistance tube 1. As shown<br />

in Figure 4D, the insertion of the shorter peristaltic pumping tube<br />

doubled the operation efficiency of the pump and a further<br />

improvement of 10% was possible using the longer peristaltic<br />

pumping tube. However, it was observed that the flow was delayed<br />

and lasted several seconds after finalization of the pumping<br />

operation.<br />

The compliance of the elastic tube, i.e., the volumetric<br />

expansions of the elastic tube in function of the inner pressure<br />

was an experimental parameter. It was measured by connecting<br />

the elastic tube and the former flow resistance tube 1 to the<br />

syringe pump and applying different flow rates. When the flow<br />

stops, the lasting flow out was collected and weighed. A linear<br />

relationship between pressure, calculated from the applied flow<br />

rates and resistance, and volume was found in the range of 0.3 to<br />

1.2 bar, leading to a compliance of 1055 µL/bar per meter of the<br />

elastic tube.<br />

It was not possible to measure neither the response time of<br />

the check valve nor the flow path dimensions during the flow<br />

out. Values used for the flow model 2 were estimated to be 1<br />

mm flow path length, 0.4 mm open diameter, and a delay time<br />

of 50 ms. The values have influence on the flow out velocity<br />

from the SMP into the elastic tube but little impact on the flowout<br />

through the flow resistance and would not affect the<br />

simulation of the contribution of the elastic tube of the OE of<br />

the SMP significantly.<br />

The simulation of four consecutive pressure pulses with a<br />

frequency of 0.33 Hz is shown in Figure 6 for neglecting any<br />

elasticity (A) and considering the 30 cm peristaltic pumping tube<br />

(B) for pump 4 and the flow resistance tube 1 applying time steps<br />

of 10 and 2 ms, respectively.<br />

It becomes clear that a single pressure pulse is insufficient<br />

to describe and explain why better OE was achieved using the<br />

elastic tube. However, under repeated operation and allowing<br />

a flow out after stopping the SMP operation renders higher<br />

OE values. Considering further a leakage through the inlet<br />

check valve or higher OE, the difference between both<br />

operation modes (with and without the elastic tube) would even<br />

become more pronounced.<br />

(16) http://www.takasago-elec.co.jp/en/product/pump.html, accessed February<br />

20, 2010.<br />

Figure 6. Modeling of the pressure and flow pulse of pump 4 for<br />

the used flow resistances (A) without elasticity included and (B) with<br />

an elastic pumping tube of 30 cm estimating a delay time of 50 ms.<br />

Further study would be required to optimize the dimensions<br />

of peristaltic pumping tubes as a function of the pulsation volume.<br />

In fact, we did not achieve better but worse operation efficiencies<br />

for the other pumps using the elastic pumping tubes of the given<br />

dimensions.<br />

Model 2 still shows some shortcomings: the aspiration step of<br />

the SMP was not included, and the flow resistances, delay times,<br />

and leakage loss of the check valves were not quantifiable but<br />

were estimated. However, both presented models were suited to<br />

describe the observations of SMP operation and thus present the<br />

first reported intent of SMP simulation.<br />

CONCLUSIONS<br />

Time adaptation and the rational use of a new economic relay<br />

card for the improvement of the operation reliability of SMP was<br />

presented. The OE being the ratio of the calibrated pulse volume<br />

and the experimental one was studied with the dependence of<br />

the connected flow resistance and activation time over a wide<br />

range of operation frequency. For the first time, the operation of<br />

SMP was simulated by pressure pulse models. The following<br />

conclusions and recommendations for future works using SMP<br />

can be made: (1) Activation and deactivation times should both<br />

exceed 150 ms. (2) The pulse volume has to be calibrated in the<br />

flow system at the desired flow rate since flow resistance and<br />

operation frequency both affect the OE considerably. (3) The use<br />

of SMP of small pulse volume at higher frequency promises higher<br />

OE than lower operation frequency using SMP with a larger pulse<br />

volume. (4) The pressure robustness of SMP passed the charac-<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6989


teristics of the producer at least twice. (5) The OE is mainly limited<br />

by the required time of pulse volume expulsion. (6) The use of<br />

elastic pumping tubes can render higher OE when the flow<br />

resistance is high.<br />

ACKNOWLEDGMENT<br />

B.H. was funded by a JAE postdoctoral fellowship from CSIC.<br />

The work was supported from Project CTQ2007-64331 funded by<br />

the MEC (Spanish Ministry of Education and Science).<br />

6990 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

SUPPORTING INFORMATION AVAILABLE<br />

Photo of the used 8-port relay card and scheme of the experiment<br />

for the determination of the back-spring constant. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 17, 2010. Accepted June 28,<br />

2010.<br />

AC101250H


Anal. Chem. 2010, 82, 6991–6999<br />

Integrated Microfluidic System for Rapid Forensic<br />

DNA Analysis: Sample Collection to DNA Profile<br />

Andrew J. Hopwood,* ,† Cedric Hurth, ‡ Jianing Yang, ‡ Zhi Cai, ‡ Nina Moran, † John G. Lee-Edghill, †<br />

Alan Nordquist, ‡ Ralf Lenigk, ‡ Matthew D. Estes, ‡ John P. Haley, † Colin R. McAlister, †<br />

Xiaojia Chen, ‡ Carla Brooks, ‡ Stan Smith, ‡ Keith Elliott, † Pieris Koumi, † Frederic Zenhausern,* ,‡<br />

and Gillian Tully †<br />

Research and Development, Forensic Science Service, Trident Court 2960 Solihull Parkway, Birmingham Business<br />

Park, Birmingham UK B37 7YN, and Center for Applied NanoBioscience and Medicine, The University of Arizona<br />

College of Medicine, 425 N. Fifth Street, Phoenix, Arizona 85004<br />

We demonstrate a conduit for the delivery of a step change<br />

in the DNA analysis process: A fully integrated instrument<br />

for the analysis of multiplex short tandem repeat DNA<br />

profiles from reference buccal samples is described and<br />

is suitable for the processing of such samples within a<br />

forensic environment such as a police custody suite or<br />

booking office. The instrument is loaded with a DNA<br />

processing cartridge which incorporates on-board pumps<br />

and valves which direct the delivery of sample and<br />

reagents to the various reaction chambers to allow DNA<br />

purification, amplification of the DNA by PCR, and collection<br />

of the amplified product for delivery to an integral<br />

CE chip. The fluorescently labeled product is separated<br />

using micro capillary electrophoresis with a resolution of<br />

1.2 base pairs (bp) allowing laser induced fluorescencebased<br />

detection of the amplified short tandem repeat<br />

fragments and subsequent analysis of data to produce a<br />

DNA profile which is compatible with the data format of<br />

the UK DNA database. The entire process from taking the<br />

sample from a suspect, to database compatible DNA<br />

profile production can currently be achieved in less than<br />

4 h. By integrating such an instrument and microfluidic<br />

cartridge with the forensic process, we believe it will be<br />

possible in the near future to process a DNA sample taken<br />

from an individual in police custody and compare the<br />

profile with the DNA profiles held on a DNA Database in<br />

as little as 3 h.<br />

DNA analysis in a forensic context relies on the extraction of<br />

DNA from a sample; quantification and normalization of the DNA;<br />

concurrent PCR-based amplification of a number of specific short<br />

tandem repeat loci and separation/detection of the PCR products,<br />

usually by capillary electrophoresis (CE); thereby allowing precise<br />

sizing of each fragment and accurate typing of the alleles in the<br />

DNA profile. This forensic process allows for the routine analysis<br />

of DNA samples from controls such as buccal swabs in 24-72 h<br />

* To whom correspondence should be addressed. E-mail: andy.hopwood@<br />

fss.pnn.police.uk (A.H.); Frederic.Zenhausern@arizona.edu (F.Z.).<br />

† Forensic Science Service.<br />

‡ The University of Arizona College of Medicine.<br />

from receipt of the sample. 1 Such processes are laboratory-based<br />

and require that the sample, once taken from a suspect in custody<br />

is transported to a laboratory for processing. For convenience,<br />

samples are usually stored and batched by the police force prior<br />

to dispatch to the lab, a process that can take between a few hours<br />

to a few weeks. Even with samples that are dispatched to the<br />

laboratory immediately, the suspect is highly likely to have been<br />

released from custody while the sample is processed. A sample<br />

taken from a suspect in the UK has a 2.3% chance of matching<br />

with a crime sample held on the National DNA Database<br />

(NDNAD) of the UK. 2 Furthermore, evidence suggests that<br />

individuals released on police bail have a high incidence of<br />

offending while on bail. 3 It would therefore be beneficial to the<br />

arresting agency for the individual to remain in custody while the<br />

DNA sample is processed and compared against the NDNAD.<br />

This would require the availability of both rapid DNA processing<br />

and real-time access to the national database.<br />

Within the laboratories of the FSS, current process allows for<br />

reference samples in urgent cases to be prioritized and, once<br />

delivered to the laboratory, these can be processed manually in<br />

as little as 8 h. However, this is a labor intensive and therefore<br />

relatively expensive process. The implementation of a rapid system<br />

whereby a reference sample can be processed within the police<br />

custody area would be of value to the law enforcement community:<br />

a suspect’s DNA sample could be processed and compared to a<br />

database of crime sample DNA profiles while the individual<br />

remains in custody, eliminating the need to locate and rearrest<br />

an individual in response to a match on the database. Rapid<br />

elimination of an individual from an investigation could also be<br />

achieved, freeing up resources for the investigation of alternative<br />

leads in a criminal case.<br />

A number of publications have reported approaches to the<br />

rapid analysis of DNA for forensic science and various methods<br />

have been described including the acceleration of specific parts<br />

(1) Hedman, J.; Albinsson, L.; Ansell, C.; Tapper, H.; Hansson, O.; Holgersson,<br />

S.; Ansell, R. Forensic Sci. Int.: Genet. 2008, 2, 184–189.<br />

(2) Annual Report of The National DNA Database 2007-2009. http://www.<br />

npia.police.uk/en/docs/NDNAD07-09-LR.pdf (Accessed March 2010).<br />

(3) Morgan, P. M.; Henderson, P. F. Home Office Report 184 1998 ISBN 1<br />

84082 074 8 http://www.homeoffice.gov.uk/rds/pdfs/hors184.pdf (Accessed<br />

Feb. 2010).<br />

10.1021/ac101355r © 2010 American <strong>Chemical</strong> Society 6991<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/15/2010


of the process such as PCR 4-6 or the optimization of the whole<br />

process which requires the intensive use of highly skilled<br />

technicians. 7 The application of Hybeacon technology which allows<br />

for a rapid analysis of STR loci from a crude lysate of buccal cells<br />

without the need for physical separation of the alleles 8-10 has<br />

shown some promise but has the drawback that the resolution of<br />

microvariant alleles is difficult and the analysis of all required<br />

alleles in single tube is not currently achievable. The development<br />

of integrated microfluidic systems, so-called micro total analysis<br />

systems, or microTAS for DNA analysis has been discussed for<br />

many years and a number of groups have demonstrated successful<br />

modules for DNA extraction, 11-14 PCR amplification, 15-18 and<br />

CE. 19-23 Integrated systems for extraction and PCR 24-26 or PCR<br />

and CE 27-29 have also been reported and illustrate that a fully<br />

integrated DNA analysis system should be feasible for forensic<br />

application. Indeed, full integration of all the constituent parts<br />

required for genetic analysis including the application of capillary<br />

(4) Vallone, P. M.; Hill, C. R.; Podini, D.; Butler, J. M. Forensic Sci. Int.: Genet.<br />

Suppl. 2009, 2, 111–112.<br />

(5) Vallone, P. M.; Hill, C. R.; Butler, J. M. Forensic Sci. Int.: Genet. 2008, 3,<br />

42–45.<br />

(6) Wang, D. Y.; Chang, C. W.; Hennessy, L. K. Forensic Sci. Int.: Genet. Suppl.<br />

2009, 2, 115–116.<br />

(7) Hopwood, A.; Fox, R.; Round, C.; Tsang, C.; Watson, S.; Rowlands, E.;<br />

Titmus, A.; Lee-Edghill, J.; Cursiter, L.; Proudlock, J.; McTernan, C.; Grigg,<br />

K.; Thornton, L.; Kimpton, C. Int. Congr. Ser. 2006, 1288, 639–641.<br />

(8) French, D. J.; McDowell, D. G.; Thomson, J. A.; Brown, T.; Debenham,<br />

P. G. Int. Congr. Ser. 2006, 1288, 707–709.<br />

(9) Gale, N.; French, D. J.; Howard, R. L.; McDowell, D. G.; Debenham, P. G.;<br />

Brown, T. Org. Biomol. Chem. 2000, 6, 4553–4559.<br />

(10) French, D. J.; Howard, R. L.; Gale, N.; Brown, T.; McDowell, D. G.;<br />

Debenham, P. G. Forensic Sci. Int.: Genet. 2008, 2, 333–339.<br />

(11) Ji, H. M.; Samper, V.; Chen, Y.; Hui, W. C.; Lye, H. J.; Mustafa, F. B.; Lee,<br />

A. C.; Cong, L.; Heng, C. K.; Lim, T. M. Sens. Actuators, A Phys. 2007,<br />

139, 139–144.<br />

(12) Wolfe, K. A.; Breadmore, M. C.; Ferrance, J. P.; Power, M. E.; Conroy,<br />

J. F.; Norris, P. M.; Landers, J. P. Electrophoresis 2002, 23, 727–733.<br />

(13) Breadmore, M. C.; Wolfe, K. A.; Arcibal, I. G.; Leung, W. K.; Dickson, D.;<br />

Giordano, B. C.; Power, M. E.; Ferrance, J. P.; Feldman, S. H.; Norris, P. M.;<br />

Landers, J. P. Anal. Chem. 2003, 75, 1880–1886.<br />

(14) Oakley, J. A.; Shaw, K. J.; Docker, P. T.; Dyer, C. E.; Greenman, J.;<br />

Greenway, G. M.; Haswell, S. J. Lab Chip 2009, 9, 1596–1600.<br />

(15) Obeid, P. J.; Christopoulos, T. K.; Crabtree, H. J.; Backhouse, C. J. Anal.<br />

Chem. 2003, 75, 288–295.<br />

(16) Yang, J.; Liu, Y.; Rauch, C.; Stevens, R. L.; Liu, R. H.; Lenigk, R.; Grodzinski,<br />

P. Lab Chip 2002, 2, 179–187.<br />

(17) Giordano, B. C.; Ferrance, J.; Swedberg, S.; Hühmer, A. F. R.; Landers,<br />

J. P. Anal. Biochem. 2001, 291, 124–132.<br />

(18) Simpson, P. C.; Roach, D.; Woolley, A. T.; Thorsen, T.; Johnston, R.;<br />

Sensabaugh, G. F.; Mathies, R. A. Proc. Natl. Acad. Sci. U.S.A. 1998, 95,<br />

2256–2261.<br />

(19) Medintz, I. L.; Paegel, B. M.; Mathies, R. A. J. Chromatogr., A 2001, 924,<br />

265–270.<br />

(20) Shi, Y.; Simpson, P. C.; Scherer, J. R.; Wexler, D.; Skibola, C.; Smith, M. T.;<br />

Mathies, R. A. Anal. Chem. 1999, 71, 5354–5361.<br />

(21) Shi, Y. Electrophoresis 2006, 27, 3703–3711.<br />

(22) Zheng, J.; Webster, J. R.; Mastrangelo, C. H.; Ugaz, V.; Burns, M. A.; Burke,<br />

D. T. Sens. Actuators, B 2007, 125, 343–351.<br />

(23) Goedecke, N.; McKenna, B.; El-Difrawy, S.; Carey, L.; Matsudaira, P.; Erlich,<br />

D. Electrophoresis 2004, 25, 1678–1686.<br />

(24) Lee, C. Y.; Lee, G. B.; Lin, J. L.; Huang, F. C.; Liao, C. S. J. Micromech.<br />

Microeng. 2005, 15, 1215–1223.<br />

(25) Legendre, L. A.; Bienvenue, J. M.; Roper, M. G.; Ferrance, J. P.; Landers,<br />

J. P. Anal. Chem. 2006, 78, 1444–1451.<br />

(26) Bienvenue, J. M.; Legendre, L. A.; Ferrance, J. P.; Landers, J. P. Forensic<br />

Sci. Int.: Genet. 2010, 4, 178–186.<br />

(27) Woolley, A. T.; Hadley, D.; Landre, P.; deMello, A. J.; Mathies, R. A.;<br />

Northrup, M. A. Anal. Chem. 1996, 68, 4081–4086.<br />

(28) Lagally, E. T.; Medintz, I.; Mathies, R. A. Anal. Chem. 2001, 73, 565–570.<br />

(29) Liu, P.; Yeung, S. H. I.; Crenshaw, K. A.; Crouse, C. A.; Scherer, J. A.;<br />

Mathies, R. A. Forensic Sci. Int. Gen. 2008, 2, 301–309.<br />

6992 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

electrophoresis has been demonstrated for the detection of<br />

Bacillus anthracis from whole blood 30 but to date we are unaware<br />

of any system which has been used to routinely produce an STR<br />

profile from DNA extraction to DNA profile with no operator<br />

intervention. Integrated microfluidic systems and their application<br />

to high-performance genetic analysis has been recently reviewed 31<br />

and will not be further discussed here.<br />

Within the UK, 75% of individuals arrested are processed and<br />

released from custody within 6 h and about 95% are processed<br />

within 24 h. 32 We defined our target for the process to be capable<br />

of processing the sample to give a DNA profile and returning any<br />

match from a database within 2 h such that the sample could be<br />

processed comfortably, and duplicated if required, within the 6 h<br />

window that a suspect is present in the custody suite.<br />

The fluidic cartridge-based system described here is built using<br />

the principle of closed architecture, which minimizes any opportunity<br />

for contamination of the samplesa critical requirement<br />

for the forensic applicationsand performs DNA extraction, DNA<br />

amplification, resolution of the STR alleles by CE and detection<br />

using laser induced fluorescence (LIF). The polycarbonate cartridge<br />

for DNA extraction, PCR and post PCR manipulation is<br />

prefilled with the reagents required for the entire process and<br />

simply clips into an adapter which also holds the glass micro<br />

capillary electrophoresis (µCE) chip used to resolve and detect<br />

the dye-labeled amplicons. The extraction-PCR cartridge is held<br />

in contact with an electronic circuit board which provides the<br />

functional control, allowing routine use by an unskilled technician<br />

with the minimum amount of training. Fluidic movement is fully<br />

automated and controlled using simple electro-chemical pumps<br />

and single use thermally activated valves as previously described,<br />

33,34 rather than more complex reusable thermally activated<br />

valves 35 avoiding the need for complex fittings and fixings<br />

to integrate the cartridge with the instrumentation, and eliminating<br />

the potential for sample contamination between the different<br />

functions of the cartridge. A servo-controlled magnet is activated<br />

as required for the collection of magnetic particles for DNA<br />

purification. Embedded resistive heaters are used to activate valves<br />

as previously described, 34 and Peltier devices employed for<br />

thermal cycling are programmed to activate in a specific sequence<br />

to facilitate processing. A miniaturized high voltage power supply<br />

is integrated to the hardware and allows for the separation of the<br />

STR amplicons through a matrix of polyvinylpyrolidone (PVP) and<br />

hydroxyethylcellulose (HEC). Data collected from the µCE is<br />

processed manually using commercially available software and<br />

the DNA profile is recorded in a format compatible with the data<br />

requirement for submission to the National DNA Database. The<br />

whole system is designed to allow simple loading of a crude cell<br />

lysate from a buccal scrape to the cartridge and robust walk-away<br />

(30) Easley, C.; Karlinsey, K.; Bienvenue, J.; Legendre, L.; Roper, M.; Feldman,<br />

S.; Hughes, M.; Hewlett, E.; Merkel, T.; Ferrance, J. Proc. Natl. Acad. Sci.<br />

U.S.A. 2006, 103, 19272–19277, Landers.<br />

(31) Liu, P.; Mathies, R. A. Trends Biotechnol. 2009, 27, 572–581.<br />

(32) Philips C. Her Majesty’s Stationary Office, 1981; ISBN 0101809212 ASIN:<br />

B002K6GPAQ.<br />

(33) Liu, R. H.; Bonanno, J.; Yang, J.; Lenigk, R.; Grodzinski, P. Sens. Actuators,<br />

B 2004, 98, 328–336.<br />

(34) Liu, R. H.; Yang, J.; Lenigk, R.; Bonanno, J.; Grodzinski, P. Anal. Chem.<br />

2004, 76, 1824–1831.<br />

(35) Pal, R.; Yang, M.; Johnson, B. N.; Burke, D. T.; Burns, M. A. Anal. Chem.<br />

2004, 76, 3740–3748.


Figure 1. Representation of the Cartridge for DNA extraction,<br />

amplification and post PCR denaturation showing P1-P4 ) electrochemical<br />

pumps; C1 ) lysate input chamber; C2 ) bead chamber;<br />

C3 ) mixing/incubation chamber; C4 ) washing and elution chamber;<br />

R ) PCR chamber; A ) DNA extract archive chamber; D )<br />

denaturation chamber; M ) bead storage chamber; B ) Binding buffer<br />

chamber; W ) wash solution storage chamber; E ) elution buffer<br />

storage chamber; F ) Formamide/ILS storage chamber; X ) output<br />

to CE. A closing valve is represented by b; an opening valve by 0;<br />

and a vent by 1.<br />

processing of the sample culminating in an STR profile which is<br />

compatible with The National DNA Database of the UK.<br />

EXPERIMENTAL SECTION<br />

Samples. Buccal samples were collected from consenting<br />

individuals using an OMNISwab (Whatman, Maidstone, UK) and<br />

processed immediately using the method described below.<br />

Cartridge Design and Fabrication. The cartridge device for<br />

DNA extraction, amplification and post PCR manipulation (Figure<br />

1) was made from 3 mm computed numerically controlled (CNC)<br />

machined polycarbonate (PC) plate stock and comprised a number<br />

of chambers for the storage of reagents and sample manipulation.<br />

The architecture of the fluidic system was designed such that<br />

outlets from one chamber to the next were typically on the bottom<br />

side of the chamber, thus taking advantage of gravity to ensure<br />

that any bubbles generated were removed from the fluid within<br />

the system. The channels linking the chambers were fitted with<br />

opening valves (O) and closing valves (V) formed from paraffin<br />

wax with a melting point of 65 °C (Sigma-Aldrich, Poole, UK) as<br />

previously described 33 which were hot dispensed into their<br />

respective positions. The exception to this was valve V15 which<br />

comprised two chambers connected vertically which were filled<br />

with different paraffin waxes; the lower chamber, closest to the<br />

Table 1. Reagents and Volumes Required for Cartridge<br />

Operation. All Reagents Were Added Prior to Loading<br />

the Cartridge into the Instrument<br />

chamber reagent volume required<br />

pump 1-3 0.25 M NaCl 600 µL<br />

pump 4 0.5 M NaCl 600 µL<br />

M ChargeSwitch (CS) beads 15 µL<br />

water 10 µL<br />

B CS purification buffer 30 µL<br />

W CS wash buffer 200 µL<br />

E CS elution buffer 150 µL<br />

C1 Buccal lysate 150 µL<br />

R PowerPlex ESI 16 multimix<br />

in ReaX beads<br />

two beads<br />

F Hi-Di formamide 45 µL<br />

ILS 500 CC5 5 µL<br />

Peltier was filled with H1 Sasol wax with a nominal melting<br />

temperature of 105 °C (Sasol Wax North America Corp., Hayward,<br />

CA), the second chamber contained the standard paraffin wax.<br />

The H1 wax prevented V15 firing prematurely from exposure to<br />

the heat generated during the thermal cycling of the Peltiers.<br />

Aluminum wire electrodes were installed into the pump chambers,<br />

sealed using Dymax 1180-M UV glue and cured using a Dymax<br />

Bluewave 200 UV Curing System (Dymax, Torrington, CT) in<br />

accordance with the manufacturer’s instructions. The PCR multimix<br />

was formulated into a ReaX bead (Q Chip Ltd., Cardiff, UK)<br />

and added to chamber R prior to binding the 0.5 mm PC cover to<br />

the 3 mm cartridge using a two sided pressure sensitive adhesive<br />

(90106 PSA, Adhesive Research, Glen Rock, PA) and pressing<br />

the assembly at 122 psi for 20 s and then 245 psi for 20 s. The<br />

ReaX bead-packaged multimix reagents are stable for at least 16<br />

weeks when stored at 4 °C (data not shown). The assembled<br />

cartridge was stored at 4 °C until required. Other reagents<br />

required for sample processing were added to the appropriate<br />

reservoirs immediately prior to use as detailed (Table 1). The<br />

single-use cartridge was inserted into a holder in a vertical position<br />

and the 140 mm reusable glass microCE chip was positioned in<br />

a horizontal plane (Figure 2A). The two components were<br />

connected via a 100 mm length of Teflon tubing with an internal<br />

diameter of 200 µm which was manually connected to the output<br />

channel of the plastic cartridge at X in Figure 1, and to the sample<br />

well of the adapter on the electrophoresis chip (Figure 2B) by<br />

sliding over short sections of PEEK tubing (IDEX Corporation,<br />

Northbrook, IL) with an outside diameter of 0.78 mm (Figure 2C)<br />

prior to loading into the instrument. The Teflon tubing was<br />

renewed for each sample.<br />

Electronic Control. The printed circuit board comprises<br />

embedded resistive heaters and control circuitry for managing<br />

the different operations required of the cartridge.<br />

The instrument houses the electronic control systems, the<br />

optical excitation and detection equipment and holds the printed<br />

circuit board to which the cartridge is precisely aligned and<br />

clamped, to allow robust connection to the electrodes for the<br />

electrochemical pumps and surface contact for valve firing by<br />

thermal transfer. The operation of the cartridge is as follows. The<br />

cells on the buccal swab sample were directly lysed by adding 1<br />

mL ChargeSwitch (Life Technologies, Inchinnen, UK) Lysis Buffer<br />

and 10 µL ChargeSwitch Proteinase K solution (20 mg/mL in 50<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6993


Figure 2. Assembly of the polycarbonate sample preparation cartridge and the borosilicate glass microCE chip. (A) The PC cartridge is positioned<br />

in a vertical plane, the glass CE chip in the horizontal plane. (B) Top image shows the adapter with wells for the electrodes. The PEEK tubing<br />

inlet can be clearly seen entering the elongated well of the sample well no. 1. Bottom image shows the adapter lid with gold-plated electrodes.<br />

(C) The adapter lid in position. The attached Teflon tubing can be seen attached to the PEEK inlet tube. (D) Representation of the adapters<br />

fitted to the CE chip and layout of the CE channel in relation to the wells.<br />

mM Tris-HCl, pH 8.5, 5 mM CaCl2, 50% glycerol) in a microcentrifuge<br />

tube and incubating at 60 °C for 15 min.<br />

An aliquot of 150 µL of the sample lysate was introduced into<br />

chamber C1 in the analysis cartridge using a manual pipet. Pump<br />

P1 was fired to pressurize the system and valve O1 was opened<br />

to direct the lysate via the binding buffer chamber B containing<br />

30 µL ChargeSwitch purification buffer, through the paramagnetic<br />

bead storage chamber M containing 25 µL of ChargeSwitch<br />

magnetic beads (1.87 mg/mL in 6 mM MES, pH 5.0, 6 mM NaCl)<br />

and into chamber C2, where the initial sample mixing occurred.<br />

The sample was then passed through two expansion chambers<br />

into chamber C3, where complete sample mixing was ensured<br />

by supplementary activation of the pump to produce a stream of<br />

bubbles which passed through the sample. The sample was<br />

incubated in this chamber for 3 min to facilitate DNA binding.<br />

The sample was then directed into chamber C4 where the<br />

ChargeSwitch magnetic beads were captured by a magnetic field<br />

of 450 ± 100 gauss generated via a servo activated magnet<br />

positioned behind the PCB. The supernatant liquid containing any<br />

unbound DNA and other lysis products was directed to the Waste<br />

chamber. The DNA-bound ChargeSwitch beads were then washed<br />

by flushing the beads with 200 µL of ChargeSwitch wash buffer<br />

from chamber W by activation of pump P2. One hundred and fifty<br />

µL of ChargeSwitch elution buffer (E5) was then directed to<br />

chamber C4 where the purified DNA was released from the beads<br />

by incubating at 60 °C for 3 min with the magnetic field removed.<br />

The beads were then recaptured by activation of the magnetic<br />

field and the eluted DNA solution was directed to the archive<br />

chamber A via the PCR chamber R by activation of pump P3.<br />

6994 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

The PCR chamber was designed to capture a 10 µL total<br />

volume of DNA solution and was preloaded with PowerPlex ESI<br />

16 PCR amplification multimix (Promega, Madison, WI) packaged<br />

in ReaX reagent beads. The PCR chamber was sealed from the<br />

rest of the cartridge prior to amplification by firing valves V13<br />

and V14.<br />

Amplification of the 16 loci multiplex PCR system was<br />

performed as follows: Activation of the PCR mix at 96 °C for 2<br />

min followed by 27 cycles of 94 °C for 30 s, 59 °C for 120 s, 72 °C<br />

for 90 s, and a final incubation at 60 °C for 45 min.<br />

The valves V13 and V14 were then reopened by heating and<br />

the liquid wax pushed through to the paraffin trap chamber O12,<br />

and valve V15 activated to close the PCR chamber bypass channel.<br />

The PCR solution was then collected by flowing 45 µL Hi-Di<br />

formamide (Applied Biosystems, Warrington, UK) containing 5<br />

µL internal size standard (ILS 500 CC5, Promega) from the<br />

formamide reservoir F using pump P4. The formamide/ILS/<br />

sample was collected in the denaturation chamber and incubated<br />

at 95 °C for 3 min prior to pumping to the CE chip.<br />

The CE microchip was custom-made from Schott Borofloat<br />

borosilicate glass (Micronit Microfluidics BV, Enschede, The<br />

Netherlands). The electrophoresis channel had a semielliptic<br />

cross-section and measured 50 µm wide and 20 µm deep. A<br />

machined polycarbonate adapter was interfaced with the powderblasted<br />

access holes on the microchip using UV-curable glue to<br />

provide buffer wells that could contain a maximum volume of 35<br />

µL, and a PEEK attachment point for the Teflon tubing from the<br />

cartridge performing DNA extraction and amplification. A pair of<br />

polycarbonate plates, one with three 1 mm diameter gold-coated


Table 2. Capillary Electrophoresis Injection and<br />

Separation Conditions. The sample is Presented to the<br />

Sample Well No. 1 and Injected into the Separation<br />

Column by a Gated Injection<br />

sample well<br />

no. 1 (V)<br />

sample<br />

waste well<br />

no. 2 (V)<br />

buffer well<br />

no. 3 (V)<br />

buffer<br />

waste well<br />

no. 4 (V) time (s)<br />

loading ground 800 ground 600 40<br />

injection ground 400 350 2000 2<br />

separation ground 400 ground 2700 1000<br />

pins (Figure 2B), one with a single pin were designed as adapter<br />

lids to clip onto the adapters to close the fluidic system (Figure<br />

2C) and connect to the control system. The reusable CE chip was<br />

prepared by flushing with water to remove the polymer from the<br />

previous run, followed by 1 M hydrochloric acid for 10 min and<br />

rinsing with water for 10 min to remove the acid prior to loading<br />

the polymer. The polymer, 3.5% w/v PVP/HEC in a 20/80 ratio<br />

in 1 × ABI 310 running buffer (Applied Biosystems) acted as both<br />

sieving and coating matrix 36 and was loaded for 25 min at 50 °C.<br />

Owing to the design of the cartridge system, the risk of<br />

contamination of the subsequent sample with PCR products from<br />

the CE chip was minimized by loading the reagents for the plastic<br />

cartridge before connecting it to the CE chip. A volume of 25 µL<br />

of 1 × ABI 310 running buffer was added to each of the wells of<br />

the electrophoresis chip prior to clipping the adapter lids in place.<br />

The sample was delivered into the sample well and injected into<br />

the separation channel using a gated injection scheme 37 prior to<br />

separation under the conditions described in Table 2 and the<br />

labeled fragments were detected at a point 110 mm from injection.<br />

The electrophoresis chip was maintained at a temperature of 50<br />

± 0.15 °C through direct contact with a heater plate.<br />

Laser-Induced Fluorescence (LIF) Detection. LIF detection<br />

was achieved with a modified confocal fluorescence setup. The<br />

excitation source, a diode-pumped solid-state (DPSS) laser (Calypso,<br />

Cobolt AB, Solna, Sweden) was cleaned-up using a 480-520<br />

nm band-pass filter (Omega Optical, Brattleboro, VT) and aligned<br />

onto a high-reflectivity elliptic mirror centered on a Raman laser<br />

reflector. The laser output was adjusted to provide 21 mW of<br />

power at 491 nm to the CE microchip and the incident light was<br />

focused onto the CE microchip using a ×40 objective (LucPLFLN,<br />

Olympus, Center Valley, PA), such that a circular 60 µm Gaussian<br />

beam hit the microchannel at a point 110 mm from the injection<br />

point. Alignment and focusing of the chip and detection system<br />

was achieved by a push-button activated servo motor and a<br />

micrometer screw respectively, on a 3 mm translation stage<br />

(Thorlabs, Newton, NJ).<br />

The emission wavelengths of the fluorescently labeled DNA<br />

fragments were collected onto a cooled charge-coupled device<br />

(CCD) camera (Newton 920, Andor, South Windsor, CT) mounted<br />

onto a high-throughput spectrometer (CP140-1605, Horiba Scientific,<br />

Newton, NJ) after a long-pass emission filter (cut-on: 520<br />

nm). The diffraction grating in the spectrometer disperses the<br />

different wavelengths in the emission spectrum over the length<br />

of the CCD, allowing a quantitative measurement of each color<br />

(36) Boulos, S.; Cabrices, O.; Blas, M.; McCord, B. R. Electrophoresis 2008,<br />

29, 4695–4708.<br />

(37) Ermakov, S. V.; Jacobson, S. C.; Ramsey, J. M. Anal. Chem. 2000, 72,<br />

3512–3517.<br />

in the sample over time. The instrument design is further<br />

described in Hurth et al. 38<br />

The raw spectral data from the CCD was deconvoluted to<br />

provide peak information for each of the five colors used in the<br />

multiplex analysis using NanoIdentity v 1.12 (SoftGenetics, State<br />

College, PA) and the data were further analyzed to provide<br />

genotype information using GeneMarker HID V 1.76 (Soft-<br />

Genetics).<br />

During optimization and testing of protocols, DNA quantification<br />

of the DNA extracted from the samples was performed using<br />

the Quantifiler Human DNA Quantification Kit (Applied Biosystems)<br />

in accordance with the manufacturer’s instructions.<br />

Electrophoresis of control samples was executed using the<br />

Applied Biosystems 3130xl in accordance with the PowerPlex ESI<br />

16 kit instructions.<br />

RESULTS AND DISCUSSION<br />

The aim of the process is to generate a DNA profile with all<br />

16 loci present in the electropherogram. To ensure this happens,<br />

a number of variables were identified as critical to the efficiency<br />

of the process: Sample collection; DNA extraction; PCR efficiency;<br />

PCR reaction recovery and robustness of the CE detection<br />

instrumentation. Each of these variables have been investigated<br />

and stabilized as far as possible.<br />

The sample input is the major cause of variation. Different<br />

samples taken from different individuals will have different<br />

quantities of cells, releasing different quantities of DNA upon lysis.<br />

For example, van Wieren-de Wijer et al observed total DNA yield<br />

ranging from 0.08 to 1078.0 µg (median 54.3 µg; mean 82.2 µg). 39<br />

Since there is no simple way to normalize the DNA concentration<br />

on the cartridge by dilution, it was essential to minimize this<br />

variation in DNA concentration extracted from the buccal swab<br />

by some means. The ChargeSwitch gDNA Normalized Buccal Cell<br />

Kit is claimed to produce a normalized yield of genomic DNA at<br />

a concentration of 1-3 ng/µL in150µL under standard conditions.<br />

40 Compared with other available DNA extraction chemistries<br />

the ChargeSwitch protocol is comparatively simple and the<br />

reagents were considered less likely to cause inhibition of the<br />

downstream processes. In our experience, with samples of freshly<br />

collected swabs extracted on cartridge as described above, a DNA<br />

concentration of between 0.47 and 1.92 ng/µL of DNA with a mean<br />

of 0.86 ± 0.41 ng/µL (n ) 27) was routinely attained.<br />

The amount of DNA produced by PCR amplification varied<br />

with a number of factors including the amount of DNA in the<br />

reaction, the total volume of the PCR and the number of cycles<br />

of PCR. The volume of DNA extract added to the reaction was<br />

governed by the size of the PCR chamber and the volume of the<br />

ReaX beads therein. The PCR chamber (Figure 3) was designed<br />

such that it is filled by the DNA solution traveling from the<br />

chamber C4 to the archive chamber A and the volume of the<br />

(38) Hurth, C.; Smith, S.; Nordquist, A.; Lenigk, R.; Surve, A.; Hopwood, A.;<br />

Haley, J.; Chen, X.; Estes, M.; Yang, J.; Cai, Z.; Lee-Edghill, J.; Moran, N.;<br />

Elliott, K.; Tully, G.; Zenhausern, F. Rev. Sci. Instrum. 2010, Submitted.<br />

(39) van Wieren-de Wijer, D. B. M. A.; Maitland-van der Zee, A. H.; de Boer, A.;<br />

Belitser, S. V.; Kroon, A. A.; de Leeuw, P. W.; Schiffers, P.; Janssen,<br />

R. G. J. H.; van Duijn, C. M.; Stricker, B. H. C. H.; Klungel, O. H. Eur. J.<br />

Epidemiol. 2009, 24, 677–682.<br />

(40) ChargeSwitch® gDNA Buccal Cell Kits. Instruction Manual. Invitrogen Life<br />

Technologies. http://tools.invitrogen.com/content/sfs/manuals/chargeswitch_<br />

gdna_buccal_man.pdf (Accessed March 2010).<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6995


Figure 3. Design of the PCR chamber fluidic system. The DNA<br />

solution enters the chamber from the right and fills the chamber. When<br />

the liquid level reaches the constriction channel in the top left-hand<br />

side of the chamber, the pressure required to pass through this<br />

constriction is much greater than that required for the solution to<br />

traverse the bypass. The DNA solution therefore passes over the<br />

bridge, leaving a precise volume of DNA in the PCR chamber<br />

(chamber highlighted by red solution). Following PCR amplification,<br />

the reaction volume is recovered by closing the bridge at Y using<br />

valve V15 and flushing the formamide through the chamber, collecting<br />

the PCR product.<br />

chamber was filled until the solution reached the restriction<br />

channel. The pressure required to breach this channel was greater<br />

than that required for the solution to flow over the bypass and<br />

consequently the flow took that direction, leaving a precisely<br />

measured volume of PCR reaction (9.96 ± 0.21 µL (n ) 20)) in<br />

the PCR chamber. The amount of DNA in the 10 µL PCR chamber<br />

therefore approximated 5.2 ± 2.5 ng, based upon the ReaX bead<br />

volume of 4 µL.<br />

Samples recovered from the PCR chamber and run on the AB<br />

3130xl gave similar peak heights to the control samples amplified<br />

in tube (data not shown).<br />

Following amplification, V15 is closed and the whole of the<br />

PCR sample is recovered from the PCR chamber and delivered<br />

with ILS 500 CC5 in formamide to the loading well of the<br />

electrophoresis chip. We rely on surface effects to smoothly<br />

remove the vast majority of the PCR reaction from the chamber.<br />

As the formamide is driven from its storage chamber, the air in<br />

the channel in front of the formamide is pushed through the PCR<br />

chamber and removes the bulk (80-90%) of the PCR reaction<br />

through the constriction. The formamide then fills the chamber,<br />

taking the remaining PCR solution with it. The surface tension at<br />

this small scale ensures that the bulk of the formamide is also<br />

flushed from the chamber leaving just one or two microlitres in<br />

the 90° corners of PCR chamber. The volume delivered to<br />

chamber D varied between cartridges with a mean volume of 25.1<br />

± 4.8 µL (n ) 25) with a range of 15-35 µL. The variation in<br />

volume was due to three things: (1) The wax in the valves 13 and<br />

14 is molten as the liquid flows by, allowing some liquid to mix<br />

with the wax. This liquid is trapped in the system as the wax<br />

solidifies. (2) Liquid is occasionally lost in the valve structures of<br />

V13 and/or V14. (3) Our channels are machined rather than<br />

injection molded, creating a rough surface that can retain 1-2<br />

µL per inch of channel. The volume of sample released from PCR,<br />

denatured in chamber D and delivered to the µCE is a variable<br />

that could have a significant effect on the final result: If only half<br />

of the PCR volume of a given PCR product was recovered we<br />

would expect to observe lower peak heights in the profile. The<br />

6996 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

system was therefore designed to push the PCR mixture out of<br />

the PCR chamber prior to flushing the chamber with formamide<br />

and we believe that the majority of the PCR product is recovered,<br />

minimizing any variability from this function of the cartridge.<br />

Sample loss in the channels, wax and valve housings in the circuit<br />

before the PCR chamber will reduce the amount of formamide<br />

and size standard reaching the denaturation chamber while the<br />

inefficiencies in the circuitry post PCR chamber would reduce<br />

the volume of PCR product and formamide in proportion to their<br />

volumetric ratios. Current laboratory practice is that the sample<br />

is heat denatured at 95 °C for 3 min and snap-cooled on ice prior<br />

to loading onto the 3130xl platform for injection to the CE. While<br />

the sample was easily heat denatured on the cartridge, no facility<br />

for snap cooling was available and the sample was loaded directly<br />

into the CE. Evaluation of the profiles demonstrated that this<br />

process was acceptable because the DNA profiles produced with<br />

no snap cooling were found to be comparable to those that had<br />

been snap cooled. The electropherogram from the internal lane<br />

standard (GeneScan 500 LIZ) diluted to a volume ratio of 9:1 with<br />

Hi-Di formamide (Applied Biosystems) and heat-denatured at 95<br />

°C for 5 min was used to assess the separation efficiency of the<br />

µCE module. The chromatographic resolution R was calculated<br />

by fitting the peaks with a Gaussian curve to obtain the migration<br />

time, t i, and the full-width at half-maximum, wi for peak i<br />

according to the following equations: 41<br />

R ) (2 ln 2) 1/2 t 2 - t 1<br />

w 1 + w 2<br />

And subsequently: Rbp ) ∆bp/R expresses the resolution in base<br />

pairs, that is, the minimum base pair separation of two peaks<br />

that would still appear separated at the baseline in the<br />

electropherogram. Figure 4A displays an example of an electropherogram<br />

which represents the “binned” data prior to deconvolution<br />

into the target wavelengths, that is, the electropherogram<br />

represents the sum of photon counts over time, and represents<br />

the whole of the signal collected by the CCD by time. Figure 4B<br />

shows the evolution of the resolution Rbp on the µCE system as<br />

a function of the known peak size (red trace) and the<br />

comparison with an electropherogram obtained on a commercial<br />

ABI 310 capillary electrophoresis analyzer (Applied<br />

Biosystems) using the same polymer matrix at 65 °C with a<br />

cathode potential of 10.3 kV. The resolution obtained on the<br />

µCE module is very close to that of a commercial CE apparatus.<br />

Typically, the resolution values obtained for a GeneScan 500<br />

LIZ varied between 1.02 and 1.24 bp for peaks within the 140-160<br />

bp range. Ongoing work aims at further improving the resolution<br />

by (1) using a pinched injection and (2) tailoring the composition<br />

of the HEC/PVP mixture for use in a microchip rather than a<br />

capillary.<br />

A series of allelic ladder runs was recorded on the CE<br />

subsystem using a CC5-labeled ILS 500 size standard (Promega)<br />

to estimate the run-to-run stability of the system. Consecutive runs<br />

were acquired on three different microchips to determine the<br />

variability between microchips. Figure 5 shows a typical sized run<br />

processed with GeneMarker HID 1.76. Inserts B though E are<br />

(41) Manabe, T.; Chen, N.; Terabe, S.; Yohda, M.; Endo, I. Anal. Chem. 1994,<br />

66, 4243–4252.


Figure 4. (A) Typical raw electropherogram obtained for a GeneScan<br />

500 LIZ run on the CE subsystem using a 140 mm microchip at 50<br />

°C. (B) Comparison of the resolution (in base pair) Rbp of the µCE<br />

system (red trace; 2) to that obtained on a 310 apparatus (Applied<br />

Biosystems) at 65 °C in a 36 cm capillary (blue trace; 1).<br />

close-ups of the [7-9] allele region (304-310 bp) on the D22S1045<br />

marker, the [10-11] region (146-150 bp) on the D18S51 marker,<br />

the [9-11] region (97-105 bp) on the TH01 marker, and the<br />

[21.2-26] region (177-195 bp) on the FGA marker respectively.<br />

Table 3 summarizes the observed shifts in base pair sizing for<br />

five different allelic ladder runs using a number of alleles from<br />

Figure 5B-E. The maximum observed variation around the<br />

tabulated value in the positioning of an allele is only about 0.2 bp<br />

on the TH01 marker, 0.7 bp on the D18S51 marker, 0.4 bp on the<br />

FGA marker, and 0.7 bp on the D22S1045 marker. The variability<br />

increases with the fragment size as expected from the absolute<br />

chromatographic resolution of the CE subsystem given in Figure<br />

4B. The allele sizing data gained from running a minimum of five<br />

allelic ladders of known designations was used to define the<br />

expected sizes of the unknown alleles in the amplified DNA<br />

samples.<br />

Integrated Sample Run. We were able to obtain STR<br />

profiles from lysed cells from buccal swabs using a disposable<br />

plastic cartridge attached to a reusable glass microchip device<br />

without any manual interruption of the programmed DNA<br />

extraction, purification, amplification, transfer and CE analysis<br />

process. A typical electropherogram including the ILS is shown<br />

in Figure 6 after applying GeneMarker HID 1.76 to size the<br />

migrated DNA fragments. The DNA profiles obtained as an<br />

integrated automated run were always similar to those obtained<br />

on an Applied Biosystems ABI 3130xl analyzer in terms of the<br />

number and position of the peaks and their relative intensities.<br />

Currently we are unable to routinely label all the peaks with a<br />

definitive allele designation as the migration of the peaks of<br />

some samples has given out of window peaks. We believe this<br />

may be due to a difference in the time base of the instrumentation<br />

and the NanoIdentity software: The software assumes an<br />

equal time base per data point while the instrumentation<br />

collects data on a variable time base which averages 0.21 s with<br />

a maximum of 0.35 s and a minimum of 0.07 s. An updated<br />

Figure 5. (A) Powerplex ESI16 allelic ladder run using CC5 ILS500 size standard (Promega, Madison, WI) sized using GeneMarker HID 1.76<br />

(SoftGenetics LLC, State College, PA) with appropriate panels and bins provided by Promega. (B) Close-up of the [7-9] allele region (304-310<br />

bp) on the D22S1045 marker. (C) Close-up of the [10-11] region (146-150 bp) on the D18S51 marker. (D) Close-up of the [9-11] region<br />

(97-105 bp) on the TH01 marker. (E) Close-up of the [21.2-26] region (177-195 bp) on the FGA marker.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6997


Table 3. Evolution of the Sizing Results for Specific Regions of an Allelic Ladder Run for Three Different Chips and<br />

Five Different Runs a<br />

D22S1045 size of allele no. 7 (bp) size of allele no. 8 (bp)<br />

chip 1 - 1st load 304.2 307.3<br />

chip 2 - 1st load 305.6 308.2<br />

chip 2 - 2nd load 304.1 307.2<br />

chip 3 - 1st load 304.3 307.2<br />

chip 3 - 2nd load 304.0 307.2<br />

D18S51 size of allele no. 10 (bp) size of allele no. 10.2 (bp) size of allele no. 11 (bp)<br />

chip 1 - 1st load 146.1 148.2 150.2<br />

chip 2 - 1st load 147.1 149.2 151.1<br />

chip 2 - 2nd load 146.3 148.3 150.2<br />

chip 3 - 1st load 145.8 147.7 149.7<br />

chip 3 - 2nd load 145.9 147.8 150.0<br />

TH01 size of allele no. 9.3 (bp) size of allele no. 10 (bp)<br />

chip 1 - 1st load 100.2 101.2<br />

chip 2 - 1st load 100.3 101.3<br />

chip 2 - 2nd load 100.2 101.1<br />

chip 3 - 1st load 100.1 101.4<br />

chip 3 - 2nd load 100.2 101.2<br />

FGA size of allele size of allele<br />

no. 21.2 (bp) no. 22 (bp)<br />

size of allele size of allele<br />

no. 22.2 (bp) no. 23 (bp)<br />

size of allele size of allele<br />

no. 23.2 (bp) no. 24 (bp)<br />

size of allele size of allele<br />

no. 24.2 (bp) no. 25 (bp)<br />

size of allele size of allele<br />

no. 25.2 (bp) no. 26 (bp)<br />

chip 1 - 1st load 176.8 178.8 180.8 182.8 184.8 186.8 188.9 190.8 192.8 194.8<br />

chip 2 - 1st load 177.2 179.3 181.0 183.0 184.2 186.0 188.0 189.8 191.6 193.3<br />

chip 2 - 2nd load 177.0 178.9 180.7 182.7 184.5 186.6 188.8 190.7 192.6 194.7<br />

chip 3 - 1st load 176.8 178.6 180.7 182.7 184.5 186.6 188.8 190.7 192.6 194.7<br />

chip 3 - 2nd load 176.6 178.5 180.6 182.5 184.7 186.6 188.6 190.6 192.7 194.5<br />

a The values represent the calculated size of the designated allele in DNA base pairs relative to the ILS in the same run. The markers are<br />

chosen to represent 1 marker per color, cover most of the regions of interest [100-300 bp], and correspond to the most challenging peaks to<br />

separate.<br />

Figure 6. Micro Capillary Electrophoresis profile from a fully integrated run. DNA was purified and amplified on-cartridge from lysed cells from<br />

a buccal swab and transferred automatically via Teflon tubing to the glass microchip. Allele designation was performed using GeneMarker HID<br />

1.76 (SoftGenetics LLC).<br />

version of the software is in development which will use the<br />

time file provided by the instrument and so align the data<br />

correctly on a true time scale.<br />

6998 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

CONCLUSION<br />

We have demonstrated a successful microTAS approach to<br />

the delivery of rapid DNA technology suitable for the forensic


field. DNA profiles were delivered from a crude DNA lysate<br />

with no manual input. The automated process is simple to use<br />

and provided DNA profiles which were concordant with the<br />

expected DNA profile of the donor. Our target is to deliver an<br />

evidential quality DNA profile from sample to profile in less<br />

than 2 h. While we currently have the capability of processing<br />

the sample in under 4 h, we believe that there are a number of<br />

areas where we can speed up the process. A major time saving<br />

(approximately 1.5 h) can be made simply by reducing the PCR<br />

cycling parameters from those published here to thermal<br />

profiles similar to those described previously. 4–6 Optimizing<br />

the cartridge operating process to allow parallel processing of<br />

valve firings and pump control will further reduce the processing<br />

time and we believe we will soon be able to meet the target<br />

time of 2 h for delivery of the genotyped DNA profile. The turn<br />

around between samples is approximately 10 min to load the<br />

solutions into the single-use plastic cartridge, and 75 min for<br />

preparation of the CE chip. This can be carried out while the<br />

previous sample is running, provided that a number of CE chips<br />

are available, reducing the turn around time to little more than<br />

5 min. The resolution obtained on the µCE module is very close<br />

to that of a commercial CE apparatus and the reproducibility<br />

of fragment migration was good: With allelic ladders the<br />

maximum variation in the positioning of an allele was around<br />

0.2 bp on the TH01 locus, 0.7 bp on the D18S51 locus, 0.4 bp<br />

on the FGA locus, and 0.7 bp on the D22S1045 locus. The<br />

migration of DNA profiles from human samples does require<br />

some improvement if the system is to be used for probative<br />

samples as we were unable to successfully type all integrated<br />

runs due to a small variation in the apparent migration of<br />

fragments. This variation is likely to be minimized by implementation<br />

of a new version of the NanoIdentity software which<br />

will take the true time file from the instrument, rather than<br />

assuming all data points are equidistant in time. The largest<br />

variation in the process remains the amount of template DNA<br />

input to the PCR reaction, and while this impacts the peak<br />

heights observed in the electropherogram, this does not affect<br />

the accuracy of the genotyping of a sample.<br />

Simple substitution of the multiplex chemistry in the cartridge<br />

will allow the production of DNA profiles suitable for loading to<br />

any established national DNA database. However, the rapid<br />

delivery of a DNA profile, whether at a crime scene, or in a<br />

custody suite is only a part of the puzzle. If the full impact of these<br />

new technologies is to be gained, the rapid delivery of chemistry<br />

must be supported by a capability to submit and compare the DNA<br />

profile in near real time. Currently neither the CoDIS database,<br />

nor the UK National DNA database have the capability to support<br />

rapid chemistry protocols.<br />

While the operation of the instrument is straightforward, a<br />

good level of training is required before the instrument can be<br />

operated but we believe that any individual with a basic scientific<br />

education could become competent in the operation of the<br />

instrument. Our vision for the future is to have a single fully<br />

integrated injection-molded cartridge with reagents preloaded with<br />

capability for DNA purification, amplification and µCE-based<br />

separation. Such a cartridge, coupled with automatic channel<br />

finding and focusing will minimize set up time and make the<br />

system usable by an individual with very basic training.<br />

ACKNOWLEDGMENT<br />

We gratefully acknowledge the technical assistance of Glen<br />

McCarty, David Nguyen, and Brett Duane for electronics design<br />

and prototyping; Baiju Thomas, Keith Burt, and Matthew Barrett<br />

for plastic device fabrication; Dr. Jian Gu and Peter East for<br />

microfabrication; and Amol Surve for instrumentation design. We<br />

also thank Bruce McCord for the provision of the separation<br />

matrix and helpful discussion around microCE and Promega<br />

Corporation for the generous gifts of PowerPlex ESI 16 STR kits<br />

and ILS 500 CC5.<br />

Received for review May 24, 2010. Accepted July 1, 2010.<br />

AC101355R<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

6999


Anal. Chem. 2010, 82, 7000–7007<br />

Visualization and Recovery of the (Bio)chemical<br />

Interesting Variables in Data Analysis with Support<br />

Vector Machine Classification<br />

Patrick W. T. Krooshof, † Bülent Üstün, ‡ Geert J. Postma, † and Lutgarde M. C. Buydens* ,†<br />

Radboud University Nijmegen, Institute for Molecules and Materials, <strong>Analytical</strong> <strong>Chemistry</strong>, P.O. Box 9010,<br />

6500 GL Nijmegen, The Netherlands, and <strong>Analytical</strong> Sciences <strong>Chemical</strong>s, Quality Unit API/Biotech, MSD Oss, P.O.<br />

Box 20, 5340 BH Oss, The Netherlands<br />

Support vector machines (SVMs) have become a popular<br />

technique in the chemometrics and bioinformatics field,<br />

and other fields, for the classification of complex data sets.<br />

Especially because SVMs are able to model nonlinear<br />

relationships, the usage of this technique has increased<br />

substantially. This modeling is obtained by mapping the<br />

data in a higher-dimensional feature space. The disadvantage<br />

of such a transformation is, however, that information<br />

about the contribution of the original variables in<br />

the classification is lost. In this paper we introduce an<br />

innovative method which can retrieve the information<br />

about the variables of complex data sets. We apply the<br />

proposed method to several benchmark data sets and a<br />

metabolomics data set to illustrate that we can determine<br />

the contribution of the original variables in SVM classifications.<br />

The corresponding visualization of the contribution<br />

of the variables can assist in a better understanding<br />

of the underlying chemical or biological process.<br />

In the past decade support vector machines (SVMs) have<br />

become a popular technique in pattern recognition and regression<br />

estimation. Applications of SVMs are among the fields of<br />

bioinformatics, 1,2 medicine, 3-6 drug discovery, 7-9 text categorizing,<br />

10 gene expression analysis, 11-13 face recognition, 14 spam<br />

* To whom correspondence should be addressed. Phone: +31 24 3653180.<br />

Fax: +31 24 3652653. E-mail: l.buydens@science.ru.nl.<br />

† Radboud University Nijmegen.<br />

‡ <strong>Analytical</strong> Sciences <strong>Chemical</strong>s.<br />

(1) Yang, Z. R. Briefings Bioinf. 2004, 5 (4), 328–338.<br />

(2) Ramo, P.; Sacher, R.; Snijder, B.; Begemann, B.; Pelkmans, L. Bioinformatics.<br />

2009, 25, 3028–3030.<br />

(3) Akay, M. F. Expert Syst. Appl 2009, 36 (2), 3240–3247.<br />

(4) Magnin, B.; et al. Neuroradiology. 2009, 51 (2), 73–83.<br />

(5) Luts, J.; Heerschap, A.; Suykens, J. A. K.; Van Huffel, S. Artif. Intell. Med.<br />

2007, 40 (2), 87–102.<br />

(6) Conforti, D.; Guido, R. Comput. Oper. Res. 2010, 37, 1389–1394.<br />

(7) Burbidge, R.; Trotter, M.; Buxton, B.; Holden, S. Comput. Chem. 2001,<br />

26, 5–14.<br />

(8) Warmuth, M. K.; et al. J. Chem. Inf. Comput. Sci. 2003, 43, 667–673.<br />

(9) Zernov, V. V.; Balakin, K. V.; Ivaschenko, A. A.; Savchuk, N. P.; Pletnev,<br />

I. V. J. Chem. Inf. Comput. Sci. 2003, 43, 2048–2056.<br />

(10) Leopold, E.; Kindermann, J. Mach. Learn. 2002, 46, 423–444.<br />

(11) Furey, T. S.; et al. Bioinformatics. 2000, 16 (10), 906–914.<br />

(12) Clarke, R.; et al. Nat. Rev. Cancer. 2008, 8, 37–49.<br />

(13) Noble, W. S. Nat. Biotechnol. 2006, 24 (12), 1565–1567.<br />

(14) Guo, G.; Li, S. Z.; Chan, K. L. Image Visualization Comput. 2001, 19, 631–<br />

638.<br />

7000 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

categorizing, 15,16 financial forecasting, 17,18 and many others.<br />

Especially because of the possibility to model complex nonlinear<br />

relationships the application of SVMs has grown substantially. 19-22<br />

By transforming the original input space into a high dimensional<br />

feature space, the nonlinear relationships can be presented in a<br />

linear form. This transformation is performed by using a specific<br />

kernel function. 6,23,24 Several kernel functions are proposed in the<br />

literature for this purpose and include variance-covariance based<br />

linear and polynomial kernels, the Euclidean distance based radial<br />

basis function (RBF) and the Pearson VII Universal Kernel (PUK)<br />

functions. 23-25 The transformation by a kernel function has also<br />

been introduced in other algorithms, such as Kernel Principal<br />

Component Analysis, 26 Kernel Partial Least Squares, 27,28 and<br />

Kernel Fisher Discrimination. 29<br />

However, the disadvantage of using such a kernel function is<br />

that the correlation between the obtained SVM model and the<br />

original input space is lost. Therefore it is not possible to<br />

determine which variables (e.g., spectral ranges) contribute to<br />

the final SVM results and a direct interpretation of the SVM model<br />

is not straightforward. 30,31 This seriously hampers the ultimate<br />

(bio)chemical interpretation of the resulting classification model.<br />

In this manuscript we propose a novel method to overcome this<br />

problem and reveal the importance of the original variables.<br />

(15) Drucker, H.; Wu, D.; Vapnik, V. N. IEEE Trans. Neural Networks 1999,<br />

10 (5), 1048–1054.<br />

(16) Guzella, T. S.; Caminhas, W. M. Expert Syst. Appl. 2009, 36 (7), 10206–<br />

10222.<br />

(17) Tay, F. E. H.; Cao, L. Omega. 2001, 29, 309–317.<br />

(18) Kim, K. J. Neurocomputing. 2003, 55, 307–319.<br />

(19) Vapnik, V. Estimation of Dependence Based on Empirical Data; Springer<br />

Verlag: New York, 1982.<br />

(20) Vapnik, V. the Nature of Statistical Learning Theory; Springer Verlag: New<br />

York, 1995.<br />

(21) Cortes, C.; Vapnik, V. Mach. Learn. 1995, 20, 273–297.<br />

(22) Vapnik, V. Statistical Learning Theory; John Wiley and Sons: New York,<br />

1998.<br />

(23) Cristianini, N.; Shawe-Taylor, J. An Introduction to Support Vector Machines<br />

and Other Kernel-Based Learning Methods; Cambridge University Press.:<br />

Cambridge, 2000.<br />

(24) Schölkopf, B.; Smola, A. J. Learning with Kernels; MIT Press.: Cambridge,<br />

2002.<br />

(25) Üstün, B.; Melssen, W. J.; Buydens, L. M. C. Chemom. Intell. Lab. Syst.<br />

2006, 81, 29–40.<br />

(26) Schölkopf, B.; Smola, A. J.; Müller, K. R. Neural Comput. 1998, 10, 1299–<br />

1319.<br />

(27) Walczak, B.; Massart, D. L. Anal. Chim. Acta 1996, 331, 177–185.<br />

(28) Rosipal, R.; Trejo, L. J. J. Mach. Learn. Res. 2001, 2, 97–123.<br />

(29) Mika, S.; Rãtsch, G.; Wetson, J.; Schölkopf, B.; Müller, K. R. Proc. NNSP′99;<br />

1999, 41–48.<br />

10.1021/ac101338y © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/20/2010


Another approach to handle this limitation is automatic relevance<br />

determination, 32,33 which was developed as a feature selection<br />

procedure in SVM models. Even though this method selects the<br />

variables that are important for the model, the relation between<br />

the variables and, for example, the class separation is not<br />

visualized. Generally, researchers report the high performance<br />

obtained by using a SVM model for classification and regression<br />

estimation, but do not comment on the relationship between the<br />

input variables and the modeled output data. SVM is often used<br />

as a black box approach.<br />

In this paper we present an innovative approach to open this<br />

black box for the SVM classifier and give insight in the transformation<br />

by the kernel function to make the SVM model more<br />

transparent. This approach is based on the nonlinear biplot<br />

principles described by Gower and Harding in 1988 34 and is used<br />

to visualize and determine the importance and influence of the<br />

input variables to the final SVM classifier. The resulting information<br />

can then be used to reduce the number of input variables to<br />

improve the performance or to reduce the complexity of the<br />

model. Furthermore, the visualization of the contribution of the<br />

original variables can assist in a better understanding of<br />

the underlying chemical or biological process. We will present<br />

the proposed approach, and illustrate and validate the methodology<br />

by applying it for classification problems: two benchmark data<br />

sets and a relevant metabolomics data set obtained from magnetic<br />

resonance spectroscopic images to diagnose human brain tumors.<br />

The effectiveness of the method is verified by a comparison of<br />

the classification performance obtained by using the entire set of<br />

input variables and a selection of variables, determined by our<br />

approach.<br />

EXPERIMENTAL SECTION<br />

Theory and Computational Strategy. As the theory of SVMs<br />

is described extensively in the literature 13,21-24 and the use of a<br />

kernel transformation results in the loss of information about the<br />

input variables, we will focus in the next section briefly on the<br />

concepts of the kernel function. Subsequently, we will explain<br />

the basic steps of the nonlinear biplot technique, as described by<br />

Gower and Harding, 34 to discuss the proposed method to<br />

determine the contribution of the input variables in SVM classifications.<br />

The full theory and algorithm of the proposed method<br />

can be found in the Supporting Information (SI).<br />

Kernel Transformations. In the various kernel-based methods<br />

a specific mapping function is used to project the original input<br />

data in a higher-dimensional feature space. 24 The data in this new<br />

feature space can subsequently be used as a new input for pattern<br />

recognition. The advantage of such an approach is that the method<br />

can deal with complex nonlinear problems. The typical (twodimensional)<br />

example to illustrate this approach is shown in the<br />

inset of Figure 1. This data set (X) contains two classes that we<br />

would like to separate.<br />

(30) Üstün, B.; Melssen, W. J.; Buydens, L. M. C. Anal. Chim. Acta 2007, 595,<br />

299–309.<br />

(31) Devos, O.; Ruckebusch, C.; Durand, A.; Duponchel, L.; Huvenne, J. P.<br />

Chemom. Intell. Lab. Syst. 2009, 96, 27–33.<br />

(32) Van Gestel, T.; Suykens, J. A. K.; De Moor, B.; Vandewalle, J. Proc. Eur.<br />

Symp. Artif. Neural Networks. 2001, 13–18.<br />

(33) MacKay, D. J. C. In Neural Networks and Machine Learning, NATO Asi<br />

Series. Series F, Computer and Systems Sciences 168; Bishop, C. M., Ed.;<br />

Springer: Berlin, 1998; pp 133-165.<br />

(34) Gower, J. C.; Harding, S. A. Biometrika 1988, 75, 445–455.<br />

Figure 1. The synthetic benchmark data set. Mapping of a twodimensional<br />

data set which contains two nonlinearly separable classes<br />

(outer and inner circles) into a three-dimensional feature space. The<br />

transformation is performed by applying the nonlinear mapping<br />

function φ(x.1, x.2) ) (x.1 2 ,�2x.1x.2, x.2 2 ), which makes the two classes<br />

linearly separable.<br />

The inner circle represents one class and the outer circle<br />

represents the other class. From the figure it is obvious that we<br />

are not able to separate the classes by a linear model. However,<br />

if we project this data in a higher dimensional space by using a<br />

mapping function, we are able to obtain the three-dimensional<br />

feature space which is represented in Figure 1. In this feature<br />

space the two classes can be linearly separated by a plane between<br />

the circles. In this case the transformation is performed by the<br />

mapping function φ. It has been shown that, since the (nonlinear)<br />

mapping function is in general unknown beforehand and is difficult<br />

to determine, the feature space can be constructed implicitly by<br />

invoking a generic kernel function (see, e.g., refs 23-25, 35). Such<br />

a kernel function is a function (K) which operates on two vectors,<br />

such that<br />

K(x i. , x j. ) ) 〈�(x i. ), �(x j. )〉 (1)<br />

where xi. and xj. are two objects in the data set and φ represents<br />

the actual nonlinear mapping function. The use of a kernel<br />

function makes it unnecessary to know the actual underlying<br />

feature map in order to be able to construct a linear model in<br />

the feature space. The application of such a kernel function<br />

will result in a square symmetric matrix: the kernel matrix K.<br />

This matrix is a weighted dissimilarity matrix, of which each<br />

position represents a dissimilarity (distance) measure between<br />

two objects. The specific kernel function and the optimal<br />

parameter settings of the function are determined in combination<br />

with the applied classification algorithm (e.g., SVM) by<br />

optimizing the classification performance. However, because<br />

the kernel function transforms the original input space into a<br />

feature space with a higher dimension, information about the<br />

original variables is not preserved.<br />

Nonlinear Biplots. To explain the concepts of nonlinear biplots,<br />

we will first describe the classical biplots technique which is used<br />

extensively in (chemometric) data analysis.<br />

(35) Gunn, S. R. Support Vector Machines for Classification and Regression.<br />

Technical Report; Image Speech and Intelligent Systems Research Group,<br />

University of Southampton: Southampton, 1997.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7001


Figure 2. Biplots of the Iris benchmark data set. The scores (representing the samples) are visualized as symbols, whereas the loadings are<br />

visualized by the vectors. The loadings are obtained by (a) projection of the first two columns in the loading matrix and (b) projection of the<br />

scores of 10 pseudosamples (for each variable).<br />

A biplot is a visualization in which the samples and the<br />

variables of a data set are represented together. This technique<br />

has been introduced by Gabriel in 1971 36 and is based on singular<br />

value decomposition (SVD) or principal component analysis<br />

(PCA) 37 of the column mean-centered data matrix X, containing<br />

n samples and m variables. By SVD X is decomposed into scores<br />

and loadings according to<br />

T<br />

X (n×m) ) U (n×r) Λ (r×r) V (m×r)<br />

T<br />

) S (n×r) L (m×r)<br />

The positions of the samples (rows of X) in a two-dimensional<br />

biplot are subsequently given by the elements in the columns of<br />

UΛ, often called the scores S. The variables (columns of X) are<br />

usually represented as vectors pointing from the origin to the<br />

coordinates that are given by the elements of the columns of V,<br />

called the loadings L. To construct a biplot for the first two<br />

singular vectors (or principal components, PCs), the elements in<br />

the first two columns of S and in the first two columns of L are<br />

used. This is illustrated for the Iris data set (see the Iris<br />

benchmark Data Set section) in Figure 2a. The coordinates of a<br />

new sample in this biplot can be obtained by premultiplying the<br />

sample as a row vector with V:<br />

x (1×m) V (m×2) ) s (1×2)<br />

The key idea leading to the nonlinear biplot is the interpretation<br />

of the loadings in the classical biplot (the representation of the<br />

variables) as projections of special so-called pseudosamples that<br />

carry all their weight in one variable. This means that these<br />

pseudosamples have a value of 0 for all variables, except for one<br />

variable. Projecting this pseudosample in the two-dimensional PCA<br />

plot yields coordinates that are equal to the loadings of the variable<br />

whose weight it carries. For example, [1, 0, 0, ..., 0] is a (1 × m)<br />

pseudosample with a value 1 for variable 1 and a value 0 for all<br />

other variables. Then<br />

[1, 0, 0, ..., 0] (1×m) V (m×2) ) s (1×2) ) v (1×2)<br />

(36) Gabriel, K. R. Biometrika. 1971, 58, 453–467.<br />

(37) Massart, D. L. Handbook of Chemometrics and Qualimetrics: Part A; Elsevier<br />

Science Publishers: Amsterdam, 1997.<br />

7002 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(2)<br />

(3)<br />

(4)<br />

where v is the first row of V and contains exactly the loading of<br />

variable 1, defining its position in the biplot. If the value of 1 is<br />

replaced by p different values z, a total of p pseudosamples are<br />

obtained. The projection of these pseudosamples in the PCA plot<br />

will result in a trajectory along the direction of the variable vector,<br />

as illustrated in Figure 2b.<br />

Gower et al. 34 extend this idea to the visualization of variable<br />

information in principle coordinate analysis. In principal coordinate<br />

analysis (or classical metrics scaling) an SVD is performed not<br />

on the original rectangular data matrix X, but on the symmetric<br />

matrix of squared Euclidean distances. This approach allows the<br />

visualization of the relative distances of the objects. It is often<br />

remarked that the information of the original variables (such as<br />

in a biplot) is lost. Gower et al., however, showed that this can be<br />

overcome by using the concept of trajectories of pseudosamples,<br />

as explained earlier. To make this approach feasible, one must<br />

be able to calculate the squared Euclidean distances of the<br />

pseudosamples with all other samples (e.g., data X). By projecting<br />

the rows of the resulting distance matrix in the principal<br />

component space, the trajectories that represent the variables are<br />

obtained. Gower et al. showed that this concept can be extended<br />

from Euclidean distances to many nonlinear distance metrics, as<br />

long as the same distance metric can be used to calculate the<br />

distances of the pseudosamples to the original data samples. In<br />

the classical linear biplot, only one pseudosample per variable is<br />

sufficient to represent the variables. For a nonlinear distance<br />

metric, the trajectory will be curved and multiple pseudosamples<br />

per variable are required.<br />

Nonlinear Biplots in SVM Classification. We propose to apply<br />

the above approach to the kernel-based SVM method, since the<br />

kernel is a (nonlinear) distance metric between objects. This<br />

procedure comprises the following steps (shortened):<br />

0. Optimize the kernel function and parameter settings for the<br />

SVM classification used (resulting in the kernel function K(xi.,xj.)).<br />

1. From the data X calculate the kernel matrix K, by using<br />

the optimized kernel function. This matrix is subsequently<br />

centered, resulting in matrix K c of size n × n.<br />

2. Apply SVD on K c and construct different score plots (for<br />

the various combinations of principal components) for the n<br />

samples in K c . Inspect these score plots to find the direction(s)


(principal components, PCs) in which maximum class separation<br />

is obtained. Note that this is not necessarily along the first<br />

PCs.<br />

3. Construct a matrix Pj for the j th variable which contains p<br />

pseudosamples. The range of these pseudosample values zj<br />

should vary between the minimum and maximum value of the<br />

original variable.<br />

4. Apply the kernel function K(pi.,xj.) to the pseudosample<br />

data Pj to obtain the kernel matrix of the pseudosamples C<br />

(i.e., calculate the kernel distances of the pseudo samples to<br />

the original samples).<br />

5. Project the rows of the centered pseudosample kernel matrix<br />

C c in the score plot that was found in step 2.<br />

6. Repeat steps 3-5 for each variable j.<br />

The resulting plot contains a trajectory of pseudosamples for<br />

each original variable. These trajectories yield information about<br />

the relative contribution of the variables to the SVM classification.<br />

The curvature of a trajectory indicates a nonlinear kernel<br />

transformation.<br />

SVM Classification and Validation Procedure. Each data set was<br />

analyzed by SVM using the RBF kernel. The RBF kernel was<br />

chosen because of its simplicity for optimization. Because the SVM<br />

application that we have used is a binary classifier, each different<br />

class in a data set that contains more than two classes was<br />

analyzed by a one-against-all approach. 38 The kernel parameters<br />

parameter σ and the SVM parameter C (C is a cost parameter 23 )<br />

were optimized by leave-10%-out cross-validation.<br />

All calculations are performed in the software program Matlab<br />

(The MathWorks Inc.) version 6.5 release 13. The nonlinear biplot<br />

approach was implemented in Matlab by using the commercially<br />

available PLS toolbox from eigenvector Research Inc.<br />

Data Sets. To illustrate the applicability of the proposed<br />

method we have used several benchmark data sets (synthetic and<br />

real) and a metabolomics data set based on Magnetic Resonance<br />

Spectroscopic Imaging (MRSI) data.<br />

Synthetic Benchmark Data Set. A synthetic data set was<br />

constructed based on the data presented in Figure 1, which is<br />

generally used to illustrate the power of SVM classification. The<br />

data set contains two variables to represent an inner and outer<br />

circle, both consisting of 50 objects. The two circles can not be<br />

separated in a linear way and therefore a kernel-based classification<br />

method is required to separate the data.<br />

To construct the Synthetic data set we added five variables<br />

containing random noise (normally distributed) to the data. The<br />

variance of these five variables was set to be about 10% of the<br />

variance of the two variables which represent the circles. Because<br />

noise contains no information, the added variables do not<br />

contribute to the classification performance. This data set was<br />

constructed to confirm that the first two variables (representing<br />

the circles) are only (equally) important in the SVM classification.<br />

Iris Data Set. A widely used benchmark data set to exemplify<br />

discriminant and cluster analysis is the data published by Fisher<br />

in 1936. 39 This Iris data set contains fifty specimens of each of<br />

the three species Iris Setosa, Iris Versicolor, and Iris Virginica,<br />

resulting in a total of 150 samples. Four properties of the species<br />

are measured to determine differences between the three classes.<br />

(38) Suykens, J. A. K.; Van Gestel, T.; De Brabanter, J.; De Moor, B.; Vandewalle,<br />

J. Least Squares Support Vector Machines; World Scientific: Singapore, 1999.<br />

(39) Fisher, R. A. Annu. Eugen. 1936, 7, 179–188.<br />

These properties (representing four variables) are Sepal Length,<br />

Sepal Width, Petal Length, and Petal Width, all measured in<br />

millimeters. The Setosa class can easily be separated from the<br />

other two classes using only one variable (either Petal Length or<br />

Petal Width). The two other classes are partly overlapping, as<br />

shown in Figure 2.<br />

We will use the Iris data set to demonstrate the applicability<br />

of our proposed method to determine the most discriminative<br />

variable for the three species.<br />

Metabolomics Data Set. To illustrate the proposed method for<br />

variable selection on a more complex data set, we have used a<br />

metabolomics data set which consists of magnetic resonance<br />

spectroscopic (MRS) spectra obtained from MRSI data. 40 This data<br />

set was constructed during a European project called Interpret,<br />

which was funded by the European Commission to develop new<br />

methodologies to automatically classify tumors in the human brain<br />

(see http://azizu.uab.es/INTERPRET). Data from a total of 24<br />

patients and 4 volunteers were acquired by MRS at different<br />

positions in the brain, according to an acquisition protocol defined<br />

by the Interpret Consortium. The study was approved by the<br />

ethical committee and followed the rules of the World Health<br />

Organization. After reaching consensus about the histopathology,<br />

three tumor types were identified according to the World Health<br />

Organization classification system. These three classes contained<br />

glial tumors with different grades: Grade II (10 cases), Grade III<br />

(4 cases), and Grade IV (7 cases). A fourth class consists of spectra<br />

acquired from patients with Meningioma (3 cases). Additionally,<br />

a class consisting of Healthy tissues was created from patient (4<br />

cases) and volunteer (4 cases) data. For each predefined class a<br />

selection of spectra from the different patients was made. Only<br />

spectra acquired at regions which clearly consisted of tissue<br />

belonging to the particular class were selected. The data for the<br />

Healthy class was selected from the volunteers or from the contralateral<br />

brain region of the patients. 41 The resulting data set<br />

contains 569 spectra, consisting of five different classes. Each<br />

spectrum contains 229 data points, covering the chemical shifts<br />

between 4.0 and 0.5 ppm. Details about the acquisition parameters<br />

and preprocessing of the data are described in Simonetti et al 42<br />

and are beyond the scope of this paper.<br />

RESULTS<br />

Synthetic Benchmark Data Set. Classification of the Synthetic<br />

data set by using SVMs resulted in a leave-10%-out crossvalidated<br />

accuracy of 100%. This accuracy (see also SI Table 1)<br />

illustrates that SVMs are able to separate the two circles. As noise<br />

contains no information, the particular noise-variables should not<br />

have contributed to the class separation. To verify the contribution<br />

of each variable in the final classification first we have to find<br />

and visualize the optimal (linear) separation between the classes<br />

in the resulting kernel matrix. Because the Synthetic data set<br />

contains one hundred objects, the feature space in which the<br />

original data is mapped (i.e., the space of the kernel matrix)<br />

consists as a consequence of one hundred dimensions. Therefore<br />

PCA is used to reduce the dimensionality of the feature space.<br />

Analysis of the score plots of combinations of only the first five<br />

(40) Barker, P. B.; Lin, D. D. M. Prog. Nucl. Magn. Reson. Spectrosc. 2006, 49,<br />

99–128.<br />

(41) Simonetti, A. W.; et al. Anal. Chem. 2003, 75, 5352–5361.<br />

(42) Simonetti, A. W.; et al. NMR Biomed. 2005, 18, 34–43.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7003


Figure 3. Principle component score plots of the synthetic benchmark data set. Pairs-plot of the first five principal components (PCs) are<br />

shown. The separation between the inner circle (blue +-symbol) and the outer circle (red o-symbol) is obtained on PC 5.<br />

Figure 4. Relevant score plot and pseudo sample trajectories for the Synthetic benchmark data set. (a) Projection of the objects of the Synthetic<br />

benchmark data set (after the kernel transformation) in the feature space spanned by PC 1 and PC 5. (b) Pseudosample trajectories projected<br />

in the same feature space. The trajectories of the two variables representing the two circles are indicated by the numbers “1” and “2”. The five<br />

noise-variables (variables “3” to “7”) are distributed around the origin of the plot and are therefore not clearly visible.<br />

PCs show that “PC 5” can be used to visualize the class separation<br />

of the two classes. The pairs-plot for the first five PCs of the<br />

Synthetic data set is given in Figure 3. Apparently, the variance<br />

which accounts for the separation between the classes is captured<br />

by PC 5. The contribution of each original variable can be<br />

determined by projecting the kernel matrix of the corresponding<br />

pseudosamples in the feature space spanned by PC 5 and any of<br />

the other PCs and by subsequently analyzing the obtained<br />

trajectories.<br />

The pseudosample matrices were constructed by varying the<br />

intensities of the individual variables over twenty objects (uni-<br />

7004 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

formly distributed). The application of the kernel function resulted<br />

in seven kernel matrices (one for each variable) of size (20 × 100).<br />

These matrices were then projected in the feature space obtained<br />

by PCA on the original kernel matrix K. The trajectories of the<br />

seven variables obtained by the proposed method are visualized<br />

for the space spanned by “PC 1” and “PC 5” in Figure 4b. As<br />

shown, the trajectories of the five noise-variables (variables 3-7)<br />

are along the direction of the class separation (PC 5), but are<br />

relatively small compared to the two trajectories of the variables<br />

representing the inner and outer circle (variables 1 and 2). This<br />

indicates that the noise-variables have a small contribution to the


Figure 5. Relevant score plots and pseudo sample trajectories for the Iris benchmark data set. Projection of the objects of the Iris data set<br />

after kernel transformation and PCA, to illustrate the optimal separation of the (a) Setosa, (b) Versicolor, and the (c) Virginica class (red o-symbols).<br />

The pseudosample trajectories are projected in the same feature space for the (d) Setosa, (e) Versicolor, and the (f) Virginica class. The<br />

trajectory of each variable is indicated by a different number: “1” for Sepal Length, “2” for Sepal Width, “3” for Petal Length, and “4” for Petal<br />

Width.<br />

class separation, which is in accordance with our hypothesis. As<br />

the trajectories of variable 1 and 2 have a similar length on PC 5,<br />

both variables are important for the class separation. A (linear)<br />

classification based on only one variable is therefore not possible.<br />

A visual inspection of the Synthetic data set (see the Theory and<br />

Computational Strategy section) already confirmed this conclusion.<br />

Inspection of the trajectories in the score plots of combinations<br />

of other PCs showed that the variables representing the two circles<br />

can have a relatively small loading in the particular feature space<br />

(results not shown). For example, in the space spanned by “PC<br />

2” and “PC 4” variable 4 (representing noise) has the largest<br />

loading compared to the other variables. However, no class<br />

separation was found in this particular score-plot (see Figure 3)<br />

and therefore this feature space is not informative.<br />

Iris Benchmark Data Set. The application of the three<br />

possible one-against-all classifications resulted in cross-validated<br />

accuracies of at least 96.7% (see also SI Table 1). After PCA was<br />

applied to the three kernel matrices (for each one-against-all<br />

classifier), we searched for the PCs resulting in the optimal<br />

separation between the classes in the pairs-plots. These optimal<br />

separations are visualized in Figure 5a-c and as shown the Setosa<br />

class is completely separated from the other classes by the<br />

variance captured by PC 1. The Versicolor and Virginica class<br />

requires two principal components to capture the variance which<br />

accounts for the class separation, that is, PC 2 and PC 5 for<br />

Versicolor and PC 1 and PC 2 for the Virginica class.<br />

To determine the contribution of the original variables, pseudosample<br />

trajectories were constructed and projected in the space<br />

spanned by the corresponding PCs. These projections are visual-<br />

ized in Figure 5d-f and as illustrated the variables Petal Length<br />

(variable 3) and Petal Width (variable 4) are most discriminative<br />

for the classifications, confirming our hypothesis. However, the<br />

length of the trajectory of Petal Width is much shorter compared<br />

to Petal Length (for unscaled data) and is therefore less important<br />

for the classifications. To confirm these observations we have<br />

applied SVM to the Iris data set by including only Petal Length.<br />

The resulting cross-validated accuracies are given in SI Table 2.<br />

Although the accuracies for the Versicolor and Virginica class are<br />

somewhat lower compared to the results where all the variables<br />

are used (95.3% versus 96.7%), the accuracies are still comparable.<br />

If one of the other variables was chosen for the classification, that<br />

is, Sepal Width, the accuracies are much lower (


Figure 6. Relevant score plot and pseudo sample trajectories for the metabolomics data set. (a) Space spanned by PC 1 and PC 3, obtained<br />

by PCA applied on the kernel matrix of the MRSI data set. The Healthy class is indicated by red o-symbols and the tumor classes by blue<br />

+-symbols. (b) Pseudosample trajectories, projected in the same feature space as in (a). Each trajectory is indicated by a different number and<br />

represents a different variable. The trajectories are also color-coded using the variable numbers, resulting in similar colors for trajectories of<br />

variables representing specific regions in the NMR spectrum.<br />

One group of trajectories consists of pseudosamples representing<br />

the variable numbers 170-184 (1.40-1.19 ppm; the trajectories<br />

in the direction of PC1, colored orange in Figure 6b), and the<br />

other group the variables 129-132 (2.03-1.98 ppm; the trajectories<br />

more or less in the direction of PC3, colored green in Figure<br />

6b). Because the group of variables within 2.03-1.98 ppm is along<br />

the direction of the class separation we postulate that these<br />

variables have a large contribution to the SVM classification. This<br />

observation corresponds to results published in the literature, in<br />

which the researchers stated that this specific region corresponds<br />

to N-acetyl-aspartate, which is a neuronal marker for viable<br />

neurons, and that the concentration is reduced or absent in most<br />

brain tumors. 43<br />

Note that the other group of trajectories, representing the<br />

variables 170-184 (1.40-1.19 ppm), corresponds to lactate and<br />

lipids regions of the spectra. These trajectories are located in the<br />

direction along PC 1 and direction corresponds to increasing<br />

tumor grade. This observation also corresponds to the results of<br />

Howe et al. 43<br />

With only the variables selected within the region of 2.03-1.98<br />

ppm (four variables), the SVM classification results in a leave-<br />

10%-out cross-validated accuracy of 95.4% (see SI Table 2). This<br />

value is in agreement with the accuracy obtained when all the<br />

variables are included in the SVM model (98.7%), indicating that<br />

these variables have a large contribution to the classification. If a<br />

set of four variables was selected randomly, the cross-validated<br />

accuracy was


CONCLUSIONS<br />

In this paper we have introduced a new method to successfully<br />

visualize and identify the important variables in the classification<br />

by a kernel-based method. By constructing a set of pseudosamples<br />

we are able to determine the effect of the kernel function to the<br />

individual variables. We have applied the proposed method to<br />

several synthetic and real data sets and have shown that our<br />

method is able to find and visualize the most discriminative<br />

variables. To confirm this conclusion, we have selected only the<br />

most discriminative variables and reapplied the proposed method.<br />

The cross-validated classification accuracies of these reduced data<br />

sets are similar to the accuracies obtained by using the data sets<br />

with the full set of variables. This illustrates the validity of our<br />

method to determine the most important variables for classifica-<br />

tion purposes by kernel-based techniques. The interpretation of<br />

SVM models with our proposed visualization method has the<br />

potential to extract relevant chemical and biological knowledge<br />

from complex data, such as omics data. This will greatly enhance<br />

the applicability of the powerful SVM classifiers.<br />

SUPPORTING INFORMATION AVAILABLE<br />

List of symbols, discussion of theory and computational<br />

strategy, two additional figures, and two additional tables. This<br />

material is available free of charge via the Internet at<br />

http://pubs.acs.org.<br />

Received for review May 26, 2010. Accepted July 7, 2010.<br />

AC101338Y<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7007


Anal. Chem. 2010, 82, 7008–7014<br />

Rapid and Sensitive Detection of Protein<br />

Biomarker Using a Portable Fluorescence<br />

Biosensor Based on Quantum Dots and<br />

a Lateral Flow Test Strip<br />

Zhaohui Li, Ying Wang, Jun Wang, Zhiwen Tang, Joel G. Pounds, and Yuehe Lin*<br />

Pacific Northwest National Laboratory, Richland, Washington 99362<br />

A portable fluorescence biosensor with rapid and ultrasensitive<br />

response for protein biomarker has been built<br />

up with quantum dots and a lateral flow test strip. The<br />

superior signal brightness and high photostability of<br />

quantum dots are combined with the promising advantages<br />

of a lateral flow test strip and result in high<br />

sensitivity and selectivity and speed for protein detection.<br />

Nitrated ceruloplasmin, a significant biomarker for cardiovascular<br />

disease, lung cancer, and stress response to<br />

smoking, was used as model protein biomarker to demonstrate<br />

the good performances of this proposed quantum<br />

dot-based lateral flow test strip. Quantitative detection of<br />

nitrated ceruloplasmin was realized by recording the<br />

fluorescence intensity of quantum dots captured on the<br />

test line. Under optimal conditions, this portable fluorescence<br />

biosensor displays rapid responses for nitrated<br />

ceruloplasmin with the concentration as low as 1 ng/mL.<br />

Furthermore, the biosensor was successfully utilized for<br />

spiked human plasma sample detection in a wide dynamic<br />

range with a detection limit of 8 ng/mL (S/N ) 3). The<br />

results demonstrate that the quantum dot-based lateral<br />

flow test strip is capable of rapid, sensitive, and quantitative<br />

detection of nitrated ceruloplasmin and hold a great<br />

promise for point-of-care and in field analysis of other<br />

protein biomarkers.<br />

Rapid and quantitative detection of protein biomarkers is<br />

extremely important for clinical diagnostics, basic discovery, and<br />

a variety of other biomedical applications. As is well-known, a host<br />

of various immunoassays for protein detection have been developed<br />

during the past years. The more established approaches<br />

include microsphere-based arrays, 1 proteome chips, 2 radioimmunoassay,<br />

3,4 surface plasma resonance, 5 microfluidic systems, 6,7<br />

* To whom correspondence should be addressed. Email: yuehe.lin@<br />

pnl.gov. Fax: 1-509-3766242. Tel: 1-509-3716241.<br />

(1) Blicharz, T. M.; Siqueira, W. L.; Helmerhorst, E. J.; Oppenheim, F. G.;<br />

Wexler, P. J.; Little, F. F.; Walt, D. R. Anal. Chem. 2009, 81, 2106–2114.<br />

(2) Zhu, H.; Bilgin, M.; Bangham, R.; Hall, D.; Casamayor, A.; Bertone, P.;<br />

Lan, N.; Jansen, R.; Bidlingmaier, S.; Houfek, T.; Mitchell, T.; Miller, P.;<br />

Dean, R. A.; Gerstein, M.; Snyder, M. Science 2001, 293, 2101–2105.<br />

(3) Wide, L.; Porath, J. Biochim. Biophys. Acta 1966, 130, 257–260.<br />

(4) Harris, E. N.; Boey, M. L.; Mackworthyoung, C. G.; Gharavi, A. E.; Patel,<br />

B. M.; Loizou, S.; Hughes, G. R. V. Lancet 1983, 2, 1211–1214.<br />

(5) Lyon, L. A.; Musick, M. D.; Natan, M. J. Anal. Chem. 1998, 70, 5177–<br />

5183.<br />

7008 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

enzyme-linked immunosorbent assay (ELISA), 8,9 surface-enhanced<br />

Raman scattering, 10 etc. Although these conventional strategies<br />

exhibit promising results for sensitive detection of proteins, there<br />

are still some hindrances including the utilization of radioactive<br />

substances, time-consuming sample purification, incubation and<br />

washing steps before analysis, and specialized equipment. Recently,<br />

a lateral flow test strip (LFTS), also called a dry-reagent<br />

strip biosensor, has been becoming a powerful tool for protein<br />

analysis and has attracted increasing attention. 11-17 In comparison<br />

to the methods mentioned above, LFTS permits a one-step and<br />

rapid low-cost analysis as well as a user-friendly format, very short<br />

assay time, and no requirement for skilled technicians. The most<br />

widely used format of LFTS is the employment of gold nanoparticles<br />

as reporters for colorimetric detection, 18-21 which was either<br />

qualitative or semiquantitative and generally utilized for analyzing<br />

proteins with relatively high concentrations. In order to meet the<br />

requirement of sensitive protein detection, more quantitative LFTS<br />

has been developed recently using various reporters, such as<br />

(6) Zhou, L. J.; Wang, K. M.; Tan, W. H.; Chen, Y. Q.; Zuo, X. B.; Wen, J. H.;<br />

Liu, B.; Tang, H. X.; He, L. F.; Yang, X. H. Anal. Chem. 2006, 78, 6246–<br />

6251.<br />

(7) Duffy, D. C.; McDonald, J. C.; Schueller, O. J. A.; Whitesides, G. M. Anal.<br />

Chem. 1998, 70, 4974–4984.<br />

(8) Engvall, E.; Perlmann, P. Immunochemistry 1971, 8, 871–874.<br />

(9) Kurita, R.; Arai, K.; Nakamoto, K.; Kato, D.; Niwa, O. Anal. Chem. 2010,<br />

82, 1692–1697.<br />

(10) Grubisha, D. S.; Lipert, R. J.; Park, H. Y.; Driskell, J.; Porter, M. D. Anal.<br />

Chem. 2003, 75, 5936–5943.<br />

(11) Millipore, C. A Short Guide: Developing Immunochromatographic Test Strips;<br />

Millipore Corp.: Millipore, MA, 1996.<br />

(12) Hjelle, B.; Jenison, S.; Torrez Martinez, N.; Herring, B.; Quan, S.; Polito,<br />

A.; Pichuantes, S.; Yamada, T.; Morris, C.; Elgh, F.; Lee, H. W.; Artsob, H.;<br />

Dinello, R. J. Clin. Microbiol. 1997, 35, 600–608.<br />

(13) Sorell, L.; Garrote, J. A.; Acevedo, B.; Arranz, E. Lancet 2002, 359, 945–<br />

946.<br />

(14) Liu, G.; Lin, Y. Y.; Wang, J.; Wu, H.; Wai, C. M.; Lin, Y. Anal. Chem. 2007,<br />

79, 7644–7653.<br />

(15) Ho, J. A. A.; Wauchope, R. D. Anal. Chem. 2002, 74, 1493–1496.<br />

(16) Lou, S. C.; Patel, C.; Ching, S. F.; Gordon, J. Clin. Chem. 1993, 39, 619–<br />

624.<br />

(17) Xu, H.; Mao, X.; Zeng, Q. X.; Wang, S. F.; Kawde, A. N.; Liu, G. D. Anal.<br />

Chem. 2009, 81, 669–675.<br />

(18) Shyu, R. H.; Shyu, H. F.; Liu, H. W.; Tang, S. S. Toxicon 2002, 40, 255–<br />

258.<br />

(19) Kim, J. H.; Cho, J. H.; Cha, G. S.; Lee, C. W.; Kim, H. B.; Paek, S. H. Biosens.<br />

Bioelectron. 2000, 14, 907–915.<br />

(20) Choi, D. H.; Lee, S. Ki.; Oh, Y. K.; Bae, B. W.; Lee, S. D.; Kim, S.; Shin,<br />

Y. B.; Kim, M. G. Biosens. Bioelectron. 2010, 25, 1999–2002.<br />

(21) Tanaka, R.; Yuhi, T.; Nagatani, N.; Endo, T.; Kerman, K.; Takamura, Y.;<br />

Tamiya, E. Anal. Bioanal. Chem. 2006, 385, 1414–1420.<br />

10.1021/ac101405a © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/19/2010


electroactive-species loaded liposome, 22 metal ion chelates, 23 and<br />

inorganic nanoparticles 24 as well as fluorescent dyes. 25,26 Because<br />

of its inherently high sensitivity, fluorescence-based LFTS (FLFTS)<br />

might be the most promising for quantifying trace amount of<br />

proteins. To date, most of the FLFTS for protein detection are<br />

mainly based on organic fluorophores. 27,28 However, these fluorophores<br />

have intrinsic limitations such as photobleaching, which<br />

seriously compromises its sensitivity and confines its further<br />

applications. Consequently, a novel fluorescent label with ideal<br />

photostability and immense brightness is highly desirable for<br />

FLFTS applications in protein analysis with ultralow concentrations.<br />

During the past decades, the development of quantum dots<br />

(Qdot) has been of considerable interest in many areas, from<br />

molecular and cellular biology to molecular imaging and medical<br />

diagnostics. 29-35 Compared with organic fluorophores, Qdot has<br />

very high levels of brightness, size-tunable fluorescence emission,<br />

narrow spectral line widths, large absorption coefficients, and<br />

excellent stability against photobleaching. 29,30,34 These unique<br />

characteristics have spurned intense interests in the use of Qdot<br />

for bioassays and biosensors. On the basis of its singular optical<br />

properties such as the superior signal brightness and high<br />

photostability, Qdot is predicted to be a robust reporter for FLFTS.<br />

Therefore, the employment of Qdot for FLFTS is going to be one<br />

of the most important applications in rapid and sensitive protein<br />

analysis. Currently, the investigations of Qdot-based FLFTS are<br />

remaining at a very early stage, and their applications on protein<br />

quantification have not been reported.<br />

In the present work, we introduced Qdot as fluorescence<br />

probes into a portable dry-reagent strip biosensor system successfully.<br />

Due to the advantages derived from Qdot and LFTS, a<br />

rapid, sensitive, selective, and one-step strategy has been developed<br />

for protein analysis. Nitrated ceruloplasmin, a main biomarker<br />

for cardiovascular disease, lung cancer, and stress response<br />

to smoking, 36-41 was used here as target analyte to test the<br />

performances of Qdot-based FLFTS. Experimental results dem-<br />

(22) Yoon, C. H.; Cho, J. H.; Oh, H. I.; Kim, M. J.; Lee, C. W.; Choi, J. W.; Paek,<br />

S. H. Biosens. Bioelectron. 2003, 19, 289–296.<br />

(23) Lu, F.; Wang, K. H.; Lin, Y. H. Analyst 2005, 130, 1513–1517.<br />

(24) Xia, X. H.; Xu, Y.; Zhao, X. L.; Li, Q. G. Clin. Chem. 2009, 55, 179–182.<br />

(25) Oh, S. W.; Kim, Y. M.; Kim, H. J.; Kim, S. J.; Cho, J. S.; Choi, E. Y. Clin.<br />

Chim. Acta 2009, 406, 18–22.<br />

(26) Choi, S.; Choi, E. Y.; Kim, D. J.; Kim, J. H.; Kim, T. S.; Oh, S. W. Clin.<br />

Chim. Acta 2004, 339, 147–156.<br />

(27) Ahn, J. S.; Choi, S.; Jang, S. H.; Chang, H. J.; Kim, J. H.; Nahm, K. B.; Oh,<br />

S. W.; Choi, E. Y. Clin. Chim. Acta 2003, 332, 51–59.<br />

(28) Khreich, N.; Lamourette, P.; Boutal, H.; Devilliers, K.; Creminon, C.; Volland,<br />

H. Anal. Biochem. 2008, 377, 182–188.<br />

(29) Chan, W. C. W.; Nie, S. M. Science 1998, 281, 2016–2018.<br />

(30) Bruchez, M.; Moronne, M.; Gin, P.; Weiss, S.; Alivisatos, A. P. Science 1998,<br />

281, 2013–2016.<br />

(31) Gao, X. H.; Cui, Y. Y.; Levenson, R. M.; Chung, L. W. K.; Nie, S. M. Nat.<br />

Biotechnol. 2004, 22, 969–976.<br />

(32) Alivisatos, A. P. Science 1996, 271, 933–937.<br />

(33) Michalet, X.; Pinaud, F. F.; Bentolila, L. A.; Tsay, J. M.; Doose, S.; Li, J. J.;<br />

Sundaresan, G.; Wu, A. M.; Gambhir, S. S.; Weiss, S. Science 2005, 307,<br />

538–544.<br />

(34) Smith, A. M.; Nie, S. M. Nat. Biotechnol. 2009, 27, 732–733.<br />

(35) Li, Z. H.; Wang, K. M.; Tan, W. H.; Li, J.; Fu, Z. Y.; Ma, C. B.; Li, H. M.;<br />

He, X. X.; Liu, J. B. Anal. Biochem. 2006, 354, 169–174.<br />

(36) Shiva, S.; Wang, X.; Ringwood, L. A.; Xu, X. Y.; Yuditskaya, S.; Annavajjhala,<br />

V.; Miyajima, H.; Hogg, N.; Harris, Z. L.; Gladwin, M. T. Nat. Chem. Biol.<br />

2006, 2, 486–493.<br />

(37) Pignatelli, B.; Li, C. Q.; Boffetta, P.; Chen, Q. P.; Ahrens, W.; Nyberg, F.;<br />

Mukeria, A.; Bruske-Hohlfeld, I.; Fortes, C.; Constantinescu, V.; Ischiropoulos,<br />

H.; Ohshima, H. Cancer Res. 2001, 61, 778–784.<br />

(38) Radi, R. Proc. Natl. Acad. Sci. U.S.A. 2004, 101, 4003–4008.<br />

onstrate that Qdot-based FLFTS has good ability for quantitative<br />

analysis of trace amount of proteins and is highly expected for<br />

portable and rapid point-of-care screening in clinical diagnostics<br />

and other biomedical applications.<br />

MATERIALS AND METHODS<br />

Reagents and Materials. Human ceruloplasmin was purchased<br />

from Genway Biotech, Inc. (San Diego, CA). Goat<br />

antinitrotyrosine antibody was obtained from Cayman <strong>Chemical</strong><br />

Company (Ann Arbor, MI) while polyclonal human ceruloplasmin<br />

antibody was ordered from Abcam Inc. (Cambridge, MA). A<br />

nitrocellulose membrane, absorbent pad, sample pad, conjugation<br />

pad, and backing cards were purchased from Millipore (Bendford,<br />

MA). Casein (1%) was provided by Bio-Rad (Hercules, CA). A Qdot<br />

585 antibody conjugation kit was obtained from Invitrogen. The<br />

reagent components in the kit include Qdot 585 nanocrystals,<br />

dithiothreitol (DTT) stock solution, succinimidyl trans-4-(Nmaleimidylmethyl)cyclohexane-1-carboxylate<br />

(SMCC) stock solution,<br />

2-mercaptoethanol, dye-labeled marker for antibody elution,<br />

separation media, and exchange buffer. All other chemicals were<br />

ordered from Sigma-Aldrich without further purification. All<br />

buffers and reagent solutions were prepared with purified water,<br />

which was produced from Barnstead Nanopure System (Kirkland,<br />

WA).<br />

Instruments. LM5000 Laminator, XYZ-3050 dispenser consisting<br />

of AirJet Quanti 3000 and BioJet Quanti 3000, and Guillotine<br />

Cutting System CM 4000 were from Biodot Ltd. (Irvine, CA). A<br />

portable fluorescence strip reader ESE-Quant FLUO was purchased<br />

from DCN Inc. (Irvine, CA). A bench-top Eppendorf 5804<br />

centrifuge was obtained from Eppendorf (Hauppauge, NY). An<br />

Ultrospec 2100 Pro UV/Visible spectrophotometer was provided<br />

by Blochrom Ltd. (Cambridge, England).<br />

Preparation of Qdot-Antinitrotyrosine Conjugate.<br />

Qdot-antinitrotyrosine conjugate was prepared according to the<br />

protocol for the Qdot antibody conjugation kit from Invitrogen.<br />

Briefly, 125 µL of Qdot 585 nanocrystals were first activated with<br />

14 µL of 10 mM SMCC at room temperature for 1 h, followed by<br />

being desalted with a NAP-5 desalting column in the presence of<br />

exchange buffer as elution solvent. The colored eluate (∼ 500<br />

µL) was then collected into a centrifuge tube. At room temperature,<br />

300 µL of 1 mg/mL goat antinitrotyrosine antibody was<br />

reduced with 20 mM DTT for 0.5 h. The resulted mixture was<br />

incubated with a dye-labeled marker and purified with a NAP-5<br />

desalting column. The colored fraction (∼500 µL) was collected<br />

and then mixed with the activated Qdot nanocrystals at room<br />

temperature for 1htoform the conjugation complex, followed<br />

by the addition of 10 µL of 10 mM 2-mercaptoethanol to quench<br />

the reaction. The quenched conjugation mixture was then split<br />

into two ultrafiltration devices and concentrated to ∼20 µL for<br />

each half mixture by centrifuging at 7000 rpm on a benchtop<br />

Eppendorf centrifuge. Following the instructions from the conjugation<br />

kit, separation media was gently loaded into the column<br />

and then conditioned with water and PBS, respectively. The<br />

concentrated conjugation mixture was then pipetted into the size-<br />

(39) Lang, J. D.; McArdle, P. J.; O’Reilly, P. J.; Matalon, S. Chest 2002, 122,<br />

314S–320S.<br />

(40) Turko, I. V.; Murad, F. Pharmacol. Rev. 2002, 54, 619–634.<br />

(41) Mitrogianni, Z.; Barbouti, A.; Galaris, D.; Siamopoulos, K. C. Am. J. Kidney<br />

Dis. 2004, 44, 286–292.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7009


Figure 1. (A) Schematic illustration of the test strip and (B1-B4) the detection of nitrated ceruloplasmin using fluorescent Qdot-based FLFTS.<br />

(B1) Aqueous sample containing nitrated ceruloplasmin is applied to sample pad. (B2) Nitrated ceruloplasmin combines with QD-antinitrotyrosine<br />

conjugate and also migrates along the porous membrane by capillary action. (B3) Nitrated ceruloplasmin is captured by anticeruloplasmin<br />

antibodies immobilized on the test line. The excess Qdot conjugates continue to migrate toward the absorption pad. (B4) Fluorescence signal<br />

of Qdot is detected using a test strip reader (solid line). As a control, ceruloplasmin without nitration cannot be recognized by Qdot-antinitrotyrosine<br />

conjugates, so no fluorescence signal can be seen on the test strip (dotted line).<br />

exclusion column followed by the addition of PBS. During the<br />

elution by gravity, the first ten drops only of colored conjugate<br />

(∼200 µL) was collected into a centrifuge tube and stored at 4 °C<br />

until use. Following the instructions from the protocol for the Qdot<br />

antibody conjugation kit from Invitrogen, the conjugate concentration<br />

was determined by measuring the absorbance density of the<br />

conjugate at 585 nm with a Ultrospec 2100 Pro UV/Visible<br />

spectrophotometer and, then, using the formula A ) εcL, where<br />

A is the absorbance, ε is the molar extinction coefficient, c is the<br />

molar concentration of Qdot conjugate, and L is the path length<br />

of the cuvette.<br />

Nitration of Ceruloplasmin. One mg/mL Human ceruloplasmin<br />

in phosphate buffered saline (pH 7.4) was nitrated by<br />

bolus addition of 1 mM authentic peroxynitrite (R&D Systems,<br />

Minneapolis, MN) according to the manufacturer’s recommendations.<br />

The volume of added peroxynitrite was


Figure 2. Qdot-based FLFTS response for (a) 100 ng/mL, (b) 10<br />

ng/mL, and (d) 0 ng/mL nitrated human ceruloplasmin, (c) 10 µg/mL<br />

human ceruloplasmin served as a control.<br />

action. After about 10 min, the cassette containing the test strip<br />

was inserted into an ESE-Quant FLUO reader, followed by the<br />

recording of fluorescence intensity from Qdot on the test line to<br />

quantify the analytes. For the detection of nitrated ceruloplasmin<br />

in human plasma, 20 times diluted plasma spiked with different<br />

quantities of nitrated ceruloplasmin was added onto the sample<br />

pad. The results were obtained by reading the optical response<br />

with the strip reader after 10 min. Meanwhile, these strips after<br />

assay were put under a UV light, and the corresponding fluorescence<br />

images were captured directly by a Sony DSLR-A300 digital<br />

camera.<br />

RESULTS AND DISCUSSION<br />

Assay Principle of Qdot-Based FLFTS. Figure 1 schematically<br />

illustrates the configuration and measuring principle of Qdotbased<br />

FLFTS. This FLFTS is composed of sample application pad,<br />

conjugation pad, absorption pad, and nitrocellulose membrane<br />

(Figure 1A). All the components were assembled onto a plastic<br />

adhesive backing card. During the assay, aqueous sample containing<br />

nitrated ceruloplasmin was applied onto the sample application<br />

pad as shown in Figure 1B1. Subsequently, the analyte migrated<br />

along the porous membrane by capillary action and then bound<br />

with Qdot-antinitrotyrosine on the conjugation zone according<br />

to the specific antibody-antigen interaction (Figure 1B2). The<br />

formed complexes continued to migrate along the membrane and<br />

were captured by anticeruloplasmin antibodies, which resulted<br />

in the accumulation of Qdot on the test line (Figure 1B3). The<br />

excess Qdot conjugates continue to flow into the absorption pad<br />

to the end of the strip. Quantitative analysis was realized by<br />

reading the fluorescence intensities of test line with a portable<br />

strip reader (Figure 1B4). The more analyte in the sample, the<br />

more Qdot conjugates would be captured to the test line, which<br />

leads to the increase of fluorescence intensity. According to the<br />

principle described above, the fluorescence intensity on the test<br />

line would be proportional to the concentration of nitrated<br />

ceruloplasmin in the samples.<br />

Figure 2 shows the typical corresponding responses of Qdotbased<br />

FLFTS to different concentrations of nitrated ceruloplasmin.<br />

Here, human ceruloplasmin without nitration served as a control.<br />

As shown in this figure, well-defined curves were observed in the<br />

Figure 3. Effect of Qdot-antinitrotyrosine conjugates concentration<br />

on the signal-to-noise ratio (S/N) of Qdot-based FLFTS for 1 µg/mL<br />

nitrated human ceruloplasmin.<br />

presence of nitrated ceruloplasmin and the peak area was getting<br />

larger as the target concentration increased from 10 to 100 ng/<br />

mL because more Qdot was captured on the test line based on<br />

the mechanism of Qdot-based FLFTS. In contrast, the presence<br />

of ceruloplasmin without nitration did not contribute to the signal<br />

and exhibited a very low background. The results indicate the<br />

great possibility of Qdot-based FLFTS for sensitive protein<br />

detections.<br />

Determination of Qdot Conjugate Concentration. Following<br />

the instructions for the Qdot antibody conjugation kit from<br />

Invitrogen, the Qdot conjugate eluting from the final column could<br />

be determined according to the Beer-Lambert law 42 by measuring<br />

the absorbance density of the conjugate at 585 nm:<br />

A ) εcL<br />

Where A is the absorbance, ε is the molar extinction coefficient<br />

(as 400 000 M -1 cm -1 provided by Invitrogen for Qdot 585), c<br />

is the molar concentration, and L is the path length of the<br />

cuvette. The result shows that the Qdot conjugate has A ) 0.6<br />

measured in a cuvette with 1 cm path length, so c ) A/ε )<br />

0.6/400 000 ) 1.5 × 10 -6 M, which is the original concentration<br />

of as-prepared Qdot conjugate stock solution.<br />

ParameterOptimization.TheamountofQdot-antinitrotyrosine,<br />

loaded on the glass fiber by physical absorption, directly affects<br />

the fluorescence response of Qdot-based FLFTS since the signal<br />

mainly depends on the amount of Qdot-antinitrotyrosine conjugates<br />

captured on the test line. To probe the optimal amount of<br />

Qdot conjugates for the assay, we diluted them into various<br />

concentrations and investigated the influence on signal-to-noise<br />

(S/N) ratio of the biosensor for 1 µg/mL nitrated human<br />

ceruloplasmin. Ten µg/mL non-nitrated ceruloplasmin served here<br />

as a control. As can be seen from Figure 3, the S/N ratio was<br />

found to be highest for dispensing 150 nM Qdot-antinitrotyrosine.<br />

However, the decrease of S/N at a higher concentration is resulted<br />

from the increase of background signal due to too high of a<br />

concentration of Qdot-antinitrotyrosine conjugate while that at<br />

(42) Ingle, J. D. J.; Crouch, S. R. Spectrochemical Analysis; Prentice Hall: Upper<br />

Saddle Rive, NJ, 1988.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7011


Figure 4. Comparison between blocked (solid line) and unblocked (dash line) strips for the detection of (A) 10 µg/mL nitrated ceruloplasmin<br />

and (B) 10 µg/mL ceruloplasmin without nitration as a control.<br />

lower concentration is ascribed to the decrease of signal due to<br />

too low of an amount of Qdot conjugate availability. Therefore,<br />

150 nM Qdot-antinitrotyrosine conjugate was routinely used as<br />

the optimal concentration throughout the entire study.<br />

Nonspecific binding is one of the likely hindrances in the<br />

development of nanoparticle-based immunoassays. In the current<br />

experiment, there was a fluorescence response obtained from the<br />

control sample (ceruloplasmin without nitration), which resulted<br />

from the nonspecific binding of Qdot conjugates on the test line.<br />

To obtain a maximum response and a minimum nonspecific<br />

absorption, we found that blocking the nitrocellulose membrane<br />

with 1% casein could eliminate the influence of nonspecific binding.<br />

Figure 4 displays the corresponding response of Qdot-based<br />

FLFTS for 10 µg/mL ceruloplasmin with (A) and without (B)<br />

nitration before (dashed line) and after (solid line) blocking. The<br />

results show that the nonspecific binding was effectively reduced<br />

after the blocking while bright fluorescence was well maintained<br />

for the target sample. The significant removal of nonspecific<br />

adsorption maybe attributed to the shield effect of casein, which<br />

was successfully absorbed onto the surface of membrane pad. As<br />

a result, the blocking with 1% casein was employed for the<br />

following experiments.<br />

It is important to evaluate the effect of polycolonal human<br />

ceruloplasmin antibody concentration on the performance of this<br />

biosensor. Consequently, we have investigated different concentrations<br />

of ceruloplasmin antibody (10 ng/mL, 100 ng/mL, 500<br />

ng/mL, 1 mg/mL, 2 mg/mL, and 3 mg/mL) and found that 1<br />

mg/mL has the best response for the detection of nitrated<br />

ceruloplasmin. Furthermore, we have studied the stability of this<br />

biosensor by periodical testing and noticed that this test strip could<br />

be stored at 4 °C (sealed) for at least 3 months and still maintain<br />

good performance.<br />

<strong>Analytical</strong> Performance of Qdot-Based FLFTS for Nitrated<br />

Ceruloplasmin Detection. To investigate the ability of Qdotbased<br />

FLFTS for protein sensitive quantification, the assay was<br />

examined with different concentrations of nitrated ceruloplasmin.<br />

The fluorescence intensity on test line was recorded and plotted<br />

as a function of various concentrations of nitrated ceruloplasmin.<br />

Figure 5 shows the typical response of this biosensor within 10<br />

min for nitrated ceruloplasmin with different concentrations of 1<br />

ng/mL, 5 ng/mL, 10 ng/mL, 40 ng/mL, 100 ng/mL, 1 µg/mL,<br />

and 10 µg/mL, respectively. Ten µg/mL ceruloplasmin without<br />

7012 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 5. Qdot-based FLFTS response for 10 µg/mL, 1 µg/mL, 100<br />

ng/mL, 40 ng/mL, 10 ng/mL, 5 ng/mL, and 1 ng/mL nitrated ceruloplasmin<br />

(curves a-g) and 10 µg/mL ceruloplasmin without nitration<br />

(curve h), respectively.<br />

nitration was used as a control. As can be seen from the figure,<br />

well-defined peaks were observed and the peak area increased<br />

along with the increasing of target concentration while no obvious<br />

fluorescence signal for the control sample could be detected.<br />

Meanwhile, trace amount of nitrated ceruloplamin as low as 1 ng/<br />

mL could be responded by this portable biosensor with a 10 min<br />

assay time.<br />

Due to the superior signal brightness and high photostability<br />

of Qdot, fluorescence imaging of this biosensor after assay could<br />

be employed as a conventional approach to qualify or semiquantify<br />

protein analytes visually and rapidly. By observing the fluorescence<br />

image directly, we can easily judge the existence or not of<br />

target proteins. As shown in Figure 6, the fluorescence band<br />

occurred clearly in the presence of nitrated ceruloplasmin. We<br />

observed proportional changes in fluorescence brightness of the<br />

test lines associated with the concentration of nitrated ceruloplasmin.<br />

Such an observation is expected because the test line<br />

captures more Qdot conjugates when the analyte concentration<br />

is higher. Furthermore, a red signal band from the target<br />

concentration as low as 10 ng/mL could be easily seen even with<br />

visual inspection. In the presence of ceruloplasmin without<br />

nitration, no obvious fluorescence band appeared, indicating a very


Figure 6. Fluorescence imaging of Qdot-based FLFTS for (A) 10 µg/mL, (B) 1 µg/mL, (C) 100 ng/mL, and (D) 10 ng/mL nitrated ceruloplasmin<br />

and (E) 10 µg/mL ceruloplasmin without nitration. The bottom curves are the corresponding readout using a strip reader.<br />

low nonspecific absorption of this biosensor. Accordingly, welldefined<br />

peaks were recorded by the strip reader shown on the<br />

bottom of Figure 6. Consequently, the strip reader and fluorescence<br />

imaging can be combined together to show the assay results<br />

of Qdot-based FLFTS with high sensitivity and specificity. Either<br />

of them could also be used alone depending on the conditions<br />

and requirements. By employing the dual reading approaches,<br />

the proposed sensing platform will be a universal and simple<br />

strategy for complicated protein analysis.<br />

Determination of Nitrated Ceruloplasmin in Human<br />

Plasma. To explore the feasibility of Qdot-based FLFTS for<br />

clinical application, the device was then applied to detect nitrated<br />

ceruloplasmin spiked in 20-fold diluted human plasma with<br />

different concentrations such as 10 µg/mL, 1 µg/mL, 100 ng/<br />

mL, 10 ng/mL, and 1 ng/mL, respectively. Ten µg/mL nonnitrated<br />

ceruloplasmin served as control. These samples were<br />

applied to Qdot-base FLFTS, and the fluorescence signals were<br />

recorded by test strip reader and digital camera after 10 min. A<br />

good calibration curve was obtained in a wide range as showed<br />

in Figure 7. The detection limit was 0.4 ng/mL (S/N ) 3), which<br />

is calculated as the concentration corresponding to 3 times the<br />

SD (standard deviation) of the background signal. Error bars are<br />

based on six duplicated measurements of nitrated ceruloplasmin<br />

at different concentrations, and the control was also run six<br />

replicates. Considering 20-fold dilution of plasma sample during<br />

the assay, the detection limit of 0.4 ng/mL is equivalent to 8 ng/<br />

mL for undiluted plasma, which is comparable to the value<br />

indicated by other techniques for nitrated protein detection such<br />

as nitrated fibrinogen and BSA. 43,44 Regarding the detection limit<br />

(43) Franze, T.; Weller, M. G.; Niessner, R.; Poschl, U. Analyst 2003, 128, 824–<br />

831.<br />

Figure 7. Qdot-based FLFTS linear response for 10 µg/mL, 1 µg/<br />

mL, 100 ng/mL, 10 ng/mL, and 1 ng/mL nitrated ceruloplasmin in<br />

human plasma.<br />

of this biosensor and the high concentration (0.5%, i.e., 2.27 µM)<br />

of ceruloplasmin in human plasma, 45 the sensitivity of this assay<br />

is sufficient to detect as low as 0.03% nitration of ceruloplasmin<br />

in human plasma samples even if only one nitration site is present<br />

in the protein. Meanwhile, fluorescence images were also easily<br />

observed as shown in Figure 8, where the fluorescence band<br />

occurred clearly in the presence of nitrated ceruloplasmin and<br />

almost no fluorescence band came up for non-nitrated ceruloplasmin.<br />

Furthermore, the fluorescence band on the test line for 10<br />

ng/mL nitrated ceruloplasmin in human plasma could be directly<br />

(44) Tang, Z.; Wu, H.; Du, D.; Wang, J.; Wang, H.; Qian, W.; Bigelow, D. J.;<br />

Pounds, J. G.; Smith, R. D.; Lin, Y. Talanta 2010, 81, 1662–1669.<br />

(45) Scheinberg, I. H.; Catlin, D. Science 1952, 116, 484–485.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7013


Figure 8. Fluorescence imaging of Qdot-based FLFTS for (A) 100<br />

ng/mL and (B) 10 ng/mL nitrated ceruloplasmin and (C) 10 µg/mL<br />

ceruloplasmin without nitration in human plasma.<br />

viewed by naked eyes, which is equivalent to 200 ng/mL for an<br />

undiluted plasma sample. Successfully detecting the spiked human<br />

plasma samples using a strip reader or fluorescence imaging<br />

displays the promise of Qdot-based FLFTS for various clinical<br />

applications in the near future.<br />

CONCLUSIONS<br />

In summary, Qdot as a promising alternative reporter was<br />

successfully integrated with lateral flow tests strip and first<br />

developed for rapid, sensitive, and one-step quantitative detection<br />

of a trace amount of nitrated ceruloplasmin. This portable<br />

7014 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

biosensor takes advantage of the speed and low cost of conventional<br />

immunochromatographic strip as well as high sensitivity<br />

and photostability of Qdot-based fluorescent immunoassay. Under<br />

optimal conditions, this proposed Qdot-based FLFTS is capable<br />

of detecting a minimum of 1 ng/mL nitrated ceruloplasmina within<br />

10 min. Furthermore, the linear relationship between peak area<br />

and the logarithm of target concentration was observed in the<br />

range of 1 ng/mL to 10 µg/mL with a detection limit of 0.4 ng/<br />

mL in a spiked plasma sample, which is equivalent to 8 ng/mL<br />

for undiluted plasma. Moreover, the presence of non-nitrated<br />

ceruoloplasmin showed no effect on the biosensor response,<br />

illustrating the good selectivity. Overall, the Qdot-based FlFTS,<br />

considered as an advance in alternative immunosensors, has a<br />

great potential for rapid, sensitive, and portable analysis of other<br />

protein biomarkers in clinical diagnostics, basic discovery, and a<br />

variety of other biomedical applications.<br />

ACKNOWLEDGMENT<br />

The work was done at Pacific Northwest National Laboratory<br />

(PNNL) supported by Grant U54 ES16015 from the National<br />

Institute of Environmental Health Sciences (NIEHS), NIH. Its<br />

contents are solely the responsibility of the authors and do not<br />

necessarily represent the official views of the federal government.<br />

PNNL is operated by Battelle for DOE under Contract DE-AC05-<br />

76RL01830.<br />

Received for review May 28, 2010. Accepted July 1, 2010.<br />

AC101405A


Anal. Chem. 2010, 82, 7015–7020<br />

Selection of Column Dimensions and Gradient<br />

Conditions to Maximize the Peak-Production Rate<br />

in Comprehensive Off-Line Two-Dimensional Liquid<br />

Chromatography Using Monolithic Columns<br />

Sebastiaan Eeltink,* ,† Sebastiaan Dolman, ‡,⊥ Gabriel Vivo-Truyols, § Peter Schoenmakers, §<br />

Remco Swart, ‡ Mario Ursem, ‡ and Gert Desmet †<br />

Department of <strong>Chemical</strong> Engineering, Vrije Universiteit Brussel, Pleinlaan 2, B-1050 Brussels, Belgium, Dionex<br />

Corporation, Abberdaan 114, 1046 AA Amsterdam, The Netherlands, and Van ’t Hoff Institute for Molecular Sciences,<br />

University of Amsterdam, Nieuwe Achtergracht 166, 1018 WV Amsterdam, The Netherlands<br />

The peak-production rate (peak capacity per unit time)<br />

in comprehensive off-line two-dimensional liquid chromatography<br />

(LC/×/LC) was optimized for the separation<br />

of peptides using poly(styrene-co-divinylbenzene) monolithic<br />

columns in the reversed-phase (RP) mode. A firstdimension<br />

( 1 D) separation was performed on a monolithic<br />

column operating at a pH of 8, followed by<br />

sequential analysis of all the 1 D fractions on a monolithic<br />

column operating at a pH of 2. To obtain the<br />

highest peak-production rate, effects of column length,<br />

gradient duration, and sampling time were examined.<br />

RP/×/RP was performed at undersampling conditions<br />

using a short 10 min 1 D gradient. The peak-production<br />

rate was highest using a 50 mm long 2 D column<br />

applying an 8-10 min 2 D gradient time and was almost<br />

a factor of two higher than when a 250 mm monolithic<br />

column was used. The best way to obtain a higher peakproduction<br />

rate in off-line LC/×/LC proved to be an<br />

increase in the number of 1 D fractions collected. Increasing<br />

the 2 D gradient time was less effective. The potential<br />

of the optimized RP/×/RP method is demonstrated by<br />

analyzing proteomics samples of various complexities.<br />

Finally, the trade-off between peak capacity and analysis<br />

time is discussed in quantitative terms for both onedimensional<br />

RP gradient-elution chromatography and the<br />

off-line two-dimensional (RP/×/RP) approach. At the<br />

conditions applied, the RP/×/RP approach provided a<br />

higher peak-production rate than the 1 D-LC approach<br />

when collecting three 1 D fractions, which corresponds<br />

to a total analysis time of 60 min.<br />

Porous polymer monolithic columns have emerged in the<br />

1990s 1,2 and have become a viable alternative for packed-column<br />

* Corresponding author. Tel.: +32 (0)2 629 3324. Fax: +32 (0)2 629 3248.<br />

E-mail: seeltink@vub.ac.be.<br />

† Vrije Universiteit Brussel.<br />

‡ Dionex Corporation.<br />

§ University of Amsterdam.<br />

⊥ Present address: Bruker Biosciences, 1/28A Albert Street, Preston VIC 3072,<br />

Australia.<br />

(1) Hjertén, S; Liao, J.-L; Zhang, R. J. Chromatogr., A 1989, 473, 273–275.<br />

technology for the gradient-elution separation of peptides and<br />

proteins. 3-5 Macroporous polymer monolithic materials were<br />

initially developed by Svec et al. in large I.D. column formats via<br />

a molding process. 6,7 Macroporous poly(styrene-co-divinylbenzene)<br />

(PS-DVB) monolithic rods were operated at high flow rates<br />

(up to 25 mL/min) for the gradient separations of proteins. 7 This<br />

group also demonstrated that the porous properties could be<br />

controlled by optimizing the polymerization mixture, i.e., type and<br />

composition of the porogenic solvent, and the percentage of<br />

difunctional monomer (“cross-linker”) in the mixture. 8,9 Huber<br />

and co-workers developed styrene-based monoliths in situ in 200<br />

µm I.D. capillary columns and applied these columns for liquidchromatographic<br />

mass-spectrometric (LC-MS) analysis of samples<br />

from proteomics and genomics studies, including peptides,<br />

proteins, single-stranded oligonucleotides, and double-stranded<br />

DNA fragments. 10-12 Excellent separation performance was, for<br />

example, demonstrated by a baseline separation of phosphorylated<br />

oligodeoxyadenylic acids, ranging in size from the 12-mer to the<br />

60-mer, and yielded peak widths of only 5.7 s when applying a<br />

7.5 min gradient, 13 resulting in a peak capacity of approximately<br />

80. Recently, we reported the use of a1mlong poly(styrene-codivinylbenzene)<br />

monolithic column yielding a peak capacity in<br />

excess of 1000 for a peptide separation when applying a 600 min<br />

gradient. 14<br />

(2) Peters, E. C.; Petro, M.; Svec, F; Frechet, J. M. J. Anal. Chem. 1997, 69,<br />

3646–3649.<br />

(3) Ivanov, A. R.; Zang, L.; Karger, B. L. Anal. Chem. 2003, 75, 5306–5316.<br />

(4) Geiser, L.; Eeltink, S.; Svec, F.; Frechet, J. M. J. J. Chromatogr., A 2008,<br />

1188, 88–96.<br />

(5) Levkin, P. A.; Eeltink, S.; Stratton, T. R.; Brennen, R.; Robotti, K.; Killeen,<br />

K.; Svec, F.; Frechet, M. J. M. J. Chromatogr., A 2008, 1200, 55–61.<br />

(6) Svec, F.; Frechet, J. M. J. Anal. Chem. 1992, 54, 820–822.<br />

(7) Wang, Q. C.; Svec, F.; Frechet, J. M. J. Anal. Chem. 1993, 65, 2243–2248.<br />

(8) Viklund, C.; Svec, F.; Frechet, J. M. J. Chem. Mater. 1996, 8, 744–750.<br />

(9) Eeltink, S.; Geiser, L; Svec, F; Frechet, J. M. J. J. Sep. Sci. 2007, 30, 407–413.<br />

(10) Toll, H.; Wintringer, R.; Schweiger-Hufnagel, U.; Huber, C. G. J. Sep. Sci.<br />

2005, 28, 1666–1674.<br />

(11) Holzl, G.; Oberacher, H.; Pitsch, S.; Stutz; Huber, C. G. Anal. Chem. 2005,<br />

77, 673–680.<br />

(12) Walcher, W.; Toll, H.; Ingendoh, A.; Huber, C. G. J. Chromatogr., A 2004,<br />

1053, 107–117.<br />

(13) Oberacher, H.; Mayr, B. M.; Huber, C. G. J. Am. Soc. Mass Spec. 2004,<br />

15, 32–42.<br />

(14) Eeltink, S.; Dolman, S.; Detobel, F.; Swart, R.; Ursem, M.; Schoenmakers,<br />

P. J. J. Chromatogr., A 2010, in press.<br />

10.1021/ac101514d © 2010 American <strong>Chemical</strong> Society 7015<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/28/2010


For the analysis of complex sample mixtures, as for example<br />

encountered in proteomics research, the peak capacities that can<br />

be realized with one-dimensional ( 1 D) LC often do not suffice to<br />

achieve complete separation of all compounds. Multidimensional<br />

separation approaches have the potential to separate<br />

thousands of components within a reasonable time. In first<br />

approximation, the maximum peak capacity in two-dimensional<br />

LC is the product of the peak capacities of the individual<br />

dimensions, provided that the two retention mechanisms are<br />

orthogonal. 15 Comprehensive two-dimensional liquid chromatography,<br />

where all essential fractions of the sample are being<br />

analyzed in both dimensions, can be subdivided into two main<br />

categories, i.e., online comprehensive two-dimensional LC (or<br />

LC×LC) and off-line comprehensive two-dimensional LC (or<br />

LC/×/LC). 16<br />

In LC×LC, column dimensions must be selected such that the<br />

eluent composition and transfer volume match the LC conditions<br />

used in the second dimension. Schoenmakers showed that the<br />

maximum flow rate in the first-dimension and, consequently, the<br />

diameter of the 1 D column depends on the maximum injection<br />

volume in the second dimension and on the second-dimension<br />

analysis time. 17 The optimization of sampling time in online<br />

two-dimensional LC has recently been reviewed by Guiochon<br />

et al. 18 When the sampling rate is too low, resolution achieved<br />

in the first dimension is partially lost. However, more time is<br />

available to achieve a high peak capacity in the second<br />

dimension. The fraction of potential peak capacity of the twodimensional<br />

combination of columns that is lost due to the<br />

selection of a too small modulation frequency is described by<br />

the “Nobuo factor” (modulation efficiency). 19 Typically, for<br />

LC×LC, two cuts per first-dimension peak provides the best<br />

trade-off between the loss of resolution in the first dimension<br />

and the analysis time in the second dimension. 19,20 However,<br />

from a practical point of view, this is difficult to achieve since<br />

modulation-phase effects and the random sampling process must<br />

be considered.<br />

The off-line approach (LC/×/LC) offers more flexibility for the<br />

selection of column dimensions and elution conditions, since the<br />

second-dimension analysis time can be optimized independently<br />

from the first-dimension sampling time. In addition, the organic<br />

modifier can be evaporated, so that fractions can be concentrated<br />

or dissolved in a different solvent prior to reinjection. 18,21 To<br />

enhance the preconcentration of peptides or proteins on a trap<br />

column, a strong ion-pairing agent can be added prior to the<br />

second-dimension separation. 21 As a result, the flow rates, transfer<br />

volumes, and the compatibility of (first and second dimension)<br />

eluent composition are less critical issues in LC/×/LC. The<br />

LC/×/LC setup can be optimized for proteomics applications. A<br />

large I.D. first-dimension column can be used, which provides<br />

(15) Giddings, J. C. Anal. Chem. 1984, 65, 1258A–1270A.<br />

(16) Schoenmakers, P. J.; Marriott, P.; Beens, J. LC-GC Eur. 2003, 16, 335–<br />

339.<br />

(17) van der Horst, A.; Schoenmakers, P. J. J. Chromatogr., A 2003, 1000, 693–<br />

709.<br />

(18) Guiochon, G.; Marchetti, M.; Mriziq, K.; Shalliker, R. A. J. Chromatogr., A<br />

2008, 1189, 109–168.<br />

(19) Horie, K.; Kimura, K.; Ikegami, T.; Iwatsuka, A.; Saad, N.; Fiehn, O.; Tanaka,<br />

N. Anal. Chem. 2007, 79, 3764–3770.<br />

(20) Davis, J. M.; Stoll, D. R.; Carr, P. W. Anal. Chem. 2008, 80, 461–473.<br />

(21) Eeltink, S.; Dolman, S.; Swart, R.; Ursem, M.; Schoenmakers, P. J.<br />

J. Chromatogr., A 2009, 1216, 7368–7374.<br />

7016 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

sufficient loadability for the analysis of samples with a broad<br />

dynamic range. A small I.D. column can be applied in the second<br />

dimension, minimizing chromatographic dilution and maximizing<br />

ionization efficiency when coupling LC with mass spectrometry<br />

via an electrospray interface. Huber and co-workers compared the<br />

sequence coverage obtained after an LC/×/LC-MS/MS separation<br />

of a tryptic digest of C. glutamicum using either a strong cationexchange<br />

(SCX) column or a reversed-phase (RP) column operated<br />

at high pH in the first dimension and a reversed-phase<br />

monolith operated at low pH in the second dimension. 22,23 They<br />

observed that the SCX/×/RP and RP/×/RP approaches complemented<br />

each other. Recently, our group compared the 1 D-LC<br />

performance of a 50 mm long monolithic column with that of<br />

an LC/×/LC approach employing a weak-anion exchange and<br />

an RP monolithic column in the first and second dimensions,<br />

respectively. 24 At the conditions applied, the WAX/×/RP<br />

approach provided a better peak-capacity-per-analysis-time ratio<br />

for separations requiring a peak capacity of 400 or higher.<br />

The present contribution discusses how to maximize the peakproduction<br />

rate in off-line comprehensive two-dimensional liquid<br />

chromatography for the reversed-phase separations (RP/×/RP)<br />

of peptides performed at high and low pH using monolithic column<br />

technology. The effects of the first-dimension ( 1 D) column<br />

dimensions (length and diameter) and LC conditions (flow rate<br />

and gradient time) on peak width and sampling volume are<br />

discussed. In addition, the effects of 2 D column length and<br />

gradient time on peak-production rate in 2 D-LC are demonstrated<br />

at “undersampling” conditions (i.e., containing fewer<br />

fractions than required to essentially maintain the first-dimension<br />

separation). Finally, the potential of RP/×/RP is demonstrated<br />

with separations of proteomic samples of varying<br />

complexity.<br />

EXPERIMENTAL SECTION<br />

<strong>Chemical</strong>s and Materials. Acetonitrile (ACN, HPLC supragradient<br />

quality), heptafluorobutyric acid (HFBA, ULC/MS quality),<br />

and trifluoroacetic acid (TFA, ULC/MS quality) were purchasedfromBiosolve(Valkenswaard,TheNetherlands).Ammonium<br />

bicarbonate (min. 99%), dithiothreitol (min. 99%), iodoacetic acid<br />

(approximately 99%), guanidine-HCl, sodium chloride (analytical<br />

reagent grade), cytochrome c (bovine heart), and apo-transferrin<br />

(bovine, g98%) were purchased from Sigma-Aldrich (Steinheim,<br />

Germany). Lysozyme (hen egg white), alcohol dehydrogenase<br />

(yeast), serum albumin (bovine, assay >96%), �-galactosidase, and<br />

sodium phosphate monobasic dihydrate (analytical reagent grade)<br />

were obtained from Fluka (Buchs, Switzerland). Escherichia coli<br />

(E. coli, strain K12) protein sample (lyophilized) was obtained<br />

from Bio-Rad Laboratories (Veenendaal, The Netherlands).<br />

Preconcentration and desalting of peptides prior to the analytical<br />

separation was performed ona5mm× 0.2 mm I.D. monolithic<br />

trap column (Pepswift RP, Dionex Benelux, Amsterdam, The<br />

Netherlands). HPLC separations were performed with 50 mm ×<br />

0.2 mm and 250 mm × 0.2 mm monolithic PepSwift RP columns<br />

(22) Delmotte, N.; Lasaosa, M.; Tholey, A.; Heinzle, E; Huber, C. G. J. Proteome<br />

Res. 2007, 6, 4363–4373.<br />

(23) Toll, H.; Oberacher, H.; Swart, R.; Huber, C. G. J. Chromatogr., A 2005,<br />

1079, 274–286.<br />

(24) Eeltink, S.; Dolman, S.; Detobel, F.; Desmet, G.; Swart, R.; Ursem, M. J.<br />

Sep. Sci. 2009, 32, 2504–2509.


Table 1. Effect of 1 D-LC Conditions on Sampling Time and Number of Fractions Collected When Applying a<br />

Modulation Ration (MR) ∼2 per 1 D Peak a<br />

I.D. (mm) flow rate (µL/min) gradient time (min) 4σ peak width (s) sampling time (s) sampling volume (µL) no. fractions<br />

0.2 2 10 5.5 2.8 0.09 214<br />

1 60 10 5.8 2.6 2.4 231<br />

a<br />

Column length is 50 mm; 1 µL injection of the six-protein digest; aqueous acetonitrile gradient from 1-26% (ACN) at pH ) 8; detection at 214<br />

nm.<br />

and with a 50 mm × 1 mm monolithic ProSwift RP-10R column<br />

(Dionex).<br />

Preparation of Tryptic Digests. The six-protein mixture<br />

prepared from transferrin, bovine serum albumin, �-galactosidase,<br />

alcohol dehydrogenase, lysozyme, cytochrome, and E. coli proteins<br />

were digested according the following procedure. Proteins<br />

were reduced for1hat60°C in the presence of 7 mol/L guanidine<br />

and 1 mol/L dithiothreitol, followed by alkylation for 30 min at<br />

room temperature by adding 1 mol/L iodoacetic acid. To consume<br />

any unreacted iodoacetic acid, 1 mol/L dithiothreitol was added.<br />

The reduced and alkylated proteins were then dialyzed against<br />

50 mM ammonium bicarbonate (pH 8) for 24 h in dialysis sacks<br />

(Sigma) with a cutoff 0.997).<br />

Therefore, the relationship 2 wb ) a + b( 2 tG) can be established,<br />

which holds at a constant length. After fitting, the a and b<br />

parameters were used to predict the peak capacity as a function<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7017


Figure 1. Effect of gradient time on peak width (A) and peak capacity<br />

(B) using a 250 mm × 0.2 mm (closed symbols) and a 50 mm × 0.2<br />

mm (open symbols) monolithic column. Sample: six-protein digest<br />

(transferrin, bovine serum albumin, �-galactosidase, alcohol dehydrogenase,<br />

lysozyme, and cytochrome c), 1 µL injection (0.5 pmol/<br />

µL); flow rate: 2 µL/min; aqueous acetonitrile gradient from 1% to<br />

35% with 0.05% TFA ion-pairing agent; column temperature: 60 °C.<br />

Detection at 214 nm using a3nLflowcell.<br />

of tG (Figure 1B). For gradient times below 60 min, the peak<br />

capacity of the 50 mm long monolith was higher than that of the<br />

250 mm monolith. This can be explained by the difference in<br />

morphology of the two monoliths. 14<br />

At undersampling conditions, the total peak capacity ( 2D nc) and<br />

total analysis time (ttot) in off-line two-dimensional LC are given<br />

by<br />

1<br />

tG<br />

2D<br />

nc )<br />

st t tot ) 1 t +<br />

×<br />

1 tG<br />

s t<br />

2 tG<br />

2 W<br />

(1)<br />

· 2 t (2)<br />

where 2 tG is the second-dimension gradient time, 2 W is the<br />

average second-dimension peak width (equivalent to 4σ), 1 t is<br />

the first-dimension analysis time, 2 t is the second-dimension<br />

analysis time, and st is the sampling time. The first-dimensional<br />

analysis time is the sum of the column holdup time ( 1 t0), the<br />

dwell time ( 1 tdwell), and the gradient time ( 1 tG). The seconddimension<br />

analysis time is the sum of 2 tdesalt, 2 t0, 2 tdwell, a wash<br />

step ( 2 twash), and the second-dimension column equilibration<br />

time ( 2 teq), which corresponds to the time needed to flush the<br />

column with three column volumes to obtain good retention<br />

time stability. The peak production rate (�, min -1 ) is defined<br />

as<br />

7018 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

� )<br />

2D nc<br />

Figure 2 illustrates the effect of the 2 tG on the � using a 50 mm<br />

and 250 mm long monolithic column. The 1 D RP separation<br />

was performed using a 50 mm × 1 mm monolithic column,<br />

applying a 10 min gradient at pH ) 8. The 2 D analysis included<br />

a preconcentration and desalting step on a monolithic trap<br />

column ( 2 tdesalt) and a RP gradient separation at pH ) 2onthe<br />

capillary column. Initially, a steep increase in � can be observed.<br />

In this region, � is dominated by the contributions of 2 t0 and<br />

2 teq time to the total analysis time in the second dimension.<br />

With increasing gradient time, the peak capacity initially<br />

strongly increases (Figure 1B) and the peak-production rate<br />

reaches a maximum. When applying even longer gradients, the<br />

peak-production rate decreases. This is because the peak capacity<br />

marginally increases with the gradient time at longer gradient<br />

duration (see Figure 1B). The existence of a maximum can also<br />

be demonstrated by taking the derivative in eq 3 with respect to<br />

2 tG. Forcing this derivative to be 0 yields<br />

2 tG,max ) � a<br />

t tot<br />

b� s t(<br />

1 td<br />

1 tG<br />

+ 1) + 2 t d<br />

where 2 tG,max refers to the gradient time in the second dimension<br />

that yields the highest value for �. 1 td ) 1 t0 + 1 tdwell, and 2 td )<br />

2 tdesalt + 2 t0 + 2 tdwell + 2 twash + 2 teq. The maximum peakproduction<br />

rate using the 50 mm long monolithic column is<br />

obtained at a 2 tG ∼ 9 min, and the maximum shifts to a higher<br />

Figure 2. Effect of 2 D gradient time on peak-production rate using<br />

250 mm (a) and a 50 mm (b) long second-dimension monolithic<br />

column with a sampling time of 30 s (solid line), 50 mm column with<br />

sampling time of 60 s (- --; c), and 50 mm column with sampling<br />

time of 120 s (---; d). First-dimension separation on a 50 mm × 1<br />

mm long column, 1 D delay time is 4 min, 10 min 1 D gradient time. 1 D<br />

gradient from 1% to 26% acetonitrile at pH ) 8 (10 mM aqueous<br />

ammonium carbonate buffer) at a flow rate of 60 µL/min. Seconddimension<br />

separation on a 250 mm column operating at a flow rate<br />

of 2 µL/min included a 1 min preconcentration, delay time of 0.4 min,<br />

t0 time of 3.53 min, 0.5 min wash step, and 10.6 min equilibration<br />

time; 2 D gradient from 1% to 35% with 0.05% TFA ion-pairing agent.<br />

Second-dimension separation on a 50 mm column operating at a<br />

flow rate of 2 µL/min included a 1 min preconcentration, delay time<br />

of 0.4 min, t0 time of 1.36 min, 0.5 min wash step, and 4.1 min<br />

equilibration time; 2 D gradient from 1% to 35% with 0.05% TFA ionpairing<br />

agent.<br />

(3)<br />

(4)


Figure 3. Peak-production rate versus total analysis time when either<br />

a 50 mm (closed symbols) or a 250 mm (open symbols) long 2 D<br />

monolithic column is used at optimal 2 tG and two or more fractions<br />

are collected. LC conditions described in Figure 2.<br />

value (20 min) when the 250 mm long monolith is used. This<br />

is because the a/b ratio increases with the column length (see<br />

Figure 1A); the slope of 2 tG vs 2 wb is lower for the 250 mm<br />

column, which implies that the b is lower. Also, the a decreases<br />

with column length.<br />

In addition, the � of the 50 mm long monolith is more than a<br />

factor of 2 higher than that of the 250 mm monolith. This is due<br />

to two effects. First, the region of 2 tG considered (from 1 to 60<br />

min) contemplates situations in which the 250 mm column<br />

yields broader peaks than the 50 mm column (see Figure 1A),<br />

reducing 2 nc for the 250 mm column compared to the 50 mm<br />

monolith for a short gradient duration (Figure 1B). Second, with<br />

increasing column length, t0 and teq increase, affecting 2 td and,<br />

hence, the � (see eq 3). Figure 2 also shows that the optimal 2 tG<br />

depends on the sampling rate. This is clearly perceived from<br />

eq 4, where 2 tG,max depends on st. The higher st, the higher the<br />

value of 2 tG,max.<br />

Effect of Sampling Time on Peak-Production Rate. Figure<br />

3 illustrates the � and total analysis time when either a 50 or a<br />

250 mm long 2 D monolithic column is used at optimal 2 tG and<br />

two or more fractions are collected.<br />

When a 50 mm 2 D column is used, first, a strong increase<br />

in � is observed. At higher sampling rates, the increase levels<br />

off and � tends to reach a maximum around 10 peaks/min.<br />

When a 250 mm 2 D column is used, � increases slightly from<br />

4.5 peaks/min (2 fractions) to a maximum of 5 peaks/min after<br />

sampling only three fractions. Apparently, when a 50 mm long<br />

2 D column is used and a slow sampling rate is applied, � is<br />

significantly affected by the contribution of the 1 D analysis time<br />

to the total analysis time. With decreasing sampling time or<br />

when longer 2 D columns (longer t0 and teq) are used, � is<br />

dominated by the 2 D analysis time.<br />

Figure 4 shows the RP(pH)8)/×/RP(pH)2) separation of a sixprotein<br />

digest when applying a sampling time of 60 and 30 s,<br />

respectively. The 1 D separation was performed using a 50 mm<br />

× 1 mm monolithic column applying a gradient time of 10 min.<br />

The 2 D separation was executed on a 50 mm × 0.2 mm long<br />

monolithic column, applying a 2 tG of 7.5 min. The maximum<br />

theoretical peak capacity of the separation shown in Figure 4A<br />

is 1200, and the separation was completed in 98 min. The<br />

Figure 4. RP (pH ) 8)/×/RP (pH ) 2) separation of a digest of six<br />

proteins showing the effect of sampling time (st ) 60 s (A); st ) 30 s<br />

(B)) on peak capacity and total analysis time. 50 mm × 1mm 1 D<br />

column and 50 mm × 0.2 mm 2 D column. Further conditions as<br />

described in Figure 2.<br />

Figure 5. Off-line 2 D-LC separation of an E. coli digest using a<br />

sampling time of 15 s. 50 mm × 1mm 1 D column and 50 mm × 0.2<br />

mm 2 D column. Further conditions as described in Figure 2.<br />

separation in Figure 4B yielded a maximum 2 D-LC peak capacity<br />

of 2400 in 164 min. Whereas the theoretical peak capacity<br />

doubles, the total analysis time increased only by 60%. It should<br />

be noted that the separation space was not completely filled.<br />

This is because the retention mechanisms of the reversedphase<br />

separations performed at pH ) 8 and 2, respectively,<br />

are not completely independent. As a consequence, the fraction<br />

of the total peak capacity that is actually used (“sample peak<br />

capacity”) is lower than the theoretical peak capacity.<br />

For the analysis of more complex samples, such as an E. coli<br />

digest, the most productive way to obtain a higher peak capacity<br />

in LC/×/LC, while still working at undersampling conditions, is<br />

to increase the number of 1 D fractions collected; see Figure 2.<br />

Increasing 2 tG is less effective. Figure 5 shows the LC/×/LC<br />

separation of E. coli digest. The sampling rate (15 s) was selected<br />

such that a maximum theoretical peak capacity of 6700 could be<br />

achieved within a total analysis time of 740 min.<br />

Trade-Off between 1 D-LC and the Optimized LC/×/LC<br />

Approach. Figure 6 illustrates the trade-off between the 1 D-LC<br />

system using a 50 mm and 250 mm long monolithic column<br />

and the RP(pH)8)/×/RP(pH)2) using a 50 mm long monolithic<br />

column in each dimension. The total 1 D-LC analysis time is<br />

the sum of the desalting time (0.5 min), the column holdup<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7019


Figure 6. Trade-off between 1 D-LC performance using 50 mm (solid<br />

line) and 250 mm (dotted line) long monolithic columns and the<br />

optimized off-line 2 D-LC approach collecting g2 fractions (solid<br />

symbols). 1 D-LC conditions described in Figure 1; 2 D-LC conditions<br />

described in Figure 2.<br />

time, the dwell time, and the gradient time. The lines depict<br />

the 1 D-LC performance. The closed circles represent the<br />

RP(pH)8)/×/RP(pH)2) performance collecting discrete numbers<br />

of fractions (two and more). It is evident that below 45 min<br />

the highest � is obtained using 50 mm long monolithic columns<br />

in the 1 D-LC mode. This yields a maximum peak capacity of<br />

320. At a total analysis time of 70 min, 50 mm and 250 mm<br />

long monolithic columns show comparable performance. For<br />

1 D-LC analysis longer than 70 min, the 250 mm monolithic<br />

column provides the �. The maximum 1 D-LC peak capacity of<br />

475 can be achieved within 3 h. However, the � is much smaller<br />

than what can be realized by LC/×/LC. A peak capacity of 480<br />

7020 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

can already be achieved in 60 min using LC/×/LC when<br />

analyzing only three 1 D fractions. The peak capacity linearly<br />

increases with the number of fractions collected and approximately<br />

doubles every 60 min.<br />

CONCLUDING REMARKS<br />

An approach to select column dimensions and LC conditions<br />

in LC/×/LC to maximize � have been demonstrated. Although<br />

the LC/×/LC optimization strategy has been performed with<br />

monolithic columns, this approach can be used when applying<br />

other types of LC columns. At undersampling conditions, the 1 D<br />

separation performance is subsidiary of sampling rate. As a<br />

result, a short first-dimension analysis time is optimal. This<br />

can be achieved using short 1 D column length and applying a<br />

short 1 tG. The optimal 2 tG depends on 2 D column length and<br />

on sampling rate. The optimal 2 tG shift to lower values when<br />

shorter 2 D columns are used yields higher values for �.<br />

For separations requiring a maximum peak capacity up to 340,<br />

1 D-LC using a 50 mm long monolithic column was found to be<br />

superior to LC/×/LC in terms of analysis time. For more<br />

demanding separations, 250 mm long monoliths can be used,<br />

applying longer gradient times (nc, max = 475). However, the �<br />

in LC/×/LC is superior for an analysis taking longer than 60<br />

min (collecting three fractions and more), and the achievable<br />

peak capacity increases linearly with analysis time.<br />

ACKNOWLEDGMENT<br />

Support of this work by a grant of the Research Foundation<br />

Flanders (G.0919.09) is gratefully acknowledged.<br />

Received for review June 8, 2010. Accepted July 17, 2010.<br />

AC101514D


Anal. Chem. 2010, 82, 7021–7026<br />

Cell-Free Expression of Soluble and Membrane<br />

Proteins in an Array Device for Drug Screening<br />

Ruba Khnouf, † Daniel Olivero, ‡ Shouguang Jin,* ,§ Matthew A. Coleman,* ,| and Z. Hugh Fan* ,†,‡<br />

Department of Biomedical Engineering, University of Florida, P.O. Box 116131, Gainesville, Florida 32611, Department of<br />

Mechanical and Aerospace Engineering, University of Florida, P.O. Box 116250, Gainesville, Florida 32611, Department of<br />

Molecular Genetics and Microbiology, University of Florida, P.O. Box 100266, Gainesville, Florida 32610, and Lawrence<br />

Livermore National Laboratory, 7000 East Avenue, Livermore, California 94551<br />

Enzymes and membrane protein receptors represent<br />

almost three-quarters of all current drug targets. As a<br />

result, it would be beneficial to have a platform to produce<br />

them in a high-throughput format for drug screening. We<br />

have developed a miniaturized fluid array device for cellfree<br />

protein synthesis, and the device was exploited to<br />

produce both soluble and membrane proteins. Two membrane-associated<br />

proteins, bacteriorhodopsin and ApoA<br />

lipoprotein, were coexpressed in an expression medium<br />

in the presence of lipids. Simultaneous expression of<br />

ApoA lipoprotein enhanced the solubility of bacteriorhodopsin<br />

and would facilitate functional studies. In addition,<br />

the device was employed to produce two enzymes, luciferase<br />

and �-lactamase, both of which were demonstrated<br />

to be compatible with enzyme inhibition assays.<br />

�-lactamase, a drug target associated with antibiotic<br />

resistance, was further used to show the capability of the<br />

device for drug screening. �-Lactamase was synthesized<br />

in the 96 units of the device and then assayed by a range<br />

of concentrations of four mock drug compounds without<br />

harvesting and purification. The inhibitory effects of these<br />

compounds on �-lactamase were measured in a parallel<br />

format, and the degree in their drug effectiveness agreed<br />

well with the data in the literature. This work demonstrated<br />

the feasibility of the use of the fluid array device<br />

and cell-free protein expression for drug screening, with<br />

advantages in less reagent consumption, shorter analysis<br />

time, and higher throughput.<br />

There is an escalating need to discover new drug targets for<br />

various diseases that impact the quality of life of countless<br />

individuals. Once a target is validated, it is necessary to find drug<br />

candidates by screening a large library of compounds. For both<br />

target discovery and drug screening, novel platforms such as highthroughput<br />

screening (HTS) are required in the modern pharmaceutical<br />

era. 1-3<br />

* To whom correspondence should be addressed. E-mail: hfan@ufl.edu<br />

(Z.H.F.). Fax: 1-352-392-7303 (Z.H.F.).<br />

† Department of Biomedical Engineering, University of Florida.<br />

‡ Department of Mechanical and Aerospace Engineering, University of Florida.<br />

§ Department of Molecular Genetics and Microbiology, University of Florida.<br />

| Lawrence Livermore National Laboratory.<br />

(1) Drews, J. Science 2000, 287, 1960–1964.<br />

(2) Carnero, A. Clin. Transl. Oncol. 2006, 8, 482–490.<br />

Although cell-based HTS has become increasingly popular due<br />

to its ability to deliver phenotype information at the cellular level,<br />

there is still a considerable need to have innovative platforms that<br />

provide biological and chemical information at the molecular<br />

level. 4 Cell-based screening can result in “off target hits” because<br />

different portions of a cellular process other than the target could<br />

be affected by library compounds. 5 In addition, those compounds<br />

that cause cytotoxicity would result in reduced signal even though<br />

they do not inhibit the desired target. 5 Galarneau et al. showed a<br />

great example that both cell-free and cell-based assays were<br />

required to provide complementary and indispensible information<br />

for their study in protein-protein interactions. 4<br />

Cell-free protein synthesis (CFPS) is a platform technology that<br />

addresses the shortcomings mentioned above for cell-based<br />

assays. CFPS utilizes an efficient, often-coupled transcription and<br />

translation reaction to produce recombinant proteins. 6-12 It<br />

eliminates the time-consuming steps of conventional cell-based<br />

protein production, including transformation, cell culture maintenance,<br />

and expression optimization. The proteins synthesized<br />

could be used for functional and structural studies, drug screening,<br />

and other applications. Since there is no cellular control mechanism,<br />

CFPS will not suffer from cytotoxicity or other challenges<br />

associated with cell-based protein expression (e.g., inclusion<br />

body). 6 A large number of soluble proteins have been synthesized<br />

using CFPS, and many of them are enzymes and functionally<br />

active. 13<br />

More importantly, CFPS is an open system and, thus, can be<br />

modified by simply adding various components. This open nature<br />

has been used for producing membrane proteins. Lipids, liposome,<br />

and lipoprotein particles have been introduced into the cell-free<br />

(3) Pereira, D. A.; Williams, J. A. Br. J. Pharmacol. 2007, 152, 53–61.<br />

(4) Galarneau, A.; Primeau, M.; Trudeau, L. E.; Michnick, S. W. Nat. Biotechnol.<br />

2002, 20, 619–622.<br />

(5) von Ahsen, O.; Bomer, U. ChemBioChem 2005, 6, 481–490.<br />

(6) Spirin, A. S.; Baranov, V. I.; Ryabova, L. A.; Ovodov, S. Y.; Alakhov, Y. B.<br />

Science 1988, 242, 1162–1164.<br />

(7) Kigawa, T.; Yabuki, T.; Yoshida, Y.; Tsutsui, M.; Ito, Y.; Shibata, T.;<br />

Yokoyama, S. FEBS Lett. 1999, 442, 15–19.<br />

(8) Madin, K.; Sawasaki, T.; Ogasawara, T.; Endo, Y. Proc. Natl. Acad. Sci.<br />

U.S.A. 2000, 97, 559–564.<br />

(9) Shimizu, Y.; Inoue, A.; Tomari, Y.; Suzuki, T.; Yokogawa, T.; Nishikawa,<br />

K.; Ueda, T. Nat. Biotechnol. 2001, 19, 751–755.<br />

(10) Angenendt, P.; Nyarsik, L.; Szaflarski, W.; Glokler, J.; Nierhaus, K. H.;<br />

Lehrach, H.; Cahill, D. J.; Lueking, A. Anal. Chem. 2004, 76, 1844–1849.<br />

(11) Jewett, M. C.; Swartz, J. R. Biotechnol. Prog. 2004, 20, 102–109.<br />

(12) Katzen, F.; Chang, G.; Kudlicki, W. Trends Biotechnol. 2005, 23, 150–156.<br />

(13) Spirin, A. S. Trends Biotechnol. 2004, 22, 538–545.<br />

10.1021/ac1015479 © 2010 American <strong>Chemical</strong> Society 7021<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/28/2010


expression media to enhance solubility of membrane proteins. 14-19<br />

Several membrane proteins have been synthesized using either<br />

commercially available CFPS systems or ones created by individual<br />

laboratories. 14-19 CFPS simplified the process and reduced<br />

the time and labor required to produce membrane proteins. Many<br />

toxic effects attributed to overproduction of recombinant proteins<br />

in cell-based systems are eliminated by cell-free expression since<br />

viable host cells are no longer required. 14,18 As a result, CFPS<br />

has become an emerging alternative tool for the high level<br />

production of membrane proteins. 14<br />

Because CFPS is able to produce both soluble and membrane<br />

proteins, it could play a significant role in supporting the drug<br />

discovery pipeline. Membrane proteins, mainly G protein-coupled<br />

receptors, represent 45% of all drug targets, while enzymes account<br />

for 28% of them. 1 Thus, these two categories make up a total of<br />

73% of all current drug targets. 1<br />

In this report, we describe the use of a miniaturized fluid array<br />

device for expression of both soluble and membrane proteins. The<br />

device was designed and optimized for high-throughput cell-free<br />

protein synthesis. The fabrication and characterization of the<br />

device have been previously reported. 20-22 We used the device<br />

for producing two membrane-associated proteins, bacteriorhodopsin<br />

and ApoA lipoprotein, in the presence of lipids. Coexpression<br />

of ApoA lipoprotein in the same CFPS addressed the solubility<br />

concern of bacteriorhodopsin.<br />

In addition, the array device was also employed to produce<br />

two enzymes, luciferase and �-lactamase, both of which were<br />

demonstrated to be compatible with enzyme inhibition assays.<br />

Further, we used �-lactamase to show the adaptability of the device<br />

for drug screening. �-Lactamase is an enzyme produced by some<br />

bacteria to generate resistance to a �-lactam class of antibiotics<br />

such as penicillin and cephalosporins. 23 All of these antibiotics<br />

have a four-atom ring structure known as �-lactam. �-lactamase<br />

breaks the amide bond in �-lactam, deactivating its antibacterial<br />

properties and causing antibiotic resistance. One common practice<br />

is to combine antibiotics with a �-lactamase inhibitor that<br />

inactivates �-lactamase. 24 Since �-lactam-associated antibiotics<br />

constitute 50% of antibiotic consumption around the world, 25 it is<br />

important to have an efficient method to identify additional<br />

�-lactamase inhibitors. Therefore, we investigated the exploitation<br />

(14) Klammt, C.; Schwarz, D.; Lohr, F.; Schneider, B.; Dotsch, V.; Bernhard, F.<br />

FEBS J. 2006, 273, 4141–4153.<br />

(15) Kalmbach, R.; Chizhov, I.; Schumacher, M. C.; Friedrich, T.; Bamberg, E.;<br />

Engelhard, M. J. Mol. Biol. 2007, 371, 639–648.<br />

(16) Cappuccio, J. A.; Blanchette, C. D.; Sulchek, T. A.; Arroyo, E. S.; Kralj,<br />

J. M.; Hinz, A. K.; Kuhn, E. A.; Chromy, B. A.; Segelke, B. W.; Rothschild,<br />

K. J.; Fletcher, J. E.; Katzen, F.; Peterson, T. C.; Kudlicki, W. A.; Bench,<br />

G.; Hoeprich, P. D.; Coleman, M. A. Mol. Cell. Proteomics 2008, 7, 2246–<br />

2253.<br />

(17) Schwarz, D.; Dotsch, V.; Bernhard, F. Proteomics 2008, 8, 3933–3946.<br />

(18) Katzen, F.; Fletcher, J. E.; Yang, J. P.; Kang, D.; Peterson, T. C.; Cappuccio,<br />

J. A.; Blanchette, C. D.; Sulchek, T.; Chromy, B. A.; Hoeprich, P. D.;<br />

Coleman, M. A.; Kudlicki, W. J. Proteome Res. 2008, 7, 3535–3542.<br />

(19) Kubick, S.; Gerrits, M.; Merk, H.; Stiege, W.; Erdmann, V. A. Membr. Protein<br />

Cryst. 2009, 63, 25–49.<br />

(20) Mei, Q.; Fredrickson, C. K.; Lian, W.; Jin, S.; Fan, Z. H. Anal. Chem. 2006,<br />

78, 7659–7664.<br />

(21) Mei, Q.; Fredrickson, C. K.; Simon, A.; Khnouf, R.; Fan, Z. H. Biotechnol.<br />

Prog. 2007, 23, 1305–1311.<br />

(22) Khnouf, R.; Olivero, D.; Jin, S.; Fan, Z. H. Biotechnol. Prog. 2010, DOI:<br />

10.1002/btpr.474.<br />

(23) Bush, K. Curr. Opin. Invest. Drugs 2002, 3, 1284–1290.<br />

(24) Poole, K. Cell. Mol. Life Sci. 2004, 61, 2200–2223.<br />

(25) Livermore, D. M. J. Antimicrob. Chemother. 1998, 41 (Suppl D), 25–41.<br />

7022 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 1. (a) Picture of a device in the format of a 96-well plate for<br />

protein expression. The middle layer (membrane) can be observed<br />

in the wells at the bottom left due to light reflection. (b) The crosssectional<br />

view of one well unit, consisting of three access holes and<br />

one reaction chamber in the top layer, a dialysis membrane in the<br />

middle layer, and a feeding chamber in the bottom layer. (c) The<br />

expression product of bacteriorhodopsin and apolipoprotein coexpressed<br />

in the miniaturized fluid array device. (d) The expression<br />

product of bacteriorhodopsin and apolipoprotein coexpressed in a<br />

conventional microplate.<br />

of CFPS to generate �-lactamase that functioned as a drug target,<br />

followed by employing four mock drug compounds to test their<br />

inhibitory effects on �-lactamase. The results demonstrated the<br />

feasibility of the use of the fluid array device for drug screening<br />

with advantages in reagent consumption, analysis time, and<br />

throughput.<br />

MATERIALS AND METHODS<br />

Device Fabrication. The details of the device fabrication have<br />

been described elsewhere. 20-22 Briefly, two polypropylene sheets<br />

were machined to have desired wells with an appropriate size. A<br />

dialysis membrane with molecular cutoff of 8 KD was sandwiched<br />

between these two sheets by an adhesive. Figure 1a shows a<br />

picture of the assembled device in the format of a 96-well plate.<br />

The cross-sectional view of one unit is shown in Figure 1b. The<br />

top layer consists of a reaction chamber and three access holes<br />

for loading the feeding solution. The bottom layer contains a<br />

feeding chamber that encompasses both the access holes and<br />

reaction chamber. The dialysis membrane was used to retain<br />

newly synthesized proteins while allowing the nutrients (e.g.,<br />

amino acids and ATP) in the feeding chamber to continuously<br />

replenish the reaction chamber. The membrane is sturdy and does<br />

not break if touched by a pipet tip by accident. The device was<br />

designed to be disposable; thus, each well was used once.<br />

Protein Expression. Luciferase and �-lactamase were expressed<br />

using an RTS 100 wheat germ kit (Roche). The vector of


luciferase (T7 control vector) was obtained from Promega while<br />

the �-lactamase vector was cloned in house using pIVEX-1.4<br />

plasmid and verified by restriction enzyme digestion and gel<br />

electrophoresis. 22 The reaction and feeding solutions were prepared<br />

according to the manufacturer’s instructions. The reaction<br />

solution was composed of 15 µL of wheat germ lysate, 15 µL of<br />

reaction mix, 4 µL of amino acids, 1 µL of methionine, and 15 µL<br />

of an individual DNA vector (2 µg). For negative controls, the<br />

DNA vector was replaced with the same volume of nuclease free<br />

water. The feeding solution was prepared by combining 900 µL<br />

of feeding mix, 80 µL of amino acids, and 20 µL of methionine.<br />

(All of these were provided in the kit and the concentration of<br />

each component was fixed by the manufacturer.)<br />

Membrane proteins were expressed in the RTS 500 E. coli kit<br />

(Roche). The reaction solution was made by mixing 525 µL ofE.<br />

coli lysate, 225 µL of reaction mix, 270 µL of amino acids without<br />

methionine, and 30 µL of methionine, 2 mg/mL 1,2-dimyristoylsn-glycero-3-phosphocholine<br />

(DMPC, Avanti Polar Lipids), 50 µM<br />

retinal (Sigma), and 5 µg/mL DNA vectors for bacteriorhodopsin<br />

and apolipoprotein. A stock solution of DMPC lipid (68 mg/mL)<br />

was prepared by adding DMPC in nuclease-free water, sonicating<br />

with a Vibra-Cell probe sonicator at a power of 6 W until the<br />

solution became clear, and then centrifuging it to retain the<br />

supernatant. The retinal solution (10 mM) was prepared in pure<br />

ethanol. DNA vectors for bacteriorhodopsin and apolipoprotein<br />

were prepared as reported previously. 16 The feeding solution was<br />

made by mixing 8.1 mL of feeding mix, 2.65 mL of amino acids<br />

without methionine, and 0.3 mL of methionine.<br />

To carry out reactions in the device for either soluble or<br />

membrane proteins, 200 µL of the feeding solution was pipeted<br />

to the feeding chamber and 10 µL of the reaction solution was<br />

added to the reaction chamber. After sealing the device using a<br />

PCR tape (to prevent evaporation), the device was placed on an<br />

orbital shaker for four hours, rotating at a speed of 30 rpm. The<br />

synthesized proteins were then analyzed as discussed below.<br />

Protein Assays. Luciferase was detected by injecting 30 µL<br />

of luciferase assay reagent (Promega) into the reaction chamber<br />

in the device, shaking the device for 2 s, and measuring<br />

luminescence over 10 s. All of these steps were carried out in a<br />

Mithras microplate reader (Berthold Technologies, Germany).<br />

�-Lactamase was measured using m-{[(phenylacetyl)glycyl]oxy}benzoic<br />

acid (PBA, Calbiochem), a chromogenic substrate that<br />

changes color upon being enzymatically cleaved. 26 PBA (90 µL,<br />

2 mM) was added into the expression product, followed by a 5<br />

min incubation and absorbance measurement at 314 nm using a<br />

BioRad spectrophotometer.<br />

Expression of the membrane proteins was indicated by color<br />

observation as discussed in the Results and Discussion. They were<br />

also verified by gel electrophoresis, followed by either Coomassie<br />

blue staining or Western blotting. For Coomassie blue staining,<br />

the sample aliquot was diluted at a ratio of 1:20 using 2× sodium<br />

dodecyl sulfate (SDS) sample loading buffer. A five µL aliquot of<br />

the resulting solution was heated at 99 °C for 10 min (to denature<br />

proteins) and then loaded onto a 15% SDS polyacrylamide gel.<br />

Electrophoresis was carried out with a Precision Plus protein<br />

standard ladder (BioRad) in a Mini-Protean III Cell system<br />

(26) Govardhan, C. P.; Pratt, R. F. Biochemistry 1987, 26, 3385–3395.<br />

(BioRad). After electrophoresis at 150 V for 3 h, the gel was<br />

transferred to a container for Coomassie blue staining.<br />

For the Western blotting, the samples were mixed in a ratio<br />

of 1:100 with the sample loading buffer. After the same electrophoresis<br />

procedure, the gel was transferred onto a nitrocellulose<br />

membrane under an electric current of 220 mA for 30 min. The<br />

membrane was then blocked with 5% fat free milk in PBS/Tween-<br />

20 buffer for 1 h, followed by overnight incubation in anti-Hisperoxidase<br />

antibody (Roche) that was prepared in the blocking<br />

solution in a ratio of 1:50 000. After washing three times, the<br />

membrane was placed in a solution of 1:40 ECL Plus Western<br />

blotting reagents (GE Healthcare) for 5 min. The image of the<br />

membrane was obtained by exposing it onto a 9 × 10 in. Kodak<br />

film.<br />

Enzyme Inhibition Assays. Proteins synthesized in the CFPS<br />

device were used directly for enzyme inhibition assays without<br />

harvesting and purification. Luciferase was chosen to be tested<br />

first using a known inhibitor D-luciferin 6′-methyl ether (LME,<br />

also known as 4,5-dihydro-2[6-methoxy-2-benzthiazolyl]-4-thiazole<br />

carboxylic acid). 27,28 LME solutions with a range of concentrations<br />

were prepared in nuclease-free water. To carry out the inhibition<br />

assay, 1 µL of a LME solution was added to the CFPS reaction<br />

product, followed by the luciferase assay as described above.<br />

For the �-lactamase inhibition assay, three clinically used<br />

drugs, tazobactam (Sigma), potassium clavulanate (Sigma), and<br />

sulbactam (Astatech), as well as an additional compound with<br />

similar chemical structure, cefotaxime (Sigma), were used to study<br />

inhibition. To carry out the inhibition assay, a series of concentrations<br />

of each compound was prepared. Five microliters of each<br />

solution was added to the synthesized �-lactamase in the reaction<br />

chamber, followed by a 15 min incubation. The resulting mixtures<br />

were analyzed using the same protein assay procedure described<br />

above for �-lactamase. The degree of inhibition was calculated<br />

relative to the positive control (no inhibitors) and the negative<br />

control (no DNA vector for protein expression).<br />

RESULTS AND DISCUSSION<br />

Miniaturized Fluid Array Device. A picture of the miniaturized<br />

fluid array device is shown in Figure 1a. The device consists<br />

of 96 units, which are in agreement with a conventional 96-well<br />

microplate. The device was found to be compatible with commercially<br />

available microplate readers and reagent dispensing<br />

apparatuses; thus, it can be used for high-throughput applications<br />

such as drug screening. For each unit (Figure 1b), the top layer<br />

contains a reaction chamber for gene transcription and protein<br />

translation, the bottom layer contains a feeding chamber for<br />

replenishing nutrients (e.g., amino acids and ATP), and the middle<br />

is a dialysis membrane that connects these two chambers. As<br />

discussed in the literature, 6,12,29 the functions of the membrane<br />

are to (1) achieve continuous supply of additional nutrients; (2)<br />

retain proteins produced and large-molecule synthesis machinery;<br />

and (3) dilute the reaction byproducts (e.g., pyrophosphates) and<br />

reduce their effects on the reaction equilibrium. Compared to the<br />

commercially available CFPS systems (e.g., RTS 500 kit), 29 one<br />

major advantage of the fluid array device is lower reagent<br />

(27) Wang, J. Q.; Pollok, K. E.; Cai, S.; Stantz, K. M.; Hutchins, G. D.; Zheng,<br />

Q. H. Bioorg. Med. Chem. Lett. 2006, 16, 331–337.<br />

(28) Barros, M. P.; Bechara, E. J. Free Radical Biol. Med. 1998, 24, 767–777.<br />

(29) Betton, J. M. Curr. Protein Pept. Sci. 2003, 4, 73–80.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7023


Figure 2. (a) Gel electrophoresis of membrane proteins. Lane 1 is<br />

a protein ladder. Lanes 2-4 are for the negative control. Lanes 5-7<br />

are for expression of bacteriorhodopsin. Lanes 8-10 are for expression<br />

of apolipoprotein. Lanes 11-13 are for coexpression of bacteriorhodopsin<br />

and apolipoprotein. The first lane in each group is for<br />

the reaction mixture; the second lane is for the pellet, and the third<br />

lane is for the supernatant. The arrow in lane 5 indicates the position<br />

of bacteriorhodopsin while the arrow in lane 8 indicates the position<br />

of apolipoprotein. (b) Western blotting of membrane proteins with the<br />

same lane designation as in (a). (c) Pictures of tubes containing<br />

expressed protein products (after centrifugation). They are in the same<br />

order as in (a) from the left to the right.<br />

consumption (10 µL versus 1 mL). The cost-savings would be<br />

substantial when a large number of arrays are required for highthroughput<br />

applications. The additional advantage of miniaturization<br />

is to enable a high-density array, which makes it possible to<br />

have high-throughput CFPS for applications such as drug screening.<br />

An alternative way to run high-throughput CFPS is in a<br />

conventional 96-well or 384-well microplate. Without a membrane<br />

in the conventional microplate, we found the synthesis yield was<br />

much lower. 20-22 For the membrane protein to be discussed<br />

below, coexpression of bacteriorhodopsin and apoA lipoprotein<br />

in the fluid array resulted in a mixture with the characteristic<br />

purple color of bacteriorhodopsin as shown in Figure 1c, demonstrating<br />

the functional production of bacteriorhodopsin. Note<br />

that the protein expression product in the device was transferred<br />

to a transparent tube for the purpose of taking a picture. In<br />

contrast, when the same expression was carried out in a<br />

conventional microplate, the amount of proteins expressed was<br />

too low to have an effect on the color of the mixture as shown in<br />

Figure 1d. The yellow color of the mixture is very similar to the<br />

negative control that contains no DNA vectors.<br />

Membrane Proteins. The fluid array device was used to<br />

produce two membrane-associated proteins, apoA lipoprotein and<br />

bacteriorhodopsin. ApoA lipoprotein can form nanoparticles due<br />

to the presence of lipid in the expression medium. 16 Its expression<br />

was confirmed by polyacrylamide gel electrophoresis (PAGE) as<br />

shown in Figure 2a (arrow in lane 8). Since the lipoprotein particle<br />

is soluble, it was in the supernatant after centrifugation (lane 10).<br />

Expression of bacteriorhodopsin in the fluid array was also<br />

confirmed by PAGE (arrow in lane 5). After centrifugation of the<br />

7024 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

mixture, most bacteriorhodopsin molecules precipitated to form<br />

pellets since they were insoluble (lane 6). 15,16,30<br />

Bacteriorhodopsin and ApoA lipoprotein could be coexpressed<br />

in a single reaction; the lipoprotein’s ability to form lipid bilayer<br />

patches from lipid micelles can be used to increase the solubility<br />

of other membrane-associated proteins such as bacteriorhodopsin.<br />

15,16,30 The parallel expression of both proteins was<br />

confirmed by PAGE (lane 11). After centrifugation of the mixture,<br />

most bacteriorhodopsin molecules stayed in the supernatant due<br />

to the increased solubility in the presence of lipoprotein particles<br />

(lane 13).<br />

Since there are many protein bands belonging to the components<br />

of the protein expression system, we carried out Western<br />

blotting to verify the presence of bacteriorhodopsin and nanolipoprotein<br />

in the samples. The blot in Figure 2b shows much<br />

cleaner protein bands for the proteins of interest. Note that<br />

Western blotting in Figure 2b is more sensitive than Coomassie<br />

blue staining in Figure 2a; hence, a minute amount of bacteriorhodopsin<br />

undetectable in lane 7 in Figure 2a was shown as a<br />

band in Figure 2b.<br />

The effect of coexpressed lipoprotein on the solubility of<br />

bacteriorhodopsin is evident from the physical appearance of the<br />

protein product solutions at the end of expression, followed by<br />

centrifugation, as shown in Figure 2c. These proteins were<br />

expressed in the device, and the product mixtures were transferred<br />

to a tube, followed by centrifugation to separate soluble<br />

and insoluble materials for visualization. The negative control with<br />

no DNA vectors shows the color of the cell-free reaction mixture.<br />

The product mixture for expression of apoA lipoprotein is<br />

homogeneous since apoA lipoprotein is soluble. However, the<br />

product mixture for expression of bacteriorhodopsin contains a<br />

purple pellet and clear supernatant after centrifugation. This is in<br />

agreement of the result of gel electrophoresis in Figure 2a that<br />

bacteriorhodopsin is in the pellet. When two proteins were<br />

simultaneously coexpressed, the solubility of bacteriorhodopsin<br />

was increased in the presence of apoA lipoproteins. Importantly,<br />

the bacteriorhodopsin sample was purple, which indicates the<br />

formation of a functional protein in our device.<br />

Soluble Proteins. As we have reported previously, several<br />

soluble proteins can be expressed in the miniaturized fluid array<br />

device, including luciferase, green fluorescent protein, �-glucoronidase,<br />

alkaline phosphatase, �-lactamase, and �-galactosidase. 20-22<br />

In this work, we chose two soluble proteins, luciferase and<br />

�-lactamase, to demonstrate the enzyme inhibition assay. Luciferase<br />

gene has been extensively used as a reporter gene for<br />

visualizing biomolecular processes in living animals, 27,31 and<br />

luciferase-based luminescence assays have been increasingly used<br />

for high-throughput drug screening. 32 As a result, we selected<br />

luciferase to verify if the protein expressed in the CFPS device<br />

can be used for enzyme inhibition assays without harvesting and<br />

purification from the protein expression mixture that contains a<br />

number of cofactors, enzymes, and others. We first studied the<br />

kinetics of luciferase assay reactions in the presence of an<br />

inhibitor, luciferin 6′-methyl ether (LME). LME is a known<br />

luciferase assay inhibitor and has been used by others in the<br />

(30) Bayburt, T. H.; Grinkova, Y. V.; Sligar, S. G. Arch. Biochem. Biophys. 2006,<br />

450, 215–222.<br />

(31) Prescher, J. A.; Contag, C. H. Curr. Opin. Chem. Biol. 2010, 14, 80–89.<br />

(32) Fan, F.; Wood, K. V. Assay Drug Dev. Technol. 2007, 5, 127–136.


Figure 3. Inhibitory effects of luciferin 6′-methyl ether on the<br />

luciferase assay. (a) The luciferase assay reaction kinetics in the<br />

presence of various concentrations of inhibitors. (b) The inhibitory<br />

effects as a function of the inhibitor concentration. The decrease in<br />

percentage of the luminescence signal in the y-axis is relative to the<br />

positive control, in which no inhibitor was added. The lines are the<br />

best fit in the power relationship among the experimental data points.<br />

The error bars are invisible, as they are smaller than the data markers.<br />

literature. 27,28 The inhibitory assay was performed using luciferase<br />

synthesized in the miniaturized fluid array device; the expression<br />

product was directly used for the assay without any treatment. Figure<br />

3a shows the reaction kinetic curves we obtained. The resemblance<br />

of these curves with those in the literature using purified luciferase 33<br />

indicates that the enzyme synthesized in the cell-free medium can<br />

be used for the enzyme inhibition assay without harvesting and<br />

purification. As a result, we can use CFPS products directly for the<br />

enzyme inhibition assay. Figure 3b shows the inhibitory effects on<br />

the luciferase assay as a function of the LME concentration.<br />

In addition, we evaluated the specificity of each enzyme assay.<br />

For example, the UV absorbance-based enzyme assay of �-lactamase<br />

using m-{[(phenylacetyl)glycyl]oxy}benzoic acid (PBA)<br />

should be specific to �-lactamase. Figure 4 shows the assay<br />

kinetics, in which signal increased over a period of time when<br />

�-lactamase was cleaving the chromogenic substrate. It leveled<br />

off to form a plateau when the reactions reached equilibrium.<br />

However, when the same assay was applied to luciferase, there<br />

was negligible signal increase, except for the initial point that was<br />

due to the addition of a solution in a way similar to a negative<br />

control. These results suggest that the �-lactamase assay is<br />

specific. The data also indicate that reliable assay results for<br />

�-lactamase should be recorded at 5 min after the assay reagents<br />

were added into the expression products.<br />

Drug Screening. To demonstrate the utility of the miniaturized<br />

fluid array for drug screening, �-lactamase was chosen as a<br />

mock drug target. �-Lactamase is an enzyme that hydrolyzes the<br />

amide bond of �-lactam antibiotics, causing the antimicrobial<br />

agents to lose their effectiveness. Three known �-lactamase<br />

(33) DeLuca, M.; Wannlund, J.; McElroy, W. D. Anal. Biochem. 1979, 95, 194–<br />

198.<br />

Figure 4. Enzyme assay kinetics of �-lactamase using a chromogenic<br />

agent, m-{[(phenylacetyl)glycyl]oxy}benzoic acid. The same assay<br />

was applied to luciferase, showing negligible increase over the<br />

background signal.<br />

Figure 5. Inhibition of clavulanate acid (open cirles), tazobactam<br />

(diamonds), and sulbactam (triangles) on the enzymatic activities of<br />

�-lactamase. The concentrations of these inhibitors in the x-axis are<br />

in the log scale. Also shown is negligible and concentrationindependent<br />

inhibition of cefotaxime (solid circles) on �-lactamase.<br />

Each data point represents an average obtained from three repeat<br />

experiments, and the error bars indicate one standard deviation. The<br />

chemical structures of all compounds tested are shown at the top;<br />

all of them have a �-lactam ring that consists of three carbon atoms<br />

and one nitrogen atom with a cyclic amide functional group.<br />

inhibitors (clavulanate acid, tazobactam, and sulbactam) and one<br />

noninhibitory chemical (cefotaxime) were used as mock compounds.<br />

All of these compounds contain a �-lactam ring that can<br />

be hydrolyzed by �-lactamase, as shown in Figure 5. However,<br />

cefotaxime is a very large molecule, difficult to be hydrolyzed by<br />

�-lactamase. �-Lactamase was produced in each unit of the array<br />

device, and a series of varying concentrations of each compound<br />

was added into several wells to obtain a response curve. The<br />

relationship between the concentrations of each compound and<br />

the level of �-lactamase inhibition is also shown in Figure 5. The<br />

percentage of inhibition is calculated against the one without any<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7025


inhibitor (positive control). The results show that clavulanate acid,<br />

tazobactam, and sulbactam had inhibitory effects on the enzymatic<br />

activities of �-lactamase, while cetotaxime did not show any<br />

inhibitory effect. The I50 values (the concentration required to<br />

produce 50% inhibition 34 ) of clavulanate acid, tazobactam, and<br />

sulbactam are 0.11, 0.87, and 24.8 µg/µL, respectively. As a<br />

result, clavulanate acid had the highest inhibitory effect among<br />

them, which agrees well with the previously published<br />

results. 35,36<br />

The result in Figure 5 suggests that protein expressed in a<br />

cell-free system can be used for drug screening just as those<br />

obtained from cell-based systems. The power of the fluid array<br />

device is evident from the fact that the inhibition-concentration<br />

curves of multiple compounds can be simultaneously obtained in<br />

one device through judicial experimental designs. Importantly,<br />

the device allowed us to determine not only if the compound is a<br />

drug candidate but also the effectiveness of the drug by comparing<br />

the various degrees of inhibition. An additional benefit of the use<br />

of the fluid array is that �-lactamase is in solution phase, so that<br />

its enzymatic activities are retained due to its appropriate threedimensional<br />

configuration in an aqueous solution. As a result, it<br />

is superior over ELISA or protein chips in which proteins are<br />

attached to a solid surface and efforts must be made to minimize<br />

the loss of biological activities of the proteins.<br />

CONCLUSION<br />

A miniaturized fluid array device has been demonstrated<br />

capable of carrying out high-throughput cell-free protein synthesis.<br />

Compared to the conventional cell-free protein synthesis using<br />

commercially available RTS 500 kits, 29 the reagent consumption<br />

is 2 orders of magnitude less (10 µL of the reaction solution in<br />

our device versus 1 mL in the commercial kits). As a result, the<br />

cost-savings would be substantial when a large array is required<br />

for high-throughput applications (e.g., drug screening).<br />

Cell-free expression of both soluble and membrane proteins<br />

has been achieved in the array device. A membrane protein<br />

(34) Duggleby, R. G. Biochem. Med. Metab. Biol. 1988, 40, 204–212.<br />

(35) Payne, D. J.; Cramp, R.; Winstanley, D. J.; Knowles, D. J. Antimicrob. Agents<br />

Chemother. 1994, 38, 767–772.<br />

(36) Bou, G.; Martinez-Beltran, J. Antimicrob. Agents Chemother. 2000, 44, 428–<br />

432.<br />

7026 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

synthesized in the device can be solubilized by coexpressing it<br />

with a lipoprotein in the presence of reagents necessary for correct<br />

folding, which would allow functional and structural studies of<br />

the membrane protein. Since membrane protein receptors represent<br />

the largest portion of drug targets, this capability could<br />

have great impact in searching for therapeutics.<br />

Proteins synthesized in the device have been exploited for<br />

enzyme inhibition assays without protein harvesting and purification.<br />

Therefore, laborious and time-consuming procedures such<br />

as magnetic bead-based separation and chromatography-based<br />

purification can be eliminated. Potential losses of proteins and<br />

possible structure changes in the separation procedures have been<br />

reduced. In addition, a solution array (proteins in solution) does<br />

not compromise the kinetics of binding 37 whereas proteins<br />

immobilized on a solid surface in conventional protein chips are<br />

likely compromised by the fact that (1) only a portion of protein<br />

structures are exposed to ligands, (2) spatial hindrance exists<br />

during interactions between immobilized proteins and ligands in<br />

a sample, and (3) protein tertiary structures may change during<br />

immobilization.<br />

The array format of the device enabled simultaneous expression<br />

of proteins, screening for a variety of inhibitors, and<br />

comparison of the inhibition levels among the inhibitors studied.<br />

We showed the capability of the device for drug screening by the<br />

use of �-lactamase, a drug target associated with antibiotic<br />

resistance, and four mock drug compounds. The inhibitory effects<br />

of these compounds on �-lactamase were measured in parallel,<br />

and their degrees in drug effectiveness were compared.<br />

ACKNOWLEDGMENT<br />

This work was supported in part by Defense Advanced<br />

Research Projects Agency (DARPA) via Micro/Nano Fluidics<br />

Fundamentals Focus Center at the University of California at<br />

Irvine, the University of Florida via UF Opportunity Fund, and<br />

the University of California Discovery Grant Program.<br />

Received for review June 10, 2010. Accepted July 18,<br />

2010.<br />

AC1015479<br />

(37) Zhou, H.; Roy, S.; Schulman, H.; Natan, M. J. Trends Biotechnol. 2001,<br />

19, S34–39.


Anal. Chem. 2010, 82, 7027–7034<br />

Patterning of Metal, Carbon, and Semiconductor<br />

Substrates with Thin Organic Films by<br />

Microcontact Printing with Aryldiazonium Salt Inks<br />

Joshua Lehr, †,‡ David J. Garrett, †,‡ Matthew G. Paulik, ‡,§ Benjamin S. Flavel, ‡,<br />

Paula A. Brooksby, ‡ Bryce E. Williamson, ‡ and Alison J. Downard* ,†,‡<br />

MacDiarmid Institute for Advanced Materials and Nanotechnology, Private Bag 4800, Christchurch, 8140, New Zealand,<br />

and Department of <strong>Chemistry</strong>, University of Canterbury, Private Bag 4800, Christchurch, 8140, New Zealand<br />

Surface modification through reduction of aryldiazonium<br />

salts to give covalently attached layers is a widely investigated<br />

procedure. However, realization of potential applications<br />

of the layers requires development of patterning<br />

methods. Here, we demonstrate that microcontact printing<br />

with poly(dimethylsiloxane) stamps inked with aqueous<br />

acid solutions of aryldiazonium salts gives stable<br />

organic layers on gold, copper, silicon, and graphitic<br />

carbon surfaces. Depending on the substrate-diazonium<br />

salt combination, the layers range from relatively irregular<br />

multilayers to smooth films with close to monolayer<br />

thickness. After printing, surface attached aminophenyl<br />

and carboxyphenyl groups retain their usual reactivity<br />

toward amide bond formation with solution species, and<br />

hence, the method is a simple route to patterned, covalently<br />

attached, reactive tether layers. Multicomponent<br />

patterned films can be prepared by printing a second<br />

modifier onto a film-coated surface. Microcontact printing<br />

using aryldiazonium salt inks is experimentally very<br />

simple and is applicable to the broad range of substrates<br />

capable of spontaneously reducing aryldiazonium salts.<br />

Localized immobilization of molecular species is an important<br />

step in the fabrication of surfaces for applications that include<br />

chemical and biological sensors, biochips, molecular electronics,<br />

and tissue engineering. A large research effort over many years<br />

has established a suite of patterning methods, compatible with<br />

the best-known surface modification strategies, particularly those<br />

based on assembly of alkanethiols at noble metal surfaces, and<br />

reactions of silanes at oxide surfaces and alkenes at silicon. A<br />

relatively recently developed surface modification method is<br />

grafting from aryldiazonium salts solutions. 1,2 The mechanism,<br />

scope, and applications of this method have been widely<br />

* To whom correspondence should be addressed. E-mail: alison.downard@<br />

canterbury.ac.nz. Fax: +64-3-3642110.<br />

† MacDiarmid Institute for Advanced Materials and Nanotechnology.<br />

‡ University of Canterbury.<br />

§ Present address: AgResearch Ltd, Private Bag 4749, Christchurch, 8140,<br />

New Zealand.<br />

Present address: School of <strong>Chemistry</strong>, Physics and Earth Sciences, Flinders<br />

University, Bedford Park, SA 5042, Australia.<br />

(1) Delamar, M.; Hitmi, R.; Pinson, J.; Saveant, J. M. J. Am. Chem. Soc. 1992,<br />

114, 5883–5884.<br />

(2) Allongue, P.; Delamar, M.; Desbat, B.; Fagebaume, O.; Hitmi, R.; Pinson,<br />

J.; Saveant, J.-M. J. Am. Chem. Soc. 1997, 119, 201–207.<br />

investigated; 3-9 however, patterning techniques appropriate for<br />

use with this approach have received little attention.<br />

Surface grafting from aryldiazonium salt solutions proceeds<br />

via reduction of the aryldiazonium cation and elimination of<br />

dinitrogen to yield an aryl radical capable of reaction with the<br />

surface. 2 For some substrates, a covalent linkage between the<br />

modifier and the surface has been directly demonstrated; 10-12 for<br />

others, covalent attachment has been inferred on the basis of the<br />

orientation of surface groups 13,14 or the stability of the layer. 2,15,16<br />

Under commonly used grafting conditions, attack by aryl radicals<br />

on pregrafted groups gives a multilayered, covalently coupled film<br />

structure. 17,18 The stability of the grafted layers is a key advantage<br />

of this modification method; other attractive features are its wide<br />

substrate compatibility (from noble and industrial metals to<br />

semiconductors 9-11,19-22 ) and the opportunity for further chemistry<br />

involving the grafted groups. Electrografting at an externally<br />

(3) Adenier, A.; Bernard, M.-C.; Chehimi, M. M.; Cabet-Deliry, E.; Desbat, B.;<br />

Fagebaume, O.; Pinson, J.; Podvorica, F. J. Am. Chem. Soc. 2001, 123,<br />

4541–4549.<br />

(4) Barriere, F.; Downard, A. J. J. Solid State Electrochem. 2008, 12, 1231–<br />

1244.<br />

(5) Downard, A. J. Electroanalysis 2000, 12, 1085–1096.<br />

(6) Fave, C.; Leroux, Y.; Trippe, G.; Randriamahazaka, H.; Noel, V.; Lacroix,<br />

J. C. J. Am. Chem. Soc. 2007, 129, 1890–1891.<br />

(7) Pinson, J.; Podvorica, F. Chem. Soc. Rev. 2005, 34, 429–439.<br />

(8) Ranganathan, S.; Steidel, I.; Anariba, F.; McCreery, R. L. Nano Lett. 2001,<br />

1, 491–494.<br />

(9) Stewart, M. P.; Maya, F.; Kosynkin, D. V.; Dirk, S. M.; Stapleton, J. J.;<br />

McGuiness, C. L.; Allara, D. L.; Tour, J. M. J. Am. Chem. Soc. 2004, 126,<br />

370–378.<br />

(10) Bernard, M. C.; Chausse, A.; Cabet-Deliry, E.; Chehimi, M. M.; Pinson, J.;<br />

Podvorica, F.; Vautrin-Ul, C. Chem. Mater. 2003, 15, 3450–3462.<br />

(11) Boukerma, K.; Chehimi, M. M.; Pinson, J.; Blomfield, C. Langmuir 2003,<br />

19, 6333–6335.<br />

(12) Combellas, C.; Kanoufi, F.; Pinson, J.; Podvorica, F. I. Langmuir 2005,<br />

21, 280–286.<br />

(13) Anariba, F.; Viswanathan, U.; Bocian, D. F.; McCreery, R. L. Anal. Chem.<br />

2006, 78, 3104–3112.<br />

(14) Ricci, A.; Bonazzola, C.; Calvo, E. J. Phys. Chem. Chem. Phys. 2006, 8, 4297–<br />

4299.<br />

(15) D’Amours, M.; Belanger, D. J. Phys. Chem. B 2003, 107, 4811–4817.<br />

(16) Shewchuk, D. M.; McDermott, M. T. Langmuir 2009, 25, 4556–4563.<br />

(17) Doppelt, P.; Hallais, G.; Pinson, J.; Podvorica, F.; Verneyre, S. Chem. Mater.<br />

2007, 19, 4570–4575.<br />

(18) Kariuki, J. K.; McDermott, M. T. Langmuir 2001, 17, 5947–5951.<br />

(19) Delamar, M.; Desarmot, G.; Fagebaume, O.; Hitmi, R.; Pinson, J.; Saveant,<br />

J. M. Carbon 1997, 35, 801–807.<br />

(20) Liang, H. H.; Tian, H.; McCreery, R. L. Appl. Spectrosc. 2007, 61, 613–<br />

620.<br />

(21) Mahmoud, A. M.; Bergren, A. J.; McCreery, R. L. Anal. Chem. 2009, 81,<br />

6972–6980.<br />

10.1021/ac101785c © 2010 American <strong>Chemical</strong> Society 7027<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/28/2010


applied potential is the most widely used strategy for the reduction<br />

step, but simple immersion of the substrate in a solution of an<br />

aryldiazonium salt can also lead to formation of surface layers.<br />

This spontaneous, or open-circuit potential (OCP), reaction has<br />

been reported for a range of metals, semiconductors, and carbon<br />

materials 4 and appears to involve electron transfer from the<br />

substrate to the aryldiazonium cation in solution.<br />

Currently, there are few examples of patterned organic layers<br />

prepared by reduction of aryldiazonium salts. In the earliest report,<br />

we used mechanical scribing with an atomic force microscope<br />

(AFM) tip to remove regions of electrografted film from a carbon<br />

substrate. 23 A second aryldiazonium salt was then electrografted<br />

to the bare regions, creating a surface with dual chemical<br />

functionality. In a soft lithographic approach, we patterned a<br />

carbon substrate by adhering a poly(dimethylsiloxane) (PDMS)<br />

mold to the surface (either bare or film-coated) to form microchannels.<br />

24 The channels were subsequently filled, either with<br />

aryldiazonium salt solution for site-specific electrografting or with<br />

reagents used for electrochemical or chemical conversion of the<br />

pre-existing surface film. Most recently, we demonstrated that<br />

conventional photolithography can be coupled with electrografting<br />

to give large areas of micrometer-sized patterns of modifiers on<br />

highly doped silicon. 22 Charlier, Palacin, and co-workers, 25 and<br />

Cougnon, Bélanger, and co-workers 26 have established that the<br />

scanning electrochemical microscope is a useful tool for localized<br />

surface grafting from aryldiazonium salts, while the former<br />

research group has also reported elegant patterning methods<br />

specific to silicon substrates. In one example, they used ionic<br />

implantation to create locally doped areas of silicon and, thus,<br />

achieved site-specific electrografting of an aryldiazonium salt. 27<br />

In another example, they illuminated p-type silicon through a mask<br />

to locally increase the substrate conductivity and allow electrografting<br />

to proceed. 28 Palacin and co-workers have also explored<br />

the use of a patterned agarose hydrogel containing an aryldiazonium<br />

salt solution sandwiched between two electrodes as an<br />

electrochemical “printing” method. 29 Finally, Corgier and Bélanger<br />

have adapted the methods of colloidal nanolithography to electrograft<br />

organic groups to the nanoscale spaces between polystyrene<br />

beads assembled on carbon and gold surfaces. 30<br />

All of the patterning methods described above involve electrochemical<br />

generation of aryl radicals at an externally applied<br />

potential. In an earlier communication, we established that<br />

spontaneous, OCP reduction of aryldiazonium salts by carbon<br />

substrates can also be used. 31 We showed that microcontact<br />

(22) Flavel, B. S.; Garrett, D. J.; Lehr, J.; Shapter, J. G.; Downard, A. J.<br />

Electrochim. Acta 2010, 55, 3995–4001.<br />

(23) Brooksby, P. A.; Downard, A. J. Langmuir 2005, 21, 1672–1675.<br />

(24) Downard, A. J.; Garrett, D. J.; Tan, E. S. Q. Langmuir 2006, 22, 10739–<br />

10746.<br />

(25) Ghorbal, A.; Grisotto, F.; Charlier, J.; Palacin, S.; Goyer, C.; Demaille, C.<br />

ChemPhysChem 2009, 10, 1053–1057.<br />

(26) Cougnon, C.; Gohier, F.; Belanger, D.; Mauzeroll, J. Angew. Chem., Int.<br />

Ed. 2009, 48, 4006–4008.<br />

(27) Charlier, J.; Palacin, S.; Leroy, J.; Del Frari, D.; Zagonel, L.; Barrett, N.;<br />

Renault, O.; Bailly, A.; Mariolle, D. J. Mater. Chem. 2008, 18, 3136–3142.<br />

(28) Charlier, J.; Clolus, E.; Bureau, C.; Palacin, S. J. Electroanal. Chem. 2008,<br />

622, 238–241.<br />

(29) Mouanda, B.; Eyeffa, V.; Palacin, S. J. Appl. Electrochem. 2009, 39, 313–<br />

320.<br />

(30) Corgier, B. P.; Belanger, D. Langmuir 2010, 26, 5991–5997.<br />

(31) Garrett, D. J.; Lehr, J.; Miskelly, G. M.; Downard, A. J. J. Am. Chem. Soc.<br />

2007, 129, 15456–15457.<br />

7028 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

printing (MCP) using PDMS stamps and aryldiazonium salt inks<br />

gave micrometer-scale patterns of modifiers on glassy-carbon-like<br />

thin films (pyrolyzed photoresist film; PPF). MCP is a very simple<br />

and relatively fast patterning method and, in these respects, has<br />

obvious advantages over the electrochemical methods outlined<br />

above. 32<br />

In this paper, we investigate the characteristics (surface<br />

concentration, thickness, homogeneity, and chemical reactivity)<br />

of layers prepared using MCP and OCP reduction of aryldiazonium<br />

salts on PPF. We demonstrate that the method can be extended<br />

to metal (Au and Cu) and semiconductor (Si) substrates and<br />

provide guidelines concerning its general applicability in terms<br />

of substrate/diazonium cation combinations and the characteristics<br />

of the resultant films. We also demonstrate a unique feature<br />

of MCP of aryldiazonium salts: covalently coupled two-component<br />

surfaces can be prepared simply by printing a second layer onto<br />

a previously modified surface.<br />

EXPERIMENTAL SECTION<br />

Materials. Aqueous solutions were prepared using Millipore<br />

Milli-Q water (>18 MΩ cm). Tetrafluoroborate salts of 4-nitrobenzenediazonium<br />

(NBD) and 4-carboxybenzenediazonium (CBD)<br />

were synthesized using standard procedures. 33 4-Aminobenzenediazonium<br />

salt (ABD) was synthesized as a 20 mM solution in<br />

0.5 M HCl. 34 Procedures for preparing citrate-capped Au nanoparticles<br />

(∼13 nm diameter), 35 PPF, 36 and planar Au films (Au/<br />

NiCr/Si), 37 and drying acetonitrile (ACN) have been described<br />

previously.<br />

Uncut single walled carbon nanotubes (SWCNTs) (Carbon<br />

Nanotechnologies Incorporated) were acid-treated by adding 25<br />

mg to 27 mL of 3:1 concentrated H2SO4 and HNO3 and sonicating<br />

for 10 h while adding ice to the ultrasonicator bath to maintain<br />

a temperature close to 20 °C. Following sonication, the solution<br />

was poured into 500 mL of distilled water. After standing<br />

overnight, the solution was filtered under suction through<br />

Millipore 0.22 µm hydrophilic polyvinylidene fluoride filter<br />

membranes and then washed with copious amounts of water.<br />

The dried SWCNT cakes were peeled from the filters and<br />

resuspended in DMSO to give a1mgmL -1 stock solution.<br />

Si(100) wafers (1-20 Ω cm, Silicon Quest and Micro Materials)<br />

were cut into ∼15 × 15 mm 2 tiles, immersed in 40% HF (Sigma-<br />

Aldrich) for 3 min (Caution: HF is hazardous; handle with care<br />

and appropriate personal protective clothing), washed with<br />

methanol, dried in a stream of N2 gas, and used within 10 min<br />

of HF treatment. Small pieces of Cu plate were immersed in<br />

16 M HNO3 for 10 s, washed with water, immersed in 17 M<br />

acetic acid for 30 s, and dried in a stream of N2 gas. 38<br />

Fabrication and solvent extraction of PDMS stamps followed<br />

previously described procedures. 24 The stamps were either<br />

nonpatterned or had a test pattern with micrometer-sized<br />

features.<br />

(32) Xia, Y. N.; Whitesides, G. M. Angew. Chem., Int. Ed. 1998, 37, 551–575.<br />

(33) Saunders, K. H.; Allen, R. L. M. Aromatic Diazo Compounds, 3rd ed.; Edward<br />

Arnold: London, 1985.<br />

(34) Lyskawa, J.; Belanger, D. Chem. Mater. 2006, 18, 4755–4763.<br />

(35) Grabar, K. C.; Freeman, R. G.; Hommer, M. B.; Natan, M. J. Anal. Chem.<br />

1995, 67, 735–743.<br />

(36) Brooksby, P. A.; Downard, A. J. Langmuir 2004, 20, 5038–5045.<br />

(37) Lehr, J.; Williamson, B. E.; Flavel, B. S.; Downard, A. J. Langmuir 2009,<br />

25, 13503–13509.<br />

(38) Chamoulaud, G.; Belanger, D. J. Phys. Chem. C 2007, 111, 7501–7507.


Electrochemistry. Electrochemical measurements were made<br />

at room temperature in an N2 atmosphere. The electrochemical<br />

cell exposed a circular area (0.18 cm 2 for PPF or 0.79 cm 2 for<br />

Au) of substrate to the cell solution, as described previously. 36<br />

The auxiliary and reference electrodes were Pt and SCE,<br />

respectively. Cyclic voltammograms were recorded with a scan<br />

rate of 100 mV s -1 .<br />

Surface concentrations of grafted nitrophenyl (NP) groups<br />

were estimated from cyclic voltammograms of modified surfaces<br />

in 0.10 M H2SO4. The area under the irreversible reduction peak<br />

at Ep,c ≈ -0.6 V and the area under the oxidation peak at Ep,a<br />

≈ 0.3 V was determined by fitting the data with polynomial<br />

baselines and mixed Lorentzian-Gaussian curves, using the<br />

Levenberg-Marquardt algorithm implemented via Linkfit software.<br />

36 The surface concentration was then calculated using<br />

Faraday’s law, as previously described, 36 with an estimated<br />

±20% uncertainty in the absolute surface concentration.<br />

Microcontact Printing. Within2hofuse, PDMS stamps were<br />

treated with O2 plasma (100 W at 0.1 Torr for 5 min). Stamps<br />

were immersed in 20 mM aryldiazonium cation ink for 2 min,<br />

dried to tackiness in a stream of N2 gas, and placed on the<br />

substrate for 30 min. Unless stated otherwise, all printed<br />

samples were ultrasonicated for 5 min in Milli-Q water and<br />

dried in a stream of N2 prior to analysis or further treatment.<br />

Each stamp was used with only one type of ink. The compositions<br />

of printing inks are denoted “diazonium cation/solvent”.<br />

Control samples were prepared in the same way but using a<br />

solution from which the aryldiazonium salt had been omitted<br />

(“blank” ink).<br />

Characterization of Films and Patterns. AFM (Digital<br />

Instruments Dimension 3100) depth-profiling measurements were<br />

performed on modified PPF samples by scratching with the AFM<br />

tip to remove a section of film, as described previously. 36 Three<br />

average line profiles were obtained from each of two 10 × 1.25<br />

µm 2 scratches per sample. Each line profile gave two thickness<br />

values: one from the step down into the scratch and the other<br />

from the step out of the scratch. Thus, the reported thickness<br />

for each sample is a mean of 12 values, and the uncertainties<br />

are two standard deviations of the mean. Condensation figures<br />

were obtained by allowing water vapor to condense on surfaces<br />

and imaging using an Olympus BX60 inverted light microscope<br />

equipped with a polarizer and an Olympus DP10 camera.<br />

Scanning electron microscopy (SEM) images were obtained<br />

using a JEOL 7000 high-resolution instrument with an accelerating<br />

voltage of 15 kV.<br />

The procedure for measuring water contact angles has been<br />

described previously. 31 Measurements were made using two 1<br />

µL drops of Milli Q water placed on each duplicate sample or<br />

control. The stated values are the means of the four measurements,<br />

and the uncertainties are two standard deviations of the<br />

mean.<br />

RESULTS AND DISCUSSION<br />

This section is divided into four parts. The first describes<br />

printing with aryldiazonium salt inks on PPF, Au, and Si substrates<br />

using nonpatterned PDMS stamps. The second part demonstrates<br />

that printing with patterned PDMS stamps gives micrometer-scale<br />

patterned layers on Au, PPF, Si, and Cu. The reactivity of the<br />

printed layers is described in the third part, and finally, we show<br />

Figure 1. First (s) and second scan (---) cyclic voltammograms in<br />

0.1MH2SO4 of a PPF surface printed with NBD/1 M H2SO4 ink for<br />

30 min.<br />

that a second modifier can be printed onto an already modified<br />

surface to create patterned, covalently coupled, two-component<br />

surfaces.<br />

Nonpatterned Printing on PPF, Gold, and Silicon. Preliminary<br />

results establishing successful MCP of PPF using DMFand<br />

1 M H 2SO4-based aryldiazonium salt inks have been<br />

reported earlier. 31 In both that work and the present study,<br />

PPF was used as the carbon substrate because aryldiazonium<br />

salts spontaneously graft to its surface at OCP from acidic<br />

aqueous solution 31 and its low surface roughness facilitates<br />

measurement of film thickness. 39 Here, two aryldiazonium<br />

cations were chosen for detailed examination of the printing<br />

process. The NBD derivative was selected as a relatively easily<br />

reduced species (in cyclic voltammograms, Ep,c ≈ 0VvsSCE<br />

in 0.1 M H2SO4 at a scan rate of 100 mV s -1 ) that is additionally<br />

convenient because the resultant grafted NP moieties can be<br />

readily detected and quantified by electroreduction. The CBD<br />

cation was selected as a more difficult to reduce species (Ep,c<br />

≈ -0.3 V vs SCE in 0.1 M H2SO4 at a scan rate of 100 mV s -1 ),<br />

but which gives grafted carboxyphenyl (CP) films whose<br />

chemical reactivity makes them useful as tether layers for<br />

subsequent coupling reagents. CP groups are not electroactive<br />

in the potential range accessible in 0.1 M H2SO4, but their<br />

presence can be detected by water contact angle measurements.<br />

Figure 1 shows consecutive cyclic voltammograms of a PPF<br />

surface printed with NBD/1 M H2SO4 ink using a nonpatterned<br />

stamp. The first scan, from 0.8 to -0.9 V, shows the characteristic,<br />

irreversible reduction (Ep,c ≈ -0.60 V) of surfaceattached<br />

NP groups to aminophenyl (eq 1) and hydroxyaminophenyl<br />

groups (eq 2) but no evidence for physisorbed<br />

aryldiazonium cations, which, if present, would be reduced at Ep,c<br />

≈ 0 V. The return scan shows the reversible oxidation (Ep,a )<br />

0.35 V) of hydroxyaminophenyl to nitrosophenyl groups (eq<br />

3), 40 with the nitrosophenyl/hydroxyaminophenyl redox couple<br />

appearing as the only significant feature in subsequent cycles.<br />

These results are consistent with immobilized NP groups,<br />

spontaneously grafted to the surface during printing. Table 1 lists<br />

electroactive-NP surface concentrations, ΓNP, determined for<br />

three nonpatterned NBD/1 M H2SO4 samples from the<br />

(39) Ranganathan, S.; McCreery, R. L. Anal. Chem. 2001, 73, 893–900.<br />

(40) Ortiz, B.; Saby, C.; Champagne, G. Y.; Belanger, D. J. Electroanal. Chem.<br />

1998, 455, 75–81.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7029


Table 1. Surface Concentration and Film Thickness<br />

Data for NP Films Printed on PPF<br />

1 (5 ± 1) × 10<br />

film<br />

thickness/nm<br />

-10 1.5 ± 0.4<br />

2 (6 ± 1) × 10-10 2.4 ± 0.6<br />

3 (16 ± 3) × 10-10 1.7 ± 0.4<br />

sample ΓNP /mol cm -2<br />

first-scan, along with the corresponding film thicknesses<br />

obtained by AFM depth profiling.<br />

Surface---C6H4-NO2 + 6H + + 6e - f<br />

Surface---C6H4-NH2 + 2H2O (1)<br />

Surface---C6H4-NO2 + 4H + + 4e - f<br />

Surface---C6H4-NHOH + H2O (2)<br />

Surface---C 6 H 4 -NHOH h Surface---C 6 H 4 -NO + 2H + + 2e -<br />

Previously, we have found that single layers of NP groups<br />

electrografted to PPF have ΓNP ) (2.5 ± 0.5) × 10 -10 mol cm -2 , 36<br />

and similar values were obtained for methylphenyl and CP<br />

groups electrografted to flat Au substrates. 41 A monolayer of<br />

vertically oriented NP groups has a calculated thickness of 0.8<br />

nm; 36 hence, both the concentration and thickness data in Table<br />

1 indicate the formation of multilayer domains with an average<br />

thickness of 2 to 3 layers. 18 Table 1 also shows significant<br />

variations between films prepared under the same conditions and<br />

no systematic relationship between ΓNP and average AFMdetermined<br />

film thickness. The initial OCPs differ significantly<br />

between PPF samples, particularly from different preparation<br />

batches, and we attribute the sample-to-sample variations in<br />

Table 1 to this variability of “activity”. We, 37 and others, 37,42,43<br />

have observed that the substrate potential increases as the<br />

spontaneous reduction of aryldiazonium cations proceeds and that<br />

film growth stops when the potential becomes too positive to<br />

sustain reduction. The initial OCP, therefore, influences the<br />

amount of charge that can be transferred before the “cutoff”<br />

potential is reached, and consequently, it helps to determine the<br />

amount of material that can be attached to the surface in the<br />

absence of an externally applied potential.<br />

The lack of a clear relationship between Γ NP and the “average”<br />

film thickness is attributed to differences in the measurement<br />

scales of the associated methods. The surface concentration<br />

data are averages over a large area (0.18 cm 2 ), whereas the<br />

AFM film thicknesses are derived from three depth profiles<br />

across two 10 × 1.25 µm 2 scratches per sample. For films of<br />

highly variable thickness, the average results of a few measurements<br />

performed over a small area may not yield results that<br />

are representative of the whole-sample average. The possibility<br />

of multilayer grafting, coupled with an inhomogeneous distri-<br />

(41) Paulik, M. G.; Brooksby, P. A.; Abell, A. D.; Downard, A. J. J. Phys. Chem.<br />

C 2007, 111, 7808–7815.<br />

(42) Le Floch, F.; Simonato, J.-P.; Bidan, G. Electrochim. Acta 2009, 54, 3078–<br />

3085.<br />

(43) Smith, R. D. L.; Pickup, P. G. Electrochim. Acta 2009, 54, 2305–2311.<br />

7030 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(3)<br />

bution of aryldiazonium cations on the inked stamp, is likely<br />

to lead to films of variable thickness, and the large uncertainties<br />

for the AFM thickness values are consistent with this. Clearly,<br />

such variability will be a limitation when highly reproducible<br />

and homogeneous films are required.<br />

Successful modification of PPF after printing with CBD/1 M<br />

H2SO4 ink was confirmed by a decrease of the water contact<br />

angles from 68 ± 2° to 31 ± 2°, consistent with the attachment<br />

of hydrophilic CP groups. In contrast, control surfaces gave<br />

an unchanged postprinting contact angle of 68 ± 11°. Cyclic<br />

voltammograms (not shown) revealed no signals between 0.8<br />

and -0.5 V, confirming the absence of physisorbed CBD<br />

(expected reduction peak at -0.3 V). AFM depth-profiling<br />

measurements on two films gave an average film thickness of<br />

1.0 ± 0.2 nm, close to that expected for a monolayer. Printed<br />

CP films are, thus, on average, thinner and significantly more<br />

uniform than printed NP films. The different morphology can<br />

be attributed to the lower reduction potential for CBD. Because<br />

CBD is more difficult to reduce than NBD, its reduction ceases<br />

at a lower substrate OCP and a correspondingly smaller amount<br />

of CP is grafted to the surface. The more limited degree of<br />

spontaneous reduction diminishes the prevalence of significant<br />

multilayer “outgrowths”, and a more uniform film thickness<br />

results. Vautrin-Ul and co-workers have reported results<br />

consistent with this interpretation, 44 finding that spontaneous<br />

reduction of NBD rapidly gave thick, nonuniform films on zinc<br />

but formed thin homogeneous layers more slowly on a lessreducing<br />

nickel surface. Hence, for substrates that act as the<br />

reducing agent for aryldiazonium cation-based grafting, when<br />

the potential driving force for reduction is large (the aryldiazonium<br />

derivatives are easily reduced in comparison with the<br />

reducing power of the substrate), thicker and more irregular<br />

films will be formed than when the driving force is lower.<br />

The feasibility of printing using other aryldiazonium salt/<br />

substrate combinations was also examined in experiments described<br />

in the Supporting Information. Electrochemical (Figures<br />

S-1, S-2) or AFM depth-profiling measurements confirmed that<br />

printing of ABD on PPF, NBD on Au, and NBD on Si gave the<br />

expected modified surfaces. For Si samples, the oxide layer was<br />

removed or significantly thinned by HF treatment prior to printing.<br />

Samples with an intact native oxide layer could not be modified.<br />

To summarize, we expect printing to be successful for all<br />

aryldiazonium salt-substrate combinations for which the grafting<br />

reaction proceeds spontaneously at OCP in solution. As is found<br />

for layers grafted from solution, the stability of attachment will<br />

depend mainly on the substrate; for example, very stable layers<br />

are formed by grafting onto graphitic carbon, 2,15 but the stability<br />

of the layers on Au is significantly less. 16<br />

Patterned Microcontact Printing of Gold, PPF, Silicon,<br />

and Cu. Having successfully demonstrated that printing with<br />

aqueous aryldiazonium salt inks leads to surface modification, we<br />

next investigated patterning of surfaces. Figures 2-4 show images<br />

of Au, PPF, Si, and Cu substrates printed with aryldiazonium salt<br />

inks and with blank inks.<br />

In Figure 2, SEM images of Au substrates patterned with NP,<br />

aminophenyl (AP), and CP groups are compared with images from<br />

(44) Adenier, A.; Cabet-Deliry, E.; Chausse, A.; Griveau, S.; Mercier, F.; Pinson,<br />

J.; Vautrin-Ul, C. Chem. Mater. 2005, 17, 491–501.


Figure 2. SEM images of Au substrates patterned by printing with<br />

(a) NBD/1 M H2SO4, (b)1MH2SO4, (c) ABD/0.5 M HCl, (d) 0.5 M<br />

HCl, and (e) CBD/1 M H2SO4 inks.<br />

the corresponding controls. The aryldiazonium salt inks give<br />

clearly defined patterns with feature sizes down to 20 µm, whereas<br />

the blank inks give faint patterns, which are attributed to PDMS<br />

residues.<br />

SEM is not a good technique for imaging organic films on<br />

carbon substrates. Consequently, the image contrast for PPF<br />

patterned with NP (Figure 3a) is just marginally better than that<br />

for the control sample (Figure 3b). To overcome this inherent<br />

limitation, alternative methods were used to image surfaces<br />

patterned with AP and CP groups. For the former, after printing<br />

with ABD/0.5 M HCl (or blank 0.5 M HCl), the surfaces were<br />

immersed for 40 min, at pH ∼ 5, in a solution of citrate-capped<br />

Au nanoparticles. Preferential assembly of nanoparticles (via<br />

electrostatic interactions) on the AP-modified areas clearly revealed<br />

the patterns (compare Figure 3c,d). For surfaces patterned<br />

with CP groups (by printing with CBD/1 M H2SO4 ink) and the<br />

corresponding blanks, condensation figures were imaged by<br />

optical microscopy (Figures 3e,f). Although the pattern is welldefined<br />

in both the CP-printed sample and the blank, the relative<br />

sizes of water droplets in the stamped and “bare” areas are the<br />

opposite in the sample and blank. Water droplets are larger in<br />

the CP areas than on bare PPF, consistent with addition of<br />

hydrophilic CP groups to the surface; in contrast, areas contacted<br />

by the stamp inked with 1MH2SO4 only are smaller than on<br />

bare PPF, suggesting hydrophobic contaminants have been<br />

transferred to the surface by the stamp. (Note that the sizes<br />

of water droplets on bare PPF in Figures 3e,f are not the same<br />

because the size depends on the extent of evaporation prior to<br />

image capture.)<br />

Figure 3. SEM images (a-d) and optical micrographs (e, f) of PPF<br />

surfaces printed with (a) NBD/1 M H2SO4, (b, f) 1 M H2SO4, (c) 20<br />

mM ABD/0.5 M HCl, (d) 0.5 M HCl, and (e) CBD/1 M H2SO4 inks.<br />

The surfaces shown in (c) and (d) were immersed in Au nanoparticle<br />

solution for 40 min before imaging; surfaces shown in (e) and (f) were<br />

treated with water vapor before imaging.<br />

Patterning of NP and AP groups on Si was also successful<br />

as revealed by the SEM images of Figure 4a-d. There is strong<br />

contrast between the grafted and bare areas in Figure 4a,c, in<br />

comparison with the faint patterns of the controls (Figure<br />

4b,d).<br />

As a final example to demonstrate the wider applicability of<br />

MCP with aqueous aryldiazonium salt inks, a Cu surface was<br />

patterned with NP groups. Copper is known to react spontaneously<br />

with NBD at OCP in both aqueous and nonaqueous conditions. 38,45<br />

The SEM image in Figure 4e shows strong contrast between NPprinted<br />

areas and bare Cu, whereas only a very faint pattern is<br />

seen on the control (Figure 4f). The roughness of the Cu surface,<br />

evident in both images, leads to incomplete contact with the stamp<br />

and accounts for the “patchy” appearance of the pattern in Figure<br />

4e.<br />

These examples confirm that MCP is a very simple route to<br />

patterning conducting surfaces. The scope of the method and the<br />

characteristics of the patterned layers will be determined by the<br />

substrate-diazonium salt combination as described in the previous<br />

section.<br />

Printed Tether Layers for Further Immobilization <strong>Chemistry</strong>.<br />

The utility of MCP can be enhanced by printing layers that<br />

act as tethers for further immobilization reactions. Two examples<br />

are demonstrated here, on the basis of coupling of secondary<br />

reagents (4-nitroaniline (NA) and SWCNTs) to primary layers of<br />

CP or AP printed on PPF. The selection of these reagents was<br />

based on their ease of detection: NA by its redox chemistry and<br />

SWCNTs by AFM imaging.<br />

(45) Hurley, B. L.; McCreery, R. L. J. Electrochem. Soc. 2004, 151, B252–B259.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7031


Figure 4. SEM images of (a-d) Si samples patterned by printing<br />

with (a) NBD/1 M H2SO4, (b)1MH2SO4, (c) ABD/0.5 M HCl, and (d)<br />

0.5 M HCl inks; (e, f) Cu samples patterned by printing with (e) NBD/1<br />

MH2SO4 and(f)1MH2SO4 inks.<br />

Table 2. Film Thicknesses for NA Layer Coupled to CP<br />

Film and Associated Controls<br />

surface film thickness/nm a<br />

PPF-CP 1.0 ± 0.3<br />

PPF-CP/SOCl2/NA 2.0 ± 0.5<br />

PPF-CP/NA 1.0 ± 0.4<br />

PPF-H2SO4/SOCl2/NA 0.4 ± 0.2<br />

a<br />

AFM line profiles are shown in Figure S-4 (Supporting Information).<br />

NA + CP on PPF. CP groups were printed on PPF using<br />

nonpatterned stamps and CBD/1 M H2SO4 ink. The films were<br />

activated by immersion in SOCl2 for 30 min and then transferred<br />

to a 20 mM NA/ACN solution at room temperature for<br />

24 h to promote coupling of NA groups to the CP layer via the<br />

formation of amide bonds. These samples are denoted PPF-CP/<br />

SOCl2/NA. Controls were also prepared: PPF-CP blanks were<br />

obtained by printing CP films onto PPF without subsequent<br />

activation or immersion in NA/ACN; for PPF-CP/NA blanks,<br />

only the activation step was omitted; and for PPF-H2SO4/<br />

SOCl2/NA, blanks were prepared by printing PPF with blank<br />

1MH2SO4, followed by “activation” and immersion in NA/<br />

ACN. Prior to their analysis, all samples and controls were<br />

sonicated for 5 min in ACN.<br />

AFM depth profiling results are shown in Table 2, and cyclic<br />

voltammograms of the modified surfaces and typical AFM line<br />

profiles are shown in Figures S-3 and S-4 (Supporting Information).<br />

Activation of the CP film and reaction with NA increased the film<br />

thickness from 1.0 ± 0.3 to 2.0 ± 0.5 nm, consistent with the<br />

coupling of NA groups to the CP layer. When the activation step<br />

7032 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 5. AFM image of VACNTs tethered to a patterned AP layer<br />

on PPF.<br />

was omitted, there was no change in film thickness, confirming<br />

that NA does not physisorb to the CP film. Interestingly, a thin<br />

surface layer of NA was detected electrochemically (Figure S-3,<br />

Supporting Information) and by AFM measurements on the<br />

PPF-H2SO4/SOCl2/NA controls. The measured thickness (0.4<br />

± 0.2 nm) of this film is less than expected for a monolayer of<br />

NA groups (0.8 nm), indicating a submonolayer coverage. We<br />

assume that NA couples directly to a low concentration of<br />

carboxylate functionalities on the (otherwise) bare PPF surface.<br />

SWCNT + AP on PPF. This second example of the utility of<br />

printed films as tethers is based on recent work in which we<br />

assembled and characterized vertically aligned carbon nanotube<br />

(VACNT) forests on AP films electrografted to PPF. 46 To test<br />

whether printed AP films could be used similarly, PPF surfaces<br />

were patterned using ABD/0.5 M HCl ink and then immersed in<br />

a DMSO solution (2 mL) of cut SWCNTs (0.2 mg mL -1 ) and N,N′dicyclohexylcarbodiimide<br />

(1 mg mL -1 )for24hat65°C. These<br />

conditions promote formation of amide bonds between surfaceimmobilized<br />

AP groups and carboxylate groups at the cut ends<br />

of the SWCNTs. The resultant surfaces were sonicated in<br />

acetone for 10 s and then in isopropyl alcohol for 10 s prior to<br />

imaging by AFM. The image shown in Figure 5 (and the SEM<br />

image in Figure S-5, Supporting Information) is similar to those<br />

previously obtained for VACNTs on electrografted AP tether<br />

layers. 46<br />

These examples demonstrate that MCP yields tether layers<br />

with their usual reactivity, and hence, the method can be used to<br />

prepare patterned substrates which form the basis of more<br />

complex structures.<br />

Buildup MCP Patterning of Two-Component Surfaces.<br />

Two- or multicomponent films in which secondary modifiers are<br />

patterned on top of a continuous base film have potential<br />

applications in sensing, where (for example) the base film is<br />

tailored to reduce nonspecific interactions with the analyte while<br />

the patterned secondary modifiers act either as tethers for<br />

attachment of recognition species or as the recognition elements<br />

themselves. This “buildup” method relies on the ability of the<br />

substrate to reduce the secondary modifiers by electron transfer<br />

across the base film. Reduction of the printed secondary modifier<br />

generates radicals which couple to the base film. Hence, a<br />

covalently coupled structure spontaneously forms in a single,<br />

simple step requiring no additional reagents. The buildup method<br />

(46) Garrett, D. J.; Flavel, B. S.; Shapter, J. G.; Baronian, K. H. R.; Downard,<br />

A. J. Langmuir 2010, 26, 1848–1854.


Table 3. Film Thickness Measurements on PPF<br />

Substrates Modified with CP Films with and without<br />

Overprinting of a NP Layer a<br />

film thickness/nm b<br />

sonication solvent and time CP region CP/NP region<br />

H2O (5 min) 1.6 ± 0.3 2.2 ± 0.3<br />

H2O (5 min) 1.4 ± 0.3 2.1 ± 0.2<br />

H2O (5 min) + ACN (30 min) 1.7 ± 0.6 2.1 ± 0.3<br />

H2O (5 min) + ACN (30 min) 1.5 ± 0.4 2.0 ± 0.2<br />

a Samples were prepared in duplicate. b AFM line profiles are shown<br />

in Figure S-7 (Supporting Information).<br />

Figure 6. SEM images of surfaces that were immersed in 10 mM<br />

CBD/0.1 H2SO4 for 30 min and subsequently printed with (a) ABD/<br />

0.5 M HCl and (b) 0.5 M HCl inks. After printing, the surfaces were<br />

immersed in Au nanoparticle solution for 40 min.<br />

was demonstrated by printing the secondary modifiers NP and<br />

AP.<br />

CP + NP on PPF. NP groups were printed onto a base CP<br />

film grafted spontaneously to PPF at OCP. Two ∼3.0 × 1.5 mm 2<br />

PPF samples were immersed in 10 mM CBD/0.1 M H2SO4<br />

solution for 30 min to form the base films. One half of each<br />

sample was then printed with a nonpatterned stamp using<br />

NBD/1 M H2SO4 ink. Cyclic voltammetry (Figure S-6, Supporting<br />

Information) confirmed that NBD had reacted in the<br />

expected manner giving an NP layer.<br />

AFM depth profiling measurements of printed and unprinted<br />

sections (Table 3) show that printing increases the film thickness,<br />

consistent with attachment of NP to the CP layer. The magnitude<br />

of the increase (∼0.4-0.7 nm) corresponds to the addition of a<br />

submonolayer of NP groups on top of the CP film. However, NP<br />

is also expected to couple within the CP film, and hence, the<br />

concentration of printed NP groups may be higher than indicated<br />

by film thickness data.<br />

To test the stability of the printed layers, the samples were<br />

sonicated in ACN for 30 min and the AFM measurements were<br />

repeated. The data in the lower part of Table 3 indicate that the<br />

film thicknesses did not change significantly, consistent with<br />

covalent attachment both to the substrate surface and between<br />

the base and printed layers.<br />

CP + AP on PPF. In the second example of the buildup<br />

printing approach, AP groups were patterned onto a spontaneously<br />

grafted layer of CP groups using a patterned stamp with ABD/<br />

0.5 M HCl ink. A blank was also prepared on a CP layer by printing<br />

with blank 0.5 M HCl ink. The printed surfaces were immersed<br />

for 40 min in a solution of Au nanoparticles and then sonicated in<br />

H 2O for 30 s prior to analysis by SEM. Figure 6 shows the SEM<br />

images where the assembled Au nanoparticles clearly reveal the<br />

patterned immobilization of AP (Figure 6a) in comparison with<br />

the control (Figure 6b).<br />

The buildup approach for preparing patterned two- or multicomponent<br />

surfaces should be applicable to all substrates at which<br />

grafting from aryldiazonium salt solutions proceeds spontaneously<br />

at OCP. However, the requirement that the second aryldiazonium<br />

cation be reduced by the substrate places some limitations on<br />

the nature and thickness of the base film and also on the<br />

aryldiazonium cation derivative. For strongly reducing substrates<br />

(such as zinc), it should be possible to print even relatively difficultto-reduce<br />

aryldiazonium cations onto thick base films. On the<br />

other hand, printing on less-reducing substrates (such as Au) is<br />

likely to be successful only with easily reduced aryldiazonium salt<br />

derivatives on thin base films. In the examples above, the base<br />

films were formed by spontaneous grafting from a aryldiazonium<br />

salt solution. There is no reason why MCP, electrografting, and/<br />

or other classes of modifiers cannot be used. Methods such as<br />

electro-oxidation of primary amines 47 or arylhydrazines; 48 electroreduction<br />

of iodonium, 49,50 sulfonium 51 salts, or vinylic compounds;<br />

52 photolytic or thermal grafting of alkenes and alkynes; 53-56<br />

or photografting of arylazides 57 would greatly widen the range of<br />

functionalities that could be added to the surface.<br />

CONCLUSION<br />

Microcontact printing using aryldiazonium salt inks has been<br />

applied to carbon, metal, and semiconductor substrates to give<br />

stable, covalently attached, thin films. The thickness and the<br />

morphology of the printed film appear to depend on the potential<br />

driving force for reduction of the aryldiazonium cation by the<br />

substrate. Aminophenyl and carboxyphenyl groups in printed<br />

layers retain the ability to form amide bonds with solution species<br />

and, consequently, provide useful tethers for more complex<br />

surface structures.<br />

Microcontact printing using aryldiazonium salts is applicable<br />

to all substrate-diazonium salt combinations for which surface<br />

modification proceeds spontaneously at open circuit potential in<br />

solution. The variable film thickness and roughness, which is<br />

substrate- and film-dependent, may be a limitation for some<br />

potential applications; however, compared with other methods for<br />

patterning layers using aryldiazonium salts, microcontact is low<br />

cost and simple to implement with no requirement for electrochemical<br />

capability. Tightly defined patterns with feature sizes<br />

tens of micrometers upward can be routinely prepared, and<br />

(47) Barbier, B.; Pinson, J.; Desarmot, G.; Sanchez, M. J. Electrochem. Soc. 1990,<br />

137, 1757–1764.<br />

(48) Malmos, K.; Iruthayaraj, J.; Pedersen, S. U.; Daasbjerg, K. J. Am. Chem.<br />

Soc. 2009, 131, 13926–13927.<br />

(49) Vase, K. H.; Holm, A. H.; Norrman, K.; Pedersen, S. U.; Daasbjerg, K.<br />

Langmuir 2007, 23, 3786–3793.<br />

(50) Vase, K. H. j.; Holm, A. H. k.; Pedersen, S. U.; Daasbjerg, K. Langmuir<br />

2005, 21, 8085–8089.<br />

(51) Vase, K. H.; Holm, A. H.; Norrman, K.; Pedersen, S. U.; Daasbjerg, K.<br />

Langmuir 2008, 24, 182–188.<br />

(52) Palacin, S.; Bureau, C.; Charlier, J.; Deniau, G.; Mouanda, B.; Viel, P.<br />

ChemPhysChem 2004, 5, 1469–1481.<br />

(53) Lasseter, T. L.; Cai, W.; Hamers, R. J. Analyst 2004, 129, 3–8.<br />

(54) Ssenyange, S.; Anariba, F.; Bocian, D. F.; McCreery, R. L. Langmuir 2005,<br />

21, 11105–11112.<br />

(55) Sun, B.; Colavita, P. E.; Kim, H.; Lockett, M.; Marcus, M. S.; Smith, L. M.;<br />

Hamers, R. J. Langmuir 2006, 22, 9598–9605.<br />

(56) Yu, S. S. C.; Downard, A. J. Langmuir 2007, 23, 4662–4668.<br />

(57) Gross, A. J.; Yu, S. S. C.; Downard, A. J. Langmuir 2010, 26, 7285–7292.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7033


multicomponent patterned surfaces can be fabricated simply by<br />

printing on top of a base film. The two-component structures were<br />

shown to be stable to prolonged sonication, indicating covalent<br />

coupling of the second modifier to the base layer. The availability<br />

of such a method will facilitate application of the aryldiazonium<br />

salt surface modification approach in areas such as the fabrication<br />

of bio- and chemical sensors.<br />

ACKNOWLEDGMENT<br />

This work was supported by the MacDiarmid Institute for<br />

Advanced Materials and Nanotechnology. J.L. and D.J.G thank<br />

the New Zealand Tertiary Education Commission for doctoral<br />

scholarships and B.S.F. thanks the Australian Government’s<br />

Endeavour Research Fellowship program. We thank Dr. John<br />

Loring for use of the Linkfit curve fitting software.<br />

7034 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

SUPPORTING INFORMATION AVAILABLE<br />

Characterizations of the following modified surfaces prepared<br />

by printing using nonpatterned stamps: AP on PPF, NP on Au,<br />

and NP on HF-treated Si; cyclic voltammograms of CP layers and<br />

bare PPF after activation with SOCl2 and reaction with 4-nitroaniline;<br />

repeat cyclic voltammograms of a PPF surface bearing<br />

a CP film, overprinted with an NP layer; figures of AFM line<br />

profiles of data presented in Tables 2 and 3; SEM image of the<br />

surface shown in Figure 5. This material is available free of charge<br />

via the Internet at http://pubs.acs.org.<br />

Received for review July 6, 2010. Accepted July 10, 2010.<br />

AC101785C


Anal. Chem. 2010, 82, 7035–7043<br />

Quantitation, Visualization, and Monitoring of<br />

Conformational Transitions of Human Serum<br />

Albumin by a Tetraphenylethene Derivative with<br />

Aggregation-Induced Emission Characteristics<br />

Yuning Hong, † Chao Feng, ‡ Yong Yu, †,‡ Jianzhao Liu, † Jacky Wing Yip Lam, † Kathy Qian Luo, ‡,§<br />

and Ben Zhong Tang* ,†,#<br />

Nano Science and Technology Program, Department of <strong>Chemistry</strong>, Institute of Molecular Functional Materials,<br />

Bioengineering Program, and Department of <strong>Chemical</strong> and Biomolecular Engineering, The Hong Kong University of<br />

Science and Technology (HKUST), Clear Water Bay, Kowloon, Hong Kong, China, School of <strong>Chemical</strong> and<br />

Biomedical Engineering, Nanyang Technological University, Singapore 637457, and Department of Polymer Science<br />

and Engineering, Institute of Biomedical Macromolecules, Key Laboratory of Macromolecular Synthesis and<br />

Functionalization of the Ministry of Education, Zhejiang University, Hangzhou 310027, China<br />

Human serum albumin (HSA) is a major protein component<br />

of blood plasma, and its assay is of obvious value to<br />

biological research. We, herein, present a readily accessible<br />

fluorescent bioprobe for HSA detection and quantitation.<br />

A nonemissive tetraphenylethene derivative named<br />

sodium 1,2-bis[4-(3-sulfonatopropoxyl)phenyl]-1,2-diphenylethene<br />

(BSPOTPE) is induced to emit by HSA, showing<br />

a novel phenomenon of aggregation-induced emission<br />

(AIE). The AIE bioprobe enjoys a broad working range<br />

(0-100 nM), a low detection limit (down to 1 nM), and a<br />

superior selectivity to albumins. The fluorescent bioassay<br />

is unperturbed by the miscellaneous bioelectrolytes in the<br />

artificial urine. The AIE luminogen can also be used as a<br />

rapid and sensitive protein stain in gel electrophoresis for<br />

HSA visualization. Utilizing the AIE feature of BSPOTPE<br />

and the Förster resonance energy transfer from HSA to<br />

BSPOTPE, the unfolding process of HSA induced by<br />

guanidine hydrochloride is monitored, which reveals a<br />

multistep transition with the involvement of molten globule<br />

intermediates. Computational modeling suggests that<br />

the AIE luminogens dock in the hydrophobic cleft between<br />

subdomains IIA and IIIA of HSA with the aid of hydrophobic<br />

effect, charge neutralization, and hydrogen bonding<br />

interactions, offering mechanistic insight into the<br />

microenvironment inside the hydrophobic cavity.<br />

Human serum albumin (HSA) is the most abundant protein<br />

in the circulatory system and plays multiple biological functions<br />

in the human body. 1 For example, it regulates water balance<br />

between blood and tissues and serves as a physiological carrier<br />

for various endogenous and exogenous substances that are<br />

* To whom correspondence should be addressed. E-mail: tangbenz@ust.hk.<br />

† Nano Science and Technology Program, Department of <strong>Chemistry</strong>, Institute<br />

of Molecular Functional Materials, and Bioengineering Program, HKUST.<br />

‡ Department of <strong>Chemical</strong> and Biomolecular Engineering, HKUST.<br />

§ Nanyang Technological University.<br />

# Zhejiang University.<br />

(1) Peters, T., Jr. Adv. Protein Chem. 1985, 37, 161.<br />

partially soluble in the bloodstream, such as bilirubin, fatty acids,<br />

and drugs. 2-5 The albumin is synthesized in the liver. A low level<br />

of albumin in the blood serum, known as hypoproteinemia, may<br />

indicate liver failure, cirrhosis, and chronic hepatitis. 6 When blood<br />

passes through healthy kidneys, the kidneys filter out waste<br />

products but keep the substances the body needs, such as albumin<br />

and other proteins. Appearance of an excess amount of proteins<br />

in urine is a sign of chronic kidney disease, which may result in<br />

diabetes, high blood pressure, and problems associated with<br />

inflammation in the kidneys. 7 Many studies have been done on<br />

diabetic nephrosis, with microalbuminuria being identified as an<br />

early sign. 8 Microalbuminuria is commonly diagnosed by elevated<br />

protein concentration (30-300 mg/L) in the urine on at least two<br />

occasions. A higher albumin value is regarded as albuminuria.<br />

The urine-testing dipsticks impregnated with pH sensitive dyes<br />

are normally used for daily health monitoring. 9 The level of<br />

albumin caused by microalbuminuria, however, cannot be accurately<br />

assayed by the dipsticks due to their low sensitivity. 10 It<br />

is, thus, of clinical value to develop effective methods for urinary<br />

protein detection and quantitation.<br />

Colorimetric methods, such as Brandford and Lowry assays,<br />

have traditionally been used for protein quantitation in solutions. 11,12<br />

These methods, however, generally lack sensitivity and accuracy,<br />

(2) Carter, D. C.; Ho, J. X. Adv. Protein Chem. 1994, 45, 153.<br />

(3) (a) Neuzil, J.; Stocker, R. J. Biol. Chem. 1994, 269, 16712. (b) Hu, Y. J.;<br />

Liu, Y.; Xiao, X. H. Biomacromolecules 2009, 10, 517.<br />

(4) Berde, C. B.; Hudson, B. S.; Simoni, R. D.; Sklar, L. A. J. Biol. Chem. 1979,<br />

254, 391.<br />

(5) Sjoholm, I.; Ekman, B.; Kober, A. Mol. Pharmacol. 1979, 16, 767.<br />

(6) Murch, S. H.; Winyard, P. J. D.; Koletzko, S.; Wehner, B.; Cheema, H. A.;<br />

Risdon, R. A.; Phillips, A. D.; Meadows, N.; Klein, N. J.; Walker-Smith, J. A.<br />

Lancet 1996, 347, 1299.<br />

(7) Hoogenberg, K.; Sluiter, W. J.; Dullaart, R. P. F. Acta Endrocrinol. 1993,<br />

129, 151.<br />

(8) (a) Mogensen, C. E.; Keane, W. F.; Bennett, P. H.; Jerums, G.; Parving,<br />

H. H.; Passa, P.; Steffes, M. W.; Striker, G. E.; Viberti, G. C. Lancet 1995,<br />

346, 1080. (b) Nomata, S.; Haneda, K.; Moriya, T.; Katayama, S.; Iwamoto,<br />

Y.; Sakai, H.; Tomino, Y.; Matsuo, S.; Asano, Y.; Makino, H. Jpn. J. Nephrol.<br />

2005, 47, 768.<br />

(9) Rose, B. D. Pathophysiology of Renal Disease; McGraw Hill: New York, 1987.<br />

(10) Martinez, A. W.; Phillips, S. T.; Butte, M. J.; Whitesides, G. M. Angew. Chem.,<br />

Int. Ed. 2007, 46, 1318.<br />

10.1021/ac1018028 © 2010 American <strong>Chemical</strong> Society 7035<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/20/2010


in addition to their lengthy procedures. 13 Biosensors based on<br />

fluorescent (FL) materials have attracted much attention due to<br />

their sensitivity, selectivity, and rapidity. Organic molecules and<br />

organometallic complexes have been developed for protein<br />

assays. 14,15 Upon complexation or conjugation with proteins, the<br />

luminophores show emission enhancements and/or spectral<br />

shifts, which enable the biopolymers to be detected and quantified.<br />

Some of the FL probes, however, are insoluble in aqueous media,<br />

while others are unstable under ambient conditions. 16 Some<br />

luminophores show spectacular performances in protein assays<br />

but are extremely expensive because of their painstaking<br />

syntheses. 14,15 These drawbacks greatly limit the scope of their<br />

real-life high-tech applications. This calls for the development of<br />

synthetically readily accessible and environmentally stable bioprobes<br />

with high sensitivity and selectivity.<br />

A thorny problem associated with the emissions of conventional<br />

luminophores in aqueous medium or physiological buffer<br />

is aggregation-caused quenching (ACQ). 17 The luminophores are<br />

in close vicinity in the aggregates suspended in the aqueous buffer,<br />

which favors the formation of such detrimental species as<br />

excimers and exciplexes and causes nonradiative relaxations of<br />

the excited states. We have recently discovered that tetraphenylethene<br />

(TPE) behaves in a way exactly opposite to the ACQ<br />

dyes: it is nonemissive in the solution state but becomes highly<br />

luminescent in the aggregate state. 18 We coined “aggregationinduced<br />

emission” (AIE) for the phenomenon because the<br />

nonemissive TPE is induced to emit by aggregate formation.<br />

Decorating TPE with ionic or polar functional groups yields watersoluble<br />

derivatives, which can be utilized as FL probes for<br />

bioanalyses. 19,20 Our and other research groups have successfully<br />

used the TPE-based AIE luminogens for nucleic acid detection,<br />

enzymatic activity assay, and metallic ion tracing. 21,22 The successes<br />

prompted us to further explore their potentials in bioanaly-<br />

(11) Bradford, M. M. Anal. Biochem. 1976, 72, 248.<br />

(12) Lowry, O. H.; Rosebrough, N. J.; Farr, A. L.; Randall, R. J. J. Biol. Chem.<br />

1951, 193, 265.<br />

(13) Haugland, R. P. Handbook of Fluorescent Probes and Research <strong>Chemical</strong>s;<br />

Molecular Probe: Leiden, 2002, p. 420.<br />

(14) (a) Giepmans, B. N. G.; Adams, S. R.; Ellisman, M. H.; Tsien, R. Y. Science<br />

2006, 312, 217. (b) Royer, C. A. Chem. Rev. 2006, 106, 1769. (c) Suzuki,<br />

Y.; Yokoyama, K. J. Am. Chem. Soc. 2005, 127, 17799. (d) Matulis, D.;<br />

Baumann, C. G.; Bloomfield, V. A.; Lovrien, R. E. Biopolymers 1999, 49,<br />

451. (e) Yarmoluk, S. M.; Kryvorotenko, D. V.; Balanda, A. O.; Losytskyy,<br />

M. Y.; Kovalska, V. B. Dyes Pigm. 2001, 51, 41. (f) Hawe, A.; Sutter, M.;<br />

Jiskoot, W. Pharm. Res. 2008, 25, 1487.<br />

(15) (a) Eryazici, I.; Moorefield, C. N.; Newkome, G. R. Chem. Rev. 2008, 108,<br />

1834. (b) Keefe, M. H.; Benkstein, K. D.; Hupp, J. T. Coord. Chem. Rev.<br />

2000, 205, 201.<br />

(16) (a) Matulis, D.; Lovrien, R. Biophys. J. 1998, 74, 422. (b) Davis, D. M.;<br />

Birch, D. J. S. J. Fluoresc. 1996, 6, 23.<br />

(17) Birks, J. B. Photophysics of Aromatic Molecules; Wiley: London, 1970.<br />

(18) Hong, Y.; Lam, J. W. Y.; Tang, B. Z. Chem. Commun. 2009, 4332.<br />

(19) Li, Z.; Dong, Y. Q.; Lam, J. W. Y.; Sun, J.; Qin, A.; Haussler, M.; Dong,<br />

Y. P.; Sung, H. H. Y.; Williams, I. D.; Kwok, H. S.; Tang, B. Z. Adv. Funct.<br />

Mater. 2009, 19, 905.<br />

(20) (a) Yuan, C. X.; Tao, X. T.; Wang, L.; Yang, J. X.; Jiang, M. H. J. Phys. Chem.<br />

C 2009, 113, 6809. (b) Zhao, M.; Wang, M.; Liu, H.; Liu, D.; Zhang, G.;<br />

Zhang, D.; Zhu, D. Langmuir 2009, 25, 676. (c) Wang, M.; Zhang, D.;<br />

Zhang, G.; Zhu, D. Chem. Commun. 2008, 4469. (d) Suzuki, Y.; Yokoyama,<br />

K. J. Am. Chem. Soc. 2005, 127, 17799.<br />

(21) (a) Tong, H.; Hong, Y.; Dong, Y.; Haussler, M.; Lam, J. W. Y.; Li, Z.; Guo,<br />

Z.; Guo, Z.; Tang, B. Z. Chem. Commun. 2006, 3705. (b) Tong, H.; Hong,<br />

Y.; Dong, Y.; Haussler, M.; Li, Z.; Lam, J. W. Y.; Dong, Y.; Sung, H. H. Y.;<br />

Williams, I. D.; Tang, B. Z. J. Phys. Chem. B 2007, 111, 11817. (c) Hong,<br />

Y.; Haussler, M.; Lam, J. W. Y.; Li, Z.; Sin, K. K.; Dong, Y.; Tong, H.; Liu,<br />

J.; Qin, A.; Renneberg, R.; Tang, B. Z. Chem.sEur. J. 2008, 14, 6428.<br />

7036 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Chart 1. <strong>Chemical</strong> Structure of Water-Soluble AIE<br />

Luminogen of BSPOTPE<br />

ses. In this work, we examined the utility of a TPE salt, sodium<br />

1,2-bis[4-(3-sulfonatopropoxyl)phenyl]-1,2-diphenylethene(BSPOTPE;<br />

Chart 1), as a bioprobe for HSA detection and quantitation. The<br />

FL “turn-on” attribute of BSPOTPE by its complexation with<br />

albumin facilitated the quantitative assay and visual observation<br />

of HSA in the aqueous buffer and gel electrophoresis, respectively.<br />

It is well-known that proper biological functions of proteins<br />

are associated with their specific strand conformations and folding<br />

structures. 23 Understanding of protein folding is of fundamental<br />

importance for proteomic and pharmaceutical research. 24 HSA is<br />

a polypeptide chain with three R-helical domains (I-III), which<br />

are further divided into two subdomains (A and B). 1,25 Its crystal<br />

structure shows that the main ligand-binding sites in the albumin<br />

are located in the hydrophobic cavities of subdomains IIA and<br />

IIIA, which are sometimes referred to as Sudlow sites I and II,<br />

respectively. 26 Conformation analyses of the hydrophobic cavities<br />

play an important role in drug development, especially pharmacokinetic<br />

and pharmacodynamic investigations. 27<br />

Study of conformation transitions of proteins in the presence<br />

of denaturants is a topic of great interest because it can offer<br />

mechanistic insights into folding and unfolding processes of the<br />

biopolymers. 28,29 Though intermediate states have been suspected<br />

to be involved in the unfolding and refolding processes of many<br />

proteins, they are often not detected due to the lack of appropriate<br />

probes. 29 Characterization of the intermediate states becomes even<br />

more complex in the multidomain proteins, such as HSA, in which<br />

each domain is capable of unfolding and refolding independently. 30<br />

Whether intermediate states are involved in the unfolding pathway<br />

of HSA has been an issue of debate. In this study, we made use<br />

of the AIE feature of BSPOTPE and investigated the unfolding<br />

process of HSA. A stable molten-globule intermediate was<br />

observed in its unfolding process induced by guanidine hydrochloride<br />

(GndHCl), a well-known denaturant. Förster resonance<br />

energy transfer (FRET) study proved the accessibility of HSA by<br />

BSPOTPE and suggested probable location of the FL bioprobe<br />

in the hydrophobic cavity in the protein folding structure.<br />

Combination of the FL technique with circular dichroism (CD),<br />

(22) (a) Wang, M.; Gu, X.; Zhang, G.; Zhang, D.; Zhu, D. Anal. Chem. 2009,<br />

81, 4444. (b) Wang, M.; Zhang, D.; Zhang, G.; Tang, Y.; Wang, S.; Zhu, D.<br />

Anal. Chem. 2008, 80, 6443.<br />

(23) Flora, K.; Brennan, J. D.; Baker, G. A.; Doody, M. A.; Bright, F. V. Biophys.<br />

J. 1998, 75, 1084.<br />

(24) Jusko, W. J.; Gretch, M. Drug Metab. Rev. 1976, 5, 43.<br />

(25) He, X. M.; Carter, D. C. Nature 1992, 358, 209.<br />

(26) (a) Sudlow, G.; Birkett, D. J.; Wade, D. N. Mol. Pharmacol. 1975, 11, 824.<br />

(b) Sudlow, G.; Birkett, D. J.; Wade, D. N. Mol. Pharmacol. 1976, 12, 1052.<br />

(27) Lavinder, J. J.; Hari, S. B.; Sullivan, B. J.; Magliery, T. J. J. Am. Chem. Soc.<br />

2009, 131, 3794.<br />

(28) Abou-Zied, O. K.; Al-Shihi, O. I. K. J. Am. Chem. Soc. 2008, 130, 10793.<br />

(29) (a) Nolting, B.; Andert, K. Protein Struct. Funct. Genet. 2000, 41, 288. (b)<br />

Santra, M. K.; Banerjee, A.; Krishnakumar, S. S.; Rahaman, O.; Panda, D.<br />

Eur. J. Biochem. 2004, 271, 1789.<br />

(30) Ahmad, B.; Ahmed, M. Z.; Haq, S. K.; Khan, R. H. Biochim. Biophys. Acta<br />

Protein Proteomics 2005, 1750, 93.


Figure 1. (A) FL spectra of BSPOTPE in the PBS buffer containing different amounts of HSA. (B) Change in the FL intensity at 475 nm with<br />

HSA concentration; I0 ) FL intensity in the absence of HSA. Inset: linear region of the binding isotherm of BSPOTPE to HSA. [BSPOTPE] )<br />

5 µM; λex ) 350 nm.<br />

differential scanning calorimetry (DSC), and molecular modeling<br />

helped draw a mechanistic picture on the protein unfolding<br />

process.<br />

MATERIALS AND METHODS<br />

General Information. BSPOTPE was prepared according to<br />

our previously published procedures. 21b HSA, GndHCl, and other<br />

biomolecules were all purchased from Sigma and used as received.<br />

Phosphate buffered saline (PBS) with pH of 7.0 was purchased<br />

from Merck. Water was purified by a Millipore filtration system.<br />

All the experiments were performed at room temperature unless<br />

otherwise specified.<br />

UV spectra were measured on a Milton Roy Spectronic 3000<br />

Array spectrophotometer, and FL spectra were recorded on a<br />

Perkin-Elmer LS 55 spectrofluorometer with a Xenon discharge<br />

lamp excitation. CD spectra were recorded on a Jasco J-810<br />

spectropolarimeter in a 1 mm quartz cuvette using a step<br />

resolution of 0.2 nm, a scan speed of 100 nm/min, a sensitivity of<br />

0.1°, and a response time of 0.5 s. Each spectrum is the average<br />

of three scans. Details about the artificial urine preparation,<br />

cytotoxicity assay, FRET study, DSC and pH-dependent FL<br />

measurements, and computation modeling are given in the<br />

Supporting Information.<br />

Sample Preparation. Stock solutions of BSPOTPE and HSA<br />

with a concentration of 1.0 mM were prepared by dissolving<br />

appropriate amounts of the luminogen and protein in the PBS<br />

buffer. The final concentration of HSA in PBS was double checked<br />

by measuring its absorbance at 279 nm. In the HSA unfolding<br />

study, HSA (0.4 µM) in PBS was incubated in the presence of<br />

different amounts of GndHCl (0.2-7.0 M) at 25 °C for 30 min.<br />

Afterward, BSPOTPE was added to the mixtures and incubated<br />

for another 30 min before spectral measurements. FL spectra were<br />

recorded in the wavelength range of 370-670 nm using 350 nm<br />

as the excitation wavelength. For the intrinsic tryptophan fluorescence<br />

study, FL spectra were collected from 310 to 570 nm<br />

using 290 nm as the excitation source. In the refolding experiment,<br />

0.2 mM HSA was incubated in 8 M GndHCl for 24 h at 25 °C to<br />

ensure that all the proteins were fully unfolded. The denatured<br />

sample was then diluted with PBS until the final concentrations<br />

of HSA and GndHCl were equal to 0.4 µM and less than 0.1 M,<br />

respectively. Fraction of refolded proteins (F r) was calculated<br />

from the following equation:<br />

F r ) 1 - I N - I R<br />

I N - I D<br />

where IN and ID are the FL intensities of native and denatured<br />

HSA-BSPOTPE complexes, respectively, and IR is the FL<br />

intensity of BSPOTPE recovered from the refolded HSA.<br />

Electrophoresis Assay. A poly(acrylamide) gel electrophoresis<br />

(PAGE) experiment was performed on a Hoefer miniVE system<br />

under nondenaturing conditions using 5% stacking and 12%<br />

resolving native poly(acrylamide) gel at 100 V for 3hat4°C.<br />

After electrophoresis, the gel was soaked in an aqueous solution<br />

of BSPOTPE (1 mg in 100 mL ddH2O) at room temperature for<br />

5 min. Alternatively, the gel was stained with a Coomassie<br />

Brilliant Blue (CBB) solution (0.1% w/v Coomassie Blue R250<br />

in an aqueous mixture containing 10% methanol and 7% acetic<br />

acid) for6htoovernight on a rotary shaker with gentle mixing,<br />

followed by destaining in an aqueous solution containing 10%<br />

methanol and 7% acetic acid for 1 to 2 h until the background<br />

of the gel became transparent. An AlphadigiDoc system with<br />

a DE-500 MultiImage II light cabinet and an ML-26 UV<br />

transilluminator (Alpha Innotech) was used for data collection<br />

and analysis.<br />

RESULTS AND DISCUSSION<br />

Protein Quantitation in Solution. BSPOTPE is soluble in<br />

water but insoluble in common organic solvents, such as acetonitrile,<br />

THF, and chloroform. Its FL quantum yield (ΦF) is<br />

increased from 0.37% in water to 17.5% in acetonitrile, where<br />

the luminogen molecules aggregate. The BSPOTPE solution<br />

in PBS is feebly luminescent at 390 nm in the absence of HSA<br />

(Figure 1A). When a small amount of HSA is added, the<br />

BSPOTPE solution becomes luminescent. Its FL intensity at 475<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(1)<br />

7037


Figure 2. (A) Binding isotherm of HSA to BSPOTPE in the artificial urine and PBS buffer. (B) Dependence of the FL intensity of BSPOTPE at<br />

470 nm on different proteins in artificial urine and PBS buffer. [BSPOTPE] ) 5 µM; [protein] ) 100 µg/mL; λex ) 350 nm.<br />

nm keeps rising with an increase in the HSA concentration. The<br />

rate of the FL enhancement is fast at lower HSA concentrations<br />

and becomes almost constant at [HSA] > 1 µM (Figure 1B). At<br />

an HSA concentration of 10 µM, the FL intensity is increased by<br />

as high as ∼300-fold. The detection limit can be squeezed to as<br />

low as 1 nM. In the protein concentration range of 0-100 nM,<br />

the plot of FL enhancement (I/I0 - 1) as a function of HSA<br />

concentration is a linear line with a high correlation coefficient<br />

(0.995).<br />

To examine the feasibility of protein assay in body fluids, we<br />

performed HSA quantitation in artificial urine. 31 As shown in<br />

Figure 2A, the binding isotherm is practically identical to that in<br />

PBS. The assay sensitivity is not affected by the presence of<br />

miscellaneous bioelectrolytes and physiological level of urea.<br />

Thanks to the intriguing AIE characteristic of BSPOTPE, the<br />

detection limit and linear range can be tuned by adjusting the<br />

luminogen concentration. 21b Besides the superior sensitivity,<br />

BSPOTPE shows excellent selectivity to albumin proteins, such<br />

as HSA and bovine serum albumin (BSA; Figure 2B). A variety<br />

of human proteins [pepsin, human immunoglobulin G (IgG),<br />

trypsin, etc.] with isoelectric points ranging from 1 to 10 and DNA<br />

[e.g., calf thymus DNA] were chosen for the binding tests. None<br />

of these biomolecules, however, can turn on the BSPOTPE<br />

emission. The selectivity of BSPOTPE toward the albumin proteins<br />

remains unperturbed in the artificial urine. These exciting results<br />

encourage us to further explore its clinical utility.<br />

Protein Visualization in PAGE. The “lighting-up” of BSPOTPE<br />

luminogen by albumin in the PBS buffer inspired us to check the<br />

possibility of the use of the AIE luminogen as a protein staining<br />

reagent in the PAGE assay. Figure 3A shows the gel images of<br />

electrophoresed HSA after staining with a BSPOTPE solution for<br />

∼5 min. The protein bands become visible under the illumination<br />

of a hand-held UV lamp. The detection limit can be down to 50<br />

(31) Brooks, T.; Keevil, C. W. Lett. Appl. Microbiol. 1997, 24, 203.<br />

7038 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Figure 3. PAGE analyses of HSA using (A) BSPOTPE and (B)<br />

Coomassie Blue R-250 as the staining reagents. Lanes correspond<br />

to the protein bands with amounts of HSA of (1) 1000, (2) 500, (3)<br />

100, (4) 50, and (5) 2.5 ng. [BSPOTPE] ) 1 mg/100 mL; staining<br />

time ≈ 5 min.<br />

ng per lane (Lane 4). CBB is one of the most commonly used<br />

protein stains in the bioassay. 32 The gel stained by CBB is shown<br />

in Figure 3B for comparison. Although we have made efforts to<br />

enhance the contrast of the image, only a few bands are<br />

discernible in the gel. Those fine bands containing trace amounts<br />

of HSA are ignored by CBB. It must be pointed out that a high<br />

CBB concentration is generally needed for staining. Evidently,<br />

BSPOTPE shows a much higher sensitivity than CBB.<br />

Another advantage of the use of BSPOTPE as gel stain is its<br />

simplicity and promptness, in contrast to the conventional methods.<br />

Commercially available protein staining reagents such as<br />

colorimetric CBB and fluorimetric SYPRO Ruby require a long<br />

staining time (normally overnight). A destaining step is often<br />

required to remove the excess dye in the gel. 33 Silver staining is<br />

expensive and entails several labor-intensive and time-sensitive<br />

steps. 13 Unlike the above probes, neither careful timing nor a<br />

destaining step is necessary for BSPOTPE staining. The protein<br />

bands can be visualized after 5 min of staining. Soaking the gel<br />

(32) (a) Blakesley, R. W.; Boezi, J. A. Anal. Biochem. 1977, 82, 580. (b)<br />

Candiano, G.; Bruschi, M.; Musante, L.; Santucci, L.; Ghiggeri, G. M.;<br />

Carnemolla, B.; Orecchia, P.; Zardi, L.; Righetti, P. G. Electrophoresis 2004,<br />

25, 1327.<br />

(33) Lopez, M. F.; Berggren, K.; Chernokalskaya, E.; Lazarev, A.; Robinson, M.;<br />

Patton, W. F. Electrophoresis 2000, 21, 3673.


Figure 4. (A) Variation in the FL intensity of BSPOTPE at 470 nm with concentration of GndHCl in the presence or absence of HSA. (B)<br />

Dependence of the FRET ratio (I470/I350) on the concentration of GdnHCl. [BSPOTPE] ) 5 µM; [HSA] ) 2 µM; λex ) 350 nm (for A), 295 nm (for<br />

B).<br />

in the luminogen solution for longer time does not cause<br />

overstaining. As the resolution of the protein band is appreciably<br />

high, no washing step is required, which vastly shortens the time<br />

and lowers the workload.<br />

Irritant acetic acid and methanol are used in the gel electrophoresis<br />

assays using CBB and SYPRO Ruby dyes as staining<br />

reagents. BSPOTPE is soluble and stable in aqueous media, and<br />

the use of toxic chemicals can, thus, be avoided. To learn whether<br />

BSPOTPE is a “safe” stain, cytotoxicity assays were performed<br />

with HeLa cells on the basis of reduced activity of methyl thiazolyl<br />

tetrazolium (MTT). As can be seen from Figure S1 (Supporting<br />

Information), the living cells grow as normally as they do in the<br />

control experiment in the absence of BSPOTPE. The luminogen<br />

neither inhibits nor promotes the growth of the cells. In other<br />

words, it is cytocompatible without interfering with the metabolisms<br />

of the living cells.<br />

Conformation Studies of HSA by BSPOTPE. A few FL dyes<br />

have been used to investigate the HSA unfolding process induced<br />

by GndHCl, but each of them gives a different result. 14c,29b,34<br />

Whether intermediate states are formed and how many steps are<br />

involved in the unfolding process remain to be a controversial<br />

issue. In this study, we tried to find out answers by utilizing the<br />

AIE characteristic of the BSPOTPE luminogen.<br />

Free BSPOTPE is practically nonluminescent in the aqueous<br />

buffer, which is virtually unaffected by the addition of different<br />

amounts of GndHCl (open circles in Figure 4A). Addition of HSA,<br />

however, turns on the BSPOTPE emission. The FL intensity at<br />

475 nm changes when the HSA is denatured by GndHCl. Analysis<br />

of the isotherm shown by the solid circles in Figure 4A reveals<br />

that three transition steps are involved in the GndHCl-induced<br />

denaturation process. The first transition step occurs at low<br />

concentrations of GndHCl (


Figure 5. (A) CD spectra of HSA in the PBS buffer in the presence of different amounts of GndHCl. (B) Plot of ellipticity of HSA at 220 nm<br />

versus GndHCl concentration.<br />

and HSA in the different steps are estimated. (See Supporting<br />

Information for details.)<br />

The change in the secondary structure of HSA during the<br />

GndHCl-induced denaturation process was monitored by far-UV<br />

CD. The CD spectrum of HSA in the PBS buffer shows an R-helix<br />

profile with two minima at 208 and 220 nm (Figure 5A). 35 When<br />

the GndHCl concentration is lower than 1.5 M, the CD signals<br />

remain almost unchanged. Afterward, the ellipticity starts to<br />

decrease progressively and continuously, indicating that the<br />

protein chains are gradually transformed from helical strands to<br />

random coils.<br />

The change in the ellipticity at 220 nm against the GndHCl<br />

concentration is plotted in Figure 5B. No intermediate states are<br />

detected by the CD technique. It is known that the Cotton effects<br />

in the far-UV region provide valuable information on the secondary<br />

structures of proteins. 30 Details on the tertiary structure, however,<br />

are not available in this spectral region. The plateau of the<br />

transition curve suggests that the secondary structure of HSA<br />

remains intact at the low GndHCl concentrations. The attenuation<br />

of the FL signals by GndHCl in this region, however, is inconsistent<br />

with the CD data. This suggests that the tertiary structure<br />

of HSA, rather than its secondary structure, is perturbed at the<br />

low GndHCl concentrations. The partial recovery of the fluorescence<br />

of BSPOTPE at ∼2.0 M GndHCl hints the formation of a<br />

stable intermediate. At the high GndHCl concentrations, the HSA<br />

strand is unfolded and the CD signal is accordingly weakened.<br />

Accompanying the unfolding process, the luminogen molecules<br />

are released back to the buffer solution, leading to the observed<br />

decrease in the FL intensity.<br />

DSC has been commonly used for the studies of stability of<br />

biological macromolecules and folding of proteins. 36 The thermal<br />

transition midpoint (Tm), where 50% of biomolecules are unfolded,<br />

can directly be obtained from the DSC measurement.<br />

The higher the Tm value, the more stable is the structure. The<br />

Tm values for HSA in the presence of 0, 0.8, 1.5, and 2.0 M<br />

(35) Sreerama, N.; Venyaminov, S. Y.; Woody, R. W. Protein Sci. 1999, 8, 370.<br />

(36) Bruylants, G.; Wouters, J.; Michaux, C. Curr. Med. Chem. 2005, 12, 2011.<br />

7040 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

GndHCl are 63.5, 46.9, 34.9, and 46.5 °C, respectively (Supporting<br />

Information, Figure S4). Note that there is no transition<br />

midpoint for HSA in the presence of 6.0 M GndHCl, where the<br />

protein folding structure is completely collapsed. From this set<br />

of data, we can elicit that the stability of the molten-globule<br />

intermediate at 2.0 M GndHCl is lower than that of the native<br />

folding structure but much higher than that of the unfolded<br />

random coil.<br />

Protein denaturation by GndHCl is reversible when GndHCl<br />

is removed. 37 Refolding of the unfolded protein was initiated by<br />

dilution in PBS. A control experiment was carried out by dilution<br />

using the 6 M GndHCl solution. Whereas BSPOTPE is almost<br />

nonluminescent at6MofGndHCl, its luminescence is revitalized<br />

with ∼98% recovery after GndHCl has been removed (Figure 6A).<br />

The refolding fraction is calculated according to eq 1 given in the<br />

Materials and Methods section. Similarly, the CD spectral data<br />

show that the secondary structure of HSA is ∼95% recovered after<br />

refolding (Figure 6B).<br />

Mechanistic Understanding. The unfolding process of HSA<br />

induced by the GndHCl denaturant has previously been investigated<br />

using intrinsic tryptophan fluorescence, bilirubin and<br />

bromophenol blue absorption, and far-UV CD analyses. The<br />

unfolding transition was proposed to follow a single-step, two-state<br />

pathway. 38 On the other hand, the studies using extrinsic FL dyes,<br />

such as Nile red and ANS, suggest that the protein denaturation<br />

occurs through a two-step, three-state transition. 29b Our study in<br />

this work, however, indicates a three-step course for the folding<br />

structure of the protein to transform to random coil.<br />

Mechanistic understanding of the emission behaviors of<br />

BSPOTPE is a prerequisite to interpret the unfolding pathway of<br />

HSA. In our previous work, we have proved that the restriction<br />

of intramolecular rotation is the main cause for the AIE effect. 18<br />

This mechanism is applicable to the case of BSPOTPE. In the<br />

(37) Fanali, G.; De Sanctis, G.; Gioia, M.; Coletta, M.; Ascenzi, P.; Fasano, M.<br />

J. Biol. Inorg. Chem. 2009, 14, 209.<br />

(38) (a) Halim, A. A. A.; Kadir, H. A.; Tayyab, S. J. Biochem. 2008, 144, 33. (b)<br />

Muzammil, S.; Kumar, Y.; Tayyab, S. Protein Struct. Funct. Genet. 2000,<br />

40, 29.


Figure 6. (A) FL spectra of BSPOTPE bound with refolded HSA in the PBS buffer and unfolded HSA in the 6 M GndHCl solution. (B) CD<br />

spectra of refolded HSA in the PBS buffer and unfolded HSA in the 6 M GndHCl solution. The spectra for the native HSA are given for comparison.<br />

aqueous solution, the luminogen molecules are well dissolved and<br />

dispersed as isolated molecules. The active torsional/rotational<br />

motions annihilate their excited states, making them nonluminescent<br />

in the solution state. Their conformations are rigidified<br />

when they are aggregated, which effectively blocks the nonradiative<br />

channels and activates their radiative transitions. We have<br />

previously observed that cationic TPE derivatives bind to DNA<br />

mainly through electrostatic interaction. 21c However, the noncanonical<br />

change in the emission of the HSA-bound BSPOTPE with<br />

pH suggests that the electrostatic attraction is not the main driving<br />

force for BSPOTPE to bind with HSA (Supporting Information,<br />

Figure S5).<br />

To gain insights into the binding mode and to better understand<br />

the favorable interactions between BSPOTPE and HSA, we<br />

conducted computation modeling according to the procedures<br />

described in the Supporting Information. The 50 conformations<br />

of the docked BSPOTPE molecules are summarized in Figure 7.<br />

BSPOTPE shows a high preference to the hydrophobic cavity<br />

surrounded by subdomains IIIA, IB, and IIA. This cavity is<br />

adjacent to W214, with 22 out of the 50 docked conformations<br />

located in it, indicative of a high probability of occurrence of the<br />

luminogen molecules in this structural region.<br />

We then compared the root-mean-square-deviation (rmsd)<br />

values between any two of the 50 docked conformations and<br />

grouped the similar conformations into clusters with a rmsd<br />

tolerance of 2.0 Å (Supporting Information, Figure S6). We chose<br />

two of the biggest clusters, namely clusters 1 and 2, in the<br />

hydrophobic cavity for the analysis of the binding sites for<br />

BSPOTPE in HSA. Since the detailed binding modes of conformations<br />

in each cluster are almost identical, the conformation with<br />

the lowest estimated binding energy in each cluster is shown as<br />

a representative to illustrate the interaction details (Figure 8).<br />

Generally speaking, the luminogen molecule docks in the same<br />

location and adopts a similar pose in the two clusters. The<br />

BSPOTPE luminogen inserts into the interdomain cleft formed<br />

by the R-helices from subdomains IIA (green helix) and IIIA (blue<br />

Figure 7. Summary of the 50 conformations of BSPOTPE docked<br />

on HSA, whose domains are denoted by different colors, with red,<br />

green, and blue for domains I, II, and III, respectively. Each of the<br />

docked conformation of BSPOTPE is represented by a sphere placed<br />

at the average position of the coordinates of all the atoms in that<br />

conformation. The tryptophan residue in domain II (W214) is marked<br />

by a purple circle. BSPOTPE shows a high preference (22 out of 50)<br />

to the hydrophobic cavity (circled by red color) adjacent to the W214<br />

residue.<br />

helix). The aromatic core of the luminogen stacks on the R-helix<br />

via hydrophobic interaction, while its sulfonate peripheral arms<br />

interact with the polar patches inside the pocket of the protein<br />

folding structure.<br />

In the binding conformations of cluster 1 (Figure 8A), Lys195,<br />

Lys199, and Arg218 in HSA serve as proton donors to form<br />

hydrogen bonds with oxygen atoms of the sulfonate group of<br />

BSPOTPE. Lys274 on the adjacent helix forms another hydrogen<br />

bond with the other sulfonate group of BSPOTPE. The nitrogen<br />

atom in the side chain of Trp214 orients toward the electrondeficient<br />

sulfonate group. On the other hand, Arg218 and Arg222<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7041


Figure 8. Details of the best binding conformations with the lowest binding free energies in clusters (A) 1 and (B) 2 (cf., Supporting Information,<br />

Figure S6). In each case, the BSPOTPE molecule is shown in a stick representation with a semitransparent van der Waals surface and with a<br />

color scheme of gray for carbon, blue for nitrogen, red for oxygen, and yellow for sulfur. The residues interacting with BSPOTPE are shown as<br />

sticks (with nonbinding apolar hydrogen atoms omitted for clarity) and labeled on R carbon. Yellow dashed lines denote hydrogen bonds.<br />

accommodate the aromatic core of BSPOTPE via cation-π<br />

interaction. A similar binding mode is found in clusters 2 (Figure<br />

8B). Additionally, Pro447 contacts the hydrophobic part of the<br />

luminogen, while His440 interacts with its sulfonate group. All<br />

the aforesaid interactions rigidify the molecular conformation of<br />

the BSPOTPE luminogen and hamper its intramolecular rotation,<br />

hence making it highly emissive in the binding state. The lowest<br />

binding energies of the conformations in clusters 1 and 2 are -8.60<br />

and -7.54 kcal/mol, corresponding to estimated inhibition constants<br />

(K i) of 0.49 and 2.98 µM, respectively. These small Gibbs<br />

free energies indicate that the interaction between BSPOTPE<br />

and HSA is highly spontaneous and energetically favorable.<br />

The detailed binding modes of the conformations from the<br />

other clusters in the hydrophobic cavity closely resemble those<br />

of clusters 1 and 2, although they are grouped into different<br />

clusters. This is not unexpected because the ligand is treated as<br />

fully flexible in the computational modeling process, and a small<br />

rmsd tolerance is used for clustering. Thus, any difference in the<br />

coordinates of an atom between two conformations will increase<br />

the rmsd value, even though the atom is not important for binding<br />

interactions.<br />

HSA is a physiological carrier in the human body, and its<br />

hydrophobic regions are capable of binding insoluble endogenous<br />

compounds like fatty acids and drugs. With close scrutinization<br />

of other proteins, such as trypsin, pepsin, and papain, one can<br />

hardly find a hydrophobic pocket as in HSA. This may explain<br />

why the BSPOTPE luminogen shows excellent selectivity toward<br />

the albumin proteins over the other types of proteins.<br />

GndHCl is known to work as a bulky ionic cosolvent. It not<br />

only weakens hydrophobic interaction but also interferes with<br />

association between charged solutes. 39 When HSA unfolds, a large<br />

number of nonpolar side chains are exposed to the aqueous<br />

medium and stabilized by GndHCl. GndHCl may also intrude into<br />

the hydrophobic pocket of HSA and break down the spatial<br />

architecture of the protein folds. CD spectral data indicate that<br />

no significant alternation in the secondary structure occurs during<br />

the first unfolding transition in the low concentration range of<br />

GndHCl (6.0 M). No secondary<br />

structure or strand helicity is retained in this region, as confirmed<br />

by the CD spectral data. The luminogen molecules are fully<br />

released from the protein, and no emission signals can be<br />

collected. The unfolding process discussed above is summarized<br />

in Scheme 1. Upon removal of the GndHCl denaturant by dilution,<br />

the protein returns to its native form and the emission of<br />

BSPOTPE is recovered, indicating that the interaction between<br />

HSA and BSPOTPE is fully reversible.<br />

CONCLUDING REMARKS<br />

In summary, in this work, we developed an environmentally<br />

stable and synthetically readily accessible FL probe for HSA<br />

detection and quantitation. The nonluminescent BSPOTPE becomes<br />

emissive in the presence of HSA. The AIE bioprobe shows<br />

a linear calibration curve at [HSA] ) 0-100 nM, enabling the<br />

protein quantitation over a wide concentration range. It enjoys a<br />

low detection limit (down to 1 nM) and a high selectively toward<br />

albumins. The FL bioassay is tolerant of the species in the artificial<br />

urine and is promising for applications in real-life urinary protein<br />

detection. Design and fabrication of simple bioassay kits for clinical<br />

diagnostics are ongoing in our laboratories.<br />

BSPOTPE is successfully employed as a protein staining<br />

reagent in the PAGE analysis. It offers a simple, rapid, effective,<br />

and economic way for visualization of protein bands in the gel<br />

assay. Studies on the interaction between BSPOTPE and HSA


Scheme 1. Proposed Mechanism for Fluorescent Probing of GndHCl-Induced HSA Unfolding Process a<br />

a BSPOTPE and HSA are denoted by stars and springs, respectively.<br />

reveal the essential role of the hydrophobic cavities of the protein<br />

folding structure. The change in the FL intensity reflects the global<br />

stability of the protein. The unfolding process of HSA induced by<br />

GndHCl monitored by the AIE luminogen reveals a three-step<br />

transition with the molten-globule intermediate involved. Studies<br />

of other biologically important processes by AIE luminogens are<br />

in progress in our laboratories.<br />

ACKNOWLEDGMENT<br />

This work reported in this paper was partially supported by<br />

the Nano Science and Technology Program at HKUST, the<br />

Research Grants Council of Hong Kong (603008 and 602706), the<br />

University Grants committee of Hong Kong (AoE/P-03/08), and<br />

the National Science Foundation of China (20974028). B.Z.T.<br />

thanks the support from Cao Guangbiao Foundation of Zhejiang<br />

University.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Text describing the procedures for artificial urine preparation,<br />

cytotoxicity assay, FRET measurement, DSC analysis, and computational<br />

modeling. Figures showing viability of the HeLa cells<br />

incubated in the presence of BSPOTPE, spectral overlap between<br />

BSPOTPE absorption and HSA emission, FL spectra of BSPOTPE<br />

and DSC thermograms of HSA in the presence of GndHCl, FL<br />

intensities of BSPOTPE at 470 nm in the presence and absence<br />

of HSA at different pH, and summary of the docked conformations<br />

of BSPOTPE clustered with a rmsd tolerance of 2.0 Å. This material<br />

is available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review July 7, 2010. Accepted July 7, 2010.<br />

AC1018028<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7043


Anal. Chem. 2010, 82, 7044–7048<br />

Direct Fluorescence Polarization Assay for the<br />

Detection of Glycopeptide Antibiotics<br />

Linliang Yu, Meng Zhong, and Yinan Wei*<br />

Department of <strong>Chemistry</strong>, University of Kentucky, Lexington, Kentucky 40506<br />

Glycopeptide antibiotics are widely used in the treatment<br />

of infections caused by Gram-positive bacteria. They<br />

inhibit the biosynthesis of the bacterial cell wall through<br />

binding to the D-alanyl-D-alanine (D-Ala-D-Ala) terminal<br />

peptide of the peptidoglycan precursor. Taking advantage<br />

of this highly specific interaction, we developed a direct<br />

fluorescence polarization based method for the detection<br />

of glycopeptide antibiotics. Briefly, we labeled the acetylated<br />

tripeptide Ac-L-Lys-D-Ala-D-Ala-OH with a fluorophore<br />

to create a peptide probe. Using three glycopeptide<br />

antibiotics, vancomycin, teicoplanin, and telavancin, as<br />

model compounds, we demonstrated that the fluorescence<br />

polarization of the peptide probe increased upon<br />

binding to antibiotics in a concentration dependent manner.<br />

The dissociation constants (Kd) between the peptide<br />

probes and the antibiotics were consistent with those<br />

reported between free D-Ala-D-Ala and the antibiotics<br />

in the literature. The assay is highly reproducible and<br />

selective toward glycopeptide antibiotics. Its detection<br />

limit and work concentration range are 0.5 µM and<br />

0.5-4 µM for vancomycin, 0.25 µM and 0.25-2 µM<br />

for teicoplanin, and 1 µM and 1-8 µM for telavancin.<br />

Furthermore, we compared our assay in parallel with<br />

a commercial fluorescence polarization immunoassay<br />

(FPIA) kit in detecting teicoplanin spiked in human<br />

blood samples. The accuracy and precision of the two<br />

methods are comparable. We expect our assay to be<br />

useful in both research and clinical laboratories.<br />

Glycopeptide antibiotics are cyclic peptides that bind to the<br />

peptidoglycan precursor D-alanyl-D-alanine to inhibit the biosynthesis<br />

of bacterial cell walls. They are widely used to treat<br />

infections caused by Gram-positive bacteria, including Methicillin<br />

Resistant Staphylococcus aureus (MRSA), and are regarded as the<br />

drug of “last resort”. 1 Vancomycin, teicoplanin, and telavancin are<br />

three representative glycopeptide antibiotics that are currently in<br />

clinical use. 2-4 Monitoring of drug levels in patients’ serum during<br />

* To whom correspondence should be addressed. Address: 305 <strong>Chemistry</strong>-<br />

Physics Building, University of Kentucky, Lexington, KY 40506-0055. Telephone:<br />

(859) 257-7085. Fax: (859) 323-1069. E-mail: yinan.wei@uky.edu.<br />

(1) Perkins, H. R. Pharmacol. Ther. 1982, 16, 181–197.<br />

(2) Levine, D. P. Clin. Infect. Dis. 2006, 42, S5–S12.<br />

(3) Delalla, F.; Nicolin, R.; Rinaldi, E.; Scarpellini, P.; Rigoli, R.; Manfrin, V.;<br />

Tramarin, A. Antimicrob. Agents Chemother. 1992, 36, 2192–2196.<br />

(4) Higgins, D. L.; Chang, R.; Debabov, D. V.; Leung, J.; Wu, T.; Krause, K. M.;<br />

Sandvik, E.; Hubbard, J. M.; Kaniga, K.; Schmidt, D. E.; Gao, Q.; Cass,<br />

R. T.; Karr, D. E.; Benton, B. M.; Humphrey, P. P. Antimicrob. Agents<br />

Chemother. 2005, 49, 1127–1134.<br />

7044 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

the treatment is critical in maximizing the effectiveness of the<br />

treatment while minimizing drug toxicities and side effects. 5-7<br />

Such data also provide important pharmacokinetic and pharmacodynamic<br />

parameters for clinical studies 8-12 and, thus, guide the<br />

optimization of therapies. 7,13,14 In addition, vancomycin has been<br />

extensively used as a model antibiotic in tissue engineering and<br />

controlled drug release studies. 15-20 Various assays have been<br />

developed to measure the concentration of glycopeptide antibiotics,includingmethodsbasedonUVabsorbance,<br />

17 immunoassay, 21-24<br />

high performance liquid chromatography (HPLC), 25-28 and agar<br />

diffusion bioassay. 29,30 Each of these existing methods suffer from<br />

(5) James, C. W.; Gurk-Turner, C. Proc. (Bayl. Univ. Med. Cent.) 2001, 14,<br />

189–190.<br />

(6) Williams, A. H.; Gruneberg, R. N. J. Antimicrob. Chemother. 1984, 14, 441–<br />

445.<br />

(7) Brink, A. J.; Richards, G. A.; Cummins, R. R.; Lambson, J.; Teicoplanin,<br />

G. U. Int. J. Antimicrob. Agents 2008, 32, 455–458.<br />

(8) Svetitsky, S.; Leibovici, L.; Paul, M. Antimicrob. Agents Chemother. 2009,<br />

53, 4069–4079.<br />

(9) Wood, M. J. J. Antimicrob. Chemother. 1997, 40, 147–147.<br />

(10) Cobo, J.; Fortun, J. J. Antimicrob. Chemother. 1996, 38, 1113–1114.<br />

(11) Wood, M. J. J. Antimicrob. Chemother. 1996, 38, 919–919.<br />

(12) Wood, M. J. J. Antimicrob. Chemother. 1996, 37, 209–222.<br />

(13) Pea, F.; Brollo, L.; Viale, P.; Pavan, F.; Furlanut, M. J. Antimicrob. Chemother.<br />

2003, 51, 971–975.<br />

(14) Mohammedi, I.; Descloux, E.; Argaud, L.; Le Scanff, J.; Robert, D. Int. J.<br />

Antimicrob. Agents 2006, 27, 259–262.<br />

(15) Adams, C. S.; Antoci, V., Jr.; Harrison, G.; Patal, P.; Freeman, T. A.; Shapiro,<br />

I. M.; Parvizi, J.; Hickok, N. J.; Radin, S.; Ducheyne, P. J. Orthop. Res. 2009,<br />

27, 701–709.<br />

(16) Antoci, V., Jr.; King, S. B.; Jose, B.; Parvizi, J.; Zeiger, A. R.; Wickstrom, E.;<br />

Freeman, T. A.; Composto, R. J.; Ducheyne, P.; Shapiro, I. M.; Hickok, N. J.;<br />

Adams, C. S. J. Orthop. Res. 2007, 25, 858–866.<br />

(17) Radin, S.; Chen, T.; Ducheyne, P. Biomaterials 2009, 30, 850–858.<br />

(18) Radin, S.; Ducheyne, P.; Kamplain, T.; Tan, B. H. J. Biomed. Mater. Res.<br />

2001, 57, 313–320.<br />

(19) Perelman, L. A.; Pacholski, C.; Li, Y. Y.; VanNieuwenhz, M. S.; Sailor, M. J.<br />

Nanomedicine 2008, 3, 31–43.<br />

(20) Lai, C. Y.; Trewyn, B. G.; Jeftinija, D. M.; Jeftinija, K.; Xu, S.; Jeftinija, S.;<br />

Lin, V. S. Y. J. Am. Chem. Soc. 2003, 125, 4451–4459.<br />

(21) Kitzis, M. D.; Goldstein, F. W. Clin. Microbiol. Infect. 2006, 12, 92–95.<br />

(22) Wilson, J. F.; Davis, A. C.; Tobin, C. M. J. Antimicrob. Chemother. 2003,<br />

52, 78–82.<br />

(23) Lee, H. B.; Kwak, B. Y.; Lee, J. C.; Kim, C. J.; Shon, D. H. J. Microbiol.<br />

Biotechnol. 2004, 14, 612–619.<br />

(24) Lam, M. T.; Le, X. C. Analyst 2002, 127, 1633–1637.<br />

(25) Valle, M. J. D.; Lopez, F. G.; Navarro, A. S. J. Pharm. Biomed. Anal. 2008,<br />

48, 835–839.<br />

(26) Abu-Shandi, K. H. Anal. Bioanal. Chem. 2009, 395, 527–532.<br />

(27) Saito, M.; Santa, T.; Tsunoda, M.; Hamamoto, H.; Usui, N. Biomed.<br />

Chromatogr. 2004, 18, 735–738.<br />

(28) Shen, J.; Jiao, Z.; Zhou, Y. N.; Zhu, H. L.; Song, Z. J. Chromatographia 2007,<br />

65, 9–12.<br />

(29) Kureishi, A.; Jewesson, P. J.; Bartlett, K. H.; Cole, C. D.; Chow, A. W.<br />

Antimicrob. Agents Chemother. 1990, 34, 1642–1647.<br />

(30) Bantar, C.; Durlach, R.; Nicola, F.; Freuler, C.; Bonvehi, P.; Vazquez, R.;<br />

Smayevsky, J. J. Antimicrob. Chemother. 1999, 43, 737–740.<br />

10.1021/ac100543e © 2010 American <strong>Chemical</strong> Society<br />

Published on Web 07/20/2010


one or more drawbacks including poor selectivity, requirement<br />

of large amount of samples, and involvement of cumbersome<br />

pretreatment processes. Fluorescence polarization immunoassay<br />

(FPIA) is the most popular method currently used in the<br />

quantification of glycopeptide antibiotics in clinical samples. 23,31<br />

For example, in the FPIA assay for vancomycin, fluorescentlabeled<br />

vancomycin forms a complex with antivancomycin antibody.<br />

Free vancomycin in test samples competitively binds to the<br />

antibody to displace the labeled one, decreases the fluorescence<br />

polarization signal, and thus allows the quantification of the<br />

analyte. 32 In this study, we developed a direct fluorescence<br />

polarization based assay, taking advantage of the specific interaction<br />

between glycopeptide antibiotics and their therapeutic target,<br />

the dipeptide D-alanyl-D-alanine (D-Ala-D-Ala). Briefly, we labeled<br />

acetylated peptide Ac-L-Lys-D-Ala-D-Ala-OH with a fluorophore and<br />

monitored the increase of fluorescence polarization of the labeled<br />

peptide in the presence of drugs. Our method is a class-specific<br />

assay, which can be used to detect glycopeptide antibiotics in<br />

general. In addition, since no protein (such as the antivancomycin<br />

antibody) is used in the assay, we expect the shelf life and storage<br />

stability of our assay to greatly exceed those of the immunoassay.<br />

In this study, we have characterized the interaction between<br />

the fluorescent peptide probes and three representative glycopeptide<br />

antibiotics, vancomycin, teicoplanin, and telavancin. We<br />

were able to measure the concentration of these drugs in serum<br />

and blood samples without complicated pretreatment. In addition,<br />

we compared the performance of our assay with a commercial<br />

FPIA kit in determining the concentration of teicoplanin in human<br />

blood samples. We expect our assay to be useful in both research<br />

and clinical applications.<br />

EXPERIMENTAL SECTION<br />

<strong>Chemical</strong>s. The acetylated Ac-L-Lys-D-Ala-D-Ala-OH peptide,<br />

teicoplanin, fluorescein isothiocyanate (FITC), HPLC grade acetonitrile,<br />

and heat inactivated, sterile-filtered fetal bovine serum<br />

(FBS) were from Sigma-Aldrich (St. Louis, MO). Vancomycin<br />

hydrochloride was from Duchefa (Haarlem, The Netherlands).<br />

Telavancin was from Astellas (Deerfield, IL). Alexa Fluor 680maleimide<br />

(AF680) was from Invitrogen (Eugene, OR). Teicoplanin<br />

FPIA kit was from LABfx (Portland, OR). Human whole blood<br />

was from Bioreclamation (Westbury, NY). All other reagents were<br />

of analytical grade and purchased from Sigma-Aldrich.<br />

Fluorescent Labeling of the Peptide. The acetylated peptide<br />

Ac-L-Lys-D-Ala-D-Ala-OH (3 mM) was incubated with fluorescein<br />

isothiocyanate (FITC, 4.5 mM) in a phosphate buffer (PBS) (10<br />

mM Na2HPO4,2mMKH2PO4, 2.7 mM KCl, 137 mM NaCl, pH<br />

7.4) at room temperature overnight. The reaction was terminated<br />

through the addition of a Tris-Cl buffer (5 mM, pH 7.4)<br />

followed by an incubated of2hatroom temperature. The FITClabeled<br />

peptide was purified using a reverse-phase HPLC<br />

(Waters, Milford, MA) on a C18 column (Waters, Milford, MA).<br />

The purified FITC-labeled peptide was dried under vacuum and<br />

then resuspended in water. The concentration of the peptide<br />

was determined based on the absorption at 515 nm. Labeling<br />

with AF680 was performed similarly. The peptide probe<br />

concentration was determined by absorption at 680 nm. The<br />

molecule weight of the labeled peptides was confirmed by mass<br />

spectroscopy analysis.<br />

Fluorescence Polarization Measurements. Fluorescence<br />

polarization measurements were performed using a Perkin-Elmer<br />

LS-55 fluorescence spectrometer (Perkin-Elmer, Waltham, MA).<br />

The temperature was kept at 20 °C. Fluorophore-conjugated<br />

peptide (FITC-peptide or AF680-peptide) was added into 400 µL<br />

of PBS or FBS to a final concentration of 1 µM. The excitation<br />

and emission wavelengths were 479 and 515 nm for FITC-peptide<br />

and 679 and 702 nm for AF680-peptide.<br />

Titration and Data Fitting. For titration experiments, FITCpeptide<br />

or AF680-peptide was added into 400 µL PBS or FBS to<br />

a final concentration of 1 µM, and then, small aliquots of<br />

glycopeptide antibiotics were titrated into the samples. The<br />

increases of fluorescence polarization upon the formation of<br />

drug-peptide complexes were recorded.<br />

The titration curve was fitted using a one-to-one binding model<br />

according to the following equation. 33<br />

∆P ) ∆Pmax( ([F] + x + Kd ) - √([F] + x + Kd ) 2 - 4[F]x)<br />

2[F]<br />

(1)<br />

Where ∆P is the monitored change of fluorescence polarization<br />

at each drug concentration, ∆Pmax is the maximum change of<br />

the fluorescence polarization upon saturation, [F] is the peptide<br />

probe concentration, Kd is the dissociation constant between<br />

the peptide probe and the drug, and x is the drug concentration.<br />

The titration data were fitted using Origin (Northampton, MA).<br />

Binding Selectivity Assay. For each antibiotic, AF680-peptide<br />

was added into 400 µL of FBS to a final concentration of 1 µM.<br />

Next, antibiotics were added to a final concentration of 8 µM. The<br />

change of fluorescent polarization upon the addition of antibiotics<br />

was recorded.<br />

Determination of Drug Concentration in Test Samples.<br />

For tests in serum, AF680-peptide was added to 400 µL of FBS to<br />

a final concentration of 1 µM. Then, known quantities of antibiotics<br />

were added into the sample. The concentrations of the drugs were<br />

determined based on the calibration curves derived in FBS. For<br />

tests in blood, three human blood samples were spiked with<br />

teicoplanin in the therapeutic range by a person not involved in<br />

this project. The exact values were kept blind to the authors during<br />

the analysis. The blood samples were centrifuged at 2000g for 10<br />

min. The supernatant (plasma) was collected and analyzed<br />

together with a series of teicoplanin standards in FBS as described<br />

above. For each plasma sample, we prepared two sets of dilutions<br />

so the final concentration of teicoplanin would be in the working<br />

concentration range of the assay. The first set of three test samples<br />

contain 10 µL of plasma and 390 µL of FBS with 1 µM of AF680peptide.<br />

The second set of three test samples contain 20 µL of<br />

plasma and 380 µL of FBS with 1 µM of AF680-peptide. The<br />

teicoplanin concentration in each set of samples was determined.<br />

The original teicoplanin concentration calculated from the set of<br />

samples falls in the working concentration range. The same<br />

(31) Urakami, T.; Maiguma, T.; Kaji, H.; Kondo, S.; Teshima, D. J. Clin. Pharm.<br />

Ther. 2008, 33, 357–363.<br />

(32) Adamczyk, M.; Grote, J.; Moore, J. A.; Rege, S. D.; Yu, Z. G. Bioconjugate<br />

Chem. 1999, 10, 176–185. (33) Yu, L. L.; Fang, J.; Wei, Y. N. Biochemistry 2009, 48, 2099–2108.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7045


plasma samples were analyzed using the teicoplanin FPIA kit<br />

following the instruction of the manufacturer.<br />

RESULT AND DISCUSSION<br />

Design and Synthesis of Peptide Probes. We used aminereactive<br />

fluorophores to label the acetylated peptide Ac-L-Lys-D-<br />

Ala-D-Ala-OH. We chose this peptide for two reasons. First, L-lys<br />

is naturally occurring in the peptidoglycan precursor peptide L-Ala-<br />

D-Glu-L-Lys-D-Ala-D-Ala. We expected the incorporation of the L-lys<br />

would not affect the interaction between the peptide and the<br />

antibiotics. Second, the side chain of Lys contains a primary amine<br />

group, which is distant from the binding site judged by the crystal<br />

structure of the vancomycin and diacetyl-L-Lys-D-Ala-D-Ala complex.<br />

34 As expected, we found that the binding affinities between<br />

the labeled peptide and antibiotics were similar to the values<br />

reported for the binding between the free peptide and antibiotics,<br />

indicating that the modification did not affect their interaction (see<br />

below).<br />

Effect of Fluorophores. The fluorescence polarization signal<br />

is determined by the Perrin Equation. 35<br />

( 1 1<br />

-<br />

P 3) ) ( 1<br />

-<br />

P0 1 kT<br />

1 + 3)( ηV τ)<br />

Where τ is the fluorescence lifetime of the fluorophore, P0 is the<br />

limiting polarization, k is the Boltzmann constant, T is the<br />

absolute temperature, η is the viscosity, and V is the molecular<br />

volume (molecular weight). The fluorescence polarization is<br />

negatively correlated with the lifetime of the fluorophores and<br />

positively correlated with the molecular volume (molecular<br />

weight). In order to maximize the observed change of polarization<br />

upon antibiotic binding, we tested two fluorophores with<br />

different lifetimes, FITC (4 ns lifetime) and AF680 (1 ns<br />

lifetime) (Figure 1). We found that AF680-peptide yielded a larger<br />

change of signal upon drug binding and, thus, exhibited a better<br />

signal-to-noise ratio.<br />

Titration Curves of Glycopeptide Antibiotics in Phosphate<br />

Buffer. The binding between the antibiotics and peptide probes<br />

was monitored following the increase of fluorescence polarization<br />

of the samples upon addition of increasing concentrations of the<br />

antibiotics (Figure 1). The dissociation constants (Kd) between<br />

the drugs and peptide probes were estimated to be around 1.1,<br />

0.2, and 1.8 µM, respectively, for vancomycin, teicoplanin, and<br />

telavancin (Table 1). The identity of the fluorophores did not have<br />

a drastic effect on the binding affinity, suggesting they were not<br />

directly involved in the interaction. The determined affinities are<br />

similar to the values reported for the bindings between these<br />

antibiotics and diacetyl-L-Lys-D-Ala-D-Ala (∼1 µM for vancomycin,<br />

and ∼0.6 µM for teicoplanin). 36-39 Compared with the FITC-<br />

(34) Nitanai, Y.; Kikuchi, T.; Kakoi, K.; Hanmaki, S.; Fujisawa, I.; Aoki, K. J.<br />

Mol. Biol. 2009, 385, 1422–1432.<br />

(35) Ye, B. C.; Ikebukuro, K.; Karube, I. Nucleic Acids Res. 1998, 26, 3614–<br />

3615.<br />

(36) Popieniek, P. H.; Pratt, R. F. Anal. Biochem. 1987, 165, 108–113.<br />

(37) Azad, M.; Hernandez, L.; Plazas, A.; Rudolph, M.; Gomez, F. A. Chromatographia<br />

2003, 57, 339–343.<br />

(38) Rao, J. H.; Yan, L.; Xu, B.; Whitesides, G. M. J. Am. Chem. Soc. 1999, 121,<br />

2629–2630.<br />

(39) van Wageningen, A. M. A.; Staroske, T.; Williams, D. H. Chem. Commun.<br />

1998, 1171–1172.<br />

7046 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

(2)<br />

Figure 1. Calibration curves of FITC-peptide (diamonds) and AF680peptide<br />

(squares) with vancomycin (A), teicoplanin (B), and telavancin<br />

(C) in PBS. The concentration of the peptide probes were 1 µM. The<br />

concentrations of the glycopeptide antibiotics increased from 0 to 7.7<br />

µM during the titration. The data were fitted using a one-to-one binding<br />

model to derive the dissociation constants listed in Table 1.<br />

Table 1. Fitting Parameters of Peptide-Drug<br />

Interaction<br />

glycopeptide<br />

antibiotics probe matrix<br />

dissociation<br />

constant Kd (µM)<br />

vancomycin FITC-peptide PBS 0.91 ± 0.06<br />

AF680-peptide PBS 1.07 ± 0.19<br />

AF680-peptide FBS 1.81 ± 0.13<br />

teicoplanin FITC-peptide PBS 0.14 ± 0.07<br />

AF680-peptide PBS 0.22 ± 0.02<br />

AF680-peptide FBS 0.87 ± 0.08<br />

telavancin FITC-peptide PBS 3.22 ± 0.38<br />

AF680-peptide PBS 1.81 ± 0.36<br />

AF680-peptide FBS 5.36 ± 0.35<br />

peptide probe, the AF680-peptide yielded a better signal-to-noise<br />

ratio. Therefore, it was used in fluorescence polarization assays<br />

for the rest of the study.<br />

To directly examine if the free fluorophores (FITC or AF680)<br />

interacted with the glycopeptide antibiotics, we compared the<br />

change of fluorescence polarization of 1 µM FITC, AF680, FITCpeptide,<br />

or AF680-peptide upon addition of 8 µM vancomycin,<br />

teicoplanin, or telavancin (Figure 2). No significant change of<br />

fluorescence polarization was observed for samples containing free<br />

fluorophores, indicating the lack of interaction between the drugs<br />

and the free fluorophores.<br />

Drug Detection in Serum. To further examine the usefulness<br />

of the current method in the detection of glycopeptide antibiotics<br />

in clinical samples, we measured the titration curves using AF680peptide<br />

for the detection of vancomycin, teicoplanin, and telavancin<br />

in FBS. The potential matrix effects from serum could affect the<br />

assay in two ways. First, the higher viscosity in serum slows down<br />

the tumbling of the fluorophores and, thus, increases fluorescence<br />

polarization. Second, certain species in the serum may interact<br />

with the peptide probe or the drugs to prevent accurate detection.


Figure 2. Free fluorophores do not interact with the antibiotics. The<br />

fluorescence polarization did not change when vancomycin (black),<br />

teicoplanin (white), or telavancin (gray) was added into PBS containing<br />

the free fluorophores (FITC or AF680). FITC and AF680 were<br />

first incubated with a Tris-Cl buffer (5 mM, pH 7.4) to block the amine<br />

reactive groups. The concentration of fluorophores or peptide probes<br />

was 1 µM. The concentration of the drugs was 8 µM.<br />

Figure 3. (A) Titration curves for vancomycin (squares), teicoplanin<br />

(triangles), and telavancin (diamonds) in FBS. AF680-peptide (1 µM)<br />

was used in the assays. (B) Intraday and interday variance of the<br />

assay. Three calibration curves of teicoplanin were collected at three<br />

different times during a day: 9:00 a.m. (filled squares), 1:00 pm (open<br />

diamonds), and 5:00 pm (open circles). A fourth curve was obtained<br />

at 9:00 a.m. of the following day (filled triangle). The same teicoplanin<br />

calibration curve as shown in Figure 3A was also plotted (gray) to<br />

illustrate the long-term reproducibility of the assay.<br />

The fluorescence polarization signals measured in serum increased<br />

both for the free peptide probes and the drug-peptide<br />

complex (Supporting Information). The increase was more dramatic<br />

for the complex, which resulted in an increase in the ∆P<br />

values compared to the ∆P values measured in PBS (Figure 3A).<br />

The binding affinities decreased slightly compared to the affinities<br />

measured in PBS (Table 1). We speculated this change of affinity<br />

was due to the interaction between certain components in the<br />

serum with the peptide probes and/or glycopeptides antibiotics,<br />

which weakened their affinities for each other through a competition<br />

mechanism.<br />

Furthermore, we examined the interday and intraday variances<br />

of the method by collecting the calibration curves of teicoplanin<br />

in serum at three different times of one day and again in the<br />

following day (Figure 3B). The curves superimposed well onto<br />

each other, indicating the assay was reproducible. At the same<br />

teicoplanin concentration, the relative variance among all four<br />

curves was less than 5%. In addition, we plotted the teicoplanin<br />

calibration curve from Figure 3A in Figure 3B again to illustrate<br />

the good reproducibility of the method over a period of 6 months,<br />

Figure 4. Selectivity of the assay. Different antibiotics (8 µM) were<br />

added into solutions containing AF680-peptide (1 µM). Antibiotics<br />

tested were ampicillin (1), gentamicin (2), tetracycline (3), nalidixic<br />

acid (4), chloramphenicol(5), vancomycin (6), telavancin (7), and<br />

teicoplanin (8). Error bars indicate the standard deviations from three<br />

independent measurements.<br />

as the trace in Figure 3A was collected approximately 6 months<br />

ago. As all binding based assays, the detection limit and working<br />

concentration range are determined by the binding affinity<br />

between the probe and the analyte, as well as the signal-to-noise<br />

ratios of the measurements. The detection limit and working<br />

concentration range of the current method were determined to<br />

be 0.5 µM and 0.5-4 µM for vancomycin, 0.25 µM, 0.25-2 µM<br />

for teicoplanin, and 1 µM and 1-8 µM for telavancin.<br />

Selectivity Study. In many situations, several antibiotics may<br />

be used together as a cocktail in research or clinical applications.<br />

It is desirable to have a method that can selectively detect a<br />

specific one or specific class of antibiotics. The interaction between<br />

the peptide probe and glycopeptide antibiotics is highly specific.<br />

We examined the binding selectivity of the AF680-peptide probe<br />

toward different antibiotics in FBS (Figure 4). The assay was<br />

highly selective toward the three glycopeptides. No significant<br />

change of fluorescence polarization was observed for other<br />

antibiotics tested.<br />

Recovery Test in Human Blood. Finally, we compared the<br />

performance of the current assay with a commercial FPIA kit in<br />

the quantification of teicoplanin in human blood. Human blood<br />

samples were spiked with different concentrations of teicoplanin<br />

based on the drug’s therapeutic concentration window. Measurement<br />

was conducted blindly using the current assay side by side<br />

with a commercial FPIA teicoplanin detection kit. For the current<br />

assay, each spiked blood sample was centrifuged to remove the<br />

blood cells. The supernatant (plasma) was collected and diluted<br />

to the working concentration range. Since the therapeutic window<br />

of teicoplanin is approximately 10-30 µg/mL, 40-42 while our<br />

working concentration range is between 0.25 and 2 µM (corresponding<br />

to 0.47-3.8 µg/mL), we diluted each plasma sample by<br />

FBS to two scales: 20-fold and 40-fold. Thus, at least one of the<br />

diluted final concentrations would fall in the working concentration<br />

range. On the basis of the calibration curve established in FBS,<br />

the antibiotic concentrations in the diluted serum solution were<br />

determined, which were then multiplied by the fold of dilution to<br />

(40) Soy, D.; Lopez, E.; Ribas, J. Ther. Drug Monit. 2006, 28, 737–743.<br />

(41) Lortholary, O.; Tod, M.; Rizzo, N. Antimicrob. Agents Chemother. 1996,<br />

40, 1242–1247.<br />

(42) Lamont, E.; Seaton, R. A.; Macpherson, M.; Semple, L.; Bell, E.; Thomson,<br />

A. H. J. Antimicrob. Chemother. 2009, 64, 181–187.<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7047


Table 2. Detection of Teicoplanin in Human Blood<br />

Using the Current Assay or a Commercial FPIA Kit<br />

spiked<br />

(µg/mL) this method recovery (%) FPIA recovery (%)<br />

9.5 9.2 ± 0.2 97 9.8 ± 0.6 103<br />

19 19.6 ± 0.4 103 18.5 ± 0.6 97<br />

38 39.0 ± 1.0 103 37.8 ± 4.4 99<br />

obtain the original drug concentration. Meanwhile, each plasma<br />

sample was also analyzed using a commercial FPIA kit. The spiked<br />

and determined concentrations were summarized in Table 2.<br />

When human blood samples from different donors were used as<br />

the background matrix for spiking, the variation between the<br />

determined drug concentrations was similar to the deviation<br />

among multiple tests performed using blood samples from the<br />

same donor. The accuracy and precision of the current assay is<br />

comparable to the commercial kit.<br />

In conclusion, we developed a simple analytical assay for the<br />

detection and quantification of selected glycopeptides antibiotics,<br />

including vancomycin, teicoplanin, and telavancin, in various<br />

samples. The method is highly selective toward glycopeptide<br />

7048 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

antibiotics and does not need complicated pretreatment for serum<br />

and blood samples. In addition, no antibody is used in our method;<br />

therefore, we expect the shelf life and storage stability of our assay<br />

to greatly exceed those of the commercial FPIA kits. We have<br />

stored the AF680-peptide probe at room temperature in the dark<br />

for 9 months. No significant loss of signal intensity or binding<br />

affinity to antibiotics could be detected.<br />

ACKNOWLEDGMENT<br />

This work was supported by a RCTF fellowship and the faculty<br />

startup fund from University of Kentucky. The authors thank Mr.<br />

Raymond Miracle from LABfx LLC for suggestions on the<br />

teicoplanin analysis using the FPIA kit.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review February 28, 2010. Accepted July 7,<br />

2010.<br />

AC100543E


Anal. Chem. 2010, 82, 7049–7052<br />

Development of a High Sensitivity Rapid Sandwich<br />

ELISA Procedure and Its Comparison with the<br />

Conventional Approach<br />

Chandra Kumar Dixit, †,‡ Sandeep Kumar Vashist, †,|,# Feidhlim T. O’Neill, † Brian O’Reilly, †<br />

Brian D. MacCraith, †,§ and Richard O’Kennedy* ,†,‡,§<br />

Centre for Bioanalytical Sciences (CBAS), National Centre for Sensor Research, Applied Biochemistry Group, School<br />

of Biotechnology, and Biomedical Diagnostics Institute (BDI), Dublin City University, Dublin 9, Ireland, and<br />

Bristol-Myers Squibb (BMS), Swords Laboratories, Watery Lane, Swords, Co. Dublin, Ireland<br />

A highly sensitive and rapid sandwich enzyme-linked<br />

immunosorbent assay (ELISA) procedure was developed<br />

for the detection of human fetuin A/AHSG (r2-HSglycoprotein),<br />

a specific biomarker for hepatocellular<br />

carcinoma and atherosclerosis. Anti-human fetuin A<br />

antibody was immobilized on aminopropyltriethoxysilanemediated<br />

amine-functionalized microtiter plates using<br />

1-ethyl-3-[3-dimethylaminopropyl]carbodiimide hydrochloride<br />

and N-hydroxysulfosuccinimide-based heterobifunctional<br />

cross-linking. The analytical sensitivity of the<br />

developed assay was 39 pg/mL, compared to 625 pg/<br />

mL for the conventional assay. The generic nature of the<br />

developed procedure was demonstrated by performing<br />

human fetuin A assays on different polymeric matrixes,<br />

i.e., polystyrene, poly(methyl methacrylate), and polycyclo-olefin<br />

(Zeonex), in a modified microtiter plate format.<br />

Thus, the newly developed procedure has considerable<br />

advantages over the existing method.<br />

Conventional enzyme-linked immunosorbent assay (ELISA)<br />

procedures have been followed for decades for the detection of<br />

analytes of importance in industrial, healthcare, and academic<br />

research. However, improvement on existing ELISA technologies<br />

are continuously attempted by many groups. 1,2 We report the<br />

development of a high-sensitivity ELISA-based assay for human<br />

fetuin A (HFA), with lower detection limits and a higher sensitivity<br />

than commercially available assays. The biological importance of<br />

HFA, a member of the cystatin superfamily, which is commonly<br />

present in the cortical plate of the immature cerebral cortex and<br />

hemopoietic matrix of bone marrow, is discussed elsewhere. 3-10<br />

There are many commercially available ELISA kits for fetuin A,<br />

* To whom correspondence should be addressed. Email:<br />

richard.okennedy@dcu.ie. Tel.: +353 1 700 7810. Fax: +353 1 700 5412.<br />

† Centre for Bioanalytical Sciences (CBAS), National Centre for Sensor<br />

Research, Dublin City University.<br />

‡ Applied Biochemistry Group, School of Biotechnology, Dublin City University.<br />

§ Biomedical Diagnostics Institute (BDI), Dublin City University.<br />

| Bristol-Myers Squibb (BMS).<br />

# Current Address: Nanoscience and Nanotechnology Initiative (NUSNNI),<br />

National University of Singapore, Engineering Drive 1, Singapore 117576.<br />

(1) Kaur, J.; Boro, R. C.; Wangoo, N.; Singh, R. K.; Suri, C. R. Anal. Chim.<br />

Acta 2008, 607 (1), 92–99.<br />

(2) Jia, C. P.; Zhong, H. Q.; Liu, M. Y.; Jing, F. X.; Yao, S. H.; Xiang, J. Q.; Jin,<br />

Q. H.; Zhao, J. L. Biosens. Bioelectron. 2009, 24 (9), 2836–2841.<br />

which are listed in Supplementary Table 1, Supporting Information.<br />

The assay was demonstrated on different commercially<br />

relevant solid supports. The developed ELISA is better than the<br />

conventional procedure in terms of greatly reduced overall assay<br />

duration, higher sensitivity, and greater reproducibility.<br />

HFA was taken as the model assay system to demonstrate the<br />

utility of our developed ELISA procedure since all the assay<br />

components were commercially available in kit form. This enabled<br />

us to do robust and highly precise comparison of the developed<br />

ELISA with the commercially existing conventional ELISA procedures,<br />

as the same assay components were used under the same<br />

conditions. The developed procedure can be employed on any<br />

commercially relevant substrate. Therefore, this approach of<br />

immobilizing antibody on chemically modified solid supports<br />

(Figure 1) has potential applications in many other assays and<br />

formats.<br />

EXPERIMENTAL SECTION<br />

Plate Preparation, Amine-Functionalization with Silane,<br />

and Cross-Linking Carboxyl Groups of Anti-HFA Antibody<br />

to the Amino Groups of the Surface of Microtiter Plate Wells.<br />

Each well of the 96-well plate was treated with 100 µL of absolute<br />

ethanol for 5 min at 37 °C and washed five times with 300 µL of<br />

deionized water (DIW). Subsequently, each well was treated with<br />

(3) Kalabay, L.; Gráf, L.; Vörös, K.; Jakab, L.; Benk|Ado˜, Z.; Telegdy, L.; Fekete,<br />

B.; Rohászka, Z.; Füst, G. BMC Gastroenterol. 2007, 7, 15.<br />

(4) Srinivas, P. R.; Wagner, A. S.; Reddy, L. V.; Deutsch, D. D.; Leon, M. A.;<br />

Goustin, A. S.; Grunberger, G. Mol. Endocrinol. 1993, 7, 1445–1455.<br />

(5) Mathews, S. T.; Srinivas, P. R.; Leon, M. A.; Grunberger, G. Life Sci. 1997,<br />

61 (16), l383–1392.<br />

(6) Kalabay, L.; Prohászka, Z.; Füst, G.; Benkõ, Z.; Telegdy, L.; Szalay, F.; Tóth,<br />

K.; Gráf, L.; Jakab, L.; Pozsonyi, T.; Arnaud, P.; Fekete, B.; Karádi, I. In<br />

Liver Cirrhosis: New Research; Chen, T. M., Ed.; Nova Science: New York,<br />

2005; pp 63-75.<br />

(7) Reynolds, J. L.; Skepper, J. N.; McNair, R.; Kasama, T.; Gupta, K.;<br />

Weissberg, P. L.; Dechent, W. J.; Shanahan, C. M. J. Am. Soc. Nephrol.<br />

2005, 16, 2920–2930.<br />

(8) Jethwaney, D.; Lepore, T.; Hassan, S.; Mello, K.; Rangarajan, R.; Dechent,<br />

W. J.; Wirth, D.; Sultan, A. A. Infec. Immun. 2005, 73 (9), 5883–5891.<br />

(9) Ix, J. H.; Shlipak, M. J.; Brandenburg, V. M.; Ali, S.; Ketteler, M.; Whooley,<br />

M. A. Circulation 2006, 113, 1760–1767.<br />

(10) Hermans, M. M. H.; Brandenburg, V.; Ketteler, M.; Kooman, J. P.; Sande,<br />

F. M. V. D.; Boeschoten, E. W.; Leunissen, K. M. L.; Krediet, R. T.; Dekker,<br />

R. T. Kidney Int. 2007, 72, 202–207.<br />

(11) Park, S. J.; Jung, W. Y. J. Colloid Interface Sci. 2002, 250, 93–98.<br />

(12) Cass, T.; Ligler, F. S. Immobilized biomolecules in analysis: a practical<br />

approach; Oxford University Press Inc.: New York, 1998.<br />

10.1021/ac101339q © 2010 American <strong>Chemical</strong> Society 7049<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

Published on Web 07/20/2010


Figure 1. Schematic of the developed sandwich ELISA procedure for the HFA/AHSG assay. Plate functionalization was performed by treatment<br />

with 1% (w/v) KOH and then silanization with 2% (v/v) APTES. KOH oxidizes the organic moiety of the polymer generating a carbonyl (-CO)<br />

or a hydroxyl (-OH) group according to the nature of the polymer. The alkoxy groups of APTES subsequently react with the carbonyl or hydroxyl<br />

groups of the surface in a hydrolysis-dependent reaction. 11,12 Anti-HFA/AHSG antibody was then captured using EDC and sulfo-NHS chemistry<br />

on the aminated 96-well plate, where the EDC-SNHS was used in a ratio of 1:100 with the antibody. The EDC-SNHS-activated antibody molecules<br />

were then immobilized on the aminated ELISA plate, where a bond was formed between amine groups on the plate and the activated carboxyl<br />

groups of the antibody.<br />

100 µL of 1.0% (w/v) KOH at 37 °C for 10 min followed by five<br />

washings with 300 µL of DIW. The KOH-treated wells were then<br />

functionalized with amino groups by incubation with 100 µL of<br />

2% (v/v) aminopropyltriethoxysilane (APTES) per well at 80 °C<br />

for 1 h inside a vacuum-desiccator, in order to achieve maximum<br />

silanization. The reaction mechanism following the KOH activation<br />

and APTES-based surface functionalization involves mild oxidation<br />

followed by a hydrolytic APTES polymerization on the oxidized<br />

surface through its alkoxy groups. 11,12 The desiccator was<br />

equilibrated to room temperature for 20 min. The amine-functionalized<br />

plate was subsequently washed five times with 300 µL<br />

of DIW in order to remove excess unbound 3-APTES from the<br />

surface. Afterward, the anti-HFA/AHSG (R2-HS-glycoprotein; 990<br />

µL of4µg/mL) was incubated with 10 µL of a premixed solution<br />

of 1-ethyl-3-[3-dimethylaminopropyl]carbodiimide hydrochloride<br />

(EDC) (4 mg/mL) and N-hydroxysulfosuccinimide (SNHS) (11<br />

mg/mL) for 15 min at 37 °C. The resulting EDC cross-linked anti-<br />

HFA/AHSG antibody solution was added to each of the functionalized<br />

wells (100 µL) and incubated for 1hat37°C. The anti-<br />

HFA/AHSG-coated wells were then washed five times with 300<br />

µL of PBS. The methodology pertaining to the conventional assay<br />

is described in detail in the Supporting Information methods<br />

section.<br />

ELISA on Anti-HFA/AHSG Antibody-Immobilized Microtiter<br />

Plates. Plates with covalent and passively immobilized anti-<br />

HFA antibody were blocked with 1% (v/v) BSA diluted in 0.1 M<br />

7050 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

phosphate buffered saline (PBS), pH 7.4, for 30 min at 37 °C and<br />

subsequently washed five times with 300 µL of PBS. Varying<br />

concentrations of HFA/AHSG (from 4.8 pg/mL to 20 ng/mL)<br />

were prepared in 0.1 M PBS, pH 7.4, and 100 µL of each of these<br />

concentrations were incubated in the antibody-coated plates for<br />

1hat37°C and, subsequently washed five times with PBS. One<br />

hundred microliters of biotinylated anti-HFA/AHSG detection<br />

antibody (200 ng/mL) was added and incubated for 1hat37°C<br />

followed by five PBS washes. HRP-conjugated streptavidin (100<br />

µL per well), at a dilution of 1:200, was added and then incubated<br />

for 20 min at 37 °C followed by five washes with PBS. The 3,3′,5,5′tetramethylbenzidine<br />

(TMB) substrate was subsequently added<br />

(as per the manufacturer’s recommendations), and the reaction<br />

was stopped after 20 min by addition of 50 µL of1NH 2SO4. The<br />

absorbance was recorded at a primary wavelength of 450 nm<br />

with a reference wavelength of 540 nm. The dual wavelength<br />

system of absorbance measurement in 96-well plates is designed<br />

to eliminate or greatly reduce the optical imperfections<br />

at the well-to-well level.<br />

All the experiments were carried out in triplicate. The control<br />

for this study was 0 ng/mL HFA in 0.1 M PBS, pH 7.4, and the<br />

absorbance of the control was subtracted from all the assay values.<br />

The respective analytical sensitivity (detection limit) of the assay<br />

was determined where analytical sensitivity was calculated using<br />

the formula [mean absorbance of blank +3σ (standard deviation<br />

of the blank)]. Additionally, data sets obtained from the modified


Figure 2. Comparative analysis of developed sandwich ELISA (black<br />

circles) with the conventional format. Anti-human fetuin A (anti-HFA)<br />

antibody was covalently bound to the amine-modified surface in the<br />

developed ELISA, whereas antibody adsorbed on the unmodified<br />

surface was used in the conventional assay format. The conventional<br />

assay was performed with the manufacturer-recommended protocol<br />

(white circles) and the protocol developed in this study (inverted<br />

triangle). The error bars represents standard deviations.<br />

Table 1. Comparison of Assay Performance<br />

Parameters (Precision and Sensitivity) for<br />

Conventional and Developed ELISAs a<br />

intraday precision<br />

range (CV %),<br />

(n ) 5)<br />

interday precision<br />

range (CV %),<br />

(n ) 3)<br />

sensitivity<br />

(pg/mL)<br />

developed ELISA 2.4-10.4 1.7-17.6 39<br />

conventional ELISA 4.7-17.4 3.6-20.0 624<br />

a Intraday precision was calculated from the five repeats (n ) 5) of<br />

the same assay on a single day but at different times, while interday<br />

precision was calculated from assay repeats on three different days (n<br />

) 3). All the assays were carried out in triplicate. <strong>Analytical</strong> sensitivity<br />

was calculated using the formula [average absorbance of the blank +<br />

3(SDblank)].<br />

and conventional ELISA were analyzed using standard curve<br />

analysis of Sigma Plot software, version 11.0.<br />

RESULTS AND DISCUSSION<br />

The range of detection of HFA/AHSG for the developed ELISA<br />

(Figure 2) was 9-20 pg/mL (r 2 ) 0.99). An analytical sensitivity<br />

of 39 pg/mL was recorded for developed ELISA, where<br />

analytical sensitivity was calculated using the formula [average<br />

absorbance of the blank + 3(SDblank)]. Assay variability<br />

parameters for this high-sensitivity assay for HFA are described<br />

in Table 1, where intraday assay variability was calculated from<br />

5 repeats performed on a single day, and interday assay variability<br />

was calculated from three repeats performed on three different<br />

days in triplicate. The percentage recovery for lower concentrations<br />

(4.88-625 pg/mL) was 50-80% while for higher concentrations<br />

(1.2-20 ng/mL) it was 75-100%. The percentage recovery<br />

was calculated as the obtained value/expected value × 100.<br />

The half-effective concentration (EC50) obtained from the<br />

dose-response curve, 13 which is a measure of analyte-ligand<br />

(13) Armbruster, D. A.; Schwarzhoff, R. H.; Hubster, E. C.; Liserio, M. K. Clin.<br />

Chem. 1993, 39, 2137–2146.<br />

interaction, was found to be on an average of 3.3 ng/mL. The<br />

EC50 for all the assay repeats was found to be in the range of<br />

3-3.5 ng/mL. A lower EC50 (3.3 ng/mL), which was obtained<br />

for the developed ELISA, suggests a strong interaction between<br />

HFA and anti-HFA antibody. This interaction behavior is<br />

attributed to the significant increase in the total amount of<br />

immobilized antibody achieved using the covalent immobilization<br />

strategy, which increased the availability of anti-HFA<br />

antibody per HFA molecule, thus increasing the resultant HFA<br />

capture on the well surface.<br />

Conversely, EC50 for the conventional ELISA format was 23<br />

ng/mL, (r 2 ) 0.99), which was significantly higher than the<br />

developed ELISA. This suggests a less sensitive conventional<br />

assay (Supplementary Table 2, Supporting Information). The<br />

most important factor that may be attributed to the enhanced<br />

sensitivity of the modified assay is the covalent immobilization of<br />

the anti-HFA/AHSG capture antibody since the covalently crosslinked<br />

antibodies should not leach out during the assay procedure<br />

in comparison to the use of passively adsorbed antibodies, where<br />

leaching may occur more easily.<br />

Performance of the developed sandwich ELISA procedure was<br />

compared with the conventional protocols (Figure 2), both<br />

involving passively adsorbed antibody on normal unmodified<br />

microtiter plates (one performed using the modified protocol and<br />

the other as per the manufacturer’s recommendations). All the<br />

assays were performed simultaneously on the same day under<br />

same set of conditions in order to minimize variability, where<br />

variability was reported as percentage coefficient of variation (%<br />

CV) (Table 1).<br />

The technique employed here for chemically modifying the<br />

polymeric surface was developed and standardized by our group,<br />

and a standard EDC-based cross-linking strategy was deployed<br />

to immobilize anti-HFA/AHSG antibody. The developed sandwich<br />

ELISA procedure decreased the overall assay duration from 20<br />

to 6 h, which is more than a 3-fold decrease. Thus, the reported<br />

procedure is a rapid ELISA format having the additional benefit<br />

of being generic in nature, i.e., it can be used with sandwich<br />

ELISAs on different polymeric matrixes for a range of different<br />

analytes. Hence, using the surface modification technique reported<br />

here, it is possible to develop rapid and high sensitivity assays<br />

on various categories of solid supports including those that are<br />

chemically inert.<br />

To the best of our knowledge, this developed ELISA procedure<br />

is the most sensitive assay reported for the detection of HFA/<br />

AHSG (Supplementary Table 1, Supporting Information), which<br />

is based on the comparison between different commercially<br />

available ELISA kits for HFA. The sensitivity of the assay<br />

developed by our group may be attributed to the use of monoclonal<br />

antibody at the capture stage, which was covalently<br />

immobilized on the surface, and a polyclonal antibody to detect<br />

the antigen. The covalently cross-linked monoclonal antibodies<br />

that are used as capture antibodies in this study were found to<br />

capture significantly higher amounts of HFA, which has improved<br />

the detection limit and the overall assay sensitivity of the<br />

developed ELISA procedure in comparison to the conventional<br />

format. However, the commercially available ELISA kits, as<br />

described in Supplementary Table 1 (Supporting Information),<br />

have different capture and detection antibody combinations, which<br />

<strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

7051


Figure 3. Performance of the developed ELISA on different solid<br />

support matrixes. Three categories of the solid support were chosen<br />

on the basis of their chemical properties, namely polystyrene (PS)<br />

which is hydrophobic (white circles); polymethyl methacrylate (PMMA),<br />

a hydrophilic polymer (black circles); and polycyclo-olefin polymer,<br />

trade name Zeonex, which is a neutral matrix (inverted black<br />

triangles). The error bars represent standard deviations.<br />

may be a factor in their reduced sensitivities. 14 The sensitivity<br />

and specificity of any immunoassay is dependent on the antigen<br />

capture efficiency of an antibody which is subsequently governed<br />

by the nature of antibody 15,16 and accounts for the different assay<br />

sensitivities of the commercially available kits employing different<br />

type of antibodies that are monoclonal or polyclonal at the capture<br />

stage in ELISA.<br />

Three different types of commercially relevant substrates,<br />

polystyrene (PS) (hydrophobic), polymethyl methacrylate (PMMA)<br />

(hydrophilic), and cyclo-olefin polymer (Zeonex) (inert), were<br />

used in the modified microtiter plate format. The HFA assay was<br />

performed on these matrixes in the modified microtiter plate<br />

format. The HFA assays performed on different solid substrates<br />

could not be directly compared because of their different thick-<br />

(14) Wild, D. The Immunoassay Handbook, 3rd ed.; Elsevier: Oxford, 2005.<br />

(15) Lacy, A.; Dunne, L.; Fitzpatrick, B.; Daly, S.; Keating, G.; Baxter, A.; Hearty,<br />

S.; O’Kennedy, R. J. AOAC Int. 2006, 89 (3), 884–892.<br />

(16) Byrne, B.; Stack, E.; Gilmartin, N.; O’Kennedy, R. Sensors 2009, 9, 4407–<br />

4445.<br />

7052 <strong>Analytical</strong> <strong>Chemistry</strong>, Vol. 82, No. 16, August 15, 2010<br />

nesses. Theoretically, the optical path length contributes toward<br />

the absorbance (A). Therefore, a thick solid support increases<br />

the path length and, hence, significantly changes the absorbance.<br />

However, the generic nature of the developed ELISA procedure<br />

was successfully demonstrated (Figure 3). The HFA assay<br />

performed on the Zeonex support, which is an inert cyclic poly<br />

olefin derivative, had an analytical sensitivity of 19.5 pg/mL, while<br />

for the assay performed on PS and PMMA, it was 78 pg/mL. The<br />

analytical sensitivities for the HFA assay and the dose-response,<br />

which determines the slope-dependent linearity range of the assay,<br />

of HFA-anti-HFA for all the three solid supports used in this study<br />

suggests that this modified protocol increases the overall assay<br />

sensitivity and is not confined to the nature of the solid support<br />

used for capturing antibody.<br />

CONCLUSIONS<br />

A rapid sandwich ELISA procedure was developed for the<br />

highly sensitive detection of HFA/AHSG. It was based on the<br />

covalent immobilization of anti-HFA capture antibody on a<br />

3-APTES-functionalized microtiter plate. The developed ELISA has<br />

comparatively better analytical sensitivity (39 pg/mL) and less<br />

variability in interday and intraday assay repeats than the<br />

conventional format, which has an analytical sensitivity of 624 pg/<br />

mL. The surface chemistry developed for this study is generic<br />

and can be employed to activate polymer matrixes irrespective of<br />

their chemical nature as demonstrated on PS (hydrophobic),<br />

PMMA (hydrophilic), and polycyclo-olefin (inert). Therefore, this<br />

strategy can also be used for rapid assay development on various<br />

commercially relevant substrates and the screening of substrates<br />

for particular biosensor/diagnostic applications.<br />

ACKNOWLEDGMENT<br />

C.K.D. and S.K.V. contributed equally to this work. We<br />

acknowledge Bristol Myers Squibb (BMS), Syracuse, USA, and<br />

Industrial Development Agency, Ireland, for the financial support<br />

under the CBAS Project Code 116294.<br />

SUPPORTING INFORMATION AVAILABLE<br />

Additional information as noted in text. This material is<br />

available free of charge via the Internet at http://pubs.acs.org.<br />

Received for review May 21, 2010. Accepted July 7, 2010.<br />

AC101339Q

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!