30.12.2014 Views

International Journal for Ion Mobility Spectrometry - B & S Analytik ...

International Journal for Ion Mobility Spectrometry - B & S Analytik ...

International Journal for Ion Mobility Spectrometry - B & S Analytik ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

INTERNATIONA L<br />

<strong>International</strong> <strong>Journal</strong><br />

<strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

SOCIETY<br />

<strong>for</strong><br />

ION<br />

MOBILITY<br />

SP ECTROMETRY<br />

5(2002)1<br />

Official publication of the<br />

<strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Table of Contents<br />

<strong>Ion</strong> <strong>Mobility</strong> Spectroemtry - Basis Elements and Applications<br />

J. Stach & J.I. Baumbach<br />

Nitric Oxide as a Reagent Gas in <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

G. A. Eiceman, K. Kelly, E.G. Nazarov<br />

Measuring the temperature of the drift gas in an ion mobility spectrometer:<br />

A technical note<br />

C. L. Paul Thomas, N. D. Rezgui, A. B. Kanu, W. A. Munro<br />

1<br />

22<br />

31<br />

Papers presented at<br />

10 th <strong>International</strong> Conference on <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>, ISIMS 2001,<br />

August 12 - 17, 2001, Wernigerode, Germany<br />

The RIP positions in dependence on the moisture<br />

H. Bensch<br />

Mobilities of halogenated compounds<br />

M.J. Leonhardt, J.W. Leonhardt, H. Bensch<br />

Reporting <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong> Data<br />

and the IUPAC/JCAMP-DX <strong>International</strong> Data Standard<br />

A.N. Davies, P. Lampen, H. Schmidt, J.I. Baumbach<br />

Negative Corona Discharge <strong>Ion</strong>ization Source <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

M. Tabrizchi & A. Abedi<br />

<strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong> of Alkali Salts<br />

M. Tabrizchi<br />

Temperature Corrections <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

M. Tabrizchi<br />

39<br />

43<br />

47<br />

51<br />

55<br />

59<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


<strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong> - Basic Elements and Applications<br />

J. Stach 1 , J.I. Baumbach 2<br />

1<br />

Bruker Saxonia <strong>Analytik</strong> GmbH, Permoserstr. 15, D-04318 Leipzig, Germany<br />

2<br />

Institut für Spektrochemie und Angewandte Spektroskopie (ISAS),<br />

Bunsen-Kirchhoff-Str. 11, D-44139 Dortmund, Germany<br />

Introduction<br />

Although first basic research on ion <strong>for</strong>mation in<br />

air has been known since the eighties of the<br />

past century [1-3], with fundamental papers on<br />

ion mobility in electrical fields published already<br />

around 1905 by Langevin [4], ion mobility<br />

spectrometry (IMS) is considered to be one of<br />

the relatively new analytical methods. Almost<br />

100 years have passed from the origins to their<br />

successful analytical applications which, along<br />

with the first commercialized instruments [5-7],<br />

were mainly attributed to the extensive work of<br />

F.W. Karasek [8-34]. Major steps of<br />

development, of course, have not been taken<br />

into consideration by this judgement. This<br />

includes the follow-up (continuing) work of<br />

Hassé on ion mobility [35,36] as well as basic<br />

research on ion- molecule reactions [37-40] or<br />

experiments with drift tubes which were<br />

published in the sixties [41-43].<br />

The principle ion mobility spectrometry is based<br />

on is very simple. <strong>Ion</strong>s <strong>for</strong>med at normal<br />

pressure migrate in an electrical field against<br />

the direction of a carrier gas, i. e. air in the<br />

easiest case. <strong>Ion</strong> acceleration, along with<br />

permanent collisions between ions and gas<br />

molecules, leads to an average ion velocity<br />

over a certain path length. <strong>Ion</strong>s of different<br />

mass and/or structure reach different velocities<br />

and, thus, get separated (Equ. 1). The quotient<br />

of ion velocity and electrical field strength is<br />

referred to as ion mobility. Under identical<br />

conditions the ion mobility constant K 0, which is<br />

corrected regarding pressure (P 0 = 101,325<br />

kPa) and temperature (T 0 = 273 K), is a<br />

substance-specific value. P 1 and T 1<br />

are the<br />

corresponding values <strong>for</strong> pressure and<br />

temperature measured in the drift tube.<br />

v = K E (1)<br />

K 0 = K (P 1/P 0) (T 0 /T 1) (2)<br />

A comparison of the basic principle of ion<br />

mobility spectrometry with other analytical<br />

methods seems very likely. Terms such as gas<br />

phase electrophoresis or plasma<br />

chromatography became popular in the<br />

relevant literature. However, the high<br />

expectations regarding the analytical<br />

per<strong>for</strong>mance of IMS resulting from these<br />

comparisons could not be confirmed [44]. It<br />

was its excellent ability of detecting chemical<br />

warfare agents [30,45,46] and, at the same<br />

time, of “real time monitoring” that, along with a<br />

clear design and a potential miniaturization of<br />

spectrometers, soon led to extensive<br />

military-oriented development programmes<br />

[47,48]. At the same time further major fields<br />

of application <strong>for</strong> the IMS were opened up. This<br />

includes the detection of explosives, drugs,<br />

pesticides or the detection of working place<br />

pollutants. Coupling techniques such as<br />

GC/IMS or IMS/MS extend the range of<br />

applications [49-56] considerably. The<br />

development of high resolution drift tubes<br />

provides new applications in life sciences<br />

espcially if Electro-Spray <strong>Ion</strong>ization (ESI)<br />

-IMS/MS is considered [57]. However, the<br />

following review covers basics and applications<br />

of IMS from the point of view of mobile<br />

analysis.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 2<br />

1 Basic Elements of <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong><br />

1.1 Function and Components<br />

The basic item of the spectrometer is the<br />

measuring tube consisting of a reaction<br />

chamber in which ions are generated by a<br />

suitable source, and a drift region at the end of<br />

which the detector is positioned. Reaction<br />

chamber and drift region are separated by a<br />

shutter grid. Normally the measuring tube is<br />

made of metal rings separated by isolators.<br />

This set-up is needed to produce an electric<br />

field with field gradients from150 und 300 V<br />

/cm. A drift gas, i. e. air in the easiest case, is<br />

pumped through the measuring tube from the<br />

Figure 1:<br />

Measuring tube (a) and components<br />

(b) of an ion mobility spectrometer<br />

detector’s side. Fig. 1 illustrates the set-up of a<br />

measuring tube and a block diagram showing<br />

the other components of an ion mobility<br />

spectrometer.<br />

Opening the shutter grid <strong>for</strong> a short moment<br />

enables a part of ions to get into the drift<br />

region. Due to the applied electrical field and<br />

the drift gas flow, ions having different masses<br />

and/or structures reach the faraday plate of the<br />

detector at different times. Thus, the recorded<br />

ion mobility spectrum shows signals registered<br />

at different drift times (ms) with the<br />

corresponding intensities (pA).<br />

1.1.1 Inlet Systems<br />

Depending on the problem of analysis various<br />

inlet systems have been decribed in literature.<br />

This includes systems with septa <strong>for</strong> a syringe<br />

injection [58], permeation or diffusion vessels<br />

[59,60] as well as inlet systems making use of<br />

thermal [61,62] or laser desorption [63]. The<br />

thermal desorption plays an important role in<br />

particular <strong>for</strong> the detection of drugs and<br />

explosives. In combination with suitable<br />

sampling and enrichment techniques drugs and<br />

explosives can be detected in the low ppt v<br />

concentration range [50]. Handheld<br />

spectrometers used to analyze the<br />

ambient air are fitted with a<br />

membrane inlet system [64,65].<br />

Direct admission of ambient air is<br />

quite a problem since the <strong>for</strong>mation<br />

of ions would be impaired by the<br />

simultaneous penetration of<br />

moisture. In instruments equipped<br />

with membrane inlet systems, the<br />

ambient air is pumped towards the<br />

membrane. Organic compounds<br />

contained in the ambient air<br />

permeate through the membrane<br />

and get into the measuring tube.<br />

This process can be described using<br />

the following equation:<br />

A (3)<br />

P i =<br />

m P r P P x<br />

F c H + P r A m P<br />

with Pi describing the partial vapor<br />

pressure of an organic compound in<br />

front of the membrane, Pr the<br />

permeability of the substance<br />

through the membrane and P the air<br />

pressure. Am is the permeability<br />

coefficient of the membrane and H<br />

its thickness. Fc describes the gass<br />

flow behind the membrane [52]. Permeability<br />

coefficients were determined <strong>for</strong> different<br />

materials [66,67]. In most cases<br />

dimethylsilicone membranes [64] are used due<br />

to the high retention capacity regarding water,<br />

which is of special importance <strong>for</strong> the <strong>for</strong>mation<br />

of negative ions. However, the membrane inlet<br />

systems have an essential disadvantage: the<br />

detection limits deteriorate due to the small<br />

percentage of the sampling molecules<br />

permeating through the membrane.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 3<br />

1.1.2 Measuring Tubes<br />

Measuring tubes of various designs have been<br />

tested and used in commercial spectrometers.<br />

First of all they vary in the dimensions of the<br />

overall measuring tube, the materials used, the<br />

shutter grid, the selected gas flow and the<br />

sample inlet systems. The design of the<br />

reaction chamber depends mainly on the<br />

selected ionization technique, with ß-emitters<br />

(e.g. 63 Ni) mainly used <strong>for</strong> this purpose [68].<br />

Further ionization techniques will be described<br />

a)<br />

b)<br />

Reaction<br />

region<br />

Reaction<br />

region<br />

Drift gas<br />

Drift gas<br />

Drift region<br />

Drift region<br />

Figure 2:<br />

Uni- (a) and bidirectional (b) flow of drift<br />

gas through the measuring tube<br />

in Section 1.3.<br />

Usually the measuring tubes are composed of<br />

a stack of metal and isolator rings [8,10,69,70].<br />

In most cases glass or ceramics is used as<br />

insulating material. The homogeneity of the<br />

electrical field depends on the radius of the<br />

metal rings and their distance [71]. Apart from<br />

the commonly used stack ring technique,<br />

ceramic tubes are used homogenously coated<br />

with conductive materials to generate an<br />

electric field [72,73]. Usually the measuring<br />

tubes have a length of about 10 cm. However,<br />

experimental setups with larger measuring<br />

tubes are also known. Brokenshire [74], <strong>for</strong><br />

example, describes a high-resultion drift tube<br />

that has a length of about 50 cm and a<br />

diameter of 10 cm.<br />

Recently a drift tube has been described [75]<br />

which, based on a common reaction chamber,<br />

has two drift tubes and, there<strong>for</strong>e, can<br />

simultaneously separate positive and negative<br />

ions. In case of conventional tubes the polarity<br />

must be changed <strong>for</strong> detection of positive or<br />

negative ions. However, the polarity of the<br />

high-voltage can be changed by using suitable<br />

switching techniques in the seconds range,<br />

which results in a quasi simultaneous detection<br />

of positive and negative ions.<br />

Two different variants of drift gas flow are<br />

known. In the easiest case gas flow through the<br />

tube is in just one direction [76]. However, drift<br />

gas flow from the direction of the detector and<br />

the inlet system to the shutter grid is possible<br />

[69,70] as well and is mainly used in<br />

spectrometers with membrane inlet systems.<br />

The two variants are illustrated in Fig. 2.<br />

1.1.3 Shutter Grid<br />

The reaction chamber is separated from the<br />

drift tube by the shutter grid. There are two<br />

different setups in use: one setup according to<br />

Bradbury and Nielsen [77], and a second one<br />

according to Tyndall [78]. The Tyndall shutter<br />

grid consists of two grids made of wires that<br />

are paralleled at a distance of about 1 mm. In<br />

the Bradbury-Nielsen grid the rods are<br />

arranged at one level [79]. The grid potential<br />

primarily depends on the applied field strength<br />

and from the field gradient and, thus, on the<br />

position of the grid within the measuring tube. If<br />

an additional field is applied between the<br />

paralleled wires of the grid (approx. 600 V/cm)<br />

which is directed perpenticularly to that of the<br />

measuring tube, the shutter grid is closed. <strong>Ion</strong>s<br />

will not pass the grid. If the additional electrical<br />

field is switched off, the ions may pass the grid.<br />

1.1.4 Detector and Aperture Grid<br />

The apperture grid is positioned approximately 1<br />

mm in front of the detector which is usually a<br />

circular disk. The apperture grid provides a<br />

capacitive decoupling between arriving ions and<br />

the detector. The main result of this decoupling<br />

is a small line width of the peaks in the ion<br />

mobility spectra. A decrease in signal intensity<br />

and grid vibrations causing an increased noise<br />

level might be disadvantageous [52].<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


⋅<br />

2<br />

0<br />

J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 4<br />

1.1.5 Spectra Acquisition<br />

The acquisition of ion mobility spectra can be<br />

carried out in different ways. A single spectrum<br />

can be recorded in 25 to 50 ms. The short time<br />

needed to record one ion mobility spectrum<br />

provides even if real time monitoring is required<br />

the possibillity of spectra accumulation.<br />

However, the improvement of the signal to noise<br />

ratio is limited since the ion currents depend on<br />

the quotient of the grid opening time t 0 and the<br />

scan time T s [52]. If the grid opening time is 0.25<br />

ms and the scan time 25 ms only 1% of the<br />

<strong>for</strong>med ions reach the detector. t 0 on the other<br />

hand is an important parameter determining the<br />

resolution of the method. Short grid opening<br />

times needed <strong>for</strong> a sufficient resolution will<br />

reduce the ion current remarkably.<br />

A considerabel improvement can be achieved if<br />

the Fourier trans<strong>for</strong>m technique is used. Due to<br />

the applied opening times of the entrance and<br />

exit grid of the drift tube 25% of the ions can be<br />

detected by theory [80]. Experimantely an<br />

intensity gain by a factor of 1.4 to 2.4 can be<br />

REDUCED MOBILITY - K - cm /V sec<br />

3.3<br />

3.2<br />

3.1<br />

3.0<br />

2.9<br />

2.8<br />

2.7<br />

2.6<br />

2.5<br />

2.4<br />

2.3<br />

2.2<br />

2.1<br />

2.0<br />

+<br />

(H0)NH<br />

2 n 4<br />

+<br />

(H0)NO<br />

2 n<br />

+<br />

(H0)H 2 n<br />

achieved. The reason <strong>for</strong> this is the increased<br />

noise level of the interferograms [52].<br />

The dependance of the ion mobility on the field<br />

strength is used by the transverse field<br />

compensation technique <strong>for</strong> ion separation [81].<br />

This method does not need a shutter grid.<br />

However, the experimental setup is different<br />

from the usual drift tubes. The analyser consists<br />

of two electrodes on which a strong high<br />

frequency field and the so called compensation<br />

voltage are applied. If the volatages are in<br />

correspondence with an particular ion mass,<br />

ions can pass the analyzer and reach the<br />

detector.<br />

1.2. <strong>Ion</strong> Formation<br />

Usually ion mobility spectrometers are equipped<br />

with a ß radiation source [10,68,73]. As a result<br />

of Atmospheric Pressure Chemical <strong>Ion</strong>ization<br />

(APCI) processes in the reaction region of the<br />

measuring cell, reactant ions with positive and<br />

negative charges are <strong>for</strong>med. Sample molecules<br />

undergo ion molecule reactions with the reactant<br />

ions resulting in the <strong>for</strong>mation of product ions.<br />

Radiation sources have a very good long term<br />

stability due to the long life time of the isotopes<br />

used. They are of special importance <strong>for</strong> hand<br />

held devices since no energy is required.<br />

Beside ionization with radiation sources<br />

ionization techniques like photo [82], laser<br />

[83,84], corona discharge [85] or electrospray<br />

ionization [85] are well known in ion mobility<br />

spectrometry. The use of thermal ionization on<br />

hot surfaces [52,86,87] and ionization due to<br />

adduct <strong>for</strong>mation with alkaline ions [52,56] was<br />

described as well.<br />

B<br />

% TOTAL IONIZATION<br />

100<br />

80<br />

60<br />

40<br />

20<br />

+<br />

(H2O) n-1H+H2O<br />

n=4<br />

+<br />

(H2O) nH<br />

n=3<br />

0<br />

10 20 40 60 80 100 120 140 160 180 200<br />

TEMPERATURE °C<br />

Figure 3:<br />

Dependence of ion mobility K o (a) and the water content (b)<br />

of the reactant ion clusters (H 2O) nH + from temperature,<br />

authorized by: S.H. Kim et al., Anal. Chem. 50 (1978)<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

n=2<br />

1.2.1 Radiation Sources<br />

63<br />

Ni radiation sources used in ion mobility<br />

spectrometers are identical with those of<br />

electron capture detectors known from gas<br />

chromatography. 63 Ni emitts ß radiation with an<br />

average energy of of 0.067 MeV and a half life<br />

time of 85 years. For some applications Tritium<br />

which emitts ß-particles isused as radiation<br />

source. The average energy is much lower and<br />

amounts to 18 keV.<br />

1.2.1.1 Formation of Reactant <strong>Ion</strong>s<br />

Electrons emitted from the 63 Ni foil collide with<br />

atoms or molecules of the drift gas. If nitrogen is<br />

used as drift gas the released energy amounts<br />

to 35 eV per collision. Nitrogen is ionized as long<br />

as the energy of the radiation is higher than its<br />

ionization potential of 15.58 eV (Equ. 4). This


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 5<br />

starting reaction which can be observed also in<br />

air initiates further competing reactions:<br />

N 2 + ß - → N 2<br />

+<br />

+ e - (4)<br />

N 2<br />

+<br />

+2N 2<br />

→ N 4<br />

+<br />

+ N 2<br />

(5)<br />

N 4<br />

+<br />

+ H 2<br />

O → H 2<br />

O + + 2N 2<br />

(6)<br />

H 2<br />

O + +H 2<br />

O → H 3<br />

O + + OH (7)<br />

H 3<br />

O + + H 2<br />

O + N 2<br />

→ (H 2<br />

O) 2<br />

H + + N 2<br />

(8)<br />

(H 2<br />

O) 2<br />

H + + H 2<br />

O + N 2<br />

→ (H2O) 3<br />

H + + N 2<br />

(9)<br />

The <strong>for</strong>mation of positively charged water<br />

clusters of the type (H 2<br />

O) n<br />

H + as described by<br />

reactions 4-9 was proved by “high-pressure”<br />

MS-investigations [88,89]. The number n of<br />

water molecules contained in the cluster<br />

depends on the temperature and water content<br />

of the drift gas. Fig. 3 illustrates a typical<br />

example of the temperature dependance of n<br />

and the resulting change of the ion mobility<br />

constant. At a temperature of 25 °C , a<br />

pressure of 700 Torr and a rel. Humidity of 20%<br />

the clusters contain 5 to 8 water molecules [90].<br />

However, IMS/MS investigations indicates that<br />

the peak <strong>for</strong>med by the reactant ions is not only<br />

<strong>for</strong>med by the water clusters described above.<br />

a)<br />

POSITIVE ION CURRENT<br />

2 -1 -1<br />

K=2.07 0 cm v s<br />

0 16.7 17.9<br />

41.0<br />

DRIFT TIME (ms)<br />

2 -1 -1<br />

K=1.93 0 cm v s<br />

As recorded IMS/MS spectra clearly show, the<br />

clusters contain beside water nitrogen or other<br />

drift gas components [59, 65, 91, 92], (see Fig.<br />

4).<br />

Considering the <strong>for</strong>mation of negatively<br />

charged ions, sample molecules are ionized by<br />

the attachment of electrons in case that<br />

nitrogen is used as drift gas. If air is used as<br />

drift gas O 2<br />

-<br />

ions as reactive species are<br />

dominant. In humid air attachment of water can<br />

be observed. The <strong>for</strong>med ions have the<br />

composition (H 2<br />

O) nO 2- . Additionally, ions with<br />

the composition (H 2<br />

O)OH - and O 4- , CO 4- , CNO - ,<br />

Cl - , CN - , partly also in hydrated <strong>for</strong>m, were<br />

detected [93-95,69,91]. Fig. 5 shows a typical<br />

example.<br />

1.2.1.2 Formation of Product <strong>Ion</strong>s<br />

The <strong>for</strong>mation of positively charged product<br />

ions depends on the protone affinity of the<br />

substance to be ionized. Since the protone<br />

affinity of water is very low, the water clusters<br />

described above ionize a great number of<br />

organic compound classes. This includes<br />

alkenes, alcoholes, thiophenes, ethers,<br />

aldehydes, ketones, esters, amines, nitriles,<br />

a)<br />

NEGATIV ION CURRENT<br />

DRIFT TIME (ms)<br />

2 -1 -1<br />

K=2.16<br />

0<br />

cm v s<br />

0 16.0<br />

41.0<br />

b)<br />

RELATIVE INTESITY (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

36<br />

+<br />

(H 2 O) 2 NH 4<br />

+<br />

(H 2 O) 3 H<br />

54<br />

55<br />

+<br />

(HO)H<br />

2 4<br />

+<br />

(HO)H<br />

2 5<br />

73<br />

82<br />

83<br />

91<br />

Figure 4:<br />

IMS/MS Coupling: <strong>Ion</strong> mobility spectrum of positively<br />

charged reactant ions (a) and corresponding MS<br />

spectrum (b), providing in<strong>for</strong>mation about all <strong>for</strong>med<br />

reactant ions, authorized by: G.E. Spangler et al., Int.<br />

J. Mass Spectrom. <strong>Ion</strong> Proc. 52(1983) 267.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

101<br />

M/z (u)<br />

+<br />

(HO)NH<br />

2 4 2<br />

+<br />

(HO)NH<br />

2 5 2<br />

119<br />

129<br />

+<br />

(HO)NH<br />

2 4 4<br />

+<br />

(HO)NH<br />

2 4 6<br />

157<br />

b)<br />

RELATIVE INTESITY (%)<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

(HO)O<br />

2 2<br />

50<br />

CO 3<br />

-<br />

O 4<br />

-<br />

60<br />

64<br />

68<br />

-<br />

(HO)O 2 2 2<br />

-<br />

(H<br />

2O)CO<br />

3<br />

78<br />

82<br />

86<br />

M/z (u)<br />

-<br />

(H<br />

2O)O<br />

4<br />

-<br />

(H<br />

2O) 3<br />

O<br />

2<br />

-<br />

(H<br />

2O)CO<br />

4<br />

-<br />

(H<br />

2O) 2CO<br />

3<br />

88<br />

94<br />

96<br />

100<br />

-<br />

(HO)CO<br />

2 2 4<br />

Figure 5:<br />

IMS/MS-Coupling: <strong>Ion</strong> mobility spectrum of negatively<br />

charged ions (a) and corresponding MS spectrum (b),<br />

authorized by: G.E. Spangler et al., Int. J. Mass<br />

Spectrom. <strong>Ion</strong> Proc. 52(1983) 267.


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 6<br />

(H 2<br />

O) n<br />

H + + AB → (AB)H + + n H 2<br />

O (10)<br />

(H 2<br />

O) n<br />

H + + AB → (AB)(H 2<br />

O) m<br />

H + + n-m H 2<br />

O (11)<br />

(AB)(H 2<br />

O) m<br />

H + + AB → (AB) 2<br />

H + + m H 2<br />

O (12)<br />

phosphororganic compounds etc. The reactions<br />

that are dominant <strong>for</strong> the <strong>for</strong>mation of positive<br />

product ions from a substance AB to be<br />

detected are summarized in Equ. 10-15<br />

[96-99]. Whereas <strong>for</strong> low concentrations only<br />

Equ. 10 and 11 apply, in case of higher<br />

concentrations the <strong>for</strong>mation of dimerice<br />

product ions according to Equ. 13 is observed.<br />

Fig. 6 shows a typical ion mobility spectrum.<br />

Clusters containing more than two sample<br />

molecules may result in dependance of the<br />

analyte, its concentration and the temperature.<br />

As IMS/MS studies have shown clustering of<br />

product ions with drift gas molecules has<br />

generally to be taken into account [46,100].<br />

Also in case of product ions the dependence of<br />

cluster <strong>for</strong>mation on temperature has to be<br />

considered. At higher temperatures proton<br />

exchange processes dominate.<br />

The reactions taking place in case of negative<br />

product ion <strong>for</strong>mation are closely related to the<br />

ionization processes observed in an ECD<br />

[34,101]. The major reactions relevant to IMS<br />

are summarized in Equ. 13-15:<br />

[(H 2<br />

O) 3<br />

O 2<br />

] - + AB → [AB] - + 3H 2<br />

O + O 2<br />

(13)<br />

[(H 2<br />

O) 3<br />

O 2<br />

] - + AB → [(AB)O 2<br />

] - + 3H 2<br />

O (14)<br />

[(H 2<br />

O) 3<br />

O 2<br />

] - + AB → A - + B + O 2<br />

+ 3H 2<br />

O (15)<br />

Figure 6:<br />

<strong>Ion</strong> mobility spectrum of positive diethyl-methylphosphonate<br />

(DEMP) ions illustrating the <strong>for</strong>mation of monomer and<br />

dimer ions<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

Halogenated<br />

hydrocarbons, but also<br />

cyano- or nitro<br />

compounds are subject<br />

to a dissociative<br />

charge transfer reaction analogous to Equ. 15.<br />

With high analyte concentrations this reaction<br />

leads to the <strong>for</strong>mation of adducts. Analogously<br />

the <strong>for</strong>mation of O 2<br />

-<br />

adducts is observed.<br />

Reactions strongly depend on the analyte<br />

concentration, temperature and on the humidity<br />

of the measuring tube. Equ. 16-18 summarize<br />

reactions of halogenated compounds. The<br />

composition of the poduct ions was confirmed<br />

by IMS/MS investigations. Tabl. 1 illustrates on<br />

the example of halothane the distribution of<br />

product ions [102] depending on the<br />

M - X + O 2<br />

- → X<br />

- + M + O2 (16)<br />

M - X + O 2<br />

- → (M - X) O2<br />

-<br />

(17)<br />

M - X + X - → (M - X) X - (18)<br />

concentration of the analyte.<br />

Proton abstraction is another typical reaction<br />

<strong>for</strong> the <strong>for</strong>mation of negative ions as observed<br />

<strong>for</strong> aromatic or acid compounds such as<br />

phenoles [31].<br />

1.2.1.3 Quantitative Aspects of <strong>Ion</strong>ization<br />

For quantitative description of ionization the<br />

following prozesses have to be taken into<br />

consideration:<br />

• primary ionization, i. e. <strong>for</strong>mation of<br />

reactant ions,<br />

• ion-molecule reactions <strong>for</strong> the <strong>for</strong>mation<br />

of product ions,<br />

• recombination of positive and negative<br />

ions,<br />

• diffusion of ions towards the sides of<br />

the measuring tube and<br />

• transport of ions from the reaction<br />

chamber due to drift gas and electric<br />

field [103].<br />

The <strong>for</strong>mation of reactant ions depends<br />

on the activity of the ß-ionization source<br />

(C A<br />

) and the radiation energy (W).<br />

Assuming that one ion pair is produced<br />

per 35 eV of primary energy each, 1 x 10 6<br />

C A W of ion pairs are produced per sec<br />

and unit volume. If no sample molecules<br />

are available in the source, the quantity of<br />

reactant ions in the ion source results<br />

from recombination reactions and the


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 7<br />

Table 1:<br />

Dependence of the distribution of halothane product<br />

ions on the concentration, determined by IMS/MS<br />

investigations at a temperature of 40 °C [102]<br />

relative occurrence<br />

product ions 10 ppb 100 ppb 500 ppb<br />

Cl -<br />

31<br />

(H 2O) Cl - 41<br />

(H 2O) 2 Cl - 55<br />

Br - 100 46 15<br />

-<br />

M O 2 24<br />

M Cl - 100 50<br />

M Br - 50 100<br />

discharge of particles due to diffusion and the<br />

drift gas. Thus, in case of usual drift gas flows<br />

and 63 Ni activities between 10 -5 to 10 3 mCi, the<br />

quantity of ions is between 10 4 and 10 11<br />

ions/cm 3 .<br />

The <strong>for</strong>mation of product ions may be<br />

described by simple rate laws assuming that<br />

the total number of charged particles remains<br />

constant. For a simple charge transfer reaction<br />

(analogous to equation 13) Equ. 19-21 can be<br />

derived, with n R standing <strong>for</strong> the number of<br />

reactant ions, n p <strong>for</strong> the number of product ions<br />

and n 0 the number of sample molecules<br />

available <strong>for</strong> the <strong>for</strong>mation of product ions, t <strong>for</strong><br />

the time and K the rate constant of the<br />

bimolecular reaction.<br />

The detection limits reached using ion mobility<br />

spectrometers with ß-ionization sources are<br />

mostly in the ppb-range. They stronly depend<br />

on the substances and vary with the<br />

Figure 7:<br />

Intensity of reactant and product ions in<br />

dependence on sample concentration <strong>for</strong> a charge<br />

transfer reaction<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

composition of the drift gas, with its water<br />

content having a predominant influence.<br />

1.2.1.4 <strong>Ion</strong>ization of Mixtures of Analytes -<br />

Use of Reactant Gases<br />

On the assumption that a stationary equilibrium<br />

has established in the reaction chamber the<br />

ionization of substance mixtures may be<br />

regarded as a result of competing reactions<br />

strongly influenced by the proton and electron<br />

affinity of the sample molecules. Compounds<br />

with a high proton or electron affinity are<br />

preferably ionized. The signal intensity of<br />

product ions depends on the product of proton<br />

or electron affinity and concentration. Thus, <strong>for</strong><br />

a substance mixture A, B, C the correlation<br />

shown in equation 22 using the proton affinity<br />

PA follows.<br />

dn P<br />

/dt = -(dn R<br />

/dt) = K n 0<br />

n R<br />

(19)<br />

n R<br />

= n R<br />

° (e -Knt ) (20)<br />

n P<br />

= n R<br />

° (1 - e Knt ) (21)<br />

ΣReactant-<strong>Ion</strong>en≈(PA A<br />

[C A<br />

] + PA B<br />

[C B<br />

] +PA C<br />

[C C<br />

])<br />

(22)<br />

In addition, in the case of substance mixtures<br />

also reactions among the sample molecules<br />

themselves (e.g. the <strong>for</strong>mation of "mixed"<br />

dimeres according to Equ. 23) may be<br />

observed. The consequence <strong>for</strong> the detection<br />

of single substances is a strong matrix<br />

dependance.<br />

[(AB)(H 2<br />

O) 2<br />

H] + + CD → [(AB)(CD)H] + + 2 H 2<br />

O<br />

(23)<br />

Due to the influence of the proton or electron<br />

affinity on the APCI processes the selectivity<br />

[105] may be increased by selecting suitable<br />

reactant gases, as illustrated in Fig. 8 by<br />

means of a hypothetical substance mixture.<br />

Acetone [106-109] or chlorinated hydrocarbons<br />

[105,110,111] are reactant gases that are<br />

frequently used e. g. <strong>for</strong> the detection of both<br />

chemical warfare agents and explosives.<br />

Whereas in the case of acetone proton bridged<br />

dimeres react as reactant ions, chlorid ions are<br />

<strong>for</strong>med by dissociative charge transfer<br />

reactions from chlorinated hydrocarbons (Equ.<br />

16). Dopant gases can also be used to prevent<br />

a potential peak interference of reactant ions<br />

and product ions as observed e. g. <strong>for</strong> the


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 8<br />

Figure 8:<br />

<strong>Ion</strong>isation of substance mixtures in dependence on the<br />

proton affinity of the reactant gas (acord. to G.A Eiceman<br />

and Z. Karpas, <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>, CRC Press,<br />

Boca Raton, 1994, p. 48)<br />

detection of hydrazine and<br />

monomethylhydrazine, if dibutyl ketone is used<br />

as dopant gas [112].<br />

1.2.2 UV-<strong>Ion</strong>ization<br />

Photoionization is a method that has been<br />

specially introduced <strong>for</strong> the detection of<br />

unsaturated or aromatic hydrocarbons<br />

[113-116]. However, using ultraviolet lamps<br />

with different radiation energies a broad<br />

spectrum of organic compounds can be<br />

ionized. Usually cylindrical gas-filled cathode<br />

lamps containing xenon (8.5 - 9.6 eV), krypton<br />

(10.0 eV), hydrogen (10.2 eV), argon<br />

(7.2 - 11.6 eV) and helium (11.3 - 20.3<br />

eV) are used. The ultraviolet lamps<br />

may be fitted vertically [82] or axially<br />

to the drift tube [117-120]. In particular<br />

in case of axial arrangement a<br />

possible ionization of the sample<br />

molecules behind the shutter grid is<br />

disadvantageous. The <strong>for</strong>mation of<br />

photoelectrons can be utilized <strong>for</strong> the<br />

detection of negative product ions.<br />

Equ. 24 describs the ionization<br />

process. The desactivation reactions<br />

summarized in Equ. 24-27 have to be<br />

considered as well due to their<br />

influence on the ion yield.<br />

In case of substance mixtures charge<br />

transfer reactions have to be taken<br />

into account, i. e. compounds with a<br />

lower ionization potential (IP) are<br />

preferred during ionization. Fig. 9<br />

illustrates this fact by means of a<br />

series of IMS spectra, in the course of which<br />

benzene, toluene and p-xylene were sampled<br />

AB + hν → AB* → AB + + e - (24)<br />

AB* → A + B (25)<br />

AB*+ C → AB + C (26)<br />

AB + + e - + C → AB + C (27)<br />

additionally[121].<br />

The range of detectable substances may be<br />

enlarged by using reactant gases. Thus, using<br />

reactant gases such as acetone the detection<br />

of phosphonates gets possible by<br />

means of UV-IMS [120,121].<br />

In comparison with ß ionization,<br />

UV-ionization has the advantage<br />

that quantitative analyses can be<br />

carried out in a large concentration<br />

range, since linearity is not limited<br />

by the restricted number of reactant<br />

ions. Moreover, the total drift time<br />

range may be used due to absence<br />

of reactant ions.<br />

Figure 9:<br />

Sequence of ion mobility spectra of positive ions 20 ppmv benzene<br />

(IP=9,2eV), 20 ppmv benzene + 20 ppmv toluene (IP=8,8eV) and<br />

20 ppmv benzene + 20 ppmv toluene + 17 ppmv p-xylen<br />

(IP=8,5eV). With the concentrations used just the signal of the<br />

compounds with the lowest ionization potential appears.<br />

1.2.3 Laser <strong>Ion</strong>ization<br />

The use of laser ionization <strong>for</strong> IMS<br />

is mainly described in the papers of<br />

Lubman et al. [83,84]. Both<br />

Nd:YAG as well as ArF-Excimer<br />

lasers were used. The dependence<br />

of IMS spectra on the radiation<br />

energy and the cross section was<br />

investigated [122]. Possibilities <strong>for</strong><br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 9<br />

selective ionization using lasers<br />

with different wave lengths were<br />

demonstrated [123].<br />

Applications of the laser<br />

ionization on surfaces [124,125]<br />

and the Matrix-Assisted Laser<br />

Desorption <strong>Ion</strong>isation (MALDI)<br />

are also known [126].<br />

1.2.4 Corona discharge<br />

Although well-known from mass<br />

spectrometry [127], corona<br />

discharge sources have only<br />

been described in a few papers<br />

[85,120,128,129] in connection<br />

with ion mobility spectrometers.<br />

A major advantage of corona<br />

discharge sources is that a<br />

higher yield of reactant ions can<br />

be gained. The processes<br />

running in the sources are nearly identical with<br />

the known APCI processes [130]. However, in<br />

comparison to usually used<br />

63<br />

Ni-radiation<br />

sources fragmentation reactions may be<br />

observed depending on the design of the<br />

corona discharge source (see Fig. 10).<br />

1.3 <strong>Ion</strong> Separation and <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Ion</strong>s <strong>for</strong>med in the reaction chamber move in<br />

the electric field against the drift gas towards<br />

the detector. According to Equ. 1 the velocity of<br />

the ions on their way from the shutter grid<br />

towards the detector is proportional to the<br />

electrical field strength and the ion mobility<br />

[131-134]. As shown by Equ. 28, K or the<br />

reduced mobility constant K 0 contain properties<br />

of the <strong>for</strong>med ions and of the used drift gas:<br />

In equation 28 q describes the number of<br />

charges ze with e = 1,602 . 10 -19 C, N stands<br />

<strong>for</strong> the density of the drift gas (molecules /cm 3 ),<br />

k is the Boltzmann constant (1,381 . 10 23 J/K)<br />

and µ is the reduced mass of an ion - drift gas<br />

molecule (atom) pair as described in Equ. 29<br />

with M the mass of a drift gas molecule or atom<br />

K = 3/16 q/N (2π/µkT) 1/2 (1 + α)/Ω D . (28)<br />

and m the ion mass.<br />

µ = m M / (m + M) (29)<br />

α in Equ. 28 is a correction value, which is<br />

under 0,02 if m > M [52]. T indicates the<br />

temperature of the drift tube in °K. Ω D is the<br />

collision cross section of ion and drift gas<br />

molecules. According to equation 28, ion<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

Figure 10:<br />

Corona discharge ion mobility spectrum of positive ions of<br />

pinacolyl-methyl-phosponate containing numerous fragment peaks.<br />

The ion mobility spectrum obtained by means of ß-ionization ( 63 Ni)<br />

contains monomere and dimere product ions only.<br />

mobility depends on the temperature, the<br />

charge, the reduced mass and the collision<br />

cross-section. Ω D is influenced by the ion or<br />

molecule size, their structure and polarizibility.<br />

If the same drift gas is used, mobility is largely<br />

controlled by the reduced mass. In case of very<br />

large ions µ is in the range of M. In this case<br />

mobility largely depends on Ω D and, thus, on<br />

ion structure. In the middle range, that means<br />

<strong>for</strong> the most IMS relevant compounds, a<br />

dependence of ion mobility on mass and<br />

structure is observed.<br />

The influence of the structure on ion mobility<br />

has been investigated in numerous studies on<br />

isomeric compounds [24, 28,135-137]. Karpas<br />

found the following classification of the<br />

influence of structure on mobility <strong>for</strong> substituted<br />

hydrocarbons [138]:<br />

• linear < branched<br />

• primary < secundary < tertiary<br />

• aliphatic compounds < aromatic compounds<br />

• amines < amides.<br />

Due to the major influence of the structure<br />

correlations between ion mobility and ion mass<br />

reveal errors up to 20%, if compounds<br />

belonging to different substance classes are<br />

considered. <strong>Ion</strong> mobility-ion-mass functions <strong>for</strong><br />

homologous series of compounds allow the<br />

determination of the ion mass with errors < 5%.<br />

Fig. 11 illustrates the mass-mobility-functions<br />

<strong>for</strong> selected classes of substances [48,138].<br />

The temperature dependence of K and Ω D is<br />

given with T -1/2 and T -2 . Thus, K 0 is to a great<br />

extent independent on the temperature of the<br />

measuring tube [133,139,140]. Due to the


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 10<br />

<strong>Ion</strong><br />

mass<br />

[amu]<br />

400<br />

Phosphor organic<br />

compounds<br />

The resolution of ion mobility spectrometers is<br />

usually determined according to Equ. 30<br />

[143]:<br />

R = t d / 2 t 1/2 (30)<br />

300<br />

200<br />

Oxygen containing<br />

compounds<br />

lg m = -0,452K+2,951<br />

0<br />

lg m = -0,524K+3,123<br />

0<br />

Aromatic<br />

hydrocarbons<br />

lg m = -0,382K+2,870<br />

0<br />

100<br />

1,0 1,5 2,0<br />

2<br />

K 0 [cm/Vs]<br />

Figure 11:<br />

<strong>Ion</strong>-mass-ion mobility functions <strong>for</strong> selected<br />

compound classes<br />

standardization of K regarding pressure and<br />

temperature (Equ. 2) the temperature<br />

dependence of the density N of the drift gas on<br />

ion mobility is compensated.<br />

As systematic investigations with nitrogen, air,<br />

argon, argon/methane mixtures and carbon<br />

dioxide have proved [84], the influence of the<br />

drift gas on ion mobility is essentially<br />

determined by the polarity of the drift gas<br />

molecules. The considerably lower ion<br />

mobilities in carbon dioxide may be explained<br />

by an enlarged cluster <strong>for</strong>mation.<br />

With the same drift gas is used the measuring<br />

results of different spectrometers should be<br />

comparable since ion mobility is standardized<br />

as regards temperature and pressure.<br />

However, small deviations in the experimental<br />

conditions, e. g. changes of the humidity of the<br />

drift or sampling gas, may also cause<br />

differences in measured K 0 values. Thus,<br />

K 0-values (e. g. [141]) published in literature are<br />

not applicable in each case. Using different<br />

calculating methods [56, 142] the calculated ion<br />

mobility constants tally to a high degree with<br />

the values determined by experiments.<br />

In Equ. 30 t d descibes the drift time of the<br />

peak and t 1/2 the peak width. Thus, the<br />

resolution strongly depends on the peak <strong>for</strong>m,<br />

which is influenced by numerous factors. This<br />

includes e. g. the grid opening time, peak<br />

broadening by diffusion or Coulomb<br />

interactions and ion molecule reactions in the<br />

drift tube [144]. The resolution may be also<br />

affected [52] by field and temperature<br />

gradients or pressure fluctuations caused by<br />

the pumps.<br />

1.4. Coupling techniques<br />

1.4.1 Coupling using chromatographic<br />

methods<br />

Due to its excellent detection per<strong>for</strong>mance<br />

IMS is predestined <strong>for</strong> coupling with<br />

chromatographic methods [145]. Beside the<br />

registration of the total ion current |I reactant ions -<br />

I product ions| selective detection is possible by<br />

acquisition of complete ion mobility spectra or<br />

by recording selected K 0 windows.<br />

Already in 1972 an interface <strong>for</strong> GC/IMS<br />

coupling was introduced by Karasek and Keller<br />

[14]. At first packed columns were used as e. g.<br />

reported <strong>for</strong> the separation of mono-, di- and<br />

trichlorotoluene [29]. Problems arised from<br />

column leakage [146] and from the long dwell<br />

times of molecules in the reaction chamber<br />

resulting in additional ion molecule reactions.<br />

Improvements were gained by using capillaries,<br />

by reducing the reaction chamber, and by using<br />

an unidirectional gas flow [76]. The influence of<br />

the drift gas composition, e. g. the oxygen<br />

content, was studied on the detection of<br />

halogenated organics [147]. Selectivity of<br />

GC/IMS has<br />

been proved by detection of<br />

2,4-dichlorophenylacetic acid in soil samles<br />

[148]. Due to the use of ceramics as insulating<br />

material it became possible to operate the IMS<br />

measuring tube even at temperatures of 300°C<br />

[149]. A further progress was made by using an<br />

axial direct coupling (s. Fig. 12) [150].<br />

In particular in the past few years miniaturized<br />

GC/IMS couplings [152-154] with optimized<br />

reaction chambers have been developed [155].<br />

The use of multi-capillary columns e. g. <strong>for</strong> the<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 11<br />

electrical lenses [103]. The design is similar to<br />

that known from mass spectrometers using<br />

APCI ion sources (APCI = Atmospheric<br />

Pressure Chemical <strong>Ion</strong>ization) [166].<br />

IMS/MS experiments have been mainly used<br />

<strong>for</strong> investigating ionization processes running<br />

in ion mobility-spectrometers. However, from<br />

this point of view it always has to be<br />

considered that modifications in the<br />

composition and structure of ions may occur<br />

due to the transfer into the vacuum system<br />

[91].<br />

Different scan-techniques can be used in<br />

IMS/MS experiments [56]:<br />

Figure 12:<br />

GC/IMS coupling techniques; lateral (a) and<br />

axial coupling (b), providing at the same<br />

time the possibility to feed a make-up gas<br />

<strong>for</strong> better eluate swirling [151]<br />

detection of explosives has also been<br />

described [156-157]. An extension of<br />

applications to nonvolatile materials can be<br />

achieved using the Pyrolysis/GC/IMS coupling.<br />

The influence of the mobile phase on the<br />

resulting drift gas composition and, thus, on the<br />

<strong>for</strong>mation of product ions has to be considered<br />

when coupling IMS with SFC (Super Fluid<br />

Chromatography) [158]. This effect can be<br />

suppressed by using measuring tubes with an<br />

unidirectional gas flow [159-162].<br />

Couplings between ion mobility spectrometers<br />

using ß-radiation sources and fluid<br />

chromatographs were implemented at first in<br />

analogy to HPLC/MS coupling by means of a<br />

moving-belt Interface [21]. A “direct” coupling is<br />

possible by using Corona Discharge or Elektro<br />

Spray <strong>Ion</strong> sources [163,164,57].<br />

1.4.2 IMS-MS Coupling<br />

Coupling of IMS with mass spectrometry has<br />

already been described in the first papers on<br />

“Plasmachromatography” [10]. For coupling<br />

usually measuring tubes are used that are<br />

provided with a “pinhole” in the centre of the<br />

detector. The interface between drift tube and<br />

mass spectrometers consists of a system of<br />

1. Acquisition of IMS spectra, recording is<br />

also possible by means of the MS detector.<br />

2. In case of a permanently opened shutter<br />

grid the IMS just ac as ion source <strong>for</strong> the<br />

MS. APCI mass spectra are recorded if<br />

radiation sources are used.<br />

3. Mass analysis of IMS peaks (see e. g. Fig.<br />

6 and 8).<br />

4. If the quadrupole is set to a certain mass<br />

and the IMS operated in the usual manner,<br />

ions of the selected mass are displayed in<br />

the ion mobility spectrum.<br />

Further results regarding the composition and<br />

structure of ions <strong>for</strong>med in the IMS can be<br />

obtained by IMS/MS/MS experiments [166].<br />

2 Applications<br />

In the past few years a large number of IMS<br />

applications have been reported. This includes<br />

the detection of chemical warfare agents,<br />

drugs, explosives, environmental pollutants as<br />

well as applications in the field of working place<br />

monitoring. Nevertheless, IMS is no universal<br />

analysis method. The complex and partly<br />

incalculable ionization mechanisms mainly with<br />

substance mixtures always require a thorough<br />

evaluation of the analysis problem.<br />

Laboratory instruments, stationary devices <strong>for</strong><br />

on-site measurements, which are programmed<br />

<strong>for</strong> selected compounds, or handheld<br />

instruments are generally available <strong>for</strong><br />

corresponding applications. Typical on-site<br />

applications of IMS are described in the<br />

following chapters.<br />

2.1 Chemical warefare agents<br />

At present the detection of chemical warefare<br />

agents is the most important and most common<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 12<br />

Figure 13:<br />

<strong>Ion</strong> mobility spectrum of sulphur mustard (spectrum of negative ions,<br />

C = 400 µg/m 3 )<br />

application of IMS. However, only a few papers<br />

have been published on this subject in literature<br />

[30,45,46,120,167].<br />

In addition to phosphororganic compounds,<br />

namely nerve agents (e. g. tabun, sarin,<br />

soman, VX), blister agents (e. g. sulphur<br />

mustard, nitrogen mustard, lewisite), choking<br />

and blood agents (e. g. phosgene, prussic acid)<br />

can be detected by means of IMS. Whereas<br />

nerve agents have a very high proton affinity<br />

and, thus, <strong>for</strong>m positive ions, blister, blood, and<br />

choking agents are strongly electrophilic and<br />

<strong>for</strong>m negative ions. Thus, modern handheld ion<br />

mobility spectrometers make use of bipolar<br />

measuring tubes in order to assure a more or<br />

less continuous detection of all groups of<br />

warefare agents. The automatically operating<br />

instruments identify and quantify the warefare<br />

agents that are stored in appropriate spectra<br />

libraries. The achieved detection limits are in<br />

the lower ppbv range. The response times of<br />

the instruments depend on the concentration, in<br />

case of concentrations near the detection limits<br />

they are in the minute’s range. High<br />

concentrations are displayed after a few<br />

seconds.<br />

In addition to military applications, the detection<br />

of chemical warfare agents is of importance in<br />

the field of CWA disposal even decades after<br />

the two world wars. As the events in Japan<br />

have shown, the use of warefare agents in<br />

terroristic attacks cannot be excluded. In both<br />

cases ion mobility spectrometers may be used<br />

both as alarm units <strong>for</strong> personal protection and<br />

<strong>for</strong> on-site analysis [168].<br />

In the field of CWA disposal IMS is also used<br />

<strong>for</strong> detection of “old” warfare agents from the<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

World War I and II. Till today<br />

large quantities of<br />

ammunition have been found<br />

on training grounds, old<br />

production sites and battle<br />

fields. Fig. 13. shows a typical<br />

ion mobility spectrum of<br />

sulphur mustard (HD) found<br />

in leaking bottles on a training<br />

ground. For ion mobility<br />

spectra of sulphur mustard<br />

the <strong>for</strong>mation of product ions<br />

of the composition HD*O 2<br />

-<br />

is<br />

typical. In addition Cl - ions are<br />

<strong>for</strong>med due to a dissociative<br />

charge-transfer eaction. As a<br />

result HD*Cl - adducts can be<br />

<strong>for</strong>med in case of higher HD<br />

concentrations as well [169]. Fig. 15 illustrates<br />

the selectivity of IMS using a ß-ionization<br />

source to warefare agents on the example of<br />

IMS and GC/MS investigations made with a<br />

sample of sarin. Sarin was found in a shell<br />

originating from the Second World War. The<br />

composition of the gas phase above the<br />

sample was determined by means of<br />

enrichment on XAD followed by thermal<br />

desorption and GC/MS. The total ion current<br />

chromatogram is shown in Fig. 14a. More than<br />

30 compounds were identified, with the major<br />

compounds listed in Fig. 14a. Fig. 14b shows<br />

the ion mobility spectrum of the same sample.<br />

The spectrum indicates just dimeric product<br />

ions [170] due to the high sarin concentration.<br />

2.2 Explosives<br />

The very good detectability of explosives and<br />

chemical warfare agents by means of IMS has<br />

already been found in the seventies. Karasek<br />

and Denney were able to detect 10 ng TNT at a<br />

signal/noise ratio >1000 [171]. The detection<br />

limit <strong>for</strong> ethyleneglycoldinitrate (EGDN) is about<br />

500 pg. If reactant gases are used which <strong>for</strong>m<br />

Cl - ions still 30 pg EGDN are detectable [172].<br />

However, the extremely low vapour pressure of<br />

explosives normally requires an enrichment.<br />

The use of particle collectors in connection with<br />

a thermal desorption of the enriched explosives<br />

has proved <strong>for</strong> this purpose[173].<br />

Explosives are identified by peaks which are<br />

related to [NO 3] - , [M-H] - , [M . Cl] - and [M . NO 3] -<br />

ions. The <strong>for</strong>mation of dimeric product ions,<br />

[M 2] - , is also possible. These processes<br />

strongly depend on temperature [174]. Adducts<br />

with Cl - ions are caused by chlorine containing


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 13<br />

Intens.<br />

x10 7<br />

4<br />

3<br />

2<br />

Figure 14:<br />

Total ion flow chromatogram of a<br />

sarin sample, obtained by<br />

enrichment on tenax with<br />

subsequent thermal desorption<br />

and GC/MS separation(a, left).<br />

<strong>Ion</strong> mobility spectrum of sarin<br />

(spectrum of positive ions,<br />

C = 600 µg/m 3 -<br />

b, below).<br />

1<br />

0<br />

0,0 2,5 5,0 7,5 10,0 12,5 15,0 17,5 Time [min]<br />

b)<br />

1: Methylchloride 2: Acetone 3: Propyl chloride<br />

4: Benzene 5: Toluene 6: Sarin<br />

7: Chlorobenzene 8: Ethylbenzene 9: o,m,p Xylene<br />

10: Diisopropylphosphonate<br />

reactant gases.<br />

Due to the low detection limits <strong>for</strong> explosives<br />

and the short analysis times IMS is a<br />

interesting method <strong>for</strong> safety areas. The<br />

detection of explosives in or on pieces of<br />

luggage, on clothing or on the skin using IMS<br />

has been described in literature [176].<br />

2.3 Drugs<br />

During the last years IMS has become an<br />

important on-site and laboratory method <strong>for</strong><br />

drug detection due to the low detection limits,<br />

its high selectivity, and short analysis times.<br />

Also in case of drugs it is necessary to enrich<br />

the sample on a carrier material (wire, teflon<br />

and others), followed by a subsequent thermal<br />

desorption. The compounds are usually<br />

detected by means of the pseudo moleculer ion<br />

or specific adduct ions. Karasek et al. [181]<br />

demonstrated by IMS/MS investigations that in<br />

the case of heroin [M] + , [M . H 2] + , [M . CH 3CO 2] +<br />

ions and with cocain [M] + , [M . C 6H 5CO 2] + and<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 14<br />

[M . CH 3CO 2] + ions are <strong>for</strong>med. Nitrogen was<br />

used as drift gas. In addition to natural drugs<br />

such as heroin or cocain, numerous synthetical<br />

drugs can be detected [177-180]. Table 4<br />

contains a selection of natural and synthetical<br />

drugs that have been investigated by means of<br />

IMS. A survey of <strong>for</strong>ensic IMS applications was<br />

published by Karpas [50].<br />

Numerous tests have proved that the detection<br />

of drugs using IMS is very reliable and that<br />

there are only a few interferents. Thus,<br />

detection is possible without further sample<br />

preparation e. g. in pieces of luggage,<br />

containers, letters [183].<br />

2.4 Environmental relevant compounds –<br />

Working place monitoring<br />

Due to the complexity of ionization<br />

mechanisms, applications of IMS in particular in<br />

the field of environmental analyses and working<br />

place monitoring require a careful evaluation of<br />

the analytical problem. Above all, the influence<br />

of matrix on peak selection and calibration has<br />

to be considered. In case of complex substance<br />

mixtures a coupling method should be generally<br />

preferred. Table 4 contains a survey of<br />

compounds or classes of compounds which<br />

can be detected by means of IMS using<br />

ß-ionization in the range between 5 ppb and 1<br />

ppm. In the following some applications will be<br />

given as examples.<br />

Monitoring of 2,4- and 2,6-toluendiisocyanate<br />

[184] is a typical application of IMS in industrial<br />

facilities. Both isomers are reported to have<br />

detection limits of about 5 ppb v. The linear<br />

measuring range is between 0 and 50 ppb v, the<br />

response time runs to a few seconds. The<br />

detection of toluendiisocyanate is interfered by<br />

amines (concentrations in the ppm range) and<br />

halogenated compounds in higher<br />

concentrations. Fig. 15 shows a typical ion<br />

mobility spectrum.<br />

Meanwhile the detection of hydrogen flouride<br />

by means of IMS has been widely used [185].<br />

As reactant gas methylsalicylate is used to<br />

avoid an overlapping of peaks of HF ions and<br />

[(H 2O)xO 2 ] - reactant ions. The detection limit is<br />

approx. 0,5 ppm v, with the linear range running<br />

to 10 ppm v. Interferences regarding NO x, HCl,<br />

chlorine or H 2S are low.<br />

Karpas et al. described a method <strong>for</strong> the<br />

detection of bromine [186]. The detection limit<br />

reached is 10 ppb v. Considering the equilibrium<br />

between Br- and Br 3- ions, which is dependant<br />

on concentration, a linear range between 10<br />

and 500 ppb v is reached.<br />

For the detection of diamide and monomethylhydrazine<br />

[187] the detection limits are reported<br />

to be < 10 ppb. The linear measuring range is<br />

between 10 ppb and 1 ppm. Using 5-nonanone<br />

Table 2:<br />

Survey on selected explosives, their vapour pressure at room temperature and the K0-values of the<br />

product ions used <strong>for</strong> identification. The values given in brackets are detection limits reached by using the<br />

corresponding product ion [175,176]<br />

Compound<br />

Letter<br />

IMS-Peaks, K 0(Detection limit /pg)<br />

Vapour pressure a)<br />

Symbol<br />

[M-H] -<br />

[M.Cl] -<br />

[M.NO 3] -<br />

Nitroglycerine<br />

NG<br />

1,45<br />

1,34 (50)<br />

1,28 (200)<br />

2,8 [mg/m 3 ]<br />

Dinitrotoluene<br />

DNT<br />

1,57 (200)<br />

-<br />

1,0 [mg/m 3 ]<br />

Trinitrotoluene<br />

TNT<br />

1,45 (200)<br />

-<br />

-<br />

56 [µg/m 3 ]<br />

Cyclotrimethylen-trinitr<br />

amine<br />

RDX<br />

-<br />

1,39 (200)<br />

1,32 (800)<br />

0,056[µg/m 3 ]<br />

Pentaerythrittetranitrate<br />

PETN<br />

1,21 (80)<br />

1,15 (200)<br />

1,10 (1000)<br />

0,090[µg/m 3 ]<br />

Ammonium nitrate<br />

-<br />

1,93 (200) a<br />

33 [µg/m 3 ]<br />

a)<br />

Detection as [NO 3] -<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 15<br />

Figure 15:<br />

<strong>Ion</strong> mobility spectrum of negative ions of 2,4-toluendiisocyanate, C = 20 ppb v<br />

Table 3:<br />

Detection of selected drugs using IMS<br />

Compound<br />

Acetylcodeine<br />

Molecular<br />

weight<br />

341<br />

K 0 [cm 2 /Vs] of<br />

important ions<br />

1,09; 1,21<br />

Carrier<br />

gas<br />

Air<br />

T [°C]<br />

220<br />

Method<br />

therm. Des.<br />

Literature<br />

178<br />

Amphetamine<br />

135<br />

1,66<br />

Air<br />

220<br />

therm. Des.<br />

178<br />

Barbital<br />

184<br />

0,99; 1,50<br />

N 2<br />

230<br />

GC/IMS<br />

177<br />

Bromazepam<br />

316<br />

1,24<br />

220<br />

therm. Des.<br />

180<br />

Cannabinol<br />

310<br />

1,06<br />

Air<br />

220<br />

therm. Des.<br />

180<br />

Chlordiazepoxid<br />

300<br />

1,18<br />

Air<br />

220<br />

therm. Des.<br />

182<br />

Cocain<br />

Codein<br />

303<br />

299<br />

1,16; 1,50; 1,84<br />

1,18; 1,21<br />

N 2, air<br />

Air<br />

153,<br />

220<br />

220<br />

therm. Des.<br />

therm. Des.<br />

181,182,<br />

178,179<br />

Diazepam<br />

Heroin<br />

Morphine<br />

285<br />

369<br />

285<br />

1,21<br />

1,05; 1,15<br />

1,04; 1,14<br />

1,22; 1,26<br />

Air<br />

N 2<br />

N 2, air<br />

Air<br />

220<br />

153<br />

220<br />

220<br />

therm. Des.<br />

therm. Des.<br />

therm. Des.<br />

therm. Des.<br />

178,179,18<br />

0<br />

181<br />

178,182<br />

178<br />

Opium<br />

1,55<br />

Air<br />

250<br />

direct, wire<br />

50<br />

Oxazepam<br />

287<br />

1,23; 1,28<br />

Air<br />

220<br />

therm. Des.<br />

180<br />

Triazolam<br />

343<br />

1,13<br />

Air<br />

220<br />

therm. Des.<br />

178<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 16<br />

Table 4:<br />

Survey of compounds detected by means of IMS using ß-ionization in the field<br />

of environmental analyses [54,56,112]<br />

Compound<br />

Chlorine<br />

Bromine<br />

Iodine<br />

Hydrogen fluoride<br />

Hydrogen chloride<br />

Hydrogen iodide<br />

Prussic acid<br />

Phosgen<br />

Sulphour dioxide<br />

Nitrogen dioxide<br />

Nitric acid<br />

Ammonia<br />

Hydrazine<br />

Hydrogen sulphide<br />

Detection limit<br />

[ppb v]<br />

100<br />

100<br />

5<br />

100<br />

100<br />

100<br />

100<br />

100<br />

100<br />

100<br />

100<br />

100<br />

10<br />

1000<br />

Classes of compounds<br />

Alcoholes<br />

Aliphatic amines<br />

Aromatic amines<br />

Ethers<br />

Esters<br />

Ketones<br />

Phenoles<br />

Chlorinated aromatic<br />

compounds<br />

Polychlorinated biphenyles<br />

Carbon tetrachloride<br />

Dry etching gases<br />

Detection limit<br />

[ppb v]<br />

100<br />

5<br />

5<br />

100<br />

10<br />

10<br />

100<br />

100<br />

100<br />

500<br />

100<br />

Acetonitrile<br />

Acetaldehyde<br />

Aniline<br />

Cyclohexanone<br />

Nitrobenzene<br />

Toluendiisocyanate<br />

Vinylacetate<br />

10<br />

100<br />

5<br />

10<br />

5 5<br />

100<br />

Figure 16:<br />

(a) Total ion chromatogram (TIC) of a benzene fraction (Kp. 65-80 °C) with 20 ppm v of highly volatile<br />

chlorinated hydrocarbons (dichloromethane, chloro<strong>for</strong>m and 1,2 dichloroethane),<br />

(b) IMS spectrum of the gaseous phase acquired above the sample.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 17<br />

as reactant gas a separation of reactant and<br />

product ion signals is achieved [112].<br />

Another advantage of IMS is illustrated in Fig.<br />

16, which shows an ion mobility spectrum of a<br />

benzene fraction (bp. 65-80 °C) containing 20<br />

ppm v of highly volatile chlorinated<br />

hydrocarbons. The chlorinated hydrocarbons<br />

<strong>for</strong>m a Cl- peak with K 0 = 2,78 cm 2 /Vs due to a<br />

dissociative charge transfer reaction in the IMS.<br />

Thus, this signal may be used <strong>for</strong> detection of<br />

the total amount of volatile chlorinated aliphatic<br />

compounds [188]. Since the detection limits <strong>for</strong><br />

chlorinated hydrocarbons strongly depend on<br />

the compound [189], the standard mixtures<br />

required <strong>for</strong> calibration have to be adapted to<br />

the analytical problem. The respective matrix<br />

has to be considered also.<br />

The determination of certain classes of<br />

substances is only possible if the respective<br />

compounds, e. g. by means of a dissociative<br />

charge transfer reaction, <strong>for</strong>m the same type of<br />

ions. Typical examples are chlorinated,<br />

brominated and iodized hydrocarbons or some<br />

cyano compounds.<br />

Further applications are known from the<br />

semiconductor industry (detection of etch<br />

gases [190]) and from electronics (detection of<br />

de-gassing from electronic equipment [191]).<br />

Investigations of the use of IMS <strong>for</strong><br />

determination of narcotisvon have also been<br />

made [102]. The headspace analysis of soil<br />

samples <strong>for</strong> benzine by means of IMS using UV<br />

ionization was described by Eiceman et al.<br />

[192].<br />

3 Summary<br />

Approximately since 1970 ion mobility<br />

spectrometry has undergone a continuous<br />

development. The continuous growth in<br />

publications may be evidence <strong>for</strong> it. Efficient<br />

laboratory instruments and rugged,<br />

miniaturized and automated devices <strong>for</strong> on-site<br />

analysis are available. As far as the latter are<br />

concerned both stationary devices <strong>for</strong> building<br />

and plant monitoring and hand-held<br />

spectrometers can be used. Applications from<br />

many sectors of the analytical chemistry are<br />

known.<br />

Device components have been optimized<br />

according to the respective applications. IMS<br />

systems to be used <strong>for</strong> fast on-site analysis<br />

have a membrane inlet system. Portable or<br />

laboratory devices are fitted with a shutter grid<br />

system <strong>for</strong> sampling. Sample enrichment<br />

followed by a thermal desorption is typical, <strong>for</strong><br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

example, <strong>for</strong> the detection of drugs or<br />

explosives.<br />

<strong>Ion</strong>ization is mainly achieved by using<br />

ß-radiation sources. They do not need energy,<br />

which is especially important <strong>for</strong> hand-held<br />

spectrometers, and they are stable over many<br />

years. <strong>Ion</strong>ization of compounds is steered by<br />

their proton and electron affinities. A possible<br />

influence of the matrix may be bypassed by<br />

using a reactant gas. The detection<br />

per<strong>for</strong>mance of the instruments allows most of<br />

the compounds to be determined. Chemical<br />

warfare agents, explosives or drugs can be<br />

detected in the lower ppb-range. The linear<br />

range <strong>for</strong> the detection of many compounds is<br />

very limited. There are alternative ionization<br />

techniques available such as UV or corona<br />

discharge ionization. UV ionization depends on<br />

the ionization potential of the substances to be<br />

analyzed and on the energy of the radiation<br />

released by the lamp. The linear range <strong>for</strong> the<br />

detection of the compounds is not limited.<br />

Using reactant gases APCI processes can be<br />

initialized resulting in the ionization of<br />

compounds which, normally, cannot be<br />

recorded by means of UV ionization. By using<br />

corona discharge ionization a higher ion output<br />

can be gained. However, fragmentation<br />

reactions have an unfavorable influence.<br />

The <strong>for</strong>med ions are separated in bipolar<br />

measuring tubes, i. e. positive and negative<br />

ions are detected by voltage reversal. Also<br />

known are measuring tubes which, proceeding<br />

from a common reaction chamber, have two<br />

drift tubes and, thus, are able to simultaneously<br />

separate positive and negative ions.<br />

Separation of <strong>for</strong>med ions in the drift tube of<br />

the measuring chambers depends on their<br />

mass and structure. In comparison to gas<br />

chromatography or mass spectrometry IMS has<br />

a low resolution capacity of IMS is low. This<br />

may be compensated <strong>for</strong>, as e. g. in the case of<br />

matrix influences, by selecting a suitable<br />

ionization technique. The detection of complex<br />

mixtures requires coupling techniques made<br />

available by GC/IMS and IMS/MS.<br />

IMS covers manifold fields of applications.<br />

Major detectable substances are chemical<br />

warfare agents, drugs, explosives or some<br />

ecologically relevant compounds that can be<br />

efficiently ionized using ß-radiation sources. In<br />

most cases mobile IMS are used <strong>for</strong> these<br />

applications.


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 18<br />

Literature<br />

[1] W.C. Röntgen, Science 3 (1886) 726.<br />

[2] J.S. Townsend, Philos. Trans. R. Soc. London<br />

A193 (1899) 129.<br />

[3] J.J. Thomson, G.P. Ruther<strong>for</strong>d, "Conduction of<br />

Electricity Through Gases",Dover, New York,<br />

1928.<br />

[4] P. Langevin, Ann. de Chim. Phys. 5 (1905) 245.<br />

[5] D.I. Carroll, M.J. Cohan, R.F. Wernlund, U.S.<br />

Patent 3,626,180, 1971.<br />

[6] D.I. Carroll, U.S. Patent 3,668,383, 1972.<br />

[7] M.J. Cohen, D.I. Carroll, R.F. Wernlund, und<br />

W.D. Kilpatrick, U.S. Patent 3,699,333, 1972.<br />

[8] M.J. Cohan und F.J. Karasek, J. Chromatogr. Sci.<br />

8 (1970) 330.<br />

[9] F.W. Karasek, Res. Dev. 21 (1970) 25.<br />

[10] F.W. Karasek, Res. Dev. 21 (1970) 34.<br />

[11] F.W. Karasek, W.D. Kilpatrick und M.J. Cohan,<br />

Anal. Chem. 43 (1971) 1441.<br />

[12] F.W. Karasek, Anal. Chem. 43 (1971) 1982.<br />

[13] F.W. Karasek, M.J. Cohan und D.I. Carroll, J.<br />

Chromatogr. Sci. 9 (1971) 390.<br />

[14] F.W. Karasek und R.A. Keller, J. Chromatogr.<br />

Sci. 10 (1972) 626.<br />

[15] F.W. Karasek, Int. J. Environ. Anal. Chem. 44<br />

(1972) 157.<br />

[16] F.W. Karasek und D.M. Kane, J. Chromatogr.<br />

Sci. 10 (1972) 673.<br />

[17] F.W. Karasek und O.S. Tatone, Anal.Chem. 44<br />

(1972)1758.<br />

[18] F.W. Karasek, O.S. Tatone und D.M. Kane,<br />

Anal.Chem. 45 (1973) 1210.<br />

[19] F.W. Karasek, O.S. Tatone und D.W. Denney, J.<br />

Chromatogr. 87 (1973) 137.<br />

[20] F.W. Karasek, Can. Res. Dev. 6 (1973) 19.<br />

[21] F.W. Karasek und D.W. Denney, Anal. Lett. 11<br />

(1973) 993.<br />

[22] F.W. Karasek, Anal. Chem. 46 (1974) 710R.<br />

[23] F.W. Karasek, D.W. Denney und E.H. Dedecker,<br />

Anal.Chem. 46 (1974) 970.<br />

[24] F.W. Karasek und D.M. Kane, Anal. Chem. 46<br />

(1974) 780.<br />

[25] F.W. Karasek und D.M. Kane, J. Chromotogr. 93<br />

(1974) 125.<br />

[26] F.W. Karasek und D.W. Denney, Anal. Chem. 46<br />

(1974) 1312.<br />

[27] F.W. Karasek, A. Maican und O.S. Tatone, J.<br />

Chromatogr. 110 (1975) 295.<br />

[28] F.W. Karasek und S.H. Kim, Anal. Chem. 47<br />

(1975) 1166.<br />

[29] F.W. Karasek, H.H. Hill, Jr., S.H. Kim und S.<br />

Rokushika, J.Chromatogr.135 (1977) 329.<br />

[30] J.M. Preston, F.W. Karasek und S.H. Kim, Anal.<br />

Chem. 49 (1977) 1746.<br />

[31] F.W. Karasek, S.H. Kim und H.H. Hill, Jr., Anal.<br />

Chem. 48 (1978) 1133.<br />

[32] S.H. Kim, F.W. Karasek und S. Rokushika, Anal.<br />

Chem. 50 (1978) 152.<br />

[33] F.W. Karasek, S.H. Kim und S. Rokushika, Anal.<br />

Chem. 50 (1978) 2013.<br />

[34] F.W. Karasek und G.E. Spangler, ”Theory and<br />

Practice in Chromatography” in Electron Capture,<br />

Herausg. Zlatis, A.; Poole, C.F., Elsevier,<br />

Amsterdam, 1981, S. 377.<br />

[35] H.R. Hassé, Phil. Mag. 1 (1926) 139.<br />

[36] H.R. Hassé und W.R. Cook, Phil. Mag. 12 (1931)<br />

554.<br />

[37] H.D. Smyth, Phys. Rev., 25 (1925) 452.<br />

[38] P. Harnwell, Phys. Rev. 29 (1927) 683.<br />

[39] T.R. Hogness und R.W. Harkness, Phys.Rev. 32<br />

(1928) 936.<br />

[40] M.M. Mann, R. Hustrulid und J.T. Tate, Phys.<br />

Rev. 58 (1940) 340.<br />

[41] E.A. Mason, H.W. Schamp, Jr., Ann. Phys. (N.Y.)<br />

4 (1958) 233.<br />

[42] W.E. McDaniel, Collisional Phenomena in <strong>Ion</strong>ized<br />

Gases, John Wiley& Sons, New York, 1964.<br />

[43] D.L. Albriton und E.W. McDaniel, J. Phys. Chem.<br />

171 (1968) 94.<br />

[44] M.M. Metro und R.A. Keller, J. Chromatogr. Sci<br />

12 (1974) 673.<br />

[45] H.A. Moye, J. Chromatogr. Sci. 13 (1975) 285.<br />

[46] S.H. Kim und G.E. Spangler, Anal. Chem. 57<br />

(1985) 567.<br />

[47] J.P. Carrico, A.W. Davis, D.N. Campbell, J.E.<br />

Roehl, G.R. Sima, G.E. Spangler, K.N. Vora und<br />

R.J. White, Am. Lab. 1986, 152.<br />

[48] V. Starrock, A, Krippendorf and H.-R.Döring,<br />

Second <strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Quebec City, Canada,<br />

15.-18.8.1993.<br />

[49] M.A. Morrissey und H. Widmer, Chimia 43 (1989)<br />

268.<br />

[50] Z. Karpas, Forensic Sci. Rev. 1 (1989) 103.<br />

[51] H.H. Hill, Jr., W.F. Siems, R.H.St. Louis und D.G.<br />

McMinn, Anal. Chem. 62 (1990) 201A.<br />

[52] R.H.St. Louis und H.H. Hill, Jr., Crit. Rev. Anal.<br />

Chem. 21 (1990) 321.<br />

[53] G.A. Eiceman, Crit. Rev. Anal. Chem. 22 (1991)<br />

471.<br />

[54] J.E. Roehl, Appl. Sectr. Rev. 26 (1991 1.<br />

[55] G.A. Eiceman, R.E. Clement und H.H. Hill,Jr.,<br />

Anal. Chem. 64 (1992) 170R.<br />

[56] G.A. Eiceman und Z. Karpas, <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, CRC Press, Boca Raton,1994.<br />

[57] J.H. Cross, Th. P. Limero und J.T. James,<br />

Proceedings of the Fourth <strong>International</strong><br />

Workshop on <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>,<br />

Cambridge, England 6.-9.8.1995.<br />

[58] M.M. Metro, und R.A. Keller, J. Chromatogr. Sci.<br />

11 (1973) 520.<br />

[59] A.E. O’Keefe und G.C. Ortman, Anal. Chem. 38<br />

(1966) 760.<br />

[60] R.L. Grob, Herausg., ”Modern Practice of Gas<br />

Chromatography”, John Wiley & Sons, New<br />

York, 1977.<br />

[61] T.W. Carr und C.D. Needham, in Surface<br />

Contamination, Herausg., K.L. Mittal, Plenum<br />

Press, New York, 1979.<br />

[62] A.A. Nanji, A.H. Lawrence und N.Z. Mikhael, Clin.<br />

Toxicol. 25 (1987) 501.<br />

[63] S.D. Huang, L. Kolaitis und D.M. Lubman, Appl.<br />

Spectrosc. 41 (1987) 1371.<br />

[64] G.E. Spangler und C.I. Collins, Anal. Chem. 47<br />

(1975) 393.<br />

[65] G.E. Spangler und J.P. Carrico, J. Mass<br />

Spectrom. <strong>Ion</strong> Phys. 52 (1983) 267.<br />

[66] G.R. Youngquist, ”Flow through Potous Media”,<br />

Nunje, R.T., Herausgeb., Washington D.C., 1970.<br />

[67] G.E. Spangler, Am. Lab. 7 (1975) 36.<br />

[68] T.W. Carr, Herausg., ”Plasma Chromatography,<br />

Plenum Press, New York, 1984.<br />

[69] G.E. Spangler und J.P. Carrico, Int. J. Mass<br />

Spectrom. <strong>Ion</strong> Phys. 52 (1983) 267.<br />

[70] Bruker-Saxonia, US Patent 5,280,175,<br />

18.1.1994.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 19<br />

[71] G.E. Spangler und M.J. Cohan, Instrument<br />

Design and Description, in Plasma<br />

Chromatography, Herausg. W.T. Carr, Plenum<br />

Press, New York, 1984, S.1.<br />

[72] J.O. Carrico, D.W. Sickenberger, G.E. Spangler<br />

und K.N. Vora, J. Phys. E 16 (1983) 1059.<br />

[73] G.E. Spangler, D.N. Campbell, K.V. Vora und J.P.<br />

Carrico, J.P., ISA Trans.23 (1984)17.<br />

[74] J.L. Brokenshire, FACSS Meeting, Anaheim, CA,<br />

Oktober 1991.<br />

[75] A. Brittain et al., 5th Int. Symp. on Protection<br />

Against Chemical and Biological Warfare Agents,<br />

Stockholm, 11.-16.6.1995.<br />

[76] M.A. Baim und H.H. Hill, Jr. Anal. Chem. 54<br />

(1982) 38.<br />

[77] N.E. Bradbury und R.A. Nielsen, Phys. Rev. 49<br />

(1936) 388.<br />

[78] A.M. Tyndall, The <strong>Mobility</strong> of Positive <strong>Ion</strong>s in<br />

Gases, University Press, Cambridge, 1938.<br />

[79] Bruker-Saxonia, Patent DE 4310106 C1,<br />

6.10.1994.<br />

[80] J.E. Knorr, R.L. Etherton, W.F. Siems und H.H.<br />

Hill, Jr., Anal. Chem. 57 (1985) 402.<br />

[81] B. Carnahan, S. Day, V. Kouznetsov und A.<br />

Tarassov, Proceedings of the Fourth <strong>International</strong><br />

Workshop on <strong>Ion</strong> <strong>Mobility</strong> Spektrometry,<br />

Cambridge, England, 6.-9.8.1995.<br />

[82] M.A. Baim, R.L. Eatherton und H.H. Hill, Jr., Anal.<br />

Chem. 55 (1983) 1761.<br />

[83] D.M. Lubman und M.N. Kronick, Anal. Chem. 54<br />

(1982) 1546/2289.<br />

[84] L. Kolaitis und D.M. Lubman, Anal. Chem. 58<br />

(1986) 1993.<br />

[85] C.B. Shumate und H.H. Hill, Jr., Anal. Chem. 1<br />

(1989) 601.<br />

[86] H. Wohltzer, U.S. Patent 581398, 3.8.1984.<br />

[87] U.Kh. Rasulev, E.G. Nazarov und V.V. Palitsin,<br />

Proceedings of the Fourth <strong>International</strong><br />

Workshop on <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>,<br />

Cambridge, England, 6.-9.8. 1995.<br />

[88] M.M. Shahin, J. Chem. Phys. 45 (1965) 2600.<br />

[89] A. Good, D.A. Dueden und P. Kebarle, J. Chem.<br />

Phys. 52 (1970) 212.<br />

[90] J. Sumner, G. Nicol und P. Kebarle, Anal. Chem.<br />

60 (1988) 1300.<br />

[91] G.E. Spangler, Proceedings of the Third<br />

<strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Galveston, Texas, 16. - 19.10.<br />

1994, S. 115.<br />

[92] S.H. Kim, K.R. Betty und F.W. Karaseck, Anal.<br />

Chem. 50 (1978) 2006.<br />

[93] T.W. Carr, Anal. Chem. 49 (1977) 828.<br />

[94] T.W. Carr, Anal. Chem. 51 (1979) 705.<br />

[95] M.W. Siegel und W.L. Fite, J.Chem. Phys. 80<br />

(1976) 2871.<br />

[96] P. Kebarl, Ann. Rev. Phys. Chem. 28 (1977) 445.<br />

[97] G. Nicol, J. Sunner und P. Kebarl, Int. J. Mass<br />

Spectrom. <strong>Ion</strong> Proc. 84 (1988) 135.<br />

[98] J. Sunner, G. Nicol und P. Kebarl, Anal. Chem.<br />

60 (1988) 1300.<br />

[99] J. Sunner, G. Nicol und P. Kebarl, Anal. Chem.<br />

60 (1988) 1308.<br />

[100] P. Watts, Anal. Proc. 28 (1991) 328.<br />

[101] E.P. Grimsrud, Mass Spectrom. Rev. 10 (1992)<br />

457.<br />

[102] G.A. Eicman, D.B. Shoff, C.S. Harden, A.P.<br />

Snyder, P.M. Martinez, M.E. Fleisher und M.L.<br />

Watkins, Anal. Chem. 61 (1989) 1093.<br />

[103] M.W. Siegel, ”Atmospheric Pressure <strong>Ion</strong>ization” in<br />

Plasma Chromatography,. T.W. Carr,<br />

Herausg, Plenum Press, New York, 1984.<br />

[104] J.C. Tou und G.U. Boggs, Anal. Chem. 48 (1976)<br />

1352.<br />

[105] C.J. Proctor und J.F.J. Todd, Anal. Chem. 56<br />

(1984) 1794.<br />

[106] G.E. Spangler, D.N. Campbell und J.P. Carrico,<br />

Pittsburgh Conf. , Atlantic City, NJ, 1983.<br />

[107] D.A. Blyth, Proceedings of the <strong>International</strong><br />

Symposium on Chemical Warfare Agents,<br />

Stockholm, 1983.<br />

[108] R.B. Turner und J.L. Brokenshire, Trends in Anal.<br />

Chem. 13 (1994) 275.<br />

[109] G.A. Eiceman, A.P. Snyder und D.A. Blyth, Int. J.<br />

Environ. Anal. Chem. 38 (1990) 415.<br />

[110] G.E. Spangler, J.P. Carrico und D.N. Campbell, J.<br />

Test Eval. 13 (1985) 234.<br />

[111] A.H. Lawrence und P. Neudorfl, Anal. Chem. 60<br />

(1988) 104.<br />

[112] G.A. Eiceman, M.R. Salazar, M.R. Rodriguez, Th.<br />

F. Limero, St. W. Beck, J.H. Cross, R. Young,<br />

und J.T. James, Anal. Chem. 65 (1993) 1696.<br />

[113] U. Boesl, J.J. Neusser und W.E. Schlag, J.<br />

Chem. Phys. 72 (1980) 4327.<br />

[114] M. Seaver, J.W. Hudgens und J.J. DeCorpo, Int.<br />

J. Mass Spectrom. <strong>Ion</strong> Phys. 34 (1980) 159.<br />

[115] A.N. Freedman, J. Chromatogr. 190 (1980) 263.<br />

[116] J.S. Hayhurst und J.N. Driscoll, Anal. Proc.<br />

[London] 30 (1993) 90.<br />

[117] C.S. Leasure, M.E. Fleischer, G.K. Anderson und<br />

G.A. Eiceman, Anal. Chem. 58 (1986) 2142.<br />

[118] J.W. Leonhardt, H. Bensch, D. Berger, M. Nolting<br />

und J.I.Baumbach, Proceedings of the Third<br />

<strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Galveston, Texas, 16. -<br />

19.10.1994, S. 49.<br />

[119] J.I. Baumbach, D. Berger, J.W. Leonhardt und D.<br />

Klockow, Int. J. Environm. Anal. Chem. 52 (1993)<br />

189.<br />

[120] J. Stach, J. Adler, M. Brodacki und H.-R. Döring,<br />

Proceedings of the Third <strong>International</strong> Workshop<br />

on <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>, Galveston, Texas,<br />

16. - 19.10. 1994, S. 71.<br />

[121] Environmental Technology Group Inc., Patent<br />

WO93/22033.<br />

[122] G.A. Eiceman, V.T. Vandiver, C.S. Leasure, G.K.<br />

Anderson, T.T. Tiee und W.C. Danen, Anal.<br />

Chem. 58 (1986) 1690.<br />

[123] D.M. Lubman und M.N. Kronick, Anal. Chem. 55<br />

(1983) 867.<br />

[124] G.A. Eiceman, G.K. Anderson, W.C. Danen, M.J.<br />

Ferris und J.J. Tiee, Anal. Letters 21 (1988) 539.<br />

[125] J. Phillips und J. Gormally, Int. J. Mass Spectrom.<br />

<strong>Ion</strong> Proc. 112 (1992) 205.<br />

[126] A.W.T. Bristow, C.S. Creaser und J.W. Stygall,<br />

Proceedings of the Fourth <strong>International</strong><br />

Workshop on <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>,<br />

Cambridge, England, 6.-9.8. 1995.<br />

[127] R.F.D. Bradshaw, U.K. Patent 1,606,926, 1978.<br />

[128] J.I. Baumbach, A.v. Irmer, D. Klockow, S.M.<br />

Alberti Segundo, St. Sielmann, O. Soppart und E.<br />

Trindade, Proceedings of the Fourth <strong>International</strong><br />

Workshop on <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>,<br />

Cambridge, England, 6.-9.8. 1995.<br />

[129] D.I. Carrol, I. Dzidic, E.C. Horning und R.N.<br />

Stillwell, Appl. Spectros. Rev. 17 (1981) 337.<br />

[130] G.A. Eiceman, J.H. Kramer, A.P. Snyder und J.K.<br />

Tofferi, J. Eniron. Anal. Chem. 33 (1988) 161.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 20<br />

[131] E.W. McDaniel und E.A. Mason, The <strong>Mobility</strong> and<br />

Diffusion of <strong>Ion</strong>s in Gases, Wiley & Sons, New<br />

York, 1973.<br />

[132] H.E. Revercomb und E.A. Mason, Anal. Chem. 47<br />

(1975) 970.<br />

[133] E.A. Mason, ”<strong>Ion</strong> <strong>Mobility</strong>: Its Role in Plasma<br />

Chromatography” in Plasma Chromatography,<br />

Herausg. T.W. Carr, Plenum Press, New York,<br />

1984, S 43.<br />

[134] E.A. Mason und E.W. McDaniels, Transport<br />

Properties of <strong>Ion</strong>s in Gases, Wiley & Sons, New<br />

York, 1987.<br />

[135] D.F. Hagen, Anal. Chem. 51 81979) 1979.<br />

[136] Z. Karpas, M.J. Cohan, R.M. Stimac und R.<br />

Wernlund, Int. J. Mass Spectrom. <strong>Ion</strong> Proc. 74<br />

(1986) 153.<br />

[137] Z. Karpas, R.M. Stimac und Z. Rappoport, Int. J.<br />

Mass Spectrom. <strong>Ion</strong> Proc. 83 (1988) 163.<br />

[138] G.W. Griffin, I. Dzidic, O.I. Carroll, R.N. Stillwell<br />

und E.C. Horning, Anal. Chem. 45 (1973) 1204.<br />

[139] D.C. Perent und M.T. Bowers, Chem. Phys. 60<br />

(1981) 257.<br />

[140] D.M. Lubman, Anal. Chem. 56 (1984) 1298.<br />

[141] C. Shumate, R.H. St. Louis und H.H. Hill, Jr.,<br />

Chromatogr. 373 (1986) 141.<br />

[142] P.C. Jurs und M.D. Wessel, Proceedings of the<br />

Fourth <strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Cambridge, England, 6.-9.8. 1995.<br />

[143] S. Rokushika, H. Hatano, M.A. Baim und H.H.<br />

Hill, Jr., Anal. Chem. 67 (1985) 1902.<br />

[144] P. Watts und A. Wilders, Int. J. Mass Sperctrom.<br />

<strong>Ion</strong> Proc. 112 (1992) 179.<br />

[145] H.H. Hill und D.G. McMinn, Chem. Analysis 121<br />

(1992) 297.<br />

[146] T. Ramstad, T.J. Nestrick und J.C. Tou, J.<br />

Chromatogr. Sci. 16 (1978) 240.<br />

[147] M.A. Baim und H.H. Hill, Jr., J.High Resolution<br />

Chromatogr. Chromatogr. Commun. 6 (1983) 4.<br />

[148] M.A. Baim und H.H. Hill, Jr., J. Chromatogr. 279<br />

(1983) 631.<br />

[149] R.L. Eatherton, W.F. Siems und H.H. Hill, Jr.,<br />

J.High Resolution Chromatogr. Chromatogr.<br />

Commun. 9 (1986) 44.<br />

[150] R.H. St. Luis, W.F. Siems und H.H. Hill, Jr., J.<br />

Chromatogr. 479 (1988) 221.<br />

[151] St. J. Taylor, WO 93/03360; CA 118, 204481x.<br />

[152] A.P. Snyder, C.S. Harden, A.H. Brittain, M.-G.<br />

Kim, N.S. Arnold und H.L.C. Meuzelaar, Am. Lab.<br />

1992, 32B-H.<br />

[153] A.P. Snyder, C.S. Harden, A.H. Brittain,<br />

M.-G.Kim, N.S. Arnold und H.L.C. Meuzelaar,<br />

Anal. Chem. 65 (1993) 299.<br />

[154] J.D. Dworzanski, M.-G. Kim, A.P. Snyder, N.S.<br />

Arnold und H.L.C. Meuzelaar, Anal. Chim. Acta<br />

293 (1994) 219.<br />

[155] N.S. Arnold, D.L. Hall, R. Wilson, S.J. Taylor,<br />

A.H. Brittain und A.P. Snyder, Proceedings of the<br />

Fourth <strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Cambridge, England, 6.-9.8. 1995.<br />

[156] A. Mercado und P. Mardson, Proceedings of the<br />

Third <strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Galveston, Texas, 16. - 19.10.<br />

1994, S. 168.<br />

[157] A.P. Snyder, S.N. Thornton, J.P. Dworzanski,<br />

W.H. McClenen und H.L.C. Meuzlaar,<br />

Proceedings of the Fourth <strong>International</strong><br />

Workshop on <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>,<br />

Cambridge, England, 6.-9.8. 1995.<br />

[158] S. Rokushika, . Hatano und H.H. Hill, Jr., Anal.<br />

Chem. 59 (1987) 8.<br />

[159] R.L. Eatherton, M.A. Morissey, W.F. Siems und<br />

H.H. Hill, Jr., J. High Resolution Chromatogr.<br />

Chromagr. Commun. 9 (1986) 154.<br />

[160] R.L. Eatherton, M.A. Morissey und H.H. Hill, Jr.,<br />

Anal. Chem. 60 (1988) 2240.<br />

[161] M.X. Huang, K.E. Markides und M.L. Lee,<br />

Chromatographia 31 (1991) 163.<br />

[162] M.A. Morrissey und H.M. Widmer, J. Chromatogr.<br />

552 (1991) 551.<br />

[163] J. Geniec, L.L. Mack, K. Nakamae, C. Gupta, V.<br />

Kumar und M. Dole, Biomed Mass Spectrom. 11<br />

(1984) 295.<br />

[164] D.G. McMinn, J.A. Kinzer, C.B. Shumate, W.F.<br />

Siems, H.H. Hill, Jr., J. Mikrocolumn Sep. 2<br />

(1990) 188.<br />

[165] D.I. Caroll, I. Dzidic, E.C. Horning und R.N.<br />

Stillwell, Appl. Spectr. Rev. 17 (1981) 337.<br />

[166] D.B. Shoff und C.S. Harden, Proceedings of the<br />

Fourth <strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Cambridge, England, 6.-9.8. 1995.<br />

[167] Z. Karpas und Y. Pollevoy, Anal. Chim. Acta 259<br />

(1992) 333.<br />

[168] Kölbel-Bölke, A. Loudon, A. Adler und J. Stach,<br />

Analusis 23 (1995) M22.<br />

[169] A.J. Bell und P. Watts, Proceedings of the Fourth<br />

<strong>International</strong> Workshop on <strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>, Cambridge, England, 6.-9.8. 1995.<br />

[170] J. Stach, K. Miersch, D. Schaper und H.-J.<br />

Döring, 5th Int. Symp. on Protection Against<br />

Chemical and Biological Warfare Agent,<br />

Stockholm, 11.-16.6.1995.<br />

[171] F.W. Karasek und D.W. Denney, J. Chromatogr.<br />

93 (1974) 141.<br />

[172] A.H. Lawrence und P. Neudorfl, Anal. Chem. 60<br />

(1988) 104.<br />

[173] D.D. Fetterolf und F.W. Whitehurst, 39th Conf.<br />

Am. Soc. Mass Spectrom., Nashville, TN, 1991.<br />

[174] L.L. Danylewych-May, Proc. 1st Int. Symp.<br />

Explosion and Detection Technolgy, Atlantic-City<br />

NJ, Nov. 1991, Vortrag C-10.<br />

[175] R.K. Ritchie, F.J. Kuja, R.A. Jackson, A.J.<br />

Loveless und L.L. Danylewych-May, Proc. Int.<br />

Symp. on Substance Identification Technologies,<br />

Insbruck, Österreich, 4.-6.10.<br />

[176] D.D. Fetterolf in Advances in Analysis and<br />

Detection of Explosives, Herausg. E. J. Yinon,<br />

Kluwer Academic Publishers, 1993, S.117-132.<br />

[177] D.S. Ithakissios, J. Chromatogr. Sci. 18 (1980)<br />

88.<br />

[178] A.H. Lawrence, Anal. Chem. 58(1986) 1269.<br />

[179] A.H. Lawrence, Forensic Sci. Int. 34 (1987)73.<br />

[180] A.H. Lawrence, Anal. Chem. 61 (1989) 343.<br />

[181] F.W. Karasek, H.H. Hill, Jr. und S.H. Kim, J.<br />

Chromatogr. 117 (1976) 327.<br />

[182] A.H. Lawrence und L. Elias, Bull. Narcot. 37<br />

(1985) 3.<br />

[183] M. Chauhan, J. Harnois, j. Kovar und P: Pilon,<br />

Can. Soc. Forensic Sci. J. 24 (1991) 43.<br />

[184] J.L. Brokenshire, V. Dharmarajan, L.B. Coyne<br />

und J. Keller, J. Cell. Plast. 29 (1990) 123.<br />

[185] A.T. Bacon, R. Getz und J. Reategui, J. Chem.<br />

Eng. Prog. 6 (1991).<br />

[186] Z. Karpas, Y. Pollevoy, und S. Melloul, Anal.<br />

Chim. 249 (1991) 503.<br />

[187] C.S. Leasure und G.A, Eiceman, Anal. Chem. 57<br />

(1985) 1890.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


J.Stach and J.I. Baumbach.: "<strong>Ion</strong> mobility spectrometry..”, IJIMS 5(2002)1, 1-21, p. 21<br />

[188] J. Stach, J. Flachowski, M. Brodacki und H.-R.<br />

Döring, Fourth Int. Symp. on Field Screening<br />

Methods on Hazardous Wastes and Toxic<br />

Chemicals, Las Vegas, 22.-24.2.1995, P84.<br />

[189] Z. Karpas, Y.-F. Wang und G.A. Eiceman, Anal.<br />

Chim. Acta 282 (1993) 19.<br />

[190] T.W. Carr, Thin Solid Films, 45 (1977) 115.<br />

[191] K. Budde, Electrochem. Soc. 90 (1990) 315.<br />

[192] G.A. Eiceman, M.E. Fleischer und C.S. Leasure,<br />

Int. J. Environ. Aal. Chem. 2 (1987) 279.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


NITRIC OXIDE AS A REAGENT GAS IN<br />

ION MOBILITY SPECTROMETRY<br />

G. A. Eiceman, K. Kelly, and E.G. Nazarov<br />

Chemistry and Biochemistry, New Mexico State University, Las Cruces, New Mexico 88003, USA<br />

ABSTRACT<br />

Nitric oxide was introduced into the ion source<br />

of an ion mobility spectrometer in air at ambient<br />

pressure so conventional hydrated proton<br />

(H 3O + (H 2O) n) ion chemistry could be replaced<br />

with ionization reactions based on charge<br />

exchange, hydride abstraction, and adduct<br />

<strong>for</strong>mation. The addition of NO(g) from 1 to 60<br />

mg/m 3 at temperatures from 125 to 250°C<br />

resulted in progressive replacement of<br />

H 3O + (H 2O) n with NO + (H 2O) n reaching a plateau<br />

of ~0.9 <strong>for</strong> [NO + (H 2O) n]/[NO + (H 2O) n +<br />

H 3O + (H 2O) n]. <strong>Mobility</strong> spectra <strong>for</strong> three<br />

chemical substances were obtained through a<br />

range of temperatures from 100 to 250°C at<br />

50°C increments with 5 to 10 values <strong>for</strong> [NO] at<br />

each temperature. <strong>Ion</strong>s were mass-identified<br />

using a mobility spectrometer/mass<br />

spectrometer. The chemicals were 2,4-lutidine,<br />

di-tert-butyl-pyridine (DTBP), and<br />

dimethylmethylphosphonate (DMMP). The<br />

core product ion <strong>for</strong> 2,4-lutidine with a reagent<br />

ion of H 3O + (H 2O) n was mass identified as MH +<br />

while the product ion with NO(H 2O) + was an<br />

adduct ion M*NO + . The adduct ion of<br />

2,4-lutidine was thermally unstable above<br />

150°C and underwent dissociation to M and<br />

NO + . The mobility spectrum <strong>for</strong> DMMP was<br />

MH + with hydrated proton reactant ions and<br />

chemistry was unaltered with the addition of<br />

NO(g) as reagent gas. The product ion <strong>for</strong><br />

DTBP was MH + with H 3O + (H 2O) n and addition of<br />

NO(g) resulted in <strong>for</strong>mation of fragment ions<br />

(M-CH 3) + , (M-(t-butyl)) + and (M-(t-butyl))H + (NO).<br />

These results demonstrate that NO(g) can<br />

serve as a reagent gas <strong>for</strong> the <strong>for</strong>mation of<br />

product ions through reactions other than<br />

proton transfer using a conventional<br />

beta-source. At temperatures below 125 o C, the<br />

hydrated proton could not be replaced even<br />

partially with NO(H 2O) + though the level of<br />

NO(g) reached 140 mg/m 3 .<br />

INTRODUCTION<br />

In ion mobility spectrometry (IMS), hydrated<br />

protons H 3O + (H 2O) n have been used commonly<br />

as the reservoir of charge <strong>for</strong> atmospheric<br />

pressure chemical ionization (APCI) reactions<br />

(1). The reactions between analyte and the<br />

hydrated protons or reactant ions leads to<br />

product ions at ambient pressure through what<br />

might be described best as a displacement<br />

reaction as shown in Equation 1:<br />

In addition to the hydrated proton, other<br />

reactant ions can be observed at minor<br />

intensities in a clean IMS drift tube and these<br />

include NH 4+ (H 2O) n, and NO + (H 2O). The level<br />

of hydration, i.e. value <strong>for</strong> n, is dependent upon<br />

temperature and moisture. The ratios of<br />

intensities <strong>for</strong> these reactant ions in scrubbed<br />

air or nitrogen at temperatures of 150+ o C are<br />

roughly 1:1:10 <strong>for</strong> NH 4+ (H 2O) n, and NO + (H 2O),<br />

M<br />

+<br />

H 3O + (H 2O) n<br />

MH + (H 2O) n-1<br />

+<br />

H 2O<br />

(1)<br />

sample<br />

neutral<br />

reactant ion<br />

product ion<br />

water<br />

neutral<br />

M<br />

+<br />

MH + (H 2O) n-1<br />

M 2H + (H 2O) n-1)<br />

+<br />

H 2O<br />

(2)<br />

sample<br />

neutral<br />

protonated<br />

monomer<br />

proton bound dimer<br />

water<br />

neutral<br />

Received <strong>for</strong> review April 30, 2002, Accepted July 15, 2002<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 23<br />

and H 3O + (H 2O) n, respectively (2). The<br />

advantages of APCI reactions include both high<br />

selectivity and detection limits of µg/m 3 or<br />

lower. Also, ions are near-thermal energies in<br />

APCI reactions so mobility spectra are<br />

comparatively simple and are comprised largely<br />

of the intact protonated sample. In instances<br />

when the sample vapor levels are large, proton<br />

bound dimers (M 2H + (H 2O) n-1) can <strong>for</strong>m as<br />

shown in Equation 2.<br />

However, the simplicity of APCI reactions and<br />

the principles that underlie selectivity can<br />

present disadvantages <strong>for</strong> chemical<br />

measurements. For example, protonated<br />

monomers or proton bound dimers provide little<br />

structural in<strong>for</strong>mation about a sample. Also,<br />

substances that exhibit low proton affinities<br />

may not be ionized or detected in the presence<br />

of matrices or interferences that have elevated<br />

proton affinities. In short, reactions through<br />

proton transfers in APCI can be considered<br />

somewhat inflexible <strong>for</strong> user-tailored response<br />

in analytical measurements. One means to<br />

impart some control over the selectivity of APCI<br />

reactions is to use alternate reagent gases (3).<br />

This is generally accomplished by adding a<br />

reagent gas into the ion source so the<br />

conventional reactant ions can be replaced by<br />

alternate reactant ions. The most common<br />

approach has been to employ reagent gases<br />

that have relatively high proton affinities and<br />

these gases are added to the ionization region<br />

at concentration of ~500 µg/m 3 . Alternate<br />

reactant ions may exclude from ion <strong>for</strong>mation<br />

those molecules with proton affinities below<br />

that of the reagent gas <strong>for</strong> comparable<br />

concentrations. Thus, substances that are not<br />

ionized should not appear as interferences (4).<br />

Apart from the hydrated protons, the next most<br />

intense reactant ion, NH 4+ (H 2O) n, can serve as<br />

an alternate reactant ion with elevated proton<br />

affinities but un<strong>for</strong>tunately introduces<br />

complications largely through the <strong>for</strong>mation of<br />

cluster ions.<br />

Another reactant ion available <strong>for</strong> ion <strong>for</strong>mation<br />

in IMS is NO + (H 2O) n and the basis <strong>for</strong> ionization<br />

with NO + (H 2O) n would be principally charge<br />

exchange. Thus, patterns of relative response<br />

to a sample using source chemistry with<br />

NO + (H 2O) n might be changed from that<br />

observed with hydrated proton reagent ions.<br />

Studies in chemical ionization mass<br />

spectrometry (CI-MS) have shown that small<br />

amounts of NO(g) in the ion source can result<br />

in the nitryl ion [NO + ] as the predominant<br />

reactant ion and that the chemistry of the nitryl<br />

ion was substantially different from the<br />

hydrated proton clusters. Though such<br />

reactions were studied at pressures of 1-10<br />

torr, the findings have relevance <strong>for</strong> IMS<br />

studies at ambient pressure. The three reaction<br />

paths identified through CI-MS studies included<br />

hydride abstraction, charge exchange, and<br />

adduct <strong>for</strong>mation as shown in Equations 3 to 5,<br />

respectively:<br />

These reactions are dependent upon the<br />

properties of the sample (M) and some patterns<br />

in selection of pathway <strong>for</strong> ionization can be<br />

discerned. For example, substances with high<br />

ionization energies (>9.3 eV) will not undergo<br />

charge exchange (Equation 4) and instead may<br />

react via Equations 3 and 5. Hydride<br />

abstraction is seen with alkanes and with<br />

molecules containing heteroatoms or<br />

aromaticity. Molecules with high electron<br />

density can be expected to <strong>for</strong>m adduct ions.<br />

Little if any fragmentation has been observed in<br />

CI-MS studies when NO was used as a reagent<br />

gas (5-10) except in isolated instances such as<br />

with highly branched molecules (11). Because<br />

of the variety of reactions possible, it was even<br />

suggested that nitric oxide will ionize most<br />

compounds regardless of the ionization<br />

potential (12). Thus, possibilities exist <strong>for</strong><br />

M +<br />

NO +<br />

(M-H) +<br />

+<br />

HNO<br />

(3)<br />

sample neutral<br />

reactant ion<br />

hydride abstracted<br />

product ion<br />

neutral<br />

M +<br />

NO +<br />

M +<br />

+<br />

NO<br />

(4)<br />

sample neutral<br />

reactant ion<br />

product ion<br />

M<br />

+<br />

NO +<br />

(M*NO) +<br />

(5)<br />

sample neutral<br />

reactant ion<br />

adduct ion<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 24<br />

addition control in selectivity of response and<br />

this motivated a study of NO + as a reactant ion<br />

<strong>for</strong> IMS. A potential complication to this<br />

concept (not seen at low pressure) is the<br />

<strong>for</strong>mation of ion clusters with water or other<br />

small neutrals at ambient pressure in IMS and<br />

APCI-mass spectrometry.<br />

The chemistry of NO + (H 2O) n has been<br />

described previously in IMS measurements <strong>for</strong><br />

only a single chemical (13) and the main<br />

product ion was an adduct per Equation 5.<br />

Nonetheless, several environmentally important<br />

chemicals might be detected preferentially<br />

through charge exchange rather than proton<br />

transfer reactions. Such chemicals include<br />

benzene and substituted benzenes. If NO(g) is<br />

a suitable reagent gas, drift tubes containing<br />

conventional ion sources (i.e., 10 mCi of 63-Ni)<br />

might be easily adapted to selective detection<br />

of benzene, toluene and xylenes. A second<br />

interest is to determine if NO-based ionization<br />

chemistry can be used to resolve or clarify<br />

response to mixtures when the components<br />

may under ionization through different and<br />

distinctive paths. For example, the product<br />

ions <strong>for</strong> two substances with the same drift time<br />

from a traditional proton-based ionization<br />

sources, may differentiated by <strong>for</strong>mation of<br />

cluster ions, proton abstracted ions or M + ions<br />

simultaneously.<br />

In this present study, the mobility spectra <strong>for</strong><br />

several well-known chemicals in IMS were<br />

studied using hydrated proton and NO + with<br />

mass-identification of ions using IMS/MS.<br />

<strong>Mobility</strong> spectra were obtained through a range<br />

of concentrations of NO(g) and concentrations<br />

of samples. Also, the temperatures needed to<br />

obtain reagent ions from NO(g) were<br />

determined.<br />

Table 1. The drift tube temperature was set<br />

using an model 6100 controller (Omega<br />

Engineering, Inc., Stam<strong>for</strong>d, CT) and a model<br />

MIS-2 meter (Panametric, Inc., Waltham, MA)<br />

was used to monitor both temperature (via an<br />

Omega series CCT-23 0/400C transducer) and<br />

moisture (Panametric probe). Moisture was<br />

kept between 5-10 µg/m 3 throughout the<br />

experiments through the use of in-house<br />

supplied air treated using an Addco model 737<br />

Pure Air Generator (Clearwater, FL), several<br />

scrubbing towers containing 5Å molecular<br />

sieve, and a commercial drier (R&D<br />

Separations GC-2 Disposable Moisture Getter,<br />

Alltech Associates, IL) and monitored using a<br />

Panametric (Waltham, MA) “Moisture Image<br />

Series 2” hygrometer probe.<br />

The mass spectrometer was a model API-III<br />

tandem mass spectrometer (PE-SCIEX,<br />

Toronto, Canada) and the drift tube was a<br />

conventional discrete ring design with the<br />

following specific features: insulating rings of<br />

high-temperature stable Teflon (PTFE);<br />

press-fit seals on the Teflon rings to provide a<br />

first level of pneumatic isolation; an aluminum<br />

shell (drift tube casing) with Teflon seals to<br />

provide a second level of pneumatic isolation;<br />

access to electric utilities at an end cap; and<br />

preheated drift gas. The drift tube was<br />

equipped with a dual shutter design though<br />

generally the shutters were full open to gain<br />

maximum ion intensity in the MS. <strong>Ion</strong>s<br />

reaching the end of the drift tube were passed<br />

through an additional distance of 3cm equipped<br />

with a radiator surface and allowed the body of<br />

the IMS to be cooled. The end of this length<br />

was fitted into a Teflon socket placed into the<br />

high voltage flange of the API-III.<br />

EXPERIMENTAL SECTION<br />

Instrumentation<br />

The GC-IMS was a<br />

Hewlett-Packard (Avondale, PA)<br />

model 5880A gas chromatograph<br />

equipped with a 30 m long, ID 0.25<br />

mm, df 0.25 µm RTX-50 capillary<br />

column (Restek Corporation,<br />

Bellefonte, PA), split/ splitless<br />

injector, flame ionization detector<br />

(FID) and ion mobility detector.<br />

The ion mobility spectrometer was<br />

designed and built at New Mexico<br />

State University in Las Cruces, NM<br />

and specifications are listed in<br />

2,4-Lutidine<br />

C 7H 9N<br />

MM 107.15<br />

CAS Registry<br />

Number: 108-47-4<br />

DMMP<br />

C 3H 9O 3P<br />

MM 124.08<br />

CAS Registry<br />

Number: 756-79-6<br />

DTBP<br />

C 13H 21N<br />

MM 191.31<br />

CAS Registry<br />

Number: 585-48-8<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 25<br />

Chemical and Reagents<br />

All chemicals were reaction or analytical grade<br />

and were used as received. The solvent <strong>for</strong><br />

samples was methylene chloride (Fisher<br />

Scientific, Houston, TX). Standards were<br />

prepared from three chemicals including<br />

2,4-lutidine (Sigma Chemical Company, St.<br />

Louis, MO), 2,6-di-tert-butylpyridine or DTBP<br />

(Aldrich Chemical Co.), and<br />

dimethylmethylphosphonate or DMMP (Aldrich<br />

Chemical Co.).<br />

Two stock solutions were made <strong>for</strong> chemicals<br />

at concentrations of 130 mg/m 3 in methylene<br />

chloride. These were a mixture of 2,4-lutidine<br />

and DTBP and a second mixture of DMMP and<br />

DTBP. These were necessary owing to the<br />

poor chromatographic resolution between<br />

2,4-lutidine and DMMP. A 5% mixture of nitric<br />

oxide in nitrogen (Matheson Gas,<br />

Montgomeryville, PA) was scrubbed over 5A<br />

molecular sieves to remove water and was<br />

introduced directly into the source region of the<br />

IMS via a 10 cm length of uncoated fused silica<br />

capillary column. The flow rate of the nitric<br />

oxide was measured using a microliter scale<br />

bubblemeter as 5 to 400 µl/min providing a final<br />

concentration in the drift tube of 1-60 mg/m 3<br />

with a drift gas flow rate of 300 ml/min.<br />

Procedures<br />

Retention times and purity of solutions were<br />

first determined using the GC-FID. With the ion<br />

mobility spectrometer as the detector, flow<br />

rates of nitric oxide were selected on the basis<br />

of the relative intensity of the NO + peak in the<br />

mobility spectrum. The initial value <strong>for</strong> NO +<br />

was determined and NO(g) was gradually<br />

increased until the hydrated proton peak was<br />

barely visible. Under these conditions, the<br />

system was kept from excessive levels of<br />

NO(g). The instrument was allowed to stabilize<br />

between each increment of nitric oxide and all<br />

TABLE 1:<br />

Parameters <strong>for</strong> gas chromatograph-mobility spectrometer<br />

Gas Chromatograph<br />

Initial Temperature (/C)<br />

Initial Time (min)<br />

Program Rate (/C/min)<br />

Final Temperature (/C)<br />

Final Time (min)<br />

Splitless time (min)<br />

Split time (min)<br />

60<br />

0<br />

20<br />

200<br />

5<br />

0<br />

.75<br />

<strong>Mobility</strong> Spectrometer<br />

Length of Drift Region (cm)<br />

5.2<br />

Length of <strong>Ion</strong> Source Region (cm)<br />

1.0<br />

Insulator Ring Inner Diameter (cm)<br />

Insulator Ring Outer Diameter (cm)<br />

Insulator Ring Thickness (cm)<br />

Conducting Ring Inner Diameter (cm)<br />

Conducting Ring Outer Diameter (cm)<br />

Conducting Ring Thickness (cm)<br />

Shutter Diameter (mm)<br />

Collector Diameter (mm)<br />

Aperture Diameter (mm)<br />

Detector Aperture Distance (mm)<br />

1.9<br />

5.1<br />

1<br />

1.9<br />

4<br />

0.2<br />

13<br />

9.5<br />

10.3<br />

~1<br />

Amount of 63 Ni in Source (mCi)<br />

Electric Field in Drift Region (V/cm)<br />

Frequency of shutter pulse (Hz)<br />

Drift Gas (air) Flow Rate (ml/min)<br />

Temperature of Drift Gas (°C)<br />

(preheated)<br />

Delay in spectra (ms)<br />

Shutter Pulse width(µs)<br />

No. of points digitized per spectrum<br />

Collector-Aperture Voltage (V)<br />

No. of averages per spectrum<br />

10<br />

*<br />

216<br />

34<br />

200<br />

150-250<br />

2<br />

208<br />

640<br />

60<br />

64<br />

* = The electric field changes slightly with temperature due to thermal expansion of the Teflon. V/cm<br />

at 250/C = 216, V/cm at 200/C = 217.6, V/cm at 150/C = 219<br />

Insulating Rings: High Temperature Teflon - Conducting Rings: 303 Stainless Steel<br />

Shutters and detector were obtained from Graseby Dynamics and used without modification.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 26<br />

Intensity<br />

0 2 4 6 8 10<br />

Figure 1:<br />

<strong>Mobility</strong> spectra <strong>for</strong> reactant ions with different flows of<br />

NO(g) into the ion source at 250°C.<br />

other flows and moisture remained constant<br />

throughout the study. One-microliter injections<br />

of solutions were made into the chromatograph<br />

injector by splitless methods. Spectra were<br />

taken from the time of injection until the last<br />

chemical eluted. Background spectra were<br />

taken from the time prior to the first chemical<br />

elution. The maximum spectra <strong>for</strong> each eluted<br />

chemical was found using WASP software and<br />

interface card (Graseby Dynamics, Wat<strong>for</strong>d,<br />

UK) and spectra was deconvoluted using Peak<br />

Fit ver 5.0 (SPSS Science, Chicago, IL)<br />

software to determine peak areas and<br />

positions. This procedure was used with<br />

temperatures of 250°C, 200°C, 150°C, 125°C,<br />

and 100°C <strong>for</strong> all chemicals in both solutions.<br />

Mass spectra <strong>for</strong> the ions were obtained using<br />

a diffusion vapor generator and a steady level<br />

of vapor flow into the source region of the drift<br />

tube. The last ring was floated at 400V DC and<br />

the interface plate <strong>for</strong> the MS was 200V DC<br />

enabling the facile transfer of ions to the mass<br />

spectrometer. The temperature of the drift tube<br />

and ionization region was 180°C. Air obtained<br />

from a zero air generator (Whatman) was used<br />

as drift gas and carrier gas. The mass spectra<br />

were all obtained with 500 accumulations due<br />

to the low ion concentrations and a mass range<br />

NO(g)<br />

NO(g)<br />

NO(g)<br />

NO + (H 2 O) n<br />

+<br />

+<br />

+<br />

of 10 and 250 amu.<br />

H + (H 2 O) n<br />

NH + 4 (H 2 O) n 10 mg/m 3<br />

Drift Time (ms)<br />

H 2O +<br />

+<br />

N 4<br />

e -<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

8.3 mg/m 3<br />

5.8 mg/m 3<br />

4.0 mg/m 3<br />

3.0 mg/m 3<br />

0 mg/m 3<br />

RESULTS AND DISCUSSION<br />

Reactant <strong>Ion</strong> Peaks and NO Reagent<br />

Gas<br />

The mobility spectra <strong>for</strong> positive polarity<br />

with clean air, without any NO reagent<br />

gas, at 250°C showed the expected<br />

patterns of reactant ions as identified by<br />

Karasek et al as NH 4+ (H 2O) n, NO + (H 2O) n,<br />

and H 3O + (H 2O) n where n was calculated<br />

here to be 2-3 based upon comparisons<br />

of reduced mobilities as shown in Figure<br />

1 (bottom spectrum). The intensities of<br />

NH 4+ (H 2O) n, NO + (H 2O) n, were low here<br />

compared to those of others and had<br />

reached these low levels after months of<br />

flow under only cleaned gas and<br />

temperatures of 250°C. An addition of<br />

NO(g) from 1 to 60 mg/m 3 in the<br />

ionization region of the drift tube caused<br />

and increase in intensity <strong>for</strong> the peak <strong>for</strong><br />

NO + (H 2O) n as shown (and labelled) in<br />

spectra in Figure 1. This response toward<br />

NO(g) was proportional (bottom to top spectra<br />

in Figure 1) and the increase in peak height <strong>for</strong><br />

NO + (H 2O) n was accompanied by decreases in<br />

the peak area <strong>for</strong> H 3O + (H 2O) n. This can be<br />

understood as conservation of charge with<br />

charge transferred or relocated in the<br />

NO + (H 2O) n peak; however, these results do not<br />

disclose the pathway(s) <strong>for</strong> this change.<br />

Presumably, the initial steps to <strong>for</strong>m NO + (H 2O) n,<br />

could arise by one of several reactions as<br />

shown in Equations 6-8:<br />

Though the exact pathway is not known <strong>for</strong> the<br />

<strong>for</strong>mation of NO + (H 2O) n in these studies, the<br />

practical importance is that the NO(g) could be<br />

controlled and an alternate reactant ion,<br />

NO + (H 2O) n, could replace up to 90% of the<br />

standard reactant ion H 3O + (H 2O) n.<br />

The yield or efficiency of converting NO(g) into<br />

NO + was found to be temperature dependent<br />

as shown in Figure 2 in plots of peak intensity<br />

<strong>for</strong> NO + versus NO(g) in the ion source<br />

atmosphere. At 250°C, the response curve <strong>for</strong><br />

NO + was sharp suggesting a high yield with the<br />

reaction(s) responsible <strong>for</strong> ionization of NO(g).<br />

As the temperature was decreased, the slope<br />

<strong>for</strong> plots of peak intensity versus NO(g) also<br />

NO +<br />

NO +<br />

NO +<br />

+<br />

+<br />

+<br />

H 2O<br />

2 N 2<br />

2 e -<br />

(6)<br />

(7)<br />

(8)<br />

decreased. Initially, this change was gradual;


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 27<br />

Intensity of NO +<br />

T=250°C<br />

T=225°C<br />

T=200°C<br />

T=150°C<br />

T=175°C<br />

T=100°C<br />

0 33 66 100 132 166<br />

NO(g) (mg/m 3 )<br />

T=125°C<br />

Figure 2:<br />

Peak intensity <strong>for</strong> NO + versus NO(g) at temperatures<br />

between 100 to 250°C.<br />

however, a dramatic change occurred in the<br />

region between 125 and 100°C where the<br />

response curve became nearly flat. In all these<br />

experiments, total charge in the reactant ions<br />

was conserved and the peak intensity <strong>for</strong><br />

H 3O + (H 2O) n declined as the peak intensity<br />

increased <strong>for</strong> NO + (H 2O) n. At temperatures<br />

below 125°C, attempts to <strong>for</strong>m NO + (H 2O) n as a<br />

reagent ion, even with excessive levels of<br />

reagent gas (140 mg/m 3 ), were unsuccessful<br />

as shown in Figure 3. This lack of <strong>for</strong>mation of<br />

NO + (H 2O) n at low temperature may be<br />

explained using these possibilities:<br />

1. The precursor ions needed to <strong>for</strong>m NO +<br />

were not available at low temperature. That<br />

is, the precursor ions were not <strong>for</strong>med in<br />

the source region of the drift tube.<br />

2. The NO + (H 2O) n may have been <strong>for</strong>med in<br />

the source region but underwent<br />

decomposition in the drift region. At low<br />

temperatures, residence times in the drift<br />

region increased so such losses might be<br />

pronounced at low temperatures.<br />

The first possibility was explored by adding<br />

2,4-lutidine (which <strong>for</strong>ms adduct ions) at 100°C<br />

to the drift tube even though nitric oxide<br />

reactant ions were not evident. The<br />

expectation was that even short lived reactant<br />

ions would have opportunity to react with<br />

sample vapors. No adduct ions were found<br />

and this suggested that the ions needed <strong>for</strong><br />

reactions were not being <strong>for</strong>med. The second<br />

possibility was explored by extending residence<br />

times <strong>for</strong> ions at 125°C (where NO + (H 2O) n was<br />

observable) by lowering the electric field<br />

strength on the drift region. <strong>Ion</strong>s of NO + (H 2O) n<br />

were observed even with comparatively long<br />

residence times in the drift region discounting<br />

the second possibility, namely, ion instability<br />

with increased residence times, was not too<br />

much of a consideration. Since NO + has been<br />

<strong>for</strong>med under a range of temperatures in mass<br />

spectrometry studies, the absence of NO+ in<br />

air at ambient pressure below 125°C likely<br />

occurred through the gas phase ion molecule<br />

reactions that precede or compete with<br />

<strong>for</strong>mation of NO + (H 2O) n in the ion source<br />

region. For example, ion-cluster <strong>for</strong>mation is<br />

promoted at low temperatures and necessary<br />

precursors to <strong>for</strong>m NO + in a beta source, i.e.,<br />

+<br />

N 4+ , N 2 or H 2O + may have been removed<br />

through reactions with water at 100°C at a rate<br />

faster than production of NO + was possible.<br />

Regardless of the causes <strong>for</strong> this ion behavior,<br />

these results offer practical guidelines <strong>for</strong><br />

operating IMS drift tubes with NO reagent gas:<br />

nitric oxide is not a viable reagent gas at<br />

temperatures below 125°C. Thus, further<br />

studies to ascertain the usefulness of<br />

NO + (H 2O) n as an alternate reactant ion were<br />

made with drift tube temperatures ≥125°C.<br />

<strong>Mobility</strong> Spectra <strong>for</strong> Prospective Chemical<br />

Standards in IMS Using NO Reagent Gas<br />

Reduced mobilities today are generally<br />

calculated from <strong>for</strong>mulas <strong>for</strong> electric field<br />

strength, drift tube dimensions, ambient<br />

pressure, drift tube temperature and others.<br />

Since some of these terms may have<br />

significant uncertainty, chemical standards<br />

have been proposed as a means of calibrating<br />

drift scales worldwide (14). In the discussion<br />

below, findings are described <strong>for</strong> three<br />

prospective chemical standards (15) and the<br />

effects of NO + (H 2O) n on mobility spectra and<br />

gas phase ion chemistry are shown.<br />

2,4-Lutidine (2,4-Dimethyl pyridine)<br />

The reaction chemistry of 2,4-lutidine proceeds<br />

via proton transfer to <strong>for</strong>m a protonated<br />

monomer MH + (H 2O) n (another product ion, a<br />

proton bound dimer, is thermally stable only at<br />

temperatures below 100°C and when vapor<br />

levels of 2,4-lutidine are increased). Thus, the<br />

product ion peak at 200°C as seen in Figure 4<br />

(bottom trace) was mass-identified as<br />

MH + (H 2O) n with a drift time of 6.58 ms (Table<br />

2). The calculated value <strong>for</strong> K o was 2.06<br />

cm 2 /Vsec and this comparable to accepted<br />

values of 1.95 cm 2 /Vsec; this demonstrates the<br />

magnitude of uncertainty in calculations of K o<br />

values and the value of chemical standards <strong>for</strong><br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 28<br />

calibration of mobility axes (15). Addition of<br />

NO(g) in the ion source resulted in the<br />

<strong>for</strong>mation of increased intensity of NO + (H 2O) n<br />

as shown in the second trace of Figure 4 and in<br />

a new product ion peak that is partially resolved<br />

~6.93 ms. As the amount of NO(g) was<br />

increased, both the NO + (H 2O) n and this<br />

additional product ion increased in intensity. At<br />

the highest level of NO(g) used in this<br />

experiment, the new product ion was more<br />

intense than the protonated monomer as seen<br />

in Figure 4 (top plot). The mobility <strong>for</strong> this ion<br />

(1.85 cm 2 /Vsec) was smaller than that <strong>for</strong> the<br />

protonated monomer suggesting a ion size<br />

larger than that <strong>for</strong> MH + (H 2O) n. Since NO has a<br />

strong binding energy with pyridines, an adduct<br />

ion of M*NO + was considered plausible as<br />

shown in Equation 5 and was mass identified<br />

along with(M-H) + NO as shown in Table 2. This<br />

last ion may have <strong>for</strong>med in the supersonic<br />

expansion region where collision induced<br />

dissociation is known to occur and may have<br />

arisen via Equation 9:<br />

Attempts to isolate the MH + from MNO + using<br />

the second shutter of the drift tube was<br />

unsuccessful in injection specific ions into the<br />

MS owing to close drift times <strong>for</strong> these two ions.<br />

had a drift time of 9.19 ms or K o value of 1.40<br />

cm 2 /Vsec (referenced to 2,4-lutidine). This was<br />

mass-identified as MH + as shown in Table 2.<br />

When NO(g) was used as the reagent gas,<br />

additional peaks appeared at 8.08 and 8.87 ms<br />

or K o values of 1.59 and 1.45 cm 2 /Vsec,<br />

respectively (referenced to 2,4-lutidine). These<br />

peaks increased in intensity with increased<br />

levels of nitric oxide and correspondingly, the<br />

intensity of the protonated monomer declined.<br />

The identities of these peaks were<br />

mass-identified as fragment ions of the<br />

compound and included (M-CH 3) + , M-t(butyl) + ,<br />

and M-t(butyl) + HNO (see Table 2). There was<br />

no peaks with drift times >10 ms, thus, there is<br />

no evidence <strong>for</strong> an adduct ion as seen with<br />

2,4-lutidine or rather no adduct ion had a<br />

lifetime sufficiently long to survive the drift or<br />

ion source regions. Tertiary butyl groups have<br />

been found to undergo fragmentation with<br />

NO(g) in chemical ionization MS studies and so<br />

agree with these findings.<br />

In summary, DTBP undergoes fragmentation<br />

through hydrocarbon chain losses and this<br />

apparently occurs in the reaction region (i.e.,<br />

kinetically fast compared to ion drift). This is<br />

notably more complex than 2,4-lutidine and<br />

NO M NO +<br />

(M-H)*NO +<br />

+<br />

HNO<br />

(9)<br />

cluster ion<br />

adduct ion<br />

Di-t-Butyl Pyridine<br />

The mobility spectrum <strong>for</strong> DTBP with hydrate<br />

proton reactant ions is shown in Figure 6<br />

(bottom trace) <strong>for</strong> 200°C where the product ion<br />

creates possibilities of a temperature<br />

TABLE 2:<br />

Mass-Identified product ions from IMS and IMS/MS of test chemicals with reagent ions of hydrated<br />

proton, H 3O + (H 2O) n, and of NO + (H 2O) n.<br />

Chemical<br />

2,4-Lutidine<br />

DTBP<br />

H 3O + (H 2O) n<br />

m/z (amu)<br />

108<br />

192<br />

<strong>Ion</strong><br />

MH +<br />

MH +<br />

m/z (amu)<br />

106<br />

107<br />

136<br />

137<br />

136<br />

164<br />

177<br />

192<br />

223<br />

NO + (H 2O) n<br />

<strong>Ion</strong><br />

M-H +<br />

M +<br />

M-H + (NO)<br />

MH + (NO)<br />

(M-t-butyl) +<br />

(M-t-butyl) + NO<br />

+<br />

M-CH 3<br />

MH +<br />

MH + (NO) (trace)<br />

DMMP<br />

125<br />

249<br />

MH +<br />

M 2H +<br />

125<br />

249<br />

MH +<br />

M 2H +<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 29<br />

Intensity<br />

Intensity<br />

H + (H NO + 2 O) n (H 2 O) n<br />

35.7 mg/m 3<br />

2 4 6 8 10 12 14 16 18<br />

H + (H 2 O) n<br />

NO + (H 2 O) n<br />

Drift Time (ms)<br />

2 4 6 8 10 12 14 16 18<br />

Drift Time (ms)<br />

15.0 mg/m 3<br />

4.9 mg/m 3<br />

0 mg/m 3<br />

Figure 3: <strong>Mobility</strong> spectra <strong>for</strong> 2,4-Lutidine at 200°C with<br />

varying concentrations of nitric oxide.<br />

35.7 mg/m 3<br />

15.0 mg/m 3<br />

4.9 mg/m 3<br />

0 mg/m 3<br />

Figure 4:<br />

<strong>Mobility</strong> spectra <strong>for</strong> DTBP at 200°C with varying<br />

concentrations of nitric oxide.<br />

dependence. The mobility spectra were<br />

explored up to 250°C without substantial<br />

changes in profiles.<br />

the product ion was unchanged suggesting<br />

that neither cluster ions nor fragment ions<br />

were <strong>for</strong>med. Since binding energies are<br />

not available <strong>for</strong> NO + with DMMP or related<br />

compounds, there is no precedents or<br />

supporting observations <strong>for</strong> these results.<br />

However, the ionization potential of this<br />

molecule is well above the recombination<br />

potential of nitric oxide and this makes<br />

charge exchange impossible. Molecules with<br />

ether groups have been seen to react by<br />

hydride abstraction, but the proximity of the<br />

phosphorous makes these predictions<br />

unreliable. When analyzed, DMMP showed<br />

no change in either intensity or drift time of<br />

the product peak when nitric oxide was<br />

introduced. Studies with mass-identification<br />

of the ions by IMS/MS (Table 2) confirmed<br />

that no product ions with DMMP were<br />

<strong>for</strong>med or were sufficiently stable to survive<br />

passage through the ion source and drift<br />

regions. Indeed, peak shape suggests that<br />

the any ion <strong>for</strong>med between DMMP and NO +<br />

such as M*NO + would have decomposed<br />

be<strong>for</strong>e injection into the drift region.<br />

CONCLUSIONS<br />

The reagent ion chemistry of a mobility<br />

spectrometer was changed from ordinary<br />

proton based ion chemistry to reactions<br />

based upon charge exchange, adduct<br />

<strong>for</strong>mation and hydride abstraction.<br />

Consequently, traditional drift tube might be<br />

configured <strong>for</strong> the ionization of chemicals<br />

best accomplished by these reactions rather<br />

than proton based reactions while still<br />

maintaining the common beta emitting<br />

NO + (H 2 O) n<br />

H + (H 2 O) n 35.7 mg/m 3<br />

Dimethylmethylphosphonate (DMMP)<br />

<strong>Mobility</strong> spectra <strong>for</strong> DMMP are shown in<br />

Figure 5 and the reaction chemistry with the<br />

reaction ion H + (H 2O) n is known to produce<br />

protonated monomers and proton bound<br />

dimers. A protonated monomer is apparent in<br />

Figure 5 (bottom trace) at a drift time of 6.61<br />

ms or K o values of 1.94 cm 2 /Vs and the<br />

proton bound dimer is seen at drift time of<br />

9.37 ms or K o values of 1.37 cm 2 /Vs<br />

(referenced to 2,4-lutidine). Increases in<br />

NO(g) at 250°C had the expected influence<br />

on the reactant ion peak but the drift time <strong>for</strong><br />

Intensity<br />

2 4 6 8 10 12 14 16 18<br />

Drift Time (ms)<br />

15.0 mg/m 3<br />

4.9 mg/m 3<br />

0 mg/m 3<br />

Figure 5:<br />

<strong>Mobility</strong> spectra <strong>for</strong> DMMP at 200°C with varying<br />

concentrations of nitric oxide.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


Eiceman, G.A. et al.: "Nitric Oxide as Reagent Gas ...”, IJIMS 5(2002)1, 22-30, p. 30<br />

sources such as 63-Ni. The creation of the<br />

NO + reagent ion was unfavorable below 125°C<br />

and concentrations of 10-600 µg/m 3 <strong>for</strong> NO(g)<br />

were needed in the source <strong>for</strong> other<br />

temperatures to obtain the desired ionization<br />

chemistry.<br />

REFERENCES<br />

[1] F.W. Karasek, "Plasma Chromatography", Anal.<br />

Chem. 46(8), 710A (1974)<br />

[2] S.H. Kim, K.R. Betty and F.W. Karasek, Anal.<br />

Chem. 50(14), 2006 (1978)<br />

[3] C.J. Proctor and J.F.J. Todd, Anal. Chem. 56(11),<br />

1794-1797 (1984).<br />

[4] G.A. Eiceman, Y-F. Wang, L. Garcia-Gonzalez,<br />

C.S. Harden, and D.B. Shoff, Anal. Chim. Acta<br />

306(1), 21-33 (1995).<br />

[5] F. Jelus, and B. Munson, Anal. Chem., 46, 729,<br />

(1974)<br />

[6] D. Hunt, C. N. McEwen and T. M. Harvey, Anal.<br />

Chem., 11, 1730, (1975)<br />

[7] I. Jardine and C. Fenselau, Anal. Chem., 47, 730,<br />

(1975)<br />

[8] D. Hunt and T. M. Harvey, Anal. Chem., 47, 2136,<br />

(1975)<br />

[9] N. Einolf and B. Munson, Int. J. Mass Spectrom.<br />

<strong>Ion</strong> Phys., 9, 141, (1972)<br />

[10] B. Jelus, B. Munson, and C. Fenselau, Biomedical<br />

Mass <strong>Spectrometry</strong>, 1, 96, (1974)<br />

[11] D. Hunt, T. M. Harvey, Anal. Chem., 47, 1965,<br />

(1975)<br />

[12] S.C. Subba Rao and C. Fenselau, Anal. Chem. 50,<br />

511, (1978)<br />

[13] F.W. Karasek and D.W. Denney, Anal. Chem., 46,<br />

633, (1974).<br />

[14] G.A. Eiceman, <strong>International</strong> Workshop in <strong>Ion</strong><br />

<strong>Mobility</strong> Spectometry Mesalaro NM 1992<br />

[15] J. Stone, E.G. Nazarov, and G.A. Eiceman,<br />

submitted.<br />

ACKNOWLEDGEMENTS<br />

We are grateful to Don Shoff and Steve Harden<br />

<strong>for</strong> cooperation in early studies by mass<br />

spectrometry.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


MEASURING THE TEMPERATURE OF THE DRIFT GAS IN AN ION<br />

MOBILITY SPECTROMETER: A TECHNICAL NOTE<br />

C. L. Paul Thomas 1 , N. D. Rezgui 1 , A. B. Kanu 1 , W. A. Munro 2<br />

1 DIAS, UMIST, PO BOX 88, Manchester, M60 1QD, UK.<br />

2 Graseby Dynamics Ltd, Park Avenue, Bushey, Wat<strong>for</strong>d, Hert<strong>for</strong>dshire, WD2 2BW, UK.<br />

ABSTRACT<br />

The importance of accurate temperature<br />

control in ion mobility spectrometry is<br />

emphasised along with the advantages of<br />

temperature programmed ion mobility<br />

spectrometry. The problems inherent in<br />

inferring a homogeneous temperature<br />

distribution in the drift gas from a temperature<br />

sensor mounted in the wall of the drift tube are<br />

considered. An alternative method <strong>for</strong> the<br />

measurement of drift gas temperature based<br />

upon reduced mobility standards is proposed.<br />

The drift time of the proton bound dimer of<br />

2,4-lutidine (K 0 = 1.43 cm 2 V -1 s -1 ) in the<br />

positive mode was used to measure the<br />

temperature of the drift gas as it was heated<br />

and cooled between 323 K and 473 K.<br />

Differences between the temperature of the<br />

drift gas and the temperature of the instrument<br />

varied between – 16 K and 26 K, corresponding<br />

to errors in the assignment of reduced mobility<br />

of up to 6.3 %.<br />

Other methods of measuring drift gas<br />

temperature by inserting thermocouples into<br />

the gas drift gas exhaust and pressure<br />

measurement ports were found to be<br />

unreliable. Finally, the effect of the use of<br />

heated transfer lines and gas preheaters ion<br />

the drift gas temperature was investigated and<br />

no significant effect on the temperature of the<br />

drift gas was observed through the use of<br />

either or both of these devices.<br />

Introduction<br />

Temperature is acknowledged as an essential<br />

factor in ion mobility spectrometry. Indeed<br />

useful additional in<strong>for</strong>mation about the chemical<br />

processes effecting the product ions in the<br />

mobility spectrum may be obtained by studying<br />

their mobilities at different temperatures [1].<br />

Temperature programming an ion mobility<br />

spectrometer enables experiments of rich<br />

complexity and subtlety to be undertaken to a<br />

highly efficient manner [2]. The accurate<br />

monitoring of temperature of the drift gas,<br />

there<strong>for</strong>e, is an important element in the<br />

operation of ion mobility spectrometers. Most<br />

importantly, because at a fundamental level<br />

accurate temperature measurement is needed<br />

to properly define reduced mobilities. However,<br />

generating a temperature programme,<br />

accurately and reproducibly, in the drift gas of<br />

an ion mobility spectrometer is a non-trivial<br />

exercise. Consideration of the construction of<br />

ion mobility spectrometers quickly reveals why<br />

this is so. Typically a heater surrounds the<br />

ionisation source, drift tube and detector of an<br />

ion mobility spectrometer. The sensor used to<br />

monitor and control the instrument’s<br />

temperature is normally placed close to or in<br />

direct proximity to the heating element. In using<br />

such an arrangement, the assumption is often<br />

made that the temperature of the drift gas is the<br />

same as that of the control thermocouple. This<br />

is not likely to be correct. The internal<br />

structures and mixtures of materials used to<br />

fabricate an ion mobility spectrometer, and the<br />

flows of gasses through the system coupled<br />

with the gas supply connections and insulation<br />

used, suggest that:<br />

• not only will the drift gas temperature be<br />

different to the temperature of the control<br />

thermocouple; but also that,<br />

• the drift gas temperature is likely to vary<br />

throughout the instrument.<br />

Under isothermal conditions it is possible to<br />

overcome this problem by calibrating the<br />

instrument against a standard material and<br />

deducing a cell constant, (Z (T,P)), that may be<br />

applied to enable reduced mobilities of other<br />

Received <strong>for</strong> review April 5, 2002, Accepted July 12, 2002<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>


C.L.P. Thomas et al.: Measuring the temperature of the drift gas ...”, IJIMS 5(2002)1,31-36, p. 32<br />

compounds to be estimated. In essence<br />

heterogeneity in the drift gas temperature<br />

distribution is accounted <strong>for</strong> at the same time<br />

as the heterogeneity in the electric field along<br />

the drift tube.<br />

The reduced mobility of an ion (K 0) in an ion<br />

mobility spectrometer may be related to: the<br />

observed drift time (t d); the drift length (d); the<br />

electric field strength (E); temperature (T); and,<br />

pressure (P) by the expression [3]<br />

K 0 = d<br />

Et d<br />

& 273<br />

T & p<br />

760<br />

(1)<br />

Under pressure regulated isothermal conditions<br />

this expression may be simplified to<br />

K 0 = z (t,p)<br />

t d<br />

(2)<br />

as d, E, and T and P are constant throughout<br />

the experiment. If the temperature of the<br />

system is changed then ideally another<br />

calibration is needed and a new value <strong>for</strong> the<br />

cell constant invoked. However, time and<br />

budget pressures, may tempt hard pressed<br />

laboratories to adopt a single value <strong>for</strong> a cell<br />

constant, Z, and apply a temperature and<br />

pressure correction, after allowing the system<br />

to stabilise to the new operating conditions,<br />

Equation 3.<br />

K 0 = z<br />

t d<br />

N 2<br />

& p T<br />

Labview TM data processing<br />

P, T<br />

Test atmosphere generator<br />

Figure 1:<br />

A schematic diagram of the experimental arrangement used. The ion<br />

mobility spectrometer was mounted on the top of a converted gas<br />

chromatography oven. The mixing flask was fitted inside the oven and<br />

the gas lines connecting the test atmosphere generator to the switching<br />

valve, and hence to the flask and ion mobility spectrometer were<br />

silco-steel heated to 423 K.<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

(3)<br />

<strong>Ion</strong> mobility spectrometer<br />

In such a situation, the accurate measurement<br />

of the temperature of the drift gas becomes a<br />

determining factor in the reliability of the<br />

experiment. Indeed, if temperature<br />

programming is to be used then the issue<br />

becomes more important still as the validity of<br />

the assumption that drift gas and heater<br />

temperatures are the same becomes less and<br />

less credible. Consequently, this study was<br />

concerned with studying the drift gas<br />

temperature during a temperature programmed<br />

ion mobility spectrometer experiment. The<br />

research sought to develop a calibration<br />

methodology <strong>for</strong> drift gas temperature<br />

measurement, and control, during temperature<br />

programmed ion mobility spectrometry. Further,<br />

as part of this study the effects of preheating<br />

the drift gas and using heated sample transfer<br />

lines on the temperature of the drift gas were<br />

also investigated.<br />

Instrumentation<br />

The approached used has been described<br />

be<strong>for</strong>e, [1], and Figure 1 is a diagram of the<br />

experimental system. Nitrogen (obtained from<br />

boiling off liquid nitrogen stored in a stainless<br />

steel Dewar flask) was passed through gas<br />

purification media (molecular sieve and<br />

charcoal adsorbent tubes) into a distribution<br />

manifold. Gas flows were controlled by needle<br />

valves and monitored with<br />

calibrated flowmeters. The<br />

Exhaust<br />

Heated<br />

zone<br />

Mixing flask<br />

2,4-lutidine vapour was<br />

produced using a permeation<br />

source housed in a temperature<br />

controlled stainless steel vapour<br />

generator. If required the<br />

concentration of the probe<br />

could be programmed through<br />

the use of exponential mixing; a<br />

4-way valve was used to switch<br />

the vapour in and out of the<br />

mixing flask. Permeation<br />

sources were constructed from<br />

PTFE tubing (4 cm long, o.d.<br />

6.25 mm with a wall thickness<br />

of approximately 0.8 mm.) [4].<br />

The tubing was sealed with 5<br />

mm diameter glass beads.<br />

Vapour sources were<br />

conditioned separately in a<br />

purpose built assembly,


C.L.P. Thomas et al.: Measuring the temperature of the drift gas ...”, IJIMS 5(2002)1,31-36, p. 33<br />

Table 1:<br />

Experimental and instrument parameters<br />

Parameter<br />

Heated transfer line temperature<br />

Drift gas flow rate<br />

Inlet gas flow rate<br />

Gas identity<br />

Vapour generator temperature<br />

2,4-lutidine concentration<br />

2,4-lutidine purity<br />

2,4-lutidine monomer reduced mobility<br />

2,4-lutidine dimer reduced mobility<br />

IMS cell temperature<br />

Drift tube length<br />

Inlet potential<br />

Source type<br />

Source potential<br />

Shutter 1 potential, moving grid<br />

Shutter 2 potential, fixed grid<br />

Number of field defining electrodes<br />

Potential dropped across cell<br />

Screen grid potential<br />

Detector potential<br />

Number of field defining electrodes<br />

Potential dropped across cell<br />

Screen grid potential<br />

maintained at 323 K <strong>for</strong> several days be<strong>for</strong>e<br />

they were inserted into the vapour generator.<br />

Care was taken when calibrating and<br />

transferring the permeation sources to avoid<br />

contaminating them. The conditions under<br />

which the test atmospheres were generated are<br />

summarised in Table 1.<br />

A high temperature ion mobility spectrometer<br />

was used <strong>for</strong> this work (Graseby Dynamics Ltd,<br />

Bushy, Wat<strong>for</strong>d, UK). The ceramic drift tube of<br />

the cell was approximately 4.25 cm long and<br />

the cell was capable of operation at<br />

temperatures of up to 523 K. Field defining<br />

electrode were wound around the outside of the<br />

ceramic tube, and a Type k thermocouple (RS<br />

components Ltd, Corby, Northamptonshire,<br />

NN17 9RS, UK) was attached to exterior wall<br />

drift tube approximately at the mid point in<br />

between the field defining electrodes.<br />

Surronding the whole assembly was a<br />

cylindrical heater, with an air gap of ca<br />

1 cm between the drift tube and the<br />

heater. Circuler flanges mounted at<br />

each end of the drift tube held the<br />

Faraday plate detector and screening<br />

grid at one end and the shutter<br />

electrodes, reaction region and<br />

ionisation source at the other. The<br />

cylindrical heater clamped around<br />

these two flanges. The thermocouples<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

Setting<br />

150<br />

200 to 220<br />

9 to 11.4<br />

Nitrogen<br />

50<br />

0.497<br />

99<br />

1.95<br />

1.43<br />

373 to 523<br />

4.25<br />

0<br />

63Ni, 370<br />

0<br />

60<br />

100<br />

8<br />

1000<br />

1063<br />

1100<br />

8<br />

1000<br />

1063<br />

used in this work were Type K<br />

thermocouples with a welded tip<br />

and insulated with glass fibre. The<br />

thermocouple wires were 0.3 mm<br />

°C in diameter, and with insulation the<br />

cm 3 min -1<br />

whole was 1.5 mm in diameter and<br />

cm 3 min-1<br />

approximately 2 m long. The<br />

°C<br />

temperature rating of the unit was<br />

g m -3 223 K to 533 K.<br />

% The operating parameters used<br />

cm 2 V -1 s -1 are given in Table 1. <strong>Mobility</strong><br />

cm 2 V -1 s -1 spectra were captured along with<br />

K drift tube pressure and<br />

cm temperature with a data acquisition<br />

V<br />

card (National Instruments) and<br />

MBq<br />

processed using Labviewä. <strong>Ion</strong><br />

V<br />

V mobility spectra were collected at<br />

V regular intervals and saved. The<br />

parameters used in the signal<br />

V processing can be found in Table<br />

V 2. As well as the temperature of<br />

V the exterior drift tube wall other<br />

temperature measurements of the<br />

V<br />

ion mobility spectrometer were<br />

V<br />

obtained from Type K<br />

thermocouples located :<br />

• in the pressure measurement port sited in<br />

the detector flange; and,<br />

• the drift gas exhaust port, sited in the inlet<br />

flange assembly.<br />

Given that both ends of the drift tube contained<br />

delicate structures (Shutter grids and detector<br />

assemblies <strong>for</strong> example) no attempt was made<br />

to directly insert a thermocouple in to the drift<br />

tube region.<br />

Experimental<br />

Experiments were undertaken based on a<br />

programmed temperature step. In each case<br />

the temperature of the drift tube was varied<br />

either from 473 K to 323 K or vice versa.<br />

Throughout the temperature programme<br />

temperatures and mobility spectra were<br />

recorded and analysed <strong>for</strong> the correlations<br />

between them. Figure 2 shows examples of the<br />

Table 2:<br />

Data acquisition parameters.<br />

Parameter<br />

Number of spectra averaged<br />

Number of data points sampled per scan<br />

Frequency of data acquisition<br />

Width of ion packet<br />

Delay from start of acquisition<br />

Setting<br />

100<br />

512<br />

50 KHz<br />

180 µs<br />

0 µs


C.L.P. Thomas et al.: Measuring the temperature of the drift gas ...”, IJIMS 5(2002)1,31-36, p. 34<br />

Instrument temperature / K<br />

470<br />

440<br />

410<br />

380<br />

350<br />

Heat up<br />

programme<br />

0 500 1000 1500 2000 2500 3000<br />

Time / s<br />

Figure 2:<br />

An example of two temperature programmes used in the course<br />

of this work. Note that the heating and cooling rates were not<br />

high with experiments typically lasting 30 to 40 minutes.<br />

temperature programmes used in the cooling<br />

and heating stages of the experiment.<br />

Comparison of direct temperature<br />

measurement methodologies.<br />

The first experiment compared the temperature<br />

of the drift tube thermocouple to the<br />

temperatures recorded by thermocouples in the<br />

drift-gas exhaust and pressure measurement<br />

ports. The temperature of the ion mobility<br />

spectrometer was initially set at 473 K and then<br />

programmed to reduce to 323 K.<br />

Measurement of the drift gas temperature<br />

through the use of 2,4-lutidine reduced<br />

mobility standard.<br />

Drift gas temperatures were estimated from the<br />

drift times of 2,4-lutidine, through the use of the<br />

relationship below derived from Equation 1,<br />

T = 0.359 dP<br />

EK 0 t d<br />

. (3)<br />

The ion mobility spectrometer was stabilised at<br />

a temperature of 473 K and 2, 4-lutidine<br />

introduced into the ionisation region at a<br />

constant concentration of 0.497 g m -3 . The drift<br />

tube temperature was then reduced to 363 K,<br />

while the temperature from the heater<br />

thermocouple was monitored and continuously<br />

recorded along with the pressure of the drift<br />

tube. During this time ion mobility spectra were<br />

also continuously recorded. Once the drift tube<br />

temperature had stabilised at 363 K the<br />

process was reversed, and the temperature<br />

was increased to 473 K; once again with<br />

spectra, pressure and temperature<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

Cool down<br />

programme<br />

continuously monitored. 2,4-lutidine<br />

was chosen as it a widely reported and<br />

generally accepted reduced mobility<br />

standard, with a stable reduced<br />

mobility value across the temperature<br />

range used in this work.<br />

Effects of heated gas supply and<br />

sample transfer lines<br />

In the third experiment 2,4-lutidine was<br />

used to:<br />

• investigate the effect on the<br />

temperature of the drift gas of<br />

incorporating a pre-heater in the<br />

drift gas supply line. This preheater<br />

consisted of 1m of 1/8’’ o.d. copper<br />

tubing wound around the outside of<br />

the heater that surrounded the drift<br />

tube and insulated with mineral<br />

wool; and,<br />

• connecting a heated sample transfer line to<br />

the inlet port of the instrument,<br />

approximately 20 cm long of 1/8” PTFE<br />

tubing traced with heating cord and<br />

maintained to 150 °C.<br />

In these experiments the temperature of the<br />

drift gas was monitored through the use of<br />

2,4-lutidine during temperature programmed<br />

ion mobility spectrometry in a manner similar to<br />

that described <strong>for</strong> the second experiment<br />

Results<br />

Comparison of direct temperature<br />

measurement methodologies.<br />

Figure 3 shows the relationship of the<br />

estimates of gas temperatures obtained from<br />

the thermocouples placed in the exhaust and<br />

pressure ports of the instrument during the<br />

temperature programmed cool down. It shows<br />

that the thermocouples placed in the gas<br />

streams were substantially cooler than the<br />

thermocouple mounted in the drift tube wall.<br />

Although the thermocouples were mounted in<br />

insulated gas lines the discrepancies between<br />

them and the drift tube wall are most likely to<br />

be attributable to cooling of the thermocouple<br />

junction through conduction of heat down the<br />

thermocouple wires. Note that as the<br />

temperature differential between the drift gas<br />

temperature and the ambient laboratory<br />

temperature, 296 K, reduces as the experiment<br />

progresses, this effect becomes less<br />

pronounced. Despite several modifications to<br />

this approach, involving extensive insulation of<br />

the thermocouple wires and heated sheath


C.L.P. Thomas et al.: Measuring the temperature of the drift gas ...”, IJIMS 5(2002)1,31-36, p. 35<br />

Estimated temperature / K<br />

Figure 3:<br />

A comparison of the estimated temperatures of the drift gas obtained from<br />

thermocouples located in the exhaust line directly at the point at which the drift<br />

gas exits the instrument, and the pressure measurement port located at the<br />

rear of the instrument, against the instrument temperature as indicated by the<br />

cell wall thermocouple temperature. Note the solid line indicates an exact<br />

agreement between the thermocouple in the cell wall and those positioned in<br />

the gas lines.<br />

gases, a practicable, reliable and accurate<br />

direct measurement of the drift gas<br />

temperature was not achieved. Consequently,<br />

this approach was abandoned.<br />

Measurement of the drift gas temperature<br />

with 2,4-lutidine as a reduced mobility<br />

standard.<br />

Figure 4 shows the ion mobility spectra<br />

obtained during the cool-down and heat up<br />

Signal / V<br />

(with 1.5 V offset between spectra)<br />

273<br />

273 323 373 423 473 523<br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

523<br />

473<br />

423<br />

373<br />

323<br />

8<br />

6<br />

4<br />

2<br />

Pressure Port<br />

Exhaust Port<br />

0<br />

0 5 10 15 20 25<br />

Drift time / ms<br />

463.9 K<br />

456.4 K<br />

448.0 K<br />

437.6 K<br />

426.6 K<br />

416.9 K<br />

408.0 K<br />

397.2 K<br />

387.6 K<br />

376.7 K<br />

366.9 K<br />

Cell Temperature / K<br />

Signal / V<br />

(with 1.5 V offset between spectra)<br />

Figure 4:<br />

Example ion mobility spectra obtained during the cool-down (A) and<br />

heat-up (B) stages of the experiment<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong><br />

20<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

stages of this<br />

experiment. Two peaks<br />

are discernible and these<br />

are the monomer and<br />

dimer <strong>for</strong>ms of a<br />

protonated 2,4-lutidine<br />

molecule. As the<br />

temperature reduces<br />

during cool-down,<br />

bridging between the two<br />

peaks, and peak<br />

broadening, become<br />

evident at about 380 K.<br />

These phenomena are<br />

also seen during the heat<br />

up stage, but the<br />

distortion of the peaks is<br />

evident to a higher<br />

temperature, about 415<br />

K.<br />

Perhaps, as the<br />

temperature of the<br />

instrument decreases<br />

2,4-lutidine starts to<br />

adsorb onto the inner surfaces of the<br />

instrument and the resultant build up leads to<br />

neutral species entering the drift region. During<br />

the heat-up stage of the study, the adsorbed<br />

species require significantly higher<br />

temperatures be<strong>for</strong>e they are removed from the<br />

system. (Following this experiment it was<br />

necessary to dismantle the whole of the test<br />

atmosphere generation system and clean it to<br />

remove the last traces of<br />

2,4-lutidine from the<br />

assembly. Further, the<br />

ion mobility spectrometer<br />

464.8 K<br />

455.0 K<br />

446.3 K<br />

434.3 K<br />

425.7 K<br />

415.8 K<br />

406.2 K<br />

395.1 K<br />

384.0 K<br />

377.2 K<br />

0 5 10 15 20 25<br />

Drift time / ms<br />

required several days at<br />

ca. 473 K with gas<br />

flowing through it to<br />

remove the 2, 4-lutidine<br />

residues)<br />

The<br />

pressure<br />

measurements taken<br />

during these experiments<br />

showed no sign of<br />

change as the<br />

temperature of the cell<br />

varied. The drift tube is<br />

made of a temperature<br />

resistant ceramic and as<br />

such the increase in the<br />

length of the drift tube<br />

due to thermal expansion<br />

was assumed to be


C.L.P. Thomas et al.: Measuring the temperature of the drift gas ...”, IJIMS 5(2002)1,31-36, p. 36<br />

Temperature difference / K<br />

0<br />

-20<br />

-40<br />

-60<br />

-80<br />

negligible. At this stage it is helpful to note that<br />

the electrical field controls are isolated from the<br />

drift tube, and as such were not effected by the<br />

changes in temperature of the drift tube. Thus,<br />

the change in the drift time of the<br />

2,4-lutidine-monomer peak was attributed<br />

solely to the change in the drift gas<br />

temperature.<br />

At the beginning of the run the system had<br />

been run at a stable temperature <strong>for</strong> over 25<br />

hours and as such a cell constant was derived<br />

based on Equation 3, see Equation 4.<br />

T = A t d<br />

300 350 400 450 500<br />

Transfer line and preheat<br />

Preheat<br />

Transfer line<br />

No temperature conditioning<br />

Cell Temperature / K<br />

Figure 6:<br />

Comparison of the temperature differences observed between the<br />

cell wall temperature and the drift gas temperature under different<br />

gas-temperature conditioning regimes within the experiment as the<br />

instrument was heated from 323 K to 473 K.<br />

(4)<br />

Here the term, A, denotes the cell constant<br />

used to estimate the average temperature of<br />

the drift gas in the instrument. The variation of<br />

the drift time of the 2,4-lutidine monomer during<br />

the cool-down and heat-up stages was then<br />

used to estimate the drift gas temperature.<br />

Figure 5 is a comparison of temperatures<br />

estimated from drift time against the indicated<br />

instrument temperature obtained from the<br />

thermocouple temperature mounted in the drift<br />

tube wall. These data show a significant<br />

hysteresis between the cell wall temperature<br />

and the estimated drift gas temperatures.<br />

During cool-down the drift gas is hotter than the<br />

cell wall, and the opposite is the case during<br />

heat up. Further, the differences in<br />

temperature are not trivial during cool down the<br />

difference is between 6 and 26 K, compared to<br />

a – 2 to -16 K difference on<br />

heat-up. Such errors in temperature<br />

measurement correspond to errors<br />

in the estimate of reduced mobility<br />

of between 1 % and 6.3 %.<br />

Heated gas supply and heated<br />

sample transfer lines.<br />

Figure 6 shows the effects of<br />

incorporating a heated sample<br />

transfer line, residence time<br />

approximately 2 s, or preheating the<br />

drift gas with an approximately 1 s<br />

residence time, or adopting both of<br />

these procedures on the<br />

temperature of the drift gas as<br />

determined by 2,4-lutidine during a<br />

heat-up stage. No significant effect<br />

was observed across the<br />

temperature range 50 to 200 °C in<br />

either the heat-up, or cool-down,<br />

stages. In these tests the heat<br />

capacity of the drift tube and the surrounding<br />

structures was significantly greater than the gas<br />

inputs to the system and no discernible heating<br />

or cooling effect was observed either through<br />

the use of a gas pre-heater or heated sample<br />

transfer lines.<br />

Summary<br />

The behaviours reported in the tests described<br />

above have been observed in a range of<br />

different experiments over a period of 2 years,<br />

and with other reduced mobility standards, in<br />

particular dimethylmethylphosphonate. The<br />

central finding is that the temperature of the<br />

drift gas in ion mobility spectrometers may be<br />

obtained through the use of reduced mobility<br />

standards, under conditions of highly regulated<br />

drift gas purity. Further, in the experimental<br />

system used in this study a reliable calibration<br />

could be obtained that related the cell wall<br />

temperature to the estimated drift gas<br />

temperature.<br />

References<br />

[1] Munro AM, Thomas CLP, and Lang<strong>for</strong>d ML, Anal<br />

Chim Acta., 375 (1998) 49-63.<br />

[2] Ewing RG, Eiceman GA, and Stone JA. Int. J. Mass<br />

Spec. 193 (1999) 57-68.<br />

[3] Eiceman GA and Karpas Z, “<strong>Ion</strong> <strong>Mobility</strong><br />

<strong>Spectrometry</strong>” CRC Press, Boca Raton, Fl, USA<br />

(1994).<br />

[4] Nelson GO, “Gas mixtures preparation and control,<br />

Lewis Publishers, Chelsea, MI, USA (1992).<br />

Copyright © 2002 by <strong>International</strong> Society <strong>for</strong> <strong>Ion</strong> <strong>Mobility</strong> <strong>Spectrometry</strong>

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!