20.06.2013 Views

controlled vapor deposition of azide-terminated siloxane monolayers

controlled vapor deposition of azide-terminated siloxane monolayers

controlled vapor deposition of azide-terminated siloxane monolayers

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.

CONTROLLED VAPOR DEPOSITION OF AZIDE-TERMINATED SILOXANE<br />

MONOLAYERS: A PLATFORM FOR TAILORING OXIDE SURFACES<br />

A DISSERTATION<br />

SUBMITTED TO THE DEPARTMENT OF CHEMICAL ENGINEERING<br />

AND THE COMMITTEE ON GRADUATE STUDIES<br />

OF STANFORD UNIVERSITY<br />

IN PARTIAL FULFILLMENT OF THE REQUIREMENTS<br />

FOR THE DEGREE OF<br />

DOCTOR OF PHILOSOPHY<br />

Randall Dewayne Lowe, Jr.<br />

August 2011


© 2011 by Randall Dewayne Lowe, Jr. All Rights Reserved.<br />

Re-distributed by Stanford University under license with the author.<br />

This work is licensed under a Creative Commons Attribution-<br />

Noncommercial 3.0 United States License.<br />

http://creativecommons.org/licenses/by-nc/3.0/us/<br />

This dissertation is online at: http://purl.stanford.edu/pp613jc4816<br />

ii


I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree <strong>of</strong> Doctor <strong>of</strong> Philosophy.<br />

Zhenan Bao, Primary Adviser<br />

I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree <strong>of</strong> Doctor <strong>of</strong> Philosophy.<br />

Christopher Chidsey, Primary Adviser<br />

I certify that I have read this dissertation and that, in my opinion, it is fully adequate<br />

in scope and quality as a dissertation for the degree <strong>of</strong> Doctor <strong>of</strong> Philosophy.<br />

Approved for the Stanford University Committee on Graduate Studies.<br />

Curtis Frank<br />

Patricia J. Gumport, Vice Provost Graduate Education<br />

This signature page was generated electronically upon submission <strong>of</strong> this dissertation in<br />

electronic format. An original signed hard copy <strong>of</strong> the signature page is on file in<br />

University Archives.<br />

iii


Abstract<br />

The <strong>controlled</strong> <strong>deposition</strong> <strong>of</strong> mixed, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong><br />

onto oxide surfaces provides a platform for the covalent attachment <strong>of</strong> alkyne-<br />

<strong>terminated</strong> species such as oligonucleotides using the copper-catalyzed <strong>azide</strong>-alkyne<br />

cycloaddition (CuAAC) reaction.<br />

A convenient method for the <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> dense <strong>siloxane</strong> <strong>monolayers</strong><br />

onto oxide surfaces was developed. A selectively deuterated silane, (CD3O)3-Si-<br />

(CH2)13-CH3, was synthesized and then <strong>vapor</strong> deposited in the presence <strong>of</strong> <strong>controlled</strong><br />

amounts <strong>of</strong> water <strong>vapor</strong> from the in situ dehydration <strong>of</strong> MgSO4·7H2O at 110°C.<br />

Monolayer densification was characterized using ellipsometry, Fourier transform<br />

infrared (FTIR) spectroscopy, water contact angle, and electrochemical capacitance<br />

measurements. The CD3 stretching mode <strong>of</strong> the selectively deuterated silane was<br />

monitored using FTIR spectroscopy to probe the hydrolysis <strong>of</strong> methoxy groups on<br />

adsorbed silanes while varying the <strong>deposition</strong> time, amount <strong>of</strong> water, and amount <strong>of</strong><br />

silane. The formation <strong>of</strong> dense, completely hydrolyzed <strong>monolayers</strong> required excess<br />

amounts <strong>of</strong> silane and water reactants.<br />

The optimized <strong>vapor</strong> <strong>deposition</strong> conditions were used to form mixed, <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong>. Dilution <strong>of</strong> the <strong>azide</strong> groups with unreactive methyl groups<br />

was achieved from the co-<strong>deposition</strong> <strong>of</strong> two silanes. Enrichment <strong>of</strong> the <strong>monolayers</strong><br />

with the higher <strong>vapor</strong> pressure silane reactant was observed as measured using FTIR<br />

spectroscopy, XPS, and water contact angle goniometry. The enrichment was<br />

attributed to fractionation <strong>of</strong> the silane <strong>vapor</strong> in contact with the surface from the<br />

liquid silane source. Quantitative CuAAC reactions <strong>of</strong> the <strong>azide</strong> groups in mixed<br />

iv


<strong>monolayers</strong> were demonstrated with <strong>azide</strong> surface compositions up to approximately<br />

40%, which suggested that the components in the <strong>monolayers</strong> were uniformly mixed.<br />

Steric hindrance presumably prevented the quantitative reaction <strong>of</strong> a purely <strong>azide</strong>-<br />

<strong>terminated</strong> monolayer.<br />

Alkyne-<strong>terminated</strong> oligonucleotides were covalently attached to <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong> on glass followed by hybridization with complementary<br />

oligonucleotides. However, the <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> were susceptible to the<br />

nonspecific adsorption <strong>of</strong> oligonucleotides even after thorough rinsing in aqueous<br />

detergents and buffers. A two-step reaction sequence was developed that involved one<br />

CuAAC reaction to attach oligonucleotides at low density followed by a second<br />

CuAAC reaction with ethynyl phosphonic acid to make the surface more resistant to<br />

nonspecific adsorption. Bulk and single molecule fluorescence measurements were<br />

used to characterize the glass surfaces before and after exposure to dye-labeled<br />

oligonucleotides.<br />

v


Acknowledgements<br />

I first thank my thesis advisor, Pr<strong>of</strong>essor Christopher E. D. Chidsey, for his<br />

patience and guidance in training me as a scientist and engineer over the past five<br />

years. I have learned a great deal from Pr<strong>of</strong>. Chidsey but am most thankful for his<br />

teaching me how to approach scientific problems and how to effectively communicate<br />

the results <strong>of</strong> my research. I appreciate Pr<strong>of</strong>essor Chidsey‟s entrusting me with the role<br />

<strong>of</strong> laboratory safety coordinator and his allowing me to mentor undergraduate students<br />

and local high school teachers. I also thank Pr<strong>of</strong>essor Zhenan Bao and Pr<strong>of</strong>essor Curt<br />

Frank for acting as readers on my thesis committee and for the experience from having<br />

performed research rotations in both <strong>of</strong> their laboratories during my first year <strong>of</strong><br />

graduate school. I thank Pr<strong>of</strong>essor Stacey Bent for acting as the non-reader on my<br />

committee and for the experience that I received during a teaching assistantship for an<br />

undergraduate laboratory course that she instructed. I thank Pr<strong>of</strong>essor Juan Santiago<br />

for acting as the chair <strong>of</strong> my committee and for the experience I gained while<br />

collaborating with one <strong>of</strong> his former students, Dr. Cullen Buie.<br />

I thank Dr. Tim Harris and Dr. Erin Artin for their excitement and hard work<br />

during our collaboration with Helicos BioSciences. I am thankful for having had the<br />

privilege <strong>of</strong> working with and learning from both Dr. Harris and Dr. Artin during the<br />

early years <strong>of</strong> my graduate career. I also thank Helicos BioSciences for funding.<br />

I thank Dr. Todd Eberspacher for his assistance and advice over the years. Dr.<br />

Eberspacher was particularly helpful in resolving safety issues in the laboratory. I<br />

thank the past and current members <strong>of</strong> the Chidsey group with whom I have had the<br />

privilege <strong>of</strong> working over the years: Jonathan Prange, Ali Hosseini, Josh Ratchford,<br />

vi


Charles McCrory, Anando Devadoss, Neal Devaraj, Alex Neuhausen, Jacob Woodruff,<br />

Sujatha Raghu, David Lapham, Ryan Vu, and Jim Harriss. I also thank members from<br />

the Stack research group: Matthew Pellow, Pratik Verma, Brian Smith, Eric<br />

Stenehjem, Andrew Thomas, Tim Storr, and Bolin Lin.<br />

I have learned very much from performing research and taking classes while in<br />

graduate school, but I have also learned as much or more from the many great friends I<br />

have made during this time. I thank each <strong>of</strong> them for their friendship and time. I look<br />

forward to carrying our friendships into the future.<br />

My family deserves special recognition and thanks for their continuous love<br />

and support throughout my life. You have each helped me so much, and I feel there is<br />

no way that I can ever repay you. I thank Shoko for her love and commitment while<br />

standing by me for so many years.<br />

vii


Table <strong>of</strong> Contents<br />

Abstract………………………….…………………………………………………….iv<br />

Acknowledgements………………………..………………………………………......vi<br />

Table <strong>of</strong> Contents……………………………..……………………………….…......viii<br />

List <strong>of</strong> Tables……………………………..……………………………………..….....xi<br />

List <strong>of</strong> Figures…………..………………..…………………………………….…......xii<br />

Chapter 1: Introduction………………………………………………………………...1<br />

Thesis objective……………….………………………………………………..1<br />

Background…………………………………………………………………….2<br />

Analytical Techniques………………………………………………………...10<br />

Collaborations…………….…………………….…………………………….15<br />

References..…………………………………………………………………...18<br />

Chapter 2: Deposition <strong>of</strong> Dense Siloxane Monolayers from Water and<br />

Trimethoxyorganosilane Vapor……………………..……………………………......26<br />

Preface………………………………………………………………………...26<br />

Abstract……………………………………………………………………….27<br />

Introduction…………………………………………………………………...28<br />

Experimental...………………………………………………………………..30<br />

Results………………………………………………………………………...35<br />

Discussion.……………………………………………………………………42<br />

Conclusions...…………………………………………………………………47<br />

Acknowledgements……………………………………….…………………..48<br />

Supporting Information………………………………………………….........49<br />

viii


References…………………………………………………………………….56<br />

Chapter 3: Vapor Deposition <strong>of</strong> Mixed, Azide-Terminated Siloxane Monolayers: A<br />

Modular Strategy for Modifying Oxide Surfaces..…………………………………...63<br />

Preface………………………………………………………………………...63<br />

Abstract……………………………………………………………………….64<br />

Introduction…………………………………………………………………...65<br />

Experimental………………………………………………………………….69<br />

Results………………………………………………………………………...72<br />

Discussion…………………………………………………………………….85<br />

Conclusions…………………………...………………………………………89<br />

Acknowledgements…………………………………………………………...90<br />

Supporting Information………….……………………………………………91<br />

References…..……………………………………………………………..….93<br />

Chapter 4: Attachment and Hybridization <strong>of</strong> Oligonucleotides on Azide-Terminated<br />

Siloxane Monolayers with Resistance to Nonspecific Adsorption.………………......99<br />

Preface………………………………………………………………………...99<br />

Abstract….…………………………………………………………………..100<br />

Introduction………………………………………………………………….101<br />

Experimental………………………………………………………………...105<br />

Results and Discussion………………………………………………...……109<br />

Conclusions………………………………………………………………….120<br />

Acknowledgements...………………………………………………………..120<br />

Supporting Information……………………………………………………...121<br />

ix


References……………………………………………………………….......131<br />

x


List <strong>of</strong> Tables<br />

Chapter 2, Table S1. Calculated CH3OH:H2O Ratio in Deposition Chamber<br />

Assuming Complete Hydrolysis and Condensation…….…………………………….54<br />

Chapter 4, Table S1. Nonspecific Adsorption <strong>of</strong> Dye-Labeled Mononucleotides onto<br />

Azide-Terminated Monolayers Reacted with Alkyne-Terminated Oligonucleotides and<br />

Propargyl Alcohol during Single Molecule Sequencing by Synthesis……………...129<br />

Chapter 4, Table S2. Nonspecific Adsorption <strong>of</strong> Dye-Labeled Mononucleotides onto<br />

Epoxide-Terminated Monolayers Reacted with Amine-Terminated Oligonucleotides<br />

and Phosphate during Single Molecule Sequencing by Synthesis ……………..…...130<br />

xi


List <strong>of</strong> Figures<br />

Chapter 1, Figure 1. Schematic <strong>of</strong> the platform developed in this thesis to tailor the<br />

properties <strong>of</strong> oxide surfaces for the covalent attachment and hybridization <strong>of</strong><br />

oligonucleotides with minimal nonspecific adsorption..……………..………………..3<br />

Chapter 1, Figure 2. Chemical structures <strong>of</strong> two trimethoxysilane adsorbates used in<br />

this thesis.……………..………………………………………………………………..6<br />

Chapter 2, Figure 1. Schematic <strong>of</strong> the glass vacuum chamber that was used for<br />

<strong>siloxane</strong> monolayer <strong>vapor</strong> <strong>deposition</strong>s onto oxide surfaces from liquid silane and<br />

MgSO4·7H2O reactants…………………………………...…………..………………32<br />

Chapter 2, Figure 2. Time dependence <strong>of</strong> 100 µL (CD3O)3-Si-(CH2)13-CH3<br />

monolayer <strong>vapor</strong> <strong>deposition</strong> onto silicon oxide without and with 0.5 g MgSO4·7H2O<br />

present …………………………………...…………..…………………….…………36<br />

Chapter 2, Figure 3. Carbon-deuterium and carbon-hydrogen stretching regions <strong>of</strong><br />

the p-polarized Brewster‟s angle transmission FTIR spectra <strong>of</strong> <strong>monolayers</strong> <strong>vapor</strong>-<br />

deposited onto silicon oxide for 12 h from 100 µL (CD3O)3-Si-(CH2)13-CH3 without<br />

and with 0.5 g MgSO4·7H2O present….....…………..…………………….…………37<br />

xii


Chapter 2, Figure 4. Plot <strong>of</strong> the ellipsometric thickness and the integrated FTIR<br />

symmetric CD3 stretching absorbance versus mass <strong>of</strong> MgSO4·7H2O added for 12 h<br />

<strong>deposition</strong>s onto silicon oxide with various initial volumes <strong>of</strong> (CD3O)3-Si-(CH2)13-<br />

CH3. ….....…………..…………………………………………..………….…………39<br />

Chapter 2, Figure 5. Plot <strong>of</strong> the electrochemical capacitance versus scan rate during<br />

cyclic voltammetry after 12 h <strong>deposition</strong>s <strong>of</strong> 100 µL (CD3O)3-Si-(CH2)13-CH3 onto<br />

ITO electrodes without and with 0.5 g MgSO4·7H2O present.....………….…………41<br />

Chapter 2, Figure S1. Attenuated total reflectance (ATR) FTIR spectrum <strong>of</strong> the bulk,<br />

liquid tetradecyl-tri(deuteromethoxy)silane used for <strong>siloxane</strong> monolayer <strong>vapor</strong><br />

<strong>deposition</strong>s………………………………………………….......………….…………49<br />

Chapter 2, Figure S2. NMR spectrum <strong>of</strong> tetradecyl-tri(deuteromethoxy)silane that<br />

was synthesized and used for <strong>siloxane</strong> monolayer <strong>vapor</strong> <strong>deposition</strong>s….….…………50<br />

Chapter 2, Figure S3. Plot <strong>of</strong> the current versus applied potential measured during<br />

cyclic voltammetry for a <strong>siloxane</strong> monolayer on ITO exposed to 0.1 M NaClO4 in<br />

water at a scan rate <strong>of</strong> 1000 mV/s…………………………………...….….…………51<br />

xiii


Chapter 2, Figure S4. Carbon-hydrogen stretching region <strong>of</strong> the p-polarized<br />

Brewster‟s angle transmission FTIR spectra <strong>of</strong> <strong>siloxane</strong> <strong>monolayers</strong> on silicon oxide<br />

after one 12 h <strong>vapor</strong> <strong>deposition</strong> and two repeated 12 h <strong>vapor</strong> <strong>deposition</strong>s...…………52<br />

Chapter 2, Figure S5. Time dependence <strong>of</strong> the dehydration <strong>of</strong> MgSO4·7H2O at<br />

110°C under reduced pressure starting with 0.5 g MgSO4·7H2O…..……...…………53<br />

Chapter 3, Figure 1. Modular strategy for the covalent modification <strong>of</strong> oxide surfaces<br />

using mixed, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> followed by the Cu-catalyzed<br />

<strong>azide</strong>-alkyne cycloaddition (CuAAC) reaction with terminal alkynes to attach species<br />

<strong>of</strong> interest….………………………………………………………….…...…..………68<br />

Chapter 3, Figure 2. Azide asymmetric and carbon-hydrogen stretching regions <strong>of</strong><br />

the p-polarized Brewster‟s angle transmission FTIR spectrum <strong>of</strong> a χ N3 ,liquid = 1.0<br />

monolayer on silicon oxide.………………………………………….…...…..………73<br />

Chapter 3, Figure 3. High-resolution Si 2p and N 1s XPS spectra for a χ N3 ,liquid = 1.0<br />

monolayer on silicon oxide after 12 h <strong>of</strong> <strong>vapor</strong> <strong>deposition</strong> from 100 µL <strong>of</strong> 11-<br />

azidoundecyltrimethoxysilane with 0.5 g MgSO4·7H2O present.……………………74<br />

xiv


Chapter 3, Figure 4. Time dependence <strong>of</strong> <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> χ N3 ,liquid = 1.0<br />

<strong>monolayers</strong> onto silicon oxide surfaces from 100 µL <strong>of</strong> 11-<br />

azidoundecyltrimethoxysilane with and without 0.5 g MgSO4·7H2O present…….…77<br />

Chapter 3, Figure 5. Carbon-fluorine and <strong>azide</strong> asymmetric stretching regions <strong>of</strong> the<br />

p-polarized Brewster‟s angle transmission FTIR spectrum <strong>of</strong> (a) χ N3 ,liquid = 1.0 and (b)<br />

χ N3 ,liquid = 0.75 <strong>monolayers</strong> on silicon oxide before and after 1 h <strong>of</strong> a CuAAC reaction<br />

with 4-ethynyl-α,α,α-trifluorotoluene………………………………………..…….…79<br />

Chapter 3, Figure 6. Relative surface compositions <strong>of</strong> mixed <strong>monolayers</strong> <strong>vapor</strong><br />

deposited from the azidosilane and alkyltrimethoxysilane diluents.………..…….…82<br />

Chapter 3, Figure 7. High-resolution XPS spectra <strong>of</strong> the C 1s, N 1s, and F 1s<br />

photoelectron regions for a χ N3 ,liquid = 0.75 monolayer on silicon oxide before and after<br />

1 h <strong>of</strong> a CuAAC reaction with 4-ethynyl-α,α,α-trifluorotoluene..……….…..…….…85<br />

Chapter 3, Figure S1. NMR spectrum <strong>of</strong> the synthesized 11-<br />

azidoundecyltrimethoxysilane used for <strong>siloxane</strong> monolayer <strong>vapor</strong> <strong>deposition</strong>s.......…91<br />

xv


Chapter 3, Figure S2. Chemical structure drawing <strong>of</strong> TTMA, the copper ligand used<br />

for CuAAC reactions…………………………………………………………........…92<br />

Chapter 4, Figure 1. Schematic <strong>of</strong> the platform developed for the attaching and<br />

hybridizing dye-labeled oligonucleotides on oxide surfaces.…….……………....…104<br />

Chapter 4, Figure 2. Cy3 fluorescence emission spectra <strong>of</strong> <strong>vapor</strong>-deposited <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong> on glass excited at 530 nm before and after exposure to 10 nM<br />

Cy3-RG2-amine …………………………………………………….…………....…111<br />

Chapter 4, Figure 3. Cy3 bulk fluorescence emission spectra and single molecule<br />

fluorescence images <strong>of</strong> <strong>vapor</strong>-deposited <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass before<br />

and after a 1 h exposure to Cy3-RG2-alkyne with and without the Cu(I)<br />

catalyst………………...…………………………………………….…………....…114<br />

Chapter 4, Figure 4. Cy5 fluorescence emission spectra <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong><br />

<strong>monolayers</strong> on glass exposed to hybridization conditions for Cy5-labeled<br />

oligonucleotides.……...…………….……………………………….…………....…117<br />

Chapter 4, Figure S1. Reaction scheme for the synthesis <strong>of</strong> the alkyne-<strong>terminated</strong><br />

oligonucleotide that was used in this study for CuAAC reactions with <strong>azide</strong>s.….…121<br />

xvi


Chapter 4, Figure S2. Cy3 fluorescence emission spectrum <strong>of</strong> 10 nM <strong>of</strong> a Cy3-<br />

labeled oligonucleotide, Cy3-RG2-amine, in water excited at 530 nm. Cy3<br />

fluorescence emission spectra <strong>of</strong> <strong>vapor</strong>-deposited <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on<br />

glass before and after exposure to 10 nM Cy3-alkyne with and without the Cu(I)<br />

catalyst for the CuAAC reaction………………………………………………….…123<br />

Chapter 4, Figure S3. Chemical structure <strong>of</strong> the alkyne-<strong>terminated</strong> Cy3 dye (Cy3-<br />

alkyne) that was covalently attached to <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> using the<br />

CuAAC reaction.………………………………………………………………….…124<br />

Chapter 4, Figure S4. High-resolution XPS spectra <strong>of</strong> the P 2p and N 1s<br />

photoelectron regions for an <strong>azide</strong>-<strong>terminated</strong> monolayer on silicon oxide before and<br />

after a CuAAC reaction with 10 mM ethynyl phosphonic acid…………....…….…125<br />

Chapter 4, Figure S5. Fluorescence emission spectra <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong><br />

on glass before and after a 1 h exposure to 10 nM <strong>of</strong> the Cy3-RG2-alkyne without and<br />

with Cu(I) present followed by propiolic acid CuAAC…………………....…….…127<br />

xvii


Chapter 1: Introduction<br />

Thesis objective<br />

The objective <strong>of</strong> my thesis is to develop a platform for tailoring the properties<br />

<strong>of</strong> oxide surfaces. I developed a method for the <strong>controlled</strong> <strong>deposition</strong> <strong>of</strong> self-<br />

assembled <strong>monolayers</strong> presenting <strong>azide</strong> groups at tunable density onto oxide surfaces<br />

from trimethoxysilane and water <strong>vapor</strong>. The <strong>azide</strong> groups on the surface can react with<br />

terminal alkynes using the Cu-catalyzed <strong>azide</strong>-alkyne cycloaddition reaction to<br />

covalently attach species to tailor the surface properties. I applied this general platform<br />

to the immobilization <strong>of</strong> oligonucleotides onto glass surfaces and their subsequent<br />

hybridization with complementary oligonucleotides while resisting nonspecific<br />

adsorption. This platform may enable the further development <strong>of</strong> complex surfaces for<br />

biosensing and other applications.<br />

1


Background<br />

Glass surfaces are interesting substrates for the surface immobilization <strong>of</strong><br />

species <strong>of</strong> interest because they are relatively cheap, unreactive to many chemicals,<br />

stable at elevated temperature, optically transparent, and they exhibit low background<br />

fluorescence. 1,2,3 Some technologies rely on chemistry or measurements that are most<br />

easily performed at fixed sites on surfaces. For example, many biosensors are based on<br />

the immobilization <strong>of</strong> bioanalytical probes such as oligonucleotides 4 and proteins 5 to<br />

surfaces such as glass. The work in this thesis was originally motivated by a<br />

collaboration with Helicos BioSciences, a company commercializing a next-<br />

generation, single molecule DNA sequencing by synthesis technology on glass<br />

surfaces. 6,7<br />

My goal in the collaboration was to produce a robust platform on glass<br />

surfaces for the covalent attachment <strong>of</strong> oligonucleotide targets at controllable density<br />

that could undergo hybridization and capture complementary oligonucleotides with<br />

minimal nonspecific adsorption. The inherently low background fluorescence <strong>of</strong> the<br />

high quality glass substrates needed to be maintained after the surface treatments to be<br />

useful for single molecule fluorescence measurements. The general platform used in<br />

this thesis to tailor the properties <strong>of</strong> oxide surfaces is shown schematically in Figure 1.<br />

This strategy involves the <strong>controlled</strong> <strong>deposition</strong> <strong>of</strong> mixed <strong>siloxane</strong> <strong>monolayers</strong> with<br />

tunable <strong>azide</strong> densities followed by Cu-catalyzed <strong>azide</strong>-alkyne cycloaddition<br />

(CuAAC) reactions with terminal alkynes on oligonucleotides and other species to<br />

resist nonspecific adsorption. This platform may be used for the immobilization <strong>of</strong><br />

other alkyne-<strong>terminated</strong> species besides oligonucleotides. The inspiration for this<br />

2


approach comes from the extensive work <strong>of</strong> Devaraj et al. in which mixed, <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong> adsorbed from thiols onto gold were found to be excellent<br />

platforms for tailoring electrode surfaces. 8,9,10<br />

1.<br />

2.<br />

Oxide Surface Oxide Surface<br />

Figure 1. Schematic <strong>of</strong> the platform developed in this thesis to tailor the properties <strong>of</strong><br />

oxide surfaces for the covalent attachment and hybridization <strong>of</strong> oligonucleotides with<br />

minimal nonspecific adsorption. The strategy first involves the <strong>controlled</strong> <strong>deposition</strong><br />

<strong>of</strong> mixed, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> followed by CuAAC reactions with<br />

alkyne-<strong>terminated</strong> oligonucleotides and other species (represented by the black<br />

squares) to tune the properties <strong>of</strong> the surface.<br />

3


Cu-Catalyzed Azide-Alkyne Cycloaddition (CuAAC) Reaction<br />

A key component <strong>of</strong> the platform is the CuAAC reaction, discovered<br />

independently by Sharpless and Meldal in 2002. 11,12 This cycloaddition reaction<br />

between <strong>azide</strong> groups and terminal alkyne groups is considered the „cream <strong>of</strong> the crop‟<br />

<strong>of</strong> the family <strong>of</strong> reactions termed „click‟ reactions. 13 The CuAAC reaction is an ideal<br />

candidate for surface attachment because it is selective (tolerant to other functional<br />

groups), quantitative, and proceeds rapidly in aqueous and organic solvents under<br />

ambient atmospheric conditions. The product <strong>of</strong> the CuAAC reaction is a 1,4-<br />

disubstituted 1,2,3-triazole that is thermally, oxidatively, reductively and<br />

hydrolytically stable. 14 The versatility <strong>of</strong> the CuAAC reaction has been demonstrated<br />

through the attachment <strong>of</strong> a variety <strong>of</strong> alkyne-<strong>terminated</strong> substituents such as:<br />

oligonucleotides, 10,15,4,16 proteins, 17,18,19 electroactive compounds, 9,20 fluorescent<br />

dyes, 21,22 various polymers, 23,24 and particles 25 to <strong>azide</strong>-<strong>terminated</strong> surfaces. The Cu(I)<br />

catalyst is water-soluble and can be maintained in the +1 oxidation state by use <strong>of</strong> an<br />

appropriate Cu(I)-stabilizing ligand and an appropriate water-soluble reductant such as<br />

ascorbate. Cu-free cycloaddition reactions between highly strained disubstituted<br />

alkynes and <strong>azide</strong>s have recently been discovered for biological applications in which<br />

copper cannot be tolerated, 26 but the utility <strong>of</strong> this reaction on surfaces is not explored<br />

in this thesis. The CuAAC reaction was chosen over the multitude <strong>of</strong> other interfacial<br />

chemistries that have been used to couple species to <strong>monolayers</strong> 27 based on the<br />

characteristics mentioned above.<br />

Before using the CuAAC reaction to attach oligonucleotides and other species<br />

to surfaces, I needed to first prepare <strong>azide</strong>-<strong>terminated</strong> surfaces. Azide-<strong>terminated</strong><br />

4


surfaces were chosen over alkyne-<strong>terminated</strong> surfaces because <strong>azide</strong>s can be probed<br />

relatively easily using FTIR spectroscopy and XPS. 8 By decoupling the attachment <strong>of</strong><br />

the oligonucleotides from <strong>siloxane</strong> monolayer formation, I was able to independently<br />

optimize conditions for CuAAC reactions and for the <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong> as the only reactive functionality on an otherwise passive glass<br />

or silicon oxide surface. 28<br />

Deposition <strong>of</strong> Siloxane Monolayers<br />

Siloxane <strong>monolayers</strong> can be used to tailor the properties and reactivity <strong>of</strong><br />

silicon oxide and other oxide surfaces. Since Sagiv et al. systematically studied the<br />

solution <strong>deposition</strong> <strong>of</strong> self-assembled <strong>siloxane</strong> <strong>monolayers</strong> onto glass surfaces, 29 a<br />

large amount <strong>of</strong> research has followed to better understand <strong>siloxane</strong> monolayer<br />

<strong>deposition</strong> and to use the <strong>monolayers</strong> for various applications as summarized in more<br />

thorough literature reviews. 30,31,32,28,33,34<br />

Siloxane <strong>monolayers</strong> are deposited by exposing oxide surfaces to reactants<br />

with silane headgroups for surface attachment and with compatible tailgroups such as<br />

alkyl chains that can promote self-assembly through Van der Waals attractions. 33 The<br />

tailgroups can also be <strong>terminated</strong> with organic functional groups to be presented on the<br />

surface. These functional groups should not interact strongly or bond with the oxide<br />

surface nor should they react with the silane headgroups to interfere with monolayer<br />

formation. 35 Examples <strong>of</strong> two silane reactants that are discussed in this thesis are<br />

shown in Figure 2.<br />

5


Figure 2. Chemical structures <strong>of</strong> two trimethoxysilane adsorbates used in this thesis.<br />

Alkylsilane reactants can be mono, di, or trifunctional depending on the<br />

number <strong>of</strong> hydrolyzable chloro or alkoxy groups present. 34 The formation <strong>of</strong> <strong>siloxane</strong><br />

<strong>monolayers</strong> from trimethoxysilanes, such as the ones used in this study, is a stepwise<br />

process. Trimethoxysilanes can react directly with hydroxyl groups on the surface to<br />

form <strong>siloxane</strong> linkages, or they can undergo hydrolysis reactions with water to<br />

produce hydroxyl groups on the silanes. Methanol is a byproduct <strong>of</strong> both reactions.<br />

Condensation reactions can then occur between hydroxyl groups on silanes and on the<br />

surface or between hydroxyl groups on neighboring silanes to produce <strong>siloxane</strong><br />

linkages and water. As these reactions proceed to completion, a dense <strong>siloxane</strong><br />

monolayer is formed that can be described as a two-dimensional network polymer<br />

with occasional <strong>siloxane</strong> linkages to the oxide surface. 36<br />

Two general methods for depositing <strong>siloxane</strong> <strong>monolayers</strong> are solution 29 and<br />

<strong>vapor</strong>-phase <strong>deposition</strong>. 37 The <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> <strong>siloxane</strong> <strong>monolayers</strong> was chosen<br />

over solution <strong>deposition</strong> because <strong>vapor</strong> <strong>deposition</strong>s <strong>of</strong>ten lead to <strong>monolayers</strong> with less<br />

6


particulate contamination from <strong>siloxane</strong> oligomers on the surface than solution<br />

<strong>deposition</strong>s as inferred from atomic force microscopy images. 38 This particulate<br />

contamination could interfere with monolayer formation or with subsequent surface<br />

reactions on the monolayer. Variations <strong>of</strong> solution <strong>deposition</strong> such as spin-coating<br />

have been used to produce ultrasmooth <strong>siloxane</strong> <strong>monolayers</strong>. 39<br />

Hong et al. showed that dense <strong>siloxane</strong> <strong>monolayers</strong> similar to those deposited<br />

from solution could be achieved using <strong>vapor</strong> <strong>deposition</strong>s with the deliberate addition<br />

<strong>of</strong> water as a reactant. 40 Water not only acts as a reactant during hydrolysis reactions,<br />

but it is also a product <strong>of</strong> condensation reactions between hydroxyl groups. Overall,<br />

the expected stoichiometric ratio <strong>of</strong> water to trimethoxysilane during <strong>siloxane</strong><br />

monolayer formation is 1.5 to 1 assuming complete hydrolysis and condensation. 41<br />

Therefore, control over the amount <strong>of</strong> water as a reactant should be important for the<br />

<strong>deposition</strong> <strong>of</strong> dense <strong>monolayers</strong> from trimethoxysilanes in both the solution phase and<br />

the <strong>vapor</strong> phase. Although <strong>vapor</strong> <strong>deposition</strong>s have been performed<br />

with 40,42,43,44,45,46,47,48,49 and without 50,51,52,53,54,55,56,57,58 the deliberate addition <strong>of</strong> water,<br />

the effect <strong>of</strong> water addition during <strong>vapor</strong> <strong>deposition</strong> has not been investigated to the<br />

level <strong>of</strong> detail as for solution <strong>deposition</strong>. One approach to studying the effect <strong>of</strong> water<br />

addition is to monitor the presence <strong>of</strong> unhydrolyzed chloro or alkoxy groups on<br />

adsorbed silanes under various <strong>deposition</strong> conditions.<br />

The hydrolysis <strong>of</strong> chlorosilanes has been monitored using FTIR 59,60 and<br />

XPS 61,62,63,50 to probe for the presence <strong>of</strong> Si-Cl bonds on oxide surfaces during<br />

monolayer <strong>deposition</strong>. However, there is not an equivalent method for clearly<br />

monitoring the hydrolysis <strong>of</strong> alkoxysilanes. For example, the presence <strong>of</strong><br />

7


unhydrolyzed ethoxy groups during the solution <strong>deposition</strong> <strong>of</strong> <strong>monolayers</strong> from<br />

aminopropyltriethoxysilane onto silicon oxide was inferred from peaks in FTIR<br />

spectra assigned to ethoxy groups, but these peaks were obscured by other stretching<br />

modes. 64 I sought to provide a clearer method for monitoring methoxy group<br />

hydrolysis on oxide surfaces to study the effect <strong>of</strong> water addition during <strong>vapor</strong><br />

<strong>deposition</strong>. As discussed in Chapter 2, this was accomplished by synthesizing and<br />

<strong>vapor</strong> depositing a selectively deuterated silane, (CD3O)3-Si-(CH2)13-CH3, and then<br />

monitoring C-D stretching modes attributable to the deuteromethoxy groups.<br />

Due to the low <strong>vapor</strong> pressure <strong>of</strong> alkylsilane reactants, <strong>vapor</strong> <strong>deposition</strong>s are<br />

usually performed at elevated temperature and under reduced pressure. However,<br />

experimental setups for performing <strong>vapor</strong> <strong>deposition</strong>s range widely from complex<br />

commercial vacuum systems 65,57 to simpler vessels operated at atmospheric pressure. 55<br />

For simplicity, I used a greaseless, glass vacuum desiccator into which the reactants<br />

and surfaces were placed remotely from each other. This was followed by pumping<br />

and sealing the chamber before heating it in an oven. A schematic <strong>of</strong> the chamber is<br />

provided in Chapter 2. This method was adapted from Ashurst et al. except their<br />

desiccator chamber was more complex with external reactant feeds. 42 Because there<br />

are no reactant feeds on the <strong>deposition</strong> chamber used in this study, a water source that<br />

would survive pumping to reduced pressure was needed. This problem was solved by<br />

using the in situ dehydration <strong>of</strong> MgSO4·7H2O to deliver <strong>controlled</strong> amounts <strong>of</strong> water.<br />

Mixed, Azide-Terminated Siloxane Monolayers<br />

Once conditions for <strong>vapor</strong> depositing dense, methyl-<strong>terminated</strong> <strong>siloxane</strong><br />

<strong>monolayers</strong> were established, the <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> mixed, <strong>azide</strong>-<strong>terminated</strong><br />

8


<strong>monolayers</strong> was studied. Azide-<strong>terminated</strong> <strong>monolayers</strong> are <strong>of</strong>ten formed by reacting<br />

bromide-<strong>terminated</strong> <strong>monolayers</strong> with sodium <strong>azide</strong>. 66,22 However, the nucleophilic<br />

substitution reaction does not always result in quantitative conversion <strong>of</strong> the bromides<br />

to <strong>azide</strong>s. This problem was avoided by first synthesizing and then <strong>vapor</strong> depositing an<br />

<strong>azide</strong>-<strong>terminated</strong> silane. Once a dense monolayer <strong>of</strong> <strong>azide</strong> groups is formed, steric<br />

hindrance may prevent their quantitative conversion using the CuAAC reaction<br />

depending on the alkyne-<strong>terminated</strong> species. 8 Therefore, the dilution <strong>of</strong> the <strong>azide</strong><br />

groups with unreactive groups such as methyl groups may be desirable.<br />

Mixed <strong>siloxane</strong> <strong>monolayers</strong> are not only attractive for controlling the spacing<br />

between groups on the surface, but they can also be used to add additional complexity<br />

to oxide surfaces for more precise surface engineering. Binary mixed <strong>monolayers</strong> are<br />

usually formed by the simultaneous co-<strong>deposition</strong> <strong>of</strong> two adsorbates from<br />

solution, 67,35,68,69 but they can also be formed by sequential <strong>deposition</strong>. 70,71,72 I opted to<br />

use the co-<strong>deposition</strong> <strong>of</strong> trimethoxysilane adsorbates from the <strong>vapor</strong> phase to study<br />

mixed monolayer <strong>deposition</strong> as reported in Chapter 3. The ability to probe the relative<br />

surface compositions <strong>of</strong> mixed <strong>monolayers</strong> is desirable because the surface<br />

compositions are <strong>of</strong>ten different than the compositions <strong>of</strong> the initial reactant mixtures<br />

from which they were deposited. 73,69,68,74<br />

Another method for producing mixed <strong>monolayers</strong> involves the <strong>deposition</strong> <strong>of</strong> a<br />

pure or mixed monolayer <strong>terminated</strong> with reactive functional groups followed by<br />

surface reactions with more than one species. 73,75 This approach to forming mixed<br />

<strong>monolayers</strong> is used in Chapter 4 to attach oligonucleotides and other species to resist<br />

nonspecific adsorption. Common methods for reducing the nonspecific adsorption <strong>of</strong><br />

9


iomolecules include rinsing in buffers and detergents 76 and the addition <strong>of</strong> functional<br />

groups or molecules known to resist nonspecific adsorption. 77,78,79,80,81,82,83,84 These<br />

functionalities are <strong>of</strong>ten hydrophilic in nature and may not be compatible as tailgroup<br />

terminations for <strong>controlled</strong> <strong>siloxane</strong> monolayer <strong>deposition</strong>. The platform developed in<br />

this thesis enables the attachment <strong>of</strong> functionalities that are incompatible with<br />

<strong>controlled</strong> monolayer <strong>deposition</strong> onto oxide surfaces.<br />

Analytical Techniques<br />

Brewster’s Angle Transmission Fourier Transform Infrared (FTIR)<br />

Spectroscopy<br />

FTIR spectroscopy is a surface-sensitive technique that I have used to identify<br />

chemical functional groups on silicon oxide surfaces after various treatments such as<br />

monolayer <strong>vapor</strong> <strong>deposition</strong>s and CuAAC reactions. FTIR spectra were obtained by<br />

transmitting p-polarized IR light through planar, double-side polished silicon oxide<br />

surfaces that were oriented at the Brewster‟s angle for silicon, 74°, 85 using a custom-<br />

built sample holder. Infrared light is absorbed by samples as vibrational modes are<br />

excited, and the absorbance <strong>of</strong> the impinging light as a function <strong>of</strong> its frequency in<br />

wavenumbers is measured with a spectrometer. A deuterated triglycine sulfate<br />

(DTGS) detector was used for all measurements. Chabal et al. popularized the use <strong>of</strong><br />

Brewster‟s angle transmission FTIR spectroscopy on silicon surfaces. 86<br />

X-ray Photoelectron Spectroscopy (XPS)<br />

XPS is a surface-sensitive spectroscopy that typically yields elemental<br />

information about a surface. XPS spectra are obtained by irradiating a surface with x-<br />

ray photons in an ultrahigh vacuum chamber. For all spectra in this thesis, a<br />

10


monochromator was used to select Al Kα incident radiation, which has a photon<br />

energy <strong>of</strong> 1486.6 electron volts (eV). The impinging photons excite electrons from<br />

core energy levels into vacuum where they pass through an analyzer that selects the<br />

electrons based upon their kinetic energies before sending them to a detector.<br />

Although the spectrometer actually measures the kinetic energies <strong>of</strong> the photoemitted<br />

electrons, spectra are usually plotted in units <strong>of</strong> electron counts versus binding energy.<br />

The binding energy <strong>of</strong> the electrons is calculated by subtracting the kinetic energy<br />

measured from the incident photon energy, 1486.6 eV for Al Kα incident radiation.<br />

Survey spectra are obtained by measuring electron counts in the 0-1100 eV<br />

binding energy range, where most elements exhibit one or more photoelectron peaks.<br />

High-resolution XPS spectra are used to more accurately determine binding energies<br />

in order to make inferences about the chemical environment around a particular<br />

element in the sample. For example, the binding energy <strong>of</strong> an atom will exhibit a<br />

slightly greater binding energy if the atom is bound to a more electron-withdrawing<br />

species. 87<br />

Fluorescence Spectroscopy<br />

Fluorescence spectroscopy was used to detect the presence <strong>of</strong> fluorescent dyes<br />

on glass surfaces in this thesis. The fluorescent dyes used in this study, Cyanine 3<br />

(Cy3) and Cyanine 5 (Cy5), absorb and emit light in the visible region <strong>of</strong> the<br />

electromagnetic spectrum. A fluorescent dye can absorb a photon <strong>of</strong> specific energy<br />

which results in the excitation <strong>of</strong> an electron from its ground state into an excited<br />

singlet state that may be composed <strong>of</strong> a number <strong>of</strong> vibrational states. Once in an<br />

excited state, the dye undergoes vibrational relaxation to the lowest vibrational level <strong>of</strong><br />

11


the excited state. Next, the excited dye can relax back to any vibrational level <strong>of</strong> the<br />

ground electronic state. This results in the fluorescent emission <strong>of</strong> a photon<br />

corresponding to the energy difference between the two states. Because the electron<br />

relaxes to the lowest vibrational level <strong>of</strong> the excited state before fluorescing and<br />

returning to the ground state, the emitted light is usually lower in energy than the<br />

excitation light. This phenomenon is known as the Stokes shift and can be measured<br />

by performing excitation and emission scans <strong>of</strong> the fluorescent dye. 88<br />

Fluorescence spectra were obtained by exposing glass surfaces to excitation<br />

light from a 450 watt Xenon lamp source after passing through a monochromator and<br />

a bandpass filter. Emitted light was collected using front-face detection from the<br />

surface after passing the light through another bandpass filter to reduce stray<br />

excitation light and another set <strong>of</strong> monochromators before reaching a detector. Front-<br />

face detection <strong>of</strong> the emitted light was used to reduce the detection <strong>of</strong> reflected and<br />

scattered excitation light.<br />

Ellipsometry<br />

Ellipsometry is an optical measurement that can be used to determine the<br />

thickness <strong>of</strong> a thin layer on a surface. 33 A helium-neon laser emits 632.8 nm<br />

unpolarized light that passes through a polarizer to become plane-polarized with the<br />

polarization vector oriented 45° from the sample surface. Plane-polarized light<br />

consists <strong>of</strong> s- and p-polarized light components, which are characterized by the<br />

orientations <strong>of</strong> their polarization vectors being either parallel (s) or perpendicular (p)<br />

with respect to the surface. The plane polarized light undergoes a single reflection<br />

from the surface with the s- and p-polarized components being reflected to different<br />

12


extents and with different phase delays. After reflection, the relative phase and<br />

amplitude <strong>of</strong> the s- and p-components <strong>of</strong> the light are determined. Ellipsometry<br />

measures two parameters, Ψ and Δ, which are related to the complex refractive indices<br />

<strong>of</strong> the film and <strong>of</strong> the substrate. Ψ is the angle, the arctangent <strong>of</strong> which is the ratio <strong>of</strong><br />

the amplitudes <strong>of</strong> the s- and p-polarized components, and Δ is the shift in the phase<br />

between the s- and p-polarized components before and after reflection from the<br />

surface. For films with thicknesses less than 50 Å, such as the <strong>monolayers</strong> in this<br />

thesis, the thickness can be calculated from the measured Ψ and Δ values with an<br />

appropriate estimate <strong>of</strong> the film refractive index. A film refractive index <strong>of</strong> 1.46 was<br />

used for all calculations in this thesis. The thickness calculation is normally performed<br />

as part <strong>of</strong> the measurement by the ellipsometry s<strong>of</strong>tware on the instrument computer.<br />

Water Contact Angle Goniometry<br />

Water contact angle goniometry can be used to probe the relative<br />

hydrophobicity or hydrophilicity <strong>of</strong> a surface. In practice, a water droplet is dispensed<br />

from a needle in close contact to the surface and the water contact angle between the<br />

solid surface and the bottom <strong>of</strong> the sessile droplet is quickly measured on both sides <strong>of</strong><br />

the droplet using a measuring telescope with a protractor in the eye piece. Young‟s<br />

equation describes the balance <strong>of</strong> forces between a surface and a water droplet,<br />

cos <br />

(1)<br />

LV<br />

where γLV, γSV, γSL, are the interfacial surface tensions at the liquid-<strong>vapor</strong>, solid-<strong>vapor</strong>,<br />

and solid-liquid interfaces, respectively. 33 A contact angle <strong>of</strong> 180° represents no<br />

interaction with the surface, while a contact angle <strong>of</strong> 0° indicates spreading <strong>of</strong> the<br />

liquid.<br />

13<br />

SV<br />

SL


Electrochemical Capacitance Measurements<br />

Electrochemical capacitance measurements were performed to probe the<br />

packing density <strong>of</strong> the alkyl chains in <strong>vapor</strong>-deposited <strong>monolayers</strong>. Since a conductive<br />

substrate is needed to perform this measurement, <strong>monolayers</strong> were <strong>vapor</strong> deposited<br />

onto indium tin oxide (ITO) surfaces, a conductive metal oxide. Cyclic voltammetry<br />

was used to estimate the electrochemical capacitance <strong>of</strong> <strong>monolayers</strong> <strong>vapor</strong> deposited<br />

onto ITO. Cyclic voltammetry experiments were conducted in a three-electrode<br />

electrochemical cell in which the exposed area was defined by a cylindrically-bored<br />

Teflon cone pressed onto the surface. The surface was exposed to an electrolyte<br />

solution <strong>of</strong> 0.1 M NaClO4 in water inside the cone. During cyclic voltammetry, a<br />

potential is applied to the ITO working electrode relative to the Ag/AgCl/saturated<br />

KCl reference electrode while electrical current into the working electrode is<br />

measured. 89 The potential is swept at a scan rate with units <strong>of</strong> potential (volts) per unit<br />

time (seconds). A potentiostat is used to measure the current while controlling the<br />

potential and scan rate during cyclic voltammetry. The resulting cyclic voltammogram<br />

is typically plotted in units <strong>of</strong> current density (microamps per unit area) versus<br />

potential (V). Electrochemical capacitances were calculated by averaging the<br />

magnitudes <strong>of</strong> the currents measured at 0.1 V during the forward and reverse scans<br />

and then dividing by both the scan rate and the electrode area.<br />

An electrical double layer is formed at the solid-solution interface during<br />

cyclic voltammetry. The double layer is modeled by Helmholtz theory as a parallel<br />

plate capacitor:<br />

14


εε0<br />

C (2)<br />

d<br />

where C is the capacitance per unit area, d is the effective thickness <strong>of</strong> the separating<br />

medium, ε is the dielectric constant <strong>of</strong> the medium, and ε0 is the vacuum<br />

permittivity. 90 If a monolayer is not densely-packed and has defects, the effective<br />

barrier thickness, d, will be less than expected from calculating molecular lengths <strong>of</strong><br />

the adsorbates because ions in the electrolyte solution can penetrate through the<br />

defects to the electrode surface resulting in a higher measured capacitance. Less<br />

permeation <strong>of</strong> ions through the monolayer and therefore a lower capacitance will be<br />

measured if the monolayer is densely-packed and defect-free.<br />

Collaborations<br />

I have had the privilege <strong>of</strong> collaborating with several people during the course<br />

<strong>of</strong> my doctoral research as outlined in this section.<br />

The cyclic voltammetry measurements that were used to estimate the<br />

electrochemical capacitances reported in Chapter 2 were performed in collaboration<br />

with Mr. Matthew Pellow, a graduate student in Pr<strong>of</strong>. Dan Stack‟s research group. I<br />

was present and assisted during all <strong>of</strong> the cyclic voltammetry experiments. Mr. Pellow<br />

and Mr. Eric Stenehjem, another graduate student in the Stack group, have been using<br />

my method to <strong>vapor</strong> deposit <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> onto ITO electrodes as<br />

platforms for covalently attaching alkyne-<strong>terminated</strong>, inorganic metal ligand<br />

complexes for electrocatalytic screening.<br />

The work in Chapter 3 was performed in collaboration with Ms. Sujatha Raghu,<br />

a local middle school teacher who enthusiastically performed research in the Chidsey<br />

15


group over two summers in 2006 and 2008. Ms. Raghu was involved with the first<br />

experiments that I performed on <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> as her first day in the<br />

Chidsey group was also my first day. Ms. Raghu was particularly involved with the<br />

ellipsometry and water contact angle measurements.<br />

This thesis is the result <strong>of</strong> a collaboration that was established between Dr. Tim<br />

Harris then at Helicos BioSciences and Pr<strong>of</strong>. Chidsey at Stanford before I joined the<br />

Chidsey group. As a result <strong>of</strong> this collaboration with Helicos, I have had the privilege<br />

<strong>of</strong> working with two outstanding collaborators, Dr. Erin Artin and Dr. Tim Harris. Dr.<br />

Harris acted as a co-advisor for me during the collaboration and provided technical<br />

guidance and support in parallel to my principal advisor, Pr<strong>of</strong>. Chidsey. Dr. Artin<br />

synthesized alkyne-<strong>terminated</strong> oligonucleotides and fluorescent dyes and shipped<br />

them to me for experiments at Stanford. My role in the collaboration was the<br />

development <strong>of</strong> best practices for the reproducible <strong>deposition</strong> <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong><br />

<strong>monolayers</strong> onto glass substrates while maintaining a low fluorescent background. I<br />

used bulk fluorescence measurements at Stanford to optimize CuAAC reaction<br />

conditions for the covalent attachment <strong>of</strong> oligonucleotides and to develop methods to<br />

reduce nonspecific adsorption. Technology developed at Stanford was transferred to<br />

Helicos through onsite visits and correspondence. Dr. Harris and Dr. Artin performed<br />

more sensitive single molecule fluorescence measurements on glass surfaces using<br />

instruments at Helicos.<br />

Mr. David Lapham, an undergraduate student in the Chidsey group, showed<br />

that the covalently attached oligonucleotides were viable for hybridization with<br />

complementary oligonucleotides as measured using bulk fluorescence spectroscopy.<br />

16


Furthermore, Dr. Artin performed single molecule sequencing runs at Helicos on glass<br />

surfaces modified with <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong>. Preliminary experiments<br />

suggested that the <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> exhibited higher resistance to the<br />

nonspecific adsorption <strong>of</strong> dye-labeled nucleotides after reaction with propargyl alcohol<br />

than the commercially-produced, epoxy-<strong>terminated</strong> <strong>monolayers</strong> that Helicos uses as<br />

their standard substrate. Unfortunately, this collaboration came to an abrupt and<br />

untimely end in December 2008. Chapter 4 discusses some <strong>of</strong> the work performed<br />

during the collaboration with Helicos.<br />

Mr. Minsub Chung, a graduate student in Pr<strong>of</strong>. Steve Boxer‟s group in the<br />

Chemistry department at Stanford, established an interesting collaboration with me. I<br />

attached alkyne-<strong>terminated</strong> oligonucleotides and ethynyl phosphonic acid to <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong> on glass surfaces. Mr. Chung used these surfaces as supports<br />

for tethered lipid bilayers by rupturing vesicles using hybridization between<br />

complementary oligonucleotides displayed on the surface and on the vesicles. This<br />

work is not discussed in this thesis but resulted in a publication on which I am second<br />

author.<br />

I also collaborated with a Chidsey group student, Mr. Jonathan Prange, on the<br />

solution <strong>deposition</strong> <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> phosphonic acid <strong>monolayers</strong> onto ITO<br />

electrode surfaces. I synthesized p-azidophenylphosphonic acid and performed initial<br />

experiments with it. Mr. Prange followed up on this work upon joining the Chidsey<br />

group, which is discussed in detail in his thesis.<br />

17


References<br />

1. Vong, T.; ter Maat, J.; van Beek, T. A.; van Lagen, B.; Giesbers, M.; van Hest,<br />

J. C. M.; Zuilh<strong>of</strong>, H. Langmuir 2009, 25, 13952-13958.<br />

2. Laib, S.; Krieg, A.; Rankl, M.; Seeger, S. Applied Surface Science 2006, 252,<br />

7788-7793.<br />

3. Zammatteo, N.; Jeanmart, L.; Hamels, S.; Courtois, S.; Louette, P.; Hevesi, L.;<br />

Remacle, J. Analytical Biochemistry 2000, 280, 143-150.<br />

4. Seo, T. S.; Bai, X.; Ruparel, H.; Li, Z.; Turro, N. J.; Ju, J. PNAS 2004, 101,<br />

5488-5493.<br />

5. Pascal Jonkheijm; Weinrich, D.; Schröder, H.; Niemeyer, Christ<strong>of</strong> M.;<br />

Waldmann, H. Angew. Chem. Int. Ed. 2008, 47, 9618-9647.<br />

6. Harris, T. D.; Buzby, P. R.; Babcock, H.; Beer, E.; Bowers, J.; Braslavsky, I.;<br />

Causey, M.; Colonell, J.; DiMeo, J.; Efcavitch, J. W.; Giladi, E.; Gill, J.; Healy, J.;<br />

Jarosz, M.; Lapen, D.; Moulton, K.; Quake, S. R.; Steinmann, K.; Thayer, E.; Tyurina,<br />

A.; Ward, R.; Weiss, H.; Xie, Z. Science 2008, 320, 106-109.<br />

7. Fuller, C. W.; Middendorf, L. R.; Benner, S. A.; Church, G. M.; Harris, T.;<br />

Xiaohua, H.; Jovanovich, S. B.; Nelson, J. R.; Schloss, J. A.; Schwartz, D. C.;<br />

Vezenov, D. V. Nature Biotechnology 2009, 27, 1013-1023.<br />

8. Collman, J. P.; Devaraj, N. K.; Eberspacher, T. P. A.; Chidsey, C. E. D.<br />

Langmuir 2006, 22, 2457-2464.<br />

9. Collman, J. P.; Devaraj, N. K.; Decreau, R. A.; Yang, Y.; Yan, Y.-L.; Ebina,<br />

W.; Eberspacher, T. A.; Chidsey, C. E. D. Science 2007, 315, 1565-1568.<br />

18


10. Devaraj, N. K.; Miller, G. P.; Ebina, W.; Kakaradov, B.; Collman, J. P.; Kool,<br />

E. T.; Chidsey, C. E. D. J. Am. Chem. Soc. 2005, 127, 8600-8601.<br />

11. Tornoe, C. W.; Christensen, C.; Meldal, M. Journal <strong>of</strong> organic chemistry 2002,<br />

67, 3057.<br />

12. Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angewandte<br />

Chemie (International ed.) 2002, 41, 2596-2599.<br />

13. Wang, Q.; Chittaboina, S.; Barnhill, H. N. Letters in Organic Chemistry 2005,<br />

2, 293-301.<br />

14. Meldal, M.; Tornoe, C. W. Chem. Rev. 2008, 108, 2952-3015.<br />

15. Fischler, M. M. Chemical communications 2007, 169.<br />

16. Chung, M.; Lowe, R. D.; Chan, Y.-H. M.; Ganesan, P. V.; Boxer, S. G.<br />

Journal <strong>of</strong> Structural Biology 2009, 168, 190-199.<br />

17. Gauchet, C.; Labadie, G. R.; Poulter, C. D. Journal <strong>of</strong> the American Chemical<br />

Society 2006, 128, 9274-9275.<br />

18. Po-Chiao, L.; Shau-Hua, U.; Mei-Chun, T.; Jia-Ling, K.; Kuo-Ting, H.; Sheng-<br />

Chieh, Y.; Avijit Kumar, A.; Yu-Ju, C.; Chun-Cheng, L. Angewandte Chemie<br />

International Edition 2006, 45, 4286-4290.<br />

19. Schlossbauer, A.; Schaffert, D.; Kecht, J.; Wagner, E.; Bein, T. Journal <strong>of</strong> the<br />

American Chemical Society 2008, 130, 12558-12559.<br />

20. Collman, J. P.; Devaraj, N. K.; Chidsey, C. E. D. Langmuir 2004, 20, 1051-<br />

1053.<br />

21. Rozkiewicz, D. I.; Janacuteczewski, D.; Verboom, W.; Ravoo, B. J.;<br />

Reinhoudt, D. N. Angew. Chem. Int. Ed. 2006, 45, 5292-5296.<br />

19


22. Claudia, H.; Stephanie, H.; Ulrich, S. S. Nanotechnology 2008, 19, 035703.<br />

23. Prakash, S.; Long, T. M.; Selby, J. C.; Moore, J. S.; Shannon, M. A. Analytical<br />

Chemistry 2007, 79, 1661-1667.<br />

24. Ostaci, R.-V.; Damiron, D.; Capponi, S.; Vignaud, G.; Leger, L.; Grohens, Y.;<br />

Drockenmuller, E. Langmuir 2008, 24, 2732-2739.<br />

25. Im, S. G.; Kim, B.-S.; Tenhaeff, W. E.; Hammond, P. T.; Gleason, K. K. Thin<br />

Solid Films 2009, 517, 3606-3611.<br />

26. Codelli, J. A.; Baskin, J. M.; Agard, N. J.; Bertozzi, C. R. JACS 2008.<br />

27. Li, J.; Thiara, P. S.; Mrksich, M. Langmuir 2007, 23, 11826-11835.<br />

28. Haensch, C.; Hoeppener, S.; Schubert, U. S. Chemical Society Reviews 2010,<br />

39, 2323-2334.<br />

29. Sagiv, J. J. Am. Chem. Soc. 1980, 102, 92-98.<br />

30. Ulman, A. Chem. Rev. 1996, 96, 1533-1554.<br />

31. Schwartz, D. K. Annu. Rev. Phys. Chem. 2001, 52, 107-137.<br />

32. Onclin, S.; Ravoo, B. J.; Reinhoudt, D. N. Angew. Chem., Int. Ed. 2005, 44,<br />

6282-6304.<br />

33. Ulman, A., An Introduction to Ultrathin Organic Films: From Langmuir-<br />

Blodgett Films to Self-Assembly. Academic Press: San Diego, CA, 1991.<br />

34. Plueddemann, E. P., Silane Coupling Agents. 2nd ed.; Plenum Press: New<br />

York, 1991; p 253.<br />

35. Fryxell, G. E.; Rieke, P. C.; Wood, L. L.; Engelhard, M. H.; Williford, R. E.;<br />

Graff, G. L.; Campbell, A. A.; Wiacek, R. J.; Lee, L.; Halverson, A. Langmuir 1996,<br />

12, 5064-5075.<br />

20


36. Brzoska, J. B.; Azouz, I. B.; Rondelez, F. Langmuir 1994, 10, 4367-4373.<br />

37. Haller, I. J. Am. Chem. Soc. 1978, 100, 8050-8055.<br />

38. Jung, G. Y.; Li, Z.; Wu, W.; Chen, Y.; Olynick, D. L.; Wang, S. Y.; Tong, W.<br />

M.; Williams, R. S. Langmuir 2005, 21, 1158-1161.<br />

39. Ito, Y.; Virkar, A. A.; Mannsfeld, S.; Oh, J. H.; Toney, M.; Locklin, J.; Bao, Z.<br />

Journal <strong>of</strong> the American Chemical Society 2009, 131, 9396-9404.<br />

40. Hong, J.; Porter, D. W.; Sreenivasan, R.; McIntyre, P. C.; Bent, S. F. Langmuir<br />

2007, 23, 1160-1165.<br />

41. Loy, D. A.; Baugher, B. M.; Baugher, C. R.; Schneider, D. A.; Rahimian, K.<br />

Chemistry <strong>of</strong> Materials 2000, 12, 3624-3632.<br />

42. Ashurst, W. R.; Carraro, C.; Maboudian, R.; Frey, W. Sens. Actuators, A 2003,<br />

104, 213-221.<br />

43. Mayer, T. M.; de Boer, M. P.; Shinn, N. D.; Clews, P. J.; Michalske, T. A. J.<br />

Vac. Sci. Technol. B 2000, 18, 2433-2440.<br />

44. H<strong>of</strong>fmann, P. W.; Stelzle, M.; Rabolt, J. F. Langmuir 1997, 13, 1877-1880.<br />

45. Fiorilli, S.; Rivolo, P.; Descrovi, E.; Ricciardi, C.; Pasquardini, L.; Lunelli, L.;<br />

Vanzetti, L.; Pederzolli, C.; Onida, B.; Garrone, E. J. Colloid Interface Sci. 2008, 321,<br />

235-241.<br />

46. Crampton, N.; Bonass, W. A.; Kirkham, J.; Thomson, N. H. Langmuir 2005,<br />

21, 7884-7891.<br />

47. Saini, G.; Sautter, K.; Hild, F. E.; Pauley, J.; Linford, M., R. J. Vac. Sci.<br />

Technol., A 2008, 26, 1224-1234.<br />

21


48. Ek, S.; Iiskola, E. I.; Niinisto, L.; Vaittinen, J.; Pakkanen, T. T.; Keranen, J.;<br />

Auroux, A. Langmuir 2003, 19, 10601-10609.<br />

49. Kurth, D. G.; Bein, T. Langmuir 1995, 11, 3061-3067.<br />

50. Fadeev, A. Y.; McCarthy, T. J. Langmuir 2000, 16, 7268-7274.<br />

51. Anderson, A. S.; Dattelbaum, A. M.; Montano, G. A.; Price, D. N.; Schmidt, J.<br />

G.; Martinez, J. S.; Grace, W. K.; Grace, K. M.; Swanson, B. I. Langmuir 2008, 24,<br />

2240-2247.<br />

52. Ressier, L.; Viallet, B.; Grisolia, J.; Peyrade, J. P. Ultramicroscopy 2007, 107,<br />

980-984.<br />

53. Hozumi, A.; Ushiyama, K.; Sugimura, H.; Takai, O. Langmuir 1999, 15, 7600-<br />

7604.<br />

54. Du, Y.; George, S. M. J. Phys. Chem. C 2007, 111, 8509-8517.<br />

55. Li, J.-R.; Garno, J. C. Nano Lett. 2008, 8, 1916-1922.<br />

56. Li, J.-R.; Lusker, K. L.; Yu, J.-J.; Garno, J. C. ACS Nano 2009, 3, 2023-2035.<br />

57. Zhang, F.; Sautter, K.; Larsen, A. M.; Findley, D. A.; Davis, R. C.; Samha, H.;<br />

Linford, M. R. Langmuir 2010, 26, 14648-14654.<br />

58. George, A.; Blank, D. H. A.; ten Elsh<strong>of</strong>, J. E. Langmuir 2009, 25, 13298–<br />

13301.<br />

59. Tripp, C. P.; Hair, M. L. Langmuir 1991, 7, 923-927.<br />

60. Tripp, C. P.; Hair, M. L. Langmuir 1995, 11, 149-155.<br />

61. Angst, D. L.; Simmons, G. W. Langmuir 1991, 7, 2236-2242.<br />

62. Silberzan, P.; Leger, L.; Ausserre, D.; Benattar, J. J. Langmuir 1991, 7, 1647-<br />

1651.<br />

22


63. Wasserman, S. R.; Tao, Y.-T.; Whitesides, G. M. Langmuir 1989, 5, 1074-<br />

1087.<br />

64. Pasternack, R. M.; Rivillon Amy, S.; Chabal, Y. J. Langmuir 2008, 24, 12963-<br />

12971.<br />

65. Prakash, S.; Long, T. M.; Selby, J. C.; Moore, J. S.; Shannon, M. A. Analytical<br />

Chemistry 2007, 79, 1661-1667.<br />

66. Lummerstorfer, T.; H<strong>of</strong>fmann, H. J. Phys. Chem. B 2004, 108, 3963-3966.<br />

67. Offord, D. A.; Griffin, J. H. Langmuir 2002, 9, 3015-3025.<br />

68. Tong, Y.; Tyrode, E.; Osawa, M.; Yoshida, N.; Watanabe, T.; Nakajima, A.;<br />

Ye, S. Langmuir 2011, 27, 5420-5426.<br />

69. Heise, A.; Stamm, M.; Rauscher, M.; Duschner, H.; Menzel, H. Thin Solid<br />

Films 1998, 327-329, 199-203.<br />

70. Fadeev, A. Y.; McCarthy, T. J. Langmuir 1999, 15, 7238-7243.<br />

71. Choi, I.; Kim, Y.; Kang, S. K.; Lee, J.; Yi, J. Langmuir 2006, 22, 4885-4889.<br />

72. Kumar, N.; Maldarelli, C.; Steiner, C.; Couzis, A. Langmuir 2001, 17, 7789-<br />

7797.<br />

73. Azam, M. S.; Fenwick, S. L.; Gibbs-Davis, J. M. Langmuir 2011, 27, 741-750.<br />

74. Bain, C. D.; Evall, J.; Whitesides, G. M. Journal <strong>of</strong> the American Chemical<br />

Society 1989, 111, 7155-7164.<br />

75. Hudalla, G. A.; Murphy, W. L. Langmuir 2010, 26, 6449-6456.<br />

76. Vandenberg, E.; Elwing, H.; Askendal, A.; Lundström, I. Journal <strong>of</strong> Colloid<br />

and Interface Science 1991, 143, 327-335.<br />

23


77. Jung, S.; Angerer, B.; Löscher, F.; Niehren, S.; Winkle, J.; Seeger, S.<br />

ChemBioChem 2006, 7, 900-903.<br />

78. Holland, N. B.; Qiu, Y.; Ruegsegger, M.; Marchant, R. E. Nature 1998, 392,<br />

799-801.<br />

79. Vikholm-Lundin, I.; Piskonen, R.; Albers, W. M. Biosensors and<br />

Bioelectronics 2007, 22, 1323-1329.<br />

80. Alexander, N. A.; Igor, Y. S.; Philip, B. O. In Effect <strong>of</strong> surface electrostatics<br />

on adsorption behavior and biospecific interactions <strong>of</strong> DNA oligonucleotides, 1999;<br />

Gerald, E. C.; John, C. O., Eds. SPIE: 1999; pp 170-177.<br />

81. Kuga, S.; Yang, J.-H.; Takahashi, H.; Hirama, K.; Iwasaki, T.; Kawarada, H.<br />

Journal <strong>of</strong> the American Chemical Society 2008, 130, 13251-13263.<br />

82. Cárdenas, M.; Campos-Terán, J.; Nylander, T.; Lindman, B. Langmuir 2004,<br />

20, 8597-8603.<br />

83. Prime, K. L.; Whitesides, G. M. Science 1991, 252, 1164-1167.<br />

84. Herne, T. M.; Tarlov, M. J. Journal <strong>of</strong> the American Chemical Society 1997,<br />

119, 8916-8920.<br />

85. Michalak, D. J.; Rivillon, S.; Chabal, Y. J.; Estève, A.; Lewis, N. S. J. Phys.<br />

Chem. B 2006, 110, 20426-20434.<br />

86. Chabal, Y., J. ; Hines, M. A.; Feijoo, D. J. Vac. Sci. Technol., A 1995, 13,<br />

1719-1727.<br />

87. Cicero, R. L. Organic Monolayers on Silicon Surfaces: Synthesis,<br />

Characterization, and Studies Toward Their Effect on Charge Injection at the<br />

Silicon/Organic Solid Interface Stanford University, 2000.<br />

24


88. Lapin, N. A. Chemical and Molecular Functionalization <strong>of</strong> Silicon Surfaces for<br />

Biosensing Applications. Rutgers, The State University <strong>of</strong> New Jersey, 2009.<br />

89. Bard, A. J.; Faulkner, L. R., Electrochemical Methods: Fundamentals and<br />

Applications. Second ed.; John Wiley and Sons: 2004; p 833.<br />

90. Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D. J. Am. Chem. Soc.<br />

1987, 109, 3559-3568.<br />

25


Chapter 2: Deposition <strong>of</strong> Dense Siloxane Monolayers from Water and<br />

Trimethoxyorganosilane Vapor*<br />

Preface<br />

This chapter <strong>of</strong> my dissertation is from a manuscript that was recently accepted<br />

to the scientific journal Langmuir. I am the first author <strong>of</strong> this publication and am<br />

responsible for most <strong>of</strong> the writing and preparation. The other authors are Matthew A.<br />

Pellow, Pr<strong>of</strong>essor T. Daniel P. Stack, and Pr<strong>of</strong>essor Christopher E. D. Chidsey. I was<br />

involved in all <strong>of</strong> the experiments that were conducted for this manuscript. My main<br />

contributions to this manuscript include: pursuing <strong>vapor</strong> <strong>deposition</strong>, synthesis <strong>of</strong><br />

tetradecyltri(deuteromethoxy)silane, using MgSO4·7H2O as a <strong>controlled</strong> water source<br />

and studying the effect <strong>of</strong> its addition, and designing the sample holder for Brewster‟s<br />

angle transmission FTIR spectroscopy.<br />

Pr<strong>of</strong>. Chidsey proposed using cyclic voltammetry to estimate electrochemical<br />

capacitance as a measure <strong>of</strong> monolayer densification on indium tin oxide (ITO)<br />

electrodes. Matthew A. Pellow performed the cyclic voltammetry experiments, but I<br />

was present and assisted with all experiments. Pr<strong>of</strong>. Chidsey and I developed the idea<br />

<strong>of</strong> using the selectively deuterated silane for studying the hydrolysis <strong>of</strong><br />

trimethoxysilanes. Pr<strong>of</strong>. Chidsey also provided significant guidance in interpreting and<br />

planning experiments as well as in the writing and preparation <strong>of</strong> the manuscript.<br />

* Reproduced with permission from Lowe, R.D.; Pellow, M.A.; Stack, T.D.P.;<br />

Chidsey, C.E.D. “Deposition <strong>of</strong> Dense Siloxane Monolayers from Water and<br />

Trimethoxyorganosilane Vapor,” Langmuir, 2011, 27(16), 9928-9935. ©Copyright<br />

2011 American Chemical Society.<br />

26


Abstract<br />

A convenient, laboratory-scale method for the <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> dense<br />

<strong>siloxane</strong> <strong>monolayers</strong> onto oxide substrates was demonstrated. This method was<br />

studied and optimized at 110°C under reduced pressure with the <strong>vapor</strong> <strong>of</strong><br />

tetradecyltri(deuteromethoxy)silane, (CD3O)3-Si-(CH2)13-CH3, and water from the<br />

dehydration <strong>of</strong> MgSO4·7H2O. Ellipsometric thicknesses, water contact angles, Fourier<br />

transform infrared (FTIR) spectroscopy, and electrochemical capacitance<br />

measurements were used to probe monolayer densification. The CD3 stretching mode<br />

in the FTIR spectrum was monitored as a function <strong>of</strong> the <strong>deposition</strong> time and amounts<br />

<strong>of</strong> silane and water reactants. This method probed the unhydrolyzed methoxy groups<br />

on adsorbed silanes. Excess silane and water were necessary to achieve dense,<br />

completely hydrolyzed <strong>monolayers</strong>. In the presence <strong>of</strong> sufficient silane, an excess <strong>of</strong><br />

water above the calculated stoichiometric amount was necessary to hydrolyze all<br />

methoxy groups and achieve dense <strong>monolayers</strong>. The excess water was partially<br />

attributed to the reversibility <strong>of</strong> the hydrolysis <strong>of</strong> the methoxy groups.<br />

27


Introduction<br />

We have developed a convenient procedure for <strong>vapor</strong>-depositing dense <strong>siloxane</strong><br />

<strong>monolayers</strong> onto oxide surfaces and a Fourier transform infrared (FTIR) spectroscopic<br />

probe for monitoring the hydrolysis <strong>of</strong> methoxy groups during <strong>deposition</strong>. Siloxane<br />

<strong>monolayers</strong> have been used in several applications ranging from chromatography 1 to<br />

the surface attachment <strong>of</strong> bioanalytical probes. 2,3 These <strong>monolayers</strong> control the<br />

properties <strong>of</strong> oxide surfaces by introducing organic functionality. 4-6 This is<br />

accomplished using silane monomers with headgroups that bond to the surface and<br />

crosslink among themselves and tailgroups that present desired chemical moieties. The<br />

ability to investigate and control the underlying interfacial chemistry allows the<br />

reproducible <strong>deposition</strong> <strong>of</strong> <strong>siloxane</strong> <strong>monolayers</strong> <strong>of</strong> <strong>controlled</strong> density suitable for<br />

specific applications.<br />

Siloxane monolayer formation is a multistep process beginning with organosilane<br />

monomers possessing one to three hydrolyzable groups, such as chloro or alkoxy<br />

groups. 7 These hydrolyzable groups react directly with surface hydroxyl groups to<br />

form <strong>siloxane</strong> or other silicon to metal oxide linkages. The hydrolyzable groups also<br />

react with water to produce silanols on the organosilanes. 8 These silanols can undergo<br />

condensation reactions with surface hydroxyl groups or with silanols on neighboring<br />

organosilanes to form additional <strong>siloxane</strong> linkages and water as a byproduct.<br />

Monolayers formed from trifunctional silanes, such as the ones used in this study,<br />

have been described as two-dimensional, network polymers with occasional bonds to<br />

the oxide surface. 9<br />

28


Because reaction <strong>of</strong> the hydrolyzable groups should be necessary to form surface-<br />

attached, two-dimensional network polymers, the amount <strong>of</strong> water is known to be an<br />

important variable for the reproducible <strong>deposition</strong> <strong>of</strong> <strong>siloxane</strong> <strong>monolayers</strong> from<br />

trifunctional silane monomers. With insufficient water available, unhydrolyzed alkoxy<br />

groups on the silanes could sterically hinder dense monolayer formation. This would<br />

be similar to the case <strong>of</strong> mon<strong>of</strong>unctional alkyldimethylsilanes where the bulky<br />

dimethylsilyl groups prevented tight alkane chain packing. 10<br />

Numerous studies have discussed the importance <strong>of</strong> water in depositing <strong>siloxane</strong><br />

<strong>monolayers</strong> from solutions. 4,11,12 When sufficient water was present, larger<br />

ellipsometric thicknesses 10,13 and contact angles 14 were measured and indicated the<br />

<strong>deposition</strong> <strong>of</strong> denser <strong>monolayers</strong> compared to more anhydrous cases. However, those<br />

measurements did not provide detailed information about the chemical bonds at the<br />

interface. X-ray photoelectron spectroscopy (XPS) 10,14-16 and FTIR spectroscopy 17,18<br />

have been used to confirm the hydrolysis <strong>of</strong> Si-Cl bonds during solution <strong>deposition</strong> <strong>of</strong><br />

trichlorosilanes. Using FTIR spectroscopy, Pasternack et al. found unhydrolyzed<br />

ethoxy groups after solution <strong>deposition</strong> <strong>of</strong> triethoxysilanes under anhydrous<br />

conditions. 19 However, the FTIR peaks assigned to residual C-H stretching modes <strong>of</strong><br />

the ethoxy groups were obscured by peaks associated with Si-O-Si and other C-H<br />

stretching modes. In this paper, we describe a clearer method <strong>of</strong> monitoring the<br />

hydrolysis <strong>of</strong> alkoxy groups. Deuteration <strong>of</strong> hydrolyzable methoxy groups gives a C-D<br />

stretching mode that is shifted into an unobscured region <strong>of</strong> FTIR spectra.<br />

Siloxane <strong>monolayers</strong> have been deposited from dilute solutions 20 and from the <strong>vapor</strong><br />

phase. 21 Monolayer <strong>deposition</strong> from the <strong>vapor</strong> phase is attractive because there is less<br />

29


<strong>siloxane</strong> oligomer particle contamination than during solution-phase <strong>deposition</strong>. 22-24<br />

As in solution-phase <strong>deposition</strong>, the amount <strong>of</strong> water is an important variable for<br />

<strong>vapor</strong>-phase <strong>siloxane</strong> monolayer <strong>deposition</strong>. Vapor-phase <strong>deposition</strong> with 22,25-33 and<br />

without 16,34-42 the deliberate addition <strong>of</strong> water <strong>vapor</strong> has been reported, but the effect<br />

<strong>of</strong> water addition has not been studied as thoroughly as in the solution case.<br />

In this paper, we present a simple method for <strong>vapor</strong>-depositing <strong>siloxane</strong> <strong>monolayers</strong><br />

with <strong>controlled</strong> amounts <strong>of</strong> water <strong>vapor</strong> and show that dense <strong>siloxane</strong> <strong>monolayers</strong> are<br />

not deposited unless sufficient water is present to hydrolyze the methoxy groups and<br />

sufficient silane is present to densify the <strong>monolayers</strong>. The hydrolysis <strong>of</strong><br />

tetradecyltri(deuteromethoxy)silane, (CD3O)3-Si-(CH2)13-CH3, is monitored after<br />

<strong>vapor</strong>-phase <strong>deposition</strong> using FTIR spectroscopy on silicon oxide surfaces.<br />

Specifically, we study how the <strong>deposition</strong> duration and the initial amounts <strong>of</strong> water<br />

and silane reactants effect methoxy group hydrolysis and monolayer completion.<br />

Ellipsometric thicknesses, FTIR spectroscopy, water contact angles, and<br />

electrochemical capacitance measurements are used to probe monolayer densification.<br />

Experimental<br />

Silane Synthesis<br />

Tetradecyltri(deuteromethoxy)silane was prepared from n-tetradecyltrichlorosilane<br />

(Gelest) via methanolysis <strong>of</strong> the trichloro groups following published procedures for<br />

making similar alkoxysilanes. 43,44 2.00 g <strong>of</strong> tetradecyltrichlorosilane (6 mmol) in 15<br />

mL <strong>of</strong> dry pentane was slowly added to a solution <strong>of</strong> 0.67 g (19 mmol) deuterated<br />

methanol, CD3OD (99.6+ atom % D, Acros), and 1.85 g (19 mmol) <strong>of</strong> dry pyridine in<br />

75 mL <strong>of</strong> dry pentane while stirring under nitrogen and cooled in an ice water bath.<br />

30


The solution was allowed to stir overnight at room temperature under a nitrogen<br />

atmosphere, filtered to remove pyridinium chloride, and then concentrated by rotary<br />

e<strong>vapor</strong>ation. The crude product was purified by two successive vacuum distillations.<br />

See the Supporting Information for nuclear magnetic resonance (NMR) and FTIR<br />

characterization.<br />

Surface Cleaning<br />

Double and single-side polished Si(100) wafers were used as received or after prior<br />

<strong>deposition</strong> experiments and cleaned in an oxygen plasma (GaLa Instrumente Plasma<br />

Prep5) for 10 min prior to <strong>deposition</strong> using 50 percent relative power with a dioxygen<br />

flow rate <strong>of</strong> 100 std. mL/min at a pressure <strong>of</strong> 0.25 mbar. These surfaces were exposed<br />

to atmospheric air for approximately 15 min after cleaning in order to determine<br />

effective refractive indices <strong>of</strong> the substrates with an ellipsometer. This cleaning<br />

method resulted in hydrophilic surfaces with water contact angles approaching 0°.<br />

Indium tin oxide (ITO) films supported on glass substrates had sheet resistances <strong>of</strong> 8-<br />

12 ohms (Delta Technologies). The ITO surfaces were plasma cleaned with the same<br />

procedure as the silicon oxide surfaces.<br />

Monolayer Vapor Deposition<br />

Siloxane <strong>monolayers</strong> were <strong>vapor</strong>-deposited onto silicon oxide and ITO surfaces in<br />

o-ring sealed, glass vacuum desiccators fitted with Teflon stopcocks (Jencons part<br />

number 250-048) with internal volumes <strong>of</strong> about 600 mL (Figure 1). These chambers<br />

were oven dried for a minimum <strong>of</strong> 4 h at 140°C in air and cooled for no more than 10<br />

min in air before use. Neat silane (100 µL) was pipetted onto and absorbed into 42.5<br />

mm diameter Whatman filter paper in the bottom <strong>of</strong> the desiccator. Various masses <strong>of</strong><br />

31


MgSO4·7H2O (Fisher biochemical grade) were placed in a foil boat in the bottom <strong>of</strong><br />

the desiccator as a water source for the hydrolysis reaction. Plasma-cleaned surfaces<br />

were placed on a metal rack supported above the silane liquid and hydrated salt in the<br />

glass vacuum chamber. The desiccator was then evacuated through a rubber hose on a<br />

glass vacuum line with a liquid nitrogen-trapped mechanical pump for approximately<br />

60 seconds. The final pressure at the trap was 1 Torr. The Teflon valve on the<br />

desiccator was closed, and the chamber was placed in a 110°C preheated oven (Forma<br />

Scientific) for various periods <strong>of</strong> time from 1 to 24 h. After <strong>deposition</strong>, the valve was<br />

opened to ambient air to return the chamber to atmospheric pressure and remove the<br />

samples. Ashurst et al. used a related but more elaborate setup. 25 Their chamber was<br />

modified with water and silane reactant inputs from external sources similar to other<br />

systems that have been used to introduce water during <strong>vapor</strong> <strong>deposition</strong>s.<br />

Deposition Chamber<br />

Oxide Surfaces<br />

RSi(OCD 3 ) 3 (l) MgSO 4 ·7H 2 O(s)<br />

32<br />

H 2 O(g)<br />

RSi(OCD 3 ) 3 (g)<br />

- CD 3 OH(g)<br />

Figure 1. Schematic <strong>of</strong> the glass vacuum chamber that was used for <strong>siloxane</strong><br />

monolayer <strong>vapor</strong> <strong>deposition</strong>s onto oxide surfaces from liquid silane and MgSO4·7H2O<br />

reactants. An incomplete monolayer with unhydrolyzed methoxy groups is shown in<br />

the schematic in the middle. A dense, completely hydrolyzed monolayer (drawing


adapted from Ulman 8 ) deposited in the presence <strong>of</strong> sufficient silane and MgSO4·7H2O<br />

is shown on the right.<br />

Treatment <strong>of</strong> Monolayers with Methanol Vapor<br />

Substrates onto which <strong>monolayers</strong> had been previously <strong>vapor</strong> deposited were placed<br />

in a clean, dry desiccator chamber. The glass desiccator was evacuated to 1 Torr, and<br />

the Teflon valve was closed as in the monolayer <strong>deposition</strong> experiments. A rubber<br />

septum was then placed on the Teflon valve. The septum was then punctured with a<br />

needle on a syringe containing pure deuterated methanol or a premixed deuterated<br />

methanol-water liquid solution. The Teflon valve on the vacuum desiccator was then<br />

opened, which evacuated the contents <strong>of</strong> the syringe into the desiccator under vacuum.<br />

The Teflon valve was then closed, and the rubber septum and syringe were removed<br />

from the top <strong>of</strong> the desiccator. The desiccator was then placed into a preheated 110°C<br />

oven for 12 h.<br />

Surface Characterization<br />

FTIR spectra were obtained with a Bruker Vertex 70 spectrometer using a KBr<br />

beam splitter and a deuterated triglycine sulfate (DTGS) detector. Spectra were<br />

collected in transmission mode through the silicon with the samples at the silicon<br />

Brewster‟s angle (74°) 45 using p-polarized light from a wire-grid polarizer (Specac<br />

KRS-5). For each spectrum, 1024 scans were collected at 4 cm -1 resolution.<br />

Background spectra were collected from freshly oxygen plasma-cleaned, double side<br />

polished Si(100) crystals in the same orientation.<br />

Ellipsometry measurements were performed using a Gaertner L116 ellipsometer at<br />

70° angle <strong>of</strong> incidence. Effective real and imaginary refractive indices <strong>of</strong> the Si(100)<br />

33


substrates were determined immediately after plasma cleaning but before <strong>siloxane</strong><br />

monolayer <strong>vapor</strong> <strong>deposition</strong>. Measurements after <strong>vapor</strong> <strong>deposition</strong>, rinsing with<br />

toluene and isopropyl alcohol, and drying in a stream <strong>of</strong> nitrogen were used to<br />

determine monolayer thicknesses. A film refractive index <strong>of</strong> 1.46 was used for all<br />

thickness calculations. At least three measurements were performed on each sample to<br />

obtain average substrate refractive indices or sample thicknesses.<br />

Advancing sessile drop water contact angles were measured with a Ramé-Hart<br />

model 100 goniometer using water from a four-bowl Millipore purification system.<br />

Water droplets were dispensed from a syringe with a flat-tipped needle. Fresh water<br />

was obtained from the Millipore system for each measurement session. At least two<br />

measurements were made on each sample on both sides <strong>of</strong> the droplets and averaged.<br />

Cyclic voltammetry experiments were conducted using either a Bioanalytical<br />

Systems model CV-50W or a Pine Instrument Company model AFCBP1 potentiostat.<br />

ITO surfaces with an exposed area <strong>of</strong> 0.21 cm 2 were used as the working electrodes in<br />

a three-electrode electrochemical cell. Platinum wire was used as the counter electrode,<br />

and all potentials measured were with respect to a Ag/AgCl/saturated KCl reference<br />

electrode. Cyclic voltammograms were collected over a scan range <strong>of</strong> -0.1 V to 0.4 V<br />

versus the reference electrode (Figure S3 in the Supporting Information). The reported<br />

electrochemical capacitances were calculated by averaging the magnitudes <strong>of</strong> the<br />

currents measured at 0.1 V during the forward and reverse scans and then dividing by<br />

both the scan rate and the electrode area. Measurements were performed under<br />

ambient laboratory conditions using an electrolyte solution <strong>of</strong> 0.1 M NaClO4 in water.<br />

34


Results<br />

Time Dependence <strong>of</strong> Ellipsometric Thickness, C-D Stretch and Water Contact<br />

Angle<br />

Figure 2a is a plot <strong>of</strong> ellipsometric thickness versus <strong>deposition</strong> time for <strong>monolayers</strong><br />

<strong>vapor</strong>-deposited from 100 µL tetradecyltri(deuteromethoxy)silane, without and with<br />

0.5 g MgSO4·7H2O present. At each time point, the ellipsometric thicknesses are<br />

greater with the hydrated salt than without. A limiting thickness is reached at 12 h in<br />

both cases, but at different values; 14.1 Å without MgSO4·7H2O and 18.2 Å with<br />

MgSO4·7H2O.<br />

Figure 3 shows that a peak is present at 2073 cm -1 in the FTIR spectrum after a 12 h<br />

<strong>deposition</strong> with 100 µL silane but without MgSO4·7H2O. This peak is assigned to the<br />

symmetric CD3 stretching mode <strong>of</strong> the deuteromethoxy groups 46,47 and is used to<br />

monitor methoxy group hydrolysis. With 0.5 g MgSO4·7H2O present during<br />

<strong>deposition</strong>, the CD3 absorbance is no longer observed.<br />

Figure 2b shows the integrated absorbance <strong>of</strong> the CD3 mode at various <strong>deposition</strong><br />

times. The CD3 absorbance is observed at all time points without MgSO4·7H2O<br />

present, although it decreases with increasing <strong>deposition</strong> time until reaching a limiting<br />

value at about 4 h. In contrast, the CD3 absorbance is not observed at any time point<br />

with 0.5 g MgSO4·7H2O.<br />

Figure 2c is a plot <strong>of</strong> the water contact angle versus <strong>deposition</strong> time. Before<br />

<strong>deposition</strong>, the freshly plasma-cleaned silicon oxide surfaces exhibit water contact<br />

angles approaching zero. After <strong>deposition</strong>, water contact angles are greater than 100°<br />

in all cases after 1 h, but are lower for <strong>monolayers</strong> <strong>vapor</strong>-deposited without<br />

35


MgSO4·7H2O than for those deposited with MgSO4·7H2O. The contact angles plateau<br />

after 12 h at 104° without MgSO4·7H2O and 109° with MgSO4·7H2O.<br />

Integrated CD 3 Absorbance / cm -1<br />

20<br />

16<br />

12<br />

8<br />

4<br />

0<br />

1.6x10 -3<br />

1.2x10 -3<br />

8.0x10 -4<br />

4.0x10 -4<br />

Water Contact Angle / °<br />

Ellipsometric Thickness / Å<br />

0<br />

120<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

(a)<br />

(b)<br />

(c)<br />

without MgSO 4 ·7H 2 O<br />

with 0.5 g MgSO 4 ·7H 2 O<br />

0 5 10 15 20 25<br />

36<br />

Time / h<br />

Figure 2. Time dependence <strong>of</strong> 100 µL (CD3O)3-Si-(CH2)13-CH3 monolayer <strong>vapor</strong><br />

<strong>deposition</strong> onto silicon oxide without and with 0.5 g MgSO4·7H2O present monitored


using (a) ellipsometric thickness, (b) integrated area <strong>of</strong> the CD3 FTIR symmetric<br />

stretch, and (c) water contact angle. The lines are guides to the eye. Error bars<br />

represent one standard deviation.<br />

C-H Stretch<br />

Figure 3 also shows the C-H stretching region for <strong>siloxane</strong> <strong>monolayers</strong> formed by<br />

12 h <strong>vapor</strong> <strong>deposition</strong>s. The intensities <strong>of</strong> the peaks in the 2800-3000 cm -1 range are<br />

lower for <strong>deposition</strong>s performed without MgSO4·7H2O than for those with<br />

MgSO4·7H2O. In addition, the peak frequencies for the symmetric and asymmetric<br />

methylene stretches shift down 3 cm -1 from 2854 and 2924 cm -1 without MgSO4·7H2O<br />

to 2851 and 2921 cm -1 with 0.5 g MgSO4·7H2O.<br />

Absorbance<br />

5.0x10 -5<br />

0<br />

without MgSO 4 ·7H 2 O<br />

with 0.5 g MgSO 4 ·7H 2 O<br />

2000 2050 2100 2150 2800 2850 2900 2950 3000<br />

Frequency / cm -1<br />

37<br />

Frequency / cm -1<br />

1.0x10 -3<br />

5.0x10 -4<br />

Figure 3. Carbon-deuterium and carbon-hydrogen stretching regions <strong>of</strong> the p-<br />

polarized Brewster‟s angle transmission FTIR spectra <strong>of</strong> <strong>monolayers</strong> <strong>vapor</strong>-deposited<br />

onto silicon oxide for 12 h from 100 µL (CD3O)3-Si-(CH2)13-CH3 without and with 0.5<br />

g MgSO4·7H2O present.<br />

0<br />

Absorbance


Effect <strong>of</strong> Mass <strong>of</strong> MgSO4·7H2O<br />

The plateau <strong>of</strong> the ellipsometric thickness at long <strong>deposition</strong> time with 100 L<br />

silane varies with the mass <strong>of</strong> MgSO4·7H2O (Figure 4a, circles). As the mass <strong>of</strong><br />

MgSO4·7H2O is increased, the ellipsometric thickness increases until a plateau is<br />

reached with 0.5 g MgSO4·7H2O. Figure 4b shows the integrated absorbance <strong>of</strong> the<br />

CD3 symmetric stretch versus the mass <strong>of</strong> MgSO4·7H2O. The mass required to<br />

decrease the CD3 stretch to zero matches the mass required to reach the plateau in<br />

ellipsometric thickness.<br />

Effect <strong>of</strong> Amount and Surface Area <strong>of</strong> the Silane Source<br />

To test whether the amount <strong>of</strong> silane limits monolayer growth, experiments were<br />

performed with varying volumes <strong>of</strong> silane. Above 50 µL <strong>of</strong> silane, the ellipsometric<br />

thickness is independent <strong>of</strong> the amount <strong>of</strong> silane and follows the dependence on<br />

MgSO4·7H2O described above (Figure 4a). However, with only 10 µL silane, the<br />

thickness is equal to the thickness observed with 100 L silane in the absence <strong>of</strong><br />

MgSO4·7H2O but does not increase with added MgSO4·7H2O. Reduced CD3<br />

absorbances are observed for 50 µL and less silane with 0.05 g MgSO4·7H2O present<br />

as shown in Figure 4b. With 0.5 g MgSO4·7H2O, the CD3 absorbance is reduced to the<br />

level <strong>of</strong> noise for all silane volumes.<br />

To test for any role <strong>of</strong> the surface area <strong>of</strong> the liquid silane, its exposed area was<br />

reduced by at least a thousand fold by placing the silane into a glass capillary sealed at<br />

one end and held with the open end upwards during <strong>deposition</strong>. Experiments using 10<br />

µL silane with 0.05 g MgSO4·7H2O and 100 µL silane with 0.5 g MgSO4·7H2O yield<br />

38


ellipsometric thicknesses and CD3 integrated absorbances equal to the values when the<br />

silane is absorbed into filter paper.<br />

Figure 4. (a) Plot <strong>of</strong> the ellipsometric thickness versus mass <strong>of</strong> MgSO4·7H2O added<br />

for 12 h <strong>deposition</strong>s onto silicon oxide with various initial volumes <strong>of</strong> (CD3O)3-Si-<br />

(CH2)13-CH3. (b) Integrated FTIR symmetric CD3 stretching absorbance versus mass<br />

<strong>of</strong> MgSO4·7H2O added for 12 h <strong>deposition</strong>s with various initial volumes <strong>of</strong> the silane.<br />

The lines are guides to the eye. Error bars represent one standard deviation.<br />

39


Treatment <strong>of</strong> Monolayers with Methanol Vapor<br />

To assess the reversibility <strong>of</strong> the hydrolysis <strong>of</strong> the methoxy groups, completely<br />

hydrolyzed, partial <strong>monolayers</strong> deposited from 10 µL silane and 0.05 g MgSO4·7H2O<br />

for 12 h were treated with a large excess <strong>of</strong> pure CD3OD or mixtures <strong>of</strong> CD3OD and<br />

H2O <strong>vapor</strong> at 110°C for 12 h. 48 After treatment with pure CD3OD <strong>vapor</strong>, CD3<br />

stretching is observed with an integrated absorbance <strong>of</strong> 3.1x10 -4 cm -1 and an<br />

ellipsometric thickness <strong>of</strong> 14 Å. Exposure to 10:1 and 3:1 molar ratios <strong>of</strong> CD3OD:H2O<br />

result in CD3 integrated absorbances <strong>of</strong> 2.3x10 -4 and<br />

1.1x10 -4 cm -1 , respectively. Negligible CD3 absorbance is observed after 12 h<br />

exposure to 1:1 CD3OD:H2O.<br />

Electrochemical Capacitance<br />

To assess the permeability <strong>of</strong> <strong>monolayers</strong> formed by silane <strong>vapor</strong> <strong>deposition</strong>, cyclic<br />

voltammetry was performed on conductive ITO samples that were plasma cleaned and<br />

<strong>vapor</strong> deposited with the silane in the same manner. Figure 5 is a plot <strong>of</strong> the<br />

electrochemical capacitance versus scan rate determined by cyclic voltammetry.<br />

Capacitances are highest for <strong>monolayers</strong> deposited without the addition <strong>of</strong><br />

MgSO4·7H2O. When <strong>monolayers</strong> are deposited in the presence <strong>of</strong> 0.5 g MgSO4·7H2O,<br />

the measured capacitances decrease. A further decrease in the capacitances is observed<br />

for surfaces exposed to a second 12 h <strong>deposition</strong> with 100 L <strong>of</strong> fresh silane and 0.5 g<br />

<strong>of</strong> fresh MgSO4·7H2O. The decrease in capacitance is accompanied by an increase in<br />

the intensity <strong>of</strong> the asymmetric methylene stretch by less than 10 percent in the FTIR<br />

spectrum after the second <strong>deposition</strong> with no discernible changes to the peak positions<br />

or widths <strong>of</strong> the methylene stretching modes (Figure S4 in the Supporting<br />

40


Information). The ellipsometric thicknesses and the water contact angles are also<br />

unchanged after the second <strong>deposition</strong>. Exposure <strong>of</strong> the surfaces to three repeated 12 h<br />

<strong>deposition</strong>s with fresh silane and fresh MgSO4·7H2O results in the same capacitance<br />

as two repeated <strong>deposition</strong>s, suggesting that a minimum limiting value <strong>of</strong> the<br />

capacitance is reached after two repeated <strong>deposition</strong>s.<br />

Capacitance / F/cm 2<br />

20 1x without MgSO 4 ·7H 2 O<br />

15<br />

10<br />

5<br />

0<br />

1x with 0.5 g MgSO 4 ·7H 2 O<br />

2x with 0.5 g MgSO 4 ·7H 2 O<br />

3x with 0.5 g MgSO 4 ·7H 2 O<br />

10 100 1000<br />

Scan Rate / mV/s<br />

Figure 5. Plot <strong>of</strong> the electrochemical capacitance versus scan rate during cyclic<br />

voltammetry after 12 h <strong>deposition</strong>s <strong>of</strong> 100 µL (CD3O)3-Si-(CH2)13-CH3 onto ITO<br />

electrodes without and with 0.5 g MgSO4·7H2O present. The squares and triangles<br />

represent the capacitances measured after two and three repeated 12 h <strong>deposition</strong>s,<br />

each with 100 µL <strong>of</strong> fresh silane and fresh 0.5 g MgSO4·7H2O. Error bars represent<br />

one standard deviation.<br />

41


Discussion<br />

Measurements <strong>of</strong> ellipsometric thickness show that incomplete <strong>monolayers</strong> are<br />

formed if insufficient MgSO4·7H2O is present. Dense <strong>monolayers</strong> are deposited with<br />

100 µL silane and 0.5 g MgSO4·7H2O as inferred from the plateau <strong>of</strong> the ellipsometric<br />

thickness at 18.2 Å independent <strong>of</strong> further increases in <strong>deposition</strong> time, amount <strong>of</strong><br />

silane or amount <strong>of</strong> MgSO4·7H2O. This <strong>vapor</strong>-deposited thickness is consistent with<br />

literature values which range from 18 Å to 20 Å for <strong>monolayers</strong> solution-deposited<br />

from tetradecyltrichlorosilane. 14,15 We infer that dense <strong>monolayers</strong> form only when<br />

sufficient water to hydrolyze all methoxy groups is present, and sufficient silane is<br />

present for the time required to densify the monolayer. When sufficient silane is<br />

present, incomplete <strong>monolayers</strong> can nonetheless result from the steric crowding <strong>of</strong><br />

unhydrolyzed methoxy groups or from insufficient water to form additional <strong>siloxane</strong><br />

bonds.<br />

FTIR spectroscopy was used to monitor the hydrolysis <strong>of</strong> methoxy groups during<br />

<strong>deposition</strong>. From the observation <strong>of</strong> CD3 stretching in FTIR spectra measured on<br />

<strong>monolayers</strong> prepared in the absence <strong>of</strong> sufficient MgSO4·7H2O, we infer that<br />

unhydrolyzed methoxy groups remain on some <strong>of</strong> the adsorbed silanes. These may<br />

prevent monolayer completion. The decrease <strong>of</strong> the CD3 integrated absorbance with<br />

increased <strong>deposition</strong> time without MgSO4·7H2O present (Figure 2b) suggests that<br />

methoxy groups are hydrolyzed over time by adventitious water in the chamber or on<br />

the silicon oxide surfaces. However, adventitious water is insufficient to hydrolyze all<br />

methoxy groups as indicated by the residual CD3 absorbance for 100 µL silane. The<br />

deliberate addition <strong>of</strong> a water source is necessary to deposit dense <strong>monolayers</strong>.<br />

42


Increasing the mass <strong>of</strong> MgSO4·7H2O above the approximately 0.5 g required for<br />

complete removal <strong>of</strong> the methoxy groups on the surface yields no increase in the<br />

ellipsometric thickness and provides further support for the possibility that unreacted<br />

methoxy groups are responsible for incomplete monolayer <strong>deposition</strong>.<br />

The hydrolysis <strong>of</strong> methoxy groups and completion <strong>of</strong> the monolayer are correlated<br />

with an observed shift <strong>of</strong> methylene stretching to lower frequency. It is known that, as<br />

the packing <strong>of</strong> alkyl chains becomes denser, the positions <strong>of</strong> the symmetric and<br />

asymmetric methylene stretches shift to lower frequency. 49-52 In the absence <strong>of</strong><br />

MgSO4·7H2O, the methylene stretching frequencies are characteristic <strong>of</strong> liquid-like<br />

packing <strong>of</strong> alkyl chains. With the addition <strong>of</strong> sufficient MgSO4·7H2O for complete<br />

hydrolysis, the methylene stretching frequencies shift to lower frequency and are<br />

similar to previous reports 53,54 for dense <strong>monolayers</strong> deposited from solutions <strong>of</strong><br />

alkyltrichlorosilanes <strong>of</strong> similar chain length. This further supports our conclusion that<br />

incomplete <strong>monolayers</strong> are formed without sufficient MgSO4·7H2O present.<br />

Water contact-angle measurements on <strong>monolayers</strong> deposited from 100 µL silane<br />

with sufficient MgSO4·7H2O are equal to those measured for dense alkyl<strong>siloxane</strong><br />

<strong>monolayers</strong> deposited from solution 15 and provide additional evidence for improved<br />

packing <strong>of</strong> the <strong>monolayers</strong>. Lower contact angles indicative <strong>of</strong> less dense <strong>monolayers</strong><br />

are measured without deliberate addition <strong>of</strong> MgSO4·7H2O.<br />

The amount <strong>of</strong> silane required to form a dense monolayer when there is sufficient<br />

MgSO4·7H2O is approximately 100 µL (Figure 4a), which is 2 orders <strong>of</strong> magnitude<br />

above the calculated amount needed to coat all surfaces in the <strong>deposition</strong> chamber<br />

with a dense monolayer <strong>of</strong> the silane. 55 The volatile silane is presumably depleted by<br />

43


hydrolysis and condensation reactions in the reservoir <strong>of</strong> liquid silane leading to cross-<br />

linked <strong>siloxane</strong>s with low <strong>vapor</strong> pressure. The lifetime <strong>of</strong> the silane source must be<br />

sufficiently large to provide volatile silane for the time required for monolayer<br />

densification. This is illustrated by experiments that use 10 µL silane with 0.05 g and<br />

0.5 g MgSO4·7H2O. In those experiments, complete methoxy group hydrolysis is<br />

observed, but ellipsometric thicknesses are measured that are similar to <strong>monolayers</strong><br />

deposited from 100 µL silane without MgSO4·7H2O. Experiments with the silane<br />

liquid source contained in a glass capillary, rather than being absorbed into filter paper,<br />

do not result in observable differences in the <strong>monolayers</strong>, suggesting that the surface<br />

area <strong>of</strong> the silane source is not limiting its lifetime.<br />

To further assess the lifetime <strong>of</strong> the silane source, silicon oxide and ITO surfaces<br />

were subjected to two repeated 12 h <strong>vapor</strong> <strong>deposition</strong>s. Fresh reactants, 100 µL silane<br />

and 0.5 g MgSO4·7H2O, were used in each <strong>deposition</strong>. Exposure to a second<br />

<strong>deposition</strong> results in an increase in the intensity <strong>of</strong> the asymmetric methylene stretch<br />

<strong>of</strong> less than 10 percent with no discernible changes to the methylene stretching peak<br />

positions or widths, the ellipsometric thicknesses, or the water contact angles <strong>of</strong><br />

<strong>monolayers</strong> deposited onto silicon oxide surfaces. The work <strong>of</strong> Hong et al. 26<br />

motivated us to pursue a more sensitive measure <strong>of</strong> monolayer densification. They<br />

reported no changes in the ellipsometric thicknesses or water contact angles measured<br />

on <strong>vapor</strong>-deposited <strong>monolayers</strong> by increasing the <strong>deposition</strong> time beyond 12 h.<br />

However, an increase in the blocking <strong>of</strong> atomic layer <strong>deposition</strong> was observed on<br />

<strong>monolayers</strong> deposited for longer than 12 h.<br />

44


Electrochemical capacitance measurements were performed to test the resistance <strong>of</strong><br />

<strong>vapor</strong>-deposited <strong>siloxane</strong> <strong>monolayers</strong> on ITO to ion permeation. The measured<br />

capacitances decrease with increasing scan rate (Figure 5). We attribute this decrease<br />

to a decrease in the time for permeation with higher scan rates. 56 Electrochemical<br />

capacitances are highest for <strong>monolayers</strong> deposited without MgSO4·7H2O, from which<br />

we infer that these <strong>monolayers</strong> are most permeable to ions in solution. This<br />

observation agrees with the ellipsometric, FTIR spectroscopic, and water contact angle<br />

measurements indicating incomplete <strong>monolayers</strong>. The denser <strong>monolayers</strong> formed with<br />

0.5 g MgSO4·7H2O are more resistant to ion permeation. Exposing <strong>monolayers</strong><br />

formed with 0.5 g MgSO4·7H2O to a second 12 h <strong>deposition</strong> with fresh 100 µL silane<br />

and fresh 0.5 g MgSO4·7H2O results in even greater resistance to ion permeation. We<br />

infer that the electrochemical capacitance is a more sensitive probe for monolayer<br />

densification than the other measurements used because a difference in the<br />

electrochemical capacitance is clearly observed after the second <strong>deposition</strong> while only<br />

a subtle increase in the asymmetric methylene stretching intensity in the FTIR<br />

spectrum is observed with no discernible changes to the methylene stretching peak<br />

positions or widths, ellipsometric thicknesses, or water contact angles.<br />

The lowest capacitance measured, 2.0 µF·cm -2 , is higher than the value <strong>of</strong> 1.2<br />

µF·cm -2 , for thiol <strong>monolayers</strong> <strong>of</strong> the same chain length on gold. 51 We infer that <strong>vapor</strong>-<br />

deposited <strong>siloxane</strong> <strong>monolayers</strong> on ITO contain more pinhole and gauche defects that<br />

allow ion permeation than their thiol counterparts on gold. Previous reports <strong>of</strong><br />

capacitance values measured on <strong>siloxane</strong> <strong>monolayers</strong> solution-deposited onto<br />

electrode surfaces vary widely 57-59 from 1.1 to 24 µF·cm -2 , highlighting the sensitivity<br />

45


<strong>of</strong> the monolayer packing in <strong>siloxane</strong> <strong>monolayers</strong> to the details <strong>of</strong> the <strong>deposition</strong><br />

method.<br />

We also find that an excess <strong>of</strong> water above the calculated stoichiometric minimum<br />

amount <strong>of</strong> water is required to achieve complete hydrolysis and dense <strong>monolayers</strong>. For<br />

example, the complete hydrolysis <strong>of</strong> 100 µL silane (0.27 mmol) should<br />

stoichiometrically require 1.5 times as much water (0.40 mmol). 60 However, we<br />

observe in Figures 4a and 4b that dense monolayer formation requires approximately<br />

0.5 g MgSO4·7H2O which, upon partial dehydration at 110°C, 61-63 yields 9 mmol <strong>of</strong><br />

water, or about twenty times as much water as required for complete hydrolysis <strong>of</strong> the<br />

silane.<br />

What determines the amount <strong>of</strong> water required to form dense <strong>monolayers</strong> when<br />

sufficient silane is present? Equilibrium effects might explain the excess <strong>of</strong> water<br />

required to achieve complete hydrolysis. Methanol generated as a byproduct <strong>of</strong><br />

hydrolysis reactions accumulates in the <strong>deposition</strong> chamber over time and could drive<br />

the reverse methoxylation reaction. We have confirmed the reversibility <strong>of</strong> the<br />

hydrolysis <strong>of</strong> methoxy groups by exposing completely hydrolyzed, partial <strong>monolayers</strong><br />

to pure CD3OD and CD3OD/H2O <strong>vapor</strong> mixtures and observing the CD3 stretch in<br />

FTIR spectra. We attribute the reappearance to methoxylation <strong>of</strong> free silanols on the<br />

surface. The CD3 stretch is observed after exposure to CD3OD:H2O <strong>vapor</strong> mixtures<br />

with molar ratios <strong>of</strong> 3:1 and higher but not with a 1:1 ratio. From these data, we<br />

estimate that the equilibrium for methoxy group hydrolysis is mildly favorable with an<br />

equilibrium constant <strong>of</strong> approximately 4. 64 This is in agreement with determinations <strong>of</strong><br />

46


the equilibrium constant for alkoxy group hydrolysis in solution on the order <strong>of</strong> 10. 65-<br />

70<br />

While the equilibrium argument predicts the qualitative trend <strong>of</strong> increased density<br />

with increased water, the equilibrium constant we calculate from reversing the<br />

hydrolysis does not quantitatively explain the data. For instance, with 50 µL silane<br />

and 0.05 g MgSO4·7H2O, the methanol to water ratio is expected to be 0.6, well below<br />

the level required to inhibit the hydrolysis reaction (Table S1 in the Supporting<br />

Information). One possibility is that water release from the MgSO4·7H2O is rate<br />

limiting. In that case, the liquid silane may remove most <strong>of</strong> the water from the <strong>vapor</strong><br />

phase with the result that the silanes in the monolayer remain methoxylated until the<br />

liquid silane is exhausted and monolayer growth terminates. Increasing the mass <strong>of</strong><br />

MgSO4·7H2O and the amount <strong>of</strong> water released per unit time might maintain a greater<br />

fraction <strong>of</strong> free silanols in the monolayer at early time and promote monolayer growth.<br />

Conclusions<br />

In summary, we showed that the addition <strong>of</strong> water was necessary during <strong>siloxane</strong><br />

monolayer <strong>vapor</strong> <strong>deposition</strong> to hydrolyze methoxy groups and form dense <strong>monolayers</strong>.<br />

Water was deliberately introduced to <strong>vapor</strong> <strong>deposition</strong>s through the dehydration <strong>of</strong><br />

MgSO4·7H2O. We developed a probe to measure the hydrolysis <strong>of</strong> deuterium-labeled<br />

methoxysilanes. This allowed the monitoring <strong>of</strong> CD3 stretching by FTIR spectroscopy<br />

as a function <strong>of</strong> time, mass <strong>of</strong> MgSO4·7H2O, and amount <strong>of</strong> silane. With sufficient<br />

silane present, ellipsometric thicknesses consistent with dense <strong>monolayers</strong> were<br />

measured when a ten-to-twenty fold excess <strong>of</strong> water above the stoichiometric amount<br />

was present. Lower frequency methylene stretching modes, higher water contact<br />

47


angles, and lower electrochemical capacitances indicating denser <strong>monolayers</strong> were<br />

achieved when enough silane and water were present. Electrochemical capacitance<br />

measurements were the most sensitive to changes in the density <strong>of</strong> the <strong>siloxane</strong><br />

<strong>monolayers</strong>. The excess water required was at least in part due to the reversibility <strong>of</strong><br />

the hydrolysis <strong>of</strong> the methoxy groups.<br />

We provide a simple laboratory method to deposit dense <strong>siloxane</strong> <strong>monolayers</strong> using<br />

optimized amounts <strong>of</strong> silane and water and repeated <strong>deposition</strong> cycles if needed. We<br />

are currently using this method to deposit mixed, <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> onto<br />

oxide surfaces, which can be further functionalized with a broad array <strong>of</strong> groups by the<br />

Cu(I)-catalyzed <strong>azide</strong>-alkyne cycloaddition reaction.<br />

Acknowledgment. This work was supported by funding from Helicos BioSciences<br />

Corporation. The authors acknowledge NSF grant CHE-0639053 for support <strong>of</strong> FTIR<br />

equipment. The authors also acknowledge NSF grant DMR-0213618 to the Center on<br />

Polymer Interfaces and Macromolecular Assemblies for use <strong>of</strong> the plasma cleaner and<br />

ellipsometer. Useful discussions with Dr. Erin Artin and Dr. Tim Harris are<br />

acknowledged.<br />

Supporting Information Available: NMR and FTIR spectra <strong>of</strong> tetradecyl-<br />

tri(deuteromethoxy)silane. Cyclic voltammogram and FTIR spectrum <strong>of</strong> a <strong>siloxane</strong><br />

monolayer after two repeated 12 h <strong>vapor</strong> <strong>deposition</strong>s. Time dependence <strong>of</strong><br />

MgSO4·7H2O dehydration. Table <strong>of</strong> expected CH3OH:H2O ratios in the <strong>deposition</strong><br />

48


chamber based on stoichiometry. This material is available free <strong>of</strong> charge via the<br />

Internet at http://pubs.acs.org/.<br />

Supporting Information<br />

Figure S1. Attenuated total reflectance (ATR) FTIR spectrum <strong>of</strong> the bulk, liquid<br />

tetradecyl-tri(deuteromethoxy)silane used for <strong>siloxane</strong> monolayer <strong>vapor</strong> <strong>deposition</strong>s.<br />

49


Figure S2. NMR spectrum <strong>of</strong> tetradecyl-tri(deuteromethoxy)silane that was<br />

synthesized and used for <strong>siloxane</strong> monolayer <strong>vapor</strong> <strong>deposition</strong>s. The spectrum was<br />

obtained on a 300 MHz Varian Inova spectrometer in CDCl3 solvent.<br />

50


Current Density / A·cm -2<br />

10<br />

5<br />

0<br />

-5<br />

-10<br />

-0.2 -0.1 0.0 0.1 0.2 0.3 0.4 0.5<br />

Potential / V vs. Ag/AgCl/saturated KCl<br />

Figure S3. Plot <strong>of</strong> the current versus applied potential measured during cyclic<br />

voltammetry for a <strong>siloxane</strong> monolayer on ITO exposed to 0.1 M NaClO4 in water at a<br />

scan rate <strong>of</strong> 1000 mV/s. The <strong>siloxane</strong> monolayer was formed from two repeated 12 h<br />

<strong>vapor</strong> <strong>deposition</strong>s, each with 100 µL <strong>of</strong> fresh (CD3O)3-Si-(CH2)13-CH3 and fresh 0.5 g<br />

MgSO4·7H2O.<br />

51


Absorbance<br />

1.0x10 -3<br />

5.0x10 -4<br />

0<br />

2x with MgSO 4 ·7H 2 O<br />

1x with MgSO 4 ·7H 2 O<br />

2800 2850 2900 2950 3000<br />

Frequency / cm -1<br />

Figure S4. Carbon-hydrogen stretching region <strong>of</strong> the p-polarized Brewster‟s angle<br />

transmission FTIR spectra <strong>of</strong> <strong>siloxane</strong> <strong>monolayers</strong> on silicon oxide after one 12 h<br />

<strong>vapor</strong> <strong>deposition</strong> (solid line) and two repeated 12 h <strong>vapor</strong> <strong>deposition</strong>s (dotted line),<br />

each with 100 µL <strong>of</strong> fresh (CD3O)3-Si-(CH2)13-CH3 and fresh 0.5 g MgSO4·7H2O.<br />

52


equiv H 2 O : equiv MgSO 4 ·7H 2 O<br />

6<br />

5<br />

4<br />

3<br />

2<br />

1<br />

0<br />

0 2 4<br />

Time / h<br />

Figure S5. Time dependence <strong>of</strong> the dehydration <strong>of</strong> MgSO4·7H2O at 110°C under<br />

reduced pressure starting with 0.5 g MgSO4·7H2O. The dehydration <strong>of</strong> the salt was<br />

monitored gravimetrically. The plateau at about 4.5 equivalents <strong>of</strong> water released per<br />

equivalent <strong>of</strong> initial MgSO4·7H2O was used for all stoichiometry calculations. The<br />

line is a guide to the eye.<br />

53


Table S1. Calculated CH3OH:H2O Ratio in Deposition Chamber Assuming Complete<br />

Hydrolysis and Condensation<br />

volume silane<br />

mass MgSO4·7H2O<br />

10 µL<br />

50 µL<br />

0 g 0.005 g 0.050 g 0.25 g 0.50 g 1.0 g<br />

0.092<br />

0.56<br />

54<br />

0.0088<br />

0.045<br />

100 µL ∞ ∞ 1.6 0.19 0.092 0.045<br />

300 µL<br />

∞<br />

Table S1 shows the calculated CH3OH:H2O ratio expected at long <strong>deposition</strong> time in<br />

the silane <strong>deposition</strong> chamber assuming a water:trimethoxysilane stoichiometric ratio<br />

<strong>of</strong> 1.5:1 for conditions used in this study. Entries in the table with „∞‟ indicate pure<br />

methanol should remain because there was enough silane present to consume all <strong>of</strong> the<br />

water in the system. Therefore, it is not surprising that CD3 stretching was observed<br />

after those experiments with less than stoichiometric water present. However, CD3<br />

stretching was still observed from experiments with water in stoichiometric excess,<br />

100 µL silane and 0.05 g MgSO4·7H2O, for example. A CH3OH:H2O ratio <strong>of</strong> 1.6 in<br />

the <strong>deposition</strong> chamber is predicted under those conditions. With the estimated<br />

equilibrium constant <strong>of</strong> 4.3 for the hydrolysis reaction, a SiOCD3:SiOH ratio <strong>of</strong> ~ 0.36<br />

is predicted. Decreasing the amount <strong>of</strong> silane to 50 µL with the same amount <strong>of</strong> water<br />

from 0.05 g MgSO4·7H2O still resulted in CD3 stretching. Under these conditions, a<br />

SiOCD3:SiOH ratio <strong>of</strong> ~ 0.13 is predicted. Although the amount water present under<br />

these conditions was greater than stoichiometric requirements, equilibrium predicts<br />

that a fraction <strong>of</strong> available silanols would be methoxylated as confirmed by the<br />

observation <strong>of</strong> CD3 stretching. No CD3 stretching was observed after lowering the<br />

0.3


amount <strong>of</strong> initial silane to 10 µL with 0.05 g MgSO4·7H2O. Less driving force for<br />

reversing the hydrolysis would be expected with a calculated CH3OH:H2O ratio <strong>of</strong><br />

0.092 and a SiOCD3:SiOH ratio <strong>of</strong> 0.021. This same ratio is expected for experiments<br />

using 100 µL silane with 0.5 g MgSO4·7H2O which resulted in dense, completely<br />

hydrolyzed <strong>monolayers</strong>. Dense, completely hydrolyzed <strong>monolayers</strong> were also obtained<br />

from experiments using 300 µL silane with 0.5 g MgSO4·7H2O with higher calculated<br />

CH3OH:H2O and SiOCD3:SiOH ratios <strong>of</strong> 0.3 and ~ 0.07, respectively.<br />

55


References and Notes<br />

(1) Sander, L. C.; Lippa, K. A.; Wise, S. A. Anal. Bioanal. Chem. 2005, 382, 646-<br />

668.<br />

(2) Devaraj, N. K.; Miller, G. P.; Ebina, W.; Kakaradov, B.; Collman, J. P.; Kool,<br />

E. T.; Chidsey, C. E. D. J. Am. Chem. Soc. 2005, 127, 8600-8601.<br />

(3) Sassolas, A.; Leca-Bouvier, B. D.; Blum, L. J. Chem. Rev. 2007, 108, 109-139.<br />

(4) Ulman, A. Chem. Rev. 1996, 96, 1533-1554.<br />

(5) Schwartz, D. K. Annu. Rev. Phys. Chem. 2001, 52, 107-137.<br />

(6) Onclin, S.; Ravoo, B. J.; Reinhoudt, D. N. Angew. Chem., Int. Ed. 2005, 44,<br />

6282-6304.<br />

(7) Plueddemann, E. P. Silane Coupling Agents, 2nd ed.; Plenum Press: New York,<br />

1991.<br />

(8) Ulman, A. An Introduction to Ultrathin Organic Films: From Langmuir-<br />

Blodgett Films to Self-Assembly; Academic Press: San Diego, CA, 1991.<br />

(9) Brzoska, J. B.; Azouz, I. B.; Rondelez, F. Langmuir 1994, 10, 4367-4373.<br />

(10) Angst, D. L.; Simmons, G. W. Langmuir 1991, 7, 2236-2242.<br />

(11) Vallant, T.; Kattner, J.; Brunner, H.; Mayer, U.; H<strong>of</strong>fmann, H. Langmuir 1999,<br />

15, 5339-5346.<br />

(12) Krasnoslobodtsev, A. V.; Smirnov, S. N. Langmuir 2002, 18, 3181-3184.<br />

(13) Le Grange, J. D.; Markham, J. L.; Kurkjian, C. R. Langmuir 1993, 9, 1749-<br />

1753.<br />

(14) Silberzan, P.; Leger, L.; Ausserre, D.; Benattar, J. J. Langmuir 1991, 7, 1647-<br />

1651.<br />

56


(15) Wasserman, S. R.; Tao, Y.-T.; Whitesides, G. M. Langmuir 1989, 5, 1074-<br />

1087.<br />

(16) Fadeev, A. Y.; McCarthy, T. J. Langmuir 2000, 16, 7268-7274.<br />

(17) Tripp, C. P.; Hair, M. L. Langmuir 1991, 7, 923-927.<br />

(18) Tripp, C. P.; Hair, M. L. Langmuir 1995, 11, 149-155.<br />

(19) Pasternack, R. M.; Rivillon Amy, S.; Chabal, Y. J. Langmuir 2008, 24, 12963-<br />

12971.<br />

(20) Sagiv, J. J. Am. Chem. Soc. 1980, 102, 92-98.<br />

(21) Haller, I. J. Am. Chem. Soc. 1978, 100, 8050-8055.<br />

(22) Jung, G. Y.; Li, Z.; Wu, W.; Chen, Y.; Olynick, D. L.; Wang, S. Y.; Tong, W.<br />

M.; Williams, R. S. Langmuir 2005, 21, 1158-1161.<br />

(23) Ashurst, R. W.; Carraro, C.; Maboudian, R. IEEE Trans. Device Mater. Reliab.<br />

2003, 3, 173-178.<br />

(24) Kong, X.; Kawai, T.; Abe, J.; Iyoda, T. Macromolecules 2001, 34, 1837-1844.<br />

(25) Ashurst, W. R.; Carraro, C.; Maboudian, R.; Frey, W. Sens. Actuators, A 2003,<br />

104, 213-221.<br />

(26) Hong, J.; Porter, D. W.; Sreenivasan, R.; McIntyre, P. C.; Bent, S. F. Langmuir<br />

2007, 23, 1160-1165.<br />

(27) Mayer, T. M.; de Boer, M. P.; Shinn, N. D.; Clews, P. J.; Michalske, T. A. J.<br />

Vac. Sci. Technol. B 2000, 18, 2433-2440.<br />

(28) H<strong>of</strong>fmann, P. W.; Stelzle, M.; Rabolt, J. F. Langmuir 1997, 13, 1877-1880.<br />

57


(29) Fiorilli, S.; Rivolo, P.; Descrovi, E.; Ricciardi, C.; Pasquardini, L.; Lunelli, L.;<br />

Vanzetti, L.; Pederzolli, C.; Onida, B.; Garrone, E. J. Colloid Interface Sci.<br />

2008, 321, 235-241.<br />

(30) Crampton, N.; Bonass, W. A.; Kirkham, J.; Thomson, N. H. Langmuir 2005,<br />

21, 7884-7891.<br />

(31) Saini, G.; Sautter, K.; Hild, F. E.; Pauley, J.; Linford, M., R. J. Vac. Sci.<br />

Technol., A 2008, 26, 1224-1234.<br />

(32) Ek, S.; Iiskola, E. I.; Niinisto, L.; Vaittinen, J.; Pakkanen, T. T.; Keranen, J.;<br />

Auroux, A. Langmuir 2003, 19, 10601-10609.<br />

(33) Kurth, D. G.; Bein, T. Langmuir 1995, 11, 3061-3067.<br />

(34) Anderson, A. S.; Dattelbaum, A. M.; Montano, G. A.; Price, D. N.; Schmidt, J.<br />

G.; Martinez, J. S.; Grace, W. K.; Grace, K. M.; Swanson, B. I. Langmuir 2008,<br />

24, 2240-2247.<br />

(35) Ressier, L.; Viallet, B.; Grisolia, J.; Peyrade, J. P. Ultramicroscopy 2007, 107,<br />

980-984.<br />

(36) Hozumi, A.; Ushiyama, K.; Sugimura, H.; Takai, O. Langmuir 1999, 15, 7600-<br />

7604.<br />

(37) Du, Y.; George, S. M. J. Phys. Chem. C 2007, 111, 8509-8517.<br />

(38) Li, J.-R.; Garno, J. C. Nano Lett. 2008, 8, 1916-1922.<br />

(39) Li, J.-R.; Lusker, K. L.; Yu, J.-J.; Garno, J. C. ACS Nano 2009, 3, 2023-2035.<br />

(40) Zhang, F.; Sautter, K.; Larsen, A. M.; Findley, D. A.; Davis, R. C.; Samha, H.;<br />

Linford, M. R. Langmuir 2010, 26, 14648-14654.<br />

58


(41) George, A.; Blank, D. H. A.; ten Elsh<strong>of</strong>, J. E. Langmuir 2009, 25, 13298–<br />

13301.<br />

(42) Fadeev, A. Y.; McCarthy, T. J. Langmuir 1999, 15, 3759-3766.<br />

(43) Bianco, A.; Maggini, M.; Nogarole, M.; Scorrano, G. Eur. J. Org. Chem. 2006,<br />

2006, 2934-2941.<br />

(44) Shimojima, A.; Liu, Z.; Ohsuna, T.; Terasaki, O.; Kuroda, K. J. Am. Chem.<br />

Soc. 2005, 127, 14108-14116.<br />

(45) Chabal, Y., J. ; Hines, M. A.; Feijoo, D. J. Vac. Sci. Technol., A 1995, 13,<br />

1719-1727.<br />

(46) Michalak, D. J.; Rivillon, S.; Chabal, Y. J.; Estève, A.; Lewis, N. S. J. Phys.<br />

Chem. B 2006, 110, 20426-20434.<br />

(47) Yang, C. S. C.; Richter, L. J.; Stephenson, J. C.; Briggman, K. A. Langmuir<br />

2002, 18, 7549-7556.<br />

(48) An excess <strong>of</strong> <strong>vapor</strong> was used during these post-treatments. At 110°C, a total<br />

pressure <strong>of</strong> about 480 Torr is expected in the <strong>deposition</strong> chamber. The total<br />

number <strong>of</strong> moles <strong>of</strong> <strong>vapor</strong> was kept constant at 0.012 mol for all mixtures used.<br />

The following masses <strong>of</strong> CD3OD and H2O were used in these experiments:<br />

0.440 g CD3OD and 0 g H2O (pure CD3OD), 0.400 g CD3OD and 0.020 g H2O<br />

(10:1 molar ratio), 0.330 g CD3OD and 0.055 g H2O (3:1 molar ratio), and<br />

0.216 g CD3OD and 0.108 g H2O (1:1 molar ratio).<br />

(49) Snyder, R. G.; Strauss, H. L.; Elllger, C. A. J. Phys. Chem. 1982, 86, 5145-<br />

5150.<br />

59


(50) Snyder, R. G.; Maroncelli, M.; Strauss, H. L.; Hallmark, V. M. J. Phys. Chem.<br />

1986, 90, 5623-5630.<br />

(51) Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D. J. Am. Chem. Soc.<br />

1987, 109, 3559-3568.<br />

(52) Cicero, R. L.; Linford, M. R.; Chidsey, C. E. D. Langmuir 2000, 16, 5688-5695.<br />

(53) H<strong>of</strong>fmann, H.; Mayer, U.; Krischanitz, A. Langmuir 1995, 11, 1304-1312.<br />

(54) Tillman, N.; Ulman, A.; Schildkraut, J. S.; Penner, T. L. J. Am. Chem. Soc.<br />

1988, 110, 6136-6144.<br />

(55) A dense <strong>siloxane</strong> monolayer should exhibit an approximate coverage <strong>of</strong> 5x10 14<br />

molecules per cm 2 based on each alkylsilane occupying an area <strong>of</strong> about 20<br />

Å 2 . 15 Therefore, 100 µL (0.27 mmol) <strong>of</strong> the silane should be enough to cover<br />

3x10 5 cm 2 with a monolayer. The upper limit <strong>of</strong> surface area inside the silane<br />

<strong>vapor</strong> <strong>deposition</strong> chamber is estimated to be approximately 10 3 cm 2 .<br />

(56) Chidsey, C. E. D.; Loiacono, D. N. Langmuir 1990, 6, 682-691.<br />

(57) Hillebrandt, H.; Tanaka, M. J. Phys. Chem. B 2001, 105, 4270-4276.<br />

(58) Kulkarni, S. A.; Vijayamohanan, K. P. Surf. Sci. 2007, 601, 2983-2993.<br />

(59) Chen, C.; Hutchison, J. E.; Postlethwaite, T. A.; Richardson, J. N.; Murray, R.<br />

W. Langmuir 1994, 10, 3332-3337.<br />

(60) The calculation <strong>of</strong> 0.4 mmol water to hydrolyze 0.27 mmol (100 µL) <strong>of</strong> the<br />

silane assumes a water:trimethoxysilane stoichiometric ratio <strong>of</strong> 1.5:1, which is<br />

derived from the overall reaction <strong>of</strong> complete trimethoxysilane hydrolysis<br />

(requiring water) followed by complete condensation (releasing water) to form<br />

a poly<strong>siloxane</strong> network. The expected upper limit for the<br />

60


water:trimethoxysilane stoichiometry is 3:1 from complete hydrolysis <strong>of</strong> the<br />

methoxy groups and no water-producing condensation to form <strong>siloxane</strong>s. In<br />

this case, 0.8 mmol water would be needed to hydrolyze the silane.<br />

(61) Experiments were conducted to determine the amount <strong>of</strong> water released from<br />

the dehydration <strong>of</strong> MgSO4·7H2O, at 110°C. Specifically, 0.5 g MgSO4·7H2O<br />

was placed into the same <strong>deposition</strong> chamber and 110°C oven used in the<br />

<strong>vapor</strong> <strong>deposition</strong> experiments for various periods <strong>of</strong> time. The measured mass<br />

difference was used to determine how much water was liberated (Figure S5 in<br />

the supporting information). Our experimental determination <strong>of</strong> the hydrated<br />

salt releasing about 4.5 waters per formula unit agrees well with previous<br />

reports. 62,63 In addition, loss <strong>of</strong> vacuum in the chamber was observed after the<br />

dehydration <strong>of</strong> masses <strong>of</strong> MgSO4·7H2O slightly greater than 1.0 g as expected<br />

from the measured dehydration stoichiometry.<br />

(62) Paulik, J. J.; Paulik, F.; Arnold, M. Thermochim. Acta 1981, 50, 105.<br />

(63) Emons, H.; Ziegenbalg, G.; Naumann, R.; Paulik, F. J. Therm. Anal. 1990, 36,<br />

1265.<br />

(64) The equilibrium constant for hydrolysis was estimated by interpolation <strong>of</strong> the<br />

measured CD3 absorbances after exposure to CD3OD/H2O mixtures. An<br />

interpolation was performed between the 10:1 and 3:1 CD3OD:H2O data points<br />

to estimate the ratio where an integrated CD3 absorbance <strong>of</strong> 1.55x10 -4 cm -1 ,<br />

one-half the absorbance after exposure to pure CD3OD, should be measured.<br />

The result <strong>of</strong> this interpolation was a CD3OD:H2O molar ratio <strong>of</strong> 4.3:1, which<br />

is the equilibrium constant for the reaction: RSiOCD3 + H2O ↔ RSiOH +<br />

61


CD3OH. It is assumed that all available free silanols that would participate in<br />

the equilibrium reaction would be methoxylated after a 12 h treatment with<br />

excess pure CD3OD.<br />

(65) Ro, J. C.; Chung, I. J. J. Non-Cryst. Solids 1989, 110, 26-32.<br />

(66) Rankin, S. E.; McCormick, A. V. Macromolecules 2000, 33, 7743-7750.<br />

(67) van Beek, J. J.; Seykens, D.; Jansen, J. B. H. J. Non-Cryst. Solids 1992, 146,<br />

111-120.<br />

(68) Rankin, S. E.; Sefcík, J.; McCormick, A. V. Ind. Eng. Chem. Res. 1999, 38,<br />

3191-3198.<br />

(69) Sanchez, J.; McCormick, A. J. Phys. Chem. 1992, 96, 8973-8979.<br />

(70) Sefcík, J.; Rankin, S. E.; Kirchner, S. J.; McCormick, A. V. J. Non-Cryst.<br />

Solids 1999, 258, 187-197.<br />

62


Chapter 3: Vapor Deposition <strong>of</strong> Mixed, Azide-Terminated Siloxane Monolayers: A<br />

Modular Strategy for Modifying Oxide Surfaces<br />

Preface<br />

This chapter <strong>of</strong> my dissertation is from a manuscript that will be submitted to<br />

the scientific journal Langmuir. I am the first author <strong>of</strong> this publication and am<br />

responsible for most <strong>of</strong> the writing and preparation. The other authors are Sujatha<br />

Raghu, and Pr<strong>of</strong>essor Christopher E. D. Chidsey. I was involved in all <strong>of</strong> the<br />

experiments that were conducted for this manuscript. My main contributions to this<br />

manuscript include: synthesis <strong>of</strong> the 11-azidoundecyltrimethoxysilane and<br />

tetradecyltri(deuteromethoxy)silane adsorbates, optimization <strong>of</strong> the <strong>vapor</strong> <strong>deposition</strong><br />

conditions, demonstrating the importance <strong>of</strong> water addition, determining mixed<br />

monolayer compositions, and optimization <strong>of</strong> copper-catalyzed <strong>azide</strong>-alkyne<br />

cycloaddition (CuAAC) reaction conditions.<br />

Sujatha Raghu‟s contributions were in performing ellipsometry and water<br />

contact angle measurements and in assisting with the first mixed monolayer <strong>vapor</strong><br />

<strong>deposition</strong> experiments. Pr<strong>of</strong>. Chidsey proposed the idea <strong>of</strong> studying <strong>azide</strong>-<strong>terminated</strong><br />

<strong>siloxane</strong> <strong>monolayers</strong> on silicon oxide as part <strong>of</strong> the collaboration with Helicos<br />

BioSciences. He also provided significant guidance in interpreting and planning<br />

experiments as well as in the writing and preparation <strong>of</strong> the manuscript.<br />

63


Abstract<br />

Pure and mixed <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> are deposited from water <strong>vapor</strong><br />

and the <strong>vapor</strong> <strong>of</strong> pure and mixed trimethoxyorganosilanes. Pure <strong>azide</strong>-<strong>terminated</strong><br />

<strong>monolayers</strong> are deposited from 11-azidoundecyltrimethoxysilane <strong>vapor</strong>. Denser<br />

<strong>monolayers</strong> are achieved when MgSO4·7H2O is present as a deliberate water source<br />

than without it, as inferred from ellipsometry, Fourier transform infrared (FTIR)<br />

spectroscopy, and x-ray photoelectron spectroscopy (XPS) measurements. Mixed<br />

<strong>monolayers</strong> are <strong>vapor</strong> deposited from 11-azidoundecyltrimethoxysilane and an<br />

alkyltrimethoxysilane diluent. The relative surface compositions <strong>of</strong> the mixed<br />

<strong>monolayers</strong> are enriched with the silane reactant with the higher <strong>vapor</strong> pressure as<br />

probed using FTIR spectroscopy, XPS, and water contact angle goniometry. Uniform<br />

mixing in the mixed <strong>monolayers</strong> is inferred from the quantitative Cu-catalyzed <strong>azide</strong>-<br />

alkyne cycloaddition (CuAAC) reactions <strong>of</strong> the <strong>azide</strong> groups with 4-ethynyl-α,α,α-<br />

trifluorotoluene with <strong>azide</strong> coverages up to approximately 40% <strong>of</strong> a monolayer.<br />

Quantitative reactions are not observed on dense <strong>monolayers</strong> <strong>of</strong> only <strong>azide</strong> groups,<br />

presumably due to steric hindrance.<br />

64


Introduction<br />

We have developed a modular method for the reproducible, covalent<br />

attachment <strong>of</strong> a species <strong>of</strong> interest to oxide surfaces with tunable coverage. Our<br />

method involves the <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> mixed, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong><br />

onto oxide surfaces followed by the Cu-catalyzed <strong>azide</strong>-alkyne cycloaddition<br />

(CuAAC) reaction to attach alkyne-<strong>terminated</strong> species <strong>of</strong> interest as shown in Figure 1.<br />

The surface properties <strong>of</strong> oxide surfaces are <strong>of</strong>ten tailored with chemisorbed<br />

<strong>monolayers</strong> presenting desired functionality at the surface. 1,2,3 These <strong>monolayers</strong> can<br />

be formed from adsorbates with silane 4 and less frequently with phosphonic acid 5 or<br />

carboxylic acid 6 headgroups that can form durable bonds to many metal oxide surfaces.<br />

One approach for covalently modifying oxide surfaces is the synthesis and<br />

<strong>deposition</strong> <strong>of</strong> unique adsorbates each time one wants to change the species or<br />

functionality presented at the surface. However, simply tethering an adsorbate group,<br />

such as a silane, to a species <strong>of</strong> interest does not guarantee its successful anchoring. A<br />

more modular strategy is the formation <strong>of</strong> a monolayer presenting a specific functional<br />

group onto which a variety <strong>of</strong> species can be covalently attached using a predictable<br />

coupling chemistry. 7 This latter approach is advantageous because it avoids extensive<br />

chemical syntheses in attaching a compatible adsorbate group to each species, and it<br />

allows independent optimization <strong>of</strong> each step.<br />

We chose the CuAAC reaction for our coupling chemistry over the multitude<br />

<strong>of</strong> other interfacial chemistries that have been used to covalently attach species to<br />

<strong>monolayers</strong>. 8 The CuAAC reaction is the most well-known „click‟ reaction and was<br />

discovered independently by Sharpless 9 and Meldal 10 in 2002. The CuAAC reaction<br />

65


is an ideal candidate for surface coupling because it is selective, quantitative, and<br />

proceeds rapidly in aqueous and mixed aqueous organic solutions under ambient<br />

laboratory conditions. 11 The sole product <strong>of</strong> the CuAAC reaction is a stable 1,4-<br />

disubstituted 1,2,3-triazole.<br />

Before using the CuAAC reaction for surface coupling, terminal <strong>azide</strong> or<br />

alkyne functional groups must be introduced onto the surface. We deliberately chose<br />

to introduce <strong>azide</strong> functional groups onto the surface because <strong>azide</strong> groups can be<br />

probed relatively easily using Fourier transform infrared spectroscopy (FTIR)<br />

spectroscopy 12 as well as x-ray photoelectron spectroscopy (XPS). 13 The versatility <strong>of</strong><br />

the CuAAC reaction on <strong>azide</strong>-<strong>terminated</strong> surfaces has been demonstrated through the<br />

attachment <strong>of</strong> a variety <strong>of</strong> alkyne-<strong>terminated</strong> species. 14,15,16,17,18,19,20,21,22,23,24,25 Azide-<br />

<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> have <strong>of</strong>ten been formed by the solution <strong>deposition</strong> <strong>of</strong><br />

bromide-<strong>terminated</strong> <strong>monolayers</strong> followed by SN2 substitution <strong>of</strong> the bromide groups<br />

with <strong>azide</strong> groups. 26 This substitution requires long reaction times (~48 h) to approach<br />

completion and sometimes leads to mixtures <strong>of</strong> bromide and <strong>azide</strong> functional groups<br />

on the surface. 27,28<br />

To avoid this problem, we have synthesized and purified 11-<br />

azidoundecyltrimethoxysilane as a preformed <strong>azide</strong>-<strong>terminated</strong> adsorbate for <strong>vapor</strong><br />

<strong>deposition</strong> onto silicon oxide surfaces. The primary advantage <strong>of</strong> <strong>vapor</strong> <strong>deposition</strong><br />

over solution <strong>deposition</strong> is to eliminate the formation <strong>of</strong> polymeric <strong>siloxane</strong>s in the<br />

phase contacting the surface. The reduced contamination <strong>of</strong> surfaces with particles <strong>of</strong><br />

these <strong>siloxane</strong> polymers has been demonstrated using <strong>vapor</strong> <strong>deposition</strong> compared to<br />

solution <strong>deposition</strong>. 29,30,31 We recently reported a method for <strong>deposition</strong> <strong>of</strong> dense,<br />

66


methyl-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> onto oxide surfaces from<br />

alkyltrimethoxysilane and water <strong>vapor</strong>s and showed that incomplete <strong>monolayers</strong> were<br />

deposited without both sufficient silane and sufficient water present. 32 That method is<br />

used to deposit <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> in this study. Though the <strong>vapor</strong><br />

<strong>deposition</strong> <strong>of</strong> <strong>siloxane</strong> <strong>monolayers</strong> presenting other reactive, terminal functional<br />

groups such as bromide 16 and alkyne 17 groups has been reported, the <strong>vapor</strong> <strong>deposition</strong><br />

<strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> has not been reported to our knowledge.<br />

Dilution <strong>of</strong> the <strong>azide</strong> groups as one component <strong>of</strong> a mixed monolayer has<br />

several advantages. Among these is the ability to sufficiently dilute the <strong>azide</strong> groups to<br />

overcome potential steric limitations in the coupling <strong>of</strong> species to all <strong>azide</strong> groups on<br />

the surface. 33 Control over the dilution <strong>of</strong> the <strong>azide</strong> groups also allows one to tune the<br />

spacing <strong>of</strong> and limit the interactions between the species once coupled to the surface.<br />

Mixed <strong>siloxane</strong> <strong>monolayers</strong> are usually deposited by coadsorption from organic<br />

solutions containing more than one silane reactant. 34,35,27 The coadsorption <strong>of</strong> two<br />

silanes from the <strong>vapor</strong> phase is studied in this paper. The surface compositions <strong>of</strong><br />

mixed <strong>monolayers</strong> are generally different than the initial adsorbate mixtures from<br />

which they were deposited. 36,37,28,38 Therefore, the ability to probe the surface<br />

compositions <strong>of</strong> mixed <strong>monolayers</strong> is important, and the <strong>azide</strong> group provides such a<br />

spectroscopic handle. 37,39<br />

In this paper, we show that denser <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> are <strong>vapor</strong><br />

deposited with a deliberate water source, MgSO4·7H2O, present than without it. Mixed<br />

<strong>monolayers</strong> are deposited from the <strong>vapor</strong>s <strong>of</strong> 11-azidoundecyltrimethoxysilane and<br />

alkyltrimethoxysilanes with the water source present. We observe enrichment <strong>of</strong> the<br />

67


mixed <strong>monolayers</strong> with the higher <strong>vapor</strong> pressure silane reactant. It is inferred that the<br />

composition <strong>of</strong> the mixed <strong>monolayers</strong> is determined by the composition <strong>of</strong> the silane<br />

<strong>vapor</strong> in contact with the oxide surface rather than the remote, liquid silane source.<br />

The quantitative CuAAC reaction <strong>of</strong> the <strong>azide</strong> groups in a mixed monolayer and 4-<br />

ethynyl-α,α,α-trifluorotoluene is demonstrated; whereas, the quantitative CuAAC<br />

reaction <strong>of</strong> a dense monolayer presenting only <strong>azide</strong> groups is presumably prevented<br />

by steric hindrance.<br />

<strong>azide</strong><br />

species <strong>of</strong> interest<br />

alkyne<br />

Cu(I) catalyst<br />

Cu-catalyzed <strong>azide</strong>alkyne<br />

cycloaddition<br />

Oxide Surface Oxide Surface<br />

Figure 1. Modular strategy for the covalent modification <strong>of</strong> oxide surfaces using<br />

mixed, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> followed by the Cu-catalyzed <strong>azide</strong>-<br />

alkyne cycloaddition (CuAAC) reaction with terminal alkynes to attach species <strong>of</strong><br />

interest (red circles).<br />

68


Experimental<br />

Trimethoxysilane Reactants<br />

Decyltrimethoxysilane and hexyltrimethoxysilane were purchased from Alfa-<br />

Aesar and used as received.<br />

11-azidoundecyltrimethoxysilane was prepared from 11-<br />

bromoundecyltrimethoxysilane (Gelest) via displacement <strong>of</strong> the bromides with sodium<br />

<strong>azide</strong> following published procedures. 40,41 1.00 g (15 mmol) <strong>of</strong> sodium <strong>azide</strong> (Sigma-<br />

Aldrich) was added to a solution containing 2.00 g (5.6 mmol) <strong>of</strong> 11-<br />

bromoundecyltrimethoxysilane (Gelest) in 30 mL <strong>of</strong> dry dimethylformamide (Acros).<br />

The solution was allowed to stir for 24 h at room temperature under a nitrogen<br />

atmosphere and then filtered. The filtrate was extracted 3 times with dry pentane. The<br />

pentane was then removed by rotary e<strong>vapor</strong>ation. The crude product was purified by<br />

two successive vacuum distillations. CAUTION: Because <strong>azide</strong> groups can<br />

decompose violently to release nitrogen gas, we limited the mass <strong>of</strong> the azidosilane to<br />

no more than 2 g and the temperature <strong>of</strong> the oil bath heat source to no more than<br />

150°C behind a blast shield. In addition, only about 90% <strong>of</strong> the azidosilane in the pot<br />

was distilled to avoid distilling to dryness. See Figure S1 in the Supporting<br />

Information for a nuclear magnetic resonance (NMR) spectrum.<br />

Tetradecyltri(deuteromethoxy)silane was prepared as previously reported. 32<br />

Surface Cleaning<br />

Single or double-side polished Si(100) wafers, either newly received or after<br />

prior experiments, were cleaned in an oxygen plasma cleaner (Harrick) for 10 min at<br />

medium power with a dioxygen flow rate <strong>of</strong> 50 std. mL/min at a pressure <strong>of</strong> 250<br />

69


mTorr. This cleaning method resulted in hydrophilic surfaces with water contact<br />

angles approaching 0°. 42 Surfaces were exposed to atmospheric air for up to 15 min<br />

after cleaning while determining effective refractive indices <strong>of</strong> the substrates by<br />

ellipsometry.<br />

Monolayer Vapor Deposition<br />

Siloxane <strong>monolayers</strong> were <strong>vapor</strong>-deposited onto silicon oxide surfaces in o-<br />

ring sealed, glass vacuum desiccators fitted with Teflon stopcocks (Jencons part<br />

number 250-048) with internal volumes <strong>of</strong> about 600 mL. These chambers were oven<br />

dried for a minimum <strong>of</strong> 4 h at 140°C in air and cooled on the bench top for no more<br />

than 10 min in air at room temperature before use. Neat silane (100 µL), either pure<br />

azidosilane or a silane mixture, was pipetted onto and absorbed into 42.5 mm diameter<br />

Whatman filter paper dried in the same way in the bottom <strong>of</strong> the desiccator. 0.5 g <strong>of</strong><br />

MgSO4·7H2O (Fisher biochemical grade) were placed in a foil boat in the bottom <strong>of</strong><br />

the desiccator to serve as a water source for the hydrolysis reaction. The plasma-<br />

cleaned surfaces were placed on a metal rack in the chamber above the silane liquid<br />

and hydrated salt. The desiccator was then evacuated through a rubber hose to a glass<br />

vacuum line with a liquid nitrogen-trapped mechanical pump for approximately 60<br />

seconds. The final pressure at the trap was 1 Torr. The Teflon valve on the desiccator<br />

was closed, and the chamber was placed in a 110°C preheated oven (Forma Scientific)<br />

for various periods <strong>of</strong> time from 1 to 24 h. After <strong>deposition</strong>, the valve was opened to<br />

ambient air to return the chamber to atmospheric pressure and to remove the samples.<br />

The following mixtures were used for the mixed monolayer <strong>deposition</strong>s: 0.250 g<br />

azidosilane and 0.620 g decyltrimethoxysilane (χ N3 ,liquid = 0.25), 0.500 g azidosilane<br />

70


and 0.413 g decyltrimethoxysilane (χ N3 ,liquid = 0.50), 0.750 g azidosilane and 0.207 g<br />

decyltrimethoxysilane (χ N3 ,liquid = 0.75), 0.750 g azidosilane and 0.162 g<br />

hexyltrimethoxysilane (χ N3 ,liquid = 0.75), 0.750 g azidosilane and 0.258 g<br />

tetradecyltri(deuteromethoxy)silane (χ N3 ,liquid = 0.75).<br />

Surface Characterization<br />

FTIR spectra were obtained with a Bruker Vertex 70 spectrometer using a KBr<br />

beam splitter and a deuterated triglycine sulfate (DTGS) detector. Spectra were<br />

collected in transmission mode through the silicon with the samples at the silicon<br />

Brewster‟s angle (74°) using p-polarized light from a wire-grid polarizer (Specac<br />

KRS-5). For each spectrum, 1024 scans were collected at 4 cm -1 resolution.<br />

Background spectra were collected from freshly oxygen plasma-cleaned, double side<br />

polished Si(100) crystals in the same orientation.<br />

Ellipsometry measurements were performed using a Gaertner L116<br />

ellipsometer at 70° angle <strong>of</strong> incidence. Effective real and imaginary refractive indices<br />

<strong>of</strong> the Si(100) substrates were determined immediately after plasma cleaning but<br />

before <strong>siloxane</strong> monolayer <strong>vapor</strong> <strong>deposition</strong>. Measurements after <strong>vapor</strong> <strong>deposition</strong>,<br />

rinsing with toluene and isopropyl alcohol, and drying in a stream <strong>of</strong> nitrogen were<br />

used to determine monolayer thicknesses. A film refractive index <strong>of</strong> 1.46 was used for<br />

all thickness calculations. At least three measurements were performed on each<br />

sample to obtain average substrate refractive indices or sample thicknesses.<br />

Advancing sessile drop water contact angles were measured with a Ramé-Hart<br />

model 100 goniometer using water from a four-bowl Millipore purification system.<br />

71


Water droplets were dispensed from a syringe with a flat-tipped needle. Fresh water<br />

was obtained from the Millipore system for each measurement session. At least two<br />

measurements were made on each sample on both sides <strong>of</strong> the droplets and averaged.<br />

X-ray photoelectron spectroscopy (XPS) was performed using a Physical<br />

Electronics Inc. 5000 Versa Probe spectrometer with Al Kα radiation (1486.6 eV).<br />

High-resolution spectra were the result <strong>of</strong> the co-addition <strong>of</strong> 10 scans collected at a<br />

step size <strong>of</strong> 0.1 eV, an analyzer pass energy <strong>of</strong> 117 eV, and a take<strong>of</strong>f angle <strong>of</strong> 45°.<br />

Cu-Catalyzed Azide-Alkyne Cycloaddition<br />

CuAAC reactions were performed by exposing <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> to<br />

3:2 dimethyl sulfoxide (DMSO):H2O solutions containing 1 mM 4-ethynyl-α,α,α-<br />

trifluorotoluene, 500 µM CuSO4·5H2O, 500 µM <strong>of</strong> 1,1',1''-Tris(1H-1,2,3-triazol-4-yl-<br />

1-acetic acid ethyl ester) trimethylamine (TTMA; CAS#736958-00-2; see Supporting<br />

Information for structure), and 5 mM sodium ascorbate for various periods <strong>of</strong> time.<br />

Next, the surfaces were rinsed with chlor<strong>of</strong>orm, toluene, isopropyl alcohol, water, and<br />

isopropyl alcohol again. The surfaces were then sonicated for 1 minute in chlor<strong>of</strong>orm<br />

and rinsed with isopropyl alcohol and dried in a stream <strong>of</strong> nitrogen. TTMA was<br />

synthesized following a published procedure. 43<br />

Results<br />

Azide Functional Group as a Spectroscopic Handle<br />

The ability to monitor the <strong>azide</strong> groups on surfaces provides a way to follow<br />

the progress <strong>of</strong> the CuAAC reaction. Figure 2 shows a p-polarized Brewster‟s angle<br />

transmission FTIR spectrum <strong>of</strong> a monolayer <strong>vapor</strong> deposited from the azidosilane for<br />

12 h with 0.5 g MgSO4·7H2O present. The peak at 2100 cm -1 is assigned to the<br />

72


asymmetric stretching mode <strong>of</strong> the <strong>azide</strong> group, 44 which can be used to monitor<br />

relative <strong>azide</strong> surface compositions. In addition to the <strong>azide</strong> absorbance, two other<br />

peaks are present at 2855 cm -1 and 2925 cm -1 . These peaks correspond to the<br />

symmetric and asymmetric methylene stretching modes <strong>of</strong> the alkyl chains in the<br />

monolayer. 45<br />

Figure 2. Azide asymmetric and carbon-hydrogen stretching regions <strong>of</strong> the p-<br />

polarized Brewster‟s angle transmission FTIR spectrum <strong>of</strong> a χ N3 ,liquid = 1.0 monolayer<br />

on silicon oxide after a 12 h <strong>vapor</strong> <strong>deposition</strong> from 100 µL <strong>of</strong> 11-<br />

azidoundecyltrimethoxysilane with 0.5 g MgSO4·7H2O present.<br />

XPS can also be used to monitor the presence <strong>of</strong> <strong>azide</strong> groups on the surface.<br />

In the case <strong>of</strong> <strong>azide</strong> groups, the XPS spectrum yields distinct bonding information in<br />

the N 1s region. As shown in Figure 3, an asymmetric doublet is present in the high-<br />

resolution N 1s XPS spectrum <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer that was <strong>vapor</strong><br />

deposited onto silicon oxide. The higher binding energy XPS peak is assigned to the<br />

73


electron-deficient, central nitrogen atom in the <strong>azide</strong> group, while the lower binding<br />

energy peak is assigned to the two outer nitrogen atoms in the <strong>azide</strong> group as shown<br />

by the peak fitting. 46 Figure 3 also shows the Si 2p photoelectron region. The lower<br />

binding energy peak is assigned to the bulk silicon atoms in the substrate, and the<br />

higher binding energy peak is assigned to the silicon atoms bonded to electron-<br />

withdrawing oxygen atoms. The ratios <strong>of</strong> the integrated areas <strong>of</strong> the N 1s and Si 2p<br />

peaks are used to probe the relative <strong>azide</strong> surface compositions after <strong>vapor</strong> <strong>deposition</strong>s<br />

and after CuAAC reactions.<br />

Figure 3. High-resolution Si 2p (95 - 110 eV) and N 1s (395 - 410 eV) XPS spectra<br />

for a χ N3 ,liquid = 1.0 monolayer on silicon oxide after 12 h <strong>of</strong> <strong>vapor</strong> <strong>deposition</strong> from 100<br />

µL <strong>of</strong> 11-azidoundecyltrimethoxysilane with 0.5 g MgSO4·7H2O present. The<br />

asymmetric doublet characteristic <strong>of</strong> the <strong>azide</strong> group in the high-resolution N 1s<br />

spectrum is fitted by three Gaussian peaks, one for each nitrogen atom in the <strong>azide</strong><br />

group.<br />

74


Time Dependence <strong>of</strong> Monolayer Formation<br />

Figure 4a shows the time dependence <strong>of</strong> ellipsometric thicknesses measured on<br />

<strong>monolayers</strong> deposited at 110°C from the <strong>vapor</strong> <strong>of</strong> 100 µL <strong>of</strong> 11-<br />

azidoundecyltrimethoxysilane without and with 0.5 g MgSO4·7H2O present in the<br />

<strong>deposition</strong> chamber. Larger thicknesses are observed for the <strong>monolayers</strong> deposited<br />

with MgSO4·7H2O. Limiting ellipsometric thicknesses are observed after 12 h in both<br />

cases. The limiting thickness without 0.5 g MgSO4·7H2O is 12.1 Å; whereas, it is 14.4<br />

Å with 0.5 g MgSO4·7H2O.<br />

Figure 4b is a plot <strong>of</strong> the integrated <strong>azide</strong> absorbance (Figure 2) versus<br />

<strong>deposition</strong> time. The <strong>azide</strong> integrated absorbance increases with increasing <strong>deposition</strong><br />

time until reaching a limiting value after about 12 h <strong>of</strong> <strong>deposition</strong>. The <strong>azide</strong><br />

absorbance is higher for <strong>deposition</strong>s performed in the presence <strong>of</strong> MgSO4·7H2O than<br />

for <strong>deposition</strong>s without MgSO4·7H2O.<br />

Figure 4c is a plot <strong>of</strong> the ratio <strong>of</strong> the areas <strong>of</strong> the N 1s and Si 2p XPS peaks<br />

(Figure 3) as a function <strong>of</strong> time. The N 1s:Si 2p ratios increase with time and are<br />

higher at all time points for <strong>deposition</strong>s performed with 0.5 g MgSO4·7H2O than for<br />

those without. Limiting values are reached after about 12 h <strong>of</strong> <strong>deposition</strong>.<br />

The time dependence <strong>of</strong> the water contact angle is plotted in Figure 4d for<br />

<strong>deposition</strong>s with and without 0.5 g MgSO4·7H2O. The water contact angles increase<br />

with increasing <strong>deposition</strong> time before reaching the same limiting value after 4 h. No<br />

difference in the limiting values <strong>of</strong> the water contact angles is observed between<br />

<strong>deposition</strong>s performed with or without MgSO4·7H2O. This is in contrast to our work<br />

75


with methyl-<strong>terminated</strong> <strong>monolayers</strong> where larger water contact angles were measured<br />

with 0.5 g MgSO4·7H2O than without it. 32<br />

76


Ellipsometric Thickness / Å<br />

Integrated Azide Absorb. / cm -1<br />

XPS N 1s:Si 2p<br />

Ellipsometric Thickness / Å<br />

Water Contact Angle / °<br />

0.04<br />

0.03<br />

0.02<br />

0.01<br />

0<br />

0.45<br />

0.30<br />

0.15<br />

16<br />

0<br />

16<br />

12<br />

8<br />

4<br />

0<br />

12<br />

60<br />

80 (d)<br />

8<br />

40<br />

4 20<br />

0<br />

0<br />

(a)<br />

(b)<br />

(c)<br />

without MgSO 4 ·7H 2 O<br />

with 0.5 g MgSO 4 ·7H 2 O<br />

with 0.5 g MgSO 4 ·7H 2 O<br />

without MgSO 4 ·7H 2 O<br />

0 0 55 10 15 20 20 25 25<br />

77<br />

Time / h<br />

Figure 4. Time dependence <strong>of</strong> <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> χ N3 ,liquid = 1.0 <strong>monolayers</strong> onto<br />

silicon oxide surfaces from 100 µL <strong>of</strong> 11-azidoundecyltrimethoxysilane with and


without 0.5 g MgSO4·7H2O present. The time dependence was monitored using (a)<br />

ellipsometry, (b) FTIR spectroscopy, (c) XPS, and (d) water contact angle goniometry.<br />

CuAAC Reactions on Dense, Azide-Terminated Monolayers<br />

Once conditions for <strong>vapor</strong> depositing dense, <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> were<br />

established, the reactivity <strong>of</strong> the <strong>azide</strong> groups with terminal alkynes using the CuAAC<br />

reaction was assessed using FTIR spectroscopy. Figure 5a shows FTIR spectra <strong>of</strong> a<br />

pure azidosilane monolayer before and after a 1 h exposure to a solution containing 4-<br />

ethynyl-α,α,α-trifluorotoluene and the Cu(I) catalyst for the CuAAC reaction. After 1<br />

h <strong>of</strong> reaction, the intensity <strong>of</strong> the <strong>azide</strong> stretch is reduced but is still observable. In<br />

addition, a new peak at 1328 cm -1 is present in the spectrum, which is assigned to the<br />

asymmetric C-F stretching mode. 37,47,48 The inset in Figure 5a shows the time<br />

dependence <strong>of</strong> the <strong>azide</strong> integrated absorbance during the CuAAC reaction with 4-<br />

ethynyl-α,α,α-trifluorotoluene. The <strong>azide</strong> absorbance decreases with increasing time<br />

until reaching a minimum limiting value after 1 h <strong>of</strong> reaction. Increasing the reaction<br />

time to 6 h did not result in any further decrease <strong>of</strong> the <strong>azide</strong> absorbance.<br />

78


Figure 5. Carbon-fluorine and <strong>azide</strong> asymmetric stretching regions <strong>of</strong> the p-polarized<br />

Brewster‟s angle transmission FTIR spectrum <strong>of</strong> (a) χ N3 ,liquid = 1.0 and (b) χ N3 ,liquid =<br />

0.75 <strong>monolayers</strong> on silicon oxide before and after 1 h <strong>of</strong> a CuAAC reaction with 4-<br />

ethynyl-α,α,α-trifluorotoluene. The <strong>monolayers</strong> were <strong>vapor</strong>-deposited for 12 h from<br />

100 µL <strong>of</strong> liquid silane with 0.5g MgSO4·7H2O present. The inset in (a) shows the<br />

time dependence <strong>of</strong> the integrated <strong>azide</strong> FTIR absorbance during CuAAC reactions<br />

between 4-ethynyl-α,α,α-trifluorotoluene and χ N3 ,liquid = 1.0 <strong>monolayers</strong> on silicon<br />

oxide.<br />

79


Mixed Monolayers<br />

Mixed <strong>monolayers</strong> were <strong>vapor</strong> deposited from azidosilane and<br />

alkyltrimethoxysilane liquid sources <strong>of</strong> known composition and 0.5 g MgSO4·7H2O.<br />

Figure 6a shows the integrated <strong>azide</strong> absorbance versus the number fraction <strong>of</strong> the<br />

azidosilane (χ N3 ,liquid) in the liquid silane source. The <strong>azide</strong> absorbance monotonically<br />

increases with increasing χ N3 ,liquid. The straight, dashed line labeled with a „1‟<br />

interpolates between the integrated absorbance for χ N3 ,liquid = 0, where no <strong>azide</strong><br />

absorbance is observed, and χ N3 ,liquid = 1, where maximum <strong>azide</strong> absorbance is<br />

observed. With decyltrimethoxysilane as the diluent (circles), the integrated <strong>azide</strong><br />

absorbances are all less than the straight, dashed line.<br />

The surface compositions <strong>of</strong> the mixed <strong>monolayers</strong> are also assessed using N<br />

1s:Si 2p XPS ratios (Figure 6b). Increasing χ N3 ,liquid results in an increase in the N<br />

1s:Si 2p ratio until reaching the maximum at χ N3 ,liquid = 1. The N 1s:Si 2p ratios for all<br />

<strong>monolayers</strong> deposited from mixtures <strong>of</strong> the azidosilane and decyltrimethoxysilane<br />

(circles) are less than the straight, dashed line that interpolates between the two<br />

limiting values. Figure 6c shows the dependence <strong>of</strong> the water contact angles on<br />

χ N3 ,liquid for <strong>monolayers</strong> deposited from mixtures <strong>of</strong> the azidosilane and<br />

decyltrimethoxysilane. The contact angles range from 80° for χ N3 ,liquid = 1 to 109° for<br />

χ N3 ,liquid = 0.<br />

The effect <strong>of</strong> varying the alkyl chain length <strong>of</strong> the alkyltrimethoxysilane<br />

diluent was explored for χ N3 ,liquid = 0.75 using FTIR spectroscopy and XPS.<br />

80


Monolayers <strong>vapor</strong> deposited from mixtures <strong>of</strong> the azidosilane and<br />

hexyltrimethoxysilane exhibit lower <strong>azide</strong> integrated absorbances and lower N 1s:Si<br />

2p XPS ratios (Figures 6a and b, diamonds) than those deposited from mixtures <strong>of</strong> the<br />

azidosilane and decyltrimethoxysilane. Larger <strong>azide</strong> absorbances and N 1s:Si 2p ratios<br />

(Figures 6a and b, squares) are measured with tetradecyltrimethoxysilane as the<br />

diluent compared to decyltrimethoxysilane.<br />

81


Integrated <strong>azide</strong> absorbance / cm -1<br />

water contact angle / °<br />

XPS N 1s: Si 2p<br />

0.04<br />

0.02<br />

0<br />

0.4<br />

0.2<br />

0<br />

110<br />

100<br />

90<br />

80<br />

(a)<br />

(b)<br />

(c)<br />

C14 diluent<br />

C10 diluent<br />

C6 diluent<br />

0 0.25 0.50<br />

N3 ,liquid<br />

0.75 1.00<br />

Figure 6. Relative surface compositions <strong>of</strong> mixed <strong>monolayers</strong> <strong>vapor</strong> deposited from<br />

the azidosilane and alkyltrimethoxysilane diluents as inferred from: (a) integrated<br />

<strong>azide</strong> FTIR absorbances, (b) N 1s:Si 2p XPS ratios, and (c) water contact angles. The<br />

82<br />

1<br />

2<br />

5<br />

10<br />

20<br />

50<br />

1<br />

2<br />

5<br />

10<br />

20<br />

50


C6, C10, and C14 alkyltrimethoxysilane diluents are hexyltrimethoxysilane<br />

decyltrimethoxysilane, and tetradecyltri(deuteromethoxy)silane, respectively. χ N3 ,liquid<br />

is the mole fraction <strong>of</strong> 11-azidoundecyltrimethoxysilane in the liquid silane source<br />

used during the <strong>vapor</strong> <strong>deposition</strong>s. The <strong>monolayers</strong> were <strong>vapor</strong>-deposited for 12 h<br />

from 100 µL <strong>of</strong> the silane mixtures with 0.5 g MgSO4·7H2O present. The dashed and<br />

dotted lines in (a) and (b) represent Raoult‟s law predictions for various <strong>vapor</strong> pressure<br />

ratios <strong>of</strong> the diluent silane to the azidosilane. 49 The numbers 1, 2, 5, 10, 20 and 50<br />

represent the diluent silane to azidosilane <strong>vapor</strong> pressure ratios for each <strong>of</strong> the Raoult‟s<br />

law predictions. The solid line in (c) is a guide to the eye.<br />

CuAAC Reactions on Mixed Monolayers<br />

Figure 5b shows the <strong>azide</strong> asymmetric stretching region <strong>of</strong> FTIR spectra<br />

measured on a χ N3 ,liquid = 0.75 monolayer with a decyltrimethoxysilane diluent before<br />

and after a 1 h CuAAC reaction with 4-ethynyl-α,α,α-trifluorotoluene. The <strong>azide</strong><br />

absorbance at 2100 cm -1 is reduced to the level <strong>of</strong> noise after 1 h <strong>of</strong> reaction,<br />

confirming quantitative conversion <strong>of</strong> the <strong>azide</strong> groups to triazoles.<br />

XPS is also used to monitor changes on the surface after the CuAAC reaction.<br />

Figure 7c is a high-resolution XPS spectrum <strong>of</strong> the N 1s region measured on a χ N3 ,liquid<br />

= 0.75 monolayer with a decyltrimethoxysilane diluent, and the asymmetric doublet<br />

characteristic <strong>of</strong> the <strong>azide</strong> group is observed. After 1 h <strong>of</strong> a CuAAC reaction with 4-<br />

ethynyl-α,α,α-trifluorotoluene, the higher binding energy <strong>azide</strong> peak is no longer<br />

83


present (Figure 7d). Instead, one broad peak at 401 eV is observed and is assigned to<br />

the three nitrogen atoms in the 1,2,3-triazole product <strong>of</strong> the CuAAC reaction.<br />

In addition to the disappearance <strong>of</strong> the higher binding energy N 1s peak after<br />

the CuAAC reaction, changes are observed in other regions <strong>of</strong> the XPS spectrum and<br />

provide further evidence for the covalent attachment <strong>of</strong> 4-ethynyl-α,α,α-<br />

trifluorotoluene. After the CuAAC reaction, a peak is observed in the F 1s region at<br />

689 eV, assignable to the fluorine atoms in the trifluoromethyl group as shown in<br />

Figure 7f, while no peak is observed in the spectrum <strong>of</strong> the mixed monolayer before<br />

reaction (Figure 7e). 37 A small peak at 293 eV, assignable to the electron-deficient<br />

carbon atom in the trifluoromethyl group, is present in the C 1s spectrum after reaction.<br />

The lower binding energy C 1s peak at 285 eV that is present before and after reaction<br />

is assigned to the other carbon atoms in the monolayer.<br />

84


Figure 7. High-resolution XPS spectra <strong>of</strong> the C 1s [(a) and (b)], N 1s [(c) and (d)], and<br />

F 1s [(e) and (f)] photoelectron regions for a χ N3 ,liquid = 0.75 monolayer on silicon<br />

oxide before and after 1 h <strong>of</strong> a CuAAC reaction with 4-ethynyl-α,α,α-trifluorotoluene.<br />

The monolayer was <strong>vapor</strong>-deposited for 12 h from 100 µL <strong>of</strong> a χ N3 ,liquid = 0.75 mixture<br />

<strong>of</strong> 11-azidoundecyltrimethoxysilane and decyltrimethoxysilane with 0.5 g<br />

MgSO4·7H2O present. The N 1s spectra are fitted by three Gaussian peaks, one for<br />

each nitrogen atom in the <strong>azide</strong> group, as shown in panels (c) and (d).<br />

Discussion<br />

Mixed, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> provide an ideal platform for the<br />

covalent modification <strong>of</strong> oxide surfaces with a variety <strong>of</strong> substituents because the<br />

<strong>azide</strong> composition is reproducible and can be conveniently probed using FTIR<br />

spectroscopy and XPS.<br />

85


Effect <strong>of</strong> Water Addition<br />

With MgSO4·7H2O present as a water source during <strong>vapor</strong> <strong>deposition</strong>, a<br />

higher coverage <strong>of</strong> the <strong>monolayers</strong> is obtained as inferred from the larger limiting<br />

values <strong>of</strong> the integrated <strong>azide</strong> FTIR absorbances and N 1s: Si 2p XPS ratios compared<br />

to <strong>monolayers</strong> deposited without MgSO4·7H2O (Figure 4). The methylene stretching<br />

modes in the 2800-3000 cm -1 range provide further evidence that complete<br />

<strong>monolayers</strong> are deposited with MgSO4·7H2O present. The positions <strong>of</strong> the symmetric<br />

and asymmetric methylene stretches shift to lower frequency as the packing <strong>of</strong> alkyl<br />

chains becomes denser. 50,51 The observation <strong>of</strong> the symmetric and asymmetric<br />

methylene FTIR stretching modes at 2855 and 2925 cm -1 for <strong>monolayers</strong> <strong>vapor</strong>-<br />

deposited for 12 h from the azidosilane with MgSO4·7H2O present is consistent with<br />

previous reports for the solution <strong>deposition</strong> <strong>of</strong> dense <strong>azide</strong> 52 and methyl-<strong>terminated</strong> 53<br />

<strong>siloxane</strong> <strong>monolayers</strong> with 11-carbon alkyl chains from trichlorosilanes.<br />

In addition, larger ellipsometric thicknesses are observed on <strong>monolayers</strong> <strong>vapor</strong><br />

deposited with MgSO4·7H2O present than without it. The thicknesses we observe with<br />

MgSO4·7H2O present, 14.4 Å, are comparable to literature values 52,23 that range from<br />

13.5 Å to 14.8 Å for <strong>monolayers</strong> solution-deposited from bromoundecyltrichlorosilane<br />

followed by SN2 displacement <strong>of</strong> the bromide groups on the surface with <strong>azide</strong> groups.<br />

We infer that complete <strong>monolayers</strong> are <strong>vapor</strong> deposited with MgSO4·7H2O present<br />

and that incomplete <strong>monolayers</strong> are deposited without MgSO4·7H2O. We have<br />

previously observed a similar difference in ellipsometric thicknesses for <strong>monolayers</strong><br />

<strong>vapor</strong>-deposited from tetradecyltri(deuteromethoxy)silane with and without 0.5 g<br />

MgSO4·7H2O present and found that unhydrolyzed methoxy groups remained on<br />

86


adsorbed silanes without sufficient MgSO4·7H2O. 32 Although the larger <strong>azide</strong><br />

absorbances, N 1s:Si 2p ratios, and ellipsometric thicknesses observed with<br />

MgSO4·7H2O present are qualitatively consistent, the increase in the <strong>azide</strong> absorbance<br />

and N 1s:Si 2p ratio with MgSO4·7H2O present is larger than the increase observed in<br />

the ellipsometric thickness. We do not have an explanation for this discrepancy.<br />

CuAAC Reactions on Pure and Mixed Azide-Terminated Monolayers<br />

The residual <strong>azide</strong> absorbance observed for χ N3 ,liquid = 1 <strong>monolayers</strong> after long<br />

exposure to CuAAC reaction conditions (Figure 5a) with 4-ethynyl-α,α,α-<br />

trifluorotoluene can be attributed to steric crowding on the surface. Devaraj et al.<br />

found that steric crowding prevents the quantitative reaction <strong>of</strong> the <strong>azide</strong> groups in<br />

dense <strong>monolayers</strong> deposited from <strong>azide</strong>-<strong>terminated</strong> thiols. 39 To overcome steric<br />

crowding on the surface, mixed <strong>monolayers</strong> were <strong>vapor</strong> deposited to dilute the <strong>azide</strong><br />

functionality with unreactive methyl groups. The quantitative reaction <strong>of</strong> the <strong>azide</strong><br />

groups was obtained with <strong>azide</strong> surface coverages up to approximately 40 percent as<br />

inferred from the disappearance <strong>of</strong> the <strong>azide</strong> absorbance in the FTIR spectrum (Figure<br />

5b) and from the disappearance <strong>of</strong> the higher binding energy <strong>azide</strong> peak in the high-<br />

resolution N 1s XPS spectrum after 1 h <strong>of</strong> reaction. This quantitative reaction <strong>of</strong> the<br />

<strong>azide</strong> groups at relatively high <strong>azide</strong> surface coverage suggests that uniform mixing <strong>of</strong><br />

the <strong>azide</strong> and methyl-<strong>terminated</strong> silanes, as opposed to segregation, is occurring in the<br />

mixed <strong>monolayers</strong>.<br />

Tuning the Azide Surface Composition with Mixed Monolayers<br />

When the <strong>azide</strong> groups are sufficiently isolated on the surface for quantitative<br />

CuAAC reactions, knowledge <strong>of</strong> the relative <strong>azide</strong> surface composition permits the<br />

87


predictable attachment <strong>of</strong> an alkyne-<strong>terminated</strong> species at a desired relative <strong>azide</strong><br />

surface composition. FTIR spectroscopy, XPS, and contact angle goniometry tell us<br />

that the <strong>vapor</strong>-deposited mixed <strong>monolayers</strong> are enriched with methyl-<strong>terminated</strong> alkyl<br />

chains from the decyltrimethoxysilane diluent because the <strong>azide</strong> integrated<br />

absorbances and N 1s: Si 2p XPS ratios for the mixtures are less than the dashed lines<br />

(Figures 6a and b, circles), which represent surface compositions equal to χ N3 ,liquid. The<br />

water contact angles also suggest enrichment <strong>of</strong> the unsubstituted<br />

decyltrimethoxysilane on the surface. Why are the surface compositions different than<br />

χ N3 ,liquid?<br />

We hypothesize that the relative <strong>azide</strong> surface composition is the same as the<br />

composition <strong>of</strong> the <strong>vapor</strong> that contacts the surface. We further hypothesize that the<br />

<strong>vapor</strong> is in equilibrium with the liquid silane source and that Raoult‟s law predicts the<br />

composition <strong>of</strong> the <strong>vapor</strong> in equilibrium with the liquid mixture. Predictions for<br />

integrated <strong>azide</strong> absorbances and N 1s:Si 2p XPS ratios from <strong>vapor</strong> compositions<br />

based upon Raoult‟s law are represented in Figures 6a and b by the dotted and dashed<br />

lines. 49 Predictions are plotted for <strong>vapor</strong> pressure ratios <strong>of</strong> the n-alkyltrimethoxysilane<br />

diluent to the azidosilane from 1 to 50. With decyltrimethoxysilane as the diluent<br />

(circles), both the FTIR data and the XPS data approximately resemble the Raoult‟s<br />

law prediction for a 5:1 <strong>vapor</strong> pressure ratio <strong>of</strong> the diluent to the azidosilane. A larger<br />

<strong>vapor</strong> pressure ratio <strong>of</strong> approximately 10 is expected by comparing the ratios <strong>of</strong> the<br />

<strong>vapor</strong> pressures <strong>of</strong> n-alkanes with comparable molecular weights to the silanes, 54,55,56<br />

suggesting that more enrichment <strong>of</strong> the monolayer with the decyltrimethoxysilane<br />

diluent is expected than obtained from experiment. Perhaps the adsorption <strong>of</strong> the<br />

88


longer chain azidosilane is slightly preferred over the shorter decyltrimethoxysilane<br />

diluent. 36 Experiments in which the alkyl chain length, and therefore <strong>vapor</strong> pressure,<br />

<strong>of</strong> the alkyltrimethoxysilane diluent was varied (Figures 6a and b) result in the<br />

expected trend. When the higher <strong>vapor</strong> pressure hexyltrimethoxysilane is used as the<br />

diluent, a greater enrichment <strong>of</strong> methyl functionality on the surface is observed<br />

compared to decyltrimethoxysilane; whereas, a lesser enrichment <strong>of</strong> methyl<br />

functionality is obtained when the lower <strong>vapor</strong> pressure tetradecyltrimethoxysilane<br />

diluent is used as inferred from integrated <strong>azide</strong> absorbances and N 1s:Si 2p ratios.<br />

Conclusions<br />

Pure and mixed <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> were <strong>vapor</strong> deposited<br />

onto silicon oxide surfaces from 11-azidoundecyltrimethoxysilane and n-<br />

alkyltrimethoxysilane diluents. The Cu-catalyzed <strong>azide</strong>-alkyne cycloaddition<br />

(CuAAC) reaction was used to covalently attach an alkyne-<strong>terminated</strong> species, 4-<br />

ethynyl-α,α,α-trifluorotoluene, to the <strong>azide</strong>-<strong>terminated</strong> surfaces. The <strong>azide</strong> group is a<br />

sensitive spectroscopic probe for surface analysis using FTIR spectroscopy and XPS<br />

to monitor relative <strong>azide</strong> surface compositions. Denser <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong><br />

were achieved from <strong>deposition</strong>s with 0.5 g MgSO4·7H2O present as a water source as<br />

inferred from the larger ellipsometric thicknesses, integrated <strong>azide</strong> FTIR absorbances,<br />

and N 1s:Si 2p XPS ratios that were measured with MgSO4·7H2O than without it.<br />

The relative <strong>azide</strong> surface compositions in mixed <strong>monolayers</strong> were not equal to<br />

the initial, liquid silane compositions. Enrichment <strong>of</strong> the higher <strong>vapor</strong> pressure silane<br />

reactant was observed in the mixed <strong>monolayers</strong> and was attributed to the difference<br />

between the silane <strong>vapor</strong> composition in contact with the surface and the remote liquid<br />

89


source composition. Quantitative reaction <strong>of</strong> the <strong>azide</strong> groups on the surface was<br />

achieved on mixed <strong>monolayers</strong> with the <strong>azide</strong> functionality diluted to a surface<br />

composition <strong>of</strong> approximately 40% with unreactive methyl groups, which suggests<br />

uniform mixing <strong>of</strong> the two components in the <strong>monolayers</strong>. Steric constraints<br />

presumably prevented the quantitative reaction <strong>of</strong> the <strong>azide</strong> groups in the pure <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong>.<br />

Acknowledgments<br />

This work was supported by funding from Helicos BioSciences Corporation. The<br />

authors acknowledge NSF grant CHE-0639053 for support <strong>of</strong> FTIR equipment. The<br />

authors also acknowledge NSF grant DMR-0213618 to the Center on Polymer<br />

Interfaces and Macromolecular Assemblies for use <strong>of</strong> the ellipsometer. Useful<br />

discussions with Dr. Erin Artin, Dr. Tim Harris, and Dr. Neal Devaraj are<br />

acknowledged.<br />

90


Supporting Information<br />

Figure S1. NMR spectrum <strong>of</strong> the synthesized 11-azidoundecyltrimethoxysilane used<br />

for <strong>siloxane</strong> monolayer <strong>vapor</strong> <strong>deposition</strong>s. The spectrum was obtained on a 300 MHz<br />

Varian Inova spectrometer in CDCl3 solvent.<br />

91


Figure S2. Chemical structure drawing <strong>of</strong> TTMA, the copper ligand used for CuAAC<br />

reactions.<br />

92


References and Notes<br />

1. Ulman, A. An Introduction to Ultrathin Organic Films: From Langmuir-<br />

Blodgett Films to Self-Assembly; Academic Press: San Diego, CA, 1991.<br />

2. Onclin, S.; Ravoo, B. J.; Reinhoudt, D. N. Angew. Chem., Int. Ed. 2005, 44,<br />

6282-6304.<br />

3. Schwartz, D. K. Annu. Rev. Phys. Chem. 2001, 52, 107-137.<br />

4. Plueddemann, E. P. Silane Coupling Agents, 2nd ed.; Plenum Press: New<br />

York, 1991.<br />

5. Mingalyov, P. G.; Lisichkin, G. V. Russian Chemical Reviews 2006, 75, 541-<br />

557.<br />

6. Karsi, N.; Lang, P.; Chehimi, M.; Delamar, M.; Horowitz, G. Langmuir 2006,<br />

22, 3118-3124.<br />

7. Haensch, C.; Hoeppener, S.; Schubert, U. S. Chemical Society Reviews 2010,<br />

39, 2323-2334.<br />

8. Li, J.; Thiara, P. S.; Mrksich, M. Langmuir 2007, 23, 11826-11835.<br />

9. Rostovtsev, V. V.; Green, L. G.; Fokin, V. V.; Sharpless, K. B. Angewandte<br />

Chemie (International ed.) 2002, 41, 2596-2599.<br />

10. Tornoe, C. W.; Christensen, C.; Meldal, M. Journal <strong>of</strong> organic chemistry<br />

2002, 67, 3057.<br />

11. Meldal, M.; Tornoe, C. W. Chem. Rev. 2008, 108, 2952-3015.<br />

12. Luo, L.; Frisbie, C. D. Journal <strong>of</strong> the American Chemical Society 2010, 132,<br />

8854-8855.<br />

93


13. Cao, P.; Xu, K.; Heath, J. R. Journal <strong>of</strong> the American Chemical Society 2008,<br />

130, 14910-14911.<br />

14. Collman, J. P.; Devaraj, N. K.; Decreau, R. A.; Yang, Y.; Yan, Y.-L.; Ebina,<br />

W.; Eberspacher, T. A.; Chidsey, C. E. D. Science 2007, 315, 1565-1568.<br />

15. Collman, J. P.; Devaraj, N. K.; Chidsey, C. E. D. Langmuir 2004, 20, 1051-<br />

1053.<br />

16. Prakash, S.; Long, T. M.; Selby, J. C.; Moore, J. S.; Shannon, M. A. Analytical<br />

Chemistry 2007, 79, 1661-1667.<br />

17. Ostaci, R.-V.; Damiron, D.; Capponi, S.; Vignaud, G.; Leger, L.; Grohens, Y.;<br />

Drockenmuller, E. Langmuir 2008, 24, 2732-2739.<br />

18. Devaraj, N. K.; Miller, G. P.; Ebina, W.; Kakaradov, B.; Collman, J. P.; Kool,<br />

E. T.; Chidsey, C. E. D. Journal <strong>of</strong> the American Chemical Society 2005, 127, 8600-<br />

8601.<br />

19. Fischler, M. M. Chemical Communications 2007, 169.<br />

20. Gauchet, C.; Labadie, G. R.; Poulter, C. D. Journal <strong>of</strong> the American Chemical<br />

Society 2006, 128, 9274-9275.<br />

21. Po-Chiao, L.; Shau-Hua, U.; Mei-Chun, T.; Jia-Ling, K.; Kuo-Ting, H.; Sheng-<br />

Chieh, Y.; Avijit Kumar, A.; Yu-Ju, C.; Chun-Cheng, L. Angewandte Chemie<br />

International Edition 2006, 45, 4286-4290.<br />

22. Schlossbauer, A.; Schaffert, D.; Kecht, J.; Wagner, E.; Bein, T. Journal <strong>of</strong> the<br />

American Chemical Society 2008, 130, 12558-12559.<br />

23. Rozkiewicz, D. I.; Janacuteczewski, D.; Verboom, W.; Ravoo, B. J.;<br />

Reinhoudt, D. N. Angew. Chem. Int. Ed. 2006, 45, 5292-5296.<br />

94


24. Claudia, H.; Stephanie, H.; Ulrich, S. S. Nanotechnology 2008, 19, 035703.<br />

25. Im, S. G.; Kim, B.-S.; Tenhaeff, W. E.; Hammond, P. T.; Gleason, K. K. Thin<br />

Solid Films 2009, 517, 3606-3611.<br />

26. Lummerstorfer, T.; H<strong>of</strong>fmann, H. J. Phys. Chem. B 2004, 108, 3963-3966.<br />

27. Fryxell, G. E.; Rieke, P. C.; Wood, L. L.; Engelhard, M. H.; Williford, R. E.;<br />

Graff, G. L.; Campbell, A. A.; Wiacek, R. J.; Lee, L.; Halverson, A. Langmuir 1996,<br />

12, 5064-5075.<br />

28. Heise, A.; Stamm, M.; Rauscher, M.; Duschner, H.; Menzel, H. Thin Solid<br />

Films 1998, 327-329, 199-203.<br />

29. Ashurst, R. W.; Carraro, C.; Maboudian, R. IEEE Trans. Device Mater. Reliab.<br />

2003, 3, 173-178.<br />

30. Kong, X.; Kawai, T.; Abe, J.; Iyoda, T. Macromolecules 2001, 34, 1837-1844.<br />

31. Jung, G. Y.; Li, Z.; Wu, W.; Chen, Y.; Olynick, D. L.; Wang, S. Y.; Tong, W.<br />

M.; Williams, R. S. Langmuir 2005, 21, 1158-1161.<br />

32. Lowe, R. D.; Pellow, M. A.; Stack, T. D. P.; Chidsey, C. E. D. Langmuir 2011,<br />

27, 9928-9935.<br />

33. Guo, F.; Jankova, K.; Schulte, L.; Vigild, M. E.; Ndoni, S. Langmuir 2009, 26,<br />

2008-2013.<br />

34. Fadeev, A. Y.; McCarthy, T. J. Langmuir 1999, 15, 7238-7243.<br />

35. Offord, D. A.; Griffin, J. H. Langmuir 2002, 9, 3015-3025.<br />

36. Bain, C. D.; Whitesides, G. M. Journal <strong>of</strong> the American Chemical Society<br />

1989, 111, 7164-7175.<br />

37. Azam, M. S.; Fenwick, S. L.; Gibbs-Davis, J. M. Langmuir 2011, 27, 741-750.<br />

95


38. Tong, Y.; Tyrode, E.; Osawa, M.; Yoshida, N.; Watanabe, T.; Nakajima, A.;<br />

Ye, S. Langmuir 2011, 27, 5420-5426.<br />

39. Collman, J. P.; Devaraj, N. K.; Eberspacher, T. P. A.; Chidsey, C. E. D.<br />

Langmuir 2006, 22, 2457-2464.<br />

40. Fu, Y.-S.; Yu, S. J. Angewandte Chemie (International ed.) 2001, 40, 437-440.<br />

41. Alauzun, J.; Besson, E.; Mehdi, A.; Rey; xe; Catherine; Corriu, R. J. P. Chem.<br />

Mater. 2008, 20, 503-513.<br />

42. Some <strong>of</strong> the surfaces were subsequently cleaned in a solution consisting <strong>of</strong> 10<br />

mL <strong>of</strong> 30% hydrogen peroxide (Kanto) and 10 mL <strong>of</strong> 29% ammonium hydroxide<br />

(Kanto) in 60 mL Millipore water 70°C. Surfaces were rinsed copiously with<br />

deionized water and then dried by rinsing with isopropanol and blow drying in a<br />

stream <strong>of</strong> nitrogen. The surfaces were then subjected to another oxygen plasma clean<br />

under the same conditions before using them. No difference in the resulting<br />

<strong>monolayers</strong> was observed between <strong>monolayers</strong> that were deposited onto surfaces that<br />

were only plasma cleaned and <strong>monolayers</strong> deposited onto surfaces that were subjected<br />

to the additional basic solution and plasma cleaning steps. .<br />

43. Zhou, Z.; Fahrni, C. J. JACS 2004, 126, 8862-8863.<br />

44. Lieber, E.; Rao, C. N. R.; Chao, T. S.; H<strong>of</strong>fman, C. W. W. Analytical<br />

Chemistry 1957, 29, 916-918.<br />

45. Porter, M. D.; Bright, T. B.; Allara, D. L.; Chidsey, C. E. D. Journal <strong>of</strong> the<br />

American Chemical Society 1987, 109, 3559-3568.<br />

46. Wollman, E. W.; Kang, D.; Frisbie, C. D.; Lorkovic, I. M.; Wrighton, M. S.<br />

Journal <strong>of</strong> the American Chemical Society 1994, 116, 4395-4404.<br />

96


47. Nomaru, K.; Gorelik, S. R.; Kuroda, H.; Nakai, K. Nuclear Instruments and<br />

Methods in Physics Research Section A: Accelerators, Spectrometers, Detectors and<br />

Associated Equipment 2004, 528, 619-622.<br />

48. Qu, L.; Xin, Z. Langmuir 2011, 27, 8365-8370.<br />

49. The dotted and dashed lines in Figure 6 are predictions <strong>of</strong> the silane <strong>vapor</strong><br />

mixture composition based upon Raoult‟s law for various ratios <strong>of</strong> the <strong>vapor</strong> pressure<br />

<strong>of</strong> the n-alkyltrimethoxysilane diluent (P<strong>vapor</strong>,diluent) to the <strong>vapor</strong> pressure <strong>of</strong> the<br />

azidosilane (P<strong>vapor</strong>,N 3 ). The predicted mole fraction <strong>of</strong> the azidosilane in the <strong>vapor</strong><br />

phase ( χ N 3 ,<strong>vapor</strong>) was then multiplied by the integrated <strong>azide</strong> absorbance and N 1s:Si 2p<br />

XPS ratio measured for <strong>monolayers</strong> <strong>vapor</strong> deposited from the pure azidosilane<br />

( χ N 3 ,liquid = 1) to convert the predicted <strong>vapor</strong> composition to an equivalent predicted<br />

surface composition. The Raoult‟s law equation used to make the predictions is:<br />

<br />

<br />

<br />

<br />

P <br />

N , liquid <br />

<strong>vapor</strong>,<br />

N <br />

3 <br />

3 <br />

<br />

N , <strong>vapor</strong><br />

3 <br />

<br />

<br />

<br />

P 1<br />

<br />

<br />

<br />

<br />

<br />

P<br />

N , liquid <strong>vapor</strong>,<br />

N N , liquid<br />

<br />

<br />

<br />

<br />

<strong>vapor</strong>,<br />

diluent<br />

3<br />

3 3<br />

50. Snyder, R. G.; Strauss, H. L.; Elllger, C. A. J. Phys. Chem. 1982, 86, 5145-<br />

5150.<br />

51. Snyder, R. G.; Maroncelli, M.; Strauss, H. L.; Hallmark, V. M. The Journal <strong>of</strong><br />

Physical Chemistry 1986, 90, 5623-5630.<br />

52. Lummerstorfer, T.; H<strong>of</strong>fmann, H. J. Phys. Chem. B 2004, 108, 3963-3966.<br />

53. H<strong>of</strong>fmann, H.; Mayer, U.; Krischanitz, A. Langmuir 1995, 11, 1304-1312.<br />

97


54. The expected ratio <strong>of</strong> 10:1 for the <strong>vapor</strong> pressures <strong>of</strong> decyltrimethoxysilane<br />

(MW = 262.5) to azidoundecyltrimethoxysilane (MW = 317.5) results from a<br />

comparison between nonadecane (MW = 268.5) with a <strong>vapor</strong> pressure <strong>of</strong> 0.37 Torr<br />

and tricosane (MW = 324.6) with a <strong>vapor</strong> pressure <strong>of</strong> 0.033 Torr at the same<br />

temperature used for the <strong>vapor</strong> <strong>deposition</strong>s, 110°C. These <strong>vapor</strong> pressures were<br />

calculated from Antoine equation parameters available on the online NIST<br />

WebBook. 55 Literature reports for the vacuum distillation <strong>of</strong> decyltrimethoxysilane 56<br />

and azidoundecyltrimethoxysilane 31 near 110°C suggest a larger <strong>vapor</strong> pressure ratio<br />

<strong>of</strong> approximately 70. Vacuum distillations are <strong>of</strong>ten performed under dynamic vacuum<br />

and the specific details are <strong>of</strong>ten omitted in synthetic preparations, which may<br />

invalidate the assumption that a reported distillation pressure is an accurate<br />

measurement <strong>of</strong> <strong>vapor</strong> pressure.<br />

55. P.J. Linstrom and W.G. Mallard, Eds., NIST Chemistry WebBook, NIST<br />

Standard Reference Database Number 69, National Institute <strong>of</strong> Standards and<br />

Technology, Gaithersburg MD, 20899, http://webbook.nist.gov, (retrieved May 2011).<br />

56. Ameduri, B.; Boutevin, B.; Moreau, J. J. E.; Moutaabbid, H.; Chi Man, M. W.<br />

Journal <strong>of</strong> Fluorine Chemistry 2000, 104, 185-194.<br />

98


Chapter 4: Attachment and Hybridization <strong>of</strong> Oligonucleotides on Azide-Terminated<br />

Siloxane Monolayers with Resistance to Nonspecific Adsorption<br />

Preface<br />

This chapter <strong>of</strong> my dissertation is derived from a manuscript that may be<br />

submitted to a scientific journal for publication. I am the first author <strong>of</strong> this manuscript<br />

and am responsible for most <strong>of</strong> the writing and preparation. The other authors are Dr.<br />

Erin Artin, David G. Lapham, Dr. Timothy D. Harris, and Pr<strong>of</strong>. Christopher E. D.<br />

Chidsey. My contributions to this work include: <strong>vapor</strong> <strong>deposition</strong> <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong><br />

<strong>monolayers</strong> onto glass and silicon oxide, synthesis <strong>of</strong> 11-<br />

azidoundecyltrimethoxysilane and ethynyl phosphonic acid, bulk fluorescence<br />

measurements on glass surfaces, optimization <strong>of</strong> copper-catalyzed <strong>azide</strong>-alkyne<br />

cycloaddition (CuAAC) reaction conditions, and acquisition <strong>of</strong> X-ray photoelectron<br />

spectra.<br />

Dr. Artin synthesized the alkyne-<strong>terminated</strong> oligonucleotide and Cy3-alkyne<br />

dye and contributed to the optimization <strong>of</strong> CuAAC conditions. She also performed the<br />

single molecule fluorescence measurements. David Lapham performed the<br />

hybridization experiments shown in Figure 4. Dr. Harris and Pr<strong>of</strong>. Chidsey originally<br />

proposed the idea to use the CuAAC reaction to attach oligonucleotides to <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong> on glass surfaces for single molecule sequencing by synthesis<br />

applications and contributed intellectually by providing guidance throughout the<br />

progression <strong>of</strong> this work. Pr<strong>of</strong>. Chidsey, Dr. Harris, and Dr. Artin also assisted in the<br />

preparation and the writing <strong>of</strong> the manuscript.<br />

99


Abstract<br />

A method is presented to covalently attach dye-labeled oligonucleotides to<br />

glass surfaces followed by hybridization with complementary oligonucleotides with<br />

minimal nonspecific adsorption. Azide-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> are first <strong>vapor</strong><br />

deposited onto glass surfaces and are found to be susceptible to the nonspecific<br />

adsorption <strong>of</strong> oligonucleotides. The nonspecific adsorption is reduced after the<br />

attachment <strong>of</strong> ethynyl phosphonic acid to the surface using the copper-catalyzed <strong>azide</strong>-<br />

alkyne cycloaddition (CuAAC) reaction. A two-step CuAAC reaction sequence is<br />

used to first attach Cy3-labeled, alkyne-<strong>terminated</strong> oligonucleotides followed by a<br />

second CuAAC reaction with ethynyl phosphonic acid as an agent to reduce<br />

nonspecific adsorption. After reaction, complementary oligonucleotides are captured<br />

on the surface by hybridization to the surface-attached oligonucleotides. Bulk and<br />

single-molecule fluorescence measurements are used to characterize the surfaces<br />

before and after exposure to dye-labeled oligonucleotides.<br />

100


Introduction<br />

In an effort to capture oligonucleotide targets by specific hybridization<br />

interactions between complementary nucleotides, a platform was developed to present<br />

covalently attached oligonucleotide probes surrounded by blocking agents to resist<br />

nonspecific adsorption on glass surfaces. The covalent attachment <strong>of</strong> biochemically<br />

viable oligonucleotides to surfaces is important for many technologies such as next-<br />

generation DNA sequencing and biochemical sensing. 1,2,3 Glass surfaces are attractive<br />

substrates for oligonucleotide immobilization because they are relatively cheap and<br />

compatible with optical measurements such as fluorescence spectroscopy where a low<br />

background fluorescence is desired. 4,5,6<br />

Oligonucleotides can be covalently bound to glass and other oxide surfaces by<br />

either appending an appropriate anchoring group, such as a silane, that will bond with<br />

the surface 7 or by introducing compatible reactive groups to the oligonucleotide and<br />

oxide surface and then allowing them to react. 6 The first method has disadvantages<br />

because tethering an anchoring group to an oligonucleotide may require difficult<br />

syntheses each time one wants to change the oligonucleotide sequence. Furthermore,<br />

the attachment <strong>of</strong> an anchoring group to a bulky oligonucleotide strand does not<br />

guarantee that the oligonucleotide will reproducibly form a monolayer or be viable for<br />

hybridization with complementary oligonucleotides once adsorbed.<br />

Decoupling the surface attachment <strong>of</strong> oligonucleotides into an adsorption step<br />

to introduce reactive functionality on the surface followed by a cross-coupling reaction<br />

to covalently attach oligonucleotides allows independent optimization at each step and<br />

is especially advantageous when the cross-coupling chemistry between the surface and<br />

101


oligonucleotides is robust. 8 The copper-catalyzed, <strong>azide</strong>-alkyne cycloaddition<br />

(CuAAC) reaction has been shown to be a nearly ideal cross-coupling chemistry for<br />

attaching species to surfaces. 9 Because the CuAAC reaction is such an ideal surface<br />

reaction, it has already been used to covalently attach functionalized oligonucleotides<br />

to <strong>azide</strong> 10,11,12,13<br />

and alkyne-<strong>terminated</strong> <strong>monolayers</strong>. 14,15,16 Azide-<strong>terminated</strong><br />

<strong>monolayers</strong> are favored because <strong>azide</strong> groups can be easily probed using Fourier<br />

transform infrared (FTIR) and x-ray photoelectron spectroscopy (XPS). 17 Azide-<br />

<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> can be used to introduce <strong>azide</strong> functionality onto glass<br />

and other oxide surfaces as previously reported. 18<br />

The use <strong>of</strong> a dense, hydrophobic monolayer and a selective surface chemistry<br />

such as the CuAAC reaction does not make surfaces immune to a major problem<br />

encountered when attaching biomolecules to them, nonspecific adsorption. Complex<br />

biomolecules such as oligonucleotides <strong>of</strong>ten nonspecifically adsorb onto hydrophobic<br />

monolayer surfaces from their bulk solutions upon exposure. 19 This nonspecific<br />

adsorption is undesirable for most applications and should be minimized. Various<br />

methods are used to reduce the nonspecific adsorption <strong>of</strong> biomolecules such as rinsing<br />

procedures 20 and the introduction <strong>of</strong> functional groups on the surface such as<br />

polyethylene glycols (PEGs), 21,22,23 acrylamide, 24 polysaccharides, 25,26 or charged<br />

species 27,28,29,30 that can add electrostatic or hydrophilic character to the surface to<br />

resist nonspecific adsorption. A mixed monolayer consisting <strong>of</strong> oligonucleotides<br />

surrounded by functional groups that can add nonspecific adsorption resistance to the<br />

surface during hybridization may be attainable using two CuAAC reactions as outlined<br />

in Figure 1.<br />

102


Fluorescence spectroscopy is a useful technique for monitoring the attachment<br />

<strong>of</strong> dye-labeled oligonucleotides to surfaces and their subsequent hybridization with<br />

dye-labeled complementary strands. 31,32 Single molecule fluorescence is a particularly<br />

sensitive technique for characterizing surfaces modified with dye-labeled<br />

oligonucleotides, and it has recently been used for sequencing by synthesis<br />

applications. 33,34 One <strong>of</strong> the problems encountered with performing fluorescence<br />

measurements is producing reactive, optical surfaces with sufficiently low background<br />

fluorescence to easily discern fluorescence attributable to the attachment <strong>of</strong> a dye-<br />

labeled species such as a biomolecule. 35,36 Aleman et al. recently reported single<br />

molecule fluorescence measurements <strong>of</strong> Cy3-labeled oligonucleotides that were<br />

covalently attached to solution-deposited, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong> on<br />

glass substrates using the CuAAC reaction. 14<br />

In this study, we demonstrate the covalent attachment <strong>of</strong> alkyne-<strong>terminated</strong>,<br />

dye-labeled oligonucleotides onto <strong>vapor</strong>-deposited, <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong><br />

<strong>monolayers</strong> using the CuAAC reaction. Bulk and single-molecule fluorescence<br />

spectroscopies are used to characterize the surfaces. The nonspecific adsorption <strong>of</strong><br />

dye-labeled oligonucleotides onto the <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> is reduced by<br />

forming a mixed monolayer <strong>of</strong> oligonucleotides and ethynyl phosphonic acid groups<br />

using two, sequential CuAAC reactions. This platform is used to capture<br />

complementary oligonucleotides on the surface via specific hybridization to covalently<br />

attached oligonucleotides with minimal nonspecific adsorption.<br />

103


(a) (b)<br />

1. CuAAC<br />

Oxide Surface Oxide Surface<br />

(d) (c)<br />

3. Hybridization<br />

104<br />

2. CuAAC<br />

Oxide Surface Oxide Surface<br />

Figure 1. Schematic <strong>of</strong> the platform developed for the attaching and hybridizing dye-<br />

labeled oligonucleotides on oxide surfaces. First, an <strong>azide</strong>-<strong>terminated</strong> monolayer is


<strong>vapor</strong> deposited as shown in (a) and then a CuAAC reaction is performed with alkyne-<br />

<strong>terminated</strong> oligonucleotides which results in covalently attached and nonspecifically<br />

adsorbed oligonucleotides as shown in (b). This is followed by rinsing and a CuAAC<br />

reaction with ethynyl phosphonic acid and another rinse to reduce nonspecific<br />

adsorption as shown in (c), and finally hybridization with a complementary<br />

oligonucleotide while still resisting nonspecific adsorption as shown in (d).<br />

Experimental<br />

Reagents<br />

11-azidoundecyltrimethoxysilane was synthesized as previously described in<br />

Chapter 3. TTMA (1,1',1''-Tris(1H-1,2,3-triazol-4-yl-1-acetic acid ethyl ester)<br />

trimethylamine) was synthesized following a published procedure. 37 Ethynyl<br />

phosphonic acid was synthesized according to a published procedure. 38,39,40<br />

Oligonucleotide Sources<br />

The Cy3-labeled, alkyne and amine-<strong>terminated</strong> oligonucleotides consisted <strong>of</strong><br />

32 bases with the following sequence: 5‟ - TCC ACT TAT CCT TGC ATC CAT CCT<br />

CTG CCC TG – 3‟. This sequence is designated RG2. Cy3 dyes were attached to the<br />

3‟ end, and terminal alkynes and amines were attached to the 5‟ end. For simplicity,<br />

oligonucleotides in this study are named Cy3-RG2-alkyne and Cy3-RG2-amine. The<br />

Cy3-RG2-amine oligonucleotide was purchased from Integrated DNA Technologies<br />

(IDT). The Cy3-RG2-alkyne oligonucleotide was synthesized and purified by HPLC<br />

at Helicos BioSciences by coupling an alkyne-<strong>terminated</strong> NHS (N-hydroxy<br />

succinimide) linker to the Cy3-RG2-amine oligonucleotide. Figure S1 in the<br />

Supporting Information shows the linker on this oligonucleotide. The Cy5-labeled<br />

105


complementary oligonucleotide consisted <strong>of</strong> 32 bases with the following sequence: 5‟<br />

- CAG GGC AGA GGA TGG ATG CAA GGA TAA GTG GA – 3‟. The Cy5 dye<br />

was attached to the 3‟ end. The noncomplementary oligonucleotide had a Cy5 dye<br />

attached to the 5‟ end and consisted <strong>of</strong> the following sequence: 5'- ACT GCT ACT<br />

GCT ACT GCT ACT GCT ACT GCC CAC AAA CCA AAA GCC CAG ACA CCC<br />

GGA GTT TTT TTT – 3‟.<br />

Surface Cleaning<br />

„Ultraclean‟ glass surfaces 40 mm in diameter (Erie Scientific) and single-side<br />

polished Si(100) wafers were cleaned in an oxygen plasma cleaner (GaLa Instrumente<br />

Plasma Prep5) for 10 min prior to <strong>vapor</strong> <strong>deposition</strong> with an oxygen flow rate <strong>of</strong> 100<br />

std. mL/min using 50 percent relative power at a pressure <strong>of</strong> 0.25 mbar. This cleaning<br />

method resulted in hydrophilic surfaces with water contact angles approaching 0°.<br />

Monolayer Vapor Deposition<br />

Siloxane <strong>monolayers</strong> were <strong>vapor</strong>-deposited onto silicon oxide surfaces in o-<br />

ring sealed, glass vacuum desiccators fitted with Teflon stopcocks (Jencons part<br />

number 250-048) with internal volumes <strong>of</strong> about 600 mL. Neat<br />

azidoundecyltrimethoxysilane (100 µL) was pipetted onto and absorbed into 42.5 mm<br />

diameter Whatman filter paper in the bottom <strong>of</strong> the desiccator. Water was deliberately<br />

added to the <strong>deposition</strong>s as a reactant for the hydrolysis reaction by adding 3 Å<br />

molecular sieves (Acros) with adsorbed water into the bottom <strong>of</strong> the desiccator in a<br />

foil boat. Water was adsorbed into the sieves by submerging 1.5 g <strong>of</strong> the as-received<br />

sieves in water for 5 min., filtering the excess water using a fritted glass filter, and<br />

then pumping on the sieves at room temperature for 5 minutes with a liquid nitrogen-<br />

106


trapped mechanical pump. The plasma-cleaned surfaces were placed on a metal rack<br />

in the chamber above the filter paper with absorbed silane. The desiccator was then<br />

evacuated through a rubber hose to a glass vacuum line with a liquid nitrogen-trapped<br />

mechanical pump for approximately 60 seconds. The final pressure at the trap was 1<br />

Torr. The Teflon valve on the desiccator was closed, and the chamber was placed in a<br />

110°C preheated oven for 12 h. After <strong>deposition</strong>, the valve was opened to ambient air<br />

to return the chamber to atmospheric pressure and to remove the samples.<br />

Surface Characterization<br />

Bulk fluorescence emission spectra were collected from the glass surfaces<br />

using a Horiba Jobin Yvon Fluorolog-3 spectr<strong>of</strong>luorometer equipped with a 450 watt<br />

Xenon lamp. Cy3 emission spectra were collected from 545-700 nm using 530 nm<br />

wavelength excitation light after passing through a bandpass filter and reaching the<br />

glass surface at normal incidence. Emission light was collected using front face<br />

detection and after passing through a bandpass filter. The bandpass on the emission<br />

and excitation monochromators was 5 nm for all experiments. Data were collected<br />

using an integration time <strong>of</strong> 0.5 s at 1 nm increments. Cy5 emission spectra were<br />

collected under the same conditions except 640 nm excitation light and bandpass<br />

filters for the Cy5 dye were used.<br />

Single molecule fluorescence images were obtained using a custom-built<br />

instrument at Helicos Biosciences 33 using 532 nm light for Cy3 excitation and 647 nm<br />

light for Cy5 excitation.<br />

107


Cu-Catalyzed Azide-Alkyne Cycloaddition<br />

CuAAC reactions were performed by exposing <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> to<br />

aqueous solutions containing 10 nM <strong>of</strong> an alkyne-<strong>terminated</strong> oligonucleotide or<br />

fluorescent dye, 500 µM CuSO4·5H2O, 500 µM TTMA, and 5 mM sodium ascorbate<br />

in 3:2 dimethyl sulfoxide (DMSO):water for 1 h at room temperature. Next, the<br />

surfaces were rinsedaccording to the standard buffer rinsing procedure outlined below.<br />

Ethynyl phosphonic acid CuAAC conditions were as follows: 1 mM ethynyl<br />

phosphonic acid, 500 µM CuSO4·5H2O, 500 µM TTMA, 5 mM sodium ascorbate in<br />

3:2 DMSO:water solvent, for 15 minutes at room temperature. Conditions for the no<br />

Cu(I) control experiments were the same except the CuSO4·5H2O was not added to the<br />

reaction mixture.<br />

Detergent and Buffer Rinsing Procedure<br />

After exposure to oligonucleotides, the <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> were<br />

subjected to a standard rinse procedure: water, 1 time in 3X sodium saline citrate<br />

(SSC), 2 times in 150 mM HEPES / 0.1% sodium dodecyl sulfate (SDS) / 1X SSC, 2<br />

times in 150 mM HEPES/NaCl, and water. The solutions were heated to 37°C while<br />

the water rinses were at ambient laboratory temperature. Approximately 100 mL <strong>of</strong> the<br />

solutions were used for each rinse. This rinsing procedure will be called the standard<br />

rinse in the rest <strong>of</strong> the paper.<br />

Hybridization<br />

Oligonucleotide hybridization experiments were performed by exposing<br />

surfaces to 10 nM oligonucleotide solutions in 3X SSC at 37°C for 15 min. After<br />

108


hybridization, the surfaces were rinsed 1 time in 3X SSC, 2 times in 150 mM HEPES/<br />

0.1% SDS/ 1X SSC, and 2 times in 150 mM HEPES/NaCl.<br />

Results and Discussion<br />

Nonspecific Adsorption <strong>of</strong> Oligonucleotides onto Azide-Terminated Monolayers<br />

Distinguishing between covalently attached and nonspecifically adsorbed<br />

biomolecules is a common problem when verifying their attachment to surfaces. To<br />

assess the nonspecific adsorption <strong>of</strong> dye-labeled oligonucleotides onto <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong>, we exposed the <strong>monolayers</strong> to unreactive Cy3-labeled, amine-<br />

<strong>terminated</strong> oligonucleotides (Cy3-RG2-amine). Figure 2a shows fluorescence<br />

emission spectra <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass before and after a 1 h<br />

exposure to 10 nM Cy3-RG2-amine in water. After exposure, an increase in counts<br />

above the background spectrum is observed even though the surfaces were rinsed with<br />

detergent and buffer solutions according to the standard rinsing procedure. This<br />

increase in fluorescence is attributed to Cy3 emission from nonspecifically adsorbed<br />

Cy3-labeled oligonucleotides because it is observed at the same wavelengths observed<br />

in the emission spectra <strong>of</strong> the Cy3-RG2-amine oligonucleotide in solution and for a<br />

Cy3-alkyne dye covalently attached to a glass surface (Figures S2a and b in the<br />

Supporting Information). In addition, the <strong>azide</strong> groups in the monolayer are unreactive<br />

with the functional groups present on the Cy3-RG2-amine oligonucleotides.<br />

Ethynyl Phosphonic Acid as a Blocking Agent to Reduce Nonspecific Adsorption<br />

We hypothesized that the introduction <strong>of</strong> negative charge on the surface might<br />

reduce the nonspecific adsorption <strong>of</strong> oligonucleotides by making the surface more<br />

hydrophilic and by electrostatically repelling the negatively charged phosphate<br />

109


ackbones <strong>of</strong> the oligonucleotides. Ethynyl phosphonic acid was synthesized and then<br />

covalently attached to <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> using the CuAAC reaction after<br />

exposure to Cy3-labeled oligonucleotides. As shown in Figure 2b, no increase in<br />

fluorescence above the background for the <strong>vapor</strong>-deposited <strong>azide</strong>-<strong>terminated</strong><br />

monolayer on glass was observed after a 1 h exposure <strong>of</strong> the monolayer to 10 nM<br />

Cy3-RG2-amine in water, standard rinse, 15 min CuAAC reaction with ethynyl<br />

phosphonic acid, and then another standard rinse. From this experiment, we infer that<br />

the introduction <strong>of</strong> the ethynyl phosphonic acids onto the surface was successful in<br />

reducing the nonspecific adsorption <strong>of</strong> Cy3-labeled oligonucleotides to the level <strong>of</strong><br />

noise in bulk fluorescence spectra.<br />

XPS was used to confirm the covalent attachment <strong>of</strong> the ethynyl phosphonic<br />

acid to an <strong>azide</strong>-<strong>terminated</strong> monolayer (see Figure S4 in the Supporting Information).<br />

Commercially available propiolic acid was also successful as a blocking agent to react<br />

with <strong>azide</strong> groups in the monolayer and to resist the nonspecific adsorption <strong>of</strong><br />

oligonucleotides as shown in Figure S5 in the Supporting Information.<br />

110


counts per second<br />

counts per second<br />

1.0x10 3<br />

7.5x10 2<br />

5.0x10 2<br />

2.5x10 2<br />

0.0<br />

1.0x10 3 (b)<br />

7.5x10 2<br />

5.0x10 2<br />

2.5x10 2<br />

0.0<br />

(a)<br />

Background<br />

After Cy3-RG2-amine Exposure<br />

Background<br />

After Cy3-RG2-amine Exposure<br />

and Ethynyl Phosphonic Acid<br />

CuAAC Reaction<br />

550 600 650 700<br />

Wavelength / nm<br />

Figure 2. Cy3 fluorescence emission spectra <strong>of</strong> <strong>vapor</strong>-deposited <strong>azide</strong>-<strong>terminated</strong><br />

<strong>monolayers</strong> on glass excited at 530 nm. (a) Emission <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> monolayer<br />

before (dots) and after (solid) exposure to 10 nM Cy3-RG2-amine and standard rinsing.<br />

(b) Emission <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> monolayer before (dots) and after (solid) exposure to<br />

10 nM Cy3-RG2-amine followed by a standard rinse, 15 min ethynyl phosphonic acid<br />

CuAAC reaction, and another standard rinse.<br />

Covalent Attachment <strong>of</strong> Alkyne-Terminated Oligonucleotides to Monolayers<br />

Once a method for reducing the nonspecific adsorption <strong>of</strong> Cy3-labeled<br />

oligonucleotides onto <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> was developed, the <strong>azide</strong> groups<br />

111


were reacted with alkyne-<strong>terminated</strong>, Cy3-labeled oligonucleotides (Cy3-RG2-alkyne)<br />

using CuAAC reaction conditions. Figure 3a shows emission spectra <strong>of</strong> an <strong>azide</strong>-<br />

<strong>terminated</strong> monolayer on glass before and after a 1 h CuAAC reaction with 10 nM<br />

Cy3-RG2-alkyne in water followed by a standard rinse, a 15 min CuAAC reaction<br />

with ethynyl phosphonic acid, and another standard rinse. A peak is observed in the<br />

fluorescence spectrum after the CuAAC reactions and rinsing, and it is assigned to<br />

Cy3 emission from covalently attached Cy3-labeled oligonucleotides. To further test<br />

against nonspecific adsorption, emission spectra were collected from an <strong>azide</strong>-<br />

<strong>terminated</strong> monolayer before and after a 1 h exposure <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong><br />

monolayer to the same CuAAC reaction and rinsing conditions used in Figure 3a but<br />

without the Cu(I) catalyst. Figure 3b shows that no increase in fluorescence above the<br />

background is observed after exposure without the Cu(I) catalyst providing further<br />

confirmation that the increase in fluorescence with Cu(I) present is attributable to<br />

covalently attached Cy3-labeled oligonucleotides.<br />

In addition to bulk fluorescence spectroscopy, single molecule fluorescence<br />

spectroscopy was used to characterize the <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass after<br />

exposure to Cy3-labeled oligonucleotides. Figure 3c shows a single molecule<br />

fluorescence image after a 1 h CuAAC reaction with 0.1 nM Cy3-RG2-alkyne<br />

followed by a standard rinse, 15 min ethynyl phosphonic acid CuAAC reaction, and<br />

another standard rinse. The white objects in the image are attributable to fluorescence<br />

emission on the surface resulting from the 532 nm excitation light. Quantification <strong>of</strong><br />

the objects results in approximately 1589 fluorescent objects per 1000 µm 2 . A single<br />

molecule fluorescence image after a 1 h exposure <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer to<br />

112


0.1 nM Cy3-RG2-alkyne under the same conditions but without the Cu(I) catalyst is<br />

shown in Figure 3d. The number density <strong>of</strong> fluorescent objects is 6 times less without<br />

the Cu(I) catalyst present than with it, approximately 250 objects per 1000 µm 2 , which<br />

is similar to the background number density <strong>of</strong> fluorescent objects measured on glass<br />

surfaces after monolayer <strong>vapor</strong> <strong>deposition</strong>. The larger number density <strong>of</strong> fluorescent<br />

objects in the single molecule image after exposure to the Cy3-labeled<br />

oligonucleotides with the Cu(I) catalyst present is consistent with the bulk<br />

fluorescence experiments and provides additional evidence that the alkyne-<strong>terminated</strong><br />

oligonucleotides are being covalently attached. The oligonucleotide concentration<br />

used during the CuAAC reaction was decreased 100-fold from the 10 nM used for<br />

bulk fluorescence measurements to 0.1 nM for the single molecule measurements to<br />

control the number density <strong>of</strong> attached oligonucleotides so that single molecules could<br />

be resolved. This experiment also demonstrates the utility <strong>of</strong> the <strong>vapor</strong>-deposited,<br />

<strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> for single molecule fluorescence measurements because<br />

their background fluorescence is low enough to clearly observe fluorescence emission<br />

from dyes on the surface.<br />

113


counts per second<br />

counts per second<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

(a)<br />

(b)<br />

Background<br />

After Cy3-RG2-alkyne CuAAC<br />

Background<br />

After Cy3-RG2-alkyne Exposure<br />

550 600 650 700<br />

Wavelength / nm<br />

114<br />

(c)<br />

0.1 nM, with Cu(I), 1589 objects/1000 µm 2<br />

(d)<br />

0.1 nM, no Cu(I), 256 objects/1000 µm 2<br />

Figure 3. Cy3 bulk fluorescence emission spectra and single molecule fluorescence<br />

images <strong>of</strong> <strong>vapor</strong>-deposited <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass. (a) Emission <strong>of</strong> an<br />

<strong>azide</strong>-<strong>terminated</strong> monolayer before (dots) and after (solid) a 1 h CuAAC reaction with<br />

10 nM Cy3-RG2-alkyne. (b) Emission <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer before (dots)<br />

and after (solid) a 1 h exposure to 10 nM Cy3-RG2-alkyne without the Cu(I) catalyst.<br />

(c) Single molecule fluorescence image <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer after a 1 h<br />

CuAAC reaction with 0.1 nM Cy3-RG2-alkyne. (d) Single molecule fluorescence<br />

image <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer after a 1 h exposure to 0.1 nM Cy3-RG2-<br />

alkyne without the Cu(I) catalyst. Each CuAAC reaction and exposure was followed


y a standard rinse, a 15 min ethynyl phosphonic acid CuAAC reaction, and another<br />

standard rinse. The bulk fluorescence spectra were collected using 530 nm excitation<br />

light while 532 nm excitation light was used for the single molecule fluorescence<br />

images.<br />

Hybridization <strong>of</strong> Complementary Oligonucleotides to Covalently Attached<br />

Oligonucleotides<br />

To assess the viability <strong>of</strong> the single-stranded oligonucleotides after covalent<br />

attachment to <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> using the CuAAC reaction, hybridization<br />

experiments with a Cy5-labeled complementary oligonucleotide were conducted.<br />

Figure 4a shows Cy5 bulk fluorescence emission spectra <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong><br />

monolayer on glass that had been reacted with the Cy3-RG2-oligonucleotide and<br />

ethynyl phosphonic acid before and after a 15 min exposure to 10 nM <strong>of</strong> a Cy5-labeled<br />

complementary oligonucleotide in 3x SSC followed by rinsing. After exposure, an<br />

increase in fluorescence above the background is observed that is consistent with Cy5<br />

emission. 41 From the increase in fluorescence, we infer that Cy5-labeled,<br />

complementary oligonucleotides hybridized to the covalently attached<br />

oligonucleotides on the surface.<br />

Two control experiments were performed to confirm the specific hybridization<br />

<strong>of</strong> the Cy5-labeled complementary oligonucleotide on the surface. For the first control,<br />

a nominally identical glass surface to the one used above was prepared by covalently<br />

attaching the Cy3-RG2-alkyne and reacting the remaining <strong>azide</strong> groups with ethynyl<br />

phosphonic acid. This surface was then exposed to 10 nM <strong>of</strong> a non-complementary,<br />

115


Cy5-labeled oligonucleotide using the same hybridization conditions as used for the<br />

complementary oligonucleotide. As shown in Figure 4b, no increase in fluorescence<br />

was observed after exposure and rinsing. This result suggests that the non-<br />

complementary oligonucleotide did not hybridize as expected and that it did not<br />

nonspecifically adsorb onto the surface as measured using bulk fluorescence.<br />

A second control experiment was performed that tested the nonspecific<br />

adsorption <strong>of</strong> the Cy5-labeled complementary oligonucleotide onto an <strong>azide</strong>-<br />

<strong>terminated</strong> monolayer that had been reacted with ethynyl phosphonic acid. Figure 4c<br />

shows that the Cy5 fluorescence emission spectra are the same before and after<br />

exposure <strong>of</strong> the phosphonic acid-<strong>terminated</strong> monolayer to 10 nM <strong>of</strong> the Cy5-labeled<br />

complementary oligonucleotide followed by standard rinsing. It is inferred that the<br />

Cy5-labeled oligonucleotide did not nonspecifically adsorb onto the surface because<br />

no increase in fluorescence was observed after exposure. Overall, the experiments<br />

represented in Figure 4 demonstrate the utility <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong><br />

<strong>monolayers</strong> on glass for the covalent attachment and subsequent hybridization <strong>of</strong><br />

oligonucleotides while resisting nonspecific adsorption.<br />

116


counts per second<br />

counts per second<br />

counts per second<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

(a)<br />

(b)<br />

(c)<br />

650 700 750 800<br />

Wavelength / nm<br />

117<br />

Background<br />

After Hybridization<br />

Background<br />

After Exposure<br />

Background<br />

After Exposure<br />

Figure 4. Cy5 fluorescence emission spectra <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass<br />

exposed to hybridization conditions for Cy5-labeled oligonucleotides. (a) Azide-<br />

<strong>terminated</strong> monolayer reacted for 1 h with 10 nM Cy3-RG2-alkyne before and after


exposure to the Cy5-labeled complementary oligonucleotide and standard rinsing. (b)<br />

Azide-<strong>terminated</strong> monolayer reacted for 1 h with 10 nM Cy3-RG2-alkyne before and<br />

after exposure to a Cy5-labeled non-complementary oligonucleotide and standard<br />

rinsing. (c) Azide-<strong>terminated</strong> monolayer reacted with ethynyl phosphonic acid before<br />

and after exposure to the Cy5-labeled complementary oligonucleotide and standard<br />

rinsing. 640 nm excitation light was used for all spectra.<br />

Resistance to Nonspecific Adsorption during Single Molecule Sequencing by<br />

Synthesis<br />

The strategy developed in this paper for resisting the nonspecific adsorption <strong>of</strong><br />

dye-labeled oligonucleotides was applied to resisting the nonspecific adsorption <strong>of</strong><br />

dye-labeled mononucleotides during single molecule sequencing by synthesis. A<br />

major challenge for sequencing by synthesis technologies is the development <strong>of</strong><br />

surfaces that resist the nonspecific adsorption <strong>of</strong> oligo- and mono-nucleotides to<br />

achieve predictable rinsing on the surface. The inability to rinse unincorporated<br />

mononucleotides from the surface can lead to increased substitution and error rates<br />

during sequencing. 2<br />

To test our strategy, <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> were first reacted with an<br />

alkyne-<strong>terminated</strong> oligonucleotide followed by a second reaction <strong>of</strong> the <strong>azide</strong>s with<br />

propargyl alcohol as a blocking agent. Next, oligonucleotides templates were captured<br />

on the surface via hybridization to covalently attached oligonucleotides. Using the<br />

single molecule sequencing by synthesis platform developed by Helicos<br />

BioSciences, 33 the surfaces were then sequentially exposed to dye-labeled<br />

mononucleotides in the presence <strong>of</strong> DNA polymerase for the template-driven<br />

118


enzymatic extension at the 3‟-end <strong>of</strong> the covalently attached oligonucleotide strands.<br />

The surfaces were then rinsed according to the same aqueous and buffer rinses used in<br />

the oligonucleotide hybridization experiments. Single molecule fluorescence<br />

spectroscopy was used to confirm the template-driven enzymatic addition <strong>of</strong> the dye-<br />

labeled mononucleotides to the surfaces and to quantify unincorporated dye-labeled<br />

mononucleotides that were not removed by rinsing.<br />

Table S1 in the Supporting Information shows the number <strong>of</strong> uncorrelated<br />

fluorescent objects measured during single molecule sequencing on <strong>azide</strong>-<strong>terminated</strong><br />

<strong>monolayers</strong> that were reacted with alkyne-<strong>terminated</strong> oligonucleotides and propargyl<br />

alcohol. The uncorrelated objects are attributed to unincorporated, nonspecifically<br />

adsorbed dye-labeled mononucleotides that remained on the surface after rinsing. For<br />

comparison, Table S2 shows the number <strong>of</strong> unincorporated mononucleotides for<br />

commercially produced epoxide surfaces that were reacted with amine-<strong>terminated</strong><br />

oligonucleotides and phosphate. The number <strong>of</strong> unincorporated mononucleotides<br />

above the background fluorescence is approximately 4-5 times lower on the <strong>azide</strong>-<br />

<strong>terminated</strong> <strong>monolayers</strong> than for the epoxide-<strong>terminated</strong> <strong>monolayers</strong>. Perhaps there is a<br />

higher density <strong>of</strong> hydroxyl groups, which have been shown to resist the nonspecific<br />

adsorption <strong>of</strong> nucleotides, 42 on the <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> after reaction with<br />

propargyl alcohol than on the epoxide surfaces after reaction with phosphate. 43 The<br />

longer, 11-carbon alkyl chains in the <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> should also<br />

promote denser, more uniform <strong>monolayers</strong> compared to the shorter, 3-carbon alkyl<br />

chains in the epoxide-<strong>terminated</strong> <strong>monolayers</strong>. The increase in monolayer density and<br />

uniformity may be important for resisting the nonspecific adsorption <strong>of</strong> nucleotides.<br />

119


Conclusions<br />

In summary, we have demonstrated a strategy for covalently attaching alkyne-<br />

<strong>terminated</strong> oligonucleotides onto <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass surfaces that<br />

were viable for hybridization with complementary strands in solution with minimal<br />

nonspecific adsorption. CuAAC reactions between <strong>azide</strong> groups and ethynyl<br />

phosphonic acid reduced the nonspecific adsorption <strong>of</strong> dye-labeled oligonucleotides to<br />

the level <strong>of</strong> noise in bulk fluorescence spectra. An optimized procedure involved an<br />

initial CuAAC reaction to covalently attach alkyne-<strong>terminated</strong> oligonucleotides<br />

followed by a second CuAAC reaction with ethynyl phosphonic acid to create surfaces<br />

that resist nonspecific adsorption but allow complementary oligonucleotide targets to<br />

be captured via hybridization. Bulk and single molecule fluorescence measurements<br />

were used to confirm the presence <strong>of</strong> dye-labeled oligonucleotides on the surfaces.<br />

Acknowledgments<br />

This work was supported by funding from Helicos BioSciences Corporation. The<br />

authors acknowledge NSF grant CHE-0639053 for support <strong>of</strong> the spectr<strong>of</strong>luorometer.<br />

The authors also acknowledge NSF grant DMR-0213618 to the Center on Polymer<br />

Interfaces and Macromolecular Assemblies for use <strong>of</strong> the plasma cleaner.<br />

120


Supporting Information<br />

Figure S1. Reaction scheme for the synthesis <strong>of</strong> the alkyne-<strong>terminated</strong> oligonucleotide<br />

that was used in this study for CuAAC reactions with <strong>azide</strong>s. The synthesis was<br />

performed by reacting an NHS-activated alkyne with the primary amine on the 5‟ end<br />

<strong>of</strong> a Cy3-labeled oligonucleotide as shown. .<br />

Cy3 Fluorescence in Solution and on Glass Surfaces<br />

Figure S2a shows a fluorescence emission spectrum <strong>of</strong> an aqueous solution<br />

containing 10 nM <strong>of</strong> the Cy3-labeled oligonucleotide, Cy3-RG2-amine, excited at 530<br />

nm. A peak is observable in the emission spectrum at 565 nm, consistent with Cy3<br />

fluorescence. 41 A Cy3 dye with a terminal alkyne (Cy3-alkyne) was synthesized to test<br />

the detection <strong>of</strong> Cy3 fluorescence on glass surfaces. Its chemical structure is shown in<br />

Figure S3. The CuAAC reaction was used to covalently attach the Cy3-alkyne to<br />

<strong>azide</strong>-<strong>terminated</strong> <strong>siloxane</strong> <strong>monolayers</strong>. Figure S2b shows fluorescence emission<br />

spectra <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer on glass before and after a 1 h CuAAC<br />

121


eaction performed by exposing the surface to 10 nM <strong>of</strong> the Cy3-alkyne in water with<br />

the Cu(I) catalyst present followed by a standard rinse, CuAAC reaction with ethynyl<br />

phosphonic acid, and another standard rinse. After the reaction, an increase in<br />

fluorescence above the background spectrum is observed in the same wavelength<br />

region as observed for the Cy3-labeled oligonucleotide in solution. This increase in<br />

fluorescence is attributed to emission from covalently attached Cy3 dye molecules.<br />

Figure S2c shows fluorescence emission spectra <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer<br />

before and after a 1 h exposure to 10 nM <strong>of</strong> the Cy3-alkyne without the Cu(I) catalyst<br />

followed by the standard rinsing procedure. No increase in fluorescence is observed<br />

without the Cu(I) catalyst providing further evidence that the Cy3-alkyne is covalently<br />

attached to the surface when the Cu(I) catalyst is present and that fluorescence<br />

spectroscopy can be used to detect the presence <strong>of</strong> Cy3 dyes on glass surfaces.<br />

122


counts per second<br />

counts per second<br />

counts per second<br />

3.0x10 5<br />

2.5x10 5<br />

2.0x10 5<br />

1.5x10 5<br />

1.0x10 5<br />

5.0x10 4<br />

0.0<br />

3.0x10 3<br />

2.5x10 3<br />

2.0x10 3<br />

1.5x10 3<br />

1.0x10 3<br />

5.0x10 2<br />

0.0<br />

3.0x10 3<br />

2.5x10 3<br />

2.0x10 3<br />

1.5x10 3<br />

1.0x10 3<br />

5.0x10 2<br />

0.0<br />

(a)<br />

(b)<br />

(c)<br />

550 600 650 700<br />

Wavelength / nm<br />

123<br />

10 nM Cy3-RG2-amine<br />

in water<br />

Background<br />

After Cy3-alkyne CuAAC<br />

Background<br />

After Cy3-alkyne exposure<br />

Figure S2. (a) Cy3 fluorescence emission spectrum <strong>of</strong> 10 nM <strong>of</strong> a Cy3-labeled<br />

oligonucleotide, Cy3-RG2-amine, in water excited at 530 nm. Cy3 fluorescence<br />

emission spectra <strong>of</strong> <strong>vapor</strong>-deposited <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass before<br />

(dots) and after (solid) exposure to 10 nM Cy3-alkyne with (b) and without (c) the


Cu(I) catalyst for the CuAAC reaction present. The glass surfaces were excited with<br />

530 nm light. After exposure and reaction, the surfaces were subjected to: a standard<br />

rinse, a CuAAC reaction with ethynyl phosphonic acid, and another standard rinse.<br />

Figure S3. Chemical structure <strong>of</strong> the alkyne-<strong>terminated</strong> Cy3 dye (Cy3-alkyne) that<br />

was covalently attached to <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> using the CuAAC reaction.<br />

XPS Confirmation <strong>of</strong> Reaction with Ethynyl Phosphonic Acid<br />

X-ray photoelectron spectroscopy (XPS) provides a convenient probe for<br />

monitoring CuAAC reactions on <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong>, 17,44 and it was used to<br />

confirm the covalent attachment <strong>of</strong> ethynyl phosphonic acid. Figure S4 shows a high-<br />

resolution N 1s spectrum <strong>of</strong> an <strong>azide</strong>-<strong>terminated</strong> monolayer on silicon oxide. An<br />

asymmetric doublet is present because the <strong>azide</strong> group contains nitrogen atoms in two<br />

chemically distinct environments. The higher binding energy peak is attributed to the<br />

electron-deficient, central nitrogen atom while the lower binding energy peak is<br />

assigned to the outer, more electron-rich nitrogen atoms. 45 Therefore, the peak ratio<br />

should ideally be 1:2 as shown here. After a CuAAC reaction with ethynyl phosphonic<br />

124


acid, the higher binding energy peak disappears leaving only one N 1s peak present in<br />

the high-resolution XPS spectrum (Figure S4d), which is attributable to the three<br />

nitrogen atoms in the triazole product. The quantitative reaction <strong>of</strong> the <strong>azide</strong> groups is<br />

inferred from the disappearance <strong>of</strong> the higher binding energy N 1s photoelectron peak.<br />

In addition, a new peak is observed in the high-resolution P 2p spectrum at 133 eV<br />

after the CuAAC reaction with ethynyl phosphonic acid; whereas, no P 2p was present<br />

before reaction as shown in Figures S4a and b. The presence <strong>of</strong> the P 2p peak after the<br />

reaction is attributed to the phosphorus atoms in the covalently attached ethynyl<br />

phosphonic acid molecules. 46<br />

Counts<br />

Counts<br />

(a)<br />

100<br />

(b)<br />

100<br />

P 2p before<br />

CuAAC<br />

P 2p after<br />

CuAAC<br />

125 130 135 140 390 400 410<br />

Binding Energy / eV<br />

125<br />

(c)<br />

500<br />

(d)<br />

500<br />

N 1s before<br />

CuAAC<br />

Binding Energy / eV<br />

N 1s after<br />

CuAAC<br />

Figure S4. High-resolution XPS spectra <strong>of</strong> the P 2p [(a) and (b)] and N 1s [(c) and<br />

(d)] photoelectron regions for an <strong>azide</strong>-<strong>terminated</strong> monolayer on silicon oxide before<br />

and after a CuAAC reaction with 10 mM ethynyl phosphonic acid in 3:2 DMSO:H2O.<br />

Counts<br />

Counts


Spectra were obtained using a Surface Science Model 150 spectrometer with Al Kα<br />

radiation (1486.6 eV). 10 scans were collected at 0.1 eV resolution at a take<strong>of</strong>f angle<br />

<strong>of</strong> 35° from the surface.<br />

Propiolic Acid as a Blocking Agent to Reduce Nonspecific Adsorption<br />

Figure S5 shows fluorescence emission spectra for <strong>azide</strong>-<strong>terminated</strong><br />

<strong>monolayers</strong> on glass exposed to 10 nM <strong>of</strong> the alkyne-<strong>terminated</strong> oligonucleotide, Cy3-<br />

RG2-alkyne. Figure S5a shows an increase in fluorescence above the background<br />

when an <strong>azide</strong>-<strong>terminated</strong> monolayer was exposed to 10 nM <strong>of</strong> the oligonucleotide for<br />

1 hour without Cu(I) present followed by standard rinsing without propiolic acid.<br />

Figure S5b shows that reacting propiolic acid with the <strong>azide</strong> groups in the monolayer<br />

after exposure to the Cy3-RG2-alkyne and standard rinsing reduces the nonspecific<br />

adsorption <strong>of</strong> oligonucleotides to the level <strong>of</strong> the background. Figure S5c shows<br />

fluorescence emission spectra before and after CuAAC reaction with 10 nM <strong>of</strong> the<br />

Cy3-RG2-alkyne followed by a standard rinse, CuAAC reaction with propiolic acid,<br />

and another standard rinse. As in the Cy3-alkyne studies, addition <strong>of</strong> the Cu(I) catalyst<br />

results in covalent attachment <strong>of</strong> the alkyne-<strong>terminated</strong> oligonucleotide as inferred<br />

from the increase in the fluorescence emission above the background.<br />

126


counts per second<br />

counts per second<br />

counts per second<br />

1000<br />

750<br />

500<br />

250<br />

0<br />

1000<br />

750<br />

500<br />

250<br />

0<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

(a)<br />

(b)<br />

(c)<br />

550 600 650 700<br />

Wavelength / nm<br />

127<br />

Background<br />

After Exposure<br />

Background<br />

After Exposure<br />

Background<br />

After CuAAC<br />

Figure S5. Fluorescence emission spectra <strong>of</strong> <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong> on glass<br />

before and after a 1 h exposure to 10 nM <strong>of</strong> the Cy3-RG2-alkyne: (a) without Cu(I)<br />

present and standard rinsing; (b) without Cu(I) present followed by a standard rinse,


propiolic acid CuAAC, and another standard rinse; (c) with Cu(I) present followed by<br />

a standard rinse, propiolic acid CuAAC, and another standard rinse.<br />

Monolayer Resistance to Nonspecific Adsorption During Single Molecule<br />

Sequencing by Synthesis<br />

Once the nonspecific adsorption <strong>of</strong> dye-labeled oligonucleotides was reduced<br />

to the level <strong>of</strong> noise as measured by bulk fluorescence and single molecule<br />

fluorescence (Figure 3), the nonspecific adsorption <strong>of</strong> dye-labeled mononucleotides<br />

onto the <strong>monolayers</strong> was characterized using single molecule fluorescence<br />

spectroscopy, a more sensitive probe. Table S1 shows the average number <strong>of</strong><br />

uncorrelated fluorescent objects per 1000 µm 2 for 180 measurements during single<br />

molecule sequencing by synthesis on <strong>vapor</strong>-deposited, <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong><br />

that were reacted with alkyne-<strong>terminated</strong> oligonucleotides and commercially available<br />

propargyl alcohol as a blocking agent followed by a standard rinse. The uncorrelated<br />

objects represent nonspecifically adsorbed, dye-labeled nucleotides on the surface that<br />

were not incorporated into a growing primer-template strand. It is inferred that the<br />

surfaces exhibit high resistance to the nonspecific adsorption <strong>of</strong> dye-labeled<br />

nucleotides because the number <strong>of</strong> uncorrelated objects is only a few tens <strong>of</strong> counts<br />

above the background.<br />

128


Table S1. Nonspecific Adsorption <strong>of</strong> Dye-Labeled Mononucleotides onto Azide-<br />

Terminated Monolayers Reacted with Alkyne-Terminated Oligonucleotides and<br />

Propargyl Alcohol during Single Molecule Sequencing by Synthesis<br />

Average rinse for each nucleotide base (objects per 1000 µm 2 )<br />

Background (X) Adenine (A) Cytosine (C) Guanine (G) Uracil (U)<br />

145.5 188.6 168.3 169 190.2<br />

Average rinse minus background (objects per 1000 µm 2 )<br />

A-X C-X G-X U-X<br />

43.1 22.8 23.5 44.8<br />

Number <strong>of</strong> positions (cycles) sampled for calculation<br />

390 180 180 180 180<br />

For comparison, a similar experiment was performed on epoxide-<strong>terminated</strong><br />

<strong>siloxane</strong> <strong>monolayers</strong> that were commercially <strong>vapor</strong> deposited onto glass surfaces and<br />

then reacted with amine-<strong>terminated</strong> oligonucleotides followed by a standard rinse and<br />

blocking in a phosphate solution. 43 Table S2 shows the average number <strong>of</strong><br />

uncorrelated fluorescent objects per 1000 µm 2 for 240 measurements on these surfaces<br />

during single molecule sequencing by synthesis. The <strong>azide</strong>-<strong>terminated</strong> <strong>monolayers</strong><br />

reacted with propargyl alcohol are more resistant to the nonspecific adsorption <strong>of</strong> dye-<br />

labeled mononucleotides because the numbers <strong>of</strong> uncorrelated objects minus the<br />

129


ackgrounds for each base are lower than the number <strong>of</strong> uncorrelated objects<br />

measured on the epoxide-<strong>terminated</strong> <strong>monolayers</strong>.<br />

Table S2. Nonspecific Adsorption <strong>of</strong> Dye-Labeled Mononucleotides onto Epoxide-<br />

Terminated Monolayers Reacted with Amine-Terminated Oligonucleotides and<br />

Phosphate during Single Molecule Sequencing by Synthesis<br />

Average rinse for each nucleotide base (objects per 1000 µm 2 )<br />

Background (X) Adenine (A) Cytosine (C) Guanine (G) Uracil (U)<br />

50.2 246.6 198.9 147.5 288.4<br />

Average rinse minus background (objects per 1000 µm 2 )<br />

A-X C-X G-X U-X<br />

196.4 148.7 97.3 238.2<br />

Number <strong>of</strong> positions (cycles) sampled for calculation<br />

510 240 240 240 240<br />

130


References<br />

1. Nandivada, H.; Lahann, J. Copper-Catalyzed „Click‟ Chemistry for Surface<br />

Engineering. In Click Chemistry for Biotechnology and Materials Science; J. Lahann,<br />

Ed.; John Wiley & Sons, Ltd: Chichester, UK, 2009, pp 291-307.<br />

2. Fuller, C. W.; Middendorf, L. R.; Benner, S. A.; Church, G. M.; Harris, T.;<br />

Xiaohua, H.; Jovanovich, S. B.; Nelson, J. R.; Schloss, J. A.; Schwartz, D. C.;<br />

Vezenov, D. V. Nature Biotechnology 2009, 27, 1013-1023.<br />

3. Niedringhaus, T. P.; Milanova, D.; Kerby, M. B.; Snyder, M. P.; Barron, A. E.<br />

Analytical Chemistry 2011, 83, 4327-4341.<br />

4. Laib, S.; Krieg, A.; Rankl, M.; Seeger, S. Applied Surface Science 2006, 252,<br />

7788-7793.<br />

5. Vong, T.; ter Maat, J.; van Beek, T. A.; van Lagen, B.; Giesbers, M.; van Hest,<br />

J. C. M.; Zuilh<strong>of</strong>, H. Langmuir 2009, 25, 13952-13958.<br />

6. Zammatteo, N.; Jeanmart, L.; Hamels, S.; Courtois, S.; Louette, P.; Hevesi, L.;<br />

Remacle, J. Analytical Biochemistry 2000, 280, 143-150.<br />

7. Venkatesh, A. G.; Herth, S.; Becker, A.; Reiss, G. Nanotechnology 2011, 22,<br />

145301.<br />

8. Haensch, C.; Hoeppener, S.; Schubert, U. S. Chemical Society Reviews 2010,<br />

39, 2323-2334.<br />

9. Meldal, M.; Tornoe, C. W. Chem. Rev. 2008, 108, 2952-3015.<br />

10. Qing, G.; Xiong, H.; Seela, F.; Sun, T. Journal <strong>of</strong> the American Chemical<br />

Society 2010, 132, 15228-15232.<br />

131


11. Devaraj, N. K.; Miller, G. P.; Ebina, W.; Kakaradov, B.; Collman, J. P.; Kool,<br />

E. T.; Chidsey, C. E. D. J. Am. Chem. Soc. 2005, 127, 8600-8601.<br />

12. Rozkiewicz, D. I.; Gierlich, J.; Burley, G. A.; Gutsmiedl, K.; Carell, T.; Ravoo,<br />

B. J.; Reinhoudt, D. N. ChemBioChem 2007, 8, 1997-2002.<br />

13. Chung, M.; Lowe, R. D.; Chan, Y.-H. M.; Ganesan, P. V.; Boxer, S. G.<br />

Journal <strong>of</strong> Structural Biology 2009, 168, 190-199.<br />

14. Alemán, E. A.; Pedini, H. S.; Rueda, D. ChemBioChem 2009, 10, 2862-2866.<br />

15. Ciampi, S.; James, M.; Michaels, P.; Gooding, J. J. Langmuir 2011, 27, 6940-<br />

6949.<br />

16. Seo, T. S.; Bai, X.; Ruparel, H.; Li, Z.; Turro, N. J.; Ju, J. PNAS 2004, 101,<br />

5488-5493.<br />

17. Collman, J. P.; Devaraj, N. K.; Eberspacher, T. P. A.; Chidsey, C. E. D.<br />

Langmuir 2006, 22, 2457-2464.<br />

18. Lummerstorfer, T.; H<strong>of</strong>fmann, H. J. Phys. Chem. B 2004, 108, 3963-3966.<br />

19. Love, J. C.; Estr<strong>of</strong>f, L. A.; Kriebel, J. K.; Nuzzo, R. G.; Whitesides, G. M.<br />

Chemical Reviews 2005, 105, 1103-1170.<br />

20. Vandenberg, E.; Elwing, H.; Askendal, A.; Lundström, I. Journal <strong>of</strong> Colloid<br />

and Interface Science 1991, 143, 327-335.<br />

21. Prime, K. L.; Whitesides, G. M. Science 1991, 252, 1164-1167.<br />

22. Sigal, G. B.; Mrksich, M.; Whitesides, G. M. Journal <strong>of</strong> the American<br />

Chemical Society 1998, 120, 3464-3473.<br />

23. Cattani-Scholz, A.; Pedone, D.; Blobner, F.; Abstreiter, G.; Schwartz, J.;<br />

Tornow, M.; Andruzzi, L. Biomacromolecules 2009, 10, 489-496.<br />

132


24. Vikholm-Lundin, I.; Piskonen, R.; Albers, W. M. Biosensors and<br />

Bioelectronics 2007, 22, 1323-1329.<br />

25. Jung, S.; Angerer, B.; Löscher, F.; Niehren, S.; Winkle, J.; Seeger, S.<br />

ChemBioChem 2006, 7, 900-903.<br />

26. Holland, N. B.; Qiu, Y.; Ruegsegger, M.; Marchant, R. E. Nature 1998, 392,<br />

799-801.<br />

27. Alexander, N. A.; Igor, Y. S.; Philip, B. O. In Effect <strong>of</strong> surface electrostatics<br />

on adsorption behavior and biospecific interactions <strong>of</strong> DNA oligonucleotides; Gerald,<br />

E. C.; John, C. O., Eds.; Proc. SPIE; 1999; 3603, pp 170-177.<br />

28. Kuga, S.; Yang, J.-H.; Takahashi, H.; Hirama, K.; Iwasaki, T.; Kawarada, H.<br />

Journal <strong>of</strong> the American Chemical Society 2008, 130, 13251-13263.<br />

29. Cárdenas, M.; Nylander, T.; Lindman, B. Colloids and Surfaces A:<br />

Physicochemical and Engineering Aspects 2005, 270-271, 33-43.<br />

30. Cárdenas, M.; Campos-Terán, J.; Nylander, T.; Lindman, B. Langmuir 2004,<br />

20, 8597-8603.<br />

31. Oh, S. J.; Cho, S. J.; Kim, C. O.; Park, J. W. Langmuir 2002, 18, 1764-1769.<br />

32. Schlapak, R.; Pammer, P.; Armitage, D.; Zhu, R.; Hinterdorfer, P.; Vaupel, M.;<br />

Frühwirth, T.; Howorka, S. Langmuir 2005, 22, 277-285.<br />

33. Harris, T. D.; Buzby, P. R.; Babcock, H.; Beer, E.; Bowers, J.; Braslavsky, I.;<br />

Causey, M.; Colonell, J.; DiMeo, J.; Efcavitch, J. W.; Giladi, E.; Gill, J.; Healy, J.;<br />

Jarosz, M.; Lapen, D.; Moulton, K.; Quake, S. R.; Steinmann, K.; Thayer, E.; Tyurina,<br />

A.; Ward, R.; Weiss, H.; Xie, Z. Science 2008, 320, 106-109.<br />

133


34. Pushkarev, D.; Neff, N. F.; Quake, S. R. Nature Biotechnology 2009, 27, 847-<br />

850.<br />

35. Korlach, J.; Marks, P. J.; Cicero, R. L.; Gray, J. J.; Murphy, D. L.; Roitman, D.<br />

B.; Pham, T. T.; Otto, G. A.; Foquet, M.; Turner, S. W. Proceedings <strong>of</strong> the National<br />

Academy <strong>of</strong> Sciences 2008, 105, 1176-1181.<br />

36. Lee, J.-Y.; Li, J.; Yeung, E. S. Anal. Chem. 2007, 79, 8083-8089.<br />

37. Zhou, Z.; Fahrni, C. J. Journal <strong>of</strong> the American Chemical Society 2004, 126,<br />

8862-8863.<br />

38. Iorga, B.; Eymery, F.; Carmichael, D.; Savignac, P. European Journal <strong>of</strong><br />

Organic Chemistry 2000, 2000, 3103-3115.<br />

39. Burt, D. W.; Simpson, P. Journal <strong>of</strong> the Chemical Society C: Organic 1969,<br />

2273-2276.<br />

40. McKenna, C. E.; Schmidhuser, J. Journal <strong>of</strong> the Chemical Society, Chemical<br />

Communications 1979, 739-739.<br />

41. Berlier, J. E.; Rothe, A.; Buller, G.; Bradford, J.; Gray, D. R.; Filanoski, B. J.;<br />

Telford, W. G.; Yue, S.; Liu, J.; Cheung, C.-Y.; Chang, W.; Hirsch, J. D.; Beechem, J.<br />

M.; Haugland, R. P.; Haugland, R. P. The Journal <strong>of</strong> Histochemistry & Cytochemistry<br />

2003, 51, 1699–1712.<br />

42. Herne, T. M.; Tarlov, M. J. Journal <strong>of</strong> the American Chemical Society 1997,<br />

119, 8916-8920.<br />

43. Pocker, Y.; Ronald, B. P. The Journal <strong>of</strong> Physical Chemistry 1984, 88, 4185-<br />

4191.<br />

44. Azam, M. S.; Fenwick, S. L.; Gibbs-Davis, J. M. Langmuir 2011, 27, 741-750.<br />

134


45. Wollman, E. W.; Kang, D.; Frisbie, C. D.; Lorkovic, I. M.; Wrighton, M. S.<br />

Journal <strong>of</strong> the American Chemical Society 1994, 116, 4395-4404.<br />

46. Spori, D. M.; Venkataraman, N. V.; Tosatti, S. G. P.; Durmaz, F.; Spencer, N.<br />

D.; Zurcher, S. Langmuir 2007, 23, 8053-8060.<br />

135

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!