analytical separations using packed and open-tubular

171
Louisiana State University LSU Digital Commons LSU Doctoral Dissertations Graduate School 2004 Analytical separations using packed and open- tubular capillary electrochromatography Constantina P. Kapnissi Louisiana State University and Agricultural and Mechanical College, [email protected] Follow this and additional works at: hps://digitalcommons.lsu.edu/gradschool_dissertations Part of the Chemistry Commons is Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion in LSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please contact[email protected]. Recommended Citation Kapnissi, Constantina P., "Analytical separations using packed and open-tubular capillary electrochromatography" (2004). LSU Doctoral Dissertations. 595. hps://digitalcommons.lsu.edu/gradschool_dissertations/595

Upload: others

Post on 12-May-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Analytical separations using packed and open-tubular

Louisiana State UniversityLSU Digital Commons

LSU Doctoral Dissertations Graduate School

2004

Analytical separations using packed and open-tubular capillary electrochromatographyConstantina P. KapnissiLouisiana State University and Agricultural and Mechanical College, [email protected]

Follow this and additional works at: https://digitalcommons.lsu.edu/gradschool_dissertations

Part of the Chemistry Commons

This Dissertation is brought to you for free and open access by the Graduate School at LSU Digital Commons. It has been accepted for inclusion inLSU Doctoral Dissertations by an authorized graduate school editor of LSU Digital Commons. For more information, please [email protected].

Recommended CitationKapnissi, Constantina P., "Analytical separations using packed and open-tubular capillary electrochromatography" (2004). LSUDoctoral Dissertations. 595.https://digitalcommons.lsu.edu/gradschool_dissertations/595

Page 2: Analytical separations using packed and open-tubular

ANALYTICAL SEPARATIONS USING PACKED AND OPEN-TUBULAR CAPILLARY ELECTROCHROMATOGRAPHY

A Dissertation

Submitted to the Graduate Faculty of the

Louisiana State University and Agricultural and Mechanical College

in partial fulfillment of the requirements for the degree of

Doctor of Philosophy

in

The Department of Chemistry

by Constantina P. Kapnissi

B.S., University of Cyprus, 1999 August 2004

Page 3: Analytical separations using packed and open-tubular

ii

Copyright 2004 Constantina Panayioti Kapnissi All rights reserved

Page 4: Analytical separations using packed and open-tubular

iii

DEDICATION

I would like to dedicate this work to my husband Andreas Christodoulou, my parents

Panayiotis and Eleni Kapnissi, and my sisters Erasmia, Panayiota and Stella Kapnissi. I

want to thank all of you for helping me, in your own way, to finally make one of my

dreams come true. Thank you for encouraging me to continue and achieve my goals.

Thank you for your endless love, support, and motivation. Andreas, thank you for your

continuous patience and for being there for me whenever I needed you during this difficult

time. I will never forget what you did for me.

Page 5: Analytical separations using packed and open-tubular

iv

ACKNOWLEDGMENTS

I would like to acknowledge the following people for their contributions in this work:

Dr. Isiah M. Warner for his guidance and excellent advice. Thank you for encouraging

me to finish what I started. I want to also thank you for the financial and moral support.

You are an excellent mentor, and you prove that every day with your constant

encouragement and support. Thank you again for everything.

Dr. Rezik A. Agbaria for his helpful scientific discussions during my general exam and

my project with laser scanning confocal microscopy.

Dr. Charles William Henry III for training me how to use the capillary electrophoresis

instrument and the packing apparatus. Dr. Henry and Dr. Shahab Shamsi helped me

understand the most important concepts of my research.

Dr. Joseph B. Schlenoff and his group at Florida State University in Tallahassee,

Florida for training me on the polyelectrolyte multilayer coating procedure.

Dr. Lei Geng and Mark Lowry at University of Iowa in Iowa City, Iowa for helping

me conduct some experiments on the laser scanning confocal microscope, and for their

helpful discussions about this part of my research.

Dr. Akbay Cevdet and Bertha C. Valle for providing the polymeric surfactants I used in

my research work.

Warner research group and Warner family for their encouragement and their total

support.

Bertha C. Valle for her friendship, her advice, and the long useful discussions we had

about research.

Page 6: Analytical separations using packed and open-tubular

v

Dr. Nikki Gill, Dr. Xiaofeng Zhu, and Dr. Abdul Numan for editing a few of my

reports.

My family and friends in Cyprus for believing in me, and supporting all my choices.

Page 7: Analytical separations using packed and open-tubular

vi

TABLE OF CONTENTS

DEDICATION .....................................................................................................................iii

ACKNOWLEDGMENTS ...................................................................................................iv

LIST OF TABLES ...............................................................................................................ix

LIST OF FIGURES ..............................................................................................................x

LIST OF ABBREVIATIONS ............................................................................................xiv

ABSTRACT .................................................................................................................... xviii

CHAPTER 1. INTRODUCTION .........................................................................................1 1.1 Capillary Electrophoresis ....................................................................................1 1.2 Micellar Electrokinetic Chromatography ...........................................................9 1.3 Capillary Electrochromatography .....................................................................15

1.3.1 Packed-Capillary Electrochromatography ...............................................17 1.3.2 Open-Tubular Capillary Electrochromatography ....................................18

1.3.2.1 Adsorption ...................................................................................20 1.3.2.2 Covalent Bonding and/or Crosslinking .......................................29 1.3.2.3 Porous Layers ..............................................................................31 1.3.2.4 Chemical Bonding After Etching ................................................32 1.3.2.5 Sol-Gel Technique .......................................................................35

1.4 Chirality ............................................................................................................38 1.5 Chiral Selectors in Open-Tubular Capillary Electrochromatography ..............40

1.5.1 Derivatized Cyclodextrins .......................................................................40 1.5.2 Cellulose ..................................................................................................42 1.5.3 Proteins ....................................................................................................43 1.5.4 Molecular Imprinted Polymers ................................................................44 1.5.5 Polymeric Surfactants ..............................................................................46

1.6 Laser Scanning Confocal Microscopy ..............................................................47 1.7 Scope of Dissertation ........................................................................................51 1.8 References .........................................................................................................54

CHAPTER 2. SEPARATION OF BENZODIAZEPINES BY USE OF CAPILLARY ELECTROCHROMATOGRAPHY .....................................65 2.1 Introduction .......................................................................................................65 2.2 Experimental .....................................................................................................68

2.2.1 Reagents and Chemicals ..........................................................................68 2.2.2 Instrumentation and Conditions ...............................................................68 2.2.3 Sample and Buffer Preparation ................................................................69 2.2.4 Preparation and Conditioning of Packed Capillary Columns ..................69

2.3 Results and Discussion .....................................................................................70

Page 8: Analytical separations using packed and open-tubular

vii

2.3.1 Effect of the Nature of Organic Modifier ................................................70 2.3.2 Effect of Mobile-Phase Composition ......................................................73 2.3.3 Effect of Applied Voltage ........................................................................73 2.3.4 Effect of Tris Concentration ....................................................................75 2.3.5 Effect of Column Temperature ................................................................76 2.3.6 CEC Separation of Drugs from a Urine Sample ......................................78

2.4 Conclusion ........................................................................................................81 2.5 References .........................................................................................................81

CHAPTER 3. ANALYTICAL SEPARATIONS USING MOLECULAR MICELLES IN OPEN-TUBULAR CAPILLARY ELECTROCHROMATOGRAPHY .....................................85 3.1 Introduction .......................................................................................................85 3.2 Experimental .....................................................................................................87

3.2.1 Apparatus and Conditions ........................................................................87 3.2.2 Reagents and Chemicals ..........................................................................87 3.2.3 Sample and Buffer Preparation ................................................................88 3.2.4 Synthesis of Monomeric and Polymeric Surfactant ................................88 3.2.5 Procedure for Polyelectrolyte Multilayer (PEM) Coating .......................88

3.3 Results and Discussion .....................................................................................89 3.3.1 Endurance of PEM Coating .....................................................................89 3.3.2 Stability of PEM Coating .........................................................................91 3.3.3 Reproducibilities ......................................................................................92 3.3.4 Voltage Study ..........................................................................................92 3.3.5 Temperature Study ...................................................................................92 3.3.6 Comparison Between Monomeric and Polymeric Surfactants ................94

3.4 Conclusion ........................................................................................................95 3.5 References .........................................................................................................98

CHAPTER 4. CHIRAL SEPARATIONS USING POLYMERIC SURFACTANTS AND POLYELECTROLYTE MULTILAYERS IN OPEN-TUBULAR CAPILLARY ELECTROCHROMATOGRAPHY .................................. 101 4.1 Introduction .................................................................................................... 101 4.2 Experimental .................................................................................................. 103

4.2.1 Apparatus and Conditions ..................................................................... 103 4.2.2 Reagents and Chemicals ....................................................................... 104 4.2.3 Sample and Background Electrolyte Preparation ................................. 104 4.2.4 Synthesis of Polymeric Surfactant ........................................................ 105 4.2.5 Procedure for Polyelectrolyte Multilayer Coating ................................ 106

4.3 Results and Discussion .................................................................................. 106 4.3.1 Effect of Additives in Polymer Deposition Solutions .......................... 107 4.3.2 Effect of NaCl Concentration ............................................................... 107 4.3.3 Effect of Column Temperature ............................................................. 108 4.3.4 Effect of Bilayer Number ..................................................................... 108

Page 9: Analytical separations using packed and open-tubular

viii

4.3.5 Reproducibilities ................................................................................... 113 4.3.6 Chiral Separation of Analytes ............................................................... 113

4.4 Conclusion ..................................................................................................... 118 4.5 References ...................................................................................................... 118

CHAPTER 5. INVESTIGATION OF THE STABILITY OF POLYELECTROLYTE MULTILAYER COATINGS IN OPEN-TUBULAR CAPILLARY ELECTROCHROMATOGRAPHY USING LASER SCANNING CONFOCAL MICROSCOPY ............................................ 121 5.1 Introduction .................................................................................................... 121 5.2 Experimental .................................................................................................. 123

5.2.1 Apparatus and Conditions ..................................................................... 123 5.2.2 Reagents and Chemicals ....................................................................... 124 5.2.3 Sample and Background Electrolyte Preparation ................................. 124 5.2.4 Synthesis of Polymeric Surfactant ........................................................ 125 5.2.5 Procedure for Polyelectrolyte Multilayer Coating ................................ 125

5.3 Results and Discussion .................................................................................. 126 5.3.1 Open-Tubular Capillary Electrochromatography ................................. 126

5.3.1.1 Coating Stability ....................................................................... 126 5.3.1.2 Capillary Recovery and Coating Regeneration ........................ 130

5.3.2 Laser Scanning Confocal Microscopy .................................................. 133 5.4 Conclusion ..................................................................................................... 137 5.5 References ...................................................................................................... 138

CHAPTER 6. CONCLUSIONS AND FUTURE STUDIES .......................................... 141 6.1 References ...................................................................................................... 147

VITA ................................................................................................................................ 148

Page 10: Analytical separations using packed and open-tubular

ix

LIST OF TABLES

Table Page

1.1 Classification of surfactant molecules used in MEKC ...........................................11

3.1 Reproducibilities of PEM capillary coating. Conditions: same as Figure 3.3 ........94

4.1 Run-to-run reproducibilities of PEM capillary coating. Conditions: same as Figure 4.2, except NaCl concentration was varied ...........................................114 4.2 Capillary-to-capillary reproducibilities of PEM capillary coating. Conditions: same as Figure 4.2, except NaCl concentration was varied...............116 4.3 Reproducibilities of PEM capillary coating. Conditions: same as Figure 4.2, except NaCl concentration, 0.1 M.......................................................116 5.1 Stability of PEM capillary coating. Conditions: same as Figure 5.1 ................... 130

Page 11: Analytical separations using packed and open-tubular

x

LIST OF FIGURES

Figure Page

1.1 Schematic diagram of a CE system ........................................................................ 2

1.2 Origin of EOF – Electrical Double Layer .................................................................5

1.3 (a) EOF flow profile, (b) Pressure driven flow profile and their corresponding solute zone profiles ...........................................................................6

1.4 Solute migration ........................................................................................................7

1.5 Structure of a surfactant molecule ..........................................................................10

1.6 Equilibria between a surfactant molecule, a surface monolayer, and a micelle .....12

1.7 Elution time window for neutral analytes in MEKC ..............................................15

1.8 Origin of EOF in a packed capillary .......................................................................19

1.9 (a) Intrinsic, (b) Extrinsic charge compensation .....................................................26

1.10 Scheme of the PEM-coated capillary .....................................................................27

1.11 Model for the aperture-based near-field optical microscope ..................................49

1.12 Schematic diagram of a confocal microscope ........................................................52

2.1 Structures of the seven benzodiazepine analytes ....................................................71

2.2 Effect of the nature of organic modifier on the CEC separation of benzodiazepines. Conditions: C18 stationary phase; 40 cm packed x 100 µm i.d.; electrolyte, 10 mM Tris (pH 8)-ACN (40:60), 10 mM Tris (pH 8)-MeOH (30:70), 10 mM Tris (pH 8)-THF (55:45); applied voltage, 30 kV; electrokinetic injection, 15 kV for 5 s; temperature, 25 °C; UV detection, 220 nm ...............................................74

2.3 Effect of mobile-phase composition on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the mobile phase composition (ACN-Tris) was varied ...........................................................................................75

Page 12: Analytical separations using packed and open-tubular

xi

2.4 Effect of applied voltage on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the applied voltage was varied; electrolyte, 10 mM Tris (pH 8)-ACN (60:40) ........................................................77

2.5 Effect of Tris concentration on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the Tris buffer concentration was varied; electrolyte, Tris (pH 8)-ACN (60:40); applied voltage, 20 kV ..............................78

2.6 Effect of column temperature on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the temperature was varied; electrolyte, 30 mM Tris (pH 8)-ACN (60:40); applied voltage, 20 kV .....................................79

2.7 CEC separation of drugs from a urine sample. (a) Without spiking. Conditions: C18 stationary phase; 40 cm packed x 100 µm i.d.; electrolyte, 30 mM Tris (pH8)-ACN (60:40); applied voltage, 20 kV; electrokinetic injection, 15 kV for 5 s; temperature, 15 °C; UV detection, 220 nm. (b) With spiking. Separation conditions are the same as above, except electrokinetic injection (urine sample), 20 kV for 10 s; electrokinetic injection (2 mg/ml standard oxazepam), 15 kV for 5 s .........................................................80

3.1 Structures of the (a) monomeric SUG and (b) polymeric SUG ..............................90

3.2 Scheme of the PEM-coated capillary .....................................................................91

3.3 Stability studies of PEM coating. Conditions: 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SUG) with 0.2 M NaCl; pressure injection, 0.1 psi for 1 s; electrolyte, 50 mM Na2HPO4 (pH 9.2); applied voltage, 20 kV; temperature, 25 °C; capillary, 57 cm (50 cm effective length) x 50 µm i.d.; detection, 214 nm ...........................93

3.4 Effect of applied voltage on the OT-CEC separation of benzodiazepines. Conditions: same as Figure 3.3, except applied voltage was varied ......................................................................................96

3.5 Effect of temperature on the OT-CEC separation of benzodiazepines. Conditions: same as Figure 3.3, except the temperature was varied ..........................................................................97

Page 13: Analytical separations using packed and open-tubular

xii

3.6 Comparison between monomeric and polymeric surfactants for OT-CEC separation of benzodiazepines. Conditions: same as Figure 3.3, except applied voltage, 30 kV. (a) 0.5% (w/v) poly (L-SUG) (b) 0.5% (w/v) mono (L-SUG) ...............................98

4.1 Structure of poly (L-SULV) ................................................................................. 105

4.2 Chiral separation of BNP. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV); pressure injection, 0.5 psi for 3 s; background electrolyte, 100 mM Tris and 10 mM Na2B4O7 (pH 10.0); applied voltage, 30 kV; temperature, 25 °C; capillary, 57 cm (50 cm effective length) x 50 µm i.d.; detection, 214 nm. (a) 0.01 M NaCl; (b) 0.01 M 1E-3MI-HFP; (c) 0.01 M 1B-3MI-TFB ................ 109

4.3 Effect of NaCl concentration on the chiral separation of BNP. Conditions: same as Figure 4.2, except NaCl concentration was varied ............. 110

4.4 Effect of column temperature on the chiral separation of BNP. Conditions: same as Figure 4.2, except temperature was varied ......................... 111

4.5 Effect of bilayer number on the chiral separation of BNP. Conditions: same as Figure 4.2, except bilayer number was varied .................... 112

4.6 Illustration of a run-to-run reproducibility for the chiral separation of BNP. Conditions: same as Figure 4.2, except NaCl concentration, 0.05 M ................................................................................. 115

4.7 (a) Chiral separation of BOH. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV) with 0.1 M NaCl; pressure injection, 0.5 psi for 3 s; background electrolyte, 100 mM Tris and 10 mM Na2B4O7 (pH 10.0); detection, 214 nm. Chiral separation of temazepam. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV) with 0.1 M NaCl; pressure injection, 0.9 psi for 7 s; background electrolyte, 25 mM Tris and 25 mM Na2B4O7 (pH 8.5); detection, 220 nm. (c) Chiral separation of secobarbital, and (d) Chiral separation of pentobarbital. Conditions for (c) and (d): 3 bilayers; 0.5% (w/v) PDADMAC and 2.5% (w/v) poly (L-SULV) with 0.2 M NaCl; pressure injection, 0.5 psi for 3 s; background electrolyte, 300 mM H3BO3 and 30 mM Na2HPO4 (pH 7.2); detection, 214 nm. Other conditions are the same as Figure 4.2 ......................... 117

Page 14: Analytical separations using packed and open-tubular

xiii

5.1 Dependence of electroosmotic mobility on the NaOH flushing time. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV) with 0.1 M NaCl; pressure injection, 0.5 psi for 3 s; background electrolyte, 100 mM Tris and 10 mM Na2B4O7 (pH 10.0); applied voltage, 30 kV; temperature, 25 °C; capillary, 57 cm (50 cm effective length) x 50 µm i.d.; detection, 214 nm ................................................................................................. 128

5.2 Dependence of selectivity on the NaOH flushing time. Conditions: same as Figure 5.1 ............................................................................ 129

5.3 Coating Regeneration - Dependence of electroosmotic mobility on the NaOH flushing time. Conditions: same as Figure 5.1 .............................. 131

5.4 Dependence of selectivity on the NaOH flushing time. Conditions: same as Figure 5.1 ............................................................................ 132

5.5 The z-Section images of a 20-bilayer capillary (left) and a fused-silica capillary (right). Optical slice thickness: 3.3 µm. z-Step size 1.65 µm. Every fourth image is shown. All images are 204 x 512 pixels (460.6 µm x 115.1 µm) .......................................... 134

5.6 z-Section images of a portion of the same 20-bilayer capillary, as in Figure 5.5. Exposure to 1.0 M NaOH. Optical slice thickness: µm. z-Step size 1.65 µm. Every third image is shown. All images are 512 x 512 pixels (65.8 µm x 65.8 µm) .............................................. 135

5.7 LSCM images of different regions within the same 20-bilayer capillary, as in Figure 5.6. Optical slice thickness: 3.3 µm. Both images are 512 x 820 pixels (115.1 µm x 184.4 µm) ................................. 136

5.8 LSCM image of the same 20-bilayer capillary, as in Figure 5.7. Longer exposure to 1.0 M NaOH. Optical slice thickness: 3.3 µm. The image is 512 x 2048 pixels (115.1 µm x 460.6 µm) ........................................................................................ 137

6.1 Structures of the six β-blocker analytes ............................................................... 142

Page 15: Analytical separations using packed and open-tubular

xiv

LIST OF ABBREVIATIONS

Abbreviation Name

AFM atomic force microscopy

BGE background electrolyte

BNP 1,1’-binaphthyl-2,2’-dihydrogenphosphate

BOH 1,1’-bi-2-naphthol

CDCPC cellulose tris(3,5-dichlorophenylcarbamate)

CE capillary electrophoresis

CEC capillary electrochromatography

CGE capillary gel electrophoresis

CIEF capillary isoelectric focusing

CITP capillary isotachophoresis

CMC critical micellar concentration

CTAB cetyltrimethylammonium bromide

CTAC cetyltrimethylammonium chloride

C18-TEOS n-octadecyltriethoxysilane

CZE capillary zone electrophoresis

DAPS N-dodecyl(N,N-dimethyl-3-ammonio-1-propane sulfonate)

Dioxo[13]aneN4 1,4,7,10-tetraazacyclotridecane-11,13,dione

DMPCC 3,5-dimethylphenylcarbamoyl cellulose

DS dextran sulfate

DTAB dodecyltrimethylammonium bromide

ELP extended light path

Page 16: Analytical separations using packed and open-tubular

xv

EOF electroosmotic flow

FESI field-enhanced sample injection

FRET fluorescence resonance energy transfer

GC gas chromatography

H3BO3 boric acid

HCl hydrochloric acid

HPLC high performance liquid chromatography

H2TPFPP 5,10,15,20-tetrakis(penta fluorophenyl)porphyrin

LC liquid chromatography

LODs limits of detection

LSCM laser scanning confocal microscopy

MEKC micellar electrokinetic chromatography

MIPs molecular imprinted polymers

Mono (L-SUG) mono (sodium N-undecenoyl-L-glycinate)

MS mass spectrometry

NA numerical aperture

Na2B4O7 sodium borate

NaCl sodium chloride

Na2HPO4 sodium phosphate

NaOH sodium hydroxide

NIR near-infrared

NSOM near-field scanning optical microscopy

ODS octadecylsilane

Page 17: Analytical separations using packed and open-tubular

xvi

OT-CEC open-tubular capillary electrochromatography

OT-CLC open-tubular capillary liquid chromatography

Packed-CEC packed capillary electrochromatography

PAHs polycyclic aromatic hydrocarbons

PAPS lithium 3’-phosphoadenosine 5’-phosphosulfate

PB polybrene

PDADMAC poly (diallyldimethylammonium chloride)

PEI polyethyleneimine

PEM polyelectrolyte multilayer

pI isoelectric point

PMBC para-methylbenzoyl cellulose

Poly (L-SUG) poly (sodium N-undecanoyl-L-glycinate)

Poly (L-SULV) poly (sodium N-undecanoyl-L-leucylvalinate)

Poly (SUS) poly (sodium undecylenic sulfate)

PSS poly (styrene sulfonate)

PVA poly (vinylamine)

R6G rhodamine 6G

RSD relative standard deviation

SDS sodium dodecyl sulfate

SFC supercritical fluid chromatography

SMIL successive multiple ionic-polymer layer

TEOS n-octadecyltriethoxysilane

THF tetrahydrofuran

Page 18: Analytical separations using packed and open-tubular

xvii

Tris tris(hydroxymethyl)aminomethane

1B-3MI-TFB 1-butyl-3-methylimidazolium tetrafluoroborate

1E-3MI-HFP 1-ethyl-3-methyl-1H-imidazolium hexafluorophosphate

Page 19: Analytical separations using packed and open-tubular

xviii

ABSTRACT

The goal of the research reported in this dissertation is to develop packed and open-tubular

capillary electrochromatographic methods for improved achiral and chiral separations of

various classes of analytes. The first part of this research involves the separation of seven

benzodiazepines by the use of the packed mode of capillary electrochromatography (CEC)

and a 40 cm packed bed of Reliasil 3 µm C18 stationary phase. Optimal conditions were

established by varying the mobile phase, the amount of organic modifier, the buffer

concentration, the applied voltage, and the column temperature. The second part of this

research focuses on the open-tubular mode of CEC and the polyelectrolyte multilayer

(PEM) coating approach. In the first study of this part, poly (diallyldimethylammonium

chloride), PDADMAC, was used as the cationic polymer and poly (sodium N-undecanoyl-

L-glycinate), poly (L-SUG), was used as the anionic polymer for the construction of the

PEM coating. The performance of the modified capillaries as a separation medium was

evaluated by use of seven benzodiazepines as analytes. In the second study, the anionic

polymeric surfactant poly (sodium N-undecanoyl-L-leucylvalinate), poly (L-SULV), was

used as the chiral discriminator for the separations of several drug analytes.

Reproducibility of the PEM coating was evaluated by computing the relative standard

deviation (RSD) values of the electroosmotic flow (EOF). The PEM-coated capillaries

were remarkably robust with excellent reproducibilities and high stabilities against extreme

pH values. The stability of the capillary surface was further investigated after exposure to

NaOH solutions. The structural changes of these coatings were monitored using laser

scanning confocal microscopy (LSCM). These changes were discussed in terms of

separations using open-tubular CEC (OT-CEC). In addition, the electropherograms

Page 20: Analytical separations using packed and open-tubular

xix

obtained from the chiral separation of 1,1’-binaphthyl-2,2’-dihydrogenphosphate (BNP) in

OT-CEC allowed the measurements of both selectivity and electroosmotic mobility

changes after long exposure to NaOH.

Page 21: Analytical separations using packed and open-tubular

1

CHAPTER 1.

INTRODUCTION

1.1 Capillary Electrophoresis

Capillary electrophoresis (CE) is a family of related techniques that use narrow-bore

fused-silica capillaries to perform high efficiency separations of both small and large

molecules. Although CE is originally considered for the analysis of biological

macromolecules, it has also been utilized for the separation of other compounds such as

chiral drugs, vitamins, pesticides, dyes, inorganic ions, organic acids, and surfactants [1,

2]. CE offers a number of advantages when compared with chromatographic techniques:

(1) extremely small amount of sample; (2) high separation efficiency and resolution; (3)

rapid and quantitative separation; (4) automated instrumentation; (5) various modes to vary

selectivity; and (6) simple separation mechanism [2, 3].

As mentioned above, one of the main advantages of CE is the overall simplicity of the

instrumentation. Figure 1.1 illustrates a schematic diagram of a typical CE system. The

basic components of this system are a fused-silica capillary whose ends are placed in

buffer reservoirs, a high-voltage power supply, a UV lamp, a photodiode-array detector, a

sample reservoir, and two electrodes that are used to make electrical contact between the

high-voltage power supply and the capillary. In CE, the separation of analytes is achieved

by replacing the inlet buffer reservoir with a sample reservoir, and by applying either an

electric field or an external pressure. After the sample is loaded onto the capillary, the

sample is replaced by the inlet buffer reservoir, the electric field is applied, the ions in the

sample move along the capillary, and the separation is performed.

Page 22: Analytical separations using packed and open-tubular

2

Sample ReservoirInlet Buffer Reservoir

(-) Electrode

Anode

Outlet Buffer Reservoir

High-Voltage Power Supply

UV-Lamp

Diode-Array Detector

Capillary

(+) Electrode

Cathode

Figure 1.1 Schematic diagram of a CE system.

The utility of CE is also derived from its six different modes of operation. These modes

of CE include capillary zone electrophoresis (CZE), capillary gel electrophoresis (CGE),

capillary isoelectric focusing (CIEF), capillary isotachophoresis (CITP), micellar

electrokinetic chromatography (MEKC), and capillary electrochromatography (CEC). CZE

is the simplest form and the most widely used technique in CE [1, 2]. The separation

principle of CZE is the difference in charge-to-size ratio and the difference in

electrophoretic mobilities of solutes that result in different velocities. Electrophoretic

Page 23: Analytical separations using packed and open-tubular

3

mobility is the mobility of an ion when an electric field is applied across the capillary. The

velocity of an ion is expressed by:

Eeµν = (1.1)

where ν is the ion velocity, eµ is the electrophoretic mobility, and E is the applied

electric field. The electric field is a function of the applied voltage and the capillary length

with units of V/cm. When a constant electric field is applied, ionic species undergo an

electrostatic force, eF :

qEFe = (1.2)

where q is the charge of a particular ion. This force causes ions to accelerate towards the

oppositely charged electrode. In a viscous medium the electrostatic force is balanced by its

frictional force, fF , that slows the mobility of ions. According to Stokes’ law for spherical

particles, the frictional force can be given by Equation 1.3:

νπηrFf 6= (1.3)

where η is the dynamic viscosity of the solution, and r is the radius of the particle or the

ion. During electrophoresis a steady state is achieved, and at this point the forces are equal.

The combination of the above equations yields an equation that describes the

electrophoretic mobility as follows:

rq

e πηµ

6= (1.4)

The electrophoretic mobility is a property of a given ion and medium, and it is a

physical constant for that ion. Equation 1.4 demonstrates that small species with high

charges have high mobilities, whereas large, minimally charged species have low

Page 24: Analytical separations using packed and open-tubular

4

mobilities. In addition, neutral species do not possess an electrophoretic mobility ( 0=q ).

Therefore, they are not separated by use of CZE. Another important characteristic that is

derived from the above equation is that the electrophoretic mobility of ionic species

decreases as the viscosity of the background electrolyte (BGE) increases [1, 5, 6].

A significant factor that causes the movement of both neutral and charged species in

solution is the electroosmotic flow (EOF). The origin of this flow is the electrical double

layer that is formed at the solid-liquid interface, between the bulk liquid and the inner

capillary surface (Figure 1.2). Under alkaline conditions, the inside wall of a fused-silica

capillary is negatively charged due to ionization of the surface silanol groups to negatively

charged silanoate groups. Electrolyte cations accumulate adjacent to the negative surface

of the capillary, they absorb on it by electrostatic attraction, and they form an immobilized

layer that is called the Stern layer. The remaining ions constitute the diffuse layer, which

extends into the bulk liquid. This arrangement of ions develops the electric double layer.

When an electric field is applied across the column, the cations, which predominate in the

diffuse layer, migrate in the direction of the cathode. Since ions are solvated by water, the

bulk solution in the capillary is dragged along by the migrating charge.

This generates the EOF, which can be described as follows:

EEOF

=ηεζν (1.5)

where EOFν is the electroosmotic velocity, ζ is the zeta potential, and ε is the dielectric

constant of the BGE. The zeta potential is the corresponding potential across the layers,

and it depends on the thickness of the diffuse layer and the surface charge on the capillary

wall. The EOF can be also expressed in terms of mobility by the equation:

Page 25: Analytical separations using packed and open-tubular

5

ηεζµ =EOF (1.6)

where EOFµ is the electroosmotic mobility.

Surface of capillaryStern layer

Diffuse layer

Bulk solutionEOF

Surface of capillarySurface of capillaryStern layerStern layer

Diffuse layerDiffuse layer

Bulk solutionBulk solutionEOFEOF

Figure 1.2 Origin of EOF – Electrical Double Layer.

A unique characteristic of the EOF is that it is uniformly distributed along the capillary.

This results in a flat plug-like profile (Figure 1.3a) rather than a parabolic profile that is

generated by a pressure-driven system (Figure 1.3b), such as in high performance liquid

chromatography (HPLC). A flat profile of EOF reduces band broadening, enhances fast

elution rates, and yields highly efficient peaks.

Another distinctive feature of the EOF under normal conditions, where the

capillary surface is negatively charged, is that it causes movement of nearly all species in

the same direction, regardless of charge. In CZE, the migration of charged species is

Page 26: Analytical separations using packed and open-tubular

6

determined by the EOF and by the electrophoretic mobility of the solutes. This can be

expressed by the following equation:

EOFepp µµµα += (1.7)

where ppαµ is the apparent mobility. The value of electrophoretic mobility of an ion

depends on the charge of that ion, and it can be either positive or negative. Therefore,

cations migrate fastest since both EOF and electrophoretic mobility are in the same

direction. Neutral molecules move along with the EOF, because they do not have

electrophoretic mobility, and they are not separated. Finally, anions migrate slowest since

their electrophoretic mobility is in the opposite direction of EOF. However, they still move

toward the detector, because the magnitude of EOF is much greater than their

electrophoretic mobility. This is clearly illustrated in Figure 1.4.

EOF

Flat profile

Pressure Driven Flow

Parabolic profile

Figure 1.3 (a) EOF flow profile, (b) Pressure driven flow profile and their corresponding solute zone profiles.

Page 27: Analytical separations using packed and open-tubular

7

Capillary electrophoretic separations are also performed using another mode of CE,

called CGE. This form of CE is performed in a porous gel polymer matrix, and the

separation of analytes is based on their charge and size. CGE has been mostly employed

for the analysis of large biomolecules, such as proteins and nucleic acids. The most

common sieving gels used in CGE are cross-linked polyacrylamide and agarose. The

former gel has smaller pore size, and it is used for protein separations, while the latter, with

a larger pore size, is mostly applied to DNA analysis [2, 4, 5].

EOF

N

N

N

NN

N

EOF

N

N

N

NN

N

EOFEOF

NN

NN

NN

NNNN

NN

Figure 1.4 Solute migration.

CIEF is an electrophoretic technique used to separate amino acids, peptides, and

proteins based on their pI (isoelectric point) values. In CIEF, a pH gradient is formed

within the capillary using ampholytes. Ampholytes, or else zwitterions, are molecules that

contain both an acidic and a basic moiety. They have pI values between pH 3 and 9. The

Page 28: Analytical separations using packed and open-tubular

8

gradient is formed when the capillary is filled with sample and ampholytes. A basic

solution is placed at the cathode, and an acidic solution is placed at the anode. When the

electric field is applied, the charged ampholytes and the sample components migrate

through the medium until they reach their pI value, at which they become uncharged. This

procedure is known as “focusing”, and it is indicated by a decrease in current. After this,

the contents of the capillary are mobilized to the detector by pressure or by adding an

electrolyte into one of the reservoirs [1, 4, 5].

CITP is another electrophoretic technique, in which a combination of two

electrolytes causes all analyte bands to migrate at the same velocity. This technique can

separate either cations or anions, but not both at the same time. During a separation

procedure, an analyte mixture is injected between a leading electrolyte containing ions of

higher mobility (Cl-) than any of the analyte ions and a terminating electrolyte with ions

that have lower mobility (heptanoate) than the sample ions. When an electric field is

applied, the analyte ions migrate in bands according to their unique mobilities. After all

analyte ions are separated into different bands, they move at the same velocity [1, 4].

MEKC and CEC are hybrid techniques that combine the best features of both

electrophoresis and chromatography. Both electrophoretic techniques can be used for the

separation of neutral as well as charged compounds. MEKC involves the introduction of a

surfactant at a concentration above the critical micellar concentration (CMC), at which

micelles are formed. The separation is based on the hydrophobic and ionic interactions of

the analytes with the micelles that act as a moving stationary phase, or else a

pseudostationary phase. CEC uses a stationary phase rather than a micellar

pseudostationary phase. As in MEKC, the mechanism of separation depends on the

Page 29: Analytical separations using packed and open-tubular

9

partitioning between the two phases. However, in the case of CEC, the partitioning occurs

between the packed or coated stationary phase and the mobile phase. In addition, when the

analytes are charged, the separation also depends on their electrophoretic mobilities [3-7].

Both approaches are discussed in detail in the following sections.

1.2 Micellar Electrokinetic Chromatography

MEKC, which was first introduced by Terabe et al. in 1984 [8], is the second most

commonly used CE technique. Although MEKC is a form of CE, its separation principle is

more similar to HPLC than to CE. As stated earlier, the separation mechanism is based on

the electrophoretic mobility and the partitioning of the analytes between the mobile phase

and the micellar pseudostationary phase. However, for neutral species, it is only the

partitioning that effects the separation. The driving force for the partitioning of analytes is

hydrophobicity. In addition, hydrogen bonding, dipole-dipole, and dispersive interactions

can contribute to the solute partitioning between the two phases. The most commonly used

pseudostationary phase in MEKC is the micelle [1, 5, 9-11]. A more detailed discussion

about surfactants and micelles is given in the succeeding paragraphs.

Surfactants, which are also known as amphiphiles, are organic compounds that

consist of a polar or ionic head group and a hydrocarbon tail group (Figure 1.5). The head

group has hydrophilic properties, and it is compatible with the aqueous environment, while

the tail group is hydrophobic and not friendly with the aqueous environment. Surfactants

are classified into different groups according to the nature of their head group (anionic,

cationic, nonionic, and zwitterionic). Table 1.1 reports these four groups along with the

structures of different head groups and a few examples.

Page 30: Analytical separations using packed and open-tubular

10

The surfactant molecules, at a low concentration, are dispersed in solution.

However, at a concentration above the CMC, and above a certain temperature, known as

the Krafft temperature, the surfactant molecules begin to build up their own structures.

They aggregate together to form micelles in the interior of the aqueous solution and

monolayers at the surface of the solution (Figure 1.6).

Hydrophobic Tail Group

Hydrophilic Head Group

Hydrophobic Tail Group

Hydrophilic Head Group

Figure 1.5 Structure of a surfactant molecule.

The CMC is a very important characteristic of each surfactant. Its value needs to be

determined in order to understand some of the characteristics and the properties of a

surfactant or a micelle. The CMC value may be attained graphically by plotting a

physicochemical property of the solution, i.e. surface tension, conductivity, turbidity,

osmotic pressure, fluorescence, and light scattering, versus the concentration of the

surfactant. The value of the CMC is the discontinuity in the plot, or the sudden variation in

the above relation. In an aqueous medium, above the CMC, surfactant molecules aggregate

together to form a micelle. As the surfactant concentration increases, the shape of the ionic

Page 31: Analytical separations using packed and open-tubular

11

micelle changes from spherical to cylindrical, to hexagonal, and finally to lamellar. For

nonionic micelles, as the concentration increases the shape changes from spherical directly

to lamellar [6]. The number of aggregated molecules in a micelle is called the aggregation

number, n , and it is usually around 50 to 100. It can be determined by several analytical

techniques, such as light scattering, fluorescence, and NMR sedimentation.

Table 1.1 Classification of surfactant molecules used in MEKC.

ExamplesStructureHydrophilic group

SDSR-O-SO-3Anionic

CTAB, CTAC, DTABR-N+(CH3)3Cationic

R: alkyl groupFor abbreviations see pages xiv, xvi

PAPS, DAPSR-N+(CH3)2-(CH2)3-SO-3Zwitterionic

Triton X-100-RSR-Cyclohexyl-O-PEG

Octyl glucosideR-O-GlucoseNonionic

ExamplesStructureHydrophilic group

SDSR-O-SO-3Anionic

CTAB, CTAC, DTABR-N+(CH3)3Cationic

R: alkyl groupFor abbreviations see pages xiv, xvi

PAPS, DAPSR-N+(CH3)2-(CH2)3-SO-3Zwitterionic

Triton X-100-RSR-Cyclohexyl-O-PEG

Octyl glucosideR-O-GlucoseNonionic

The MEKC separation process for neutral analytes is only based on their

partitioning in and out of the micelle. When anionic micelles are used, the more the analyte

interacts with the micelle the longer its migration time. This is because anionic micelles

have electrophoretic mobilities that are in the opposite direction of the EOF. The more

Page 32: Analytical separations using packed and open-tubular

12

hydrophobic analytes interact more strongly with the micelle, and they are retained longer

since the micelle partially offsets the effects of the EOF.

SurfactantMonolayer

Micelle

SurfactantMonolayer

Micelle

Figure 1.6 Equilibria between a surfactant molecule, a surface monolayer, and a micelle.

The separation mechanism in MEKC is chromatographic, and it can be described

by the use of modified chromatographic relationships. How effective a chromatographic

column is in separating two analytes partially depends on the relative rates at which the

analytes are eluted. These rates are determined by the magnitude of the equilibrium

constants (partition coefficients), K , which are defined as:

M

S

ccK = (1.8)

where Sc is the molar concentration of the analyte in the micellar pseudostationary phase

and Mc is its molar concentration in the mobile phase. The retention factor, or capacity

factor, 'k , describes the migration rates of analytes in the column, or the ratio of moles of

analyte in the micelle to those in the mobile phase:

Page 33: Analytical separations using packed and open-tubular

13

( )

=

−=

M

S

m

r

r

VVK

ttt

ttk1

'

0

0 (1.9)

where rt is the retention time of the analyte, 0t is the retention time of the unretained

analyte that moves with EOF, mt is the retention time of the micelle, SV is the volume of

the micellar phase, and MV is the volume of the mobile phase. The above equation is partly

different from the normal chromatographic 'k because of the movement of the micellar

pseudostationary phase. If the micelle becomes the stationary phase the term mt becomes

infinite, and the Equation 1.9 is converted into the conventional chromatographic equation.

The selectivity factor α of a column for two analytes A and B is determined by the

equation:

( )( ) 0

0

tttt

Ar

Br

−−

=α (1.10)

where ( )Brt and ( )Art are the retention times of the more strongly and less strongly

retained analytes B and A, respectively. This definition dictates that the selectivity is

always greater than or equal to unity.

The resolution, SR , of a column describes its ability to separate two analytes. This

column resolution is defined as:

+

=

'1

0

0

'2

'2

2/1

1

1

11

4k

tttt

kkNR

m

mS α

α (1.11)

Page 34: Analytical separations using packed and open-tubular

14

where N is the number of theoretical plates. This number of theoretical plates is widely

used as a quantitative measure of column efficiency, and is defined by use of the following

equation:

2

2/1

54.5

=

WtN r (1.12)

where 2/1W is the width of the peak at half its maximum height.

In Equation 1.11, the first derivative corresponds to efficiency, the second to

selectivity, and the third and fourth to retention. This suggests that resolution can be

improved by optimizing efficiency, selectivity, and/or the capacity factor. Since the

capacity factor is proportional to the concentration of the surfactant, it can be easily

modified by varying this concentration. In addition, resolution can be improved by

increasing the elution range, or time window, which is the time between 0t and mt (Figure

1.7). Neutral analytes elute between this elution range. The analytes that elute with the

EOF are hydrophilic and do not interact with the micelles. The analytes that are completely

retained by the micelles elute with the micelles. Maximum resolution is obtained when mt

t0

is very small (large time window). Selectivity can also be easily modified by varying the

size, charge, and geometry of the micelle, and particularly the surfactant [1, 3].

In chromatography and electrophoresis, resolution is given by:

( ) ( )[ ]BA

ArBrS WW

ttR

+−

=2

(1.13)

where AW and BW are the widths of the base of peaks A and B, respectively. A resolution

of 1.5 gives a baseline separation of the two analytes. The resolution can be improved by

Page 35: Analytical separations using packed and open-tubular

15

increasing the length of the column, thus, increasing the number of theoretical plates

(column efficiency). However, this results in an increase in separation time [4].

0 t0 tr1 tr2 tr3 tm

Time

Time window

Analytes

EOF

Micelle

0 t0 tr1 tr2 tr3 tm

Time

Time window

Analytes

EOF

Micelle

Figure 1.7 Elution time window for neutral analytes in MEKC.

1.3 Capillary Electrochromatography

In recent years, CEC has become an important member of the arsenal of tools

available in separation science. Although the CEC approach was first introduced by

Pretorius et al. [12] in 1974, it received renewed interest during the 1990s [13-19]. This

revival of CEC is because it is a microcolumn electroseparation technique that combines

the best features of HPLC with those of CE. In essence, CEC mainly couples the high

Page 36: Analytical separations using packed and open-tubular

16

selectivity of HPLC and the high separation efficiency of CE [20-26]. This is because in

CEC, the mobile phase solvent is transported through a capillary by use of the EOF. In

addition, CEC provides high resolution, short analysis time, ruggedness, and low sample

and mobile phase consumption. It also offers wider selectivity and facilitates the separation

of both neutral and charged compounds.

As discussed earlier, the latter characteristic is because the separation of analytes is

based on their interactions with the stationary phase and, when charged, their

electrophoretic mobility [15, 25-32]. Therefore, the apparent mobility for a charged analyte

in CEC is influenced by its electrophoretic mobility, its partitioning interaction and the

EOF. However, for a neutral analyte, the apparent mobility depends only on the

partitioning interaction with the stationary phase, while its elution is driven by the EOF.

Several manuscripts have described the capacity factor for separation of an analyte

in CEC [33-36]. Briefly, the capacity factor in CEC ( CECk ′ ) can be expressed as:

+

=

EOF

e

EOF

e

CEC

kk

µµµµ

1

'

' (1.14)

This equation incorporates both the electrophoretic and chromatographic separation

mechanisms in CEC. It should be noted that the mechanism of separation in CEC is highly

dependent upon the analyte. For neutral analytes, eµ is zero, and CECk ′ is equal to k ′ .

Thus, neutral analytes are solely separated on the basis of a chromatographic mechanism.

Finally, charged analytes are separated by a combination of chromatographic and

electrophoretic mechanisms, provided these analytes interact with the stationary phase.

Page 37: Analytical separations using packed and open-tubular

17

In CEC, as in every chromatographic system, the capillary column is the most

important component. It serves as a vessel to transport the mobile phase and as a

separation channel [15, 28]. Therefore, preparation of the column, and particularly the

stationary phase packed or immobilized on the inner walls of the capillary, is critical for

CEC. The capillary columns in CEC are usually classified into three main formats [21, 22,

28, 37-44]: (i) packed-CEC [23-27, 45-52], (ii) open-tubular CEC (OT-CEC) [39-43, 53-

64], and (iii) monolithic (or continuous rod) CEC [20, 38, 65-85]. In this dissertation, the

first two main formats are discussed in detail.

1.3.1 Packed-Capillary Electrochromatography

In packed-CEC, a fused-silica capillary is filled with a typical HPLC packing

material (i.e. octadecyl silica). The mobile phase in packed-CEC is driven by the EOF that

is induced by applying an electric field across the capillary column. The origin of this flow

is the electrical double layer that is formed at the solid-liquid interface of a charged surface

in contact with the BGE (Figure 1.8).

In a capillary packed with silica particles the surfaces of the capillary wall and the

particles are negatively charged due to the dissociation of silanol groups. The velocity of

EOF in packed-CEC is shown in the following equation:

η

εεσµ

EcF

RTro

EOF

2/1

32

= (1.15)

where σ , which is proportional to zeta potential, is the charge density at the surface, oε is

the permittivity of vacuum (8.85 x 10-12 C2N-1µ-2), rε is the dielectric constant of the

mobile phase, R is the gas constant, T is temperature, c is the concentration of the

Page 38: Analytical separations using packed and open-tubular

18

electrolyte, F is Faraday’s constant, η is the viscosity of the mobile phase, and E is the

electric field strength. The above equation illustrates that the velocity of EOF depends on

the charge density on the surfaces of the capillary walls and silica particles, the dielectric

constant, the viscosity, and the concentration of the BGE, and the temperature.

However, there are some problems that need to be solved in order for this

conventional form of CEC to be a viable alternative to both CE and HPLC. One of the

problems of packed-CEC is the requirement to fabricate frits, which are needed to retain

the packed particles inside the capillary column. Another problem in packed capillaries is

the tendency to form bubbles around the packing material or at the frit. This problem often

results in an unstable baseline, variable migration times, and current breakdown.

Pressurization of both ends of the column is required, and the mobile phase must be

thoroughly degassed in order to reduce the possibility of bubble formation. Another major

drawback of conventional CEC is that the packing procedure is generally more difficult

than for HPLC due to the narrow inner bore of the capillary and the small diameter of the

particles (5 µm or less). Finally, basic compounds are difficult to separate in packed-CEC

due to the presence of silanol groups that are needed to generate an adequate EOF [21, 39-

41, 86-93]. In order to circumvent the problems mentioned above, continuous bed-type

columns, i.e. open-tubular and monolithic columns, have been suggested as alternatives to

packed-CEC.

1.3.2 Open-Tubular Capillary Electrochromatography

In the OT-CEC format, the stationary phase is deposited on the inner walls of the

capillary. Preparation of the stationary phase, or coating, is a crucial step in any

chromatographic system. In OT-CEC, this coating needs to be stable in order to provide

Page 39: Analytical separations using packed and open-tubular

19

efficient chromatographic separations and a reproducible EOF [92]. In OT-CEC, there are

six general approaches used for modifying the capillary column. These include (i)

adsorption, (ii) covalent bonding and/or crosslinking, (iii) porous silica layers, (iv)

chemical bonding after etching, (v) sol-gel, and (vi) molecular imprinting [39-42, 91-94].

Although OT-CEC is considered an alternative to packed-CEC, its phase ratio and sample

capacity are relatively low due to the typical small surface area of the coating. Several

options have been proposed to increase the surface area, and therefore, to increase the

interactions between the analyte and the coated phase. These options include polymer

coatings, porous silica layers, etching, and sol-gel techniques [22, 41, 89, 94].

Stationary phase particles

EOF

Electrical double layersCapillary wall

Stationary phase particles

EOF

Electrical double layersCapillary wall

Figure 1.8 Origin of EOF in a packed capillary.

Page 40: Analytical separations using packed and open-tubular

20

1.3.2.1 Adsorption

Adsorption of chemicals to solid surfaces is a common phenomenon observed in

separation science. The adsorption of proteins, peptides, and basic compounds on the silica

capillary wall usually causes serious problems, such as loss in efficiency, peak tailing,

unstable baseline, nonreproducible migration times, and reduction of capillary lifetime. In

order to reduce or eliminate the analyte-wall interactions, different methods have been

developed. The most common methods shield the negatively charged silanol groups with

another layer, the adsorbed stationary phase [28, 39, 61, 95]. According to the strength of

adsorption, the adsorbed stationary phases can be divided into two groups: dynamically

adsorbed and physically adsorbed stationary phases [28, 30, 92]. If the adsorption of the

modifier on the capillary wall is weak, the stationary phase is called dynamically adsorbed

stationary phase. In this case, the modifier is added to the mobile phase. In contrast, if the

modifier is strongly adsorbed, the stationary phase is termed a physically adsorbed

stationary phase. In this case, the addition of the modifier to the mobile phase is not

necessary. Several modifiers have been used for the preparation of adsorbed stationary

phases [17, 18, 39, 59, 61, 62, 92, 93, 95-121]. These modifiers can be grouped into three

main categories: (i) cationic surfactants, (ii) polymeric surfactants, and (iii) charged

polymers.

(i) Cationic Surfactants

In 1990, Pfeffer and Yeung [17] developed a method for the separation of several

polycyclic aromatic hydrocarbons (PAHs) by use of open-tubular capillary liquid

chromatography (OT-CLC). The capillaries were first coated with a polymer solution of

0.9% PS-264 (polyvinylsiloxane). The velocity of the EOF on these columns was low due

Page 41: Analytical separations using packed and open-tubular

21

to the polymer coating that blocked the silanol groups on the capillary wall. The cationic

surfactant cetyltrimethylammonium bromide (CTAB) was then added to the mobile phase

at low concentrations (µM). When this eluent was used in a polymer coated, reversed-

phase open-tubular column, a substantial EOF was developed. This, in turn, provided

highly efficient and fast separations of neutral compounds. In 1991, Pfeffer and Yeung

[18] used OT-CLC for the separation of anions. This separation involved partitioning of

the anions with the help of the ion-pairing agent, tetrabutylammonium (TBA) cation, being

adsorbed onto the surface of an open-tubular column. The separation of the anions was

based on differences in retention, rather than differences in electrophoretic mobility. The

ion pairing agent, which was added in the mobile phase at low concentrations (600 µM),

was used to control the affinity of the anions for the stationary phase.

Garner and Yeung [95] proposed the use of an ion-exchange mechanism for the

separation of compounds with similar mobilities, such as 4-amino-1-naphthalenesulphonic

acid and 5-amino-2-naphthalenesulphonic acid. Capillaries coated with a hydrophobic

stationary phase were shown to be dynamic ion exchangers when the quaternary

ammonium compound, CTAB, was added to the mobile phase. CTAB was dynamically

adsorbed to the column, forming a charged double layer. Sample ions were then retained

due to their coulombic attraction to the double layer. In this study, they showed that the

retention of the sample ions could be varied in this system by changing various parameters

such as the concentration of the buffer ion, the addition of an organic modifier, and the

concentration of the cationic surfactant CTAB.

Liu et al. [105-107] developed a novel preparation method for a physically

adsorbed stationary phase in OT-CEC for the separation of both achiral and chiral

Page 42: Analytical separations using packed and open-tubular

22

compounds. The adsorbed stationary phases were prepared by simply rinsing the capillary

with a mobile phase containing a cationic surfactant, such as CTAB or a basic chiral

selector, such as protein, peptide or amino acid. After adsorption of the stationary phase,

the capillary was rinsed again with a mobile phase, which did not contain either CTAB or a

chiral selector. Five alkyl substituted benzenes were baseline separated within 6 min [105].

The run-to-run reproducibility of retention time was good with relative standard deviation

(RSD) values of less than 2.3%. However, the day-to-day reproducibility was not as good

with RSD values of less than 10% [106, 107].

Baryla et al. [100] described the adsorption mechanisms and aggregation properties

of the cationic surfactants CTAB and didodecyldimethylammonium bromide (DDAB) that

were used for the preparation of dynamically adsorbed stationary phases. Atomic force

microscopy (AFM) was used for the elucidation of the coating morphology. In addition,

separation of the basic proteins lysozyme, ribonuclease A, α-chymotrypsinogen A,

cytochrome c, and myoglobin was performed. Two of the five proteins were irreversibly

adsorbed to the wall. All three peaks that were present were broad and tailed due to wall

adsorption. The use of a CTAB-coated capillary gave a baseline separation of three of the

proteins in less than 15 min with high efficiencies. When a DDAB-coated capillary was

used all five proteins were separated in less than 6 min. This was due to the increased

surface coverage provided by DDAB.

(ii) Polymeric Surfactants

Although micelles have been successfully used in separations, they have some

significant limitations that cause problems in MEKC. The most important problem

associated with conventional micelles is the dynamic equilibrium between the surfactant

Page 43: Analytical separations using packed and open-tubular

23

molecules and the micelle. The presence of dynamic equilibrium lowers the stability of the

pseudostationary phase. PAHs, which are considered extremely hydrophobic compounds,

are difficult to separate because they co-migrate with the micelle. This problem can be

solved by either increasing the concentration of the surfactant or by adding an organic

modifier in the BGE solution. However, organic modifiers usually perturb the micelle

structure. In addition, an increase in the concentration of the surfactant may cause longer

migration times and Joule heating, which has detrimental effects on MEKC separations [6,

7, 11]. In order to compensate the above problems, new pseudostationary phases have been

developed by use of chiral and achiral polymeric surfactants.

In 1994, Wang and Warner [120] reported the use of a polymeric surfactant added

to the BGE in MEKC. Polymeric surfactants offer several distinct advantages over

conventional micelles [121-126]. Firstly, polymerization of the surfactant eliminates the

dynamic equilibrium due to the formation of covalent bonds between the surfactant

aggregates. This, in turn, enhances stability and improves resolution. Secondly, polymeric

surfactants can be used at low concentrations because they do not depend on the CMC.

This usually provides higher efficiencies and rapid analysis.

Harrell et al. [119] developed a dynamic coating by use of a novel nonionic micelle

polymer, poly (n-undecyl-α-D-glucopyranoside). They applied this coating to the

separation of seven tricyclic antidepressants. Although physical and dynamic adsorptions

have simple and rapid coating procedures and good reproducibilities, they usually have

short lifetimes and limited pH ranges [92, 112, 113]. Both coatings are adsorbed to the

capillary wall via electrostatic interactions and hydrogen bonding. These interactions are

weaker than covalent bonds. However, multiple electrostatic interactions within a coating

Page 44: Analytical separations using packed and open-tubular

24

that involves a layer-by-layer deposition process provides greater stability and longer

lifetime [92, 93, 112-114, 121].

A layer-by-layer coating is termed a polyelectrolyte multilayer (PEM), which is

usually constructed in situ by use of alternating rinses of positively and negatively charged

polymers [114, 127-130], where the negatively charged polymer may be a polymeric

surfactant [92, 93, 121]. A layer of polymer adds to the oppositely charged surface,

reversing the surface charge and priming the film for the addition of the next layer via

electrostatic forces. The advantages of such coatings are two-fold. First, less consumption

of the polymer is required, since it is adsorbed onto the capillary wall. Second, there is less

detection interference between the polymer and the analyte of interest, which in turn,

makes the system more amenable to coupling with mass spectrometry (MS).

Although the basic idea of the PEM coating is simple, a theoretical description is

relatively complex due to the long range of the coulombic interaction between layers. In

addition, several techniques have been employed (neutron reflectometry, atomic force

microscopy, infrared spectrometry) in order to better understand the mechanism of the

PEM coating formation. However, it is still not very well understood. The PEM coating is

constructed by the use of the simple procedure described above. The first layer should have

high adsorptivity in order to strongly attach the multilayer to the substrate. Several studies

have shown that after the first four or five layers, the thickness of the coating increases

linearly with the number of layers, and the zeta potential has a value of around zero. The

thickness of each layer depends on the amount of salt added to the polymeric solutions,

and not on the surface charge of the substrate. However, the total thickness of the coating

increases with a surface charge increase. In addition, it is observed that salt has the

Page 45: Analytical separations using packed and open-tubular

25

strongest effect on the polyelectrolyte layer thickness. Polymer concentration, polymer

type, molar mass, salt type, deposition time, and solvent are less important variables [127,

131, 132].

All polymer deposition solutions contain an amount of salt, usually NaCl. Salt

moderates the electrostatic interactions that occur between the oppositely charged

polyelectrolyte sections, Pol+ and Pol-, according to the following chemical equation:

Pol+Pol-m + Na+

aq + Cl-aq ↔ Pol+Cl-

m + Pol-Na+m

where m refers to the multilayer phase. The charges on a polymer can be balanced by

either those on the oppositely charged polymer or by the salt counterions within the film.

In the first case, where the positive charge of one polymer is balanced by the negative

charge of another polymer, the compensation is called intrinsic (Figure 1.9a). In the second

case, the compensation is extrinsic, since much of the polymer charge is balanced by the

salt ions that are added in the polymer deposition solution (Figure 1.9b) [127, 131].

In this dissertation, the PEM coating approach is explored, and its performance is

evaluated by use of different drug compounds. For the achiral studies, the polymeric

surfactant poly (sodium N-undecanoyl-L-glycinate), poly (L-SUG), was used as the anionic

polymer and poly (diallyldimethylammonium chloride), PDADMAC, was used as the

cationic polymer [92]. A diagrammatic scheme of the PEM coated capillary is illustrated in

Figure 1.10. This diagram is not intended to give an actual structural representation of the

bilayer, but rather a representation of the order of polymer deposition. More details

concerning this work are given in Chapter 3.

Page 46: Analytical separations using packed and open-tubular

26

+

-

- --

- --

----

--

-

+++

++

++++++

++

- --

- --

----

--

-

++

++ + + +

+ ++ + +

+

-

+

Na+

Na+

Na+

Na+

Cl- Cl- Cl-Cl-

(a) Intrinsic

(b) Extrinsic

+

-

- --

- --

----

--

-

+++

++

++++++

++

- --

- --

----

--

-

++

++ + + +

+ ++ + +

+

-

+

Na+

Na+

Na+

Na+

Cl- Cl- Cl-Cl-

(a) Intrinsic

(b) Extrinsic

Figure 1.9 (a) Intrinsic, (b) Extrinsic charge compensation.

Kamande et al. [93] investigated the use of a PEM coating for the separation of

phenols and benzodiazepines. For their studies, they used the polymeric surfactant, poly

(sodium undecylenic sulfate), poly (SUS), and PDADMAC in a single bilayer PEM

coating. The run-to-run and capillary-to-capillary reproducibilities were very good, and the

RSD values of EOF were less than 1.5%. The endurance of the coating was more than 100

runs. In addition, this study demonstrated the importance of the PEM coating by

comparing the separation of benzodiazepines using CZE and MEKC.

Page 47: Analytical separations using packed and open-tubular

27

N+N+N+N+N+N+N+N+N+N+ N+N+N+N+N+N+N+N+N+N+N+ N+

CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO-CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NH

C O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO-

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-

Silicate surface Si

O-O(deprotonated

silanol groups)

Capillary wallSi

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-O

Si

O-

Silicate surface Si

O-O

Si

O-O(deprotonated

silanol groups)

Capillary wall

BilayerBilayer

Figure 1.10 Scheme of the PEM-coated capillary.

(iii) Charged Polymers

Erim et al. [104] reported the use of a simple method for the preparation of a

polyethyleneimine (PEI) coating on the inner surface of fused-silica capillaries. The PEI

layer was constructed by flushing the capillary with a solution that contained high-

molecular-mass PEI. These authors investigated the reproducibility and the long-term

stability of the PEI coating by measuring the plate numbers and retention times of some

basic proteins. The RSD values in migration times for run-to-run reproducibilities ranged

from 0.5 to 1.5%, and for column-to-column reproducibilities ranged from 1.9 to 2.8%.

The stability of the coating was tested by flushing the capillary with a solution at pH 11.0

for 60 hours. The migration times of all proteins after 60 hours slightly increased, and the

efficiencies were not altered.

A stable modification of the inner wall was also developed by using a simple

coating procedure, i.e. successive multiple ionic-polymer layer (SMIL) coating. Katayama

Page 48: Analytical separations using packed and open-tubular

28

et al. [112, 113] used polybrene (PB) as the cationic polymer, and dextran sulfate (DS),

alginic acid, or hyaruronic acid as an anionic polymer. In their first study [112], they

established an anion-modified capillary (SMIL-DS capillary) by first attaching the cationic

polymer to the capillary wall, and then the anionic polymer to the cationic polymer layer.

Efficient separation of acidic proteins was achieved even at pH values under 7.4. The

ability to prevent the adsorption of the acidic proteins to the SMIL-DS capillary wall was

due to the presence of sulfonic groups in the DS layer. These groups provided strong

repulsion between the protein and the capillary wall. However, the SMIL-DS capillary was

not suitable for the separation of basic proteins since the capillary wall was negatively

charged through pH 2-11. In a pH range of 2-11, the SMIL-DS capillary exhibited a pH-

independent EOF from anode to cathode. The endurance of the SMIL-DS coated capillary

was more than 100 runs and the capillary-to-capillary variation was less than 1%. In their

next study [113], this group further modified the SMIL coating and developed a cation-

modified capillary (SMIL-PB capillary) by attaching the cationic polymer to the anionic

polymer layer. The SMIL-PB capillary was applied to the separation of basic proteins. The

separation was possible even when the pH of the mobile phase was near the pI of the

protein. The SMIL-PB capillary demonstrated strong endurance with the achievement of

600 runs and excellent chemical stability against 1 M sodium hydroxide (NaOH) and 0.1

M hydrochloric acid (HCl). The RSD values of the run-to-run, day-to-day, and capillary-

to-capillary reproducibilities were below 1%.

Poly (styrene sulfonate), PSS, is another anionic polymer that has been used in a

PEM coating procedure. Graul et al. [114] used this coating for the separation of a series of

basic proteins. The coatings used for protein separations and reproducibility studies

Page 49: Analytical separations using packed and open-tubular

29

consisted of 6.5 layer pairs. A layer pair is a layer of cationic polymer plus a layer of

anionic polymer, also termed a bilayer. The polymer deposition solutions for the

construction of the first 3.5 bilayers contained no salt, and for the last 3 contained 0.5 M

sodium chloride (NaCl). They observed an excellent run-to-run stability over a wide range

of pH. These capillaries were also proven to be very stable to extremes of pH (pH 12.0 and

pH 2.0) and ionic strength, and to dehydration/rehydration. The RSD values of capillary-

to-capillary reproducibilities were less than 2% when both pH 6.0 and pH 4.0 were used.

In addition to the reproducibility control of EOF, these authors achieved stable flow rates

immediately after exposure of the column to mobile phase, as well as reverse flow.

1.3.2.2 Covalent Bonding and/or Crosslinking

Fixation of a layer by covalent bonding and/or crosslinking is another approach

used for modifying the capillary [58, 63, 101, 102, 133-142]. Although this approach

offers a long capillary lifetime, it usually requires a more complicated coating procedure.

Rehder et al. [58] studied the electrochromatographic separation of bovine β-lactoglobulin

variants A and B by covalently attaching DNA aptamer, which form a G-quartet

conformation to a capillary surface. Aptamers are short, single-stranded oligonucleotides

that are recognized for their high affinity and specific binding to target molecules. Their

use as a stationary phase offers many advantages. They are stable, easy and simple to

produce, manipulate, and attach to silica surfaces, and they are compatible with a wide

range of separation conditions and mobile phases. For the preparation of the stationary

phase, oligonucleotides were covalently attached to the inner capillary surface using the

organic linker molecule, sulphosuccinimidyl 4-(N-maleimidomethyl) cyclohexane-1-

carboxylate.

Page 50: Analytical separations using packed and open-tubular

30

A porphyrin derivative, 5,10,15,20-tetrakis(penta fluorophenyl)porphyrin,

H2TPFPP, was also used as a capillary wall modifier in OT-CEC by Charvatova et al.

[102]. H2TPFPP was physically adsorbed or covalently bonded to the capillary surface for

the separation of aromatic carboxylic acids. In both covalently bonded and physically

adsorbed phases, the capillaries were filled with a solution of the porphyrin derivative in

dichloromethane. In the first case, both ends of the capillary were closed and the capillary

was left overnight at room temperature. In the second case, the capillary was left in the

vacuum oven for 2 hours at 60 °C. The synthetic strategy used for the preparation of

chemically bonded phases was based on the generation of anionic silanol groups on the

capillary wall, which were then employed for nucleophilic substitution at the para position

of H2TPFPP. The main advantage of H2TPFPP is that it is capable of increasing the EOF.

The results from this study showed that coating of the capillary with H2TPFPP gave a

better resolution at all pH values tested. Although both covalent bonding and physical

adsorption resulted in similar results at pH 8.5 and pH 6.0, covalently bonded capillaries

offered better results. At pH 5.0, physically adsorbed coatings lead to results, which were

not reproducible.

In another study, Chiari et al. [101] reported the use of poly (vinylamine), PVA, for

the separation of polyanionic acids. The coating procedure consisted of two steps: (i)

dynamic adsorption of PVA, and (ii) crosslinking and derivatization of PVA by

nucleophilic addition of primary amino groups to the conjugated double bonds of N, N-

methylenebisacrylamide and N, N, N-trimethylaminoethylacrylamide. The best stability of

this coating was achieved with the crosslinked PVA. There was no change in the EOF even

after 120 hours of continuous use at pH 4.0. RSD values and average transit times of ten

Page 51: Analytical separations using packed and open-tubular

31

injections were taken every 6 hours during the 120-hour study. The values between 0.2 and

1% indicated excellent reproducibility.

Calixarene-coated capillaries were also used in OT-CEC. Such capillaries are

expected to provide more efficient separations of larger compounds [139-143]. Most

recently, Wu et al. [140] have successfully used p-tert-butylcalix[8]arene bonded

capillaries for the separation of o-, m-, and p-benzenediols, α- and β-naphthols, and α- and

β-naphthylamines. These capillaries were prepared with the use of γ-

glycidoxypropyltrimethoxysilane as a bridge. The bonded capillaries showed good

separation selectivity, suggesting significant interactions between the analytes and the

bonded phase. In addition, for the evaluation of the stability of the capillary surface, five

batches prepared over a period of six months were tested. All capillaries exhibited high

stability and reproducibility.

1.3.2.3 Porous Layers

Another approach to increasing the phase ratio, the surface area, and the loading

capacity of a capillary is by use of porous-layer open-tubular columns [13, 143-146].

Huang et al. [13] introduced the utility of such columns with a positively charged

hydrocarbonaceous porous layer as the stationary phase for the separation of basic proteins

and peptides. The porous layer was crosslinked and prepared by in situ polymerization of

vinylbenzyl chloride and divinylbenzene in the presence of 2-octanol as a porogen inside a

silanized fused-silica capillary. The chloromethyl groups at the surface of the porous layer

were reacted with N, N-dimethyldodecylamine to obtain a positively charged surface with

fixed C12 alkyl chains. Stability was monitored daily by measuring the electroosmotic

mobility. They observed that the migration times remained constant for more than a week.

Page 52: Analytical separations using packed and open-tubular

32

A porous silica layer chemically modified with C18 groups was used as a stationary

phase by Crego et al. [143]. The 9.60 µm i.d. capillary used in this work had a thin porous

silica layer (0.70 µm) with C18 groups chemically bonded onto the prepared layer. These

authors investigated the effects of several experimental parameters, such as the volume

fraction and the type of organic modifier in the mobile phase, the concentration, type and

pH of the buffer on the electroosmotic mobility, the retention behavior of test analytes, and

the column efficiency. In their studies, they used a group of PAHs as a test mixture.

Consequently, they were able to optimize the separation of this group of analytes by

varying all of the operating parameters identified above.

The separation of charged analytes was also examined on porous-layer open-

tubular capillaries [146]. The preparation of the stationary phase was performed by use of

the sol-gel technique. This technique is discussed in detail later in this section. The

successful separation of acidic diuretics and N-alkylanilines as basic compounds at low pH

indicated that it is possible to separate charged compounds on porous-layer open-tubular

columns. Neutral aryl alkyl ketones were also successfully separated with high

efficiencies. However, basic pharmaceutical drugs demonstrated severe peak tailing and

were poorly separated by use of these columns. In addition, strongly acidic analytes were

not detected due to their high electrophoretic mobilities.

1.3.2.4 Chemical Bonding After Etching

Etching is another approach to overcoming the major problems associated with OT-

CEC. The etching process increases the surface area of the capillary by a factor of up to

1000 [147]. Therefore, more stationary phase can be attached to the wall, which in turn,

increases the loading capacity of the column. In addition, during the etching process,

Page 53: Analytical separations using packed and open-tubular

33

dissolution and redeposition of silica material occurs. This creates radial extensions from

the wall that decrease the distance an analyte has to travel in order to interact with the

stationary phase [88]. The etching process was first introduced by Onuska et al. [147] for

gas chromatography (GC), and it was then modified for OT-CEC by Pesek et al. [14, 39,

41, 88-91, 148-156].

The entire procedure includes etching followed by chemical bonding. The etching

process uses ammonium hydrogen difluoride as the etching agent which, under controlled

conditions of temperatute and reaction time, can increase the surface area. The method of

bonding an organic moiety to the etched surface of a capillary utilizes the

silanization/hydrosilation method. In the first step, which involves silanization, the etched

surface is reacted with triethoxysilane (TES) to produce a hydride layer that is deposited on

the surface. In the second step, an organic moiety is attached to the hydride intermediate

via hydrosilation. This is accomplished by passing a solution that contains a terminal

alkene and a catalyst, typically hexachloroplatinic acid, through the capillary [92].

In a 1997 study, Pesek et al. [148] attached two organic moieties, octadecyl and

diol, to the etched capillary wall. Peptide (angiotensins) and protein samples were used for

comparison studies between these two columns and a bare capillary. The migration times

for each of the analytes increased in the following order: bare<diol<C18. Longer migration

times were observed when the C18-modified etched capillary was used. This was probably

due to stronger analyte-bonded phase interactions. From these studies, the authors

concluded that the bonded etched capillaries have significantly different retention

characteristics from a bare capillary. In addition, the separation capabilities of the two

Page 54: Analytical separations using packed and open-tubular

34

modified capillaries varied due to differences in the chemical properties of the two organic

moieties.

Another study from Pesek et al. focused on a comparison of the migration behavior

of two antibiotics (ampicillin and gentamycin) between several types of etched open-

tubular capillaries [90]. The types of columns used included (i) etched and unmodified, (ii)

etched and bonded with an octadecyl moiety, (iii) etched and bonded with a cholesterol

moiety, (iv) etched and coated with anionic fluorosurfactant (FSA), and (v) etched and

coated with neutral zwitterionic fluorosurfactant (FSK). Each of the antibiotic samples

consisted of several components, and the resolution capabilities of the different types of

columns were compared. Their studies showed that each type of column gave a different

elution pattern due to the stationary phase effect. A number of variables had to be

optimized in order to get the maximum separation for these samples. Optimization was

achieved by varying several experimental conditions, such as the type of stationary phase,

the pH and composition of the buffer, and the applied voltage.

Matyska et al. [89] investigated the performance characteristics of n-octadecyl and

cholesterol- modified capillaries for the analyses of synthetic peptides. The results

indicated that the resolution and selectivity of peptides could be affected by varying the

type of chemically bonded group and the nature of the surface chemistry used to modify

the capillary wall. In addition, other experimental variables, such as the type and

percentage of organic solvent modifier, and the pH of the buffer system can affect the

retention of peptides. For the evaluation of the reproducibility and the stability of both

types of open-tubular columns, four batches were prepared over a period of two years. For

this study, a mixture of two small basic compounds, serotonin and tryptamine, and a

Page 55: Analytical separations using packed and open-tubular

35

mixture of two basic proteins, chicken and turkey lysozyme, were used. The measurement

of the selectivities and efficiencies of these analytes showed that both the reproducibility

and stability were excellent. All of the etched modified capillaries used in this study still

performed well after more than 200 runs.

1.3.2.5 Sol-Gel Technique

As noted earlier, the preparation procedure for columns with expansive surface

areas and large loading capacities involves two steps: (i) etching or laying down a porous

layer, and (ii) attaching functional groups to the intermediate layer by chemical bonding

[30]. However, this two-step procedure may be time consuming. Another approach, which

involves the construction of polymeric stationary phases, involves good column stability

and high surface ratio. However, the main problem associated with these columns is the

poor column efficiency as a result of the slow diffusivity of analytes in the layers [145]. In

order to circumvent these drawbacks, the sol-gel technique was introduced for the

synthesis of organic-inorganic hybrid materials that are used as stationary phases in OT-

CEC. Recently, the sol-gel technique, which combines the synthesis of a bonded phase and

a supporting porous silica film in a single step, has gained much attention [29, 40, 145,

146, 157-164].

The advantages of the sol-gel processed stationary phases include high stability,

high mass loadability, great column efficiency, large surface area leading to higher

retentions, and a relatively simple preparation procedure [157, 159-161]. The sol-gel

process consists of five major steps: (i) hydrolysis, (ii) condensation and polycondensation

of sol-gel precursors, (iii) casting of the sol, (iv) aging, and (v) drying. In the first step, a

metal alkoxide is hydrolyzed under acidic or basic conditions to form the corresponding

Page 56: Analytical separations using packed and open-tubular

36

metal hydroxide, which condenses and polycondenses to form “sol” particles. These “sol”

particles are then crosslinked to form a wet “gel,” which can take any desired shape. A dry

“gel,” or a “xerogel” is finally formed after the loss of solvent through aging and drying.

By controlling different processing parameters, such as temperature, pressure, alcohol

solvent, and pH during the different steps of the process, one can fabricate sol-gel derived

materials with specific structures [15, 157, 160-162].

In 1995, Guo and Colon [145] introduced the organosilicon sol-gel fabricated

coatings in OT-CEC. They synthesized a porous glass film in such a way that the

stationary phase (octyl groups) is incorporated into the glass matrix during the glass

formation process. This film was then used in capillaries, and functioned as a stationary

phase. The performance of this sol-gel derived stationary phase was evaluated by use of a

group of PAHs. Its stability was also studied under acidic and basic conditions. In order to

test the stability at low and high pH values, the capillary was washed with 1%

trifluoroacetic acid (pH ≈ 0.3) and with the mobile phase methanol/1 mM phosphate buffer

(70:30) at pH 11.4. Their results indicated that a retentive layer was still on the surface of

the capillary, even after exposure to high- and low-pH conditions. Finally, the mass

loadability of the column was evaluated by injecting different concentrations of

naphthalene in the capillary. The fact that the efficiency started to degrade at a

concentration of 100 mM indicated overloading. In a conventional column, the efficiency

deteriorated at concentrations below 10 mM.

Rodriguez and Colon [162] used n-octadecyltriethoxysilane (C18-TEOS) and

tetraethoxysilane (TEOS) as the precursors to fabricate an organic-inorganic hybrid

material by use of the sol-gel method. Fused-silica capillaries were coated with the C18-

Page 57: Analytical separations using packed and open-tubular

37

TEOS/TEOS mixture and were tested using OT-CEC. A test mixture containing toluene,

ethylbenzene, biphenyl, dimethylnaphthalene and amybenzene was separated by using two

columns coated with (i) TEOS, and (ii) C18-TEOS/TEOS sol-gel derived materials.

Additionally, eight PAHs were separated by use of a C18-TEOS/TEOS-coated column.

These compounds were baseline separated only in the columns containing the organically

modified composite (C18-TEOS/TEOS).

A novel sol-gel approach for fabrication of open-tubular columns was described by

Hayes and Malik [158]. A surface-bonded octadecylsilane (ODS) stationary phase coating

was created by sol-gel chemistry. The use of a deactivating reagent, phenyldimethylsilane

(PheDMS), in the sol-gel solution was evaluated. The performance of both nondeactivated

and deactivated sol-gel columns was also evaluated with the test mixtures of PAHs,

benzene derivatives, and aromatic aldehydes and ketones. The column deactivation

resulted in a higher separation efficiency and better resolution. In addition, the RSD values

for the run-to-run reproducibility of the sol-gel columns were less than 0.70% and 0.80%

for the aromatic hydrocarbons and the aromatic carbonyl compounds, respectively.

Wang et al. [161] used 1,4,7,10-tetraazacyclotridecane-11,13, dione

(dioxo[13]aneN4) for the first time in the sol-gel approach for the preparation of

dioxo[13]aneN4 modified capillaries. In comparison with columns prepared by the sol-gel

process with just TEOS, the sol-gel derived macrocyclic dioxopolyamine columns were

able to give better separations of a mixture of isomeric nitrophenols and benzenediols, a

mixture of isomeric aminophenols and diaminobenzenes, and a mixture of four

neurotransmitters. The reproducibilities of migration times and plate numbers were

Page 58: Analytical separations using packed and open-tubular

38

satisfactory with RSD values of less than 2% and less than 8.5% for the respective run-to-

run and column-to-column reproducibilities.

1.4 Chirality

Chirality is the geometric property that is responsible for the nonidentity of an

object with its mirror image [165]. In other words, an object is chiral if it is non-

superimposable on its mirror image. In 1848, Pasteur [166] marked the beginning of

chirality and chiral separation. He discovered that the resolution of the racemic mixture of

ammonium sodium tartrate yielded two enantiomorphic crystals. The separation was

perfomed by use of a pair of tweezers and a hand lens. He also observed that the crystals

gave a left and a right rotation of polarized light. In addition, he proposed that since the

difference of the optical rotation was examined in solution, the molecules are the ones that

are mirror images of each other [167].

Chiral compounds are molecules that relate to each other like a pair of hands. The

word chiral was derived from the Greek word cheir, which means hand. The chiral

molecules are also termed optically active because they have the ability to rotate plane-

polarized light. Optically active molecules that rotate light to the left are called

levorotatory (L), and they have the negative sign (-). If the molecules rotate the light to the

right, they are said to be dextrorotatory (D) or positive (+). In addition, chiral compounds

are also termed as molecules with non-superimposable mirror images. Molecules that are

mirror images of each other are also called enantiomers or optical isomers. These

molecules have an asymmetric tetrahedral carbon atom with four different substituents. If

the priority of the substituents is in a clockwise direction, the configuration is called R

(right or rectus). If the priority is in a counterclockwise direction, the configuration of the

Page 59: Analytical separations using packed and open-tubular

39

chiral center is S (left or sinister). These differences are due to the asymmetric element in

the chiral molecule that can be a center, an axis, or a plane of asymmetry. An equimolar

mixture of two enantiomers is called a racemate, and it is designated by the symbol (±) [6,

165, 168].

Although enantiomers rotate plane-polarized light in opposite directions, these

molecules have identical physical properties such as boiling and melting points, and

spectroscopic properties such as nuclear magnetic resonance (NMR) spectra. However,

enantiomers in a racemic drug usually have different biological activities. For example,

dextromethorphan is an over-the-counter cough suppressant, whereas levomethorphan is a

controlled narcotic. Thalidomide is a sedative drug that was prescribed to pregnant women

in the early 1960’s. When this drug was taken during the early stages of pregnancy, it

prevented the normal growth of the fetus. This resulted in serious birth defects in

thousands of children around the world. It was later observed that the R-enantiomer was a

sedative, whereas the S form was the one that caused the foetal abnormalities. The

examples mentioned above illustrate the different pharmakokinetic characteristics and

pharmacological activities of each enantiomer in a racemic drug. Therefore, the

development of analytical techniques for the separation of chiral bioorganic molecules has

become very important [6, 168, 169].

In order to distinguish between two enantiomers in a racemic drug, a chiral selector

has to be added to the BGE solution. However, the mechanism of chiral discrimination is

not very well understood. In general, the “three point rule,” which was illustrated by

Easson and Stedman [170], describes the interactions that are necessary for chiral

discrimination. Chiral separation can be achieved if a minimum of three simultaneous

Page 60: Analytical separations using packed and open-tubular

40

interactions occur between the chiral selector and one of the enantiomers. Due to spatial

restrictions, the other enantiomer should attain at least two of these interactions. These

interactions can be hydrophobic interactions between the hydrophobic core of the polymer

and the analyte, and electrostatic interactions between the polar head group of the polymer

and the analyte. Another type of interaction is the dipole-dipole forces, such as hydrogen

bonding between the polar group of the chiral selector and the analyte. In addition, other

secondary interactions can occur, such as π-π interactions, ion-dipole bonds, and Van der

Waals forces.

1.5 Chiral Selectors in Open-Tubular Capillary Electrochromatography

Chiral separation has received considerable attention in many industries,

particularly the pharmaceutical industry. Several analytical techniques have already been

developed for the separation of chiral bioorganic molecules. One of them is OT-CEC,

where the chiral selector is coated and immobilized onto the inner surface of a capillary. In

1992, Mayer and Schurig [171] reported the first enantiomeric separation in OT-CEC by

using capillaries coated with immobilized Chirasil-Dex (permethyl-β-cyclodextrin

chemically linked to dimethylpolysiloxane). In addition to derivatized cyclodextrins [164,

171-178], cellulose [55, 60], proteins [42, 105-107, 179-183], molecular imprinted

polymers (MIPs) [184-195], and polymeric surfactants [121] were also used.

1.5.1 Derivatized Cyclodextrins

Mayer et al. [171-174] used an immobilized Chirasil-Dex for the separation of a

number of chiral compounds. Chirasil-Dex is a permethyl-β- or γ-cyclodextrin covalently

bonded to a dimethylpolysiloxane via an octamethylene spacer. Chirasil-Dex modified

capillaries proved to be buffer-resistant and stable under neutral and acidic conditions.

Page 61: Analytical separations using packed and open-tubular

41

Schurig et al. [175, 176] also investigated the use of an immobilized Chirasil-Dex

(mono-6-O-octamethylenepermethyl-β-cyclodextrin chemically linked to

dimethylpolysiloxane) in chromatography. The concept of unified enantioselective

chromatography was demonstrated for the chiral separation of hexobarbital by use of GC,

supercritical fluid chromatography (SFC), liquid chromatography (LC), and CEC [175].

The results from this study indicated that CEC is superior to GC, SFC and LC in regards to

several parameters, which are important in chiral separation. These parameters include the

chiral separation factor, α, peak resolution, Rs, and efficiency, N. In addition, Chiralsil-Dex

demonstrated a long lifetime, as well as configurational and thermal stability.

A related approach was also reported by Armstrong et al. [177] for the separation

of racemic mephobarbitol by OT-CEC. The capillary was coated with permethylated-β-

cyclodextrin that was covalently linked to dimethylpolysiloxane. This stationary phase

appeared to be stable and relatively unchanged by high temperature. Further, it could not

be removed from the capillary by conventional polar and nonpolar solvents or by

supercritical fluids.

Another approach for immobilization of β-cyclodextrin was developed by Pesek et

al. [178]. These authors etched the capillaries with ammonium hydrogendifluoride and

then modified them with a chiral selector by use of the silanization/hydrosilation method,

which was described earlier. The types of selectors evaluated include lactone, β-

cyclodextrin, and naphthylethylamine. For the etched cyclodextrin-modified capillary (2-

hydroxy-3-methacryloyloxypropyl-β-cyclodextrin), benzodiazepines were the only class of

compounds resolved among the test analytes used. They obtained partial resolution of the

two isomers for oxazepam, temazepam, chlorodiazepoxide, and diazepam. The stability of

Page 62: Analytical separations using packed and open-tubular

42

the column was also good since it was used for at least 200 injections with no observed

changes in enantiomeric resolution.

Wang et al. [164] coated columns with 2,6-dibutyl-β-cyclodextrin (DB-β-CD)

using the sol-gel technique. This new technology provided columns that demonstrated

good resolution, stability, and reproducibility on separation of some positional isomers and

chiral compounds. The chiral compounds used in this study were ibuprofen and

binaphthol. The two isomers of both analytes gave partial resolution.

1.5.2 Cellulose

Capillaries coated with cellulose derivatives were demonstrated by Francotte et al.

[55] for the separation of a number of chiral pharmaceuticals. Their columns were prepared

by coating 3,5-dimethylphenylcarbamoyl cellulose (DMPCC) or para-methylbenzoyl

cellulose (PMBC) on 50 µm i.d. fused-silica capillaries. The capillaries were coated at 35-

40 °C and 0.3 mbar. A film thickness of about 0.025 µm was obtained by using a filtered

solution of the cellulose derivative in dichloromethane or tetrahydrofuran (THF) for

DMPCC or PMBC, respectively. Although the enantioselectivity of the coating was

excellent, the column stability was low, and its lifetime was relatively short. Unfortunately,

column lifetime rarely exceeded one hundred injections. These problems were largely due

to the fact that the cellulose coating was not immobilized on the inner wall of the capillary.

Recently, another study was performed by Wakita et al. [60], in order to examine

the enantioseparation ability of cellulose tris(3,5-dichlorophenylcarbamate), (CDCPC), in

OT-CEC. The capillary was filled with a solution containing CDCPC derivative, styrene,

dry THF, and a solution of 2,2-azobisisobutyronitrile. The column was then heated at 60

°C for 20 hours to produce copolymerization. The covalently bound CDCPC was used as a

Page 63: Analytical separations using packed and open-tubular

43

chiral selector for the separations of trans-stilbeneoxide, laudanosine, etozolin, and

piprozolin. The enantiomers of the first two analytes were partially resolved and the

enantiomers of the last two analytes were baseline resolved on this CDCPC coated

capillary.

1.5.3 Proteins

Liu et al. [105-107] prepared chiral stationary phases for OT-CEC by use of a

method based upon the physical adsorption of a basic chiral selector such as protein,

peptide, and amino acid on the wall of a capillary. A number of chiral compounds were

successfully resolved using this system. The adsorbed protein, lysozyme, showed high

chiral selectivity, and the resolution ranged from 1.74 to 2.05.

Liu, Otsuka, and Terabe [42, 180-182] studied chiral separations by OT-CEC with

avidin as a chiral stationary phase. This coating was prepared by use of the physical

adsorption method that was proposed by Liu et al. [105-107]. A total of sixteen

enantiomeric compounds were separated with avidin adsorbed on the capillary wall [42,

180]. Among them, ten pairs of enantiomers were baseline separated, while the other

enantiomers were partly resolved with resolutions of 0.8-1.2. However, due to the low

phase ratio of OT-CEC, only enantiomers that have strong interactions with the stationary

phase can be separated. The RSD values for the run-to-run, day-to-day, and column-to-

column reproducibilities were less than 2.2%, 2.3%, and 1.1%, respectively. The column-

to-column reproducibility was better than the day-to-day reproducibility due to the small

loss of the stationary phase during the day-to-day reproducibility measurements. Therefore,

the capillary was rinsed daily with an avidin solution for 10 min prior to experiments.

Page 64: Analytical separations using packed and open-tubular

44

The feasibility of OT-CEC for quantitative analysis of enantiomers was also

examined with an avidin adsorbed capillary [181]. With conventional injection, the lowest

detectable concentration was 10-6 M. The limits of detection (LODs) for enantiomers were

below 1%. The detection sensitivity was improved by use of the field-enhanced sample

injection (FESI) technique. In this technique, a water plug is introduced hydrodynamically

into the capillary inlet end, and then, the sample solution is introduced with electrokinetic

injection. Using the FESI technique, the LOD was reduced to a 10-9 M level, and the

enantiomeric ratio was improved to 0.3%. In addition to the sensitivity, both the efficiency

and resolution were improved.

Another approach to improving the sensitivity of OT-CEC is to use an extended

light path (ELP) capillary, which has a bubble cell at the detection point. A physically

adsorbed avidin stationary phase was used to evaluate an ELP capillary and an etched

capillary [182]. With an ELP capillary, the peak height was enhanced by a factor of as

much as 10. However, the peak efficiency and resolution decreased. When an etched

capillary was used, the phase ratio was slightly enhanced by a factor of 1.64, as compared

to an unetched capillary.

1.5.4 Molecular Imprinted Polymers

Over the last few years, molecular imprinting has attracted considerable attention as

an approach for the preparation of chiral stationary phases in OT-CEC. This is due to the

high enantioselectivity and predetermined elution order that characterizes the MIPs [44].

The selectivity of the resultant MIP is predetermined by the choice of template (imprint)

molecule used for the imprint preparation. Monomers are selected for their ability to

interact with the template molecule by mainly electrostatic interactions, hydrogen bonding,

Page 65: Analytical separations using packed and open-tubular

45

dipole-dipole interactions, and hydrophobic interactions. The template molecule can either

be the analyte of interest or a structural analogue [194-196]. MIPs are prepared by

polymerization of a mixture that contains a template molecule, functional monomers, a

crosslinking agent, and an initiator in a nonpolar solvent (porogen). In the

prepolymerization mixture, the functional monomers interact with the imprint molecule,

and orient around it in a specific way to form complexes. During polymerization, in the

presence of a crosslinking agent, a rigid polymer is produced. After polymerization, the

imprint molecule can be removed by solvent extraction. This gives rise to a material that is

filled with cavities, and one that is complementary in size, shape, and chemical

functionality to the imprint molecule. These cavities enable the polymer to rebind the

original template molecule [189, 191, 193-197].

Schweitz and coworkers [189-192, 194, 195] prepared chiral stationary phases by

using molecular imprinting of the (R) and (S) enantiomers of propranolol. In 1997, Nilsson

et al. [189] performed the enantiomeric separation of β-adrenergic antagonists by using

three different capillary electrochromatographic methods. In these methods, they used

different cyclodextrins added to the mobile phase, a crosslinked protein gel, and a

molecularly imprinted ((R) enantiomer of propranolol) superporous polymer as chiral

selectors. All three methods were able to resolve or partially resolve all β-adenergic

antagonists into their enantiomers. Very rapid enantiomer separations of propranolol were

also studied by the use of short super-porous monolithic MIPs [194]. MIP stationary

phases were synthesized by use of an in situ photo-initiated polymerization reaction.

Propranolol enantiomers were resolved in less than 1 min. In addition, more MIP coatings

were synthesized by the use of a surface-coupled radical initiator [195]. A

Page 66: Analytical separations using packed and open-tubular

46

prepolymerization mixture was prepared by dissolving the template molecule (S)

propranolol, the functional monomer methacrylic acid, and the crosslinking monomer

1,1,1-tris(hydroxymethyl)propane trimethacrylate in either toluene, dichloromethane, or

acetonitrile. The prepolymerization mixture was then introduced into the capillary and the

ends were sealed. The polymerization was performed by illuminating the capillary with a

UV lamp. The use of different solvents facilitates control of the coating, regarding its

morphology and appearance. Furthermore, all MIP coatings synthesized using the different

solvents, provided enantiomeric separation of propranolol when the (S) enantiomer was

used as the template molecule.

Tan et al. [194] reported a method for in situ preparation of molecular imprint

polymers as thin films inside 25 µm i.d. fused-silica capillaries. Methacrylic acid and 2-

vinyl pyridine were used as the functional monomers. Ethylene dimethacrylate or

trimethylol propane trimethacrylate was used as the crosslinker and toluene as the porogen.

Chiral separations of the enantiomers D- and L- dansyl phenylalanines were achieved in

both OT-LC and OT-CEC. The resolution in OT-CEC was much higher than in OT-LC.

1.5.5 Polymeric Surfactants

In this dissertation, the chiral separations of 1,1’-binaphthyl-2,2’-

dihydrogenphosphate (BNP), 1,1’-bi-2-naphthol (BOH), secobarbital, pentobarbital and

temazepam are investigated by using the PEM coating [121]. However, the PEM coating

procedure used in the achiral studies [92] needed to be modified in order to achieve chiral

separations. PDADMAC was used as the cationic polymer, and the polymeric surfactant

poly (sodium N-undecanoyl-L-leucylvalinate), poly (L-SULV), was used as the anionic

polymer. Thus far, poly (L-SULV) has shown the best chiral discrimination ability for a

Page 67: Analytical separations using packed and open-tubular

47

number of pharmaceutical compounds [126]. The PEM coating approach applied to chiral

separations is discussed in detail in Chapter 4.

1.6 Laser Scanning Confocal Microscopy

Since the beginning of observational science, the need to observe objects smaller

than those visible by eye alone has driven the development of microscopy. Conventional or

“far-field” light microscopy is the oldest form of microscopy. Very early in the

development of microscopy, it was realized that the spatial resolution that could be

achieved was limited by diffraction phenomena. This limit is best known as the diffraction

limit, or Abbe’s limit. Abbe showed that the theoretical resolution limit for an objective

lens is approximately half the wavelength, λ , of light employed. This relationship is

demonstrated in the following equation:

NAR λ61.0= (1.16)

where NA is the numerical aperture of the objective lens. The numerical aperture of an

objective lens is a measure of the angular size of the focusing cone of light and is defined

as:

θsinnNA = (1.17)

where θ is the acceptance angle of the lens, and n is the refractive index of the medium

(usually air, n=1.0) between the sample and objective lens. Therefore, the resolution in far-

field optical microscopes depends on the wavelength of light used. The use of a shorter

wavelength light can improve the spatial resolution. However, when ultraviolet light is

used several difficulties arise. These low wavelengths are usually absorbed by atmospheric

Page 68: Analytical separations using packed and open-tubular

48

gases, liquids and many of the materials from which conventional optical elements are

constructed [198, 199].

In order to eliminate the problems caused by the far-field optics, and to improve the

optical resolution, near-field scanning optical microscopy (NSOM) was developed. Synge,

in 1928, was the first to propose the development of NSOM as a scanned probe

microscopy. He proposed the use of a light source formed from a subwavelength-size

aperture in a metal screen (Figure 1.11). When light passes through the backside of the

aperture, fields are produced in and near the front side of the hole. Only the sample region

directly beneath the aperture is illuminated. Images can be formed by moving the aperture

and sample relative to one another (raster scanning). The signal is observed by the use of a

single-element detector. The resolution is determined by the size of the aperture and its

distance from the sample, rather than the wavelength of the light [198].

Confocal microscopy is a valuable tool for obtaining three-dimensional images of

thick objects. It is widely used in the fluorescence mode for imaging biological species and

in the reflection mode for imaging objects of different forms. In the fluorescence mode, the

objects are stained with fluorescent dyes. Fluorescent dyes are molecules that absorb light

at one wavelength and emit light at another wavelength. When these molecules absorb

light at a specific absorption wavelength, the electron rises to a higher energy level, called

the excited state. In this state, electrons are unstable; thus, they return to the ground state

by releasing energy in the form of heat and light. This emission of energy is called

fluorescence. Due to some loss in energy in the form of heat, the emitted light has less

energy. Therefore, the light is emitted at a longer wavelength than the absorbed light [200,

201].

Page 69: Analytical separations using packed and open-tubular

49

Near Field

Metal ScreenAperture

<< λλλλ

Incident Plane Waves

Sample

<< λλλλ

Far Field λ

X

Y

Z

Near Field

Metal ScreenAperture

<< λλλλ

Incident Plane Waves

Sample

<< λλλλ

Far Field λ

X

Y

Z

Y

Z

Figure 1.11 Model for the aperture-based near-field optical microscope.

Figure 1.12 is a schematic diagram of a confocal microscope. This microscope

differs from a conventional optical microscope in that it images points of light rather than

large volumes of light. In general, it illuminates and images the object one point at a time

through a pinhole. Light from a laser source passes through a small pinhole, which acts as

a spatial filter, to an objective lens. This lens focuses the light onto a small spot on the

object, at the focal plane of the objective lens. The light that is reflected back from the

illuminated object is collected by the objective lens, and it is partially reflected by a beam

splitter. Then, it is directed at a confocal aperture, or pinhole, that is placed in front of the

detector (i.e. photomultiplier tube). The pinhole rejects light that did not originate from the

Page 70: Analytical separations using packed and open-tubular

50

focal plane of the microscope objective lens. Therefore, only a very small volume of light

is focused at any time. This is what gives the system its confocal property [200, 202].

The laser scanning confocal microscope (LSCM) is the most widely used confocal

microscope in the areas of biomedicine and material science. It works in both transmission

and reflection modes. In addition, it gives higher resolution and thinner optical sections

than those obtained by use of the conventional microscope. By scanning a lot of thin

sections through the sample, a very clear three-dimensional image of the specimen can be

produced. Scanning is important in LSCM because only a small volume is illuminated at a

time, but a larger area should be collected for producing a better image. Scanning can be

accomplished by either beam scanning or stage scanning. The advantage of beam scanning

is that a total image can be generated from the small volumes of light without moving the

specimen. During beam scanning the specimen is stationary, and it is scanned by flying

spots of light. However, the advantage of stage scanning over beam scanning is that the

total light intensity is much greater.

The most important components of a typical LSCM system are a light source, an

objective lens, a scanning device, pinhole lens, a beam splitter, a detector, an analog signal

processor, a digital signal processor, an image analyzer, and a computer. In LSCM, the

light source should be a stable source of spatially coherent light. Therefore, the single

mode laser is most commonly used for this purpose. For fluorescence applications, multi-

line lasers, such as Ar, KrAr, and HeNe, can be used. The objective lens is used both to

illuminate and receive the reflected light from the illuminated spot on the specimen. After

the laser beam is converged by the objective lens and focused on the specimen, the scanner

scans the focused spot in an x, y raster pattern in order to produce an image of the sample

Page 71: Analytical separations using packed and open-tubular

51

point by point. The reflected light from the specimen is collected by the objective lens, and

reflected off by the beam splitter to the pinhole and the detector. The pinhole is an

important component for the determination of both the axial ( 2

2)(NA

nzR λ= ) and lateral

(NA

yxR λ4.0),( = ) resolution of the microscope. When the pinhole is large, it transmits

more light to the detector. This, in turn, generates a larger signal with lower resolution.

However, a smaller pinhole results in a higher resolution and a lower signal-to-noise ratio.

After the light passes through the pinhole, it is directed to the detector, which generates an

electrical signal. Then, the electrical signal is amplified and sent to a scan converter, where

it is combined with the x and y position signals in order to produce an image for display

[203-205].

1.7 Scope of Dissertation

In this dissertation, both packed and open-tubular capillary electrochromatographic

methods were developed for the achiral and chiral separations of various classes of

analytes. Chapter 2 is a report of studies on the packed mode of capillary

electrochromatography (CEC). The goal of this work was to separate seven

benzodiazepines by the use of a 40 cm packed bed of Reliasil 3 µm C18 stationary phase.

Optimal conditions were established by varying the mobile phase, the amount of organic

modifier, the buffer concentration, the applied voltage, and the column temperature. In

addition, the method developed was applied to the characterization of oxazepam in a

standard urine sample.

Page 72: Analytical separations using packed and open-tubular

52

Beam Splitter

Objective Lens

ObjectFocal Plane

Light Source

Source Pinhole

Detector Pinhole

Photodetector

In-focus light

Out-of-focus light

Figure 1.12 Schematic diagram of a confocal microscope.

In Chapter 3, the open-tubular mode of CEC, which is an alternative approach to

conventional CEC, is described. In this approach, a polyelectrolyte multilayer (PEM)

coating procedure was used to construct thin films on the capillary walls of fused silica

capillaries. For the fabrication of this coating both positively and negatively charged

polymers were utilized. Poly (diallyldimethylammonium chloride), PDADMAC, was used

as the cationic polymer and poly (sodium N-undecanoyl-L-glycinate), poly (L-SUG), was

Page 73: Analytical separations using packed and open-tubular

53

used as the anionic polymer. These capillaries were evaluated by use of seven

benzodiazepines as analytes. The run-to-run, day-to-day, week-to-week and capillary-to-

capillary reproducibilities were evaluated by computing the relative standard deviations

(RSDs) of electroosmotic flow (EOF). In addition, the chromatographic performance of the

monomeric form of the polymeric surfactant was compared for the separation of these

analytes.

Chapter 4 is an outline of an investigation of the chiral separations of 1,1’-

binaphthyl-2,2’-dihydrogenphosphate (BNP), 1,1’-bi-2-naphthol (BOH), secobarbital,

pentobarbital and temazepam by using the PEM coating approach and the polymeric

surfactant poly (sodium N-undecanoyl-L-leucylvalinate), poly (L-SULV), as the chiral

discriminator. However, the PEM coating procedure used in the achiral studies (Chapter 3)

needed to be modified in order to achieve chiral separations. The quaternary ammonium

salt PDADMAC was used as the cationic polymer, and the polymeric surfactant poly (L-

SULV) was used as the anionic polymer. Optimal conditions were established by varying

the additive (sodium chloride, 1-ethyl-3-methyl-1H-imidazolium hexafluorophosphate, 1-

butyl-3-methylimidazolium tetrafluoroborate) in the polymer deposition solutions, the salt

concentration, the column temperature, and the bilayer number. Reproducibilities were

also evaluated by using the RSD values of the EOF and the first peak (R-(+)-BNP).

Chapter 5 details an approach that utilizes open-tubular capillary

electrochromatography (OT-CEC) and laser scanning confocal microscopy (LSCM) to

further investigate the stability of the PEM coating after exposure to 0.1 M and 1.0 M

NaOH. The multilayer coatings used for these studies consisted of two and twenty layer

pairs. The structural changes of these coatings were monitored and imaged using LSCM

Page 74: Analytical separations using packed and open-tubular

54

after flushing the capillaries with 1.0 M NaOH. This technique also allowed a study of the

uniformity and discontinuity of the coating. Using OT-CEC, both electroosmotic mobility

and selectivity changes were measured after flushing the capillaries with 0.1 M and 1.0 M

NaOH. In addition, a correlation between flushing time and change in electroosmotic

mobility and selectivity was examined.

1.8 References

1. Heiger, D. N. In High Performance Capillary Electrophoresis – An Introduction, 2nd ed.; Hupe, P., Rozing, G., McManigill, D., van de Goor, T., Swedberg, S., Jegle, U., Verhoef, M., Jarke, L., Ford, J., Krull, B., Eds.; Hewlett-Packard Company: France, 1992.

2. Terabe, S.; Otsuka, K.; Nishi, H. J. Chromatogr. A 1994, 666, 295. 3. Chiari, M.; Nesi, M.; Righetti, P. G. In Capillary Electrophoresis in Analytical

Biotechnology; Righetti, P. G., Ed.; CRC Press, Inc.: Boca Raton, FL, 1996. 4. Skoog, D. A.; Holler, J. F.; Nieman, T. A. In Principles of Instrumental Analysis,

5th ed.; Sherman, M., Bortel, J., Messina, F., Eds.; Harcourt Brace College Publishers: Orlando, FL, 1998.

5. Chankvetadze, B. In Capillary Electrophoresis in Chiral Analysis; John Wiley & Sons Ltd: London, UK, 1997.

6. Williams, C. C.; Shamsi, S. A.; Warner, I. M. In Advances in Chromatography;

Brown, P. R., Grushka, E., Eds.; Marcel Dekker, Inc.: New York, NY, 1997, 363-423.

7. Yarabe, H. H.; Billiot, E.; Warner, I. M. J. Chromatogr. A 2000, 875, 179. 8. Terabe, S.; Otsuka, K.; Ichikawa, K.; Tsuchiya, A.; Ando, T. Anal. Chem. 1984,

56, 111. 9. Terabe, S.; Otsuka, K.; Ando, T. Anal. Chem. 1985, 57, 834.

10. Terabe, S.; Ozaki, K.; Otsuka, K.; Ando, T. J. Chromatogr. 1985, 332, 211. 11. Haynes, J. L.; Shamsi, S. A.;Warner, I. M. Anal. Chem. 1999, 18, 317.

12. Pretorius, V.; Hopkins, B. J., Schieke, J. D. J. Chromatogr. 1974, 99, 23.

Page 75: Analytical separations using packed and open-tubular

55

13. Huang, X.; Zhang, J.; Horvath, C. J. Chromatogr. A 1999, 858, 91. 14. Pesek, J. J.; Matyska, M. T. J. Chromatogr. A 1996, 736, 255. 15. Malik, A. Electrophoresis 2002, 23, 3973. 16. Knox, J. H.; Grant, I. H. Chromatographia 1991, 32, 317. 17. Pfeffer, W. D.; Yeung, E. E. Anal. Chem. 1990, 62, 2178. 18. Pfeffer, W. D.; Yeung, E. S. J. Chromatogr. 1991, 557, 125. 19. Hindocha, D.; Smith, N. W. J. Chromatogr. A 2000, 904, 99. 20. Chen, Z.; Hobo, T. Electrophoresis 2001, 22, 3339. 21. Liu, Z.; Otsuka, K.; Terabe, S.; Motokawa, M.; Tanaka, N. Electrophoresis 2002,

23, 2973. 22. Tsuda, T.; Matsuki, S.; Munesue, M.; Kitagawa, S., Oehi, H. J. Liq Chromatogr.

& Rel. Techn. 2003, 26, 697. 23. Chen, X.; Zou, H.; Ye, M.; Zhang, Z. Electrophoresis 2002, 23, 1246. 24. Deng, Y.; Zhang, J.; Tsuda, T.; Yu, P. H.; Boulton, A. A.; Cassidy, R. M. Anal.

Chem. 1998, 70, 1586. 25. Henry, C. W.; McCarroll, M. E.; Warner, I. M. J. Chromatogr. A 2001, 905, 319. 26. Thiam, S.; Shamsi, S. A.; Henry, C. W.; Robinson, J. W.; Warner, I. M. Anal.

Chem. 2000, 72, 2541. 27. Colon, L. A.; Guo, Y.; Fermier, A. Anal. Chem. 1997, 69, 461A. 28. Liu, Z.; Wu, R.; Zou, H. Electrophoresis 2002, 23, 3954. 29. Wu, J. T.; Huang, P.; Li, M. X.; Qian, M. G.; Lubman, D. M. Anal. Chem. 1997,

69, 320. 30. Zou, H.; Ye, M. Electrophoresis 2000, 21, 4073. 31. Lammehofer, M.; Lindner, W. J. Chromatogr. A 1999, 839, 167. 32. Lammehofer, M.; Lindner, W. J. Chromatogr. A 1998, 829, 115.

Page 76: Analytical separations using packed and open-tubular

56

33. Wu, J. T.; Huang, P.; Li, M. X.; Lubman, D. M. Anal. Chem. 1997, 69, 2908. 34. Ye, M.; Zou, H.; Liu, Z.; Ni, J.; Zhang, Y. Science in China B 1999, 42, 639. 35. Walhagen, K.; Unger, K. K.; Olsson, A. M.; Hearn, M. T. W. J. Chromatogr. A

1999, 853, 263. 36. Zou, H.; Ye, M. Electrophoresis 2000, 21, 4073. 37. Paces, M.; Kosek, J.; Marek, M.; Tallarek, U.; Seidel-Morgenstern, A.

Electrophoresis 2003, 24, 380. 38. Wistuba, D.; Schurig, V. Electrophoresis 2000, 21, 3152. 39. Pesek, J. J.; Matyska, M. T.; Sentellas, S.; Galceran, M. T.; Chirai, M.; Pirri, G.

Electrophoresis 2002, 23, 2982. 40. Constantin, S.; Freitag, R. J. Chromatogr. A 2000, 887, 253. 41. Matyska, M. T.; Pesek, J. J.; Katrekar, A. Anal. Chem. 1999, 71, 5508. 42. Liu, Z.; Otsuka, K.; Terabe, S. J. Sep. Sci. 2001, 24, 17. 43. Jinno, K.; Sawada, H.; Trend in Anal. Chem. 2000, 19, 664. 44. Wistuba, D.; Schurig, V. J. Chromatogr. A 2000, 875, 255. 45. Tobler, E.; Lammerhofer, M.; Lindner, W. J. Chromatogr. A 2000, 875, 341. 46. Ye, M.; Zou, H.; Lei, Z.; Wu, R.; Liu, Z.; Ni, J. Electrophoresis 2001, 22, 518. 47. Tegeler, T.; Rassi, Z. E. Electrophoresis 2002, 23, 1217. 48. Huber, C. G.; Choudhary, G.; Horvath, C. Anal. Chem. 1997, 69, 4429. 49. Bailey, C. G.; Yan, C. Anal. Chem. 1998, 70, 3275. 50. Pirogov, A. V.; Buchberger, W. J. Chromatogr. A 2001, 916, 51. 51. Lim, L. T.; Zare, R. N.; Bailey, C. G.; Rakestraw, D. J.; Yan, C. Electrophoresis

2000, 21, 737. 52. Wen, E.; Asiaie, R.; Horvath, Cs. J. Chromatogr. A 1999, 855, 349.

53. Charvatova, J.; Deyl, Z.; Kasicka, V.; Kral, V. J. Chromatogr. A 2003, 990, 159.

Page 77: Analytical separations using packed and open-tubular

57

54. Fujimoto, C. Electrophoresis 2002, 23, 2929. 55. Francotte, E.; Jung, M. Chromatographia 1996, 42, 521. 56. Charvatova, J.; Kral, V.; Kasicka, V. Collection Symposium Series 2001, 4, 109. 57. Pesek, J. J.; Matyska, M. T. J. Capillary Electrophoresis 1997, 4, 213. 58. Rehder, M. A.; McGown, L. B. Electrophoresis 2001, 22, 3759. 59. Tan, Z. J.; Remcho, V. T. Anal. Chem. 1997, 69, 581. 60. Wakita, T.; Chankvetadze, B.;Yamamoto, C.; Okamoto, Y. J. Sep. Sci. 2002, 25,

167. 61. Charvatova, J.; Matejka, P.; Kral, V.; Deyl, Z. J. Chromatogr. A 2001, 921, 99. 62. Sun, P.; Landman, A.; Barker, G. E.; Hartwick, R. A. J. Chromatogr. A 1994,

685, 303. 63. Mingxian, H.; Guoliang, Y.; Bradshow, J. S.; Milton, L. J. Microcol. Sep. 1993,

5, 199. 64. Guan, N.; Zeng, Z.; Wang, Y.; Fu, E.; Cheng, J. Anal. Chim. Acta 2000, 418, 145. 65. Kato, M.; Dulay, M. T.; Bennett, B.; Chen, J. R.; Zare, R. N. Electrophoresis

2000, 21, 3145. 66. Fujimoto, C.; Fujise, Y. Anal. Chem. 1996, 68, 2753. 67. Kornysova, O.; Owens, P. K.; Maruska, A. Electrophoresis 2001, 22, 3335. 68. Koide, T.; Ueno, K. J. Chromatogr. A 2000, 893, 177.

69. Gusev, I.; Huang, X.; Horvath, C. J. Chromatogr. A 1999, 855, 273. 70. Kang, J.; Wistuba, D.; Schurig, V. Electrophoresis 2002, 23, 1116. 71. Vegvari, A.; Foldesi, A.; Hetenyi, C.; Kocnegarova, O.; Schmid, M. G.;

Kudirkaite, V.; Hjerten, S. Electrophoresis 2000, 21, 3116. 72. Schmid, M. G.; Grobuschek, N.; Tuscher, C.; Gubitz, G.; Vegvari, A.;

Machtejevas, E.; Maruska, A.; Hjerten, S. Electrophoresis 2000, 21, 3141.

73. Chen, Z.; Hobo, T. Anal. Chem. 2001, 73, 3348.

Page 78: Analytical separations using packed and open-tubular

58

74. Lammerhofer, M.; Peters, E. C.; Yu, C.; Svec, F.; Frechet, J. M. J.; Lindner, W. Anal. Chem. 2000, 72, 4614.

75. Lammerhofer, M.; Svec, F.; Frechet, J. M. J.; Lindner, W. Anal. Chem. 2000, 72,

4623. 76. Koide, T.; Ueno, K. Anal. Sci. 2000, 16, 1065. 77. Koide, T.; Ueno, K. J. Chromatogr. A 2001, 909, 305. 78. Peters, E. C.; Petro, M.; Svec, F.; Frechet, J. M. J. Anal. Chem. 1997, 69, 3646. 79. Peters, E. C.; Petro, M.; Svec, F.; Frechet, J. M. J. Anal. Chem. 1998, 70, 2288. 80. Peters, E. C.; Petro, M.; Svec, F.; Frechet, J. M. J. Anal. Chem. 1998, 70, 2296. 81. Que, A. H.; Novothy, M. V. Anal. Chem. 2002, 74, 5184. 82. Wu, R.; Zou, H.; Ye, M.; Lei, Z.; Ni, J. Electrophoresis 2001, 22, 544. 83. Wu, R.; Zou, H.; Fu, H.; Jin, W.; Ye, M. Electrophoresis 2002, 23, 1239. 84. Ping, G.; Schmitt-Kopplin, P.; Hertkorn, N.; Zhang, W.; Zhang, Y.; Kettrup, A.

Electrophoresis 2003, 24, 958. 85. Legido-Quigley, C.; Marlin, N. D.; Melin, V.; Manz, A.; Smith, N. W.

Electrophoresis 2003, 24, 917. 86. Baltussen, E.; van Dedem, G. W. K. Electrophoresis 2002, 23, 1224. 87. Smith, N. W.; Carter-Finch, A. S. J. Chromatogr. A 2000, 892, 219. 88. Pesek, J. J.; Matyska, M. T. J. Chromatogr. A 2000, 887, 31.

89. Matyska, M. T.; Pesek, J. J.; Boysen, R. I.; Heam, M. T. W. Anal. Chem. 2001,

73, 5116. 90. Pesek, J. J.; Matyska, M. T.; Tran, H. J. Sep. Sci. 2001, 24, 729.

91. Jinno, K.; Sawada, H.; Catabay, A. P.; Watanabe, H.; Sabli, N. B. H.; Pesek, J. J.;

Matyska, M. T. J. Chromatogr. A 2000, 887, 479.

92. Kapnissi, C. P.; Akbay, C.; Schlenoff, J. B.; Warner, I. M. Anal. Chem. 2002, 74, 2328.

Page 79: Analytical separations using packed and open-tubular

59

93. Kamande, M. W.; Kapnissi, C. P.; Zhu, X.; Akbay, C.; Warner, I. M. Electrophoresis 2003, 24, 945.

94. Xie, M. J.; Feng, Y. Q.; Da, S. L.; Meng, D. Y.; Ren, L. W. Anal. Chim. Acta

2001, 428, 255. 95. Garner, T. W.; Yeung, E. S. J. Chromatogr. 1993, 640, 397. 96. Breadmore, M. C.; Macka, M.; Avdalovic, N.; Haddad, P. R. Anal. Chem. 2001,

73, 820. 97. Breadmore, M. C.; Palmer, A. S.; Curran, M.; Macka, M.; Avdalovic, N.;

Haddad, P. R. Anal. Chem. 2002, 74, 2112. 98. Breadmore, M. C.; Macka, M.; Haddad, P. R.; Avdalovic, N. Analyst 2000, 125,

1235. 99. Boyce, M. C.; Breadmore, M.; Macka, M.; Doble, P.; Haddad, P. R.

Electrophoresis 2000, 21, 3073. 100. Baryla, N. E.; Melanson, J. E.; McDermott, M. T.; Lucy, C. A. Anal. Chem.

2001, 73, 4558.

101. Chiari, M.; Ceriotti, L.; Crini, G.; Morcellet, M. J. Chromatogr. A 1999, 836, 81.

102. Charvatova, J.; Kasicka, V.; Deyl, Z.; Kral, V. J. Chromatogr. A 2003, 990, 111.

103. Breadmore, M. C.; Boyce, M.; Macka, M.; Avdalovic, N.; Haddad, P. R. J. Chromatogr. A 2000, 892, 303.

104. Erim, F. B.; Cifuentes, A.; Poppe, H.; Kraak, J. C. J. Chromatogr. A 1995, 708,

356.

105. Liu, Z.; Zou, H.; Ni, J. Y.; Zhang, Y. Anal. Chim. Acta 1999, 378, 73.

106. Liu, Z.; Zou, H.; Ye, M.; Ni, J.; Zhang, Y. Chinese J. Chromatogr. 1999, 17, 245.

107. Liu, Z.; Zou, H.; Ye, M.; Ni, J.; Zhang, Y. Electrophoresis 1999, 20, 2891.

108. Cifuentes, A.; Poppe, H.; Kraak, J. C.; Erime, F. B. J. Chromatogr. B 1996, 681, 21.

109. Gilges, M.; Kleemiss, M. H.; Schomburg, G. Anal. Chem. 1994, 66, 2038.

110. Cordova, E.; Gao, J.; Whitesides, G. M. Anal. Chem. 1997, 69, 1370.

Page 80: Analytical separations using packed and open-tubular

60

111. Chiu, R. W.; Jimenez, J. C.; Monnig, C. A. Anal. Chim. Acta 1995, 307, 193.

112. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 2254.

113. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 5272.

114. Graul, T. W.; Schlenoff, J. B. Anal. Chem. 1999, 71, 4007.

115. Sawada, H.; Jinno, K. Electrophoresis 1999, 20, 24.

116. Bendahl, L.; Hansen, S. H.; Gammelgaard, B. Electrophoresis 2001, 22, 2565.

117. Kohr, J.; Engelhardt, H. J. Microcol. Sep. 1991, 3, 491.

118. Li, M. X.; Liu, L.; Wu, J. T.; Lubman, D. M. Anal. Chem. 1997, 69, 2451.

119. Harrell, C. W.; Dey, J.; Shamsi, S. A.; Foley, J. P.; Warner, I. M. Electrophoresis 1998, 19, 712.

120. Wang, J.; Warner, I. M. Anal. Chem. 1994, 66, 3773.

121. Kapnissi, C. P.; Valle, B. C.; Warner, I. M. Anal. Chem. 2003, 75, 6097.

122. Yarabe, H. H.; Billiot, E.; Warner, I. M. J. Chromatogr. A 2000, 875, 179.

123. Palmer, C. P.; Tanaka, N. J. Chromatogr. A 1997, 792, 105.

124. Palmer, C. P.; Terabe, S. Anal. Chem. 1997, 69, 1852.

125. Wang, J.; Warner, I. M. J. Chromatogr. A 1995, 711, 297.

126. Shamsi, S. A.; Valle, B. C.; Billiot, F.; Warner, I. M. Anal. Chem. 2003, 75, 379.

127. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626.

128. Schlenoff, J. B.; Dubas, S. T.; Farhat, T. Langmuir 2000, 16, 9968.

129. Laurent, D.; Schlenoff, J. B. Langmuir 1997, 13, 1552.

130. Rmaile, H. H.; Schlenoff, J. B. Langmuir 2002, 18, 8263.

131. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153.

132. Castelnovo, M.; Joanny, J.-F. Langmuir 2000, 16, 7524.

Page 81: Analytical separations using packed and open-tubular

61

133. Huang, X.; Horvath, C. J. Chromatogr. A 1997, 788, 155.

134. Cifuentes, A.; de Frutos, M.; Santos, J. M.; Diez-Masa, J. C. J. Chromatogr. 1993, 655, 63.

135. Schmalzing, D.; Piggee, C. A.; Foret, F.; Carrilho, E.; Karger, B. L. J.

Chromatogr. 1993, 652, 149.

136. Towns, J. K.; Regnier, F. E. J. Chromatogr. 1990, 516, 69.

137. Bao, J. J. J. Liq. Chromatogr. & Rel. Tecnol. 2000, 23, 61.

138. Towns, J. K.; Bao, J. M.; Regnier, F. E. J. Chromatogr. 1992, 599, 227.

139. Zeng, Z.; Xie, C.; Li, H.; Han, H.; Chen, Y. Electrophoresis 2002, 23, 1272.

140. Wu, X.; Liu, H.; Liu, H.; Zhang, S.; Haddad, P. R. Anal. Chim. Acta 2003, 478, 191.

141. Li, H. B.; Zeng, Z. R.; Xie, C. H.; Chen, Y. Y. Chromatographia 2002, 55, 591.

142. Wang, Z.; Chen, Y.; Yuan, H.; Huang, Z.; Liu, G. Electrophoresis 2000, 21,

1620.

143. Crego, A. L.; Martinez, J.; Marina, M. L. J. Chromatogr. A 2000, 869, 329.

144. Bruin, G. J. M.; Tock, P. P. H.; Kraak, J. C.; Poppe, H. J. Chromatogr. 1990, 517, 557.

145. Guo, Y.; Colon, L. A. Anal. Chem. 1995, 67, 2511.

146. Dube, S.; Smith, R. M. Chromatographia 2003, 57, 485.

147. Onuska, F. I.; Comba, M. E.; Bistricki, T.; Wilkinson, R. J. J. Chromatogr. 1977,

142, 117.

148. Pesek, J. J.; Matyska, M. T.; Mauskar, L. J. Chromatogr. A 1997, 763, 307.

149. Catabay, A. P.; Sawada, H.; Jinno, K.; Pesek, J. J.; Matyska, M. T. J. Cap. Elec. 1998, 005:1/2, 89.

150. Pesek, J. J.; Matyska, M. T.; Cho, S. J. Chromatogr. A 1999, 845, 237.

151. Pesek, J. J.; Matyska, M. T.; Swedberg, S.; Udivar, S. Electrophoresis 1999, 20,

2343.

Page 82: Analytical separations using packed and open-tubular

62

152. Matyska, M. T.; Pesek, J. J.; Yang, L. J. Chromatogr. A 2000, 887, 497.

153. Matyska, M. T.; Pesek, J. J.; Sandoval, J. E.; Parkar, U.; Liu, X. J. Liq. Chrom. & Rel. Technol. 2000, 23, 97.

154. Matyska, M. T.; Pesek, J. J.; Boysen, R. I.; Hearn, M. T. W. J. Chromatogr. A

2001, 924, 211.

155. Matyska, M. T.; Pesek, J. J.; Boysen, I.; Hearn, T. W. Electrophoresis 2001, 22, 2620.

156. Pesek, J. J.; Matyska, M. T. J. Chromatogr. A 1996, 736, 313.

157. Narang, P.; Colon, L. A. J. Chromatogr. A 1997, 773, 65.

158. Hayes, J. D.; Malik, A. Anal. Chem. 2001, 73, 987.

159. Zhao, Y.; Zhao, R.; Shangguan, D.; Liu, G. Electrophoresis 2002, 23, 2990.

160. Rodriguez, S. A.; Colon, L. A. Anal. Chim. Acta 1999, 397, 207.

161. Wang, Y. C.; Zeng, Z. R.; Xie, C. H.; Guan, N.; Fu, E. Q.; Cheng, J. K.

Chromatographia 2001, 54, 475.

162. Rodriguez, S. A.; Colon, L. A. Applied Spectroscopy 2001, 55, 472.

163. Crosnier de Bellaistre, M.; Mathieu, O.; Randon, J.; Rocca, J. L. J. Chromatogr. A 2002, 971, 199.

164. Wang, Y.; Zeng, Z.; Guan, N.; Cheng, J. Electrophoresis 2001, 22, 2167.

165. Jacques, J.; Collet, A.; Wilen, S. H. In Enantiomers, Racemates, and Resolutions;

Krieger Publishing Co: Malabar, FL, 1981.

166. Pasteur, L. Ann. Chim. Phys. 1848, 24, 442.

167. Feibush, B.; Grinberg, N. In Chromatographic Chiral Separations – The History of Enantiomeric Resolution, 2nd ed.; Zief, M., Crane, L. J., Eds.; Marcel Dekker, Inc.: New York, NY, 1988.

168. Ahuja, S. In Chiral Separations – Applications and Technology; American

Chemical Society: Washington, DC, 1997.

169. http://chirality.ouvaton.org/homepage.htm

Page 83: Analytical separations using packed and open-tubular

63

170. Easson, E. H.; Stedman, E. J. Biochem. 1933, 27, 1257.

171. Mayer, S.; Schurig, V. J. High Res. Chromatogr. 1992, 15, 129.

172. Mayer, S.; Schurig, V. J. Liq. Chromatogr. 1993, 16, 915.

173. Mayer, S.; Schurig, V. Electrophoresis 1994, 15, 835.

174. Mayer, S.; Schleimer, M.; Schurig, V. J. Microcol. Sep. 1994, 6, 43.

175. Schurig, V.; Jung, M.; Mayer, S.; Fluck, M.; Negura, S.; Jakubetz, H. J. Chromatogr. A 1995, 694, 119.

176. Jakubetz, H.; Czesla, H.; Schurig, V. J. Microcol. Sep. 1997, 9, 421.

177. Armstrong, D. W.; Tang, Y.; Ward, T.; Nichols, M. Anal. Chem. 1993, 65, 1114.

178. Pesek, J. J.; Matyska, M. T.; Menezes, S. J. Chromatogr. A 1999, 853, 151.

179. Yang, J.; Hage, D. S. Anal. Chem. 1994, 66, 2719.

180. Liu, Z.; Otsuka, K.; Terabe, S. Chromatogr. 2000, 21, 302.

181. Liu, Z.; Otsuka, K.; Terabe, S. Electrophoresis 2001, 22, 3791.

182. Liu, Z.; Otsuka, K.; Terabe, S. J. Chromatogr. A 2002, 961, 285.

183. Hofstetter, H.; Hofstetter, O.; Schurig, V. J. Microcol. Sep. 1998, 10, 287.

184. Lin, J. M.; Uchiyama, K.; Hobo, T. Chromatographia 1998, 47, 625.

185. Quaglia, M.; de Lorenzi, E.; Sulitzky, C.; Massolini, G.; Sellergren, B. Analyst

2001, 126, 1495.

186. Bruggemann, O; Freitag, R.; Whitcombe, M. J.; Vulfson, E. N. J. Chromatogr. A 1997, 781, 43.

187. Lin, J. M; Nakagama, T.; Uchiyama, K.; Hobo, T. J. Liq. Chromatogr. & Rel.

Techn. 1997, 20, 1489.

188. Lin, J. M.; Nakagama, T.; Uchiyama, K.; Hobo, T. J. Pharmac. Biomed. Analysis 1997, 15, 1351.

189. Nilsson, S.; Schweitz, L.; Peterson, M. Electrophoresis 1997, 18, 884.

Page 84: Analytical separations using packed and open-tubular

64

190. Schweitz, L.; Spegel, P.; Nilsson, S. Analyst 2000, 125, 1899.

191. Schweitz, L.; Andersson, L. I.; Nilsson, S. Anal. Chem. 1997, 69, 1179.

192. Spegel, P.; Schweitz, L.; Nilsson, S. Electrophoresis 2001, 22, 3833.

193. Tan, Z. J.; Remcho, V. T. Electrophoresis 1998, 19, 2055.

194. Schweitz, L.; Andersson, L. I.; Nilsson, S. Anal. Chim. Acta 2001, 435, 43.

195. Schweitz, L. Anal. Chem. 2002, 74, 1192.

196. de Boer, T.; Mol, R.; de Zeeuw, R. A.; de Jong, G. J.; Sherrington, D. C.; Cormack, P. A. G.; Ensing, K. Electrophoresis 2002, 23, 1296.

197. Quaglia, M.; de Lorenzi, E.; Sulitzky, C.; Caccialanza, G.; Sellergren, B.

Electrophoresis 2003, 24, 952.

198. Higgins, D. A.; Mei, E. In Near-Field Scanning Optical Microscopy; Bonnell, D., Eds.; Wiley-VCH, Inc: New York, NY, 2001, 371-420.

199. http://www.jasco.co.uk/nearfield.htm

200. Sheppard, C. J. R. In Multidimensional Microscopy; Cheng, P. C., Lin, T. H.,

Wu, W. L., Wu, J. L., Eds.; Springer-Verlag New York Inc.: New York, NY, 1994, 1-31.

201. http://dept.kent.edu/projects/cell/FLUORO.HTM

202. Rochow, T. G.; Tucker, P. A. In Introduction to Microscopy by Means of Light,

Electrons, X rays, or Acoustics, 2nd ed.; Plenum Press: New York, NY, 1994.

203. Sheppard, C. J. R. In Multidimensional Microscopy; Cheng, P. C., Lin, T. H., Wu, W. L., Wu, J. L., Eds.; Springer-Verlag New York Inc.: New York, NY, 1994, 53-71.

204. Corle, T. R.; Kino, G. S. In Confocal Scanning Optical Microscopy and Related

Imaging Systems; Academic Press: San Diego, CA, 1996.

205. http://www.cs.ubc.ca/spider/ladic/intro.html

Page 85: Analytical separations using packed and open-tubular

65

CHAPTER 2.

SEPARATION OF BENZODIAZEPINES BY USE OF CAPILLARY ELECTROCHROMATOGRAPHY

2.1 Introduction

Benzodiazepines are effective medicinal compounds, which are used primarily for

the treatment of anxiety and sleep disorders [1]. In fact, these drugs are among the most

widely prescribed of all psychoactive drugs. They are important in forensic toxicology

having hypnotic, tranquillizing and anticonvulsant properties. Therefore, they are often

encountered in casework involving road traffic offences or drug overdose [2-4]. They enter

the brain rapidly and work by binding to a specific type of receptor protein, which is

widely distributed in groups of nerve cells involved in anxiety, memory, sedation, and

coordination. In recent years, there has been a growing interest regarding the side effects of

benzodiazepines. These effects include dizziness, adverse interaction with alcohol, and risk

of dependence after long-term use [5]. For these reasons, the analysis of such compounds

is vital to areas such as pharmaceutical analysis, therapeutic drug monitoring, and forensic

toxicology [6-8]. A variety of methods have been developed for the separation of

benzodiazepines, but only some of them have been applied to their identification and

determination in complex matrices such as blood [9-21], urine [22-27] and hair [28, 29].

The methods most frequently used for controlling and monitoring these drugs in

blood or urine are voltametry [30, 31], radioimmunoassay [6, 15], spectrophotometry [32],

supercritical fluid chromatography (SFC) [33], thin layer chromatography (TLC) [25, 34],

gas chromatography (GC) [10, 11, 26-28] and high-performance liquid chromatography

(HPLC) [2, 16-21, 34-38]. Among these techniques, GC and HPLC have been the most

Page 86: Analytical separations using packed and open-tubular

66

popular due to their selectivity. However, GC analysis is complicated and time consuming,

due to the need for derivatization and the thermal instability of some drugs such as

oxazepam [39]. According to the literature, the separation of benzodiazepines is mainly

performed by liquid chromatography (LC) using reversed-phase systems composed of

silica support materials and chemically bonded alkyl chains (octyl or octadecyl) [2, 34]. In

comparison with existing electrokinetic techniques such as capillary zone electrophoresis

(CZE) and micellar electrokinetic chromatography (MEKC), HPLC has low column

efficiency due to the need for pressure-driven flow.

CZE and MEKC are well established for the separation of many classes of

compounds. However, CZE is unable to resolve benzodiazepines, because the majority of

them are neutral and are of similar hydrophobicity. MEKC has been proposed as an

alternative approach. In this approach, micelles are introduced into the background

electrolyte and the separation of some neutral species is achieved [40, 41]. MEKC has been

successfully used to separate benzodiazepines [13, 22-24, 39, 42]. Boonkerd et al. [42]

have studied the migration behavior of a series of benzodiazepines in MEKC using three

kinds of surfactants, i.e., sodium dodecyl sulfate (SDS), dodecyl trimethylammonium

bromide (DoTAB) and bile salts. Renou-Gonnord and David [39] have studied the effects

of SDS, β-cyclodextrin (β-CD), urea, organic solvents and applied voltage on the

separation of nine benzodiazepines using MEKC. Although easy to implement, MEKC

lacks the selectivity and the variety of stationary phases that HPLC offers [43, 44].

Another disadvantage of MEKC is its incompatibility with mass spectrometry detection

due to the high concentrations of surfactants used [45].

Page 87: Analytical separations using packed and open-tubular

67

CEC is considered as an alternative approach to MEKC. It uses a stationary phase

rather than a micellar pseudo-stationary phase. Solutes are separated according to their

partitioning between the mobile and stationary phase and, when charged, their

electrophoretic mobility [46, 47]. The mobile phase in CEC is driven by electroosmotic

flow (EOF) induced by applying an electrical field over the column [45]. The advantages

that CEC offers over HPLC include higher efficiency, the possibility of using small

particles in beds that would create high back pressure in HPLC, and the unique selectivity

due to the superimposition of chromatographic and electrophoretic effects [48].

CEC has been successfully used for the analysis of neutral drugs [49-51]. However,

only a few studies have been performed on the separation of benzodiazepines using CEC

[44, 45, 52, 53]. Cahours et al. [45] have investigated the influence of temperature, ionic

strength and organic modifier content on electrophoretic, chromatographic and separation

performances of five benzodiazepines using CEC on a phenyl-bonded silica column. Jinno

et al. [52] have compared the separation behavior of a series of benzodiazepines in packed

CEC and open-tubular CEC using a cholesteryl silica-bonded phase.

In this chapter, the influence of several experimental parameters is described in

order to obtain improved selectivity and efficiency for the separation of seven

benzodiazepine standards. This is accomplished by use of an octadecyl silica (ODS)

stationary phase in CEC. The optimized method proved to be effective in characterizing

oxazepam in a urine sample.

Page 88: Analytical separations using packed and open-tubular

68

2.2 Experimental

2.2.1 Reagents and Chemicals

Oxazepam, lorazepam, nitrazepam, clonazepam, temazepam, flunitrazepam,

diazepam and tris(hydroxymethyl)aminomethane·hydrochloride (Tris·HCl) were purchased

from Sigma Chemical Company (St Louis, MO, USA). Acetonitrile (ACN), methanol

(MeOH) and tetrahydrofuran (THF) were obtained from Fisher (Springfield, NJ, USA).

Hydrochloric acid was purchased from Mallinckrodt & Baker (Paris, KY, USA).

Polyimide-coated fused-silica capillary columns of 100 µm i.d. × 365 µm o.d. were

obtained from Polymicro Technologies (Phoenix, AZ, USA). The columns were packed

with 3 µm Reliasil CEC C18 stationary phase that was purchased from Column Engineering

(Ontario, CA, USA). The DAT Multi-Drug High Urine Calibrator was donated by Earl K.

Long hospital (Baton Rouge, LA, USA).

2.2.2 Instrumentation and Conditions

The CEC experiments reported here were conducted using an HP3DCE capillary

electrophoresis system (Agilent Technologies, Wilmington, DE, USA). Data were

collected by use of HP3DCE Chemstation software (Agilent Technologies). The dimensions

of the capillary were 48 cm × 100 µm i.d. (40 cm to the detector). All CEC separations

were performed by pressurizing both the inlet and outlet buffer vials with 12 bar of

nitrogen to prevent bubble formation. Unless stated otherwise, the temperature of the

capillary cassette was maintained at 25 °C by the instrument thermostating system.

Detection of the analytes was performed using a photodiode array detection system set to

220 nm. All sample solutions were injected electrokinetically (15 kV for 5 s).

Page 89: Analytical separations using packed and open-tubular

69

2.2.3 Sample and Buffer Preparation

Analytical standard benzodiazepine stock solutions were prepared at concentrations

of 2 mg/ml in MeOH. A 400 µl aliquot of each analyte was mixed and the final

concentration of each benzodiazepine in the test mixture was ∼ 0.3 mg/ml. A buffer

solution of 100 mM Tris was prepared by dissolving the appropriate amount of Tris buffer

in 10 ml of deionized water and the pH was adjusted to 8.0 using 1M HCl. An appropriate

percentage of ACN, MeOH or THF (v/v) was added to an appropriate percentage of the

aqueous Tris buffer solution (v/v), and then, the final volume was adjusted with deionized

water depending on the mobile phase being studied. The final solution was filtered using a

polypropylene nylon filter with 0.45 µm pore size and sonicated for 15 min. Finally, the

urine sample was injected into the CEC capillary without any preparation.

2.2.4 Preparation and Conditioning of Packed Capillary Columns

The CEC columns were packed in our laboratory according to a standard procedure

developed elsewhere [54, 55]. The 3 µm Reliasil CEC C18 silica stationary phase was

slurried in acetone at a concentration of 0.2 g/ml. After sonication, 1 ml of slurry was

injected through a Rheodyne injector connected to a stainless steel reservoir. The injector

was connected to the pump that was a Knauer pneumatic HPLC pump (Berlin, Germany)

and the slurry reservoir was connected to the capillary. The other end of the capillary was

connected to a union containing a 0.5 µm frit. The pump pressure was set to 400 bar. When

the capillary was filled with stationary phase, the pump was turned off and the excess

slurry was removed from the reservoir. The capillary was reconnected and the pump was

set back to 400 bar for two more hours. While the pump was on, the first frit was

fabricated using an electrically heated Nichrome wire. The bottom union from the capillary

Page 90: Analytical separations using packed and open-tubular

70

was removed and the excess stationary phase was flushed out with 200 bar pump pressure.

The last procedure was repeated after the preparation of the second frit. The detection

window was placed adjacent to the outlet frit by burning off the polyimide coating. The

packed capillary column was flushed with the mobile phase for one hour and then installed

in a capillary cartridge. The column was further conditioned by applying both pressure of

12 bar to the inlet side and potential in 5 kV increments for 10 min up to 25 kV. Finally,

both the inlet and the outlet vials were pressurized, and the voltage was set to 30 kV until

the current was stabilized. This procedure was used whenever a new mobile phase was

tested. Between injections, the CEC capillary columns were conditioned for 5 min using 10

kV.

2.3 Results and Discussion

The chemical structures and the numerical designations for each of the seven

benzodiazepines used in this study are shown in Figure 2.1. The influence of operating

parameters, such as nature and amount of organic modifier, buffer electrolyte

concentration, applied voltage, and temperature was studied to optimize the CEC

separation of these benzodiazepines.

2.3.1 Effect of Organic Modifier

Mobile phases comprised of Tris (10 mM, pH 8) modified with 60%, 70% and 45%

of either ACN, MeOH or THF, respectively, were used to study the influence of the

modifier on the separations of benzodiazepines. The organic solvent content was fixed in

order to have mobile phases with the same elution strength. A nomograph, which provides

the interconversion of reversed-phase mobile phases having the same strength, can be

found elsewhere [56, 57]. Vertical lines in this figure intersect mobile phases having the

Page 91: Analytical separations using packed and open-tubular

71

same strength. For example, 70% MeOH has the same strength as 60% ACN or 45% of

THF. The buffer Tris was chosen because its low mobility would more closely match that

of the analytes, when compared to more conventional buffers such as borate and

phosphate. The low mobility of Tris also allows higher concentrations of buffer to be used

without significantly increasing the current. In addition, the relatively high ionic strength

of the buffer leads to sharper and more defined peaks. The ionic strength of each mobile

phase in this study was constant (10 mM). A pH of 8 was chosen because the EOF at this

pH had a greater stability and the analysis time was shorter as compared to lower pH

values.

N

HN

OH

O

Cl N

HN

OH

O

Cl

Cl

N

HN

O2N

O

N

HN

O

O2N

Cl

N

N

OH

O

Cl

H3C

N

NO

O2N

F

H3C

N

N

Cl

OH3C

1. Oxazepam 2. Lorazepam 3. Nitrazepam 4. Clonazepam

5. Temazepam 6. Flunitrazepam 7. Diazepam

Figure 2.1 Structures of the seven benzodiazepine analytes.

Page 92: Analytical separations using packed and open-tubular

72

For a given capillary, the EOF ( EOFµ ) is defined as:

0VtLL td

EOF =µ (2.1)

where dL is the distance from injector to detector, tL is the total capillary length, 0t is the

migration time of the electroosmotic flow marker and V is the applied voltage. The

relative EOF can be monitored by use of the values of 0t , since all the other factors are

constant as is evident from Equation 2.1. In the studies reported here, the values of 0t were

measured with MeOH.

Figure 2.2 demonstrates electrochromatograms for the separation of

benzodiazepines on a C18 stationary phase using Tris-ACN (40:60), Tris-MeOH (30:70)

and Tris-THF (55:45) binary mixtures. A reduction in electroosmotic mobility was

observed from 1.78×10-4 for the ACN-buffer mixture to 4.64×10-5 cm2V-1s-1 for THF–

buffer mixture. When ACN-buffer mixture was used, the total separation time decreased

and the peak efficiencies increased at the expense of lower resolution and retention factor

values. Using Tris-THF (55:45), the total separation was very long and the peak

efficiencies were low. Although the analytes in the mixture were not well resolved using

Tris-ACN (40:60), they eluted in 20 min. Taking into consideration the peak efficiency and

the speed of analysis, the Tris-ACN (40:60) binary mixture was used to further optimize

separation conditions for the benzodiazepines.

The retention mechanism of benzodiazepines on a C18 stationary phase with a

mobile phase of ACN/H2O is based on the differential partitioning of the analytes into the

alkyl-bonded phase. Their retention is determined by hydrophobic interactions between the

C18 stationary phase and the nonpolar moiety of each analyte, and by interactions between

Page 93: Analytical separations using packed and open-tubular

73

the polar mobile phase and the sample molecules. The migration order of benzodiazepines

for the ACN/H2O mobile phase was tr1,2< tr

3,4< tr5< tr

6< tr7. However, the elution order

changed when MeOH and THF were used. The elution order for the MeOH/H2O mobile

phase is tr3,4< tr

2< tr1< tr

6< tr5< tr

7 and tr1,2< tr

5< tr6< tr

3< tr4< tr

7 for the THF/H2O mobile

phase.

2.3.2 Effect of Mobile-Phase Composition

In an attempt to achieve baseline separation, the content of the aqueous buffer (10

mM Tris, pH 8) was increased from 30% to 60%. As depicted in Figure 2.3, both

selectivity and retention time of the analytes increase with decreasing ACN concentration.

A decrease from 70% to 40% ACN increased the selectivity between peaks 6 and 7 from

1.31 to 1.99. The increase in selectivity is due to changes in the partition coefficients as a

result of the increased polarity of the mobile phase. As the mobile phase becomes more

polar, the analytes partition more into the stationary phase, and they are significantly

retained. As a consequence of the latter, the migration times are longer, the resolution is

higher, and the efficiency is lower. A slight increase in electroosmotic mobility, at higher

concentrations of ACN, was also observed from 1.57x10-4 to 3.56x10-4 cm2V-1s-1, probably

due to an increase in the ratio of the dielectric constant to buffer viscosity.

2.3.3 Effect of Applied Voltage

The effect of the applied voltage on the CEC separation of benzodiazepines was

then investigated using a mobile phase of 10 mM Tris (pH 8)-ACN (60:40). As expected,

retention times decreased when a higher voltage was applied. Figure 2.4 demonstrates the

electrochromatograms obtained when 30 kV, 20 kV and 15 kV were applied. At 30 kV, the

analytes elute faster with lower resolution and higher efficiency. Although the resolution

Page 94: Analytical separations using packed and open-tubular

74

between analytes 2 and 3, and 5 and 6 is higher at 15 kV, the total separation time is

longer. Based on these results, 20 kV was applied to further optimize the separation

conditions.

1,2 3,4

5 6 7

3,4 2 1 6 5 7

1,2 5 6 3 4 7

min20 40 60 80 100 120

mAU

0 20 40 60 80

min20 40 60 80 100 120

mAU

0 5

10 15 20 25

min20 40 60 80 100 120

mAU

-20 0

20 40

6 0 % A C N

7 0 % M e O H

4 5 % T H F

Figure 2.2 Effect of the nature of organic modifier on the CEC separation of benzodiazepines. Conditions: C18 stationary phase; 40 cm packed x 100 µm i.d.; electrolyte, 10 mM Tris (pH 8)-ACN (40:60), 10 mM Tris (pH 8)-MeOH (30:70), 10 mM Tris (pH 8)-THF (55:45); applied voltage, 30 kV; electrokinetic injection, 15 kV for 5 s; temperature, 25 °C; UV detection, 220 nm.

Page 95: Analytical separations using packed and open-tubular

75

1,2,3,4 5 6 7 1,2 3,4 7 5 6 1 2 5,6 3 4 7

1 2 3 4

5 6 7

min20 40 60 80

mAU

0 20 40 60

min20 40 60 80

mAU

0 20 40 60 80

min20 40 60 80

mAU

0 10 20 30 40

min20 40 60 80

mAU

0 5

10

A C N = 7 0 %

A C N = 6 0 %

A C N = 5 0 %

A C N = 4 0 %

Figure 2.3 Effect of mobile-phase composition on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the mobile phase composition (ACN-Tris) was varied.

2.3.4 Effect of Tris Concentration

The influence of the ionic strength of the mobile phase was also evaluated using 10

mM, 20 mM, and 30 mM Tris (pH 8)-ACN (60:40). At a constant ACN content, as the

ionic strength increased from 10 mM to 30 mM, the retention time of the analytes

increased (Figure 2.5). In addition, the electroosmotic mobility decreased from 1.76x10-4

to 1.30x10-4 cm2V-1s-1 with increasing Tris concentration due to the interactions of Tris

Page 96: Analytical separations using packed and open-tubular

76

with the silica interface, which reduces the zeta potential. As mentioned in Chapter 1, the

electroosmotic mobility in packed-CEC is given by:

η

εεσµ

EcF

RTro

EOF

2/1

32

= (2.2)

where σ , which is proportional to zeta potential, is the charge density at the surface, oε is

the permittivity of vacuum (8.85 x 10-12 C2N-1µ-2), rε is the dielectric constant of the

mobile phase, R is the gas constant, T is temperature, c is the concentration of the

electrolyte, F is Faraday’s constant, η is the viscosity of the mobile phase, and E is the

electric field strength. According to the above equation, the electroosmotic mobility is

inversely proportional to the square root of the buffer concentration. Resolution also

increases upon increasing ionic strength due to improved stacking during electrokinetic

injection. An increase in ionic strength from 10 mM to 30 mM increased the resolution

between the analytes 5 and 6 from 1.04 to 1.39. Therefore, 30 mM Tris was used to further

optimize the separation.

2.3.5 Effect of Column Temperature

For this component of our study the temperature was varied from 45 °C to 15 °C,

and the binary mixture 30 mM Tris (pH 8)-ACN (60:40) was used as the mobile phase. As

in CE, the electroosmotic mobility increases upon increasing temperature in CEC, due to

the decrease in viscosity of aqueous-organic solvent system. When the temperature

increased from 15 °C to 45 °C, the electroosmotic mobility increased from 1.24x10-4 to

1.82x10-4 cm2V-1s-1. As is also typical for liquid chromatography, retention factors,

retention times, and resolution decrease at higher temperature. The effects of these

Page 97: Analytical separations using packed and open-tubular

77

parameters by temperature variation are shown in Figure 2.6. It is also shown that by

decreasing the temperature to 15 °C all benzodiazepines were baseline resolved.

1 23 4

6 7 5 1 2 3 4 5 6 7 1 2 3 4 5 6 7

min20 40 60 80 100 120

mAU

-2.5 0 2.5 5 7.5 10 12.5

min20 40 60 80 100 120

mAU

-10 0

10 20 30

min20 40 60 80 100 120

mAU

-5 0 5

10 15 20

V = 3 0 k V

V = 2 0 k V

V = 1 5 k V

Figure 2.4 Effect of applied voltage on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the applied voltage was varied; electrolyte, 10 mM Tris (pH 8)-ACN (60:40).

Page 98: Analytical separations using packed and open-tubular

78

12 3 4 5 6 7 1 2 3 4 5 6 7 1 2 3 4 5 6 7

min20 40 60 80 100 120

mAU

-10 0

10 20 30

min20 40 60 80 100 120

mAU

-10 0

10 20 30

min20 40 60 80 100 120

mAU

-10 0

10 20 30 40

1 0 m M T r i s

2 0 m M T r i s

3 0 m M T r i s

Figure 2.5 Effect of Tris concentration on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the Tris buffer concentration was varied; electrolyte, Tris (pH 8)-ACN (60:40); applied voltage, 20 kV.

2.3.6 CEC Separation of Drugs from a Urine Sample

The sample used in this study was a urine calibrator that contained 1000 ng/ml

oxazepam and other drugs, such as acetaminophen, amphetamines, imipramine, morphine,

cocaine, etc. Without any sample preparation the urine sample was injected into the CEC

capillary and the optimum conditions were applied. Figure 2.7a illustrates that the trace

Page 99: Analytical separations using packed and open-tubular

79

amount of oxazepam was able to be detected and separated from the other drugs. The urine

sample was injected at 15 kV for 5 s. We also proved the presence of oxazepam in the

urine sample by spiking with the standard analyte (Figure 2.7b.). For this experiment, the

urine sample was injected at 20 kV for 10 s, and the standard analyte at 15 kV for 5 s.

2,3 1 4 5,6 7

1 2 3 4 5,6 7

1 2 3 4 5 6 7 1 2 3 4 5 6 7

min20 40 60 80 100

mAU

0 20 40

min20 40 60 80 100

mAU

-20 -10 0 10 20

min20 40 60 80 100

mAU

-10 0 10 20 30

min20 40 60 80 100

mAU

-10 0

10 20

T = 4 5° C

T = 3 5° C

T = 2 5° C

T = 1 5° C

Figure 2.6 Effect of column temperature on the CEC separation of benzodiazepines. Separation conditions are the same as in Figure 2.2, except the temperature was varied; electrolyte, 30 mM Tris (pH 8)-ACN (60:40); applied voltage, 20 kV.

Page 100: Analytical separations using packed and open-tubular

80

? ? ?

1

min10 20 30 40 50

mAU

-10 0

10 20 30 40 50 60

(a)

(b) ? 1 ? ?

min0 10 20 30 40 50

mAU

0

200

400

600

800

Figure 2.7 CEC separation of drugs from a urine sample. (a) Without spiking. Conditions: C18 stationary phase; 40 cm packed x 100 µm i.d.; electrolyte, 30 mM Tris (pH8)-ACN (60:40); applied voltage, 20 kV; electrokinetic injection, 15 kV for 5 s; temperature, 15 °C; UV detection, 220 nm. (b) With spiking. Separation conditions are the same as above, except electrokinetic injection (urine sample), 20 kV for 10 s; electrokinetic injection (2 mg/ml standard oxazepam), 15 kV for 5 s.

Page 101: Analytical separations using packed and open-tubular

81

2.4 Conclusion

In this study, we have developed a new CEC method for the baseline resolution of

benzodiazepines. The optimized method proved to be effective in separating and

identifying oxazepam in urine samples that contain various concentrations of other drugs.

We conclude that this CEC method is promising for determining benzodiazepines in

forensic and clinical drug analysis without any sample preparation. We have also been able

to understand how several parameters affect the electrophoretic and separation mechanisms

of hydrophobic analytes on an ODS stationary phase. The volume fraction of ACN and the

temperature are inversely proportional to migration time and resolution. However,

electroosmotic mobility increases upon increasing the volume fraction of ACN or the

temperature. In contrast, the ionic strength of the electrolyte is proportional to the

migration time and resolution, and inversely proportional to electroosmotic mobility.

Finally, higher resolutions were obtained at 15°C when a binary mixture of 30 mM Tris

(pH 8)-ACN (60:40) was used.

2.5 References

1. Linnoila, M. In Molecular Biology to Clinical Practice; Costa, E., Ed.; Raven Press: New York, NY, 1983, 267-278.

2. Gill, R.; Law, B.; Gibbs, J. P. J. Chromatogr. 1986, 356, 37.

3. Sternbach, L. H.; Randall, L. O.; Banziger, R.; Lehr, H. In Drugs affecting central

nervous system; Nutley, N. J., Ed.; Marcel Dekker, Inc.: New York, NY, 1968, 237-264.

4. Sioufi, A.; Dubois, J. P. J. Chromatogr. 1990, 531, 459.

5. Doble, A; Martin, I. L. Trends in Pharmacological Sciences 1992, 13, 76.

6. Drummer, O. H. J. Chromatogr. B 1998, 713, 201.

Page 102: Analytical separations using packed and open-tubular

82

7. Smyth, W. F.; McClean, S. Electrophoresis 1998, 19, 2870.

8. Brettell, T. A.; Inman, K.; Rudin, N.; Saferstein, R. Anal. Chem. 2001, 73, 2735.

9. Tomita, M.; Okuyama, T. J. Chromatogr. B 1996, 678, 331.

10. Douse, J. M. F. J. Chromatogr. 1984, 301, 137.

11. Drouet-Coassolo, C.; Aubert, C.; Coassolo, P.; Cano, J. P. J. Chromatogr. 1989, 487, 295.

12. Evenson, M. A.; Wiktorowicz, J. E. Clin. Chem. 1992, 38, 1847.

13. Imazawa, M.; Hatanaka, Y. Journal of Pharmaceutical and Biomedical Analysis

1997, 15, 1503.

14. De Silva, J. A.; Bekersky, I.; Puglisi, C. V.; Brooks, M. A.; Weinfeld, R. E. Anal. Chem. 1976, 48, 10.

15. Dixon, W. R.; Earley, J.; Postma, E. J. Pharm. Sci. 1975, 64, 937.

16. Lloyd, J. B. F.; Parry, D. A. J. Chromatogr. 1988, 449, 281.

17. Mascher, H.; Nitsche, V.; Schutz, H. J. Chromatogr. 1984, 306, 231.

18. Klockowski, P. M.; Levy, G. J. Chromatogr. 1987, 422, 334.

19. Vree, T. B.; Baars, A. M.; Hekster, Y. A.; Van der Kleijn, E. J. Chromatogr. 1981,

224, 519.

20. Minder, E. I.; Schaubhut, R.; Minder, C. E.; Vonderschmitt, D. J. J. Chromatogr. 1987, 419, 135.

21. Aakerman, K. K.; Jolkkonen, J.; Parviainen, M.; Penttilae, I. Clin. Chem. 1996, 42,

1412.

22. Schafroth, M.; Thormann, W.; Allemann, D. Electrophoresis 1994, 15, 72.

23. Tomita, M.; Okuyama, T.; Sato, S.; Ishizu, H. J. Chromatogr. 1993, 621, 249.

24. Wernly, P.; Thormann, W. Anal. Chem. 1992, 64, 2155.

25. Inoue, T.; Niwaguchi, T. J. Chromatogr. 1985, 339, 163.

26. Mule, S. J.; Casella, G. A. J. Anal. Toxicol. 1989, 13, 179.

Page 103: Analytical separations using packed and open-tubular

83

27. Maurer, H.; Pfleger, K. J. Chromatogr. 1987, 422, 85.

28. Yegles, M.; Mersch, F.; Wennig, R. Forensic Science International 1997, 84, 211.

29. Tagliaro, F.; Smyth, W. F.; Turrina, S.; Deyl, Z.; Marigo, M. Forensic Science International 1995, 70, 93.

30. Brooks, M. A.; D’Arconte, L.; Hackman, M. R.; De Silva, J. A. F. J. Anal. Toxicol.

1997, 1, 179.

31. Brooks, M. A.; Bruno, J. J. B.; De Silva, J. A. F.; Hackman, M. R. Anal. Chim. Acta 1975, 74, 367.

32. Werner, I. A.; Altorfer, H.; Perlia, X. Chromatographia 1990, 30, 255.

33. Smith, R. M.; Sanagi, M. M. J. Chromatogr. 1989, 483, 51.

34. Valko, K.; Olajos, S.; Cserhati, T. J. Chromatogr. 1990, 499, 361.

35. Mura, P.; Piriou, A.; Fraillon, P.; Papet, Y.; Reiss, D. J. Chromatogr. 1987, 416,

303.

36. Guillaume, Y.; Guinchard, C. J. Liq. Chromatogr. 1993, 16, 3457.

37. Guillaume, Y.; Guinchard, C. J. Liq. Chromatogr. 1994, 17, 1443.

38. Choma, I.; Dawidowicz, A. L.; Lodkowski, R. J. Chromatogr. 1992, 600, 109.

39. Renou-Gonnord, M. F.; David, K. J. Chromatogr. A 1996, 735, 249.

40. Nishi, H.; Tsumagari, N.; Kakimoto, T.; Terabe, S. J. Chromatogr. 1989, 465, 331.

41. Terabe, S.; Otsuka, K.; Ichikawa, K.; Tsuchiya, A.; Ando, T. Anal. Chem. 1984, 56, 111.

42. Boonkerd, S.; Detaevernier, M. R.; Vindevogel, J.; Michotte, Y. J. Chromatogr. A

1996, 756, 279.

43. Yan, C.; Dadoo, R.; Zhao, H.; Zare, R. N.; Rakesraw, D. J. Anal. Chem. 1995, 67, 2026.

44. Catabay, A. P.; Sawada, H.; Jinno, K.; Pesek, J. J.; Matyska, M. T. J. Capillary

Electrophor. 1998, 5, 89.

45. Cahours, X.; Morin, P.; Dreux, M. J. Chromatogr. A 1999, 845, 203.

Page 104: Analytical separations using packed and open-tubular

84

46. Thiam, S.; Shamsi, S. A.; Henry, C. W.; Robinson, J. W.; Warner, I. M. Anal. Chem. 2000, 72, 2541.

47. Henry, C. W.; McCarroll, M. E.; Warner, I. M. J. Chromatogr. A 2001, 905, 319.

48. Moffatt, F.; Cooper, P. A.; Jessop, K. M. J. Chromatogr. A 1999, 855, 215.

49. Smith, N. W.; Evans, M. B. Chromatographia 1994, 38, 649.

50. Reilly, J.; Saeed, M. J. Chromatogr. A 1998, 829, 175.

51. Wang, J.; Schaufelberger, D. E.; Guzman, N. A. J. Chromatogr. Sci. 1998, 36, 155.

52. Jinno, K.; Sawada, H.; Catabay, A. P.; Watanabe, H.; Haji Sabli, N. B.; Pesek, J. J.;

Matyska, M. T. J. Chromatogr. A 2000, 887, 479.

53. Matyska, M. T.; Pesek, J. J.; Katrekar, A. Anal. Chem. 1999, 71, 5508.

54. Knox, J. H.; Grant, I. H. Chromatographia 1987, 24, 135.

55. Knox, J. H.; Grant, I. H. Chromatographia 1991, 32, 317.

56. Schoenmakers, P. J.; Billiet, H. A. H.; De Galan, L. J. Chromatogr. 1979, 185, 179.

57. Schoenmakers, P. J.; Billiet, H. A. H.; De Galan, L. J. Chromatogr. 1981, 218, 261.

58. Dittmann, M. M.; Rozing, G. P. J. Chromatogr. A 1996, 744, 63.

Page 105: Analytical separations using packed and open-tubular

85

CHAPTER 3.

ANALYTICAL SEPARATIONS USING POLYMERIC SURFACTANTS IN OPEN-TUBULAR CAPILLARY ELECTROCHROMATOGRAPHY

3.1 Introduction

Capillary electrochromatography (CEC) is a hybrid electroseparation technique that

couples the selectivity of high performance liquid chromatography (HPLC) and the

separation efficiency of capillary electrophoresis (CE) [1-5]. CEC also provides high

resolution, short analysis time, smaller sample and buffer consumption, and efficiencies

five to ten times higher than HPLC. The separation in CEC is based upon the

electrophoretic mobility of the solutes and their partitioning between the stationary and

mobile phases.

In the development of CEC, both packed and open-tubular column configurations

have been reported [1-10]. The packed mode of CEC utilizes a fused-silica capillary with a

typical internal diameter of 50-100 µm. This capillary is packed with a typical HPLC

stationary phase such as an octadecyl silica (ODS) stationary phase [2, 4]. However, there

are several problems that need to be solved in order for packed-CEC to be a viable

alternative to either CE or HPLC. The limitations of conventional CEC include the

necessity to fabricate frits, the tendency that packed capillaries have to form bubbles

around the packing material or at the frit, the difficult packing procedure, and the difficult

separation of basic compounds. The problems mentioned above are discussed in detail in

Chapter 1.

Open-tubular CEC (OT-CEC) is an alternative approach to packed-CEC [9]. None

of the problems mentioned above are likely to be encountered in an open-tubular format. In

Page 106: Analytical separations using packed and open-tubular

86

this CEC format, a stable coating needs to be constructed on the inner walls of the capillary

in order to provide efficient chromatographic separations and reproducible EOF [10]. The

most commonly used approaches to wall coatings for modifying the capillary include: (i)

dynamic coating performed by adding the cationic or neutral modifier to the electrolytes

[11, 12]; (ii) adsorbed cationic modifier on the capillary wall by physical adsorption [13-

16]; and (iii) fixation of the hydrophilic layer by covalent bonding and/or cross linking

[17-22]. Harrell et al. [23] achieved a baseline separation of seven tricyclic antidepressants

by use of a novel nonionic micelle polymer, poly (n-undecyl-α-D-glucopyranoside) as a

dynamic coating. However, dynamic coating is known to cause problems when CE is

coupled to mass spectrometry (CE/MS). In addition, the presence of the nonvolatile buffer

constituents may deteriorate the ionization of the analytes [24, 25]. Although physical

adsorption has a simple and rapid coating procedure and good reproducibility, it has been

shown to have a short lifetime and limited pH range [24, 26]. In contrast, some of the

covalent bonding and/or cross-linking have a long lifetime, but require a more complicated

coating procedure [24, 26]. Obviously, an ideal coating procedure would be one that is

both simple and stable.

In this chapter, an alternative to covalent linking of a polymer to silica beads is

explored. In our approach, the polymeric surfactant poly (sodium N-undecanoyl-L-

glycinate), poly (L-SUG), is used in a simple coating procedure, which involves a layer-by-

layer deposition process [27, 28]. This coating is a polyelectrolyte multilayer (PEM), and it

is constructed in situ by alternating rinses of positively and negatively charged polymers

[29-33]. Via electrostatic forces, a layer of polymer adds to the oppositely charged surface,

reversing the surface charge and priming the film for the addition of the next layer. Such

Page 107: Analytical separations using packed and open-tubular

87

coatings have been found to be robust, and thus, highly resistant to charge and

deterioration during use [24, 26, 32]. The advantages of our PEM coating are two-fold.

First, since the polymeric surfactant is coated electrostatically onto the capillary, less

consumption of the reagent is required. Second, with the polymeric surfactant coated on

the capillary, there is less detection interference with the analyte of interest, which in turn

makes the system more amenable to coupling with mass spectrometry or other detectors

where the polymeric surfactant reagent interferes.

This PEM coating approach used for fabricating columns for use in OT-CEC is

described below. The performance of the modified capillaries as a separation medium is

evaluated by use of seven benzodiazepines as analytes. The coating was found to be

remarkably stable with excellent performance for more than 200 runs.

3.2 Experimental

3.2.1 Apparatus and Conditions

Separations were performed on a Beckman P/ACE MDQ capillary electrophoresis

system with UV detection (Fullerton, CA). The fused-silica capillary, 57 cm (50 cm

effective length) x 50 µm i.d., was purchased from Polymicro Technologies (Phoenix, AZ)

and mounted in a Beckman capillary cartridge. Unless stated otherwise, the cartridge

temperature was maintained at 25 °C by use of liquid coolant. UV detection was performed

at 214 nm and the samples were injected by pressure (0.1 psi; 1 psi=6894.76 Pa) for 1 sec.

3.2.2 Reagents and Chemicals

Flunitrazepam, temazepam, diazepam, oxazepam, lorazepam, clonazepam and

nitrazepam were purchased from Sigma Chemical Company (St Louis, MO). The

structures of the analytes used in this study are shown in Figure 2.1 (Chapter 2). However,

Page 108: Analytical separations using packed and open-tubular

88

the elution order of benzodiazepines changes to flunitrazepam1 <temazepam2 <diazepam3

<oxazepam4 <lorazepam5 <clonazepam6 <nitrazepam7. Sodium phosphate (Na2HPO4 and

NaH2PO4), hydrochloric acid (HCl) and sodium chloride (NaCl) were all obtained from

Fisher Scientific (Fair Lawn, NJ). Poly (diallyldimethylammonium chloride), PDADMAC

(Mw=200,000-350,000) was obtained from Aldrich (Milwankee, WI). Other chemicals,

including L-glycine, undecylenic acid and N-hydroxysuccinimide, were also purchased

from Sigma.

3.2.3 Sample and Buffer Preparation

Analytical standard benzodiazepine stock solutions were prepared in methanol-

water (1:1) at concentrations of about 0.15 mg/ml each. A buffer solution of 50 mM

Na2HPO4 was prepared by dissolving the appropriate amount of Na2HPO4 in 10 ml of

deionized water. The solution was filtered using a polypropylene nylon filter with 0.45 µm

pore size and sonicated for 15 min before use.

3.2.4 Synthesis of Monomeric and Polymeric Surfactant

The surfactant monomer of sodium N-undecenoyl-L-glycinate, mono (L-SUG), was

synthesized from the N-hydroxysuccinimide ester of undecylenic acid according to a

previously reported procedure [34]. A 100 mM sodium salt solution of the monomer was

then polymerized by use of 60Co-γ radiation. After irradiation, the polymer was dialyzed by

use of a 2000 molecular mass cut-off and then lyophilized to obtain the dry product.

Structures of the monomeric and polymeric surfactants are illustrated in Figure 3.1.

3.2.5 Procedure for Polyelectrolyte Multilayer (PEM) Coating

PEM coating was achieved by deposition of the polymer solutions using the rinse

function on the Beckman CE system. Each polymer deposition solution contained 0.5%

Page 109: Analytical separations using packed and open-tubular

89

(w/v) polymer in 0.2 M aqueous NaCl solution. It was observed that the addition of NaCl

to the polymer solution resulted in enhanced thickness for each polyelectrolyte layer [32].

The capillary was conditioned before coating using a 5-min rinse of water in order to

remove any contaminants originating from the capillary drawing process. The column was

then conditioned with 1 M NaOH for 60 min. Pure deionized water was flushed through

the capillary for 15 more min. The first monolayer of polymer (PDADMAC) was

deposited by rinsing the solution of the cationic polymer through the capillary for 20 min

followed by a 5-min water rinse. All other polymer depositions were done with 5-min

rinses followed by 5-min water rinses. A diagrammatic scheme of the PEM-coated

capillary is shown in Figure 3.2. This diagram is not provided to give an actual structural

representation of the bilayer, but only to represent the order of polymer deposition. The

multilayer coatings used for the separation of benzodiazepines and the reproducibility

studies consisted of ten layer pairs (a layer pair is a layer of cationic polymer plus a layer

of anionic polymer; also termed a bilayer). The capillary was then flushed with buffer until

a stable current was achieved. The columns were conditioned with buffer for 2 min

between injections.

3.3 Results and Discussion

3.3.1 Endurance of PEM Coating

An important aspect of this approach, which must be considered, is the lifetime of

the stationary phase. Thus, we examined the stability of our coating by use of the following

procedure. A 10-layer pair multilayer was constructed, and nearly 50 separations were

performed within 5 days. Each separation was done with applied voltages of 15 kV to 30

kV at 25 °C. The electrolyte concentration was 50 mM of Na2HPO4 and a pH range of 9.2-

Page 110: Analytical separations using packed and open-tubular

90

11.0 was used. After these experiments, the capillary was removed from the instrument,

and the tips of the column were placed in water vials for one week. When the capillary was

placed back in the instrument, 50 replicate runs were performed using an applied voltage

of 20 kV, a temperature of 25 °C, and a 50-mM phosphate buffer (pH 9.2). The capillary

was again removed from the instrument and the tips were placed in water vials for an

additional week, after which the column was placed back into the instrument and more

than 100 runs were performed. In these studies, both the applied voltage and the

temperature were varied from 15 kV to 30 kV and from 15 °C to 35 °C, respectively.

Therefore, the aggregate performance of the PEM-coated capillary was evaluated for more

than 200 runs.

CO

NH

C O

Na +O-

CONH

C ONa+O-

CO NHCO

Na+O-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHN

CO-ONa+

CO

HNC

O

Na+O-

CO

HNC

O

Na+O-

CO

NHC

ONa+O-

(a) (b)

Figure 3.1 Structures of the (a) monomeric SUG and (b) polymeric SUG.

Page 111: Analytical separations using packed and open-tubular

91

SiO- SiO- SiO- SiO- SiO- SiO- SiO- SiO- SiO-

N+ N+ N+ N+ N+ N+ N+ N+ N+ N+N+N+ N+

N+N+N+N+N+N+N+N+N+N+ N+N+N+

Capillary wall, silanol groups

Cationic polymer, PDADMAC

Anionic polymer, poly(L-SUG) CO

NH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

COHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHC

O

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO-

CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

COHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHC

O

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO-

CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHC

O

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O

-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO-

CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHC

O

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO- CONH

C O+NaO-

CO NHC O

+NaO-

CO

NHCO

-ONa+

CO

NH

CO

-ONa+

CO

HNC

O-ONa+

C OHN

CO -ONa+

C OHNCO-ONa+

CO

HNC

O

+NaO-

CO

HNC

O+NaO-

CO

NHC

O+NaO-

Figure 3.2 Scheme of the PEM-coated capillary.

3.3.2 Stability of PEM Coating

Another important factor to consider when using PEM-coated capillaries is the

stability of the capillary surface, especially after exposure to solutions with extreme pH

values. To evaluate this parameter, two more phosphate buffers of pH 11.0 and 3.0 were

prepared. First, 30 replicate runs were performed with 50 mM phosphate buffer (pH 9.2).

The 10th run (Figure 3.3a) yielded an electroosmotic mobility of 2.39x10-3 cm2V-1s-1. The

capillary was then flushed with 50 mM phosphate buffer (pH 11.0) for 100 min and 50

mM phosphate buffer (pH 9.2) for 30 min. One of the electropherograms obtained after the

exposure to pH 11.0 (Figure 3.3b) gave an electroosmotic mobility of 2.37x10-3 cm2V-1s-1.

After this, the same procedure was followed with the 50 mM phosphate buffer (pH 3.0).

One of the runs that was performed after the last exposure (Figure 3.3c) yielded an

electroosmotic mobility of 2.37x10-3 cm2V-1s-1. Therefore, the PEM coating was

demonstrated to have extraordinary stability under extreme values of pH.

Page 112: Analytical separations using packed and open-tubular

92

3.3.3 Reproducibilities

Reproducibility of the PEM coating is also an important consideration. The

reproducibilities were evaluated by computing the relative standard deviations (RSDs) [36]

of the EOF, which are reported in Table 3.1. The run-to-run RSD was obtained from 50

consecutive electrophoresis runs; both the day-to-day and capillary-to-capillary RSDs were

obtained by use of five replicate analyses; the week-to-week RSD was obtained by use of

three replicate analyses. All RSDs of the EOF were below 1%, and thus, very good

reproducibilities were observed.

3.3.4 Voltage Study

The PEM-coated capillary was applied to the separation of seven benzodiazepines.

The influence of the applied voltage on the efficiency, resolution, and analysis time of the

benzodiazepines was evaluated using a mobile phase of 50 mM Na2HPO4 at 25 °C. As

expected, a higher voltage decreased the retention times. At 30 kV, the analytes eluted

faster with higher efficiency and lower resolution. In contrast, at 15 kV, the migration

times were longer, the resolution was higher, and the efficiency was lower. However, an

applied voltage of 15 kV did not have a major impact on analyte resolution, compared to

the electropherogram obtained when a 20 kV voltage was applied (Figure 3.4).

3.3.5 Temperature Study

The effect of temperature on the separation of benzodiazepines was also studied.

The temperature for this study was varied from 35 °C to 15 °C. As shown in Figure 3.5, the

retention time decreased at higher temperature, and peak efficiency decreased at lower

temperature. In addition, electroosmotic mobility increased when temperature increased,

likely due to a decrease in electrolyte viscosity.

Page 113: Analytical separations using packed and open-tubular

93

a

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

0.008

0.010

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

0.008

0.010

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

0.008

0.010

(a) run 10

(b) after exposure to pH 11.0 (run 40)

(c) after exposure to pH 3.0 (run 110)

123 4 5 6 7

Figure 3.3 Stability studies of PEM coating. Conditions: 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SUG) with 0.2 M NaCl; pressure injection, 0.1 psi for 1 s; electrolyte, 50 mM Na2HPO4 (pH 9.2); applied voltage, 20 kV; temperature, 25 °C; capillary, 57 cm (50 cm effective length) x 50 µm i.d.; detection, 214 nm.

Page 114: Analytical separations using packed and open-tubular

94

Table 3.1 Reproducibilities of PEM capillary coating. Conditions: same as Figure 3.3.

RSD (%)EOF Average (min)

0.789.990run-to-run

day-to-day

capillary-to-capillary

week-to-week

9.982

9.930

9.950

0.81

0.86

0.93

RSD (%)EOF Average (min)

0.789.990run-to-run

day-to-day

capillary-to-capillary

week-to-week

9.982

9.930

9.950

0.81

0.86

0.93

3.3.6 Comparison Between Monomeric and Polymeric Surfactants

Another important consideration for this study is whether molecular micelles are

needed to form an effective and stable PEM, or can the same be achieved by use of

monomeric surfactants. In an effort to compare the chromatographic performance of mono

(L-SUG) and poly (L-SUG) for the separation of hydrophobic analytes, benzodiazepines

were used as test solutes. The separation of benzodiazepines using 50 mM Na2HPO4 (pH

9.2) as the electrolyte and 0.5% (w/v) poly (L-SUG) as the anionic polyelectrolyte for the

construction of the PEM coating is shown in Figure 3.6a. Figure 3.6b is the

electropherogram of the benzodiazepines under the same conditions as in Figure 3.6a.

However, the anionic surfactant used for PEM coating construction in this figure is the

monomeric (nonpolymerized) surfactant at the concentration of 0.5% (w/v) (19 mM).

Almost no separation is noted, even though the monomeric surfactant concentration is

significantly above the normal CMC of the nonpolymerized surfactant (7 mM). Thus, it is

Page 115: Analytical separations using packed and open-tubular

95

clear from this study that the molecular micelle allows better discrimination of the

hydrophobic analytes than the conventional micelle.

In a normal (nonpolymerized) micellar system, the dynamic equilibrium that exists

between the monomers and micellar aggregates, has been demonstrated to be a

disadvantage for separations.33 This dynamic equilibrium will likely reduce the stability of

the PEM coating in OT-CEC. In such a case, poor analyte separation will be observed as

seen here. In contrast, polymeric micelles do not have such problems because the covalent

bonds formed between monomers eliminate dynamic equilibrium. Thus, the coating that is

produced will be more stable as is observed in this study.

3.4 Conclusion

A stable modified capillary has been developed by use of a simple PEM coating

procedure employing a molecular micelle. Excellent run-to-run, day-to-day, week-to-week

and capillary-to-capillary reproducibilities in separation were observed since the RSD

values of electroosmotic flow were below 1% in all cases. In addition, the PEM-coated

capillaries exhibited high stability against extreme pH values. The stability of the

capillaries allowed us to perform over 200 runs. This approach also shows that highly

efficient and reproducible peaks can be obtained from stationary phases prepared by such a

simple procedure. In addition, the chromatographic performance of the monomeric form of

the molecular micelle was compared for the separation of benzodiazepines. This study

confirmed that the polymeric surfactant allows better discrimination of hydrophobic

analytes than the nonpolymerized surfactant. We conclude that this method is a promising

alternative to conventional CEC and should prove very useful in both chiral and achiral

separations.

Page 116: Analytical separations using packed and open-tubular

96

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

0.008

0.010

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

0.008

0.010

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

0.008

0.010

30 kV

15 kV

20 kV

123 4 5 6 7

Figure 3.4 Effect of applied voltage on the OT-CEC separation of benzodiazepines. Conditions: same as Figure 3.3, except applied voltage was varied.

Page 117: Analytical separations using packed and open-tubular

97

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.002

0.004

0.006

0.008

0.010

Minutes0 2 4 6 8 10 12 14 16 18 20 22 24 26 28 30 32 34

AU

0.000

0.001

0.002

0.003

0.004

35°°°°C

25°°°°C

15°°°°C

123 4 5 6 7

Figure 3.5 Effect of temperature on the OT-CEC separation of benzodiazepines. Conditions: same as Figure 3.3, except the temperature was varied.

Page 118: Analytical separations using packed and open-tubular

98

Minutes0 2 4 6 8 10 12 14 16 18 20

AU

0.000

0.002

0.004

0.006

0.008

0.010

Minutes0 2 4 6 8 10 12 14 16 18 20

AU

0.000

0.002

0.004

0.006

0.008

0.010

(a) poly(L-SUG)

(b) mono(L-SUG)

123 4 5 6 7

Figure 3.6 Comparison between monomeric and polymeric surfactants for OT-CEC separation of benzodiazepines. Conditions: same as Figure 3.3, except applied voltage, 30 kV. (a) 0.5% (w/v) poly (L-SUG) (b) 0.5% (w/v) mono (L-SUG).

3.5 References

1. Liu, Z.; Zou, H.; Ye, M.; Ni, J.; Zhang, Y. Electrophoresis 1999, 20, 2891.

2. Henry, C. W.; McCarroll, M. E.; Warner, I. M. J. Chromatogr. A 2001, 905, 319.

3. Ye, M.; Zou, H.; Liu, Z.; Ni, J.; Zhang, Y. Anal. Chem. 2000, 72, 616.

4. Thiam, S.; Shamsi, S. A.; Henry, C. W.; Robinson, J. W.; Warner, I. M. Anal. Chem. 2000, 72, 2541.

5. Jorgenson, J. W.; Lukacs, K. D. Anal. Chem. 1981, 53, 1298.

Page 119: Analytical separations using packed and open-tubular

99

6. Jinno, K.; Sawada, H.; Catabay, A. P.; Hiroshi, W.; Sabli, N. B. H.; Pesek, J. J.; Matyska, M. T. J. Chromatogr. A 2000, 887, 479.

7. We, J.; Huang, P.; Li, M. X.; Qian, M. G.; Lubman, D. M. Anal. Chem. 1997, 69,

320.

8. Dulay, M. T.; Quirino, J. P.; Bennett, B. D.; Kato, M.; Zare, R. N. Anal. Chem. 2001, 73, 3921.

9. Matyska, M. T.; Pesek, J. J.; Katrekar, A. Anal. Chem. 1999, 71, 5508.

10. Hayes, J. D.; Malik, A. Anal. Chem. 2001, 73, 987.

11. Gilges, M.; Kleemiss, M. H.; Schomburg, G. Anal. Chem. 1994, 66, 2038.

12. Cifuentes, A.; Poppe, H.; Kraak, J. C.; Erim, F. B. J. Chromatogr. B 1996, 681, 21.

13. Erim, F. B.; Cifuentes, A.; Poppe, H.; Kraak, J. C. J. Chromatogr. A 1995, 708,

356.

14. Chiu, R. W.; Jimenez, J. C.; Monnig, C. A. Anal. Chim. Acta 1995, 307, 193.

15. Preisler, J.; Yeung, E. S. Anal. Chem. 1996, 68, 2885.

16. Cordova, E.; Gao, J.; Whitesides, G. M. Anal. Chem. 1997, 69, 1370.

17. Towns, J. K.; Regnier, F. E. J. Chromatogr. 1990, 516, 69.

18. Figeys, D.; Aebersold, R. J. Chromatogr. B 1997, 695, 163.

19. Bruin, G. J. M.; Chang, J. P.; Kuhlman, R. H.; Zegers, K.; Kraak, J. C.; Poppe, H. J. Chromatogr. 1989, 471, 429.

20. Hjerten, S.; Johansson, M. K. J. Chromatogr. 1991, 550, 811.

21. Cobb, K. A.; Dolnik, V.; Novotony, M. Anal. Chem. 1990, 62, 2478.

22. Huang, X.; Horvath, C. J. Chromatogr. A 1997, 788, 155.

23. Harrell, C. W.; Dey, J.; Shamsi, S. A.; Foley, J. P.; Warner, I. M. Electrophoresis

1998, 19, 712.

24. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 5272.

25. Niessen, W. M. A.; Tjaden, U. R.; Geef, J. J. Chromatogr. 1993, 636, 3.

Page 120: Analytical separations using packed and open-tubular

100

26. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 2254.

27. Decher, G.; Schmitt, J. Prog. Colloid Polym. Sci. 1992, 89, 160.

28. Decher, G. Science 1997, 277, 1232.

29. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626.

30. Laurent, D.; Schlenoff, J. B. Langmuir 1997, 13, 1552.

31. Schlenoff, J. B.; Li, M. Ber. Bunsen-Ges. Phys. Chem. 1996, 100, 943.

32. Graul, T. W.; Schlenoff, J. B. Anal. Chem. 1999, 71, 4007.

33. Schlenoff, J. B.; Dubas, S. T.; Farhat, T. Langmuir 2000, 16, 9968.

34. Wang, J.; Warner, I. M. Anal. Chem. 1994, 66, 3773.

35. Dittmann, M. M.; Rozing, G. P. J. Chromatogr. A 1996, 744, 63.

36. Skoog, D. A.; Holler, J. F.; Nieman, T. A. In Principles of Instrumental Analysis, 5th ed.; Sherman, M., Bortel, J., Messina, F., Eds.; Harcourt Brace College Publishers: Orlando, FL, 1998.

Page 121: Analytical separations using packed and open-tubular

101

CHAPTER 4.

CHIRAL SEPARATIONS USING POLYMERIC SURFACTANTS AND POLYELECTROLYTE MULTILAYERS IN OPEN-TUBULAR CAPILLARY

ELECTROCHROMATOGRAPHY

4.1 Introduction

The separation of chiral compounds has been of great importance in many

industries, particularly the pharmaceutical industry. This interest is due to the different

pharmakokinetic characteristics and pharmacological activities of each enantiomer in a

racemic drug [1, 2]. Thus, there has been a great demand for the development of analytical

techniques for the separation of chiral bioorganic molecules.

In recent years, different modes of capillary electrophoresis (CE) have proven to be

very efficient methods for the separation and analysis of enantiomeric compounds [3-5].

Micellar electrokinetic chromatography (MEKC), which is one of the most recent methods

employed for chiral separations, uses a pseudostationary phase that involves the addition of

a chiral micellar selector in the background electrolyte (BGE).

In 1994, Wang and Warner reported the use of a synthetic chiral polymeric

surfactant added to the BGE for chiral separations. These polymeric amino acid-based

surfactants offer several advantages over conventional micelles [6-9]. As stated in Chapter

1, the polymerization of the surfactant eliminates the dynamic equilibrium between

monomers and micelles. This, in turn, results in better chiral recognition and enantiomeric

resolution. In addition, the polymeric surfactant can be used at very low concentrations

since it does not depend on the critical micelle concentration. This usually results in higher

efficiencies and rapid analysis in CE.

Page 122: Analytical separations using packed and open-tubular

102

Recently, the use of polymeric surfactants with two amino acid functional groups

(dipeptides) at the polar head region has attracted considerable attention [10]. Various

polymeric dipeptide chiral surfactants have been synthesized and used as chiral

discriminators in MEKC [11, 12]. So far, poly (sodium N-undecanoyl-L-leucyl-valinate),

poly (L-SULV), has shown the best chiral discrimination ability for a wide variety of drug

compounds [8]. One major drawback to using these polymeric surfactants as additives in

BGE is that they are ultimately flushed out, and a fresh BGE is required for each

separation. To circumvent this problem, alternatives to MEKC have been explored [13-19].

Open-tubular capillary electrochromatography (OT-CEC) is an alternative

approach to MEKC. It requires the construction of a stable coating on the inner walls of the

capillary. This, in turn, provides efficient chromatographic separations and reproducible

electroosmotic flow (EOF) [20].

In 1995, Erim et al. [13] investigated the performance of a polyethyleneimine (PEI)

layer as a coating for the separation of basic proteins and peptides. The PEI coating was

constructed by just filling the capillary with a solution containing high-molecular-mass PEI

and flushing the capillary after a certain time. Katayama et al. [14, 15] reported the use of a

new simple coating procedure that is called successive multiple ionic-polymer layer

coating. In their studies, they developed both anion-modified and cation-modified

capillaries for the separation of acidic and basic proteins, respectively. The anion-modified

capillary was prepared by first attaching the cationic polymer to the capillary wall, and

then the anionic polymer to the cationic polymer layer. The cation-modified capillary was

established by attaching the cationic polymer to the anionic polymer layer. In 1999, Graul

and Schlenoff [16] used poly (styrene sulfonate) as the anionic polymer and poly

Page 123: Analytical separations using packed and open-tubular

103

(diallyldimethylammonium chloride) as the cationic polymer in a polyelectrolyte

multilayer (PEM) coating procedure, which involved an electrostatic layer-by-layer

deposition process [21, 22]. This coating was constructed in situ by alternating rinses of

positively and negatively charged polymers [23-30], and it was applied to the separation of

a series of basic proteins. In addition, in 2003, Rmaile and Schlenoff [17] demonstrated

that the use of optically active PEMs for chiral membrane separations allows very high

enantiomer permeation rates with encouraging selectivity. In our laboratory, a PEM

coating has been successfully employed as a separation medium for the achiral separation

of benzodiazepines (Chapter 1) and phenols [18, 19].

In this chapter, the PEM coating is applied to chiral separations by the use of the

anionic polymer poly (L-SULV) as the chiral discriminator. An examination of the

influence of several parameters that are used to obtain more efficient and more

reproducible chromatographic separations is also discussed. The PEM-coated capillary was

found to be remarkably robust with a performance of more than 230 runs.

4.2 Experimental

4.2.1 Apparatus and Conditions

All experiments were conducted using a Beckman P/ACE MDQ capillary

electrophoresis system with UV detection (Fullerton, CA). Fused-silica capillaries, 57 cm

(50 cm effective length) x 50 µm i.d., were purchased from Polymicro Technologies

(Phoenix, AZ). The applied voltage was 30 kV. The samples were injected by pressure.

1,1’-binaphthyl-2,2’-dihydrogenphosphate (BNP), 1,1’-bi-2-naphthol (BOH), secobarbital

and pentobarbital were injected at 0.5 psi (1 psi = 6894.76 Pa) for 3 s, and temazepam at

0.9 psi for 7 s. Unless stated otherwise, the temperature of the cartridge was maintained at

Page 124: Analytical separations using packed and open-tubular

104

25 °C using a liquid coolant. UV detection was performed at 214 nm for BNP, BOH,

secobarbital and pentobarbital, and 220 nm for temazepam.

4.2.2 Reagents and Chemicals

The analytes BNP, BOH, secobarbital, pentobarbital and temazepam, and

tris(hydroxymethyl)aminomethane (Tris) were purchased from Sigma Chemical Co. (St.

Louis, MO). Sodium borate (Na2B4O7) and sodium chloride (NaCl) were obtained from

Fisher Scientific (Fair Lawn, NJ). Boric acid (H3BO3) was purchased from Alfa Products

(Danvers, MA), and sodium phosphate (Na2HPO4) from Mallinckrodt & Baker, Inc. (Paris,

KY). The ionic liquids 1-ethyl-3-methyl-1H-imidazolium hexafluorophosphate (1E-3MI-

HFP) and 1-butyl-3-methylimidazolium tetrafluoroborate (1B-3MI-TFB) were purchased

from Aldrich Chemical Co. (Milwaukee, WI) and Chemada Fine Chemicals Ltd. (Nir

Itzhak, D. N. HaNegev 85455, Israel), respectively. The dipeptide L-leucylvalinate,

undecylenic acid, and N-hydroxysuccinimide were purchased from Sigma. Poly

(diallyldimethylammonium chloride) (PDADMAC; Mw = 200,000-350,000) was obtained

from Aldrich.

4.2.3 Sample and Background Electrolyte Preparation

The BGE at pH 10.0 consisted of 100 mM Tris and 10 mM Na2B4O7, the BGE at

pH 8.5 consisted of 25 mM Tris and 25 mM Na2B4O7, and the BGE at pH 7.2 consisted of

300 mM H3BO3 and 30 mM Na2HPO4. In all cases, the pH values were adjusted by using

either 1 M sodium hydroxide (NaOH) or 1 M hydrochloric acid (HCl). All solutions were

filtered using 0.45 µm polypropylene nylon filters and sonicated for 15 min before use.

The analytes BNP, BOH and secobarbital were dissolved in 50:50 methanol/water, and the

analytes pentobarbital and temazepam were dissolved in 80:20 methanol/water. The final

Page 125: Analytical separations using packed and open-tubular

105

analyte concentrations were 0.1 mg/ml for BNP and BOH, 0.2 mg/ml for secobarbital, and

0.5 mg/ml for pentobarbital and temazepam.

4.2.4 Synthesis of Polymeric Surfactant

The monomeric surfactant of sodium N-undecenoyl-L-leucylvalinate was

synthesized from the N-hydroxysuccinimide ester of undecylenic acid using a procedure

reported by Wang and Warner [7]. A 100 mM sodium salt solution of the monomer was

then polymerized by use of 60Co-γ radiation. The structure of poly (L-SULV) is shown in

Figure 4.1.

HN

C

[CH-CH]x

CNH

CO-Na+

O

O

O

Figure 4.1 Structure of poly (L-SULV).

Page 126: Analytical separations using packed and open-tubular

106

4.2.5 Procedure for Polyelectrolyte Multilayer Coating

PEM coatings were constructed in our laboratory according to a procedure

described in Chapter 3 [18]. The coatings were achieved by flushing the cationic and

anionic polymer deposition solutions through the capillaries. Polymer deposition solutions

contained 0.5% and 2.5% (w/v) polymer and 0.0-0.2 M NaCl or 0.01 M of the ionic liquid

1E-3MI-HFP or 1B-3MI-TFB. The capillary was first conditioned with water for 5 min,

with 1 M NaOH for 60 min, and then again with water for 15 min. The first layer of the

cationic polymer was deposited by flushing the PDADMAC solution through the capillary

for 5 min followed by a 5-min water rinse. All polymer depositions were done with 5 min

rinses followed by 5 min water rinses. The PEM coatings used for all the studies reported

here consisted of two layer pairs. However, for the chiral separations of the analytes

secobarbital and pentobarbital three layer pairs were constructed. Each coated capillary

was then flushed with the BGE until a stable current was achieved.

4.3 Results and Discussion

Several experimental parameters such as the type of additive in the polymer

deposition solutions, NaCl concentration, column temperature, and bilayer number were

studied to optimize the chiral separation of BNP. The influence of each parameter on chiral

separation was examined in detail as reported below. For the purpose of this work two

polymeric surfactants were examined: poly (sodium N-undecanoyl-L-valinate), poly (L-

SUV), and poly (L-SULV). The use of the first polymeric surfactant in the chiral

separation of BNP resulted in a single peak. Therefore, all studies were performed using

the polymeric surfactant poly (L-SULV).

Page 127: Analytical separations using packed and open-tubular

107

4.3.1 Effect of Additives in Polymer Deposition Solutions

Each polymer deposition solution contained a polymer and an additive. NaCl was

one of the additives used for this study. It has been reported that the addition of NaCl to the

polymer solution resulted in enhanced thickness for each layer [16, 25]. Several studies

have also shown that salt concentration is approximately proportional to the thickness of

each polyelectrolyte layer [31-33].

The ionic liquids 1E-3MI-HFP and 1B-3MI-TFB were also used as additives in the

solutions used for deposition. Ionic liquids are liquid electrolytes composed entirely of ions

with melting points at ambient temperature. In recent years, there has been a growing

interest in ionic liquids due to their distinct properties. They are good solvents for several

organic, inorganic and polymeric materials, and they have excellent chemical and thermal

stability. They are also good electrical conductors, and nonvolatile [33-35]. With these

properties in mind, the ionic liquids were explored for possible enhanced separations.

Figures 4.2a, 4.2b and 4.2c demonstrate the chiral separation of BNP using 0.01 M

NaCl, and the ionic liquids 1E-3MI-HFP or 1B-3MI-TFB as additives in the polymer

solutions, respectively. The addition of either ionic liquid increased the resolution of BNP

from 0.83 to 0.88 and 0.90. However, increasing the concentration of the ionic liquid

further resulted in the formation of aggregates in the cationic polymer solution. Therefore,

NaCl was used to further optimize the separation conditions.

4.3.2 Effect of NaCl Concentration

Polymer deposition solutions composed of 0.5% (w/v) polymer and 0.00 M, 0.01

M, 0.05 M, and 0.10 M NaCl were used to study the influence of NaCl concentration on

the chiral separation of BNP. As shown in Figure 4.3, the retention times of the analyte

Page 128: Analytical separations using packed and open-tubular

108

peaks slightly increased when NaCl was added to the solutions, probably due to an

increase in the thickness of the coating. In addition, an increase in resolution at higher

concentrations of NaCl was observed. The electroosmotic mobility decreased from 3.89 x

10-4 to 3.37 x 10-4 cm2V-1s-1 with increasing NaCl concentration. Based on these results,

0.1 M NaCl was used to further optimize the chiral separation of BNP.

4.3.3 Effect of Column Temperature

The influence of column temperature was also investigated. Figure 4.4 illustrates

the chiral separation of BNP at 35 °C, 25 °C and 15 °C. At a constant applied voltage, with

temperature decreasing, the retention times of R-(+)-BNP and S-(-)-BNP increased mainly

due to an increase in electrolyte viscosity. In addition, the electroosmotic mobility

decreased from 4.24 x 10-4 to 2.79 x 10-4 cm2V-1s-1 with decreasing temperature. In Figure

4.4, it is also shown that BNP was successfully resolved at a column temperature of 15 °C.

4.3.4 Effect of Bilayer Number

The effect of the number of bilayers on the resolution and analysis time of BNP

was also examined. It has been shown that an increase in the bilayer number resulted in an

enhanced film thickness [27, 36]. In this study, 2, 4, 8, and 12 bilayers were constructed in

the PEM coating. As shown in Figure 4.5, the analysis time, the resolution, and the

selectivity increased when more bilayers were constructed. However, two and four bilayers

did not demonstrate a change in selectivity, even though there was an increase in retention

time.

Page 129: Analytical separations using packed and open-tubular

109

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.008

-0.006

-0.004

-0.002

0.000

(a) 0.01 M NaCl R-(+)-BNP S-(-)-BNP

17

O

O

PO

OH

αααα=1.022

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.010

-0.008

-0.006

-0.004

-0.002

0.000 (c) 0.01 M 1B-3MI-TFB

17

αααα=1.0235

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.004

-0.003

-0.002

-0.001

0.000

0.001

(b) 0.01 M 1E-3MI-HFP

17

αααα=1.023

Figure 4.2 Chiral separation of BNP. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV); pressure injection, 0.5 psi for 3 s; background electrolyte, 100 mM Tris and 10 mM Na2B4O7 (pH 10.0); applied voltage, 30 kV; temperature, 25 °C; capillary, 57 cm (50 cm effective length) x 50 µm i.d.; detection, 214 nm. (a) 0.01 M NaCl; (b) 0.01 M 1E-3MI-HFP; (c) 0.01 M 1B-3MI-TFB.

Page 130: Analytical separations using packed and open-tubular

110

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.010

-0.005

0.000

0.005

0.010

0.00 M NaCl

17

αααα=1.021

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.008

-0.006

-0.004

-0.002

0.000

0.01 M NaCl

17

αααα=1.022

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.008

-0.006

-0.004

-0.002

0.000

0.05 M NaCl

17

αααα=1.025

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.005

-0.004

-0.003

-0.002

-0.001

0.10 M NaCl

17

αααα=1.026

Figure 4.3 Effect of NaCl concentration on the chiral separation of BNP. Conditions: same as Figure 4.2, except NaCl concentration was varied.

Page 131: Analytical separations using packed and open-tubular

111

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.004

-0.002

0.000

0.002 35 °C

17

αααα=1.019

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.005

-0.004

-0.003

-0.002

-0.001 25 °C

17

αααα=1.026

Minutes0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

AU

-0.004

-0.002

0.000

0.002

15 °C

17

αααα=1.044

Figure 4.4 Effect of column temperature on the chiral separation of BNP. Conditions: same as Figure 4.2, except temperature was varied.

Page 132: Analytical separations using packed and open-tubular

112

M inutes0 2 4 6 8 10 12 14 16 18 20 22 24 26

AU

-0.006

-0.005

-0.004

-0.003

2 Bilayers

αααα=1.026

M inute s0 2 4 6 8 10 12 14 16 18 20 22 24 26

AU

-0.002

-0.001

0.000

0.001

4 Bilayers

αααα=1.026

M inute s0 2 4 6 8 10 12 14 16 18 20 22 24 26

AU

-0.002

0.000

0.002

0.004

0.006

12 Bilayers

αααα=1.043

M inute s0 2 4 6 8 10 12 14 16 18 20 22 24 26

AU

-0.002

-0.001

0.000

0.001

8 Bilayers

αααα=1.040

Figure 4.5 Effect of bilayer number on the chiral separation of BNP. Conditions: same as Figure 4.2, except bilayer number was varied.

Page 133: Analytical separations using packed and open-tubular

113

4.3.5 Reproducibilities

Reproducibility of the PEM coating is an important factor for the evaluation of

column performance. Reproducibilities were evaluated by using the relative standard

deviation (RSD) [37] values of the EOF and the R-(+)-BNP peak. In both cases, the run-to-

run RSD values were obtained from 10 consecutive electrophoresis runs. For each NaCl

concentration added to the polymer solutions (0.00 M, 0.01 M, 0.05 M, and 0.10 M), three

capillaries were coated. The RSD values of the EOF ranged from 0.13% to 0.49%, and the

RSD values of the R-(+)-BNP peak ranged from 0.24% to 0.99% (Table 4.1). Figure 4.6 is

an illustration of this excellent reproducibility, and it demonstrates 10 electropherograms

obtained from 10 replicate runs when 0.05 M NaCl was used.

The capillary-to-capillary reproducibilities were obtained from 30 runs (10

consecutive runs performed in 3 different capillaries for each NaCl concentration). It was

observed that all RSD values were below 0.5% (Table 4.2). In addition, 10 more replicate

analyses were performed in 6 consecutive days. For the first two days, the RSD values

were above 0.2% (Table 4.3). However, after the second day, all RSD values dropped

below 0.2%. Therefore, conditioning of the coated capillaries significantly improves the

reproducibility.

4.3.6 Chiral Separation of Analytes

As shown in Figure 4.7, four more analytes were able to be resolved with

resolution (Rs) values ranging from 1.07 to 1.55. BOH, which is a partially anionic analyte

at pH 10, was almost baseline separated with a resolution value of 1.24 (Figure 4.7a).

Temazepam is a neutral chiral compound in the benzodiazepine class of analytes. The

racemate of this analyte was successfully and easily resolved with the resolution value of

Page 134: Analytical separations using packed and open-tubular

114

1.41 (Figure 4.7b). The barbiturates secobarbital and pentobarbital are partially anionic

compounds in a buffer solution of pH 7.2. The enantiomeric separation of both analytes

was performed using a three-bilayer PEM coating, 2.5% poly (L-SULV), and polymer

deposition solutions with 0.2 M NaCl. The chiral separations of both secobarbital (Figure

4.7c) and pentobarbital (Figure 4.7d) resulted in resolution values of 1.07 and 1.55,

respectively.

Table 4.1 Run-to-run reproducibilities of PEM capillary coating. Conditions: same as Figure 4.2, except NaCl concentration was varied.

RSD of R-BNP (%)RSD of EOF (%)

0.13

0.00 0.18

0.21

NaCl Concentration (M)

0.40

0.38

0.24

0.13

0.01 0.26

0.26

0.30

0.39

0.54

0.29

0.05 0.35

0.49

0.65

0.96

0.99

0.15

0.10 0.30

0.32

0.35

0.61

0.90

RSD of R-BNP (%)RSD of EOF (%)

0.13

0.00 0.18

0.21

NaCl Concentration (M)

0.40

0.38

0.24

0.13

0.01 0.26

0.26

0.30

0.39

0.54

0.29

0.05 0.35

0.49

0.65

0.96

0.99

0.15

0.10 0.30

0.32

0.35

0.61

0.90

Page 135: Analytical separations using packed and open-tubular

115

Minutes

4.0 5.0 6.0 7.0 8.0 9.0 10.0 11.0

AU

0.00

0.02

0.04

0.06 0.05 M NaCl

12.0

Figure 4.6 Illustration of a run-to-run reproducibility for the chiral separation of BNP. Conditions: same as Figure 4.2, except NaCl concentration, 0.05 M.

Page 136: Analytical separations using packed and open-tubular

116

Table 4.2 Capillary-to-capillary reproducibilities of PEM capillary coating. Conditions: same as Figure 4.2, except NaCl concentration was varied.

RSD of the EOF (%)

0.160.00

0.01

0.10

0.05

0.22

0.40

0.37

NaCl Concentration (M) RSD of the EOF (%)

0.160.00

0.01

0.10

0.05

0.22

0.40

0.37

NaCl Concentration (M)

Table 4.3 Reproducibilities of PEM capillary coating. Conditions: same as Figure 4.2, except NaCl concentration, 0.1 M.

RSD of the EOF (%)

0.30Day 1

Day 2

Day 4

Day 3

0.23

0.09

0.12

Day 6

Day 5 0.10

0.10

RSD of the EOF (%)

0.30Day 1

Day 2

Day 4

Day 3

0.23

0.09

0.12

Day 6

Day 5 0.10

0.10

Page 137: Analytical separations using packed and open-tubular

117

M in u te s1 0 . 0 1 0 .5 1 1 .0 1 1 .5 1 2 .0 1 2 .5 1 3 .0 1 3 . 5 1 4 . 0 1 4 . 5 1 5 .0

A U

- 0 . 0 0 3 4

- 0 . 0 0 3 2

- 0 . 0 0 3 0

- 0 . 0 0 2 8

αααα=1.029

BO H

M in ute s5 .0 5 .2 5 .4 5 .6 5 .8 6 .0 6 .2 6 .4 6 .6 6 .8 7 .0 7 .2 7 .4

A U

0 .0 0 0

0 .0 0 2

0 .0 0 4

0 .0 0 6

0 .0 0 8

0 .0 1 0

Pentobarbital

αααα=1.184

M in u te s5 .1 5 .2 5 .3 5 .4 5 .5 5 . 6 5 . 7 5 . 8 5 .9 6 . 0 6 . 1 6 . 2

A U

0 .0 0 0

0 .0 0 1

0 .0 0 2

0 .0 0 3

0 .0 0 4

Secobarbital

αααα=1.105

M in u te s1 1 . 0 1 1 .2 1 1 .4 1 1 . 6 1 1 . 8 1 2 .0 1 2 .2 1 2 .4 1 2 .6 1 2 . 8 1 3 . 0 1 3 . 2 1 3 .4 1 3 .6 1 3 .8 1 4 .0 1 4 . 2 1 4 . 4 1 4 .6

A U

0 . 0 0 0 1

0 . 0 0 0 2

0 . 0 0 0 3

0 . 0 0 0 4

N

N

O H

O

Cl

H 3 C

Temazepam

αααα=1.032

(a)

(b)

(c)

(d)

Figure 4.7 (a) Chiral separation of BOH. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV) with 0.1 M NaCl; pressure injection, 0.5 psi for 3 s; background electrolyte, 100 mM Tris and 10 mM Na2B4O7 (pH 10.0); detection, 214 nm. (b) Chiral separation of temazepam. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV) with 0.1 M NaCl; pressure injection, 0.9 psi for 7 s; background electrolyte, 25 mM Tris and 25 mM Na2B4O7 (pH 8.5); detection, 220 nm. (c) Chiral separation of secobarbital, and (d) Chiral separation of pentobarbital. Conditions for (c) and (d): 3 bilayers; 0.5% (w/v) PDADMAC and 2.5% (w/v) poly (L-SULV) with 0.2 M NaCl; pressure injection, 0.5 psi for 3 s; background electrolyte, 300 mM H3BO3 and 30 mM Na2HPO4 (pH 7.2); detection, 214 nm. Other conditions are the same as Figure 4.2.

Page 138: Analytical separations using packed and open-tubular

118

4.4 Conclusion

In this study, we were able to develop a simple PEM coating procedure for chiral

separations. The anionic polymer poly (L-SULV) proved to be a good chiral discriminator

for the separation of several analytes. The enantiomeric separations of BNP, BOH, and

temazepam were achieved using a two-bilayer PEM coating, 0.5% poly (L-SULV), and 0.1

M NaCl in the polymer deposition solutions. On the other hand, secobarbital and

pentobarbital were separated using a three-bilayer PEM coating, 2.5% poly (L-SULV), and

polymer deposition solutions with 0.2 M NaCl. In addition, the endurance of the PEM

coated capillary was more than 230 runs with RSD values of less than 1%. In all cases, the

run-to-run and capillary-to-capillary RSD values of EOF were less than 0.5%, and the run-

to-run RSD values of the R-(+)-BNP peak were less than 1%. It should be noted that the

chiral separation coating had to be modified from the procedure previously used for achiral

separations [18]. In our laboratory, further studies are ongoing to separate more chiral

compounds, and to better understand the structure of the coating.

4.5 References

1. Wang, J.; Warner, I. M. Anal. Chem. 1994, 66, 3773.

2. Jamali, F.; Mehvar, R.; Pasutto, F. M. J. Pharm. Sci. 1989, 78, 695.

3. Snopek, J.; Jelinek, I.; Smolkova-Keulemansova, E. J. Chromatogr. 1992, 609, 1.

4. Haddadian, F.; Billiot, E. J.; Shamsi, S. A.; Warner, I. M. J. Chromatogr. A 1999, 858, 219.

5. Shamsi, S. A.; Warner, I. M. Electrophoresis 1997, 18, 179.

6. Palmer, C. P.; Terabe, S. Anal. Chem. 1997, 69, 1852.

7. Wang, J.; Warner, I. M. J. Chromatogr. 1995, 711, 297.

Page 139: Analytical separations using packed and open-tubular

119

8. Shamsi, S. A.; Valle, B. C.; Billiot, F.; Warner, I. M. Anal. Chem. 2003, 75, 379.

9. Pundlett, K. L.; Armstrong, D. W. Anal. Chem. 1996, 68, 3493.

10. Shamsi, S. A.; Macossay, J.; Warner, I. M. Anal. Chem. 1997, 69, 2980.

11. Billiot, F. H.; Billiot, E. J.; Warner, I. M. J. Chromatogr. A 2001, 922, 329.

12. Billiot, E.; Warner, I. M. Anal. Chem. 2000, 72, 1740.

13. Erim, F. B.; Cifuentes, A.; Poppe, H.; Kraak, J. C. J. Chromatogr. A 1995, 708, 356.

14. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 2254.

15. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 5272.

16. Graul, T. W.; Schlenoff, J. B. Anal. Chem. 1999, 71, 4007.

17. Rmaile, H. H.; Schlenoff, J. B. J. Am. Chem. Soc. 2003, 125, 6603.

18. Kapnissi, C. P.; Akbay, C.; Schlenoff, J. B.; Warner, I. M. Anal. Chem. 2002, 74,

2328.

19. Kamande, M. W.; Kapnissi, C. P.; Zhu, X.; Akbay, C.; Warner, I. M. Electrophoresis 2003, 24, 945.

20. Hayes, J. D.; Malik, A. Anal. Chem. 2001, 73, 987.

21. Decher, G.; Schmitt, J. Prog. Colloid Polym. Sci. 1992, 89, 160.

22. Decher, G. Science 1997, 277, 1232.

23. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626.

24. Sui, Z.; Salloum, D.; Schlenoff, J. B. Langmuir 2003, 19, 2491.

25. Schlenoff, J. B.; Dubas, S. T.; Farhat, T. Langmuir 2000, 16, 9968.

26. Caruso, F.; Lichtenfeld, H.; Donath, E.; Möhwald, H. Macromolecules 1999, 32,

2317.

27. Laurent, D.; Schlenoff, J. B. Langmuir 1997, 13, 1552.

28. Smith, R. N.; Reven, L.; Barrett, C. J. Macromolecules 2003, 36, 1876.

Page 140: Analytical separations using packed and open-tubular

120

29. Rmaile, H. H.; Schlenoff, J. B. Langmuir 2002, 18, 8263.

30. Liu, Y.; Fanguy, J. C.; Bledsoe, J. M.; Henry, C. S. Anal. Chem. 2000, 72, 5939.

31. Dubas, S. T.; Schlenoff, J. B. Macromolecules 1999, 32, 8153.

32. Lösche, M.; Schmitt, J.; Decher, G.; Bouwman, W. G.; Kjaer, K. Macromolecules 1998, 31, 8893.

33. Armstrong, D. W.; He, L.; Liu, Y. Anal. Chem. 1999, 71, 3873.

34. Dupont, J.; De Souza, R. F.; Suarez, P. A. Z. Chem. Rev. 2002, 102, 3667.

35. Yanes, E. G.; Gratz, S. R.; Baldwin, M. J.; Robison, S. E.; Stalcup, A. M. Anal.

Chem. 2001, 73, 3838.

36. Schlenoff, J. B.; Dubas, S. T. Macromolecules 2001, 34, 592.

37. Skoog, D. A.; Holler, J. F.; Nieman, T. A. In Principles of Instrumental Analysis, 5th ed.; Sherman, M., Bortel, J., Messina, F., Eds.; Harcourt Brace College Publishers: Orlando, FL, 1998.

Page 141: Analytical separations using packed and open-tubular

121

CHAPTER 5.

INVESTIGATION OF THE STABILITY OF POLYELECTROLYTE MULTILAYER COATINGS IN OPEN-TUBULAR CAPILLARY

ELECTROCHROMATOGRAPHY USING LASER SCANNING CONFOCAL MICROSCOPY

5.1 Introduction

In recent years, layer-by-layer deposition process of polyelectrolytes on hydrophilic

surfaces via electrostatic interactions has attracted considerable attention. This method was

first developed by Decher et al. [1, 2]. In this approach, a substrate is alternately exposed

to solutions of positively and negatively charged polyelectrolytes. The alternate adsorption

of these oppositely charged polyelectrolytes produces fairly uniform thin films that have

been used in the areas of light-emitting devices, sensors, coatings, nonlinear optics,

catalysis, patterning, diagnostics, bioadhesion, and drug delivery [3-6].

The layer-by-layer method has also been involved in analytical separations [7-11].

This coating involves a polyelectrolyte multilayer (PEM), constructed in a fused-silica

capillary by alternating rinses of cationic and anionic polyelectrolytes [12-22]. The first

layer of the cationic polymer adds to the negatively charged surface of the capillary. The

construction of the cationic layer reverses the charge on the surface and primes the film for

the addition of the next layer of anionic polymer. These layer-by-layer coatings have

shown strong stability and high resistance to charge and deterioration during use [7, 8, 23,

24].

Katayama et al. [23, 24] established a stable modification of the inner walls of the

capillary by use of the successive multiple ionic-polymer layer (SMIL) coating. Their first

study focused on anion-modified capillaries that were achieved by first attaching the

Page 142: Analytical separations using packed and open-tubular

122

cationic polymer to the inner walls of the capillaries and then, the anionic polymer to the

cationic polymer layer [23]. The endurance of the coated capillary was more than 100 runs,

and this is due to the multiple attachment of the anionic polymer. The coating was also

tolerant to some organic and alkaline solvents. However, the coated capillary was unstable

after 0.1 M HCl and CH3CN rinsing, and the coating was detached after 1 M NaOH rinse.

In their second study, these authors developed a cation-modified capillary by attaching the

cationic polymer to the anionic polymer layer [24]. This coated capillary endured 600

replicate runs, and showed strong stability against 1 M NaOH and 0.1 M HCl with

degradation ratios of less than 2%. Degradation ratios are changes in electroosmotic flow

(EOF).

Graul and Schlenoff [7] coated fused-silica capillaries with thin films of charged

polymers by use of the PEM coating procedure. They demonstrated that these columns

were stable to extremes of pH and ionic strength, and to dehydration/rehydration. For the

stability studies against exposure to extreme pH values and ionic strengths, the 6.5-layer

pair-coated capillary was flushed twice with 0.01 M (pH 12) for 10 min, and twice with

0.01 M HCl (pH 2) for 10 min.

In the study reported in this chapter, we describe an approach that uses open-

tubular capillary electrochromatography (OT-CEC) and laser scanning confocal

microscopy (LSCM) to further investigate the stability of the PEM coating approach that

was previously described. In this approach, the polymeric surfactant poly (sodium N-

undecanoyl-L-leucylvalinate), poly (L-SULV), was used in the construction of the coating.

Using OT-CEC, both electroosmotic mobility and selectivity changes were measured after

flushing the capillaries with 0.1 M and 1.0 M NaOH. In addition, a correlation between

Page 143: Analytical separations using packed and open-tubular

123

flushing time and change in electroosmotic mobility and selectivity was investigated. The

structural changes of these coatings were also monitored and imaged after exposure to 1.0

M NaOH with LSCM. LSCM has proven to be a useful and valuable tool for

characterization of packed columns, coatings, tablets, capsules, living cells, and other

surfaces [25-39].

5.2 Experimental

5.2.1 Apparatus and Conditions

(i) Capillary Electrochromatography

All experiments were performed with a Beckman P/ACE MDQ capillary

electrophoresis system using UV detection (Fullerton, CA). Fused-silica capillaries of 57

cm (50 cm effective length) x 50 µm i.d., were purchased from Polymicro Technologies

(Phoenix, AZ). The capillaries were thermostated at 25 °C by use of liquid coolant. The

applied voltage was 30 kV, and the samples were injected by pressure at 0.5 psi (1 psi =

6894.76 Pa) for 3 s. UV detection was done at 214 nm.

(ii) Laser Scanning Confocal Microscopy

An inverted confocal microscope (Axiovert 100, LSM 510 scan module, Carl Zeiss

Inc., Thornwood, NY) was used to image the PEM-coated capillaries. The setup for

imaging has been described elsewhere [25]. Briefly, a 4-cm coated capillary was placed

between a microscope slide and a coverslip supported by two spacers consisting of two

short pieces of fused-silica capillary. The window of the coated capillary was immersed in

a refractive index matching fluid. Refractive index matching is important for obtaining

minimally distorted images by reduction of refraction and scattering from the capillary

walls. This arrangement is then placed on the microscope stage.

Page 144: Analytical separations using packed and open-tubular

124

The imaging system was equipped with a helium-neon laser (Uniphase, 1674P) that

provided excitation at 543 nm. All confocal images were obtained with a 20x (NA 0.5) dry

objective. Band-pass or long-pass filters were used for selection of the spectral range of

fluorescence emission. Fluorescence was directed through an adjustable pinhole to a

photomultiplier tube.

5.2.2 Reagents and Chemicals

The analyte 1,1’-binaphthyl-2,2’-dihydrogenphosphate (BNP), and the buffer

tris(hydroxymethyl)aminomethane (Tris) were purchased from Sigma Chemical Co. (St.

Louis, MO). Sodium borate (Na2B4O7), sodium chloride (NaCl), and sodium hydroxide

(NaOH) were obtained from Fisher Scientific (Fair Lawn, NJ). Rhodamine 6G (R6G) and

PDADMAC (Mw = 200,000-350,000) were obtained from Aldrich (Milwaukee, WI).

Refractive index matching oil (n=1.51) was purchased from Carl Zeiss Microimaging, Inc.

(Thornwood, NY). The dipeptide L-leucylvalinate, undecylenic acid, and N-

hydroxysuccinimide were also obtained from Sigma.

5.2.3 Sample and Background Electrolyte Preparation

The analyte BNP was dissolved in 50:50 methanol/water. Its final concentration

was 0.1 mg/ml. The background electrolyte (BGE) consisted of 100 mM Tris and 10 mM

Na2B4O7. The pH was adjusted to 10 by use of 1.0 M NaOH. The solution was filtered

using 0.45 µm polypropylene nylon filters and sonicated for 15 min before use. The

fluorescent dye R6G was dissolved in deionized water at concentrations of 10 µM and 100

µM.

Page 145: Analytical separations using packed and open-tubular

125

5.2.4 Synthesis of Polymeric Surfactant

The monomeric surfactant of sodium N-undecenoyl-L-leucylvalinate was

synthesized in our laboratory using the N-hydroxysuccinimide ester of undecylenic acid

according to a procedure previously described by Wang and Warner [40]. A 100 mM

sodium salt solution of the monomer was then polymerized by use of 60Co-γ radiation to

form the polymeric surfactant poly (L-SULV). The structure of poly (L-SULV) is shown in

Figure 4.1 (Chapter 4).

5.2.5 Procedure for Polyelectrolyte Multilayer Coating

The procedure for construction of the PEM coating has been described in Chapters

3 and 4 [8]. The cationic and anionic polymer deposition solutions were alternately flushed

through the capillary by use of the rinse function on the Beckman CE system. Both

solutions contained 0.5% (w/v) polymer in 0.1 M aqueous NaCl solution. First, the fused-

silica capillary was conditioned with water for 5 min, 1.0 M NaOH for 60 min, and again

with water for 15 min. Then, the first layer of the polymer PDADMAC was constructed by

rinsing the polymer solution through the capillary for 5 min followed by a water rinse for

another 5 min. All other polymer depositions were performed by flushing the capillary

with the polymer solutions for 5 min followed by 5-min water rinses. The multilayer

coatings used for the OT-CEC and LSCM studies consisted of two and twenty layer pairs.

For the LSCM experiments, after each coating was constructed, the capillary was flushed

with 100 µM R6G for 10 min in order to incorporate the fluorescent probe into the

polymer coating. For the chromatographic experiments, after the construction of the

coating, the capillary was conditioned with the BGE until a stable current was achieved.

Page 146: Analytical separations using packed and open-tubular

126

5.3 Results and Discussion

5.3.1 Open-Tubular Capillary Electrochromatography

Experiments performed in OT-CEC studied the stability of capillary surface, the

capillary recovery, and the coating regeneration. These parameters were evaluated by

computing the electroosmotic mobility, EOFµ , and the selectivity factor, α .

As stated in Chapter 2, the electroosmotic mobility is described by the following

equation:

0VtLL td

eo =µ (5.1)

where dL is the distance from injector to detector; tL is the total length of the capillary; 0t

is the migration time of the EOF marker; and V is the applied voltage. Methanol was used

as the EOF marker in this study since it is not retained by the PEM coating.

Selectivity is another important parameter that was used to study the stability and

regeneration of the coating after exposure to NaOH. The selectivity factor of a capillary

column for two species A and B, as already discussed, is defined as:

( )( ) 0

0

tttt

Ar

Br

−−

=α (5.2)

where ( )Brt and ( )Art are the retention times of B and A, respectively. In this study, B and

A are the enantiomers S-(-)-BNP and R-(+)-BNP, respectively.

5.3.1.1 Coating Stability

To study the stability of the capillary surface, two bilayers were constructed. First,

the EOF was measured before the coating was deposited on the capillary walls of the

fused-silica capillary. The measured electroosmotic mobility was 5.04 x 10-4 cm2V-1s-1.

Page 147: Analytical separations using packed and open-tubular

127

The EOF was again measured after the two bilayers were constructed in the PEM coating.

In this case, the electroosmotic mobility decreased to 3.31 x 10-4 cm2V-1s-1. Then, 0.1 M

and 1.0 M NaOH solutions were flushed through the PEM-coated capillary, and the

magnitude of the EOF was measured. Figure 5.1 demonstrates the dependence of

electroosmotic mobility on the NaOH flushing time. In addition, this figure illustrates

graphically how NaOH can increase the value of electroosmotic mobility, as compared to

the 2-bilayer capillary that is not treated with the strong base. As shown, 0.1 M NaOH did

not have a major impact on electroosmotic mobility, as compared to 1.0 M NaOH, even

after flushing the coated capillary for 100 min. A 5-min 1.0 M NaOH flushing time slightly

increased the value of electroosmotic mobility to 3.35 x 10-4 cm2V-1s-1. However, fifteen

more minutes demonstrated a significant increase that was still observed upon continuous

rinse with 1.0 M NaOH. A total of 230 min resulted in an electroosmotic mobility of 5.03

x 10-4 cm2V-1s-1. In addition, a total of 350 min gave the same EOF value obtained before

deposition of the coating (5.04 x 10-4 cm2V-1s-1). This clearly demonstrates that the coating

was completely desorbed from the capillary column. It should be noted that each data point

in this graph and the graphs that follow is the average value obtained from running each

experiment multiple times. Standard deviation values were also calculated for each data

point. However, these values are very small, and not perceptible on the graphs.

Figure 5.2 demonstrates the dependence of selectivity on the NaOH flushing time.

Before polymer deposition, the selectivity was 1.0 since no separation between the two

enantiomers of BNP was observed. The construction of two bilayers gave a selectivity

value of 1.026, which was relatively stable during the coating’s exposure to 0.1 M NaOH.

A slight decrease was observed when 1.0 M NaOH solution was flushed through the

Page 148: Analytical separations using packed and open-tubular

128

coated capillary. A total flush time of 170 min with 1.0 M NaOH resulted in a single BNP

peak. This observation also confirmed that the coating was removed after a long exposure

to NaOH.

3

3.5

4

4.5

5

5.5

0 20 40 60 80 100 120Time (minutes)

3

3.5

4

4.5

5

5.5

0 50 100 150 200 250 300 350 400

EOF

(cm

2 V-1

s-1 x

10-4

)

time (minutes)

0.1 M NaOH 1.0 M NaOH

Figure 5.1 Dependence of electroosmotic mobility on the NaOH flushing time. Conditions: 2 bilayers; 0.5% (w/v) PDADMAC and 0.5% (w/v) poly (L-SULV) with 0.1 M NaCl; pressure injection, 0.5 psi for 3 s; background electrolyte, 100 mM Tris and 10 mM Na2B4O7 (pH 10.0); applied voltage, 30 kV; temperature, 25 °C; capillary, 57 cm (50 cm effective length) x 50 µm i.d.; detection, 214 nm.

The stability was further evaluated by calculating the change in EOF, also termed

“degradation ratio” [23, 24]. This ratio was calculated by use of the following equation:

Degradation ratio = 1001

21 ×−

EOFEOFEOF

(5.3)

Page 149: Analytical separations using packed and open-tubular

129

where 1EOF and 2EOF were measured before and after flushing the capillary with the

NaOH solution, respectively. In this case, 1EOF is as mentioned above, i.e. 3.31 x 10-4

cm2V-1s-1. These degradation ratio values are reported in Table 5.1. The PEM-coated

capillary was relatively stable after rinsing with 0.1 M NaOH for 110 min, followed by a

rinse with 1.0 M NaOH for 5 min. The degradation ratios obtained until the last exposure

(1.0 M NaOH for 5 min) were all below 2%. However, after additional 1.0 M NaOH rinse,

the degradation ratios were above 2%. The coating was finally detached from the capillary

wall after long exposure to 1.0 M NaOH as evidenced by α and the EOF value.

0.99

1

1.01

1.02

1.03

1.04

1.05

0 20 40 60 80 100 120Time (minutes)

0.99

1

1.01

1.02

1.03

1.04

1.05

0 50 100 150 200 250 300 350 400

Sele

ctiv

ity

time (minutes)

0.1 M NaOH 1.0 M NaOH

Figure 5.2 Dependence of selectivity on the NaOH flushing time. Conditions: same as Figure 5.1.

Page 150: Analytical separations using packed and open-tubular

130

Table 5.1 Stability of PEM capillary coating. Conditions: same as Figure 5.1.

Degradation Ratio(%)

EOF2 Average (min)(x 10-4 cm2 V-1s-1)

5.0

Total Flushing Time(min)

0.1 M NaOH

20.0

50.0

110.0

1.0 M NaOH

5.0

20.0

50.0

110.0

170.0

230.0

350.0

3.31

3.31

3.31

3.26

0.00

0.00

0.00

1.51

3.35

3.65

3.86

4.39

4.45

5.03

5.04

1.21

10.27

16.61

32.63

34.44

51.96

52.27

5.3.1.2 Capillary Recovery and Coating Regeneration

When the coating was completely detached, and a selectivity value of 1.0 was

obtained, a 2-bilayer coating was reconstructed in the same fused-silica capillary. Figure

Page 151: Analytical separations using packed and open-tubular

131

5.3 graphically demonstrates how electroosmotic mobility gradually changes after again

flushing the coated capillary with NaOH solutions. The trend observed in this figure is the

same as obtained when the first 2-bilayer coating was constructed. The EOF after the last

coating deposition was 3.32 x 10-4 cm2V-1s-1, as compared to the value of 3.31 x 10-4 cm2V-

1s-1 that was measured after the first coating. In addition, a similar trend was observed with

the selectivity values, as shown in Figure 5.4.

3

3.5

4

4.5

5

5.5

0 20 40 60 80 100 120Time (minutes)

3

3.5

4

4.5

5

5.5

0 50 100 150 200 250 300 350 400

0.1 M NaOH 1.0 M NaOH

EOF

(cm

2 V-1

s-1 x

10-4

)

time (minutes)

Figure 5.3 Coating Regeneration - Dependence of electroosmotic mobility on the NaOH flushing time. Conditions: same as Figure 5.1.

Page 152: Analytical separations using packed and open-tubular

132

0.99

1

1.01

1.02

1.03

1.04

1.05

0 20 40 60 80 100 120Time (minutes)

0.99

1

1.01

1.02

1.03

1.04

1.05

0 50 100 150 200 250 300 350 400

Sele

ctiv

ity

time (minutes)

0.1 M NaOH 1.0 M NaOH

Figure 5.4 Dependence of selectivity on the NaOH flushing time. Conditions: same as Figure 5.1.

Another important consideration for this study is the amount of time that is needed

to completely detach the coating for capillary recovery. This allows us regeneration of the

coating by use of this simple PEM coating procedure. For purpose of this study, two and

twenty bilayers were constructed. Both electroosmotic mobility and selectivity

measurements illustrated that a 2-bilayer and a 20-bilayer PEM coating (data not shown)

could be completely detached from the capillary walls after approximately 3.5 and 9.5

hours, respectively of continuous exposure to 1.0 M NaOH.

Page 153: Analytical separations using packed and open-tubular

133

5.3.2 Laser Scanning Confocal Microscopy

Using the results from the above study, one can conclude that the PEM coating

constructed using polymeric surfactants as anionic polymers affects separations [8-10, 44].

However, we desired an imaging technique that would allow visualization of the PEM-

coated capillaries for characterization of the surface morphology. In this part of our study,

LSCM was used as an imaging technique to examine the coating. LSCM was also used to

image the structural changes of the coating after its exposure to NaOH. For the

experiments reported here, R6G proved to be a suitable fluorescent dye to examine the

coating since this positively charged probe prefers the environment of the coated phase.

Figure 5.5 is a comparison of a 20-bilayer capillary (left) and a fused-silica

capillary (right) side by side in a single image. This comparison is used to visually

demonstrate the difference between a coated and bare capillary. In both sections of the

capillaries, 10 µM R6G in water was flushed, and then, air was injected in order to wash

the excess solution out and dry the capillary. This figure illustrates z-section images that

were collected at different depths of focus of the LSCM. The pinhole size of the LSCM

was adjusted such that the optical slice was 3.3 µm. From Figure 5.5a to Figure 5.5f, the z-

sectioning goes deeper into the PEM-coated capillary. The focal plane in Figure 5.5a is at

the outer edge of the capillary wall. Figure 5.5f slices through the outer wall of the

capillary on the opposite side. Figures 5.5c and 5.5d slice though regions of the capillary

near the center plane. As shown in these images, the walls of the 20-bilayer capillary are

very bright as compared to the fused-silica capillary. Since the R6G should be retained to a

greater extent in the coated phase than on the wall of the fused-silica capillary, this

difference confirms the existence of a coating on the capillary wall.

Page 154: Analytical separations using packed and open-tubular

134

(a) (b)

(c) (d)

(e) (f)

20-bilayer capillary Fused-silica capillary

Figure 5.5 The z-section images of a 20-bilayer capillary (left) and a fused-silica capillary (right). Optical slice thickness: 3.3 µm. z-Step size 1.65 µm. Every fourth image is shown. All images are 204 x 512 pixels (460.6 µm x 115.1 µm).

Page 155: Analytical separations using packed and open-tubular

135

The same 20-bilayer capillary was exposed to 1.0 M NaOH. It was then flushed

again with 10 µM R6G and injected with air. Using the confocal microscope, we observed

a change in the structure of the coating, i.e. large portions of the polymer coating have

been removed from the wall. This observation is clearly illustrated in Figure 5.6. The

images of this figure represent again z-section images that were taken as the excitation

beam traversed deeper into the coated capillary.

(a) (b) (c)

(d) (e) (f)

Figure 5.6 The z-section images of a portion of the same 20-bilayer capillary, as in Figure 5.5. Exposure to 1.0 M NaOH. Optical slice thickness: 3.3 µm. z-Step size 1.65 µm. Every third image is shown. All images are 512 x 512 pixels (65.8 µm x 65.8 µm).

Page 156: Analytical separations using packed and open-tubular

136

Additional LSCM images of larger portions of the capillary at different locations

demonstrate similar discontinuities in the coating (Figure 5.7). The area depicted in Figure

5.7a includes the discontinuity and a large region where the coating on the walls of the

capillary does not appear to be damaged to the same extent. Figure 5.7b includes several

discontinuities along the wall of the capillary where the coating has been greatly affected.

(a) (b)

Figure 5.7 LSCM images of different regions within the same 20-bilayer capillary, as in Figure 5.6. Optical slice thickness: 3.3 µm. Both images are 512 x 2048 pixels (115.1 µm x 460.6 µm).

Page 157: Analytical separations using packed and open-tubular

137

When the same 20-bilayer capillary was exposed to additional 1.0 M NaOH, the

majority of the coating appears to be removed, leaving only several bright islands on the

wall (Figure 5.8). In addition, this last image demonstrates that the coating is removed after

long exposure to 1.0 M NaOH. The same procedure was followed with the 2-bilayer

capillary (data not shown).

Figure 5.8 LSCM image of the same 20-bilayer capillary, as in Figure 5.7. Longer exposure to 1.0 M NaOH. Optical slice thickness: 3.3 µm. The image is 512 x 820 pixels (115.1 µm x 184.4 µm).

5.4 Conclusion

The stability of PEM coatings was further investigated in this study after exposure

to NaOH solutions. The OT-CEC study, which allowed measurements of electroosmotic

mobility and selectivity, suggests that the PEM coating is relatively stable against 0.1 M

NaOH. In addition, both OT-CEC and LSCM demonstrated the ability to recover the

Page 158: Analytical separations using packed and open-tubular

138

capillaries by flushing them with 1.0 M NaOH. We were also able to visualize the coating

within the capillary, and to study its structural changes by using LSCM. This helped us

prove that the coating exists, allowed us to see the effects of NaOH on the coating, and

demonstrated that the coating can be removed after long exposure to 1.0 M NaOH.

Although LSCM provided new information, it did not provide us with all the details

regarding the structure of the coating.

5.5 References

1. Decher, G.; Schmitt, J. Prog. Colloid Polym. Sci. 1992, 89, 160.

2. Decher, G. Science 1997, 277, 1232.

3. Schlenoff, J. B.; Dubas, S. T.; Farhat, T. Langmuir 2000, 16, 9968.

4. Smith, R. N.; Reven, L.; Barrett, C. J. Macromolecules 2003, 36, 1876.

5. Schlenoff, J. B.; Ly, H.; Li, M. J. Am. Chem. Soc. 1998, 120, 7626.

6. Farhat, T. R.; Schlenoff, J. B. Langmuir 2001, 17, 1184.

7. Graul, T. W.; Schlenoff, J. B. Anal. Chem. 1999, 71, 4007.

8. Kapnissi, C. P.; Akbay, C.; Schlenoff, J. B.; Warner, I. M. Anal. Chem. 2002, 74, 2328.

9. Kamande, M. W.; Kapnissi, C. P.; Zhu, X.; Akbay, C.; Warner, I. M.

Electrophoresis 2003, 24, 945.

10. Kapnissi, C. P.; Valle, B. C.; Warner, I. M. Anal. Chem. 2003, 75, 6097.

11. Rmaile, H. H.; Schlenoff, J. B. J. Am. Chem. Soc. 2003, 125, 6602.

12. Müller, M. Biomacromolecules 2001, 2, 262.

13. Laurent, D.; Schlenoff, J. B. Langmuir 1997, 13, 1552.

14. Schlenoff, J. B.; Dubas, S. T. Macromolecules 2001, 34, 592.

Page 159: Analytical separations using packed and open-tubular

139

15. Caruso, F.; Lichtenfeld, H.; Donath, E.; Möhwald, H. Macromolecules 1999, 32, 2317.

16. Rmaile, H. H.; Schlenoff, J. B. Langmuir 2002, 18, 8263.

17. Castelnovo, M.; Joanny, J.-F. Langmuir 2000, 16, 7524.

18. Dubas, S. T.; Schlenoff, J. B. Macromolecules 2001, 34, 3736.

19. Dubas, S. T.;Schlenoff, J. B. Macromolecules 1999, 32, 8153.

20. Hayes, J. D.; Malik, A. Anal. Chem. 2001, 73, 987.

21. Farhat, T.; Yassin, G.; Dubas, S. T.; Schlenoff, J. B. Langmuir 1999, 15, 6621.

22. Sui, Z.; Salloum, D.; Schlenoff, J. B. Langmuir 2003, 19, 2491.

23. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 2254.

24. Katayama, H.; Ishihama, Y.; Asakawa, N. Anal. Chem. 1998, 70, 5272.

25. Lowry, M.; He, Y.; Geng, L. Anal. Chem. 2002, 74, 1811.

26. Zhu, H.; Ji, J.; Tan, Q.; Barbosa, M. A.; Shen, J. Biomacromolecules 2003, 4, 378.

27. Moya, S.; Donath, E.; Sukhorukov, G. B.; Auch, M.; Bäumler, H.; Lichtenfeld, H.;

Möhwald, H. Macromolecules 2000, 33, 4538.

28. Guo, H. X.; Heinämäki, J.; Yliruusi, J. International J. Pharmaceutics 1999, 186, 99.

29. Silvano, D.; Krol, S.; Diaspro, A.; Cavalleri, O.; Gliozzi, A. Microscopy Research

and Technique 2002, 59, 536.

30. Georgieva, R.; Moya, S.; Leporatti, S.; Neu, B.; Bäumler, H.; Reichle, C.; Donath, E.; Möhwald, H. Langmuir 2000, 16, 7075.

31. Voigt, A.; Lichtenfeld, H.; Sukhorukov, G. B.; Zastrow, H.; Donath, E.; Bäumler,

H.; Möhwald, H. Ind. Eng. Chem. Res. 1999, 38, 4037.

32. Lamprecht, A.; Schäfer, U. F.; Lehr, C.-M. European J. Pharmaceutics and Biopharmaceutics 2000, 49, 1.

33. Stricker, S. A.; Whitaker, M. Microscopy Research and Technique 1999, 46, 356.

Page 160: Analytical separations using packed and open-tubular

140

34. Gao, C.; Leporatti, S.; Donath, E.; Möhwald, H. J. Phys. Chem. B 2000, 104, 7144.

35. Buttino, I.; Ianora, A.; Carotenuto, Y.; Zupo, V.; Miralto, A. Microscopy Research and Technique 2003, 60, 458.

36. Khopade, A. J.; Caruso, F. Langmuir 2003, 19, 6219.

37. Astrakharchik-Farrimond, E.; Shekunov, B. Y.; Sawyer, N. B. E.; Morgan, S. P.;

Somekh, M. G.; See, C. W. Particle and Particle Systems Characterization 2003, 20, 104.

38. Missert, N.; Copeland, R. G.; Barbour, J. C.; Mikkalson, J. E. Electrochemical

Society Proceedings 2001, 22, 687.

39. Caponetti, G.; Hrkach J. S.; Kriwet, B.; Poh, M.; Lotan, N.; Colombo, P.; Langer, R. J. Pharm. Sci. 1999, 88, 136.

40. Wang, J.; Warner, I. M. J. Chromatogr. 1995, 711, 297.

41. Chiari, M.; Nesi, M.; Righetti, P. G. In Capillary Electrophoresis in Analytical

Biotechnology; Righetti, P. G., Ed.; CRC Press, Inc.: Boca Raton, FL, 1996.

42. Chankvetadze, B. In Capillary Electrophoresis in Chiral Analysis; John Wiley & Sons Ltd: London, UK, 1997.

43. Skoog, D. A.; Holler, J. F.; Nieman, T. A. In Principles of Instrumental Analysis,

5th ed.; Sherman, M., Bortel, J., Messina, F., Eds.; Harcourt Brace College Publishers: Orlando, FL, 1998.

44. Zhu, X.; Kamande, M. W.; Thiam, S.; Kapnissi, C. P.; Mwongela, S. M.; Warner, I.

M. Electrophoresis, 2004, 25, 562.

Page 161: Analytical separations using packed and open-tubular

141

CHAPTER 6.

CONCLUSIONS AND FUTURE STUDIES

In this work, two different modes of capillary electrochromatography (CEC) were

used for the achiral and chiral separations of various classes of analytes. In Chapter 2 of

this dissertation, the packed mode of CEC is described. A 40-cm packed bed of Reliasil 3

µm C18 stationary phase was able to separate seven benzodiazepines. The separation

conditions were optimized by varying the mobile phase, the amount of organic modifier,

the buffer concentration, the applied voltage, and the column temperature. A mobile phase

composition of Tris.HCl (pH 8)-acetonitrile (60:40), an electrolyte concentration of 30

mM, and a temperature of 15 °C with an applied voltage of 20 kV proved to be optimum.

In addition, the packed capillary electrochromatographic method developed here was

applied to the characterization of oxazepam in a standard urine sample that contained

various concentrations of other drugs, such as acetaminophen, amphetamines, imipramine,

morphine, cocaine, etc.

Some preliminary data were also obtained for the separation of a mixture of six β-

blockers using packed-CEC with octadecyl silica stationary phase and co-polymerized

micelle SUS/SUG (sodium undecylenic sulfate/sodium N-undecanoyl-L-glycinate)

“pseudostationary phase.” β-blockers are clinically important drugs, used for the treatment

of several disorders such as angina pectoris, cardiac arrhythmias, and hypertension. They

have also been used as doping agents for athletes in order to improve their athletic

performance in cases where high psychological pressure may cause the heart to race. The

structures of the β-blockers used in this study are shown in Figure 6.1. Co-polymerized

micelles are produced by irradiating a micellar solution of surfactants with 60Co

Page 162: Analytical separations using packed and open-tubular

142

irradiation. Their main advantage is that they can be used in a wider pH range than the

conventional polymeric surfactants, since they are soluble at and below pH 6.8. The

studies that have been performed until now showed that the presence of the co-

polymerized micelle SUS/SUG improved the resolution of the analytes. The resolution was

also proportional to the concentration of the co-polymerized micelle and the percentage of

SUS in the co-polymerized micelle. The studies on the separation of β-blockers will be

continued, and when optimum conditions are established, the method will be applied to a

real sample (urine or blood).

1. Sotalol 2. Atenolol

4. Labetalol

5. Propranolol 6. Acebutolol

3. Metoprolol

CHCH2NHCH(CH3)2CH3SO2NH

OH

OCH2CHCH2NHCH(CH3)2H2NCCH2

O OH

OCH2CHCH2NHCH(CH3)2CH3OCH2CH2

OH

HO

H2NC

CHCH2NHCHCH2CH2

O

CH3

OH

OCH2CHCH2NHCH(CH3)2

OH

CCH3

O

OCH2CHCH2NHCH(CH3)2

OH

CH3CH2CH2CNH

O

Figure 6.1 Structures of the six β-blocker analytes.

Page 163: Analytical separations using packed and open-tubular

143

Chapter 3 is an examination of the open-tubular mode of CEC, which is an

alternative approach to conventional CEC. The primary advantage of open-tubular CEC

(OT-CEC) is the elimination of problems associated with frits and silica particles in

conventional CEC. In this approach, fused silica capillaries coated with thin films of

physically adsorbed charged polymers were developed by use of a polyelectrolyte

multilayer (PEM) coating procedure. The PEM coating was constructed in situ by

alternating rinses with positively and negatively charged polymers, where the negatively

charged polymer was a polymeric surfactant. In this study, poly

(diallyldimethylammonium chloride), PDADMAC, was used as the cationic polymer and

poly (sodium N-undecanoyl-L-glycinate), poly (L-SUG), was used as the anionic polymer

for PEM coating. The performance of the modified capillaries as a separation medium was

evaluated by use of seven benzodiazepines as analytes. The run-to-run, day-to-day, week-

to-week and capillary-to-capillary reproducibilities of electroosmotic flow (EOF) were

very good with relative standard deviation (RSD) values of less than 1% in all cases. In

addition, the chromatographic performance of the monomeric form of the polymeric

surfactant was compared for the separation of these analytes. The anionic surfactant used

for PEM coating construction in this experiment was the monomeric (nonpolymerized)

surfactant at the concentration of 0.5% (w/v) (19 mM). Almost no separation was noted,

even though the monomeric surfactant concentration was above the normal critical

micellar concentration of the nonpolymerized surfactant (7 mM). Therefore, the molecular

micelle allows better discrimination of the hydrophobic analytes than the conventional

micelle. Furthermore, the PEM-coated capillary was remarkably robust with more than 200

Page 164: Analytical separations using packed and open-tubular

144

runs accomplished in this study. Strong stability against extreme pH values (pH 11.0 and

pH 3.0) was also observed.

In Chapter 4, the fused-silica capillaries were again modified using the PEM

coating procedure. The quaternary ammonium salt PDADMAC was used as the cationic

polymer, and the polymeric surfactant poly (sodium N-undecanoyl-L-leucylvalinate), poly

(L-SULV), was used as the anionic polymer. In this study, the PEM coating was applied to

investigate the chiral separations of 1,1’-binaphthyl-2,2’-dihydrogenphosphate (BNP),

1,1’-bi-2-naphthol (BOH), secobarbital, pentobarbital and temazepam. However, the PEM

coating procedure used in the achiral studies needed to be modified in order to achieve

chiral separations. Optimal conditions were established by varying the additive (sodium

chloride, 1-ethyl-3-methyl-1H-imidazolium hexafluorophosphate (1E-3MI-HFP), 1-butyl-

3-methylimidazolium tetrafluoroborate (1B-3MI-TFB)) in the polymer deposition

solutions, the salt concentration, the column temperature, and the bilayer number. The

ionic liquids 1E-3MI-HFP and 1B-3MI-TFB were explored for possible enhanced

separations. The addition of either ionic liquid to the polymer deposition solutions

increased the resolution of BNP from 0.83 to 0.88 and 0.90. The effects that ionic liquids

have on separations are not very well understood. Therefore, fluorescence studies should

be performed to better understand the mechanisms that take place in separations when

ionic liquids are used as additives. Reproducibilities were also evaluated by using the RSD

values of the EOF and the first peak (R-(+)-BNP). In all cases, the run-to-run and

capillary-to-capillary RSD values of EOF were less than 0.5%, and the run-to-run RSD

values of the R-(+)-BNP peak were less than 1%. In addition, more than 230 runs were

performed in the PEM coated capillary.

Page 165: Analytical separations using packed and open-tubular

145

The studies described above for both chiral and achiral separations have shown that

PEM-coated capillaries have excellent reproducibilities, remarkable endurance, and strong

stabilities against extreme pH values when used in open tubular capillary

electrochromatography (OT-CEC). In Chapter 5, the stability of the capillary surface was

further investigated after exposure to 0.1 M and 1.0 M NaOH. The multilayer coatings

used for the studies reported in this chapter consisted of two and twenty layer pairs. A

layer pair, i.e. a bilayer, is one layer of a cationic polymer and one layer of an anionic

polymer. PDADMAC was used as the cationic polymer, and the polymeric surfactant poly

(L-SULV) was used as the anionic polymer. The structural changes of these coatings were

monitored using laser scanning confocal microscopy (LSCM) after flushing the capillaries

with NaOH. This technique also allowed a study of the uniformities and discontinuities of

the coatings. The observed structures were discussed in terms of separations using OT-

CEC. In addition, the electropherograms obtained from the chiral separation of BNP in

OT-CEC showed a decrease in selectivity and an increase in electroosmotic mobility after

long exposure to NaOH. To study the stability of the capillary surface, two bilayers were

constructed. The stability was evaluated by computing the change in EOF (degradation

ratio). The degradation ratios obtained after rinsing the PEM-coated capillary with 0.1 M

NaOH for 110 min, and with 1.0 M NaOH for 5 min were all below 2%. However, after

additional 1.0 M NaOH rinse, the degradation ratios were above 2%. Finally, the ability to

recover the capillaries by direct exposure to NaOH was demonstrated. Measurements of

both electroosmotic mobility and selectivity showed that a 2-bilayer and a 20-bilayer PEM

coating could be completely removed from the capillary surface after approximately 3.5

and 9.5 hours, respectively, of continuous exposure to 1.0 M NaOH.

Page 166: Analytical separations using packed and open-tubular

146

The studies discussed earlier verify that the PEM coating has a significant effect on

separations. The next goal is to characterize the coating using various characterization

techniques, such as nuclear magnetic resonance (NMR) spectroscopy, ellipsometry, X-ray

diffraction, Fourier transform infrared (FTIR) spectroscopy, neutron reflectometry,

fluorescence resonance energy transfer (FRET), and near-infrared (NIR) multispectral

imaging spectrometry. Different kinds of microscopic techniques, such as atomic force

microscopy (AFM), scanning electron microscopy (SEM), transmission electron

microscopy (TEM), scanning tunneling microscopy (STM), and near-field scanning optical

microscopy (NSOM) can also be used for the characterization of the coating. In particular,

the NIR multispectral imaging technique, which can simultaneously record the spectral and

spatial information of a sample [1, 2], will be used to study the uniformity and

discontinuity of the coating, and to find the optimized flush times of both PDADMAC and

polymeric surfactant. Then, FRET and some microscopic techniques will be used to

measure the thickness of the PEM coating. FRET is a distance-dependent interaction

between the electronic excited states of two fluorophores. Excitation is transferred from a

donor fluorophore (i.e., fluorescein) to an acceptor fluorophore (i.e.,

tetramethylrhodamine) without emission of a photon. The efficiency of FRET is inversely

proportional to the sixth power of the intermolecular separation.

Although LSCM was able to demonstrate the existence of the coating and the

change in structure after rinsing the capillary with NaOH, it did not provide enough details

regarding the structure of the coating. Therefore, the techniques mentioned above can be

used to provide information about film thickness, interlayer spacings, structure, elemental

content etc [3-14].

Page 167: Analytical separations using packed and open-tubular

147

6.1 References

1. Fischer, M.; Tran, C. D. Anal. Chem. 1999, 71, 2255.

2. Khait, O.; Smirnov, S.; Tran, C. D. Anal. Chem. 2001, 73, 732.

3. Kerimo, J.; Adams, D. M.; Brbara, P. F.; Kaschak, D. M.; Mallouk, T. E. J. Phys. Chem. 1998, 102, 9451.

4. Barberi, R.; Bonvent J. J.; Bartolino, R.; Roeraade, J.; Capelli, L.; Righetti, P. G. J.

Chromatogr. B 1996, 683, 3.

5. Pullen, P. E.; Pesek, J. J.; Matyska, M. T.; Frommer, J. Anal. Chem. 2000, 72, 2751.

6. Mahltig, B.; Müller-Buschbaum, P.; Wolkenhauer, M.; Wunnicke, O.; Wiegand, S.;

Gohy, J.-F.; Jerome, R.; Stamm, M. J. Colloid and Interface Science 2001, 242, 36.

7. Perez-Salas, U. A.; Faucher, K. M.; Majkrzak, C. F.; Berk, N. F.; Krueger, S.; Chakof, E. L. 2003, 19, 7688.

8. Styrkas, D. A.; Bütün, V.; Lu, J. R.; Keddie, J. L.; Armes, S. P. Langmuir 2000, 16,

5980.

9. Oehlke, J.; Birth, P.; Klauschenz, E.; Wiesner, B.; Beyermann, M.; Oksche, A.; Biener, M. Eur. J. Biochem. 2002, 269, 4025.

10. Kharlampieva, E.; Sukhishvili, S. A. Langmuir 2003, 19, 1235.

11. Preisler, J.; Yeung, E. S. Anal. Chem. 1996, 68, 2885.

12. Kaupp, S.; Wätzig, H. J. Chromatogr. A 1997, 781, 55.

13. Kaupp, S.; Steffen, R.; Wätzig, H. J. Chromatogr. A 1996, 744, 93.

14. Advincula, R.; Wang, Y.; Park, M.; Youk, J. H.; Mays, J.; Yang, J.; Kaneko, F.;

Baba, A.; Knoll, W. Polymer Reprints 2001, 42, 281.

Page 168: Analytical separations using packed and open-tubular

148

VITA

Constantina P. Kapnissi was born in Nicosia, Cyprus, on March 22, 1977. She is

the first-born child in the family, and she has three sisters. The fourth one is ten years old.

In June of 1995, she graduated from a public high school. Two months after that she

attended the University of Cyprus. She majored in general chemistry, and she obtained a

Bachelor of Science degree in May, 1999. However, before graduation, in the summer of

1998, she started her undergraduate thesis work at ETH University in Zurich, Switzerland.

After spending a month there, she moved back to Cyprus to continue her work with the

help and direction of her advisor Dr. Epameinondas Leontidis. Her undergraduate thesis

title was “Morphology of Gold Colloids from Cationic Surfactant Solutions.” On July 11,

1999, she got engaged to Andreas Christodoulou, who had already been a student at

Louisiana State University in Baton Rouge, Louisiana. After the engagement, she moved

to United States with him to pursue a degree of Doctor of Philosophy in chemistry at the

same university, under the direction of Dr. Isiah M. Warner. Her dissertation focuses on

the development of packed and open-tubular capillary electrochromatographic methods for

improved achiral and chiral separations of various classes of analytes.

The articles she published, and the patent she submitted from her research work are

listed below:

• Warner, I. M., Kapnissi, C. P., Kamande, M., Valle, B. C., Schlenoff, J. B. Analytical Separations With Polyelectrolyte Layers, Molecular Micelles, or Zwitterionic Polymers, U.S. Patent, Submitted.

• Kapnissi, C. P., Agbaria, R. A., Lowry, M., Geng, L., Warner, I. M. Investigation

of the Stability of Polyelectrolyte Multilayer Coatings Using Laser Scanning Confocal Microscopy and Open-Tubular Capillary Electrochromatography, Submitted for publication.

Page 169: Analytical separations using packed and open-tubular

149

• Kamande, M. W., Zhu, X., Kapnissi-Christodoulou, C. P., Warner, I. M. Chiral Separations Using a Polypeptide, a Polymeric Dipeptide Surfactant, and a Polyelectrolyte Multilayer Coating in Open-Tubular Capillary Electrochromatography, Submitted for publication.

• Kapnissi, C. P., Warner, I. M. Separation of Benzodiazepines by Use of Capillary

Electrochromatography, Journal of Chromatographic Science, 2004, 42, 238-244.

• Zhu, X., Kamande, M. W., Thiam, S., Kapnissi, C. P., Mwongela, S. M., Warner, I. M. Open-Tubular Capillary Electrochromatography/Electrospray Ionization-Mass Spectrometry Using Polymeric Surfactant as a Stationary Phase Coating, Electrophoresis, 2004, 25, 562-568.

• Kapnissi, C. P., Valle, B. C., Warner, I. M. Chiral Separations Using Polymeric

Surfactants and Polyelectrolyte Multilayers in Open-Tubular Capillary Electrochromatography, Analytical Chemistry, 2003, 75, 6097-6104.

• Kapnissi-Christodoulou, C. P., Zhu, X., Warner, I. M. Analytical Separations in

Open-Tubular Capillary Electrochromatography, Electrophoresis, 2003, 24, 3917-3934.

• Kamande, M. W., Kapnissi, C. P., Akbay, C., Zhu, X., Agbaria, R. A., Warner, I.

M. Open-Tubular Capillary Electrochromatography Using a Polymeric Surfactant Coating, Electrophoresis, 2003, 24, 945-951.

• Kapnissi, C. P., Akbay, C., Schlenoff, J. B., Warner, I. M. Analytical Separations

Using Molecular Micelles in Open-Tubular Capillary Electrochromatography, Analytical Chemistry, 2002, 74, 2328-2335.

During her years in graduate school, she presented her work at a number of scientific

conferences including:

• “Chiral Separations Using Polymeric Surfactants and Polyelectrolyte Multilayers in Open-Tubular Capillary Electrochromatography - Morphologic Elucidation of Capillaries Using Laser Scanning Confocal Microscopy” Pittsburgh Conference of Analytical Chemistry and Applied Spectroscopy, Chicago, IL, March 2004. Kapnissi-Christodoulou, C. P., Agbaria, R. A., Lowry, M., Valle, B. C., Geng, L., Warner, I. M.

• “Chiral Separations Using a Polymeric Dipeptide Surfactant in Open-Tubular

Capillary Electrochromatography” Pittsburgh Conference of Analytical Chemistry and Applied Spectroscopy, Chicago, IL, March 2004. Kamande, M. W., Kapnissi-Christodoulou, C. P., Zhu, X., Warner, I. M.

Page 170: Analytical separations using packed and open-tubular

150

• “Chiral Separations Using Polymeric Surfactants and Polyelectrolyte Multilayers in Open-Tubular Capillary Electrochromatography” American Chemical Society Meeting, New York, NY, September 2003. Kapnissi, C. P., Agbaria, R. A., Lowry, M., Valle, B. C., Geng, L., Warner, I. M.

• “Analytical Separations Using Molecular Micelles in Open-Tubular Capillary

Electrochromatography: Morphologic Elucidation of Capillaries Modified by Polyelectrolyte Multilayers” American Chemical Society Meeting, New Orleans, LA, March 2003. Kapnissi, C. P., Agbaria, R. A., Lowry, M., Valle, B. C., Geng, L., Warner, I. M.

• “Separation of Phenols and Benzodiazepines Using Poly (Sodium Undecylenic

Sulfate) and Open-Tubular Capillary Electrochromatography” American Chemical Society Meeting, New Orleans, LA, March 2003. Kamande, M. W., Kapnissi, C. P., Zhu, X., Akbay, C., Agbaria, R. A., Warner, I. M.

• “Morphologic Elucidation of Capillaries Modified by Polyelectrolyte Multilayers

for Analytical Separations” Pittsburgh Conference of Analytical Chemistry and Applied Spectroscopy, Orlando, FL, March 2003. Kapnissi, C. P., Agbaria, R. A., Lowry, M., Valle, B. C., Geng, L., Warner, I. M.

• “Elution Behavior of Hydrophobic Analytes by Use of Various Stationary Phases

in Capillary Electrochromatography” Pittsburgh Conference of Analytical Chemistry and Applied Spectroscopy, Orlando, FL, March 2003. Igbinosun, O. J., Kapnissi, C. P., Kimberly Y. H., Warner, I. M.

• “Open-tubular Capillary Electrochromatography Using a Polymeric Surfactant

Coating” Pittsburgh Conference of Analytical Chemistry and Applied Spectroscopy, Orlando, FL, March 2003. Kamande, M. W., Kapnissi, C. P., Zhu, X., Akbay, C., Agbaria, R. A., Warner, I. M.

• “Open-Tubular Capillary Electrochromatography Using a Polymeric Surfactant

Coating” 16th International Symposium on Microscale Separations and Analysis, San Diego, CA, January 2003. Kamande, M. W., Kapnissi, C. P., Zhu, X., Akbay, C., Warner, I. M.

• “Novel Separations Using Polymeric Surfactants in Open-Tubular Capillary

Electrochromatography” Pittsburgh Conference of Analytical Chemistry and Applied Spectroscopy, New Orleans, LA, March 2002. Kapnissi, C. P., Schlenoff, J. B., Warner, I. M.

• “Separation of Beta-Blockers Using Capillary Electrochromatography” Pittsburgh

Conference of Analytical Chemistry and Applied Spectroscopy, New Orleans, LA, March 2002. Kapnissi, C. P., Igbinosun, O. J., Agbaria, R. A., Warner, I. M.

Page 171: Analytical separations using packed and open-tubular

151

• “Separation of Benzodiazepines by Capillary Electrochromatography” Pittsburgh Conference of Analytical Chemistry and Applied Spectroscopy, New Orleans, LA, March 2001. Kapnissi, C. P., Thiam, S., Warner, I. M.

• “Separation of Benzodiazepines by Capillary Electrochromatography” American

Chemical Society Meeting, New Orleans, LA, December 2000. Kapnissi, C. P., Thiam, S., Warner, I. M.