PLACE IN RETURN BOX to remove this checkout from your record. TO AVOID FINES return on or before date due. MAY BE RECALLED with earlier due date if requested. DATE DUE DATE DUE DATE DUE 2/05 p:/ClRC/DateDue.indd-p.1 CAPILLARY ELECTROPHORESIS FOR THERMODYNAMIC AND KINETIC STUDIES By Carl l. D. Newman A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 2006 ABSTRACT CAPILLARY ELECTROPHORESIS FOR THERMODYNAMIC AND KINETIC STUDIES By Carl l. D. Newman Capillary electrophoresis (CE) is a technique in which the individual components of complex mixtures can be separated based upon their physical characteristics of size and electrical charge. This work details the theoretical and instrumental development, simulation, validation, application, and limitations of capillary electrophoresis for investigating interactions between solute molecules. In order to understand the evolution of zone profiles in reactive CE, 3 series of stochastic (Monte Carlo) simulations were employed. These simulations were used to systematically investigate the effects of equilibrium constant, characteristic reaction lifetime, and electrophoretic mobility difference on zone profiles. Statistical moments were used to calculate the figures of merit velocity, plate height, and asymmetry of each zone profile. It was shown that the equilibrium constants are directly related to zone velocity; whereas, kinetic rate constants are directly related to zone plate height. Using conclusions drawn from the simulation studies, a theoretical description is developed of the effects of reaction on velocity and plate height in capillary electrophoresis. The developed equations are validated with the simulations performed previously. The equations for velocity and plate height are also validated by investigating peptidyI-proline isomerizations that have been studied previously. These isomerizations exemplify first-order reactions and are notorious for slow rate constants. Calculated values for equilibrium constants and rate constants are shown to be in good agreement with previously published vmues. The plate height model is also applied to the reversible, second-order inclusion of dansyl phenylalanine by B-cyclodextrin. Although equilibrium or binding constants with B-cyclodextrin have been reported previously, rate constants have remained elusive. The plate height model is used to calculate equilibrium constants and rate constants from the effects of B—cyclodextrin concentration on velocity and plate height, respectively. The effect of temperature on equilibrium constant is used to calculate molar enthalpy and molar entropy. The calculated thermodynamic quanitities are in agreement with previously published values. However, it was concluded that the rate constants for this reaction are too rapid for accurate evaluation via the plate height model. In light of this unsuccessful application, the influence of the physical properties of the instrumentation and of the solutes on the maximum calculable rate constants is discussed. Finally, avenues of future development are discussed. Potential theoretical developments include expanding the existing theoretical equations to include mathematical descriptions of solute zone evolution for reactions that are far from steady state. Potential instrumental developments include applications of the plate height model investigate solute-stationary phase interactions via chip-based, open-tube capillary electrochromatographic separations. To my family and in memory of my grandparents: George Kraemer Milo Newman Evelyn Newman ‘I uN-n. AKNOWLEDGEMENTS There are so many individuals to whom I owe thanks. My progress has been aided by so many wonderful people in a multitude of ways, that it is difficult to decide where to begin. So, I will attempt to do this semi-chronologically . . . My parents and my family for their love and unwavering support. They were there for me as l vacillated between undergraduate universities and degrees. As I pursued my education from Massachusetts to Arizona to Florida and finally to Michigan under the auspices of higher education, they were a constant connection to home. Moreover, as my interests varied from music to linguistics to anthropology to chemistry, they served as constant reminders of who I am and from where I come. My undergraduate professors: Dr. Walter Birkby, Dr. Glen Doran, and Dr. Thomas \fickers. Their patience and guidance facilitated my growth and independence as a young scientist. Through their tutelage I discovered that scientificresearch can be a viable and rewarding path and was inspired to pursue graduate education. My graduate advisor, Vicki McGuffin. Vlfithout her wisdom, expertise, and guidance this work not be possibie. Our conversations during my tenure at Michigan State University have shaped me into not only a better scientist, but a better author and person as well. I am forever in her debt for the confidence, independence, and intellectual development that I now enjoy. I have immense gratitude and. respect for Vicki personally and professionally and am forever grateful to her. My Ph.D. committee for their efforts, input and guidance. Dr. Merlin Bruening was not only second reader for my dissertation but also brought a level of insight and thoughtfulness that has helped me to gain further understanding of this work. Dr. Gary Blanchard for his moral and intellectual support. His door was always open to me. Dr. Brian Teppen and Dr. Thomas Pinnavaia for their thoughtfulness and commitment to my progress through the Ph.D. program. My friends for the‘emotional and moral support they have provided me over the years. I am grateful to my Americorps teammates who are to me a second family cf loving siblings. I am grateful to Jason Stotter, Nate Lynd, and Sam Howerton. These are cherished friends and companions for drinking, cigar smoking, and philosophizing. Our conversations regularly helped to remind me that chemistry can indeed be fun and that it is possible to remain both a chemist and a child at heart. I am thankful to Melissa Meaney and Amber Hupp for bringing new life and excitement to the group. And I am forever grateful to Haruko Yabe, my friend, confidant, and fiance. Her love, support, and confidence in my abilities nourished and sustained me through times of joy, frustration, and self-doubt. Lastly I thank the many students, faculty, and staff at‘ MSU and within the Department of Chemistry without whom my success would not be possible. Kathy Severin for her advice and assistance with the numerous departmental instruments that were used during the course of my reasearch. Paul Reed, Tom vi Carter, and Tom Atkinson of Computing Services for their valuable'assistance and insight with numerous software, hardware, and network questions. Glen Wesley and Tom Bartlett of the Machine Shop for their craftsmanship and assistance in the fabrication of numerous components for the instrumentation I developed during the course of my research. Scott Sanderson and Dave Cedarstaff of Instrument Services for their invaluable assistance and advice in the assembly and maintenance of the instrumentation I developed. The Graduate School and Department of Chemistry for financial support. Lastly, the Council of Graduate Students (COGS) and the COGS faculty advisor Marylee Davis for facilitating my growth socially and as a member of the University community. During graduate school it is easy to forget that we do not live within departmental vacuums. My time with COGS and the advice from Marylee helped to sustain. within me the philosophy I lived while in Americorps - that we should all do what we can to make each others’ lives a little better. vii ‘I met a traveler from an antique land Who said: "Two vast and trunkless legs of stone Stand in the desert... Near them, on the sand, ~ Half sunk a shattered visage lies, whose frown, And wrinkled lip, and sneer of cold command, Tell that its sculptor well those passions read Which yet survive, stamped on these lifeless things, The‘hand that mocked them and the heart that fed; And on the pedestal these words appear: My name is Ozymandius, King of Kings, Look on my works, ye Mighty, and despair! Nothing beside remains. Round the decay Of that colossal wreck, boundless and bare The lone and level sands stretch far away. Ozymandius, Percy Bysshe Shelley viii TABLE OF CONTENTS LIST OF TABLES LIST OF FIGURES CHAPTER 1 1.1 '1.2 1.3 1.5 INTRODUCTION AND BACKGROUND CAPILLARY ELECTROPHORESIS 1.1.1 1.1.2 Physical Contributions to Separation and Broadening 1.1.3 1.1.4 History 1.1.2.1 Electrical Double-Layer 1.1.2.2. Mass Transport 1.1.2.2.1 Diffusion 1 1 2.2.2 Migration 1 .1 .2.2.2.1 Electroosmotic Migration 1.1.2.2.2.1 Electrophoretic Migration 1.1.2.3 Influence of Electrical Conductivity 1 1.2. 4 Column Contributions to Broadening ..2 4.1 Diffusion ..2 4. 2 Column Geometry 1.1 1.1 1.1..2 4. 3 Thermal Gradients A 1.1.2.5 Extra-Column Contributions to Broadening 1.1.2.5.1 Injection 1.1 .2.5.2 Detection Chemical Contributions to Separation 1.1.3.1 Acid-Base Equilibria 1.1.3.2 Other Equilibria Experimental Methods for Reactive Capillary Electrophoresis 1.1.4.1 Methods for Estimating Equilibrium Constants 1.1.4.2 Methods for Estimating Rate Constants THERMODYNAMIC AND KINETIC THEORY 1.2.1 Thermodynamic Calculations 1.2.2 Kinetic Calculations CONCLUSIONS REFERENCES CHAPTER 2 STOCHASTIC SIMULATIONS OF REACTIVE 2.1 2.2 2.3 CAPILLARY ELECTROPHORESIS INTRODUCTION SIMULATION METHODS 2.2.1 2.2.2 Simulation Conditions Simulation Output and Calculations RESULTS AND DISCUSSION xiii 42 52 53 58 2.3.1 Effects of Rate Constant 58 2.3.2 Effects of Equilibrium Constant 65 2.3.3 Effects of Electrophoretic Mobility Difference 77 2.4 CONCLUSIONS 86 2.5 REFERENCES 88 CHAPTER 3 THEORETICAL PLATE HEIGHT MODEL FOR REACTIVE CAPILLARY ELECTROPHORESIS 3.1 INTRODUCTION 92 3.2 THEORY 93 3.2.1 Development of the Plate Height Model 93 3.2.2 Evaluation of the Plate Height Model 98 3.3 SIMULATION METHODS 111 3.3.1 Simulation Conditions 111 3.3.2 Simulation Output and Calculations 112 3.4 RESULTS AND DISCUSSION 113 3.4.1 Comparison of Simulation and Theory 113 3.4.1 Propagation of Random Error 123 3.5 CONCLUSIONS ' 132 3.6 REFERENCES ' 139 CHAPTER 4 EXPERIMENTAL METHODS 4.1 INTRODUCTION 140 4.2 EXPERIMENTAL SYSTEMS 140 4.2.1 Capillary Preparation 140 4.2.2 Capillary Electrophoresis Systems 141 4.2.2.1 CE System with Diode Array Absorbance Detection System 141 4.2.2.2 CE System with Laser-Induced Fluorescence Detection System 141 4.3 DATA TREATMENT AND ANALYSIS 145 4.3.1 Mathematical Functions 145 4.3.2 Statistical Moments ' ‘149 4.3.3 Figures of Merit 149 4.4 CONCLUSIONS 150 4.5 REFERENCES , 151 CHAPTER 5 THEORETICAL PLATE HEIGHT MODEL FOR THERMODYNAMIC AND KINETIC STUDIES OF REVERSIBLE, FIRST-ORDER REACTIONS 5.1 INTRODUCTION 152 5.2 THEORY 153 5.2.1 Velocity and Equilibrium Constant » 156 5.2.2 Plate Height and Characteristic Reaction Lifetime 156 5.3 . SIMULATION METHODS 157 5.3.2 Simulation Conditions 157 5.4 EXPERIMENTAL METHODS 5.4.1 Chemicals 5.4.2 Experimental System 5.4.3 Data Analysis 5.4.3.1 Theoretical Plate Height Model 5.4.3.2 ChromVlfin 5.5 RESULTS AND DISCUSSION 5.5.1 Simulation Studies 5.5.1.1 Accuracy of Theoretical Plate Height Model and ChromVIfin 5.5:2 Experimental Studies 5.5.2.1 Effects of Electric Field Strength 5.5.2.1.1 Evolution of Reactive Zone Profiles 5.5.2.1.2 Identification of lsomers 5.5.2.1.3 Comparison of Rate Constants 5.5.2.2 Effects of Temperature 5.5.2.2.1 Electrophoretic Mobility 5.5.2.2.2 Equilibrium Constant 5.5.2.221 Molar Enthalpy and Molar Entropy 5.5.2.2.3 Rate Constants f 5.5.2.231 Molar Activation Energy 5.6 CONCLUSIONS 5.7 REFERENCES CHAPTER 6 THEORETICAL PLATE HEIGHT MODEL FOR THERMODYNAMIC AND KINETIC STUDIES OF REVERSIBLE, SECOND-ORDER REACTIONS 6.1 INTRODUCTION 6.2 THEORY 6.2.1Velocity and Equilibrium Constant 6.2.2Plate Height and Rate Constant 6.3 EXPERIMENTAL METHODS 6.3.1Chemicals 6.3.2Experimental System 6.3.3Data Analysis 6.4 RESULTS AND DISCUSSION 6.4.1Velocity and Equilibrium Constant 6.4.1.1 Effects of Electric Field Strength and Concentration on Velocity 6.4.1.2 Effects of Temperature and Electric Field Strength on Equilibrium Constant 6.4.1.2.1 Molar Enthalpy and Molar Entropy 6.4.2Plate Height and Rate Constant 6.4.2.1 Effects of Electric Field Strength and xi 158 158 158 158 158 159 159 159 162 166 166 166 170 170 172 172 . 178 181 181 185 188 190 192 195 197 197 198 198 198 201 201 201 - 201 207 208 208 Concentration on Plate Height 212 6.4.2.2 Effects of Temperature and Electric Field Strength on Rate Constants 216 6.5 CONCLUSIONS 220 6.6 REFERENCES 224 CHAPTER 7 LIMITATIONS OF THE PLATE HEIGHT MODEL 7.1 INTRODUCTION 226 7.2 THEORY 227 7.3 CAPILLARY ELECTROPHORESIS 229 7.3.1 Effect of Electrophoretic Mobility Difference on kmax in Reactive CE 229 7.3.2 Effect of Electric Field Strength on kmax in Reactive CE 232 7.3.3 Effect of Equilibrium Constant on kmax in Reactive CE 232 7.3.4 Effect of Diffusion Coefficient on kmax in Reactive CE 232 7.4 CAPILLARY ELECTROCHROMATOGRAPHY 239 7.4.1 Effect of Electrophoretic Mobility Difference on km.x CEC 240 7.4.2 Effect of Electric Field Strength on kmax in CEC 240 7.4.3 Effect of Equilibrium Constant on kmax in CEC 245 7.4.4 Effect of Diffusion Coefficient on kmax in CEC 248 7.5 CONCLUSIONS 248 CHAPTER 8 CONCLUSIONS AND FUTURE DIRECTIONS 8.1 INTRODUCTION 253 8.2 EVOLUTION OF REACTIVE ZONE PROFILES 254 8.3 THEORETICAL DEVELOPMENT 255 8.4 INSTRUMENT DESIGN 256 8.5 EXPERIMENTAL DESIGN 257 8.6 REFERENCES 258 APPENDIX VALIDATION OF INSTRUMENT CONTROL PROGRAMS AND CAPILLARY ELECTROPHORESIS . POWER SUPPLY . , 260 xii TABLE 1.1: TABLE 5.1: TABLE 5.2: TABLE 5.3: TABLE 6.1: TABLE 6.2: TABLE OF TABLES Reproducibility of velocity, plate height, and skew. " Average and standard deviation calculated from four independent simulations where k. = k, = 1 s" (17 = 0.5 s) with 1x10‘5 s time increment. Other simulation conditions as given in Figure 2.3. b Average and standard deviation calculated from four independent simulations where kg = kr = 1 s’1 (1: = 0.5 s), one each at 1x10‘2, 1x10'3, 1x104, and 1x10'5 5 time increments. Other simulation conditions as given in FIGURE 2.3. 57 Damkohler numbers for simulation profiles in FIGURE 2A. 165 Comparison of Equilibrium Constants and Rate Constants of Isomerization. Forward and reverse rate constants for Ala-Pro and Phe-Pro isomerization at 10 °C calculated by plate height model in the long time regime (Da > 5), ChromWIn in the short time‘regime (Da < 1), and reported values from [11] and [16]. Buffer conditions in this work: 1.0 x 10'2 M NazB4O7 (pH 9.3) Buffer conditions for [11]: 7.0 x 10‘2 M NazB4O7 (pH 9.5) Buffer conditions for [16]: 1.0 x 10'1 M NazB407 (pH 8.4). 171 Thermodynamic and Kinetic Parameters of Isomerization. Thermodynamic changes in molar enthalpy (AH) and molar entropy (AS) calculated from Equation 5 as well as molar activation energy for the forward and reverse reactions (AEF and AE,*) for Ala-Pro and Phe-Pro. 182 Calculated Equilibrium Constants for BCD Inclusion of DnsF. Equilibrium constants for BCD inclusion of DnsF calculated from separations as a function of electric field strength and temperature. 209 Calculated Diffusion Coefficients of Free and Bound DnsF. Diffusion coefficients of free and bound DnsF with (DA and Dc) and without (D'A and DC) corrections for temperature and xiii ionic strength at 283, 285, 288, and 293 K. 213 ‘ TABLE 6.3: Calculated Rate Constants for bCD Inclusion of DnsF. Average rate constants for association (kuvg) and dissociation (kd,a,,g) of SOD inclusion of DnsF calculated from separations at 334, 368, and 418 Vlcm at 283, 285, and 288 K. 219 xiv FIGURE 1.1: FIGURE 1.2: FIGURE 1.3: FIGURE 1.4: FIGURE 1.5: FIGURE 2.1A: FIGURE 2.18: TABLE OF FIGURES Schematic diagram of the electrical double-layer structure with hydrogen ions (H+), buffer ions (+,-), and solvent molecules (open) oriented towards the negatively charged silica surface. The curved, solid line represents the decay of electrical potential from ‘ the silica surface into the bulk solution This is a simplified representation of the surface potential as it intimates a uniform distribution of charge. 5 Effect of limiting ionic conductivity on effective diffusion coeff' cient for |z.|- - 1 (El), |z.|= 2 (O), |zi|= (A), |z.|= -4 (X), and |z.|= 5 (O) Calculation conditions: 0.}: 1 .00x105 cm2,/s e= 78.,5 T= 298.15K, l=0.,02M |z,-|= |zk|=1, ,7211 (Na+)= 510. 11 cm2 0'1 equiv, 1M°(Cl)= 76.340m Q1equiv'19 Effect of limiting ionic conductivity on effective diffusion coefficient for l- - 0 002 M (O), I- - 0. 005 M (A), |=0.,02M(l:l) l=0.005M(X), and5l=0.2M(O). Calculaiton conditions. D.z° = 1. 00 x 10'5 cm 2Is, 8 = 78. 5, T= 298.15 K, |z.|= |z)|= |zk|- - 1, 2A,” 1=(Na+) 50.11 cm2 0'1 equiv, 121° (CI- -=) 76. 34 cm201equiv111 Effect of ionic strength on electrophoretic mobility forlzil1(D) lzil= 2(0) IZiI= (A) lzil= 400 and |z.|= 5 (O). Calculation conditions: p.31“ - 1. 00 x 10" cmNs a=1.0,0x101°m n= 0.997cP, e= 78.,5 T= 298.,15K l=0.02M, |z,-|=1, of (Na+)= 5.19x10‘1 cmst 17 Effect of pH on observed electrophoretic mobility of phosphate. Caalculation conditions: pKa1- =2 23, pKaz= 7.21 pig: =12.,32 po=-288x10 3cm2Ns m- =1. 18 x 10 acmst, 112' — 3. 73 x 10“ cmst, n3 - 0.00 cmst 28 Flow chart of the stochastic simulation of reactive capillary electrophoresis adapted from [55]. 49 Schematic diagram of reactive capillary electrophoresis simulation process. 51 FIGURE 2.2: FIGURE 2.3 FIGURE 2.4A: FIGURE 2.48: FIGURE 2.4C: FIGURE 2.5: FIGURE 2.6A: Illustration of the relationship between the distribution of molecules within the capillary and the resulting zone profile. 54 Simulated CE profiles at multiple detector locations of 2, 4, 6, 8, and 10 cm. Simulation conditions: L = 100 cm; R... = 50 um; N = 2000; t = 1x105 s; om = 1x10‘5 cmzls; v0 = 0.2 cm/s; V = 25 kV; 0A = —1 x 10‘1 cmst; p3=—2x10"1cm;2Ns K- 1 (A) k, k,= 1051,t= 0.058;=(B)kf kr=k151, 1:: 0.5S;=(C)kf 0=.66s'1, t=0.176$;=(D)kf=.033S'1,1:=1.5ZS;(E)kf= 0.1s1, 1:= 5s; (F)k;= k.= 0.06651, 1: 7.;58s (G)k; =k,= 0.033s ,‘t= 15.155 60 Effect of characteristic lifetime 1: on apparent zone velocity (A) at detector locations of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.3. 62 Effect of characteristic lifetime 1: on plate height (B) and skew (C) at detector locations of 2 cm (0), 4 cm (A), 6 cm (CI), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.3. 66 Effect of characteristic lifetime t on plate height (B) and skew (C) at detector locations of 2 cm (0), 4 cm (A), 6 cm (III), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.3. 67 Simulated CE profiles at multiple detector locations of 2, 4, 6, 8. and 10 cm. Simulation conditions: L = 100 cm; Rm = 50 pm; N = 2000; t = 1x10‘5 s; 0,.1 = 1x10‘5 cmzls; v0 = 0.2 cm/s; v = 25 kV; in = —1 x 10"1 cn12/\I;s [13"- —2x10" cmst;1:= 0. 5 s; (A) K= 1000 k;= 1.998s ,=k, 0.00251;(B)K=1100, kf= 1.98 1s" k.= 0.102s (,C)K= 10, in: 1818s k.= 0.";182s (D)K=1k;1s’1,k,=1s'1,(E)K=0.1kf=0.182s'1, k,= 1.818s'1' ,(F)K= 0.,01 k}: 002s" ,k.=1.98s1; (G)K= 0.001, in: 0."002s ,k,=1.998s". 59 Effect of equilibrium constant K on apparent zone velocity (A) at detector locations of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.5. 72 FIGUREZSB: FIGURE 2.6C: FIGURE 2.7: FIGURE 2.8A: FIGURE 2.88: FIGURE 2.8C: FIGURE 3.1: Effect of equilibrium constant K on plate height (B) and skew (C) at'detector locations of 2 cm (<>), 4 cm (A), 6 cm (III), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.5. 75 Effect of equilibrium constant K on plate height (B) and skew (C) at detector locations of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.5. 76 Simulated CE profiles at multiple detector locations of 2, 4, 6, 8, and 10 cm. Simulation conditions: L = 100 cm; R... = 50 pm; N = 2000; t = 1 x 10'5 s; D... = 1 x 10'5 cmzls; v0 = 0.2 cmls; v = 25 W; K = 1; k; = k. = 1 s"; (A) 0A =1.5 x 10‘1 cmst, (is = —4.5 x 10‘1 -cm2Ns; (B) m = 1.0 x 10‘1 cmst, (is = —4.0 x 10*1 cm’Ns; (C) in = 0.5 x 10‘5 cmst, [La = —3.5 x 10*1 ‘ cmst; (D) m = 0.0 x 10*1 cmst, (is = —3.0 x 10‘1 cmst; (E) m = —0.5 x 10‘1 cmst, (is = —2.5 x 10" cmst; (F) m = -1.0 x 10‘1 cmst, (is = —2.0 x 10" cmst. 79 Effect of electrophoretic mobility difference Au. on apparent zone velocity (A) at detector locations of 2 cm (0), 4 cm (A), 6 cm (Cl), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.7. 81 Effect of electrophoretic mobility difference Art on plate height (B) and skew (C) at detector locations of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (O), and 10 cm (X). Simulation conditions as given in Figure 2.7. 84 Effect of electrophoretic mobility difference Au on plate height (8) and skew (C) at detector locations of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (0), and 10 cm (X). Simulation conditions as given in Figure 2.7. 85 Effect of electroosmotic velocity and electric field strength on plate height contributions from diffusion (----) and mass transfer (- -), as well as on total plate height (—). Calculation conditions: (1.0 = 8.00x10“1 cmst, ILA = —1.00 x 10*1 cmst, [.13 = -2.00 x 10‘1 cmst, om = 1.00 x 10'5 cm2/s, K = 1, ice = k8). = 1 s". 1 00 FIGURE 3.2A: FIGURE 3.28: FIGURE 3.2C: FIGURE 3.3: FIGURE 3.4A: Effect of electroosmotic velocity on plate height for RAB =l 3) voltage steps. The digital controls labeled “kV1' - “kV6” and “uA1” - “uA6” allow the user to specify the magnitude of the applied voltage or current. The digital control labeled “T max” allows the user to . specify the maximum time for voltage or current application. The digital indicator labeled “Time” shows the elapsed experiment time. The digital indicator labeled “Output” shows the voltage output to the HVPS. 262 Schematic of the step-logi62 program. The schematic shows the data flow as logical cascade from T1 through T6 to the output channels that interface with the HVPS. 264 Typical screen view of the PUnCC program. The dialog box labeled “Input Comments” allows the user to add comments regarding each experiment to the data file. The toggle switch allows the user to monitor either the current or voltage from the HVPS. The digital controls xxiv FIGURE A.2B: FIGURE A.3: FIGURE A.4: FIGURE A.5: FIGURE A.6: labeled “Buffer size”, “Scan rate", and “Scans to read at time” allow the user to specify the number of scans stored in the data buffer, the number of scans recorded each second, and the number of scans written to file, respectively. The button labeled “New file” allows the user to choose to store the data as a new file or to append the data to an existing file. The digital controls labeled “Input limits” allows the user to specify the board gain. The digital indicators labeled “Scan backlog” and “Time Elapsed” show the number of scans in the buffer and the elapsed time, respectively. The button labeled “STOP” allows the user to stop the experiment. The waveform charts and digital indicators show the real-time response from the detectors and the monitored current or voltage from the HVPS. 266 Schematic of the PUnCC program. The schematic shows the data flow of specified inputs from the screen view to the data acquisition loop and the data file. The schematic also shows the interface to the step-logic2 program as a sub- routine (labeled “V or I Step”) within the data acquisition loop. 268 Programmed voltage versus the output voltage (El) and percent error of the output voltage (X). 271 Programmed voltage versus the measured current (El) and percent error of the measured current (X). 273 Programmed current versus the measured voltage ([3) and percent error of the measured voltage (X). 275 Input voltage versus the voltage measured by PUnCC (III), DMM (A), and percent error of the voltage measured by PUnCC (x ). 278 XXV CHAPTER 1: INTRODUCTION AND BACKGROUND 1.1 CAPILLARY ELECTROPHORESIS 1.1.1 History Electrophoresis is the separation of charged molecules or particles by differential migration in an electric field. The first reported use of electrophoresis as a separation technique was in the 1930s by Tiselius for the separation of constituent proteins from human serum [1,2]. This pioneering work eventually earned Tiselius the NObel Prize in 1948. Electrophoresis in a glass capillary was first reported by Hjerten in 1967 [3], but did not gain popularity until some 14 years later when Jorgenson and Lukacs [4] demonstrated the efficiency of the technique for separating small biological molecules. In the intervening years, capillary electrophoresis (CE) and related techniques have become a staple for fast and efficient separations of charged and neutral solutes. Applications of CE include trace analysis of metals from environmental samples and protein separations for the Human Genome Project [5]. A recent review by lssaq [6] discusses a broad range of solutes for which separation by CE has become standard operating procedure. The separation efficiency of CE is a result of differential and concurrent solute migration within the bulk solution under the influence of an applied electric field. Physical origins of this unique separation mechanism are derived from the formation of an electrical double-layer at the interface of the bulk solution and the capillary walls, combined with the mass transfer response of solute and bulk components in an electric field. Chemical origins of the separation originate from acid-base equilibria, complexation equilibria, and wall adsorption. This introductory chapter discusses some of the origins of the physical and chemical phenomena that contribute to separation in CE. 1.1.2 Physical Contributions to Separation and Broadening - 1 .1 .2.1 Electrical Double-Layer The interaction of a charged solid with an electrolyte solution results inthe formation of an electrical double-layer at the interface [7]. For a fused silica capillary in aqueous solution, the walls carry a negative charge from the deprotonation of surface silanol‘groups. Water molecules orient their dipoles to the charged surface while some cations adsorb to negatively charged sites on the capillary wall [8,9]. Hydrogen ions are bound covalently to the surface but other cations are bound ionically. This organized, initial layer of water molecules and hydrogen ions and other specifically adsorbed ions constitutes the inner Helmholtz plane (IHP). A second, less organized layer of water molecules and electrolyte ions is oriented adjacent to the IHP and is known as the outer Helmholtz plane (OHP). The region bounded by the IHP and the OHP is referred to as the compact portion of the double-layer. Because all of the negative surface charge is not compensated by specific adsorption, some cations are electrostatically attracted towards the OHP. The resulting concentration of cations is thus large close to the OHP but approaches the concentration of the bulk with increasing distance from the capillary wall. This region of decreasing concentration from the OHP and extending into the bulk is referred to as the diffuse portion of the double- Iayer. A schematic diagram of the structure of the electrical double-layer is shown in FIGURE 1.1. The compact and diffuse portions of the double-layer are characterized by an effective thickness known as the Debye length [10,11] 5= 28KBT (1) 1000NA e2l where c is the dielectric constant of the buffer, k3 is Boltzmann’s constant, T is the absolute temperature, NA is Avogadro’s number, e is the charge of an electron, and l is the ionic strength. For small surface potentials, 6 describes the distance (y) into solution at which the surface potential (We) has decayed by ~ 37% [7,11] ‘I’(y)-‘PceXP(-%) (2) ' 1.1.2.2 Mass Transport Transport phenomena may be generally considered as the response of a molecule or ion to an applied gradient. For an electrolyte solution, the response of the concentration of the i111 solute with an electrical charge 2 (Cu) to a gradient may be described mathematically by a . 11?ch - —v in - (3) where Jr; is defined as the flux of the i,z1h species, 4.40.2), <4) X FIGURE 1.1:Schematic diagram of the electrical double-layer structure with hydrogen ions (H+), buffer ions (+,-), and solvent molecules (open) oriented towards the negatively charged silica surface. The curved, solid line represents the decay of electrical potential away from the silica surface and into the bulk solution. This is a simplified representation of the surface potential as it intimates a uniform distribution of charge. ompact C (9+ :1 $5.1“ 9 013,501: 0 .6 223° 38 o O 0 FIGURE 1.1 G +'.'+'. 1'". 29751311 { .11. + . C p] + '0. + 9 + /O H H1" H+ . and the summation index X refers to the diffusion, migration, convection and thermal components of flow. This expression describes the decrease of concentration of i,z in a unit volume with time. Unwanted convective flow can arise in CE from pressure differences between the ends of the column. Unwanted thermal flow can arise from temperature gradients that develop as a result of poor heat dissipation. The use of narrow diameter capillaries can effectively mitigate both of these unwanted contributions to flow. This is because pressure differences and axial temperature gradients decrease with the square of the internal diameter of the capillary. Thus, for electroosmosis in narrow diameter capillaries, thermal and convective contributions to flow are negligible with respect to the contributions from diffusion and migration [8,12]. The effects of temperature on transport are discussed further in a later section. 1.1.2.2.1 Diffusion Diffusion is the mass transport response to a concentration gradient. Flow from diffusion may be described by (le)dif ' ”01.2 V CL2 (5) where Du is the diffusion coefficient of the i,z‘11 species. Diffusion coefficients are a measure of the ability of a molecule to drift randomly through a given medium. Smaller molecules tend to have larger diffusion coefficients than larger molecules because they experience less frictional resistance (viscous drag) to movement. The Stokes-Einstein equation illustrates the dependence of diffusion coefficient at infinite dilution (D550) on molecular size and viscosity [13] D.0 _ kBT 1'2 61mmz (6) where n is the solution viscosity and rLz is the solvated radius of the i,zth species. Diffusion coefficients of electrolytes also depend upon the electrical properties of the medium. The diffusion of ions at near-infinite dilution in an otherwise uniform electrolyte solution has been discussed by Onsager [14]. Onsager’s treatment of the confluence of diffusion and electrical conduction shows that at low total ionic strengths, the-effective diffusion coefficient of an ion in vanishly small amounts is 6 . Di,z -D€z 1-2'—8°—1—’;1i(1-(/q(u;z))z,2~fi (7) (ET-)4 where q(p;z) is a complex function of the mobilities and valencies of the ions present in solution. The function q(pi,z) is simplified to , Izil lzil’ti Izkllit (8) 11811434 rrlxi+lzl>i1lzl2~i+lzklxi where j and k represent ions of the main electrolyte and lo the limiting ionic conductivity. FIGURE 1.2 illustrates the effects of analyte valence (1 < |z;| .< 5) and limiting ionic conductivity (50 < >411 < 150 cm2 0'1 equiv‘1) on Du. It is apparent that Du decreases with increasing 14° and that the extent of deviation from D; 1° is exacerbated as |z;| increases. FIGURE 1.3 illustrates the effects of ionic strength (0.002 < I < 0.2 M) and limiting ionic conduCtivity (50 < 14° < 150 Cm2 0’1 equiv'1) on Du. Once again, it is apparent that Dr,z decreases with increasing 24°. It is also apparent that the extent of deviation from D;,z° is FIGURE 1.2: Effect of limiting ionic conductivity on effective diffusion coefficient for |z;| = 1 (El), |z;| = 2 (O), lzrl = 3 (A), |z;| = 4 (X), and |z.| = 5 (O). Calculation conditions: D;,z° = 1.00 x 10'5 cm2ls, e = 78.5, T = 298.15 K, I = 0.02 M, (2,) = (2..) = 1, 3,9 (Na+) = 50.11 cm2 0" equiv", N? (CH = 76.34 cm2 0" equiv" FIGURE 1.2 of. A7235 76 «ES >._._>_._.ODDZOO 0.20. Oz_._.=>=r_ of. o: oo on — — tr — om ooimoo . comma - cameo - comma r comes ) m . comes a St... cosmos ( . comma .. comes .. 8-8.5 coefficient I.) = 5(0). 8.5, T = cm2 0'1 - moimo._. iNEIIOlzlzlEIOO NOISszllO EIAILOEHdEI FIGURE 1.3: Effect of limiting ionic conductivity on effective diffusion coefficient for l = 0.002 M (O), l = 0.005 M (A), I = 0.02 M (El), l = 0.05 M (X), and l = 0.2 M (O). Calculation conditions: Dr,z° = 1.00x10'5 cm2/s, 5 = 78.5, T = 298.15 K, |z,| = |sz = |zk| = 1, If (Na+) = 50.11 cm2 0" equiv'1, 33° (OH = 76.34 cm2 0'1 equiv'1 10 FIGURE 1.3 p-238 cease E26328 020. 02.52: 02 02 o: 8 on 8 I \ \ I ) ( < < ) > > ‘I x x < < ’\ ’\ < ‘I > ) ) ) x < ’\ J 1.. _I. J J J \l \I .I. .J .I. ..J 1.. “H'H'N'M'I I I I I I I I I I I I I F“— E U 'I'I'i'. .I. .'. .|. .' | l | I . l . l4 s . . . . . . .l . l ‘ ‘ ‘ ‘ ‘ ‘ ‘ ‘ ‘ ‘ ‘ x a 11 J.N3IO|:I:IEIOO NOISn:I:I|CI EIAIiOElea exacerbated as I increases; however, the effect of l on Dr,z is not as great as the effect of |z;|. Thus, the effective diffusion coefficient is more strongly influenced by electronic properties intrinsic to the solute than by those of the bulk solution. 1.1.2.2.2 Migration 1 .1 .2.2.2.1 Electrocsmotic Migration Migration is the mass transport response to an applied external force. In electrophoresis, migration is induced by an electric field. Application of an electric field tangential to the capillary walls induces migration of cations in the diffuse portion of the double layer towards the cathode. The flow of free solvent molecules arises from the minimization of viscous forces within the bulk solution responding to the solvated ion migration. This bulk flow originating from the response to an electric field is referred to as electroosmotic flow (EOF). Ions closest to the capillary wall do not migrate because the frictional force resulting from attraction to the capillary wall is too great. With increasing distance from the wall, the effect of this friction decreases until it is exactly equal to the force of EOF. Beyond this distance, the EOF force becomes increasingly . greater than the retarding friction, such that a plane of shear exists within the diffuse double-layer between the stagnant and moving regimes. The potential drop from the wall to the shear plane is called the zeta potential (2;) and is related to both 0 and EOF velocity [8,15]. Dependence of C on the double-layer thickness is thus a dependence on the ionic strength of the solution. As the ionic strength decreases, fewer charged Species are available in the double-layer to induce sufficient movement of the bulk to overcome. the retarding effects of 12 friction. Consequently, the plane of shear and 8 are extended further from the wall, thereby decreasing the zeta potential and EOF [16]. 1 The velocity of electroosmotic flow (vosm) is a function of the electric field strength (E), the vacuum permittivity (so), the dielectric constant (8), and viscosity (1)) of the double-layer as well as the zeta potential (2;) [15,17] E880: TI (9) VoSm ‘ It is important to note that although t;, e, and n are all dependent upon ionic strength, the f; is the most sensitive to changes in | [11]. This expression for EOF Velocity describes the effects producing and impeding bulk flow. The principal components creating flow are the coupled E, 7;, and 8 terms, while Tl describes the friction inhibiting flow. Thus, the migration of the bulk solution in response to the application of an electric field is determined by the extent to which the ions within the double—layer are influenced by the I applied field, the intermolecular friction within the double-layer, and the distance of the plane of shear from the capillary wall. As the effective double-layer thickness decreases and the potential at the plane of shear increases, the greater will be the force of attraction experienced by solvated electrolytes to an applied electric field, and the greater the EOF velocity. However, this bulk migration velocity is impeded by intermolecular friction of solvent molecules in the double-layer and by electrostatic attraction of ions to the capillary walls. The EOF profile is different from the pressure-driven laminar flow profile observed in liquid chromatography (LC) because it is generated near the capillary walls. The velocity of EOF increases with decreasing electrolyte 13 concentration from the shear plane into the bulk, where it remains fairly uniform. In contrast, laminar flow velocity increases with the square of the distance from the wall to its maximum at the center of the capillary. Therefore, while the laminar flow profile is parabolic, the EOF profile is plug-like. This means that electrophoretic separations facilitate more narrow solute zones than chromatographic separations. 1 .1 .2.2.2.2 Electrophoretic Migration The velocity of the i,z1h species (viz) when acted upon by an electric field is described by viz = plz E (10) where Mu is the electrophoretic mobility of the species [9]. Assuming the volume occupied by the solvated species is spherical, themobility at infinite dilution ((5,211) may be estimated by O Zi e (11) From Equations 10 and 11 it is shown that the rate at which an ion travels in response to an electric field is proportional to the coupling of the applied field to the charge on the ion. However, this movement is inhibited by the drag of the solvated ion and the intermolecular friction between solvent molecules. It is worthy to note that substitution of Equation 11 into Equation 6 yields the well known Einstein relation 11:12 R T 0 D12 - zF (12) 14 ‘.A J.- - Ad ‘5 ‘- Ir .H Similar to diffusion coefficients, electrophoretic mobility is dependent upon temperature and the solution electrical properties. Li et al. have adapted the detailed conductivity theory of Pitts to describe this relationship by [18-20] If 1 (fl 1+ 5.03x109aJ(—1—) i ET I where a is the sum of the solvated radii of the analyte and its counter-ion, and q o 1.40x1062i . - .o- .82_-5_i_ . 8 “12‘2"” int/E)21z (8.0% [1111/91] (13) is a function of the analyte and counter-ion mobilities and valencies . lzil H.412. IZ‘WII Ilili‘i’ulzillriz FIGURE 1.4 illustrates the effects of analyte valence (-1 < z; < -5) and q- : (14> ionic strength (10'3 < l < 10" M) on the variation of In. with ionic strength. Although there is a complicated dependence of [1.1.1 on z; and l (Equation 13), the effect of 2; on the deviation of mi,z from the infinite dilution value is greater than that of l. However, even at z; = -5 the deviation of pi; from (5,211 is still less than 0.01%. Individual solutes within a mixture migrate according to their respective mobilities-under the influence of an electric field, inevitably forming solute zones. Each zone travels at a net velocity (vm) determined by the electroosmotic flow of the buffer electrolyte (v03...) from Equation 9 and the electrophoretic flow of the i111 solute (vi) from Equation 10 15 FIGURE 1.4: Effect of ionic strength on electrophoretic mobility for |z;| = 1 (III), |z;| = 2 (<>), [2;] = 3 (A), |z;| = 4 (X), and |z;| = 5 (0). Calculation conditions: (55° = 1.00 x 10" cm2Ns, a = 1.00 x 10111 m, n = 0.997 cP, e = 78.5, T = 298.15 K, I = 0.02 M, |sz = 1, p,° (Na+) = 5.19x10‘1 cm2Ns 16 as Ieozmmew 0.29 _..o 3.0 50.0 FIGURE 1.4 .f. . . . . . _ . . . . . i . . oo+mood . mormoo._. u u u u u u art u Am mo-m_oo.~ 6 IV 6 o o 6 o Aw. mo-m_oo.m . mormooé a. a c h a. 4.. a. momooo mo..Moo.© > ) > > > ) > r moimoofi 0 o o o O o o to moflooo morwood 17 02111 / (02111 " 21111) X 00L Finally, if it is assumed that solutes have negligible interactions with the capillary walls, the mass transport of the i,zth species along the column (Equation 3) may be simplified to a a a2 —C =v —C +D. —-C. at tz net 3X tz tza 2 1,2 (16) Equation 16 indicates that the physical contributions to solute transport are dependent upon properties of the individual solute molecules as well as properties of the bulk solution. Each solute will share contributions to transport from the electric field and the viscosity of the buffer solution. However, the diffusion coefficient, charge, and hydrated radius of a solute will contribute uniquely to its movement within the system. 1.1.2.3 Influence of Electrical Conductivity The electric field, and hence the consequential electroosmotic and electrophoretic flow processes, is dependent upon the conductivity (1c) of the medium it traverses. Further, the conductivity of the solution within the capillary is a function of the charge and concentration of ions 0) in the solution [17] (-1: Elam (17) l where the summation includes all buffer and sample ions. Therefore, zones will have contributionsito conductivity from buffer as well as solute ions (K5) and may be expressed as the sum of the buffer and sample conductivities. For neutral solutes, Ki is zero and the electric field in the solute zone will be the same as in the buffer. However, Ki is not zero for charged solutes and the electric field in the zone will vary from the electric field in the buffer to 18 compensate for the variances in conductivity, such that a constant current is maintained [8,21]. Zone conductivities greater than the buffer conductivities will result in a diminution of the electric field in the zone. Conversely, zone conductivities less than the buffer conductivity will result in an amplification of the electric field in the zone. The electric field at any point along the column (x) may therefore be expressed as [8,22]. E -—1-j+l=gog,ic, (18) ’1 xx ax where j is the current density (assumed invariant with column position), xx is the local conductivity, and the summation is the local diffusion of ions between zones. This local electric field strength has direct influence on the electroosmotic (vow), electrophoretic (vi), and net (v.5) velocities, as given in Equations 3, 10, and 16, respectively. The above discussion assumes that the contribution of the capillary surface to the passage of current is negligible. At low pH values this assumption is generally valid, because the charge of the capillary surface is minimized. However, under alkaline conditions it has been suggested that the capillary surface also contributes to conductivity [17]. This is because as the number of ionized silanol groups increases, the ability of the capillary surface to conduct current is also increased. When the conductivity of the solution is uniform along the column, the electric field strength may be assumed independent of column position. Also, because the. electric field is applied along the column axis and the resultant Migration of ions in response to that field is axial, the only sources of radial 19 transport are the result of thermal and concentration gradients as well as migration in response to the charged silica surface. When the applied electric field is sufficiently large (~30 W), the magnitude of these sources of radial transport is negligible with respect to axial migration. Further, while the passage of current through a conductor generates 'heat, the use of narrow diameter capillaries in CE mitigates the generation of convective flow in response to temperature variations in solution. The temperature gradient from the center of the capillary to the walls is described by [23,24] 2 AT- 011° - (19) 16K where clc is the column innerdiameter, K is the thermal conductivity.- of the solution, and Q is the volumetric rate at which heat is generated as described for eleCtrophoretic separations by [23,24] 0 - E2K¢ (20) where K is the conductivity of the buffer electrolyte and 4) is the total column porosity, which is equal to unity for CE. 1.1.2.4 Column Contributions to Broadening The preceding sections discuss mass transfer processes occurring on the column. Although the migration processes discussed in Section 1.1.2.2.2 contribute positively to separation and increase resolution, diffusion and thermal gradients contribute negatively to separation and decrease resolution. Ultimately, separation efficiency is limited by resolution and, hence, by all of the sources that contribute to zone broadening. These sources of broadening can 20 arise either from column processes or from extra-column processes. Extra- column contributions to broadening will be discussed later. Column contributions to broadening of importance to CE include diffusion, thermal gradients, and capillary geometry. 1.1.2.4.1 Diffusion The contribution to broadening from axial diffusion (02.13) is described by the Einstein equation [13] 0%” -20th (21) where our; represents the average distance traveled by a randomly moving molecule. For separation systems, 02.,“ contributes to symmetric broadening of the zone. 1.1.2.4.2 Column Geometry Broadening can also occur in CE as a result of the spatial geometry of the capillary. Kasicka et al. have shown» that capillary coiling can lead to inhomogeneity in the distribution of electrical current density within the capillary [25]. The expression developed to describe the contribution of coiling to zone broadening is 02 _, Lia d3 (22) °°1 165,20, where red is radius of the coil. Equation 22 shows that o2maris negligible when ran >>- dc, as is the case for most conventional-scale CE separations. However, for microchip CE systems, dc can approach the magnitude of r00... thus resulting 21 in significant additional broadening. The critical value of rear. (rem) below which capillary coiling has a greater contribution to zone broadening than diffusion is -.11°t_dc_ (23) 4 D t 1cat 12 Equation 23 shows that for fixed capillary dimensions and migration times, molecules with larger diffusion coefficients are influenced less by coiling than those with smaller diffusion coefficients. 1.1.2.4.3 Thermal Gradients Broadening can also arise in CE from Joule heating. Although a well- known concept, the effects of Joule heating on mass transfer are complicated. A simplified approach is to first consider the contribution of thermal flow to mass transport (le)therm - —C(z Du VT . (24) where D*i,z is the thermal diffusion coefficient of the i,zth species. This means that changes in velocity can arise as a result of thermally induced convective flow from regions of higher temperatures to regions of lower temperatures. Thus, additional zone broadening can occur from convective perturbations caused by temperature variations within the column. Moreover, thermal gradients can lead to additional broadening as a result of thermally induced variations in molecular properties such as mobile-phase diffusion coefficient (Equations 6 and 7) and electrophoretic mobility (Equation 13) as well as bulk properties such as dielectric constant, viscosity, and ionic strength. These potential complications notwithstanding, it has been shown that the effects of Joule heating are not 22 significant contributions to zone broadening (< 5%) when separations are performed in capillaries with internal diameters s 75 pm with electric field strengths 5 2000 Vlcm [26-28]. 1.1.2.5 Extra-Column Contributions to Broadening In addition to mass transfer processes, zone broadening has contributions from extra-column sources. Although these extra-column sources contribute (sometimes significantly) to the measured width of individual zones, they do not contribute positively to the separation. Extra-column sources of zone broadening of importance to CE include broadening from injection and detection. 1.1.2.5.1 Injection The contribution to broadening from a finite injection volume (02“) is approximated by the equation developed by Sternberg for a rectangular injection profile [26,29] '2 2 0.2. -fitJfli‘i (25) '"J 12 12 where Vin; is the volume of the injection zone,1lrn) the injection length and, A the cross-sectional area of the capillary. For hydrodynamic injections, laminar flow is used to introduce the sample via application of a temporary pressure gradient along the length of the capillary. Thus, the length of the hydrodynamic injection zone is described by the Hagen- Poiseuille equation [30] .. (26) ' n] 16 '0 Ltot 23 where AP is the induced pressure gradient, r is the capillary radius, tan; is the injection time, n is the solution viscosity, and LM is the total capillary length. Hydrodynamic injections may be performed by directly applying pressure at the capillary inlet or a vacuum at the capillary outlet. However, these types of injections are usually performed by siphoning action, from which AP can be calculated by AP-pgAh (27) where p is the solution density, 9 is the gravitational acceleration constant, and Ah is the height difference between the inlet and outlet solution levels. For electrokinetic injection, electroomotic and electrophoretic flow are used to introduce the sample via application of a temporary potential gradient along the length of the capillary. Thus, the length of the injection zone is described by the electroosmotic and electrophoretic mobilities and the injection time (“so + Hobs)Vtinj Ltot (28) .linj " 1.1.2.5.2 Detection For any detector, a finite volume is in contact with the transducer for a finite amount of time and the resulting output signal represents the average response. Detectors that monitor flowing systems have unique characteristics ‘ that contribute to zone broadening. These types of detectors (e.g. absorbance and fluorescence) have finite spatial boundaries that define a rectangular window 24 and the response of the detector is uniform along the length of the window (Ider). The broadening from such detectors for a capillary flowcell is described by [29] (29) where 1V6... is the detection volume and dd... is the inner diameter of the flowcell. For most CE applications, detection occurs along the length of the separation capillary, such that, diet is the inner diameter of the column. 1.1.3 Chemical Contributions to Separation 1.1.3.1 Acid-Base Equilibria Many solutes separated in CE are subject to acid-base equilibria. Consequently, the observed electrophoretic mobility (pm) of these solutes is pH dependent and may be described as a function of the solute acid dissociation constant (K,) values and the pH of the buffer solution Ilobs ' Zi‘iz Cli,z (30) where a is the average fraction of solute i molecules existing with charge state z n B" [H*| “i,z - N n (31) 1+ 2 811 [H] n where n is the summation index, N is the total number of exchangeable protons, and 8 is the cumulative acid formation constant N B" - r1 1 (32) n 25 8’1. FIGURE 1.5 illustrates the effect of pH on the observed electrophoretic mobility of phosphate. It is apparent that as pH increases [robs exhibits four plateau-like regions of increasing magnitude connected by three inflection regions. The corresponding plateau regions occur when (rob. is representative of a single, dominant phosphate species. The plateaus at pH < 1, pH ~ 5, pH ~ 10, and pH > 14 represent the mobilities of the triprotic, diprotic, monoprotic, and fully dissociated phosphate species, respectively. The inflections occur when pH is equal to pK.1, pKa2, and pK.‘1 for phosphate. 1.1.3.2 Other Equilibria A logical extension of the effect of acid-base equilibria on (tabs in CE is the effect of other types of chemical interactions solute molecules may undergo during separation. These generally include interactions with the capillary wall or other stationary phase [31-40] and buffer constituents [39-41]. In particular, CE has been used extensively to evaluate second and higher order reactions such as complex formation of proteins with other proteins as well as with DNA and enzymes [42-50]. 1 “The interactions of solute molecules with other mobile-phase molecules are routinely exploited in CE to achieve increased resolution as well as to study the selectivity of those interactions. Investigations of biomolecular interactions are often designed to determine binding constants between the molecules of interest. Affinity capillary electrophoresis, the Hummel—Dreyer method, vacancy affinity capillary electrophoresis, vacancy peak method, and frontal analysis are 26 " FIGURE 1.5: Effect of pH on observed electrophoretic mobility of phosphate. Calculation conditions: pKa1 = 2.23, pKa2 = 7.21, pKa3 = 12.32, [to = -2.88 x 10'3 cm2Ns, [1.1 = 1.18 x 10'3 cm2Ns, p2 = 3.73 x 10*1 cm2Ns, p3 - 0.00 27 FIGURE 1.5 3. N_. or o \I oo+wod . . vormoo - mormo._. r mormm._. .. morwocm r mormmN morwod (SNzwo) AJJ'IIBOW oar/\aasao 28 frequently used to perform these investigations. Busch and coworkers have recently reviewed the general principles and limitations of these methods [51,52]. Similar approaches have been developed for investigations of selective binding to chiral selectors such as cyclodextrins and crown ethers [53-59]. Capillary electrophoresis has also been used to investigate intramolecular processes such as isomerization and denaturation of proteins [60-65]. All of these experimental methods can be described as types of reactive separations. Consequently, the velocity of the reactive zone is dependent upon the extent of reaction (thermodynamic equilibrium constant), whereas zone broadening is dependent upon both the extent and rates of reaction (kinetic rate constant) [66]. Some estimates of equilibrium constants and rate constants have been performed via computer modeling .of electropherograms [49,61,64]; however, most have been Calculated from peak parameters such as peak height, peak area, and migration time [47-50,54-60,63]. 1.1.4 Experimental Methods for Reactive Capillary Electrophoresis 1.4.1 Methods for Estimating Equilibrium Constants Applications of CE typically used to determine equilibrium constants for binding include affinity capillary electrophoresis (ACE), the Hummel-Dreyerl (HD) method, the vacancy peak (VP) method, and vacancy affinity capillary electrophoresis (VACE) [51,52]. The ACE and HD methods are similar experimentally, in that one of the interacting species is added to the buffer and the other is injected. However, the two methods differ in the parameter measured to calculate the equilibrium constant. In ACE, the change in the net 29 electrophoretic mobility is used to calculate the equilibrium constant; on the other hand, the HD method uses the area of the vacancy peak to calculate the equilibrium constant. The VACE and VP methods are also experimentally similar. For these methods, both interacting species are added to the buffer and neat buffer is injected. Likewise, the two methods differ in the parameter measured to calculate the equilibrium constant. In VACE, the change in the net electrophoretic mobility is used to calculate the equilibrium constant, Whereas the VP method uses the area of the vacancy peak to calculate the equilibrium Constant. 1 Although each of the methods discussed above can be used to estimate equilibrium constants, they have different capabilities. Whereas the ACE method is capable of determining only the equilibrium constant, each of the other methods can be used to determine the number of binding sites as well as the equilibrium constant [41]. Accurate determination of equilibrium constants can only be obtained with the HD, VP, and VACE methods when the electrophoretic mobilities of the complex and the species not monitored are equal. Conversely, 1 accurate determination of equilibrium constants can only be obtained with the ACE method when the electrophoretic mobility of each of the reacting species differs from that of the complex [52]. 1.4.2 Methods for Estimating Rate Constants Of the aforementioned methods, ACE has been extended to calculate rate constants by matching experimental and simulated zone shapes [49]. More recently, CE methods have been developed to evaluate rate constants as well as 30 equilibrium constants of reactions represented by zones that are overlapping or connected by a plateau. In nonequilibrium capillary electrophoresis of equilibrium mixtures (NECEEM),-reactants and products are introduced to the column as an equilibrium mixture and separation is achieved with a buffer that contains none of the reacting species. Under these conditions, the observed zones have the general shape of two Gaussian-like zones connected by a plateau. One zone represents the complex, one zone represents the completely dissociated reactant, and the plateau represents the decay of the complex [67,68]. On the other hand, for electrophoretically mediated microanalysis (EMMA), reactants are injected onto the column as discrete zones, often separated by a buffer. Mixing of the zones is achieved by electromigration as both reactants and products are transported to the detector. Because reactants undergo mixing and then separation in EMMA, the reactions are considered irreversible and treated with a Michaelis-Menton model [69,70]. In each of the above methods, experimentally measured parameters such as zone area, height, and migration time are used to calculate binding constants. Additionally, rate constants are evaluated from distorted or splitting zones by matching simulated zones to the experimental data. Moreover, although each of these methods can be used to estimate [rate constants, they are not necessarily complimentary. The EMMA method is only applicable for calculation of Michaelis-Menton constants for irreversible, second- order reactions. Both the ACE and NECEEM methods can be used to calculate rate constants for reversible, second-order reactions. However, calculation of 31 .‘7 rate constants by either method is inherently limited by the accuracy of implicit assumptions of zone shape. Limitations of the ACE and NECEEM methods differ in several important attributes. The accuracy of rate constants calculated by the ACE method is limited by a priori knowledge of equilibrium constant as well as electrophoretic mobilities of the free and bound species. The accuracy of rate constants calculated by the NECEEM method is limited by the implicit assumption that the observed migration time is less than or equal to the rate of dissociation (ts 1/kd). Capillary electrophoresis has also been used to investigate first-order reactions such as protein isomerization [60-64]. For these types of investigations, evaluation of equilibrium constants and rate constants is typically confined to the time regime in which the isomers are detected individually (ts kd). Moreover, the ability to calculate rate constants for isomerization is predicated upon the ability to match simulated electropherograms to experimental data. This means that the accuracy of the rate constants is inherently limited not only by the mathematical models used to generate the simulated data, but also by the computational methods employed to perform the requisite calculations. 1.2 THERMODYNAMIC AND KINETIC THEORY 1.2.1 Thermodynamic Calculations The effect of temperature on equilibrium constant is described by standard Gibbsian thermodynamic equations. By measuring the equilibrium constants (K) 32 ‘1' at different temperatures (T). the changes in molar free energy (AG), molar enthalpy (AH) and molar entropy (AS) can be calculated via [13] -AG I K - — n RT (33) AG - AH — TAS (34) where, R is the molar gas constant. Thus a plot of In K vs 1/T is used to evaluate AH from the slope and AS from the intercept. 1.2.2 Kinetic Calculations The effect of temperature on rate constants is described by the classical Arrhenius model (Equation 35) or by the Eyring model (Equation 36) [13]. While the underlying principles behind these two approaches are related, there are some fundamental differences that distinguish them. Whereas both assume the presence of a high-energy transition state, the equations developed by Eyring presume that the transition state is an activated complex that exists in a quasi- thermodynamic equilibrium with the reactants and products. Additionally, the actiVation energy (AE5) in the Arrhenius expression is most closely related to the internal free energy, whereas the activation free energy (AG2) in the Eyring expression is representative of the free energy of the system. Finally, the Arrhenius expression assumes that the pre-exponential factor (In A) is independent of temperature. By measuring rate constants (k) at different temperatures, the activation energy (AE5) is calculated via the Arrhenius equafion :t Ink - lnA—é%— (35) 33 Altematively,'the molar activation enthalpy (AH*) and molar activation entropy (A81) are calculated via the Eyring equation k h —AG* l a: r(ken) RT (36) A61 . AH2 — TASt (37) where h is the Planck constant, v is the transmission coefficient, and kg is the Boltzmann constant. A plot of In k vs 1/T is used to evaluate AEJ from the slope and A from the intercept of Equation 35. Similarly a plot of In (le) vs 1/T is used to evaluate AHt from the slope and AS‘ from the intercept of Equation 36. In addition to the assumptions in the Eyring equation discussed above, this application of transition state theory presumes that the kinetic pathways from each of the ground states to the transition state represent elementary steps. Moreover, this application of transition state theory assumes that all molecules ' that achieve the transition state cross the high-energy barrier (v = 1), and that a Maxwell-Boltzmann distribution exists between each of the ground states and the transition state (kgT/h) [71-73]. However, these latter assumptions introduce significant potential error into the slope, such that little certainty should be assigned to calculated values for As. 1.3 CONCLUSIONS ' Although CE does not share the historical depth of liquid and gas chromatography, it has achieved prominence as a standard practice for analytical separations. Much of this success is attributable to the rapid advancement of genomics, proteomics, and biotechnology fields. The preceding sections illustrate physical and chemical foundations that contribute to processes of separation and broadening in CE. Additionally, it is the unique separation mechanism of CE that enables the calculation of true thermodynamic equilibrium Constants as well as true kinetic rate constants for molecular interactions. The work described herein provides a physical and mathematical description of how the thermodynamic and kinetic properties of reaction influence zone shape in CE. Chapters 2 and 3 constitute the simulation and theoretical development for evaluating thermodynamic and kinetic parameters in reactive ‘ CE. In Chapter 2, stochastic (Mente Carlo) simulations are used to illustrate the effects of thermodynamic equilibrium constant, characteristic reaction lifetime, and electrophoretic mobility on the shape, velocity, plate height, and asymmetry of reactive zones. In Chapter 3, the mathematical relationships between reaction and zone broadening in CE are developed and evaluated under the rubric of Giddings’ theoretical plate height model. Chapter 4 describes the experimental methods used to evaluate the thermodynamic and kinetic parameters of reversible, first-order and second-order reactions with the plate'height model developed in Chapter 3. Chapters 5 and 6 constitute the experimental validation and application of the plate height model. Chapter 5 illustrates the experimental validation of the plate height model by thermodynamic and kinetic evaluation of , previously investigated reversible, first-order reactiOns that are known to be kinetically slow. Chapter 6 illustrates an attempt to apply the plate height model to a reversible, second-order system that was not previously kinetically 35 characterized. In Chapter 7, the physical limitations of the plate height model are explored with respect to the effects of electrophoretic mobility, diffusion coefficient, equilibrium constant, and electric field strength on the maximum. calculable rate constants. Chapter 8 summarizes the results and implications of the simulation, theoretical, and experimental work discussed herein as well as potential future applications of the plate height model for reactive CE. 36 1O 11 12 13 14 15 16 1 .5 REFERENCES Tiselius, A. Thesis: Nova Acta Regiae Societatis Scientiarum Uppsaliensis, ser. IV, vol 7, Almqvist & Wlksell, Upsala, Sweden, 1930 Tiselius, A. Trans. Faraday Soc. 1937, 33, 524 Hjerten, S. Chromatogr. Rev. 1967, 9, 122 Jorgenson, J. W.; Lukacs, K. D. Anal. Chem. 1981, 53, 1298 Fung, Y.-S.; Tung, H.-S. Electrophoresis, 1999, 20, 1832 lssaq, H. J. Electrophoresis, 2000, 21, 1921 Electrical Phenomena at Interfaces: Fundamentals, Measurements, and Applications, 2nd ed., Ohshima, H. and Furusawa, K. (Ed.), Marcel Dekker, Inc., New York, NY, 1998 Electrophoresis: A Survey of Techniques and Applications, Part A, Chapter 1, Deyl, 2. (Ed.), Elsevier Scientific Publishing Company, New York, NY, 1979 Capillary Electrophoresis of Small Molecules and Ions Jandik, P. and Bonn, G. (Eds), VCH Publishers, New York, NY, 1993 Banholczer, A.; Pyell, U. J. Chromatogr. A 2000, 869, 363 Tavares, M. ‘F. M.; McGuffin, V. L. Anal. Chem. 1995, 67, 3687 Porra's, S. P.; Marziali, E.; Gas, 3.; Kenndler, E. Electrophoresis 2003, 24, 1 553 . Atkins, P. W. Physical Chemistry, 6th Ed, Freeman: New York, 1998 Robinson, R. A.; Stokes, R. H. Electrolyte Solutions, The Measurement and Interpretation of Conductance, Chemical Potential and Diffusion in Solutions of Simple Electrolytes, Butterworths Scientific Publications: London,1955 ' Colon, L. A.; Guo, Y.; Fermier, A. Anal. Chem. 1997, 69, 461A Colon, L. A.; Burgos, G.; Maloney, T. D.; Cintron, J. M.; Roderiguez, R. L. Electrophoresis, 2000, 21, 3965 37 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 35 36 McGuffin, V. L.; Tavares, M. F. M. Anal. Chem. 1997, 69, 152 Li, D.; Fu, 8.; Lucy, C. A. Anal. Chem. 1999, 71, 687 Pitts, E. Proc. R. Soc. A 1953, 217, 43 Pitts, E; Tabor, B. E; Daly, J. Trans. Faraday Soc. 1970, 66, 693 Mikkers, F. E. P.; Everaerts, F. M.; Verheggen, P. E. M. J. Chromatogr. 1979, 169, 1 Dose, E. V.; Guiochon, G. A. Anal. Chem. 1991, 63, 1063 Knox J. H. Chromatographia 1988, 26, 329 Knox, J. H.; Grant, I. H. Chromatographia 1987, 24, 135 Kasicka, V.; Prusik, Z.; Gas, 3.; Stedry, M. Electrophoresis 1995, 16, 2034 Palonen, S.; Jussila, M.; Porras, S. P.; Riekkola, M.-L. Electrophoresis _ 2004, 25, 344 Grushka, E.; McCormick, R. M.; Kirkland, J. J. Anal. Chem. 1989, 61, 241 Knox, J. H. Chromatographia1988, 26, 329 Sternberg, J. C. in Advances in Chromatography, Vol. 2, pp 205-270; Giddings, J. 0.; Keller, R. A., Eds.; Marcel Dekker, Inc.: New York, 1966 Giddings, J. C. Unified Separation Science; Wiley-lnterscience: New York, 1991 Shariff, K.; Ghosal, S. Anal. Chim. Acta 2004, 517, 87 Ghosal, S. J. Fluid Mech. 2003, 491, 285 Stastna, M.; Radko, S. P.; Chrambach, A. Electrophoresis 2001, 22, 66 Tsukagoshi, K.; Hashimoto, M.; Otsuka, M.; Nakajima, R.; Kondo, K. Bull. Chem. Soc. Japan 1998, 71,2831 Shao, X.; Shen, Y.; Lee, M.; O’Neill, K. J. Chromatogr. A 1998, 809, 173 Schomburg, G. Chem. Anal. 1998, 146, 481 38 37 38 39 40 41 42 43 45 '46 47 48 49 50 51 52 53 Ermakov, S. V.; Zhukov, M. Y.; Capelli, L.; Righetti, P. G. J. Chromatogr. A 1995, 699, 297 Phinney, K. W.; Sander, L. C. Chirality 2005, 17, S65 Corrandini, D.; Cogliandro, E.; D’Alessandro, L.; Nicoletti, l. J. Chromatogr. A 2003, 1013, 221 Okafo, G. N.; Vinther, A.; Komfelt, T.; Camilleri, P. Electrophoresis 1995, 16, 1 91 7 Busch, M. H. A.; Boelens, H. F. M.; Kraak, J. C.; Poppe, H. J. Chromatogr. A 1997, 775, 313 Heegaard, N. H. Electrophoresis 2003, 24, 3879 Heegaard, N. H.; Kennedy, R. T. Electrophoresis 1999, 20, 3122 Hamdan, I. |.; Skellern, G. G.; Waigh, R. D. Nucleic Acids Res. 1998, 26, 3053 Chu, Y.-H.; Cheng, C. C. Cellular and Molecular Life Sciences 1998, 54, 663 McKeon, J.; Holland, L. A. Electrophoresis 2004, 25, 1243 German, l.; Buchanan, D. 0.; Kennedy, R. T. Anal. Chem. 1998, 70,4540 Chu, H.Y.; Lees, W. J.; Stassinopoulos, A.; Walsh, C. T. Biochemistry 1994, 33, 10616 Avilla, L. 2.; Chu, Y.—H.; Blossey, E. C.; Whitesides, G. M. J. Med. Chem. 1993, 36, 126 ’ Mammen, M.; Gomez, F. A.; Whitesides, G. M. Anal. Chem. 1995, 67, 3526 . Busch, M. H. A.; Carels, L. B.; Boelens, H. F. M.; Kraak, J. C.; Poppe, H. J. Chromatogr. A1997, 777, 311 Busch, M. H. A.; Kraak, J. C.; Poppe, H. J. Chromatogr. A 1997, 777, 329 Kuhn, R.; Erni, .F.; _Bereuter, T.; Hausler, J. Anal. Chem. 1992, 64, 2815 Penn, S. G.; Bergstrom, E. T.; Goodall, D. M.; Loran, J. S. Anal. Chem. 1994, 66, 2866 39 55 56 57 58 59 60 61 62‘ 63 65 66 67 68 69 7O 71 72 Schaeper, J. P.; Fox, S. B.; M. Sepaniak, M. J. J. Chromatogr. Sci. 2001, 39, 411 . Guillaume, Y. C.; Peyrin, E. Anal. Chem. 1999, 71, 2046 Sabbah, S.; Scriba, G. K. E. Electrophoresis 2001, 22, 1385 Hoffman, J.; Mechref, Y.; Cassely, A.; Novotny, M. V. Electrophoresis 2003, 24, 2914 Fang, N.; Ting, E.; Chen, D. D. Y. Anal. Chem. 2004, 76, 1708 Rochu, D.; Ducret, G.; Ribes, F .; Vanin, S.; Masson, P. Electrophoresis 1999, 20, 1586 Moore, A. W.; Jorgenson, J. W. Anal. Chem. 1995, 67, 3464 ’Ma, S.; Kalman, F.; Kalman, A.; Thunecke, F.; Horvath, Cs. J. Chromatogr. A 1995, 716, 167 Thunecke, F.; Kalman, A.; Kalman, F.; Ma, S.; Rathore, A. S.; Horvath, Cs. J. Chromatogr. A 1996, 744, 259 Schoetz, G.; Trapp, 0.; Schurig, V. Electrophoresis 2001, 22,2409 Rathore, A. S.; Horvath, Cs. Electrophoresis 1997, 18, 2935 Newman, C. I. D.; McGuffin, V. L. Electrophoresis 2005, 26, 537 Berezovski, M.; Krylov, S. N. J. Am. Chem. Soc. 2002, 124, 13674 Okhonin, V.; Krylova, S. M.; Krylov, S. N. Anal. Chem. 2004, 76, 1507 Harmon, B. J.; Patterson, D. H.; Regnier, F. E. Anal. Chem. 1993, 65, 2655 Patterson, D. H.; Harmon, B. J.; Regnier, F. E. J. Chromatogr. A 1996, 732, 119 Eyring, H. Chem. Rev. 1935, 17,65 Steinfeld, J. I.; Francisco, J. 8.; Hase, W. L. Chemical Kinetics and Dynamics, Prentice Hall, Englewood Cliffs, NJ, 1989 40 73 Benson, S. W. The Foundations of Chemical Kinetics, McGraw Hill, New York, 1960 41 CHAPTER 2: STOCHASTIC SIMULATIONS OF REACTIVE CAPILLARY ELECTROPHORESIS 2.1 INTRODUCTION Although theoretical models have been widely developed for reactive separations in chromatography, there have been relatively few advances in theoretical understanding of reactive capillary electrophoresis. The models that have been develOped generally fall into one of two categories: those that solve mass balance equations and those that approach the system stochastically. Equations of mass balance can be solved analytically or by numerical methods to generate zone profiles in distance or time with the corresponding statistical moments. These models have been reviewed by Jeng and Langer [1]. One common assumption is that mass transfer processes are at equilibrium and the system response is. limited by reaction processes [26]. Conversely, another common assumption is that reaction processes are at equilibrium and the system response is limited by mass transfer processes [6-8]. Other mass balance models treat mass transfer and reaction process explicitly, but assume that axial dispersion is negligible or that it can be modeled independently of the reaction process by an effective dispersion term [8-12]. Mass balance models of transport without reaction in electrophoretic systems have been developed to address specific modes of separation, including capillary zone electrophoresis (CE) [13-18], moving boundary electrophoresis [19], isotachophoresis [17,20—24], and isoelectric focusing [2529], as well as unified models of electrophoretic 42 separations [30-33]. Mass balance models of transport with reaction in electrophoretic systems have beendeveloped to investigate irreversible enzyme- substrate interactions [34] and reversible interactions including isomerization, enantiomerization, and complexation [3538]. Unlike the mass balance models, stochastic models do not require the solution of complex differential equations and the associated simplifying assumptions. Rather, stochastic models utilize the fundamental equations of motion for diffusion, convection, and other transport processes. The solutions to these equations of motion are less complex than the solutions to equations for mass and charge balance and require few, if any, simplifying assumptions. Stochastic. models use the statistical behavior of individual molecules to develop probability distribution functions, from which zone profiles and statistical moments are generated. These models have been extended to Monte Carlo and random walk applications [39-41]. Another stochastic method applies fundamental equations of motion to individual molecules and follows the trajectories of the individual molecules through the system [4247]. The complexity of stochastic simulations can range from. models that are one- dimensional with a fixed simulation time increment [48-50] to models that are three-dimensional with a variable simulation time increment [42,43,46,47,51-53]. The work presented in this chapter uses the three-dimensional stochastic simulation developed by Hopkins and McGuffin [54] and adapted to include reactive separations by Krouskop [55]. The simulation is capable of modeling separation systems in which solute molecules undergo irreversible and reversible 43 reactions with first order and pseudo-first order kinetics. The simulations discussed herein examine the effects of equilibrium constant, kinetic rate constants, and electrophoretic mobility on the shape and statistical moments of zone profiles in reactive capillary electrophoresis separations. Insight gained from these simulations will be valuable to understand and evaluate thermodynamic and kinetic information from experimental data. 2.2 SIMULATION METHODS The simulation program was written in the FORTRAN 90 programming language and executed on a 32-processor Silicon Graphics Origin 3400 computer. The program utilizes algorithms for mass transfer and reaction to simulate the spatial as well as temporal distribution of molecules during separation [44.54.55]. Simulated mass transfer processes include diffusion, convection by laminar and electroosmotic flow, electrophoretic migration, and stationary phase partition or adsorption. Simulated reaction processes include irreversible as well as reversible first order and pseudo-first order kinetic reactions occurring in the mobile and stationary phases. These processes are ’ applied to each molecule at each time increment (t) until the totalsimulation time (T) is reached. At any specified time or spatial position, the molecular distribution and corresponding solute zone profile may be examined and characterized. Simulation of molecular diffusion is performed with a three-dimensional extension of the Einstein-Smoluchowski equation [56,57]. During each time 44 increment (t), the radial distance (p) that each molecule travels is selected randomly from the probability distribution (_ 2 P - —p—— exp J’— (1) P 2 (4ant)V \4Dmt where Dm is the diffusion coefficient of the molecule in the mobile phase. This method of simulating the diffusion process provides a variable diffusion step derived from a normal (Gaussian) distribution in which the direction of travel is randomized in three-dimensional space. This algorithm has been validated over the range of diffusion coefficients from 10‘1 to 10'10 cm2/s [46,54,55]. Molecular convection in the mobile phase occurs from electroosmotic flow. The axial distance (2) that a molecule travels during each. time increment is described by z = vt (2) where v is the linear velocity. The radial velocity profile may be assumed to be flat or may be described by the Rice-Whitehead equation [58] v- vo(1- '°("R) ] (3) 10(me) where R is the radial coordinate of the molecule, Rm is the radius of the mobile phase, x" is the Debye length, and to is the zero-'order modified Bessel function of the first kind. The maximum electroosmotic velocity (v0) may be specified as an input parameter or calculated from the Helmholtz-Smoluchowski equation [59] eCV 41mL (4) V0- 45 where V is the applied voltage, L is the column length, C is the zeta potential of the interface between the mobile and stationary phases, and e and n are the relative permittivity and viscosity of the mobile phase, respectively. The algorithm for electroosmotic convection has been validated for linear velocities over the range of 0.01 to 1.0 cm/s [54]. The electrophoretic migration velocity vp is described by V VI, - EL— (5) where the electrophoretic mobility M of each molecule is corrected for the ionic strength of the mobile phasegby a modified Onsager equation [54,60] :1 =uo -(0.23uOIZI+31.3x10'9 [2])1:fl ' (6) where Ito is the mobility at zero ionic strength and I is the ionic strength. The algorithm for migration has been validated for electrophoretic mobilities over the range of +10‘3 to —10‘3 cm2Ns [54]. First-order and pseudo-first-order irreversible or reversible chemical reactions can be modeled to occur in the mobile phase, the stationary phase, or both. It is possible to simulate systems ranging in complexity from a single irreversible reaction to three sequential reversible reactions with one reversible side reaction [55]; To simulate the reaction process, each molecule, upon entering the reacting phase, is assigned a lifetime 0.) it--i(—1log(x) . (7) 46 where k is the rate constant and X is a random number between zero and unity. Any molecule that remains in the reacting phase for an amount of time equal to or greater than the assigned lifetime is converted to the new species and assigned the appropriate diffusion coefficient and electrophoretic mobility, which are used in Equations 1 and 5, respectively. The algorithm for reaction has been . validated for kinetic rate constants over the range of 1.0 x 10'8 to 1.0 x 1010 s’1 and for equilibrium constants over the range of 1.0 x 10'18 to 1.0 x 1018 [55] FIGURE 2.1A is a flow chart of the stochastic simulation program. Input parameters for the program fall into three general categories: system parameters, molecular parameters, and simulation parameters. System parameters include capillary dimensions, applied voltage, detection locations, and electroosmotic velocity. Molecular parameters include the number and charge of solute species, electrophoretic mobility, diffusion coefficient, rate constant, and equilibrium constant. Simulation parameters include the number of molecules, simulation time step, total simulation time, and the frequency at which data are written to file. FIGURE 2.18 is a schematic representation of the simulated processes of migration, diffusion, and reaction for individual molecules in reactive capillary electrophoresis. Open and shaded spheres represent the individual reacting species. Each species exhibits a unique electrophoretic mobility, mobile phase diffusion coefficient, and rate constant. Each sphere represents a random, 3- dimensional diffusion step. The solid line extending from the center of each sphere to the edge represents the magnitude and direction of each diffusion step. 47 FIGURE 2.1A: Flow chart of the stochastic simulation of reactive capillary electrophoresis adapted from [55]. 48 FIGURE 2.1A INPUT PARAMETERS I INITIAL SPATIAL DISTRIBUTION I r—v TOTAL TIME I MOLECULE NUMBER I MOLECULE REACTS? YES NO REACTION V DIFFUSION 4 I ELECTROOSMOTIC FLOW I * ~ $ MOLECULAR DISTRIBUTION STATISTICAL MOMENTS ELECTROPHORETIC FLOW T I MOLECULE DETECTED? ‘ l YES NO TEMPORAL DISTRIBUTION INCREMENT MOLECULE H— l YES ' OUTPUT INTERVAL? No] INCREMENT TOTAL TIME I . END OUTPUT 49 FIGURE 2.1B: Schematic diagram Of reactive capillary electrophoresis simulation processes. 50 FIGURE 2.1B A 2,2“. 51 The dashed lines extending between spheres represent the cumulative convection and migration steps. 2.2.1 Simulation Conditions Capillary electrophoresis separations are simulated for the reversible reaction A B (8) The physical relevance of this type Of reaction extends not only to first-order reactions but also to pseudo-first-order reactions. lsomerization and enantiomerization are examples Of chemically interesting, reversible first-order reactions that can be described by these simulations. More broadly, reactions normally considered to be second order, such as complex formation, host-guest inclusion, and monoprotic acid-base equilibria, can be modeled as pseudo-first- order reactions when one of the reactants (i.e. ligand, host, hydrogen ion) is in sufficient excess as to be assumed constant. More complex reactions, such as polyprotic acid-base equilibria, are a natural extension of this work with the existing simulation algorithms. For the simulations described in this chapter, the kinetic rate constants (k), equilibrium constant (K), and electrophoretic mobility of species A (UA).and B ([13) are varied. Each simulation begins with 2000 molecules distributed as species A and B in equilibrium proportions. The processes of diffusion, convection, electrophoretic migration, and reaction are then applied to each molecule in time increments of 1.0 x 10'5 s. Separation occurs within a 100 cm x 100 um id 52 capillary column with an applied potential Of 25 kV, an electroosmotic velocity of 0.2 cm/s, and a flat flow profile. The resultant time distribution Of molecules is written to the output file at distances Of 2, 4, 6, 8, and 10 cm along the column. 2.2.2 Simulation Output and Calculations Output arrays are used to generate zone profiles and calculate statistical moments at each of the fixed detector locations along the column. Zone profiles at each detector are generated by binning the temporal distribution of molecules that reach each detector. FIGURE 2.2 illustrates the relationship between the temporal distribution of molecules within the capillary at 10 cm (top) and the representative zone profile (bottom). The first statistical moment (M1) describes the mean, the second statistical moment (M2) describes the variance, and the third statistical moment (M3) describes the asymmetry of the solute zone. The statistical moments are calculated as .gti . M1(t) - EN— (9) N .3 (ti _ MI)n Mn(t) - FLT— (10) where N is the total number of molecules and t; the time for each molecule to reach a specified detector location. These statistical moments are then used to calculate the Damkohler number as well as figures of merit such as velocity, plate height, and skew for the zone profiles. The Damkohler nUmber (Da) describes the ratio of total time for reaction to the characteristic lifetime for reaction (1:) [61-63]. 53 FIGURE 2.2: Illustration of the relationship between the distribution of molecules within the capillary and the resulting zone profile. 54 FIGURE 2.2 45 40~ 100100030 (OMNNPF' SH'IFIOS'IOW :IO USGWIIN 55 In O 60.9 61 .2 61 .4 61 .7 61 .9 62.2 62.4 60.7 TIME (s) In reactive CE, the Damkohler number is given by [64] Da-&(t_) , (11) T where ‘C = 1/(kf + k,). The apparent zone velocity (v) is calculated as v-—z— (12) M, where z is the distance to the detector location. Plate height (H) and skew (S) are calculated by H-I—AZ—z— (13) 3-23. (1.) M? which describe the symmetric and asymmetric broadening about the mean. In order to verify the reproducibility of the calculated figures of merit, replicate simulations have been performed with varying time increments. TABLE 2.1 shows the average and standard deviation for the velocity, plate height, and skew. For simulations with the same time increment (1.0 x 10'5 s), the relative standard deviation is 0.03% for velocity and 4% for plate height, which is consistent with previous results [46,47,55]. The relative standard deviation for skew is quite high (~ 60%) because Of the very small asymmetry of the simulated zone profiles. It is evident that the accuracy and precision are statistically indistinguishable for simulations with different time increments (1.0 x 10'2 s t s 1.0 x10'ss). 56 TABLE 2.1 Reproducibility of Velocity, Plate Height, and Skew of Simulated Zone Profiles. Simulation Time Velocity (cm/s) Plate Height (cm) Skew Increment (5) 10-5 . 1.62x10" 1.06x10'3 6.30x10’2 (:t 4.47x10'5) (:l: 4.22x10'5) (:l: 4.01x10‘2) 10.2 _ 10-5 b 1.62x1 0'1 1.10x10‘3 5.12x10’2 (:1: 6.91x10‘5) (:l: 5.13x10'5) (a: 3.99x10'2) ‘ Average and standard deviation calculated from four independent simulations where kg = k, = 1 s'1 (1: = 0.5 s) with 1.0 x 10'5 s time increment. Other simulation conditions as given in FIGURE 2.3. b Average and standard deviation calculated from four independent simulations where k. = k, = 1 s" (1: = 0.5 s), one each at 1.0 x 10’2, 1.0 x 10'3, 1.0 x 10*, and 1.0 x 10'5 s time increments. Other simulation conditions as given in FIGURE 2.3. 57 2.3 RESULTS AND DISCUSSION 2.3.1 Effects of Rate Constant The effects of kinetic rate constants on zone profiles, velocity, plate height, and skew were studied by varying the forward (kg) and reverse (kg) rate constants from 0.033 to 10 s‘1 for a fixed equilibrium constant Of unity. For each of these simulations, the mobility of species A (MA) was —1.0 x 10“ cm2Ns, the mobility of species B ([13) was —2.0 x 10“ cm2Ns, and the mobile-phase diffusion coefficient for both A and B was 1.0 x 10'5 cm2/s. The zone profiles are shown as a function of kinetic rate constants in FIGURE 2.3. For values Of kg = kr s 0.1 5", two poorly resolved zones are observed at early detector locations and merge into a single, broad zone at later detectors. For values of kg = kr > 0.1 s“, a single zone is observed at all detector locations. It is also apparent that broadening decreases with increasing rate constant and that the broadening is symmetric about a fixed mean time. This dependence of zone profiles on rate constant can be described in terms of the Damkohler number (Da). When the total time (M1) is small with respect to 1: (Da < 1), most molecules have not had sufficient time to react and the initial zones of species A and B migrate separately at their respective velocities [64]. As the total time increases (1 s Da 5 5), molecules spend time traveling as both species A and B and the zones begin to coalesce. The coalescence of reactive zones is manifest as the formation of a plateau between the zones of the individual species. As molecules continue to react and spend more time traveling as both species A 58 FIGURE 2.3: Simulated CE profiles at multiple detector locations of 2, 4, 6, 8, and 10 cm. Simulation conditions: L = 100 cm; Rm = 50 um; N = 2000; t = 1.0 x10'5 s; Dm = 1.0 x10'5 cm2/s; vo = 0.2 cmls; v = 25 kV; 11.. = —1.0 x 10“ cm2Ns; 113 = —2.0 x 10“ cm2Ns; K = 1; (A) kg = k,=10 s", 1 = 0.05 s; (B) k1: k. =1 s‘1,1: = 0.5 s; (C) k. = k, = 0.66 s", 1 = 0.76 s; (D) k1: k. = 0.33 s“, 1 = 1.52 s; (E) kg = k, = 0.1 s“, 1 = 5 s; (F) 11. = k. = 0.066 s", 1 = 7.58 s; (G) k1: k. = 0.033 s",1: =15.15 s. 59 FIGURE 2.3 cm 3 m2: _ _- ..j , 1. Iii-Til 11 141qu 4 4 4 4 4;. 1 .1. s s < < < < < 60 and B, the plateau becomes more prevalent until all of the molecules have converged into a single zone. Finally, when the total time is sufficiently large with respect to ‘5 (Da 2 20), steady state for the reaction is achieved and all of the molecules migrate as a single zone containing the proper proportions of both species A and B. The same trends for evolution of the reactive zones are Observed in FIGURE 2.3 for fixed values of total time and varying rate constant or characteristic lifetime (I). The graph of apparent velocity as a function of characteristic lifetime and distance is shown in FIGURE 2.4A. It is evident that zone velocity is invariant with rate constant and with detector location. Although the zones are only at steady state for rate constants greater than 0.1 s", their velocities are the same for all rate constants. Velocity is independent of rate constant for these profiles becauSe species A and B are initially distributed in equilibrium proportions. These proportions do not change as the reaction progresses, so each species contributes to the zone velocity consistently at each detector location for all rate constants. The graph of plate height as a function of characteristic lifetime and distance is shown in FIGURE 2.48. The plate height is relatively constant with distance for zones at steady-state (kg = kr > 0.1 5"). Conversely, plate height increases with distance for zones not at steady state (kg = .k, s 0.1 s") and eventually converges to the steady-state value. At steady state, the plate height is linearly dependent upon the characteristic lifetime. When 1: is small, all molecules spend the appropriate steady-state proportion of time traveling as 61 FIGURE 2.4: Effect of characteristic lifetime 1? on apparent zone velocity (A) and plate height (B) at detector locations of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (0), and 10 cm (X). Simulation conditions as given in FIGURE 2.3. 62 FIGURE 2.4A oo_. E mzfime: O:w_mm5:._.m“=1_ O_._.w_mw._.0> 1), the velocity is identical for each molecule. Hence, there is no contribution to broadening from differences between the velocities Of species A and B and the plate height is minimized to that resulting from diffusion alone. When molecules spend a significant fraction of time traveling as both speciesA and B, the velocity for each molecule has some statistical variability. This variability causes the plate height to reach a maximum value when K = 1. The graph of skew as a function of equilibrium constant and distance is shown in FIGURE 2.6C. From these data, it is evident that the magnitude of . skew is dependent upon K and upon distance. The magnitude of skew decreases with increasing distance because the symmetric contributions to broadening from diffusion increase throughout the separation. The trends of negative skew (fronting) for values of K > 1 and positive skew (tailing) for values Of K < 1 result because VA is greater than v3. When K > 1, molecules within the steady-state zone travel a greater fraction of time as species B than as species 'A. The manifestation of this temporal distribution Of molecules traveling at different velocities is a leading portion (fronting) to the zone. A following portion (tailing) to the zones arises similarly for K < 1. Local maxima for skew, which occur at approximately K = 0.1 and K = 10, are the result of molecules traveling significant (non-equal) fractions of time as both species A and B. However, skew 73 FIGURE 2.6: Effect of equilibrium constant K on plate height (B) and skew (C) at detector locations Of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (0), and 10 cm (X). Simulation conditions as given in FIGURE 2.5. 74 FIGURE 2.68 000 _. ._.Z<._.wZOO EDEmZEOm. 00w 0_. F _..0 _.0.0 _.00.0 - - _ — . L .. — Foocco 1 _.00.0 (wo) iHOIElH ELLV'Id _.0.0 75 FIGURE 2.60 ._.Z<._.wZOO EDEmEEOw 000—. 00_. 0_. _. _..0 _.0.0 . F000 - Fl - - — — MEDIS 76 approaches zero when molecules travel exclusively as species A (K << 1) or as species B (K >> 1), or when molecules travel equally as both species (K = 1). 2.3.3 Effects of Electrophoretic Mobility Difference The effects Of the difference between the electrophoretic mobility of species A (11A) and species B ((13) on zone profiles, velocity, plate height, and skew were studied by varying Au = [1A — [1.3 between 1x104 and 6x10‘4 cm2Ns for a fixed number average Of (1A and us Of -1.5x10“ cm2Ns. The equilibrium constant for each of these simulations is unity and the kinetic rate constants (kg and k.) are 1 s’1 (t = 0.5 s). The zone profiles are shown as a function of electrophoretic mobility difference (Ag) in FIGURE 2.7. For each value Of Au, the Damkohler number is greater. than 20 at each detector location. Consequently, each simulation is representative of reactions at steady state at all detector locations. From these profiles, it is evident that the peak mean is unchanging, while the extent of symmetric broadening increases with Au. The graph of apparent velocity as a function of electrophoretic mobility difference and distance is shown in FIGURE 2.8A. The velocity is invariant with Au and distance because each of the reactions is at steady state and the equilibrium constant is unity. As discussed for FIGURE 2.6A, the velocity of the steady-state zone is the weighted average Of the velocities of species A and B according to Equation 15. However, the equilibrium constant is unity for each of 77 FIGURE 2.7: Simulated CE profiles at multiple detector locations of 2, 4, 6, 8, and 10 cm. Simulation conditions: L = 100 cm; R1.1 = 50 11111; N = 2000; t = 1.0 x 10'5 s; O... =10 x 10'5 cm2/s; v0 = 0.2 cmls; v = 25 W; K =1;kf= kg=1s'1;(A)l~lA =1.5 x 10“ cm2Ns, pa = —4.5 x 10" cm2Ns; (B) NA = 1.0 x 10" cm2Ns, 113 =—4.0 x 104 cm2Ns; (C) 11. = 0.5 x 10'5 cm2Ns, 113 = —3.5 x 10‘4 cm2Ns; (D) 111 = 0.0 x 10* cm2Ns, 1113 = —3.0 x 10“ cm2Ns; (E) 111 = —0.5 x 104 cm2Ns, 1111 = —2.5 x 104 cm2Ns; (F) 11. = -1.0 x 10‘4 cm2Ns, 118 = —2.0 x 10“ 6111st. 78 FIGURE 2.7 79 40 60 80 TIME (s) 20 FIGURE 2.8: Effect of electrophoretic mobility difference Au on apparent zone velocity (A) at detector locations of 2 cm (0), 4 cm (A), 6 em'(I:I), 6 cm (O), and 10 cm (X). Simulation conditions as given in FIGURE 2.7. 80 FIGURE 2.8A _.00.0 b 62.5502?th E4502 9551.6050me _.000.0 f _..0 N0 (Slum) ALIOOTEIA 81 these simulations, such that the zone velocity at steady state is the number average of VA and v3. Velocity in CE is linearly dependent upon electrophoretic mobility. Thus, because the number average Of (IA and lie was held constant, the number average of VA and vs was also held constant. Therefore, velocity is independent Of Ap. for these simulations strictly because the equilibrium constant is unity. The graph of plate height as a function of electrophoretic mobility difference and distance is shown in FIGURE 2.88. The plate height is invariant with distance because each of these zones is representative Of reactions at steady state. However, plate height increases with Au because of nonequilibrium broadening as discussed fOr the characteristic lifetime (1:) in FIGURE 2.48. It is noteworthy that this broadening is linearly proportional to 1, but proportional to the square of Au. These relationships are apparent from the slopes in FIGURES 2.48 and 2.88. It is also noteworthy that the observed dependence of plate height on 1: and Ap. is closely related to what would be expected from a simple, two-velocity random walk model, wherein the plate height is proportional to 2 (V2 - v1)2 1/ (v1 + v2). As a consequence of this relationship, a greater extent of broadening is observed for greater differences in the velocities of species A and B. The graph of skew as a function of electrophoretic mobility difference and distance is shOwn in FIGURE 2.80. The skew is invariant with distance because each of these zones is representative of reactions at steady state. The skew is 82 FIGURE 2.8: Effect of electrophoretic mobility difference Au on plate height (8) and skew (C) at detector locations of 2 cm (0), 4 cm (A), 6 cm (El), 8 cm (0), and 10 cm (X). Simulation conditions as given in FIGURE 2.7. 83 FIGURE 2.88 AmZNEvaozwmquE >._._1=mO_>_ 0_._.m_m_OIn_Om._.Ow4m _.00.0 _.000.0 1 . 500.0 _.00.0 l ._ O. 0 _..0 (um) J.HOI3H ELLV'Id 84 FIGURE 2.86 AmZNEOV wozmmwnEE >.:.=mO_>_ O_._.mmOIn_Om._10w1_m _.00.0 500.0 F . - p . u - . Fl ' a . 1 H II < 85 MBMS small and independent of Au because species A and 8 are initially diStributed in equilibrium proportions with an equilibrium constant of unity. Thus, there are equal numbers of molecules traveling as species A and B for all values of Au. When the equilibrium constant is not unity, the numbers of molecules traveling as species A and B are not equal and skew is Observed, as discussed in the previous section. 2.4 CONCLUSIONS These simulations show that the effects of reaction in CE are manifest in the shape of zone profiles and can be evaluated via the velocity, plate height, and skew as calculated from the statistical moments of those zones. The zone profiles suggest that molecules introduced as a single zone in equilibrium proportions Of species A and B undergo an initial period of separation prior to reaction and coalesce towards a single zone containing both species as the reaction approaches steady state. Once at steady state, molecules continue to travel as a single zone and the width of that zone is influenced by the amount of time required to achieve steady state as well as the difference between the electrophoretic mobilities Of the reacting species. The figures of merit for these zone profiles demonstrate that the velocity at steady state is dependent upon the equilibrium constant and the individual mobilities, but independent of the kinetic rate constants. The extent of zone broadening is influenced by the equilibrium constant, kinetic rate constants, and mobilities of the reacting species. Plate height increases when not at steady 86 state, remains constant at steady state, and is greatest when the equilibrium constant is unity. Conversely, the skew is greatest for equilibrium constants of 0.1 (tailing) and 10 (fronting), exhibiting an inflection point with no asymmetry when the equilibrium constant is unity. While the equilibrium constant has different effects on plate height and skew, both can be expected to increase with electrophoretic mobility difference. Thus, the zone profiles in reactive CE arise from the confluence of contributions from reaction (1, K) and from separation (11,1 and [13). The significance of these trends is that velocity and plate height are statistically viable avenues for evaluating equilibrium constants and rate constants, respectively, from experimental data. However, because of its poor reproducibility, skew should not be used for any calculations, but only as a means to establish if K is less than or greater than unity. 87 10 11 12 13 14 15 16 . 17 18 19 20 2.5 REFERENCES Jeng, C. Y., Langer, S. H., J. Chromatogr. 1992, 589, 1 Klinkenberg, A., Chem. Eng. Sci. 1961, 15, 255 Magee, E. M., Ind. Eng. Chem. Fundam. 1963, 2, 32 Hattori, T., Murakami, Y., J. Catal. 1968, 12, 166 Schweich, 0., \fIllermaux, J., Ind. Eng. Chem. Fundam. 1976, 17, 1 Binous, H., McCoy, 8., J. Chem. Eng. Sci. 1992, 47,4333 Schweich, D., \fIllermaux, J., Sardin, M., AIChE J. 1980, 26,477 , VIllermaux, J, The Chromatographic Reactor, in Percolation Process: Theory and Applications, Roderigues, A. E., Tondeur, D. Eds., Sijthoff en Noordhoff, Alpen aan den Rijn, The Netherlands, 1981 1 Gore, F. E., Ind. Eng. Proc. Des. Devel. 1967, 6, 10 Kocirik, M., J. Chromatogr. 1967, 30, 459 Chu, C., Tsang, L. C., Ind. Eng. Chem. Proc. Des. Devel. 1971, 10, 47 Carta,G., Mahajan, A. J., Cohen, L. M., Byers, C. H., Chem. Eng. Sci. 1992, 47. 1645 Reijenga, J. C., Kenndler, E., J. Chromatogr. A 1994, 659, 403 Reijenga, J. C., Kenndler, E., J. Chromatogr. A 1994, 659,417 Cann, J. R., Goad, B. W., J. Biol. Chem. 1965, 240, 1162 ‘ Ermakov, S. V., Righetti, P. G., J. Chromatogr. A 1994, 667, 257 Dose, E. V., Guiochon, G. A., Anal. Chem. 1991, 63, 1063 Mikkers, F. E., Everaerts, F. M., Verheggen, T. P. E. M., J. Chromatogr. 1979, 169, 1 Cam, J. R., Goad, B. W., J. Biol. Chem. 1965, 240, 148 Vacik, J., Fidler, V., In Analytical lsotachophoresis, Everaerts, F. M., Ed., Elsvier: Amsterdam, The Netherlands, 1981 ‘88 21 22 23 24 25 26 27 28 29 30 31 32' 33 35 36 37 38 Moore, G. T., J. Chromatogr. 1975, 106, 1 Shimao, K., Electrophoresis 1986, 7, 121 Shimao, K., Electrophoresis 1986, 7, 297 Gebauer, P., Bocek, P., J. Chromatogr. 1984, 299, 321 Cann, J. R., Stimpson, D. I., Biophys. Chem. 1977, 7, 103 Stimpson, D. l., Cann, J. R., Biophys. Chem. 1977, 7, 115 Palusinski, O. A., Allgyer, T. T., Mosher, R. A., Bier, M., Saville, D., Biophys. Chem. 1981, 13, 193 Bier, M., Mosher, R. A., Palusinski, O. A., J. Chromatogr. 1981, 211, 313 Shimao, K., Electrophoresis 1987, 8, 14 Bier, M. Plusinski, O. A., Mosher, R. A., Graham, A., Saville, D. A., In Electrophoresis ’82, Stathokos, 0., Ed., Walter de Gruyter & 00.: Berlin, Germany, 1983 Bier, M., Palusinski, O. A., Mosher, R. A., Saville, o. A., Science 1983, 219.1281 - ’ Bier, M., Palusinski, O. A., AIChE J. 1986, 32, 207 Palusinski, O. A., Graham, A., Mosher, R. A., Bier, M., Saville, D. A., AIChE J. 1986, 32, 215 Patterson, D. H., Harmon, B. J., Regnier, F. E., J. Chromatogr. A 1996, 732, 119 Thunecke, F., Kalman, A., Kalman, F., Ma, S., Rathore, A. S., Horvath, Cs., J. Chromatogr. A. 1996, 744, 259 ' Moore, A. W., Jorgenson, J. W., Anal. Chem. 1995.67, 3464 Trapp, O.. Schurig, V., J. Chromatogr. A 2001, 911, 167 Avila, L. Z., Chu, Y.-H., Blossey, E. C., Whitesides, G. M., J. Med. Chem. 1993, 36, 126 89 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 55 56 Felinger, A., Cavazzini, A., Remelli, M, Dondi, F., Anal. Chem. 1999, 71 , 4472 Cavazzini, A., Remelli, M., Dondi, F., Felinger, A., Anal. Chem. 1999. 71, 3453 Dondi, F., Munari, P.. Remelli, M., Cavazzini, A., Anal. Chem. 2000, 72, 4353 Schure, M. R., Anal. Chem. 1988, 60, 1109 Schure, M. R.. Lenhoff, A. M., Anal. Chem. 1993, 65, 3024 McGuffin. V. L., Wu, P., J. Chromatogr. A 1996, 722, 3 McGuffin, V. L., Krouskop, P. E., Wu, P., J. Chromatogr. A 1998, 828, 37 McGuffin, v. L., Krouskop, P. E., Hopkins, D. L., In Unified ' Chromatography, ACS Symposium Series 748, Parcher, J. F., Chester, T. L. Eds.. American Chemical Society. Washington. DC, 2000 McGuffin, V. L., Electrophoresis 2001, 22, 3709 Weber, S. G., Anal. Chem. 1984, 56, 2104 Giddings, J. C., Dynamics of Chromatography, Marcel Dekker: New York, 1965 Giddings, J. C., J. Chem. Educ. 1958, 35, 588 Betteridge, D., Marczewski, C. 2., Wade, A. P., Anal. Chim. Acta 1984, 165, 227 Wentzell, P. D., Bowridge, M. R., Taylor. E. L., MacDonald, 0., Anal. Chim. Acta 1993, 278, 293 McGuffin, V. L., Wu, R, Hopkins, D. L., In Proceedings of the lntemational Symposium on Chromatography, Hatano. H., Hanai, T., Eds., World Scientific Publishing: River Edge, NJ, 1995 Hopkins, D. L., MycGuffin, V. L., Anal. Chem. 1998, 70, 1066 Krouskop, P. E. Ph.D. Dissertation 2001, Michigan State University Einstein, A., Ann. Phys. 1905, 17, 549 90 57 58 59 60 61 62 63 65 66 Reid, R. C., Prausnitz, J. M.. Sherwood, T. K., The Properties of Gases and Liquids, McGraw-Hill, New York, NY, 1977 Rice, C. L., Whitehead, R., J. Phys. Chem. 1965, 69, 4017 von Smoluchowski, M., Bull. lntem. Acad. Sci. Cracovic, 1903, 1903, 184 Robinson, R. A., Stokes, R. H., Electrolyte Solutions: The Measurement and Interpretation of Conductance, Chemical Potential. and Diffusion in Solutions of Simple Electrolytes, Butterworths, London, England, 1959 Jacobson, J., Melander, W., Gintaras, V., Horvath, C., J. Phys. Chem. 1984, 88, 4536 Thompson, J. D., Carr, P. W., Anal. Chem. 2002, 74, 1017 Kalman, A.. Thunecke, F ., Schmidt, R., Schiller, P. W., Horvath, C., J. Chromatogr. A 1996, 729, 155 Ma', S.. Kalman, F., Kalman, A., Thunecke, F., Horvath, C., J. Chromatogr. A 1995 716,167 Guillaume, Y. C., Peyrin, E., Anal. Chem. 1999.71, 2046 Seals, T. H., Sheng, C., Davis, J. M, Electrophoresis 2001. 22, 1957 91 CHAPTER 3: THE THEORETICAL PLATE HEIGHT MODEL FOR REACTIVE CAPILLARY ELECTROPHORESIS 3.1 INTRODUCTION In Chapter 2, the effects of reaction and separation on zone profiles and figures of merit in reactive capillary electrophoresis were investigated via stochastic (Monte Carlo) simulations. It was shown that velocity and plate height of the reactive zones are reasonable avenues for evaluating equilibrium constant and rate constants, respectively, from experimental data. This chapter details the theoretical development of velocity and plate height for capillary electrophoresis separations In which solute molecules undergo reactive separations. The development is consistent with chromatographic theory and the resulting equations for velocity and plate height are functionally equivalent to Chromatographic analogs. This is not the first attempt to illustrate unifying fundamental concepts of chromatographic and electrophoretic separations. Datta and Kotamarthi have previously used a Taylor dispersion approach to describe the contribution to plate height from electrokinetic broadening in capillary electrophoresis [1]. Plate height expressions have also been developed _by McEldoon and Datta for capillary electrochromatOgraphy in which all contributions to broadening are attributed to diffusion and resistance to mass transfer in the mobile phase [2]. The mobile-phase resistance to mass transfer 92 terms were developed using the generalized dispersion theory of Aris and illustrated a dependence on the Chromatographic retention factor. Rathore and Horvath have applied the chromatographic construct of retention factor to capillary electrophoresis and capillary electrochromatography [3]. However, in electrophoretic systems, this formalism lacks any thermodynamic or kinetic significance and is only useful as a peak locator [3,4]. The equations described in this chapter are developed under the rubric of Giddings’ generalized non-equilibrium theory for chromatography [5] and are validated with an independently developed stochastic (Monte Carlo) simulation [6-11]. These equations show that velocity is directly dependent on the equilibrium constant, whereas plate height is inversely proportional to rate constant. Hence, these equations provide a novel approach to extract thermodynamic and kinetic data from experimental zone profiles in reactive capillary electrophoresis. 3.2 THEORY 3.2.1 Development of the Plate Height Model The migration of solute molecules in electrophoretic systems is influenced by direct interaction with a mobile-phase or stationary-phase additive. Broadening of the solute zone is also influenced by these interactions and may be accurately described by application of the plate height model. While this model has Chromatographic origins, the general premise is equally applicable to electrophoretic separation techniques [1,2]. 93 The van Deemter equation describes plate height (H) as H-A+%+Cv (H where v is mobile-phase velocity in liquid chromatography or the zone velocity in capillary electrophoresis. The contributions to zone broadening from multiple paths (A), longitudinal diffusion (B), and resistance to mass transfer (C) are assumed to be independent and additive. For open-tubular separations like capillary electrophoresis, the contributions to zone broadening are reduced to longitudinal diffusion and resistance to mass transfer. The contribution from longitudinal diffusion is described with the mobile-phase diffusion coefficient (D...) by B - 20... (2) Similarly, the contribution from mass transfer can be expressed in terms of an effective dispersion coefficient (Dc) [5] as c-%%z - (9 In order to describe zone broadening from mass transfer, there must be consideration Of the rates of solute transfer between two or more states, the equilibrium distribution of solute molecules, and the velocity of solute molecules in each state. To begin, a general mass-balance statement is necessary CT-gq (0 where CT is the total solute concentration and C; represents the concentration of solute in state i. At steady state, each possible state will contain a fraction )0 of the total concentration 94 Ci" - CT (5) )0 where the steady-state concentration of i (Cf) is related to C; via a non- equilibrium departure term :1 Ci - cg“(1 + s.) (6) Each positive displacement from equilibrium for one state is balanced by negative displacement from equilibrium for another state. such that ZCIEI = EXISI = 0 (7) I I For the specific case of reaction between two states, "AB A Z 8 (3) I‘BA km and RM are the rate constants for transfer from species A to B and from ' species 8 to A, respectively. The mass-transfer term arising from species A (r11) is equal to the sum of all reaction paths over which the concentration (CA) may be changed rA - CB kBA - CA kAB _ (9) At steady state, the rates of enrichment and depletion of species A are equal. Hence, by substitution of Equation 6 into Equation 9 for CA and CB rA - kAB CA(8B - 8A) (10) The mass-transfer term arising from flow for species A is given by 2 d CA _dCA+V dCA 11 "‘Adzz dt Ach I) 95 where the contribution from diffusion to mass transfer is assumed negligible. At steady state, Equation 11 can be rewritten as ac; dc}; — 12 dz ”A dz ( ) I'A = -V where the steady-state zone velocity (v) is the weighted average Of the individual velocities of species A and 8 (VA and v3, respectively) V=ZXIVi-XAVA+XBVB (13) l Equations 10 and 12 may be combined to relate mass transfer from non- equilibrium to mass transfer from flow, such that kABCM6B-6A) ”(11A -v)d:zA _ (14) Equation 7 may be substituted into Equation 14, recalling that O; is proportional to CT (Equation 5), and followed by rearrangement to yield -xe (VA - v) dIan' .. "XB (VA - V) dln cT .. 15 ' kAB (XA +X3) CIZ kAB CIZ ( ) and by analogy -xA (VB -V) dlnCT a 'XA (VB -V) dInCT (16) EB - kBA (XA + X8) dZ kBA dZ The effective dispersion coefficient (Dc) is equivalent to the weighted average of zone broadening from the concerted effects of non-equilibrium and flow [5] (17) Z D __ . . . Upon substitution of Equations 15 and 16, Equation 17 is equivalent to 96 .. AVXA XS VA _ Avxi XB VB kAB kBA Be (18) where Av = V)», - V.; > 0. Further substitution of R113 = K kBA, where K is the equilibrium constant, results in the simplified expression _AVZXiXB 19 kBA ( ) DC Equations 3 and 19 may then be substituted into Equation 1 to express the contributions to zone broadening from axial diffusion and mass transfer as 20m + 2sz xi x3 _ 20... + 2Av2K H - V "BA V V kBA v(1+ K)3 (20) where 301 = 1/(1 + K), Xe = Kl(1 + K), and D... represents the weighted average of the mobile-phase diffusion coefficients of species A and B (0.1.x and Dma, respectively) Dm-ExiDmI-XADmAHCBDmB (21) I Equation 20 can be rewritten specifically for electrophoretic separations by making the appropriate substitutions for the velocity terms 11- En - 12(11... +9...) ' (22) Av =1 EAp. - E(11A -113) (23) where E is the electric field strength, I1 is the net mobility of the zone, (1..., is the electroosmotic mobility, u... is electrophoretic mobility, (IA and [1.3 are the electrophoretic mobilities of species A and 8, respectively, and Au is the mobility difference between species A and B. Substitution of Equations 22 and 23 into 97 Equation 20 yields the following expression for plate height for electrophoretic separations 20... + 2KAp2E H' 3 E“ kBA11(1+K) (24) The final plate height equation for reactive capillary electrophoresis in Equation 24 has the general form of H = BlE + CE, which is analogous to the chromatographic expression for plate height in Equation 1. It is interesting to note that Equation 20 is a generalized form of plate height from diffusion and mass transfer and is equally applicable to chromatographic as well as electrophoretic separations. Although, the contributions to broadening from multiple paths and Joule heating are neglected in this equation, additional terms may be added where appropriate. It is also noteworthy that the second term in Equation 20 resembles the form of a two-velocity random walk, sztlv, where 1: = mm; + k3..) is the characteristic reaction lifetime. When v3 = 0, the mass- transfer term in Equation 20 reduces to that for retention via single-site adsorption or partition in liquid chromatography [5,12]. Similarly, .when v3 = 0, the mass-transfer term in Equation 24 describes the contribution in capillary electrochromatography from interactions at the wall or at a particle surface. 3.2.2 Evaluation of the Plate Height Model The evaluation of Equations 20 and 24 as a function Of electroosmotic velocity or electric field strength, shown in FIGURE 3.1, is noticeably similar to its Chromatographic analogs. As the electroosmotic velocity or field strength increases, plate height passes from a diffusion-limited region to a mass-transfer 98 FIGURE 3.1: Effect of electroosmotic velocity and electric field strength on I plate height contributions from diffusion («--) and mass transfer (— —), as well as on total plate height (—). Calculation conditions: pea = 8.00 x 10“ cm2Ns, (111 = —1.00 x 10“ cm2Ns, 113 = —2.00 x 104 cm2Ns, o... = 1.00 x 10'5 cm2/s, K = 1, km; = kg). = 1 s". 99 FIGURE 3.1 E62 EOzmmB Sum 2583 82 82 com _ ace 8... com o 6260 LEOO.__m> oFoszOEOmd ed ed to No o o - Bee . «8.0 1 Bee . Bod 000.0 (um) .LHOlE-IH EILV‘ch 100 limited region with a minimum value (Hman), where the contributions of diffusion and mass transfer are equal. The field strength at Hm... (E0...) is determined by calculating the first derivative of H with respect to E and equating it with zero, such that 0... kg... (1+ K)3 K A112 Eopt = (25) Equation 25 illustrates the dependence Of E0... and, hence, Hm... on the competing effects of diffusion and mass transfer. The value of E0... increases with the mobile-phase diffusion coefficient and the rate of transfer from Species B to A. Moreover, plate heights Obtained below E... are linearly dependent upon D... whereas, plate heights. obtained above E0... are inversely proportiOnal to k3... Thus, the magnitude of E required to achieve mass-transfer limited plate heights increases with increasing rate constants. The value of E0... decreases linearly, however, as An increases. Thus. the magnitude of E required to achieve mass- transfer limited plate heights increases with decreasing Au. FIGURE 3.2A is a graph of plate height as a function of electroosmotic velocity for a range of rate constants (kAB = kg). = k, 0.01 s k s 100 s"). From this graph, it is shown that H... decreases whereas E0... increases with increasing k. Also, when H is greater than 5 times the contribution from diffusion (H 2 5 B/E), the plate height increases linearly with electric field strength. Values of plate height within this region are dominated by mass-transfer processes and are inversely proportional to k. It is also shown that the range of mass-transfer- limited plate heights increases with decreasing rate constant and that the range 101 FIGURE 3.2A: Effect of electroosmotic velocity on plate height for kAB = k3). = 0.01 s", 1 = 50 s (— —); k... = k... = 0.1 s". 1 = 5 s (m); k... = k... =1 s",1 = 0.5 s (—); k... = k... =10 s", 1 = 0.05 s (—.—); k... = k... = 100 s", 1 = 0.005 s (— - . —). Calculation conditions: .1,... = 6.00 x 10“ cm2Ns, .1,. = —1.00 x 10“ cm2Ns, .13 = -—2.00 x 10“ cm2Ns, o... = 1.00 x 10'5 cm2ls, K =1. 102 FIGURE 3.2A EEov Eood> { to Se 5 - 50.0 5000.0 500.0 50.0 5.0 _..0 (um) LHOIaH 31in 103 of accessible rate constants increases with increasing electric field strength. This means that the ability to experimentally access a broad range Of rate constants is facilitated by increasing the applied voltage, decreasing the column length, or both. FIGURE 3.28 is a graph of plate height as a function of electroosmotic velocity for a range of mobility differences (2.00 x 10'5 5 Au 5 5.00 x 10“ cm2Ns). From this graph, it is evident that plate height increases with increasing Au. It is also shown that mass-transfer-limited plate heights are achieved at decreasing magnitudes of electric field strength as Au increases. This means that the range of accessible rate constants increases with increasing Au. FIGURE 3.20 is a graph of plate height as a function of electroosmotic velocity for a range of equilibrium constants (0.01 s K s 100). From this graph, it is evident that plate height achieves a maximum value when K is unity. It is also apparent that the plate height is nearly symmetrically distributed for values of K other than unity. Accordingly, the magnitude of K has less effect on the range of accessible rate constants than either electric field strength or mobility difference. FIGURE 3.3 is a graph of velocity as a function of equilibrium constant for a range of mobility differences (2.00 x 10'5 5 Au 5 5.00 x 10" cm2Ns). From this graph, it is evident that when K << 1 the zone velocity is Similar to that of species A, and when K >> 1 the zone velocity is similar to that of species B. It is also . shown that as the magnitude of Au increases, so does the range over which velocity is linearly related to K. This means that the range of accessible equilibrium constants increases with Au. 104 FIGURE 3.28: Effect of electroosmotic velocity on plate height for Ap. = 5.00 x 104 cm2Ns (— —); Au = 2.00 x 10‘4 cm2Ns (----); Au = 1.00 x 10‘4 cm2Ns (——); Au = 5.00 x 10'5 cm2Ns (— -—); Au = 2.00 x 10‘5 cm2Ns (— . - —). Calculation conditions: .11,... = 8.00 x 10“ cm2Ns, (.1. + (13)/2 = -1.50 x 104 cm2Ns. D... = 1.00 x 10'5 cm2/s, K =1. k... = k... =1 s". 105 FIGURE 3.28 Eeov Eood> to Be 50.0 500.0 I ‘-. O (Luv) LHOIaH Elin 106 FIGURE 3.2C: Effect of electroosmotic velocity on plate height for K = 100 (— —); K = 10 (----); K =1 (—); K = 0.1 (—-—); K = 0.01 (— - . —). Calculation conditions: .1... = 8.00 x 10“ cm2Ns, .1.. = —1.00 x 10“ cm2Ns. .13 = —2.00 x 10‘4 cm2Ns, D... = 1.00 x 10'5 cm2/s, kAB = kBA = 1 8-1. 107 FIGURE 3.2C EEov EOO._m_> to :3 Sod 500.0 1 50.0 108 1 50 (um) iHOIEIH Eli‘rfld l ‘—. O IghtforK=100 (= 0.01 (... ' ,5 “As—100x = 1.00 x10 FIGURE 3.3: Effect of equilibrium constant on velocity for Au = 5.00 x 10‘5 cm2Ns, Int/Ito = 2.00 (—); Au = 2.00 x 10'5 cm2Ns, .1./.1.. = 1.25 (— —); Au = 0.00 x 10“ cm2Ns, .1./.1.. = 1.00 (m); Au = 2.00 x 10*1 cm2Ns, .1./.1.. = 0.333 (— - —); Au = 5.00 x 10‘4 cm2Ns, .1./113 = 0.167 (— - - —). Calculation conditions: .1.. = 8.00 x 10“ cm2Ns, .1.. = —1.00 x 104 cm2Ns, E = 250 Vlcm, 0... =1.00 x 10'5 cm2ls, K =1,kAB = k... =1 s". 109 FIGURE 3.3 ._.Z<._.wZOO EDEmZEOM 000V 00.. 0.. _. _..0 5.0 50.0 _ _ _ p . 00.0 I. I I I. I 1 00.0 I! / O/O . 1 0.3.0 /0 IIIIII I I I 1 / I: oo /. / 1 m_..0 0N0 = 5.00x10‘5 cm2Ns, III/P8 ‘ 1.00 1-); All = = 500 X 104 = 250 V/cm, All conditions: [lee = II , E (ser) ALIOO'IEIA 110 3.3 SIMULATION METHODS The derived equations for velocity and plate height (Equations 13 and 24, respectively) are validated with a three-dimensional stochastic (Monte Carlo) simulation developed by Hopkins and McGuffin [10] and adapted to include reactive separations by Krouskop [11]. The simulation program was written in the FORTRAN 90 programming language and executed on a 32-processor Silicon Graphics Origin 3400 computer. The program utilizes algorithms for mass transfer and reaction to simulate the spatial and temporal distribution of molecules during separation [7.10.11]. Simulated mass-transfer processes include diffusion, convection by electroosmotic flow, and electrophoretic migration. Simulated reaction processes include irreversible as well as reversible first order and pseudo-first order kinetic reactions occurring in the mobile phase. These processes are applied to each molecule at each simulation time increment until the total simulation time is reached. At any Specified time or spatial position, the molecular distribution and corresponding solute zone profile may be examined and characterized. The algorithms used to simulate mass transfer and reaction have been described in detail previously in Chapter 2. 3.3.1 Simulation Conditions Capillary electrophoresis separations are simulated for the reversible reaction shown in Equation 8, where the kinetic rate constants (kAn, kBA), equilibrium constant (K), and electrophoretic mobility of species A ((1)1)‘and B ([13) are varied. Each simulation begins with 2000 molecules distributed as species A and B in equilibrium proportions. The processes Of diffusion, convection, 111 electrophoretic migration, and reaction are then applied to each molecule in time increments of 1.00 x 10‘5 s until the total simulation time of 100 s has elapsed. Separation occurs within a 100 cm x 100 um i.d. open-tubular capillary column with fixed electroosmotic velocity of 0.2 cmls and electric field strength of 250 Vlcm. The resultant time distribution Of molecules is written to the output file at distances of 2, 4, 6, 8, and 10 cm along the column. 3.3.2 Simulation Output and Calculations Output arrays are used to generate Zone profiles and calculate statistical moments at each of the fixed detector locations along the column. The statistical moments are calculated as N 2h 11.-.... <29 N _2(t1-M1)" NI" ..= M N ' (27) where N is the total number of molecules and t. is the time for each molecule to reach a specified detector location. The first (M1) and second (M2) statistical moments describe the mean and variance, respectively, of the solute zone. These statistical moments are then used to calculate figures of merit such as Velocity and plate height for the zone profiles. The apparent zone velocity (v) and plate height (H) are calculated as v-_Z_ A . 26 M1 ‘ ( ) 112 H - M (29) where z is the distance to the detector location. 3.4 RESULTS AND DISCUSSION 3.4.1 Comparison of Simulation and Theory Simulations are performed to validate the proposed model and to examine the independent effects of equilibrium constant, kinetic rate constants, and electrOphoretic mobility in reactive capillary electrophoresis [6]. The time distribution of molecules at each detector is used to generate zone profiles and to calculate statistical moments via Equations 26 and 27. The statistical moments are then used to calculate velocity and plate height via Equations 28 and 29, respectively. Simulated and theoretical (with and without diffusion) values of plate height are shown in FIGURE 3.4A as a function of characteristic reaction lifetime (0.005 s 1 5 15.15 s). The data show that as 1: increases, plate height passes from a regime in which it is independent of ‘C (diffusion limited), to a regime in Which it is linearly related to t (mass-transfer limited). Thus, zones representative of kinetically fast reactions are more likely to exhibit diffusion- Iimited plate heights. Conversely, zones representative of kinetically slow I'eactions are more likely to exhibit mass-transfer-limited plate heights. This means that larger electric field strengths are necessary to achieve mass-transfer- - limited plate heights for kinetically fast reactions. 113 FIGURE 3.4A: Effect of characteristic reaction lifetime on plate height from theory without diffusion (----), with diffusion (—), and from simulation (El). Theoretical conditions: V.; = 0.2 cmls; II). = —1.00 x 10*1 cm2Ns; .13 = —2.00 x 104 cm2Ns; E = 250 Vlcm; D... =1.00 x 10‘5 cm2/s; K = 1; 0.033 s" (1 =15.15 s) 5 I1... = kg). 5 10 s" (1 = 0.05 s). Simulation conditions: N = 2000; t = 1.00 x 10'5 s; z = 10 cm; all other conditions same as theoretical conditions. 114 FIGURE 3.4A 00.. 0.. .3 main... zOfiO/mm .. ...0 ..0.0 50.0 ..000.0 ..0000000 [I ..00000.0 5000.0 ..000.0 50.0 5.0 ...0 (we) iHOlElH Elin 115 Simulated and theoretical (with and without diffusion) values of plate height are shown in FIGURE 3.48 as a function of electrophoretic mobility difference (1.00 x 10“ 5 Au 5 6.00 x 10" cm2Ns). The data show that plate height increases with the square of Au, owing to an increase in mass-transfer contributions. Therefore, zones representative of solutes with larger mobility differences between species A and B are more likely to exhibit mass-transfer- limited plate heights. This means that it is important to maximize mobility difference in order to evaluate kinetic rate constants reliably. For complexation reactions, this would equate to maximizing the change in size or charge of the complex. Simulated and theoretical (with and without diffusion) values of plate height are shown in FIGURE 3.4C as a function of equilibrium constant (0.01 s K s 100). The data show that the mass-transfer contribution to plate height decreases as the equilibrium constant deviates from unity. However, the overall effect of equilibrium constant on plate height is small compared to the contributions from rate constant and mobility difference. Whereas mass-transfer- Iimited plate height increases linearly with t and with the square of Au, it increases only by an order of magnitude over the entire range of K. This relatively small contribution from equilibrium constant notwithstanding. more aCcurate evaluation of kinetic rate constants is achieved when K is near unity. The simulated values of velocity with varying rate constant, mobility difference, and equilibrium constant are compared with the theoretical values calculated by using Equation 13 in FIGURE 3.5A. As shown, the theoretical 116 FIGURE 3.48: Effect of mobility difference on plate height from theory without diffusion (----), with diffusion (—), and from simulation (:1). Theoretical conditions: v3 = 0.2 cmls; —1.00 x 10“ cm2Ns 5 Ap. 5 —6.00 x 104 cm2Ns; E = 250 Vlcm; D... = 1.00 . x 10’5 cm2/s; K = 1; k... = k3,. = 1 s'1 (1: = 0.5 s). Simulation conditions: N = 2000; t = 1.00 x 10'5 s; z = 10 cm; all other conditions same as theoretical conditions. 117 FIGURE 3.48 50.0 AmZNEOV mozwmmmnzn. >1_._1__m0_>_ ..000.0 58.0 x 50.0 I - O. 0 ...0 (we) iHOIEIH EILVTd 118 FIGURE 3.4C: Effect of equilibrium constant on plate height from theory without difquion («--), with diffusion (—). and from simulation (III). Theoretical conditions: v0 = 0.2 cmls; u). = —1.00 x 10" cm2Ns; .13 = —2.00 x 10*4 cm2Ns; E = 250 Vlcm; D... = 1.00 x 10'5 cm2/s; 0.001 5 K 5 1000; I113 = k3,. = 1 s'1 (1 = 0.5 s). Simulation conditions: N = 2000; t = 1.00 x 10'5 s; z = 10 cm; all other conditions same as theoretical conditions. 119 FIGURE 3.4C 000.. .8. ezfimzoo 2235358 0.. .. — ...0 ..0.0 50.0 50000.0 5000.0 ..000.0 ..00.0 ..0.0 ...0 (mo) LHeIaH 21in 120 FIGURE 3.5A: Comparison Of theoretical velocity with simulation velocity for simulations with varying equilibrium constant ([1) ‘ and varying mobility difference (A) on the primary y-axis. The diagonal solid line represents a linear relationship between theoretical and simulation velocities. Percent error (X) is shown on the secondary y-axis. The horizontal solid line represents zero percent error. Simulation and theoretical conditions as described in Figures 3.4A and 3.48. 121 FIGURE 3.5A 80883 lNElO8E-Icl LO Q Ill) <1: I I co to <1- ‘7- 5 $9. T'. .—o O 0 0 0 0 (Slum) ALIOO'IEIA OEILVTnWIS 122 0.16 0.17 0.18 THEORETICAL VELOCITY (cm/s) 0.15 0.14 equation provides an excellent description of velocity for the reactive zone. The error is negligible (< 0.05%) for each velocity because the molecules begin in equilibrium proportions for each simulation. Therefore, each species contributes consistently to the zone velocity. I The simulated values of plate height with varying rate constant, mobility difference, and equilibrium constant are compared with the theoretical values calculated by using Equation 24 in FIGURE 3.58. As shown, the theoretical equation provides an excellent description of the plate height for the reactive zone. The error is within 10% for nearly all plate heights. However, it should be noted that two values of plate height Obtained from simulations have greater than 1 0% error. This is because the reactions associated with those data have not achieved steady state, which is a fundamental assumption Of the plate height model. Therefore, the accuracy of rate constants calculated via Equation 24 is limited to solute zones that are representative of reactions at steady state. 3.4.1 Propagation of Random Error In order to evaluate the reliability of calculated equilibrium constants and rate constants. it is important to understand the individual contributions to.error. In order to evaluate the error in equilibrium constant. Equation 13 must first be rearranged for K V _ K-A\V (30) V—VB From Equation 30, the propagation of random error in K is straight-fonivard 2 2 2 0% 1‘13) o2 +15) o2 4%) o?, (31a) OVA VA aVB V6 av 123 FIGURE 3.58: Comparison of theoretical plate height with simulation plate height for simulations with varying equilibrium constant (El). varying rate constants (I), and varying mobility difference (A) on the primary y-axis. The diagonal solid line represents a linear relationship between theoretical and Simulation plate heights. Percent error (X) is shown on the secondary y- axis. The horizontal solid line represents zero percent error. Simulation and theoretical conditions as described in Figures 3.4A — 3.4C. 124 FIGURE 3.58 0 "1' , 80883 .LNEIO81-‘Id O ‘- O O N 00 .50 EOE: meson 20:809.: ...0 ..0.0 50.0 ..000.0 . . .- 38d 1 1 ..00.0 x x x x < x ) Ix. x x x x xx vowx x - i . Ed I x m 5 (um) iHOlEIH ELLV‘ld OEILVTnWIS 125 2 2 02_(1)202+-(_VA_‘LI (.2. V‘VA -1 0%, (31b) v-vB V11 (v—v3)2 Vs (v-VB)2 v—VB The individual contributions to the error in K from v, v.., and V.; in Equation 31 are independent and additive when measured in separate experiments. It is apparent that the errors in K arising from VA and v3 are minimized when v >> v3 and that the error from V.; is further minimized when VA = v. It is also apparent that the error from v is minimized when v >> v3 and v... = v. FIGURE 3.6A shows the individual and cumulative contributions to percent error in K arising from VA, va. and v for Av = 0.025 cmls. When K < 0.1. the percent error from VA and v increases linearly as K decreases, whereas when K > 10, the percent error from v3 and v increases linearly with K. Moreover, when K < 0.1, 0.3 is large and dominated by the error from VA and v, whereas when K > 10, OK2 is large and dominated by the error from v3 and v. Only when 0.1 < K < 10 is 0.3 minimized With comparable contributions from each of the sources Of error. FIGURE 3.68 shows the effects of velocity difference (0.0025 5 AV 5 0-025 cmls) on the percent error in K for a fixed numerical average Of (V). + v3)l2 = 0.1625 cmls. It is apparent that the error in K decreases linearly as Av increases. This is a result of the ability to distinguish more accurately between the individual contributions to v from VA and V.; as Av increases. This correlates Well with the trends observed in FIGURE 3.3, where the linear regime between Velocity and K increases with Av. Therefore, increasing Av increases both the rahge and accuracy of discernable equilibrium constants. 126 FIGURE 3.6A: Individual and cumulative contributions to propagated percent error in equilibrium constant for 0.01 _<_ K 5 100: 0.12 (-—); (aK/av..)2 0.112 (— —); (ck/6113)2 o.,32 (m); (aK/av)2 0.2 (— .—). Calculation conditions: V.; = 0.2 cmls; E = 250 Vlcm; .111 = —1.00 x 10“ cm2Ns; .13 = -2.00 x 10“ cm2Ns; o... 6.3, and 0.. =1.00 %. 127 FIGURE 3.6A 00.. ._.Z<._1mZOO EDEmESOM 10.. 1 00.. 0000 v 80883 _LN3083d 128 FIGURE 3.68: Effects of mobility difference on propagated percent error in equilibrium constant for An = 1.00 x 10“ cm2Ns (—); Au = 5.00 x 10'5 cm2Ns (- —); Au = 2.00 x 10'5 cm2Ns (w); and Au = 1.00 x 10'5 cm2Ns (— - —). Calculation cOnditions: v0 = 0.2 cmls, E = 250 Vlcm, (.1.. + .13)/2 = —1.50 x 104 cm2Ns; o... Ova, and 0.. = 1.00 %. 129 FIGURE 3.68 00.. 0.. ...Z<..1.wZOO 23_mm_1._30m . .. ...0 n 5.0 10.. 1.00.. 1 0000.. ooooo. 80883 .LN3083d 130 In order to evaluate the propagated error in reaction lifetime, it is necessary to first rearrange Equation 20 for 1 (”‘QFIVWKIZ _ (111.2o...)(1.1<)2 32) 2KAv2 2KAv2 ( In From Equation 32, the propagation of random error in 1 is straight-fonrvard, albeit cumbersome 2 2 2 (33a) 61: 617 at 121-117,) “31115) 09111) “i 2 v(1+K) 02+ H(1+K) + v(1+K) I-I(1+K)2 a 2KAv2 H 2KAv2 2KAv2 2KAv2 HV ( I22 ( )( ) 2 ‘ 21+K 2Dm-Hv 1+K " m 1. * ma “’3“ ‘33”) + 2(Hv-ZDm)(1+K)_(Hv—2Dm)(1+K)2 202 2KAv2 2K2Av2 Most Of the contributions to the error in 1 are independent and additive. HOwever, the steady-state velocity and plate height are intrinsically linked and are measured in the same experiment, thus requiring consideration of their - C0Variance(o1..,). Covariance is negative when plate height is diffusion limited (H °‘ 11v) and positive when plate height is mass-transfer limited (H oc v). We have assumed positive covariance for these calculations. It is apparent that the error 131 in 1 increases with H, v, D..., and K and decreases with Av. It is also apparent that the error arising from K is minimized when K = 1. FIGURE 3.7A shows the individual and cumulative contributions to the percent error in 1 arising from H, v, D..., and Av for K = 1. The errors arising from v, D..., and Av, as well as the H-v covariance are inversely proportional to 1, whereas the error from H varies little with 1. Moreover, when 1 < 0.2 s, 0.2 is dominated by the contributions from Av and v, whereas when 1 > 2 s, 0.2 is dominated by the contribution from H. This is because of is limited by separation when 1 is small, but by reaction when 1 is large. FIGURE 3.78 shows the effects of equilibrium constant (0.1 s K s 10) on the percent error in 1. When K at 1. 0.2 varies little over the entire range of1, whereas when K = 1, 0.2 shows a strong dependence for values Of 1 < 2 s. The trends for 0.2 are mirrored about K = 1, such that errors associated with K < 1 are slightly larger than their K > 1 counterparts. These trends are a direct result of the contribution to 0.2 arising from K. When K = 1, the error arising from equilibrium constant is zero (Equation 33b). However when K at 1, the error arising from equilibrium constant is large and dominates the combined contributions from H and Av over nearly the entire range of 1. Therefore, the accuracy in 1 increases with proximity of K to unity. 3.5 CONCLUSIONS Generalized derivations are shown for velocity and plate height in reactive capillary electrophoresis using the non-equilibrium approach described] by 132 FIGURE 3.7A: Individual and cumulative contributions to propagated percent error in reaction lifetime for 0.05 5 1 5 500 s: 0.2 (—); (111/11H)? 01.2 (— —); (111/am)? 0.12 M: (at/av)2 of (— - -); (.11/60...)2 00.1.2 (— . - —); 2(a1laH)(61/av) 01.10., (——). Calculation conditions: 11.. = 0.2 cmls; E = 250 Vlcm; .11. = —1.00 x 10‘4 cm2Ns; .13 = -2.00 x 10‘ cm2Ns; K = 1; OH, 00..., o... and 0.... =1.00 %. 133 FIGURE 3.7A I , / / I r / i- < I l l l O O ‘- T“ V- \— C 80883 .LN3083d 134 100 10 0.1 0.01 REACTION LIFETIME (s) FIGURE 3.78: Effects of equilibrium constant on propagated. percent error in reaction lifetime for 0.05 5 1 5 500 s: K = 10 (- —); K = 2 ("-); K = 1 (—); K = 0.5 (_._); K = 0.1 (—--—). Calculation conditions: v1. = 0.2 cmls; E = 250 Vlcm; .1.. = —1.00 x 10" cm2Ns; .13 = —2.00 x 10‘4 cm2Ns; o... 09..., 0... and 11..v = 1.00 %. 135 FIGURE 3.78 00.. .o. 0.253.: 20:09.04. 2 F ..o ..0.0 ..0.0 1 F0 10.. 1 00.. . 82 80883 .LN3083d 136 Giddings. The resulting equations are consistent with their chromatographic analogs and are validated with an independent stochastic simulation. The equations for velocity and plate height are logical extensions of chromatographic theory and thereby illustrate fundamental similarities between chromatographic and electrophoretic separations. Moreover. these equations exemplify the shared thermodynamic (K) and kinetic (k) relationships in spite of» different separation mechanisms. Therefore, the equations for velocity and plate height I may be used to evaluate equilibrium constants and rate constants in liquid chromatography, capillary electrophoresis, and capillary electrochromatography. Velocity for reactive electrophoretic separations is dependent on equilibrium constant as well as on the mobilities of the relevant species. Velocity can be used to _detem'iine the equilibrium constants accurately when the individual mobilities are known. The range and accuracy of accessible equilibrium constants increases with mobility difference. On the other hand, plate height is dependent upon equilibrium constant. kinetic rate constant. and mobility difference. Plate height is shown to pass from . a diffusion-limited region to a mass-transfer-limited region with increasing electric field strength. It is shown that the range of accessible rate constants from mass- transfer-limited plate heights increases with electric field strength and mobility difference. , It is also shown that, at steady state, mass-transfer-limited plate heiQ hts are linearly related to kinetic rate constants. The propagated errors for K and 1 show that the range and accuracy of any Calculated equilibrium constants or rate constants will depend most on the 137 proximity of K to unity and on the magnitude of Av. For complexation reactions, this means that the most accurate rate constants will be achieved when the ligand concentration equals the inverse of K and when the overall size or Charge of the complex is substantially larger than that of the uncomplexed species. 138 10 11 12 3.6 REFERENCES Datta, R.; Kotamarthi, V. R., AIChE J. 1990, 36, 916 McEldoon, J. P.; Datta. R., Anal. Chem. 1992, 64, 227 Rathore, A. S.; Horvath, Cs., Electrophoresis 2002, 23. 1211 Knox. J. H., J. Chromatogr. A 1994, 680, 3 Giddings, J. 0., Dynamics of Chromatography, Marcel Dekker, Inc., New York, NY 1965 Newman, C. I. D.; McGuffin, V. L.. Electrophoresis 2005, 26, 537 McGuffin, V. L.; Krouskop, P. 8; Wu. P. R., J. Chromatogr. A 1998, 828, 37 - McGuffin, v. L.; Krouskop, P. 8; Hopkins, D. L., Unified Chromatography, ACS Symposium Series 748, Parcher, J. F ., Chester, T. L. Eds., American Chemical Society, Washington, DC, 2000 McGuffin, V. L.; Wu, P. R.; Hopkins, D. L., Proceedings of the lntemational Symposium on Chromatography, Hatano. H., Hanai, T., Eds., World Scientific Publishing: River Edge, NJ, 1995 Hopkins, D. L.; McGuffin. V. L., Anal. Chem. 1998, 70, 1066 Krouskop, P. 8, Ph.D. Dissertation 2002, Michigan State University Howerton. s. 8.; McGuffin, v. L., Anal. Chem. 2003, 75. 3539 ' 139 CHAPTER 4: EXPERIMENTAL METHODS 4.1 INTRODUCTION Two instrumental systems were utilized for the studies discussed in Chapters 5 and 6. In Chapter 5, a commercial capillary electrophoresis (CE) system was used for diode array absorbance detection at a single location along the capillary. In Chapter 6.6 novel CE system designed and constructed in- house was used for laser-induced fluorescence detection at multiple locations along the capillary. This chapter details the experimental systems and mathematical functions used to collect and analyze the data in those Chapters. 4.2 1 EXPERIMENTAL SYSTEMS 4.2.1 Capillary Preparation All separations were performed in W transparent, fused-silica capillaries , (Polymicro Technologies). Each capillary was cut to the appropriate length and detections windows were fabricated by removing the polyimide coating at specific locations. Following installation, the surface of each capillary was conditioned with 0.1 M NaOH and distilled water for 10 minutes each, prior to equilibration with the buffer at room temperature for 24 hours. The capillary surfaces were further equilibrated with the buffer at each therrnostatted temperature for 3 hours. 140 4.2.2 Capillary Electrophoresis Systems 4.2.2.1 CE System with Diode Array UV Absorbance Detection The separations discussed in Chapter 5 were achieved with an Agilent G1600A capillary electrophoresis system operating ChemStation software (Hewlett Packard). The capillary (id. = 75 um, o.d. = 360 um, L1... = 34 cm, L3... = 26 cm) was thermostatted (8.0 - 60.0 :I: 0.2 °C) with an internal Peltier device. Sample and buffer vials were thermostatted in situ with a NesLab RTE-10 (T hermo Electron Corporation) Circulating water bath (-25.00 — 150.00 :I: 0.01 °C). Samples were introduced to the column hydrodynamically (0.5 mbar x 30 s) and detected via UV absorbance at 210 nm. A schematic diagram of the diode-array CE system is shown in FIGURE 4.1A. 4.2.2.2 CE System with Laser-Induced Fluorescence Detection The separations discussed in Chapter 6 were achieved with a capillary electrophoresis system that was designed, constructed, and validated in-house. This system contains of a modified incubation oven (LabLine, Model 3500-DT) capable Of maintaining cOnstant temperature (5.0 - 50.0 :I: 0.2 °C). A plexiglass insert electrically insulates the oven interior in addition to physically supporting the individual components Of the separation system with a series of ventilated . shelves. The capillary (id. = 50 um, CO. = 360 um, L... = 60 cm, L3... = 22 cm and 38 cm) as well as sample and buffer vials are thermostatted within the oven. An auto-reversing high-voltage power supply (Bertan High Voltage, Model 2341-A) is used for injection and separation. The power supply can be used to apply a fixed voltage (0 - 30 W t 0.1%) and monitor the current or to apply a 141 FIGURE 4.1A: Schematic diagram of the Agilent G1600A capillary electrophoresis system. Separation system components: Pt electrodes (E), :l: 0 - 30 kV high voltage power supply (V), inlet vial (l), fused-silica capillary (0), outlet vial (O), and ground (G). Detection system components: deuterium light source (S), focusing lens (L), and UV-vis diode array detector (D). 142 FIGURE 4.1A 0-30 kV 143 fixed current (0 — 400 LIA :I: 0.1%) and monitOr the voltage. A magnetic switch (McMaster Carr, Model Schmersal BNS 33) positioned on the incubator door is a remote safety interlock for the high-voltage power supply and insures that the power supply is only enabled when the incubator door is closed. A 24 V power supply (Lambda Power, Model VS758-24) and a 5 V power supply (Lambda Power, Model VS150-5) are used to power and control the high-voltage power supply, respectively. These power supplies are housed within a) ventilated aluminum case which also provides connection between the high-voltage power supply and a computer via a 15—pin D-connector plate and cable (Black Box. Model EGM16E—0005-MF). A computer program written with commercially available software ~ (LabWew. Version 5.1, National Instruments) allows the user to simultaneously control and monitor the high-voltage power supply. .The program'enables the user to select the magnitude and duration of either the voltage or current applied from the power supply. The program is also used to monitor and store data from the power supply that represents either the current when voltage is controlled or the voltage when current is controlled. Samples are introduced to the column electrokinetically (1 W x 15 s) and detected spectroscopically via laser-induced fluorescence using a 325 nm HeCd laser (Melles Griot, Model 3074-20M). Solute molecules are excited at each detector location by transmission of laser light to the capillary via 100 um Optical fibers. Fluorescence is collected orthogonally to the incident light via 500 um optical fibers, transmitted through 400 nm long-pass optical filters to 144 monochromators (Instruments SA, Model'H-1061) and finally transmitted to photomultiplier tubes (Hamamatsu, Model R1463). The resulting photocurrent is amplified and converted to a voltage (Keithley, Model 480 picoammeter) that is recorded digitally with a data acquisition board (National Instruments, Model PCI- MlO-16XE-50) on the computer. Thus, the same LabVIew program is used to record data from the detectors as well as to cOntrol and monitor the high-voltage power supply. A schematic diagram of the laser-induced fluorescence CE system is shown in FIGURE 4.18. 4.3 DATA TREATMENT AND ANALYSIS After collection, zone profiles are extracted from electropherograms using previously established conditions for the minimum number of points and integration limits [1]. Each profile is fit by nonlinear regression with commercially available software (PeakFit, Version 3.18, SYSTAT Software). The fitted profiles are then used to calculate the statistical moments of each zone. 4.3.1 Mathematical Functions Numerous mathematical functions within PeakFit are capable of fitting the zOne prOfiles. However, the exponentially modified Gaussian (EMG) and asymmetric double-sigmoidal (ADS) functions almost exclusively provided the best fit of the experimental data. The EMG is the result of convoluting a Gaussian funCtion with an exponential decay [2-4]. This function generally provided the best fit for zones 145 FIGURE 4.18: Schematic diagram of the capillary electrophoresis system built in-house. Separation system components: Pt electrodes (E), :I: 0 - 30 kV high voltage power supply (V), inlet vial (I), fused-silica Capillary (0), outlet vial (O), and ground (0). Detection system components: HeCd laser (S), focusing lens (L), excitation optical fibers (EX), emission optical fibers (EM). optical filters (F), monochromators (M), and photomultiplier tubes (P). 146 . FIGURE 4.18 PMT P Mono M PMT P Mono EM M F<‘> EX EX 1 E G O >0. POGH , (n .J V 0-30kV 147 that exhibited tailing profiles. The mathematical form of the EMG is shown in Equation 1 A o2 te -t t-tG o ) C(t)=— exp + e - +1 (1) 2" [21:2 t «’20 «I21: where A is the area, Is is the mean time of the Gaussian component, or is the standard deviation of the Gaussian component. and 1 is the lifetime of the exponential component. The parameters of the EMG have previously been attributed physical significance for the evaluation of chromatographic zone profiles [511]. However, in this work significance is assigned to the moments of . the fitted profile rather than to the parameters used to obtain the fit. The ADS is an empirical function in which each side of the peak is represented by two distinct sigmoidal functions with different widths [12]. This function generally provided the best fit for zones that exhibited fronting profiles. The mathematical form Of the ADS is shown in Equation 2 I- h 1 (2) 1+ exp{_ t - ghosts} 1+ exp{_ t - 12:11. h C(t) = where h is the height of the zone, to is the center, t... is the separation of the two sigmoidals, w. is the width of the left sigmoidal, and wR the width of the right. (reverse) sigmoidal. I Zones that exhibited Gaussian profiles were generally fit equally well by the EMG and the ADS functions. 148 4.3.2 Statistical Moments Statistical moments for each zone were calculated at 0.01% of the maximum height. The first statistical moment (M1) describes the mean of the zone and the second statistical moment (M2) describes the variance of the zone. These values are related to the concentration of solute molecules as a function of time C(t) and may be calculated as 2tC(t)At ‘__2c(1)11 ‘3’ _ 2(1411.)2 on)... 2C(t)At (4) Mz 4.3.1 Figures of Merit The statistical moments for each zone were used to calculate the zone velocity (v) and plate height (H) Z -_ 5 v M () H-214; _ ' - (6) Mi where'z is the distance traveled. For data collected at a single detector location,. 2 is the distance from the inlet to the detection window and M1 and M2 are the moments Of the zone at the detection window. For data collected at multiple detector locations, 2 is the distance to either detection window or the distance between detection windows and M1 and M2 are the moments of the zone at or between detection windows, respectively. 149 Calculation of figures of merit by difference between multiple detectors is inherently more accurate than calculations using data from a single detector. Velocities calculated by difference are more accurate because the distance traveled between detectors is more accurately known than the distance traveled to any single detector. Plate heights calculated by-difference are more accurate because, in addition to the advantages mentioned for velocity, the broadening that occurs between detectors is strictly a result of separation processes of diffusion and mass transfer. This means that plate heights calculated from broadening between detectors selectively exclude contributions to zone variance from injection, which are independent and additive. 4.4 CONCLUSIONS The use of novel instrumentation and data treatment is necessary in order to reliably investigate reactive systems with capillary electrophoresis. This Chapter details the experimental systems and mathematical functions that were employed fOr data collection and analysis. 150 (”‘10) 10 11 12 4.5 REFERENCES Howerton. S. 8.; Lee, C.; McGuffin, V. L. Anal. Chim. Acta 2003, 478, 99 Grushka, E. Anal. Chem. 1972, 44, 1733 Grushka, E. J. Phys. Chem. 1972, 76, 2586 Yau, W. W. Anal. Chem. 1977, 49, 395 Foley, J. P.; Dorsey, J. G. J. Chromatogr. Sci. 1984, 22, 40 Jeansonne, M. S.; Foley, J. P. J. Chromatogr. Sci. 1991, 29, 258 Phillips, M. L.; White, R. L. J. Chromatogr. Sci. 1997. 35, 75 'Torres-Lapasio, J. R.; Baeza-Baeza, J. J.; Garcia-Alvarez-Coque. M. C. Anal. Chem. 1997. 69, 3822 Howerton. S. 8.; McGuffin, V. L. Anal. Chem. 2003, 75, 3539 Howerton. S. B; McGuffin,V. L. J. Chromatogr. A 2004, 1030.3 McGuffin, V. L.; Howerton. S. 8.; Li X. J. Chromatogr. A 2005, 1073, 63 Rundel, R. PeakFit Non-Linear Curve-Fitting Software: Technical Guide, Jandel Scientific, 1991 151 CHAPTER 5: THEORETICAL PLATE HEIGHT MODEL FOR THERMODYNAMIC AND KINETIC STUDIES OF REVERSIBLE, FIRST-ORDER REACTIONS 5.1 INTRODUCTION The work in this chapter utilizes the theoretical model developed in - Chapter 3 to calculate equilibrium constants and rate constants for the first-order isomerization of proline-containing dipeptides from the zone velocity and plate height, respectively. Whereas in chromatOgraphy, retention factor is proportional f to equilibrium constant, this relationship is not necessarily maintained in electrophoretic separations [1.2]. Application of the plate height model to capillary electrophoresis (CE) retains chromatographic formalism while relating velocity to equilibrium constant and plate height to rate constant [3,4]. This relationship is maintained because, under the auspice of‘ separations, zone velocity is dependent upon the individual velocities of the reactants and products and the fraction of time spent traveling at each velocity. However, zone broadening is dependent upon the individual velocities as well as the rates of reaction [5]. These relationships were validated with the independently developed stochastic (Monte Carlo) simulation in Chapter 2 [5.6-10]. Unlike other methods for evaluating rate constants in CE that are restricted to time regimes in which plateau profiles are Observed, the plate height model is restricted to time regimes in which a single zone is observed. This 152 means that the plate height model facilitates the evaluation of reactions at or near steady state, whereas other methods require that reactions be far from steady state. Moreover, only the first and second statistical moments (mean and variance) of the zone are required for calculating equilibrium constants and rate constants, respectively. Thus, no assumptions regarding the functional form of the zone profile are required. It is important to note here that the term steady state for a dynamic system (eg. separation) refers to the condition in which all molecules have reacted a sufficient number Of times such that each molecule has spent an equilibrium distribution Of time traveling as both reacting species. 5.2 THEORY A generalized development of the theoretical plate height model for reactive capillary electrophoreSiswas given in Chapter 3. Herein. that model is applied to the reversible, first-order isomerization of proline. dipeptides. The isomerization occurs via rotation about the peptide bond between the nitrogen atom of proline and the carbonyl Of the primary amino acid. This rotation supports the presence of both trans and cis isomers and can becharacterized by the equilibrium expression transJ—tcis (1) k. where k. is rate constant for the trans-cis rotation and k. is the rate constant for the cis-trans rotation. An illustration of this isomerization for Alanine-Proine and Phenylalanine-Praline is shown in FIGURE 1. 153 FIGURE 1: Illustration of the cis-trans isomerization for Alanine-Praline (top) and Phenylalanine-Proline (bottom). 154 FIGURE 1: 155 5.2.1 Velocity and Equilibrium Constant Capillary electrophoresis of these dipeptides results in a zone containing both isomers, with a velocity that is determined by the velocities of each of the individual isomers. The velocity (v) for the zone as well as that for the individual lsomers is calculated by (2) z Vi - —— M11 where z is the distance to the detector and M... is the first statistical moment or mean time to the detector for i = cis, trans, zone. In reactive CE. the velocity of the zone is equal to the weighted average Of the velocities Of the reacting species ' by [4,5] _vm, ch.s 3 V2” 1+K+1+K (-) where K = M kr is the equilibrium constant for the reaction in Equation 1. By measuring equilibrium constants at different temperatures, changes in the molar enthalpy (AH) and molar entropy (AS) are calculated via standard Gibbsian thermodynamic equations (Chapter 1, Equations 33 and 34). 5.2.2 Plate Height and Characteristic Reaction Lifetime Whereas velocity alone provides a reasonable means to estimate equilibrium constant. it is not sufficient for investigating the rates of reaction. Zone broadening, on the other hand. is directly related to resistance to mass transfer and is, therefore, a viable means for evaluating rate constants [4]. Plate height (H) is Calculated by 156 zllll2 -— 4 M? H where M2 is the second statistical moment _or variance. Application of the plate height model to peptide isomerization yields the expression 20 217K(vtrans - vols )2 H - + vzone (1+ K)2 V131... (5) where D is the average mobile-phase diffusion coefficient of the dipeptide. ~ The characteristic reaction lifetime (1 ) is given by (6) It is noteworthy that evaluation of kinetic rate constants by Equations 5 and 6 requires a priori knowledge of the equilibrium constant and the velocities of the individual isomers. By measuring rate constants at different temperatures, the activation energy (AE‘) is calculated via the Arrhenius equation (Chapter 1, Equation 35). 5.3 SIMULATION METHODS 5.3.2 Simulation Conditions A stochastic (Monte Carlo) Simulation was used to simulate capillary electrophoresis separations for the reversible, first-order A e 8 reaction with a fixed equilibrium constant (K = 1) over a range of rate constants (k1= k. = 0.001 to 10 s"). The input parameters for stochastic simulations are presented in Chapter 2. These simulation data were used as a known data set to compare 157 the efficacy of the plate height model and a Windows-based mass-balance simulation method for estimating rate constants from reactive CE zone profiles. 5.4 EXPERIMENTAL METHODS 5.4.1 Chemicals Alanine-proline (Ala-Pro), phenylalanine-proline (Phe—Pro), and mesityl oxide were purchased from Sigma. Sodium tetraborate decahydrate (Na28407) was purchased from MCB Manufacturing Chemists. The electrophoresis buffer for all separations was 1.0'x 10'2 M Na28407 (PH 9.3) prepared in distilled, deionized water. Analytical solutions of 1.0 x 10'3 M Ala-Pro and Phe—Pro were prepared in the buffer solution with 1.0 x 10'3 M mesityl oxide as the electroosmotic flow marker. 5.4.2 Experimental System Separations were achieved using the Agilent G1600A capillary electrophoresis system described in Chapter 4. 5.4.3 Data Analysis 5.4.3.1 Theoretical Plate Height Model Analysis of the experimental and Stochastic simulation data was performed .with commercially available software (PeakFit, Version 3.18, SYSTAT Software) for calculation of the first and second statistical moments (M1 and M2, respectively) from the fit parameters for the individual zones. These data were then used to calculate velocity from Equation 2 and plate height from Equation 4. The velocity data were used to calculate equilibrium constants (K) via Equation 3. 158 The plate height data were used to calculate characteristic reaction lifetimes (1) and rate constants (kg and k.) via Equations 5 and 6, respectively. 5.4.3.2 ChromWin As a measure Of comparison, thermodynamic and kinetic parameters for isomerization were also estimated With a Windows-based computer simulation (ChromWin 2004) developed by Trapp and Schurig [11-14]. The program utilizes experimental values such as retention time, zone height, plateau height. and zone width to calculate rate constants for reactive CE separations. Rate constants can be calculated by matching simulated electropherograms with experimental data as well as directly via an approximation function by inserting the datainto a spreadsheet. Prior to use for calculating rate constants from experimental data. the accuracy of ChromWIn was evaluated with the stochastic simulation data. 5.5 RESULTS AND DISCUSSION 5.5.1 Simulation Studies Stochastic simulations of reversible, first-order reactions illustrate the inflUence of rate constant and extent Of reaction on zone profiles. FIGURE 5.2A shows the zone profiles as a function 'Of rate constant and detector distance. For 0.1 2 k 2 0.033 s", two poorly resolved zones are observed at early detector distances and merge into a single, broad zone at later detectors. Whereas only a single zone is present at each detector for k 2 1 s“, two zones are present for k s 0.01 s". Retention time (from zone apexes). zone height, full width at-half 159 FIGURE 5.2A: Simulated CE profiles. Simulation conditions: N = 2000 molecules; t = 1 x 10'5 s; L1... = 100 cm; L3... = 2, 4, 6, 8, 10 cm. id. = 100 .1m; v = 25 kV; v... = 0.20 cmls; D = 1 x 10'5 cm2/s; .11 = —1.00 x 10“ cm2Ns; .13 = —2.00 x 10“ cm2Ns; K =1. 160 FIGURE 5.2A 161 maximum, and plateau height were measured at each detector after fast-Fourier transform (30%) of the zone profiles. These data were used as input parameters for calculation of rate constant via ChromVlfin. Zone velocity and plate height were calculated from the zone profiles at each detector. These data, in addition to the electrophoretic mobilities of species A and 8,.were used for calclation of rate constant via the plate height model. 5.5.1.1 Accuracy of Theoretical Plate Height Model and ChromWin FIGURE 5.28 shows the accuracy of rate constants calculated via ChromWin and Equation 5 for each zone profile. It is apparent that the accuracy of the calculated rate constants is dependent upon the extent of reaction, indicated by the Damkohler number (Da = M1/1) in TABLE 5.1. When Da < 0.2, ‘ relatively few molecules have reacted and individual zones are observed for species A and 8. In this time regime, both ChromWin and the plate height model fail to calculate rate constants accurately (> 30% error). When 0.2 < Da < 5, some molecules have spent time traveling as both species A and B and a - characteristic plateau is observed between the zones representative of those molecules traveling primarily as species A or 8. In this time regime, ChromVIfin is able to calculate rate constants quite well (2 — 25% error). whereas the plate height model is less accurate (2 25% error). When 5 < Da < 100, most molecules have spent an equilibrium proportion of time traveling as species A and B and a single zone is observed. In this time regime, ChromVlfin is incapable of calculating meaningful rate constants because only a single zone is present. However, the accuracy of the plate height model is actually greatest in this time 162 FIGURE 5.28: Effect of Damkohler number on the accuracy of rate constants I calculated by ChromWin (X) and by the plate height mod 6‘ (O) to the profiles in FIGURE 5.2A. 163 FIGURE 5.28 E: ”.9 5 O t 00 O )- OO - O O OO O 00 I 00 : (D X ' O< r O X X . O X X t 0 X X 0 X 0 X : .8 x X I O C O x _ x O X : O X Z O X ' O x - O O O O ‘- ‘- O O ‘— O v- " 80883 _I.N3083d 164 100 10 0.1 0.01 DAMKOHLER NUMBER TABLE 5.1 Damkohler Numbers for Simulation Profiles in FIGURE 5.2A. Damkohler number (Da) k (s"') 2 cm 4 cm 6 cm 8 cm 10 cm 10 2.46 x 102 4.92 x 102 7.38 x 152 9.85 x 102 1.23 x 103 1 2.46 x 101 4.92 x 101 7.38 x 101 9.85 x 10" 1.23 x 102 0.1 2.47 4.93 7.39 9.86 1.23 x 101 0.066 1.63 3.26 4.88 6.51 8.13 0.033 8.15 x 10'1 1.63 2.44 3.26 4.07 ' 0.01 2.47 x 10'1 4.94 x 10'1 7.41 x 10'1 9.88 x 10"l 1.24 0.001 2.47 x 10'? 4.94 x 10'2 7.41 x 10’2 9.88 x 10'2 1.24 x 10'T 165 regime (0.75 — 25% error). Therefore, the analysis of accurate kinetic information from zone profiles by ChromWin is limited to the time regime in which 0.2 < Da < 5 and, most importantly, is predicated upon the presence of plateau profiles. However, the analysis of accurate kinetic information from zone profiles by the plate height model is limited to the time regime in which 5 < Da < 100 and, is predicated upon the presence of a single zone. 5.5.2 Experimental Studies 5.5.2.1 Effects of Electric Field Strength Rotational isomerization about a peptide bond is a reversible. first-order reaction. Consequently, it is necessary to perform separations under two’time regimes for each temperature: one in which the individual isOmers are distinguishable (Da < 1), and one in which only a single zone containing both isomers is observed (Da > 5). Separations in the short-time regime (Da < 1) are ‘ necessary to determine the electrophoretic mobilities of the individual isomers as well as for kinetic analysis via ChromWin. These separations are also useful for evaluating equilibrium constants because (1c... 1133..., and .120... are all available from the same experiment. Electropherograms obtained in the long-time regime (Da .> 5) are representative of reactions near steady state and are used to calculate rate constants with the plate height model [4.5]. 5.5.2.1.1 Evolution of Reactive Zone Profiles I FIGURES 5.3A and 5.38 illustrate the effect of separation time on the zone profile at constant temperature (10 °C). The electropherogram in FIGURE 5.3A was achieved at 7500 V and Clearly shows both cis-Ala-Pro and 166 FIGURE 5.3: Effect of separation time on electropherograms of mesitY‘ oxide and Ala-Pro. Separation conditions: 1.0 x 10"2 M Na28407 buffer (pH 9.3), 10 °C, L... = 343111, L... = 25 cm. id. = 75 nm. (A) 7500 v, (8) 500 v. UV absorbance detected at 210 nm. 167 FIGURE 5.3A .55 ms: .4 L r... can? osxo Eco: 168 FIGURE 5.38 00F .EE. ms: 8. on O._n.1m_< case _aoos. 169 trans-Ala-‘Pro in the same zone, connected by a plateau. It is within this time regime that the electrophoretic mobilities of the individual isomers can be obtained. The electropherogram in FIGURE 5.38 was achieved at 500 V and clearly shows only a single Ala-Pro zone. It is within this time regime that the broadening of the Ala-Pro zone is proportional to the reaction lifetime. 5.5.2.1.2 Identification of lsomers For both Ala-Pro and Phe—Pro, the faster velocity is attributed to the trans isomer and the slower velocity to the cis isomer. This assignment presumes that the cis form exhibits a smaller hydrodynamic radius than the trans form, whereas 'theelectric charge of the isomers is necessarily the same. Therefore, the cis isomer has the larger charge-to-volume ratio and travels against the electroosmotic flow with a greater velocity than the trans isomer. This identification Of elution order is consistent with previous CE inVestigations Of peptidyl-proline isomerization [1 1.15.16]. 5.5.2.1.3 Comparison of Equilibrium Constants and Rate Constants « TABLE 5. 2 Is a comparison of equilibrium constants and rate constants for Ala-Pro and Phe-Pro at 10 °C from the plate height model, Chroman, and reported literature values [11,16]. Errors are reported as the standard deviation of replicate measurements. It is apparent that the equilibrium constants determined by these different methods are in good agreement and that the values calculated via the plate height model and ChromWin are statistically indistinguishable. The equilibrium constants calculated via 'zone heights for ChromWin are more precise than those calculated via the mobilities for the plate 170 Comparison of Equilibrium Constants and Rate Constants of Isomerization TABLE 5.2 Ala-Pro Phe-Pro Method 1.42 (s 0.27) 7.12 (1 2.50) Plate Height Model K 1.66 (1: 0.09) 6.00 (1: 0.08) ChromWin 1.20 5.66 [11] e 2.69 [16] 7 5.76 (1 0.30) x 10“ 1.06 (1 0.04) x 10‘3 Plate Height Model k (8..). 6.24 (11.19) x 10‘4 5.60 (:l: 4.06) x 10'5 Chromwln ' 6.9x 10*4 5.6x 10“ [11] ‘ 1.73x10“ [16] 3.91 (110.20) x 10" 1.43 (1 0.05) x 10“l Plate Height Model k (8-1) 3.76 (10.65) x 10*4 9.71 (:1 6.62) x 10’5 A ChromVVIn ' 7.4 x 10“ 9.65 x 10'5 [11] 6.42 x 10'5 [16] Forward and reverse rate constants for Ala-Pro and Phe-Pro isomerization at 10 °C calculated by plate height model in the long time regime (Da > 5), ChromWin in the short time regime (Da < 1), and reported values from [11] and [16]. Buffer conditions in this work: 1.0 x 10'2 M Na28407(pH19.3) Buffer conditions for [11]: 7.0 x 10'2 M Na28407 (pH 9.5) Buffer conditions for [16]: 1.0 x 10'1 M Na284Ov (pH 8.4). 171 height model. It is also apparent that there is excellent agreement among the rate constants for Ala-Pro, but less so_for Phe-Pro. In part, this is due to the smaller magnitude of the rate constants for Phe-Pro and the greater imprecision in the measurements. The rate constants calculated via the plate height model are more precise than those calculated via ChromWin under identical buffer and temperature conditions. This relatively poor precision for ChromVIfin may arise from uncertainty in determining plateau heights as result of the relatively large equilibrium constant for Phe-Pro. Slightly different buffer and pH conditions are used in references 11 and 16, which account for the small discrepancies observed with the literature values. It is important to recall that the plate height model was used to calculate rate constants from undistorted zones similar to those in FIGURE 5.38. Conversely, each of the other methods calculated rate constants from plateau profiles similar to those in FIGURE 5.3A. 5.5.2.2 Effects of Temperature In order to evaluate the effect of temperature on equilibrium constant, it is necessary to first evaluate the effect of temperature on electrophoretic mobility. FIGURE 5.4 shows electropherograms of Ala-Pro achieved at temperatures of 9 - 25 °C. At lower temperatures there are two zones for Ala-Pro, the resolution of which decreases with increasing temperature. ‘ This loss of resolution is attributable primarily to two sources: rate constants generally increase and equilibrium constants generally decrease with increasing temperature. 5.5.2.2.1 Electrophoretic Mobility V FIGURES 5.5A and 5.58 show the mean electrophoretic mobilities for 172 FIGURE 5.4: Effect of temperature on electropherograms of mesityl oxide and Ala-Pro. Separation conditions: 1.0 x 10’2 M Na28407 buffer (pH 9.3), L... = 34 cm, L3... = 26 cm, id. = 75 um, 7500 V. UV absorbance detected at 210 nm. 173 FIGURE 5.4 Mesityl oxid - Ala-Pro 9 °C , 10 °C w 15°C 20 °C 22 °C TIME (min) 174 FIGURE 5.5: Effect of temperature on (A) 11111111.-» (0). 113-$-11» (>< ). 3111A). and (B) Wrens-FF (0)1 Hols-FF (X)1 1le (A) separation conditions same as for FIGURE 5.4. Regression parameters (slope, intercept, R2): 113333.11: (-3.54 x 106. 8.82 x 10“, 0.9995); .1....» (-3.53 x 10", 6.63 x 10“, 0.9998); .1113 (-3.44 x 10", 8.45 x 10“, 0.9999); .1....13 (-3.20 x 10‘, 7.77 x 10*, 0.9998); .13.... (-307 x 103. 7.29 x 104. 0.9994); .133 (-3.03 x 106, 7.16 x 10“. 0.9996). 175 FIGURE 5.5A 4: N 0. 0°. ‘0. V. ‘3! ' O. (\I N \— \— ‘— r- \— (SNZLUG ..OLX) MITIBOW OIL38OHdO8iO313 176 290 295 300 TEMPERATURE (K) 285 280 FIGURE 5.58 CD or N . ( 1 I 1 l 0 °C) ‘9. ‘2 “I O. N ‘- 1- ‘- 1— ‘— SNzl-UO ..0 IX) AJJ'IIBOIN OLI.380Hd08.LO3'I3 177 ' 290 295 300 305 TEMPERATURE (K) 285 280 Ala-Pro and Phe—Pro species. respectively, as a function of temperature. Error bars are representative of the standard deviation of replicate measurements. In both figures, it is apparent that the mobilities of the individual cis and trans isomers (113.3311, (133.3311, (1.,-5-31:, and 113333.311) exhibit similar dependence on temperature. Moreover, the mobility Of the Ala-Pro zone (.1111) approaches a median value of (11.15.1111 + nmyz with increasing temperature, i.e. as K approaches unity. A similar trend is observed for the mobility Of the Phe-Pro zone (HFPI- The electrophoretic mobility has a complex temperature dependence arising from viscosity, dielectric constant, and ionic strength effects [17]. Yet the dependence of Ala-Pro and Phe-Pro mobilities on temperature within this range is well characterized by linear regression. Regression parameters are given in the figure legend, with high correlation coefficients (R2 > 0.999). Consequently, mobilities obtained at lower temperatures can be used to estimate electrophoretic mobilities at higher temperatures, where the individual isomers are indistinguishable. 5.5.2.2.2 Equilibrium Constant FIGURE 5.6A shows the effect of temperature on equilibrium constant for Ala-Pro and Phe—Pro isomerization. Error bars are representative of the propagated error, as described previously in Chapter 3. This propagated error in K is proportional to (.123... — .1....)2. Thus, the error in K is larger for Phe—Pro than for Ala-Pro because the difference between (11:11 and (13,-3-11: is smaller than the difference between 11.1..» and (13,-3.1.: owing to differences in their size. It is 178 FIGURE 5.6A: Effect of temperature on calculated equilibrium constants for Ala-Pro (A) and Phe-Pro (0). Separation conditions as described in text. 179 FIGURE 5.6A 7100 TIII I I I'll!!! T l O ‘- F iNVlSNOO Wfll88l'lan3 [IIIITI l 180 ‘_. O 3.35 3.40 3.45 3.50 3.55 1 I TEMPERATURE (x10'3 K) 3.30 3.25 apparent that the slopes are positive and that the equilibrium constant approaches unity for both Ala-Pro and Phe-Pro as temperature increases. It is also apparent that the slope for Ala-Pro is less than for Phe-Pro. The more positive slope for Phe-Pro indicates that the isomerization of Phe-Pro has a more negative Change in molar enthalpy than the isomerization of Ala-Pro. The more negative intercept for Phe—Pro indicates the isomerization of Phe-Pro has a more negative change in molar entropy than the isomerization of Ala-Pro. 5.5.2.2.2.1 Molar Enthalpy and Molar Entropy TABLE 5.3 summarizes the changes in molar enthalpy and molar entropy for Ala-Pro and Phe-Pro calculated via Equations 33 and 34 (Chapter 1). Errors represent the standard error from linear regression. It is apparent that trans-cis isomerization for both Ala-Pro and Phe-Pro is enthalpically favorable and has a strong entropically unfavorable contribution. This may be a result of a greater number of polar solvent molecules associated with the cis isomer because of decreased exposure of the non-polar side-chain at the a-carbon. It is also apparent that the magnitude of the enthalpic and entropic contributions for Ala- Pro are about one third of those for Phe—Pro. The entropic excess may be a result of the bulky aromatic ring of phenylalanine displacing more solvent molecules during the cis-trans rotation than the methyl group Of alanine. The enthalpic excess may have contributions from the breaking of hydrogen bonds as a result of displacing these solvent molecules during the rotation. 5.5.2.2.3 Rate Constants FIGURE 5.68 shows the effect of temperature on the forward rate 181 TABLE 5.3 Themodynamic and Kinetic Parameters of Isomerization Ala-Pro Phe—Pro AH (kJ mol") -14.12 (:I: 0.09) -36.34 (:1: 0.63) AS (J K" mol") -46.64 (:t 0.32) -111.62 (:t 0.21) AEg1 (kJ mol") 17.20 (:I: 1.84) -23.48 (:t 2.78) AESt (kJ mol") 31.27 (:I: 1.76) 12.86 (t 2.15) Thermodynamic changes in molarenthalpy (AH) and molar entropy (AS) as well ' as molar activation energy for the forward and reverse reactions (AE? and AE.*) for Ala-Pro and Phe-Pro. 182 ' FIGURE 5.68: Effect of temperature on calculated forward rate conStants for Ala-Pro (A) and Phe—Pro (0). Separation conditions as described in text. 183 )nshnt tions as FIGURE 5.68 (.-s) J.NV.LSNOO 31V8 184 A I |__{ 1 4 l I n 1— I V I m Tllll I l [TTIIII l l [IITIII l I V'- ‘- T- V- O O O O 0' 0 0 0 0' 0.: g 0' . O 3.35 3.40 3.45 3.50 3.55 1 ITEMPERATURE (X10‘3 K) 3.30 3.25 constants for Ala-Pro and Phe-Pro isomerization. Error bars are representative of the propagated error, as described previously in Chapter 3. This propagated error in kg has multiple contributions from (13,-s, 1133..., and K. It increases with (11m — 11...)", decreases as K approaches unity, and is proportional to the error in K. Consequently, the propagated error in kg for Ala-Pro is substantially less than that for Phe-Pro. It is also apparent that the slope for Ala-Pro is negative, whereas the slope for Phe-Pro appears positive. The more negative slope for Ala-Pro indicates that the trans-cis isomerization Of Ala-Pro has a more positive activation energy than that of Phe-Pro. The positive slope for Phe-Pro intimates a negative activation energy. suggesting a departure from Arrhenius behaviOr. FIGURE 5.60 shows the effect of temperature on the reverse rate constants for Ala-Pro and Phe—Pro isomerization. Error bars are representative of the propagated error, as described previously in Chapter 3. Once again, it is apparent that the error for Ala-Pro is less than that for Phe—Pro. This is for reasons similar to those discussed above for the propagated error in kg. It is also apparent that the slopes for both Ala-Pro and Phe-Pro are negative. The more negative slope for Ala-Pro indicates that the cis-trans isomerization has a more positive activation energy than that of Phe-Pro. 5.5.2.231 Molar Activation Energy TABLE 5.3 summarizes the activation energies for the fonlvard reaction (AE3) calculated via Equation 35 (Chapter 1). Errors represent the standard error from linearregresSion. It is shown that the activation energy for Ala-Pro is positive and that the molar activation enthalpy (AH = AE - RT) 185 FIGURE 5.6C: Effect of temperature on calculated reverse rate constants for Ala- Pro (A) and Phe-Pro (0). Separation conditions as deScribed in text. 186 FIGURE 5.6C goose 83.255202. men one mam 91m _ new one new f\ \l TIIII .50000 ..000.0 ..00.0 (.-S) .LNVLSNOO 31V8 187 contributes positively to the energy barrier for the trans-cis rotation. It is also verified. as suggested above, that the activation energy for Phe-Pro is negative. This type of departure from Arrhenius behavior has previously been attributed to a mechanistic change in the reaction pathway for protein-folding kinetics [18]. In the present case, it is unclear whether this trend is mechanistic or indicative Of a temperature dependence of the pre-exponential factor in Equation 6. Moreover, this activation energy includes zero at the 95% confidence level (AB = -23.48 :I: 35.48 kJ mol"). Thus, it would be unwise to attribute significant meaning to this value. The activation energies for the reverse reaction (AE3) calculated via Equation 9 are summarized in TABLE 5.3. Errors represent the standard error from linear regression. It is Shown that the activation energy for Ala-Pro is positive and that the mOlar activation enthalpy contributes positively to the energy barrier for the cis-trans rotation. It is also shown that the activation energy for Phe-Pro is positive, but includes zero at the 95% confidence level (AE.* = 12.86 :t 27.44 kJ mol"). Thus. little meaning should be attributed to this value. 5.6 CONCLUSIONS This is the first reported application and validation of the plate height model for thermodynamic and kinetic studies in capillary electrophoresis. This model has been used to calculate equilibrium constants and rate constants for isomerization of two proline dipeptides. Equilibrium constants are calculated directly from velocity, whereas rate constants are calculated directly from plate 188 height. The calculated values for the equilibrium constant and rate constants are in good agreement with those calculated by ChromWin as well as published literature values. It is shown that ChromWin and the plate height model are complimentary methods for extracting rate constants. ChromWin calculates rate constants more accurately for reactions that are far from steady state and exhibit plateau profiles. Alternatively, the plate height model calculates rate constants more accurately for reactions that are at or near steady state. 189 10 11 12 13 14 15 16 17 5.7 REFERENCES Rathore, A. S.; Horvath, Cs. Electrophoresis 2002, 23, 1211 'Knox. J. H.; J. Chromatogr. A 1994, 680, 3 Giddings, C. J. Dynamics of Chromatography, Marcel Dekker, New York, 1965 Newman, 0. l. D.; McGuffin, V. L. Electrophoresis 2005, 26, 4016 Newman, 0. I. D.; McGuffin, V. L. Electrophoresis 2005, 26, 537 McGuffin, V. L.; Krouskop, P. E.; Wu, P. J. Chromatogr. A 1998, 828, 37 McGuffin. V. L.; Krouskop, P. E.; Hopkins, D. L. In Unified Chromatography. ACS Symposium Series 748, Parcher, J. F., Chester, T. L. Eds, American Chemical Society, Washington, DC, 2000 McGuffin, V. L.; Wu. R; Hopkins, D. L. In Proceedings of the lntemational Symposium on Chromatography, Hatano. H., Hanai. T., Eds, World Scientific Publishing, River Edge, NJ, 1995 Hopkins,_D. L.; McGuffin, V. L. Anal. Chem. 1998, 70, 1066 Krouskop, P. E. Ph.D. Dissertation 2001, Michigan State University Schoetz. G.; Trapp, O.-; Schurig, V. Electrophoresis 2001, 22, 2409 Trapp. 0.; Trapp. G.; Schurig, V. Electrophoresis 2004, 25. 318 Trapp. 0.; Schurig, V. J. Chromatogr. A 2001, 911 , 167 Trapp. 0.; Schurig, V. Computers in Chemisry 2001, 25, 187 Ma, S.; Kalman, F.; Kalman, A.; Thunecke, F.; Horvath, Cs J. Chromatogr. A 1995, 716, 167 Thunecke, F.; Kalman, A.; Kalman, F.; Ma, S.; Rathore, A. S.; Horvath. Cs. J. Chromatogr. A 1996, 744, 259 Robinson, R. A.; Stokes, R. H. Electrolyte Solutions: The Measurement and Interpretation of Conductance, Chemical Potential and Diffusion in 190 18 Solutions of Simple Electrolytes, Buttenlvorths Scientific Publications, London,1995 Oliveberg. M.; Tan, Y. J.; Frescht. A. R. Proc. Natl. Acad. Sci. USA 1995, 92, 8926 191 CHAPTER 6: THEORETICAL PLATE HEIGHT MODEL FOR THERMODYNAMIC AND KINETIC STUDIES OF I REVERSIBLE, SECOND-ORDER REACTIONS 6.1 INTRODUCTION The theoretical plate height model was shown in Chapter 3 to accurately describe the contributions to zone broadening from diffusion and mass transfer processes in reactive capillary electrophoresis (CE). The theOreticaI equations developed in Chapter 3 illustrate that veloCity is independent of kinetic rate constant and is a direct measure of equilibrium constant. It is also illustrated in Chapter 3 that the mass-transfer contribution to plate height is inversely proportional to rate constant. These relationships were validated in Chapter 3 with stochastic (Monte Carlo) simulations as well as in Chapter 5 with well- characterized [isomerization of proline dipeptides. These successes notwithstanding, prior application of the plate height model in CE has been limited to reversible, first-order reactions. I This chapter describes an attempt to apply the plate height model to a reversible, second-order reaction. This class of reactions is ubiquitous in chemical and biochemical systems including DNA intercalation, drug-protein _ binding, cyclodextrin inclusion, and acid-base equilibria [1-9]. These reactions are characterized thermodynamically by binding (K3) Or dissociation (K3) constants and kinetically by association (kg) and dissociation (k3) rate constants. 192 Thermodynamic values are roUtinely calculated in CE from steady-state measurements of peak heights or electrophoretic mobility. Jia recently reviewed many of the methods used to perform these measurements, a number of which are discussed in more detail in Chapter 1 [10]. However, the calculatIOn of rate constants in CE has historically been limited to reactions that are far from steady state and exhibit plateau profiles. . While effective, these measurements inherently require that the plateau profiles be fit by one of a number of simulations that iteratively estimate kinetic rate constants. These methods are also described in more detail in Chapter 1. Each of these approaches inherently assumes that the simulation algorithms accurately describe the kinetic contributions to peak shape and also limits the range of calculable rate constants to less than 1 s". Application of the plate height model to reactive CE has the potential to overcome these limitations by allowing reactions to achieve steady state, thereby removing any reliance on simulation methods for calculating rate constants and simultaneously facilitating evaluations of more rapid reactions. Previous research in our laboratory (unpublished) investigated the selective Chromatographic retention Of several small drug and pesticide enantiomers by B- cyclodextrin. The dissociation rate constants for the retained solutes were found to range between 10 and 50 s"1 at room temperature. These rate constants are sufficiently rapid to preclude plateau profiles, but still expected to be within the range of measurable rate constants via the plate height model. Cyclodextrins are a popular class of compounds known to undergo selective complexation with a broad range of molecules. Cyclodextrins are 193 torroidally shaped polysaccharides Composed of D-glucose monomers connected at the 1 and 4 carbon atoms that exhibit a non-polar cavity and a polar exterior. The a, B. and y-cyclodextrins are composed of 6, 7, and 8 D-glucose units and are characterized by cavity diameters of 4.7, 6.3. and 8.3 x 10'10 m. respectively [11]. This combination Of physical and chemical properties enable cyclodextrins to selectively form inclusion complexes with a wide range of molecules. The ability to form inclusion complexes that are soluble in water positions cyclodextrins as ideal vessels to improve the solubility and bioavailability of pharmaceutical drugs [12]. The thermodynamics of cyclodextrin complexation have been studied extensively via numerous methods. In 1998 Rekharsky and lnoue published a lengthy compilation of cyclodextrin complexation thermodynamic data published or in press as of October 1997 [13]. Their results showed that structural and electronic attributes of the guest molecule such as hydrophobicity, shape, hydrogen bonding ability. aromatic substitution. and flexibility strongly affect the ' complexation thermodynamics. It was also concluded that only the portion of the guest molecule that “senses” a change in environment upon inclusion contributes appreciably to the overall thermodynamics of complexation. Although a small fraction of the reported data were collected by Chromatographic methods, and most were collected spectroscopically or by calorimetry. none were collected - electrophoretically. However, the use of CE to investigate thermodynamics of cyclodextrin complexation has been the topic of numerous research articles. Lemesle— 194 Lamache et al. used CE to investigate the complexation of native and derivatized B-cyclodextrin with a bronchodilation stimulant [14]. Penn et al. investigated the complexation of native and derivatized a— and B-cyclodextrins with a series of nitrophenolates [15]. Davis and co-workers investigated the complexation of native B-cyclodextrin with neutral and charged aliphatic dansyl amino acids, benzoate, and 4-nitrophenolate [9]. Additionally, applications of cyclodextrins for chiral CE separations have been discussed in several reviews [16-19]. This wealth of thermodynamic data notwithstanding, there is a noticeable lack of literature regarding the kinetics of cyclodextrin complexation. It is the intent of the work discussed in this chapter to employ CE to investigate therm‘odynamics and kinetics of cyclodextrin complexation via application of the theoretical plate height model. 6.2 THEORY Reversible, second-order complexation reactions can be expressed by the general chemical equation ka (1) where A and 8 are reactants and AB is the complex of A and 8, k. is the rate constant for association, and k3 is the rate constant for dissociatiOn. The reaction in Equation 1 may be characterized by the equilibrium expression k. [AB] “:Tnli ‘2’ 195 where K3 is the equilibrium constant for association. The approximation arises from the assumption that activity coefficients for A, B, and A8 are e 1. The rates for association (ra) and dissociation (rd) may be expressed as ra - ka [A] [6] (3a) rd .. k... [AB] (3b) Thus, the rate of association is second-order and is dependent upon [A] and [8], whereas the rate of dissociation is first-order and independent of [A] or [8]. Moreover, if [8] (or [A]) is sufficiently large or buffered such that it may be considered constant, then K3 and k. may be approximated as a pseudo-first- Order quantities K'b- K11 [B] (4) k'a-ka [B] (5) Therefore, controlling [8] facilitates simultaneous control of k’.. as well as the equilibrium proportions of A (GA) and A8 (011(3) GA - A] - 1 (6a) A + AB 1+Kb|BI a _ [AB] - Kb [B] (6b) AB |A|+|A8| 1+Kb|8l The studies inclusion of dansyl phenylalanine by B—cyclodextrin discussed in this chapter demonstrate reactions in which pseudo-first-order conditions are imposed by maintaining cyclodextrin concentrations in exceSs of dansyl phenylalanine. ' Investigations of the equilibrium constants and rate constants for 196 this system were performed as a detailed suite of experiments evaluating the effects of electric field strength, concentration, and temperature. 6.2.1 Velocity and Equilibrium Constant In capillary electrophoresis, the Observed velocity (V333) of the zone representative of species A interconverting between the free (A) and complexed states (AB) is equivalent to the weighed average of the individual velocities of A (VA) and AB (v.13) by Volos - 0LAVA + GABVAB (7) Substitution of Equations 6a and 6b into Equation 7 yields VA + VAB Kb [3] 1+Kb[8] 7 (8) Vobs " from which K3 may be calculated by evaluating v33. as a functiOn of [8], provided VA and v1.3 are known. 6.2.2 Plate Height and Rate Constant Whereas velocity alone provides a reasonable description of equilibrium constant, it is not sufficient for describing the rates of reaction. Zone broadening on the other hand is directly related to resistance to mass transfer and is therefore a viable means for evaluating rate constants [20.21]. Application of the plate height model deVeloped in Chapter 3 to the equilibria in Equation 1 yields 2 _ 2K3 [8](VA — 11.3)2 Vobs kd (1+ Kb [8])3 va8 DA , +DABKDIB: 1+K..[B] 1+K.,[B] H (9) 197 where, DA is the mobile-phase diffusion coefficient of A, D13 is the mobile-phase diffusion coefficient of A8, and the quantity in brackets represents the weighted average of the diffusional contributions to plate height. Thus, accurate evaluation of kinetic rate constants may be evaluated from the plate height. but require a priori knowledge of K3, DA, DAB, VA, and v.13. 6.3 EXPERIMENTAL METHODS 6.3.1 Chemicals Dansyl amide (DnsA), dansyl phenylalanine (DnsF) and B—cyclodextrin (80D) were purchased from Sigma. FIGURE 6.1 shows the molecular structures of DnsA, DnsF, and BCD. Sodium tetraborate decahydrate I (Na28407) was purchased from MCB Manufacturing Chemists. Electrophoresis buffer solutions containing 0 - 1.5 x 10'2 M 8CD were prepared with 1.0 x 10'2 M Na28407 (pH 9.3) in distilled, deionized water. Analytical solutions of 5.0 x 10‘5 M DnsF were prepared in each buffer solution with 4.0 x 10'5 M DnsA as the electroosmotic flow marker. These experiments were performed at 283, 285, 288, and 293 K. 6.3.2 Experimental System Separations were achieved using the capillary electrophoresis system with laser-induced fluorescence detection described in Chapter 4 and depicted in Figure 4.18. Fluorescence excitation was achieved at 1.3.31. = 325 nm and emission was monitored at A33... = 570 nm. 198 FIGURE 6.1: Molecular structures of dansyl amide. dansyl phenylalanine. and B-cyclodextrin. 199 FIGURE 6.1 \N/ \N/ NH2 . HN /C Dansyl amide ' \TH \OH CH; Dansyl phenylalanine o ’16 19 9° 0 go 0 0 o 0 B—cyclodextrin 200 6.3.3 Data Analysis Zone profiles were extracted from electropherograms with commercially available software (Origin, version 7.5, OriginLab) and iteratively fit for calculation of statistical moments using the procedures described in the Data Treatment and Analysis in Chapter 4 (Section 3). The presence of multiple detector locations for the laser-induced fluorescence system enables more accurate calculation of velocity and plate height by difference between detectors H_(Zz'z1)[MzI2"Mz)1] . (11) [~01 -M.).i where 21 and 22 are the distances todetector locations 1 (22.73 cm) and 2 (38.05 cm), M1). and M1)2 are the zone means at z. and 22. respectively, and M2). and M2)2 are the zone variances at 21 and z2, respectively.) By calculating velocity and plate height between detectors, extra-column contributions (Chapter 1. Section 1.2.4) to mean and variance are isolated and do not contribute to the calculated equilibrium constants or rate constants. 6.4 RESULTS AND DISCUSSION 6.4.1 Velocity and Equilibrium Constant 6.4.1.1 Effects of Electric Field Strength and Concentration on Velocity In order to accurately determine solute electrophoretic mobility and equilibrium constants for [30D inclusion, separations were performed at electric field strengths (E) Of 418, 368, and 334 Vlcm (25, 22, and 20 kV, respectively). 201 FIGURE 6.2 shows representative electropherograms of DnsA and DnsF with 0 - 1.0 x 10'2 M 6CD obtained at detector 2 with 416 Vlcm at 265 K. For each separation samples are introduced at the anode, electroosmotic migration is towards the cathode, and the electrophoretic migration of DnsF is towards the anode. As expected, the velocity of DnsF increases with 8CD concentration whereas the velocity of DnsA varies little with 6CD concentration. Addition of 80D to the electrophoresis buffer is known to effect buffer viscosity directly as well as indirectly (via buffer conductivity), thereby resulting in uncontrollable influences on both electroosmotic and electrophoretic velocities [10.14.15]. When unaccounted for, these additional contributions to velocity have been shown ‘to result in underestimations of K3 [15]. It has also been shown that the ratio of buffer viscosities with and without BCD (n and 110, respectively) is linear with the ratio of separation current with and without BCD (i [and in, respectively) [14]. Thus, the complicated effects of BCD on buffer viscosity and, hence, electrophoretic velocities (v... = v333-v33) may be corrected via [14.15] V' 311' vep(-,i—) (12) I0 where v33 is the electroosmoticvelocity and v’ep is the corrected electrophoretic velocity. FIGURE 6.3 shows the effect of electric field strength and BCD concentration on v”... at 285 K. It is apparent that v’3p is approaches a minimum when [80D] < 1 mM, experiences the greatest increase between 1 < [8CD] < 4 202 FIGURE 6.2: Effect of [ICU concentration on electropherograms of DnsA and DnsF. Separation conditions: L... = 59.85 cm, L3... = 36.05 cm. 265 K10 x 10'3 M Na28407 buffer (pH 9.3), v = 25 W, [660] = o - 1.0 x 10'2 M. Fluorescence detection conditiOns: 1.3.3.. = 325 nm, A3"... = 570 nm. 203 ns of W Ii 9.3))“ e detect?“ FIGURE 6.2 _<_—g _. J ' 2' E '42 O 1mM 2mM 4mM 10mM 204 TIME (min) FIGURE 6.3: Effect of 8CD concentration on electrophoretic velocity at 334 (O), 368 (A), and 416 (El) Vlcm. Separation conditions: L... = 59.65 cm. L... = 38.05 cm. 285 K, 1.0 x 10'3 M Na28407 buffer (pH 9.3), [BCD] = o - 1.0 x 10'2 M. Fluorescence detection conditions: 1.3.31. = 325 nm. 1.3.31. = 570 nm. 205 FIGURE 6.3 5.0 3.95 .00... See _._..P. 5000 5000.0 ...0- 1 00.01 1 00.01 1 N001 1 00.01 1 00.01 1 .4001 1 00.01 1 No.01 ..0.01 (S/UJO) A.L|3013/\ OIJ.38OHdO8.LO313 206 mM, and increases less when [BCD] > 4 mM. This sigmoidal dependence of v’33 on 8CD concentration can be generally described by three regions. The region in which v’.... is a minimum ([800] < 1 mM) corresponds to DnsF traveling predominantly as free DnsF. The region in which v’.p increases most with BCD (1 < [8CD] < 4 mM) corresponds to increasing proportions of DnsF traveling as ~ both the free and BCD inclusion complex. The region in which v’33 begins to approach a maximum ([BCD] > 4 mM) corresponds to DnsF traveling predominantly as the BCD inclusion complex. The inflection of the curve corresponds to DnsF traveling equal prOportion's of time as both free DnsF and the BCD inclusion complex. Therefore, the concentration of [ICU at the inflection is 1/K3, such that the pseudo-first-order equilibriuchnstant is unity (Equation 4). 6.4.1.2 Effects of Temperature and Electric Field Strength on Equilibrium Constant Corrected electrophoretic velocities were used to calculate equilibrium constants (K3) fOr inclusion "via non-linear regression to Equation 8. TABLE 6.1 summarizes the equilibrium constants Obtained at 283, 285, 288, and 293 K for each applied electric field strength as well as the average equilibrium Constant (ng) and standard deviation for each temperature. These values are in good agreement with previously reported binding constants for BCD with dansyl amide and dansyl amino acids (1 — 4 x 102 M") [9.11.13]. It is apparent that the data show excellent agreement between equilibrium constants calculated at the same temperature for each electric field Strength. Although the data shows minor increases in equilibrium constant with electric field strength, evaluation of the F 207 statistic for each temperature indicates that this trend is not statistically significant. It is also apparent that the data also Show a general trend of decreasing equilibrium constant with increasing temperature. 6.4.1.2.1 Molar Enthalpy and Molar Entropy The effects of temperature on K3... were used to calculate the molar enthalpy (AH) and molar entropy (AS) for inclusion via the standard Gibbsian thermodynamic relationships described in Chapter 1 (Section 2.1). FIGURE 6.4 is a plot of In K3... vs 1fl'. It is apparent that the plot is linear with a positive sIOpe and a positive intercept. The molar enthalpy for inclusion is calCulated from the slope as -12.66 kJ/mol (:I: 0.87 lemol); the molar entropy for inclusion is calculated from the intercept as -2.19 J/K-mol (i 3.06 JlK-mol). These data show that BCD inclusion Of DnsF is enthalpically driven and has a competing entropic contribution, in good agreement with previously reported trends [11 .13]. 6.4.2 Plate Height and Rate Constants I In order to differentiate the contributions of diffusion and mass transfer to plate height it is necessary to estimate the diffusion coefficients DA and DAB in Equation 9. Electrophoretic mobility can be used to estimate diffusion coefficients via the Einstein relation (D°. = (1°.RTIzF). However, this equation is only accurate at infinite dilution and zero ionic strength and, thus, does not fully account for the dependence of diffusion coefficient and electrophoretic mobility on temperature and ionic strength. As discussed in Chapter 1, Li et al. have adapted the detailed conductivity theory of Pitts to more accurately describe the dependence of electrophoretic mobility on temperature and solution electrical 208 TABLE 6.1 Calculated Equilibrium Constants for BCD Inclusion of DnsF. 334 Vlcm 358 Vlcm 4.18 Vlcm 283 K 278 277 280 278 (d: 1) 285 K 268 273 277 ‘ 272 (:t 4) 288 K 259 261 263 261 (:I: 2) 293 K 228 232 237 233 (:l: 4) Equilibrium constants for BCD inclusion of DnsF calculated from separations as a function of electric field strength and temperature. 209 FIGURE 6.4: Effect of temperature on calculated equilibrium constants for 8CD inclusion of DnsF. Experimental conditions same as given for FIGURE 6.3. 210 FIGURE 6.4 V0.0 googmmazmmosmt. _ em mew .. can 33 h n — n . Ned v.0 II< 000 F 00F . .LNVLSNOO INI'II88I'IIIIO3 211 properties [9,22-24]. From their description of electrophoretic mobility, 11°. can be calculated iteratively via the self-consistent algorithm developed by Survey et al. [9.25], which may then be used to calculate D01. Diffusion coefficients may then be corrected for temperature and ionic strength via Onsager’s treatment for diffusion of ions at near-infinite dilution discussed in Chapter 1 (Equations 7, 8, 12, and 13) [9.26]. TABLE 6.2 lists the diffusion coefficients of free and bound DnsF with (Do. and D13) and without (DA and D’AB) Proper corrections for temperature and ionic strength. It is apparent that direct application of the Einstein relation to electrophoretic mobility tends to overestimate DA and D113 by about 1%. However, this relatively small discrepancy translates into > 5% error in calculated , rate constants when the mass transfer contribution to plate height is < 20 times the contribution from diffusion. 6.4.2.1 Effects of Electric Field Strength and Concentration on Plate Height FIGURE 6.5A shows the effects of electric field strength and [30D concentration on H at 285 K. It is apparent that plate height changes little when [BCD] < 1 mM, is generally slightly greater for intermediate values of [8CD] (2 - 5 mM), and decreases with [3CD] > 5 mM. However, the error associated with these values is quite large. obscuring any Significance of these trends. It is also apparent that plate heights obtained at 334 Vlcm are generally greatest, whereas those obtained at 418 Vlcm are generally least. This means that plate height is larger at slower velocities than at faster velocities, intimating that diffusion is a 212 TABLE 6.2 Calculated Diffusion Coefficients of Free and Bound DnsF. 283 K 285 K 288 K 293 K *—=——-F 10 04cm Is) 2.35 (+ 0.01) 2.43 (+ 0.01) 2.69 (+ 0.01) 3.15 (+ 0.02) 1 OWD’A (cm2/s) 2.38 (+ 0.01) 2.46 (+ 0.01) 2.72 (+ 0.01) 3.19 (+ 0.02) %Error in D’). 1.06 1.11 1.19 1.37 10'7 013(cm’rs) 7.20 (+ 0.26) 7.84 (+ 0.20) 8.45 (+ 0.21) 10.1 (+ 0.33) 10'7 0’13 (cm2/s) 7264+ 0.26) 7.93 (+ 0.20) 8.55 (+ 0.22) 10.2 (+ 0.34) %Error in D’AB 1.06 1.11 1.19 1.37 Diffusion coefficients of free and bound DnsF with (D1. and DA...) and without (DA and D’AB) corrections for temperature and ionic strength at 283, 285, 288, and 293 K. 213 FIGURE 6.5A: Effect of BOD concentration on plate height at electric field strengths Of 334 (O), 368 (A), and 416 (El) Vlcm. Separation conditions: L... = 59.85 cm, L... = 38.05 cm, 285 K, 1.0 x 10'3 M Na28407 buffer (pH 9.3), [BCD] = 0 - 1.0 x 10'2 M. Fluorescence detection conditions: A3,... = 325 nm, Aemit = 570 nm. 214 FIGURE 6.5A 3.68 .00... ..o 8.0 .86 88.0 88.0 .F.p. EF-._. 3 H ...br. . . L .....hb _ r 00000.0 - «888 d .. 880.0 m ....— . mm - 088 o m .. H .. .. In... L 14.31.. .. .Ii -“.. . - 88cc ) ffieii ._ ... . . 289° N..000.0 215 . more dominant contribution to broadening than mass transfer from reaction. FIGURE 6.53 shows the effects of electric field strength and pCD concentration on H corrected for diffusion at 285 K. It is shown in Equation 4 that [300 concentration is proportional to K’b, thus FIGURE 6.58 also shows the effects of K’b on H corrected for diffusion. It is apparent that plate height is least when [BCD] < 1 mM (K’b < 0.27) and greatest when [BCD] > 1 mM (K’b > 0.27). It is also apparent that plate height changes little within these regimes. This is in contrast to the trends shown in FIGURE 3.4C (Chapter 3) where the mass transfer contributions to plate height are greatest at K’b = 1 (here, [fiCD] = 3.7 mM) and decreases monotonically as K deviates from unity. Thus intimating that the dominant contribution to broadening is from neither diffusion nor mass transfer. 6.4.2.2 Effects of Temperature and Electric Field Strength on Rate Constants Ignoring the obvious lack of statistical significance or dependence of plate height on [500] for these data, one can attempt to use plate heights corrected for diffusion to calculate rate constants for dissociation (kd) via Equation 9 and association (kg) via Equation 2. TABLE 6.3 summarizes the average rate constants (km,g and Ram) and standard deviations obtained at 283, 285, and 288 K. It is apparent from these data that kd are all roughly 20 -~ 30 s", increase marginally with E, and remain essentially constant with temperature. This trend is in contrast to the “rule-of-thumb” expectation that rate constants typically double for every increase of 10 K. Moreover, it was discovered sometime after 216 FIGURE 6.58: Effect of 500 concentration on plate height corrected for diffusion at electric field strengths of 334 (O), 368 (A), and 416 (El) Vlcm. Separation conditions: Ltd a 59.85 cm, Ld.‘ = 38.05 cm, 285 K, 1.0 x 10'3 M NazB407 buffer (pH 9.3), [BCD] = O - 1.0 x 10'2 M. Fluorescence detection conditions: km“ = 325 nm, Item“ = 570 nm. 217 FIGURE 6.58 _..o l-P-lb - p 3:25 303 :3 god 58o ...-L... - in.-.. . —.b_.-.. . 5000.0 ..u\; .. .1 .... \t .W...‘.: 1:3 l \ 51‘ . “V A n. “my: nn’... , H! II I. aux as o 1 525.0 - Nooood r mooood .. Vooood - mooood I 000006 - hooood mooood (we) .LHSIEIH 5mm! 218 TABLE 6.3 Calculated Rate Constants for BCD Inclusion of DnsF. kwg (M'iI S") kd.a_vg (8'1) 283 K 6.42 (+ 0.44) x 103 2.31 (4- 0.16) x 101 285 K 6.41 (+ 0.59) x 103 2.36 (+ 0.26) x 101 288 K 5.77 (+ 0.49) x 103 2.21 (+ 0.20) x 10'r Average rate constants for association (kam) and dissociation (Ram) of fiCD inclusion of DnsF calculated from separations at 334, 368, and 418 Vlcm as a function of temperature. 219 completion of these studies that theserate constants are ."103 times slower than those reported for the dissociation of [3CD with cobaltocene and ferrocene [27- 29]. It is also apparent that all k, are > 5.0 x 103 M'1 s“, increase marginally with E, and decrease as temperature increases. This trend arises in the data strictly as a consequence of calculating ka via Equation 2. Moreover, these rate constants are 104 times slower than those reported for the dissociation of BCD with cobaltocene and ferrocene and 103 times slower than those expected assuming kd ~ 10‘ 5". Thus, although the calculated equilibrium constants compare reasonably with literature values, the calculated rate constants do not, because mass transfer is not a significant contribution to broadening. Using the calculated values of k, and kd, the characteristic reaction lifetimes near K’b~ 1 are on the order of 10 — 20 ms, resulting in Damkohler numbers > 1000. As shown in Chapter 5 (FIGURE 5.13) the plate height model cannot be used to accurately determine rate constants from Damkohler numbers > 100 without correcting for diffusion. _However, correcting for diffusion only extends the viability of the plate height model by an order of magnitude to Damkohler numbers 5 1000. Therefore, the calculated values for ka and kd are not only unreasonable in contrast to published values for. BCD inclusion but also unreliable with respect to the experimental limitations of the plate height model. 6.5, CONCLUSIONS This chapter details the first attempted application of the plate height model to the reversible, second-order inclusion of dansyl phenylalanine by B- 220 cyclodextrin. Separations were performed with a dual, on-column laser- fluorescence detection CE system with varying concentrations of BCD over a range of temperatures and electric field strengths. The plate height model was used to calculate thermodynamic equilibrium constants and kinetic rate constants for inclusion from 283 - 293 K. Equilibrium constants are calculated by evaluating the effects of [SOD concentration on velocity. It is shown that the electrophoretic velocity of DnsF increases sigmoidally with BCD concentration _in agreement with the observed dependence of velocity on equilibrium constant shown in Chapter 3 (FIGURE 3.3). This trend was used to elucidate equilibrium constants via non-linear regression to Equation 8. The calculated equilibrium constants, molar activation enthalpy, and molar activation entropy {are shown to be in agreement with previously reported values for BCD inclusion of dansyl amino acids. Rate constants are calculated by evaluating the effects of BCD concentration on plate height. It is shown that plate heights are largely unaffected by BCD concentration. It is also shown that plate heights generally increase as electric field strength decreases, intimating that diffusion is a more dominant source of broadening than mass transfer from reaction. Plate heights were corrected for diffusion to increase the accuracy of calculated rate constants; however, the rate constants determined via the PHM were ~103 times slower than the only other reported rate constants for BCD inclusion in aqueous buffers. It is presumed that the calculated rate constants for BCD inclusion of DnsF are grossly underestimated by the plate height model because of both inherent 221 physical limitations of the experimental system as well as inherent physical . limitations in the mathematical model. It was shown in Chapter 3 that the range of calculable rate constants by the plate height model under ideal chemical conditions (K’b = 1) is limited by: 1) the amount of time molecules have had to react prior to detection; and 2) the difference between the velocities of products and reactants. For the 3CD inclusion of DnsF, velocity increased with Kb’; however, the differences between free and bound DnsF are quite small (< 0.05 cmls). This means that even using the maximum achievable electric field strength (501 Vlcm), the fastest rate constants calculable for the DnsF-[3CD inclusion system are ~20 — 30 s". Thus, calculation of rate constants in excess of this limit requires faster separations (smaller Damkohler numbers) via shorter detection distances, larger electric field strengths, or both. The most likely method to meet these requirements is to perform the separations on chip. The primary advantages of chip CE over conventional CE are shorter separation channels (3 - 25 cm) and a larger range of electric field strengths (~1 kV/cm). Chip CE also facilitates detection at shorter detection distances than conventional CE systems, thereby decreasing diffusional contributions to broadening. However, unreasonably large electric field strengths (> 5 kV/cm) are necessary in order to access rate constants near 104 s”. Thus, even given optimal instrumentation, this reaction is likely still too fast to calculate rate constants from the plate height model with any degree of reliability. Moreover, it is difficult to predict successful application of the plate height model for a particular reaction without some a priori knowledge of electrophoretic 222 mobilities and diffusion coefficients in addition to a reasonable estimation of at least one of the rate constants. The effects of these limiting factors on the maximum calculable rate constants are evaluated in detail in Chapter 7. 223 10 11 12 13 14 15 16 17 6.7 REFERENCES Maiti, S.; Chaudhury, N. K.; Chowdhury, S. Biochem. Biophys. Res. Comm. 2003, 310, 505 Jia, 2.; Ramstad, T.; Zhong, M. Electrophoresis 2001, 22, 1112 Jia, Z.; Ramstad, T.; Zhong, M. J. Pharm. Biomed. Anal. 2002, 30, 405 Ostergarard, J.; Schou, G; Larsen, C.; Heegaard, N. H. H. Electrophoresis 2002, 23, 2842 Ohnishi, T.; Mohamed, N. A. L.; Shibukawa, A.; Kuroda, Y.; Nakagawa, T.; Gizawy, S. E.; Askal, H. F.; Kommos, M. E. E. J. Phann. Biomed. Anal. 2002, 27, 607 Baumy, P.; Morin, P.; Dreux, M.; Viaud, M. C.; Boye, S.; Guillaumet, G. J. Chromatogr. A. 1995, 707, 311 Larsen, K. L.;Zimmem1ann, w. J. Chromatogr. A. 1999, 836, 3 Guillaume, Y. C.; Peyrin, E. Anal. Chem. 1999, 71, 2046' Seals, T. H.; Sheng, C.; Davis, J. M. Electrophoresis 2001, 22, 1957 Jia, 2. Curr. Phann. Anal. 2005, 1, 41 Catena, G. C.; Bright, F. V. Anal. Chem. 1989, 61 , 905 Stella, V. J.; Rao, V. M.; Zannou, E. A.; Zia, V. Adv. Drug Deliv. Rev. 1999, 36, 3 Rekharsky, M. V.; lnoue, Y. Chem. Rev. 1998, 98, 1875 : Lemesle-Lamache, V.; Taverna, M.; Wouessidjewe, D.; Duchene, D.; Ferrier, D. J. Chromatogr. A 1996, 735, 321 Penn, S. G.; Bergstrom, E. T.; Knights, l.; Liu, G.; Ruddick, A.; Goodall, D. M. J. Phys. Chem. 1995, 99, 3875 Chu, Y.-H.; Cheng, C. C. Cell. Mol. Life Sci. 1998, 54, 663 Britz-McKibbin, P.; Chen, D. D. Y. Electrophoresis 2002, 23, 880 224 18 19 20 21 22 23 24 25 26 .27. 28 29 Kapnissi-Christodoulou, C. P.; Zhu, X.; Warner, I. M. Electrophoresis 2003, 24, 3917 Gubitz, G.; Schmid, M. G. Electrophoresis 2004, 23, 3981 Newman, C. I. D., McGuflin, V. L. Electrophoresis 2005, 26, 573 Newman, C. l. D.‘, McGuffin, V. L. Electrophoresis 2005, 26,4016 Li, D.; Fu, 8.; Lucy, C. A. Anal. Chem. 1999, 71,687 Pitts, E. Proc. R. Soc. A 1953, 217,43 Pitts, E.; Tabor, B. E.; Daly, J. Trans. Faraday Soc. 1970, 66, 693 Survay, M. A.; Goodall, D. M.; Wren, S. A. C.; Rowe, R. C. J. Chromatogr. A 1996, 741, 99 Robinson, R. A.; Stokes, R. H. Electrolyte Solutions, The Measurement and Interpretation of Conductance, Chemical Potential and Diffusion in Solutions of Simple Electrolytes, Butterworths Scientific Publications: London,1955 lsnin, R.; Salam, C.; Kaifer, A. E. J. Org. Chem. 1991, 56, 35 Wang, Y.; Mendoza, S.; Kaifer, A. E. Inorg. Chem. 1998, 37, 317 Guyard, L.; Hapiot, P.; Jouini, M.; Lacroix, J.-C.; Lagrost, C.; Neta, P. J. Phys. Chem. 1999, 103, 4009 225 CHAPTER 7: THEORETICAL LIMITATIONS OF THE PLATE HEIGHT MODEL 7.1 INTRODUCTION The theoretical development in Chapter 3 showed that mass-transfer limited plate heights are inversely proportional to rate constant. It was also shown in Chapter 3 that plate height is mass-transfer limited when the contributions from diffusion are 5 20% the total plate height. As demonstrated previously, the diffusional contribution to plate height is dependent upon physical parameters of the solutes such as diffusion coefficient (Chapter 3, Equation 2) and electrophoretic mobility (Chapter 1, Equation 7 and Chapter 3, Equations 1 and 22) in addition to physical parameters of the experimental system such as temperature (Chapter 1, Equation 7) and ionic strength (Chapter 1, Equation 12) as well as electroosmotic mobility and electric field strength (Chapter3, Equations 1 and 22). It was also shown in Chapter 3 (Equations 20 and 22) that the mass— transfer contribution to plate height is dependent upon physical parameters such as electric field strength; electroosmotic mobility and electrophoretic mobility as well as chemical parameters such as equilibrium constant and rate constant. Therefore, the maximUm rate constants calculable via the plate height model will vary with both the physical and chemical attributes of the individual system investigated. 226 7.2 THEORY In order to evaluate the individual physical and chemical contributions to the maximum calculable rate constant, it is helpful to consider their effects on a normalized plate height (h) hJ. (1) Hb where H is the total plate height from both longitudinal diffusion (H.,) and mass- transfer (H.,). For the reversible, first-order reaction e A B (2) kAB and kBA are the rate constants for transfer from species A to B and from species B to A, respectively. From Chapter 3, the contribution to plate height from longitudinal diffusion is described by (3) H -—--— b .v v1+K+1+K 2D 2( D A K03) where v is the observed velocity of the reactive zone, D is the observed diffusion coefficient of the reactive zone, DA is the diffusion coefficient of species A, D3 is the diffusion coefficient of species B, and K is the equilibrium constant for the reaction in Equation 2. 227 The contribution to plate height from mass transfer is described in Chapter 3 by 21(sz H a c kBAv(1+K)3 (4) where Av is the velocity difference between species A and B (VA and v3, respectively). Therefore, h approaches unity when H is diffusion limited and increases with increasing contributions from mass transfer. In electrophoretic separations, velocity is dependent upon the applied electric field strength (E), the electroosmotic mobility (Pea). and the electrophoretic mobility of the individual species (HA and us) such that VA -E(p.eo+p.A) (5) v3 -E(ueo+u3) (6) Analogously to diffusion coefficient (Equation 3), the observed velocity of the reactive zone is the weighted average of VA and v3 -VE+KVB- - 119+ng 7 V 1+K E(”e°+ 1+K H In order for the reaction in Equation 2 to be evaluated by the plate height model one or both species must exist in the mobile phase. If both A and B exist in the mobile phase, then VA and v3 are non-zero and the reaction is‘evaluated with the plate height model via capillary electrophoresis (CE). However, if one species is in the mobile phase (A) and the other is immobilized (B), then only VA is non-zero and the reaction is evaluated with the plate height model via capillary electrochromatography (CEC). Consequently, the mass-transfer contributions to 228 plate height in CE and CEC influence the maximum calculable rate constants (kmax) differently. Therefore, the theoretical limitations for CE and CEC are not necessarily identical and should thus be evaluated separately. 7.3 CAPILLARY ELECTROPHORESIS As stated previously, theoretical limitations to evaluating reactions via the plate height model in CE are dictated by the physical parameters of the instrumentation and solutes as well as by the chemical parameters of reaction. Each of these factors influences H, and He differently and is therefore considered individually. Values of km, are identified by h = 5, the minimum value at which plate height is mass-transfer limited and inversely proportional to rate constant. 7.3.1 Effect of Electrophoretic Mobility Difference on km in Reactive CE FIGURE 7.1 is a graph of normalized plate height as a function of rate constant (10’2 s k s 10‘ s") and electrophoretic mobility difference (0.1 s Ap. s 2.0 x 10“ cm2Ns). It is apparent that the range of rate constants over which h 2 5 increases with Au. This means that smaller values of Au result in diffusion- limited plate heights _at slower rate constants, whereas larger values of Au permit calculation of faster km”. It is also apparent that increasing Au by 10 increases kmax by 100 and that the maximum rate constant calculable for these conditions is ~30 s". It is also apparent that for more realistic values of Au (5 1.0 x 10“ cm2Ns), the maximum rate constant calculable is s 10 3". Thus, to whatever extent possible, Au should be maximized to increase kmax. 229 FIGURE 7.1: Effect of electrophoretic mobility difference on maximum calculable rate constants via capillary electrophoresis. Calculation conditions: E = 500 Vlcm, (robs = 6.0 x 10" cm2Ns, (“Amp/2 = -1.5 x 10" cm2Ns, o = 1.0 x 10‘5 cm2/s, K =1,10'2 < k <104 s", and All. =1.o x 10*" cmst (D), 2.0 x 10'5 cm2Ns (A), 5.0 x 10'5 cm2Ns (O), '1.0 x 104 cmst (I), 2.0 x 10*4 cm2Ns (A), 5.0 x 10* cm2Ns (c). h = 5 (- -). * 230 FIGURE 7.1 be kzfimzoo ME». , 82:. 82 2: 2 F to . 5o .. cor ooov oooor oooooe iHOIElH 2|le CEIZI'IVINHON 231 7.3.2 Effect of Electric Field Strength on kmax in Reactive CE FIGURE 7.2 is a graph of normalized plate height as a function of rate constant (10'2 s k s 10‘ s") and electric field strength (100 s E s 1500 Vlcm). This range of electric field strengths encompasses the reasonable operating range for conventional scale (capillary) and chip-based electrophoresis. It is apparent that the range of rate Constants over which h 2 5 increases with E. This means that smaller values of E result in diffusion-limited plate heights at smaller rate constants, whereas larger values of E permit calculation of larger km”. It is also apparent that increasing E by a factor 3 increases kmax by 10 and that at even the largest values of E k...ax is only ~ 100 s". Thus, increasing electric field strength by employing shorter columns orlchips has a greater influence on kmax than Au. 1 - 7.3.3 Effect of Equilibrium Constant on k...." in Reactive CE _ FIGURE 7.3 is a graph of normalized plate height as a function of rate constant (10'2 s k s 104 s") and equilibrium constant (10'? s K s 102). It is I apparent that the range of rate constants over which h 2 5 increases as K approachesunity and is mirrored symmetrically as K deviates from unity. This means that kmax is greatest when K ~ 1 and, to whatever extent possible, efforts should be made to impose this condition. Although little can be done to adjust K to unity for true first-order reactions, it is possible to manipulate K to unity for * pseudo-first-order reactions by judicious control of concentration. 7.3.4 Effect of Diffusion Coefficient on km“ in ReactiveCE FIGURE 7.4 is a graph of normalized plate height as a function of rate 232 FIGURE 7.2: Effect of electric field strength on maximum calculable rate constants via capillary electrophoresis. Calculation conditions: it,” = 8.0 x 10“ cm2Ns, [LA = -1.0 x 10'5 cm2Ns, (is = .2.0 x 10'5 cm2Ns, 0 =1.0 x 10‘5 cm2/s, K =1,10'2 < k < 104 s", and E = 100 Vlcm (c), 200 Vlcm (A), 500 Vlcm (I), 1000 Vlcm (O), 1200 Vlcm (A), 1500 Vlcm (Cl). h = 5 (- -)- 233 FIGURE 7.2 oooow coo _. Ase wpzfiwzoo mEE ooe or e prP - —:-pr - p ...o —.:-n p b - 5.0 _ F o_. oo_. coo _. ooooo _. .LHOIEIH SlV'Id GEIZI'IVINHON 234 FIGURE 7.3: Effect of equilibrium constant on maximum calculable rate constants via capillary electrophoresis. Calculation conditions: E = 500 Vlcm, n... = 8.0 x 10“ cm2Ns, in = -1.0 x 10'5 cm2Ns, (is = -2.0 x 10'5 cm2Ns, o = 1.0 x 10“5 cm2/s, 10'2 < k < 10‘ s", and'K = 0.01 (o), 0.1 (A), 1 (I), 10(0), 100 (A). h = 5 (- -). 235 FIGURE 7.3 9-3 kzfimzoo m5». 82: 82 8F 2 V e to 8.0 4!... . _ l1. . C... . . .....ee. . 5:... . F m woe W2: m oooe. ooooe 236 lHOIEIH HIV—Id GEIZI'IVINEION FIGURE 7.4: Effect of diffusion coefficient on maximum calculable rate constants via capillary electrophoresis. Calculation conditions: E = 500/ Vlcm n... = 8.0 x 10* cm2Ns, m = -1.0 x 10'5 cm2Ns, (is = -2.0 x 10'5 cm2Ns, K = 1, 10'2 < k < 10‘ s", and o = 1.0 x 10“5 cm2/s (c), 2.0 x 10‘ cm2/s (A), 4.0 x 1045 cm2/s (I), 6.0 x 10*5 cm2/s (O), 8.0 x 10" cm2/s (A), 1.0 x 10'5 cm2/s (El). n = 5 (- -). 737 FIGURE 7.4 oooov ooor Ase ezfimzoo we». Do? or F ooooow .LHDIEIH SiV‘ld oaznvwao~ 238 constant (10‘2 s k s 10‘ s") and the observed diffusion coefficient of the reactive zone (0.1 s D s 1.0 x 10‘5 cm2/s). It is apparent that the range of rate constants oirer which h 2 5 increases as D decreases. This means that larger values of D result in diffusion-limited plate heights at smaller rate constants, whereas smaller values of D permit calculation of larger km. It is also apparent that decreasing D by a factor of 10‘.increases km proportionately and that the maximum rate constants calculable for these conditions is ~ 100 s". Thus, to whatever extent possible, D should be minimized to increase kmax. The preceding evaluation for CE illustrates the effects of Au, E, K and D on kw. It is shown that km, is influenced most by physical parameters (E, An, and D) and least by chemical parameters of the reaction (K). It is also shown that the largest values of kmax are only attainable when reactions are characterized by first-order or pseudo-first-order equilibrium constants of unity. Of these, the greatest values of kmax are achieved under high electric field strengths for species that exhibit large mobility differences and small diffusion coefficients. Of these influences, there is greater experimental control over E and K (when pseudo-first-order); however, application of large E should be chosen judiciously to avoid excessive Joule heating and buffer decomposition. 7.4 CAPILLARY ELECTROCHROMATOGRAPHY In this section, the individual contributions to Hi, and H.,- are again considered separately, but for the case where Equation 2 represents the transition of a molecule existing in solution (A) and at a surface (B). In CEC, this 239 transition is manifest as species A traveling in solution with VA > 0, and species B remaining on the surface with v3 = 0. As shown below, this distribution of velocities results in kmam values that are considerably larger than those obtainable via CE. 1 7.4.1 Effect of Electrophoretic Mobility Difference on k.m in CEC FIGURE 7.5 is a graph of normalized plate height as a function of rate constant (10'2 S k s 104 s") and electrophoretic mobility difference (-1.0 5 Apt 5 1.0 x 10" cm2Ns) where Ap. = [1A because its - 0. For this evaluation, it is assumed that surface diffusion is negligible (D9 = 0). It is apparent that the range . of rate constants over which h 2 5 has little dependence on Air. This is in contrast to the large dependence of km on An in CE. It is also apparent that kmax for CEC (~ 1000 s") is much greater than kmax for CE (~ 10 s") with the same values of Au and (1.0. This is because in CEC bulk flow contributes to Av (Equation 4) and this contribution dominates kmax. The minimal effect of Au notwithstanding, it can be seen that large, positive values of Au permit calculation of slightly larger km, than large, negative values of Au. Thus, to whatever extent possible, Av should be maximizedvia 090110 increase kmax. 7.4.2 Effect of Electric Field Strength on km, in CEC FIGURE 7.6 is a graph of normalized plate height as a function of rate constant (10'2 s k s 104 s“) and electric field strength (100 s E s 1500 Vlcm). For this evaluation, it is assumed that surface diffusion is negligible (D3 = 0).- It is apparent that the range of rate constants over which h 2 5 increases with E. This means that smaller values of E result in diffusion-limited plate heights at smaller 240 FIGURE 7.5: Effect of electrophoretic mobility difference on maximum calculable rate constants Via capillary electrochromatography. Calculation conditions: E = 500 Vlcm, it... = 8.0 x 104 cm2Ns, D). = 1.0 x 10'5 cm2ls, K = 1, 10'2 < k < 10" s", and a. = 1.0 x 104 cm2Ns (1:1), 5.0 x 10‘5 cm2Ns (A), 0.0 x 10'5 cm2Ns (O), -5.0 x 10'5 cm2Ns (A), .10 x 104 cm2Ns (I). h = 5 (- -). 241 FIGURE 7.5 0000—. -:-. - u - Ase Ezfiwzoo ME». 009. oo_. o_.. _. ...0 5.0 —.-..-p. b run... . n :-.-.P. - bE-pb-p. - -:-PF- - I I I I I I I I I I I I I I I I I I I I I I I I I I I I I I ll _. o _. oo _. ooo _. oooo _. ooooo F 000000 _. .LHOIEIH arena CEIZI‘IVINEION 242 FIGURE 7.6: Effect of electric field strength on maximum calculable rate constants via capillary electrochromatography. Calculation conditions: pea = 8.0 x 10“ cm2Ns, px = -1.0 x 10'5 cm2Ns, oA =10 x 10'5 cm2/s, K =1,10'2 < k <1o‘ s", and E = 100 Vlcm (O), 200 Vlcm (A), 500 Vlcm (I), 1000 Vlcm (O), 1200 Vlcm (A), 1500 Vlcm (1:1). h = 5 (- -). 243 FIGURE 7.6 1 0000000 1000000 2: 1 00000 1 0000 1 000 1 00 1 0 lHOIEH BLV'Id CIEIZI'IVINHON 244 I IIIU I I rIIIIu T I IIIIIII j I I rIIIIl I I IIIIIIr 1 00 1 000 1 0000 10 RATE CONSTANT (s“) 0.1 0.01 rate constants, whereas larger values of E permit calculation of larger kmax. It is also apparent that increasing E by a factor 3 increases km by 10, similar to the trend shown for CE. However, the largest values of E yield km...x ~ 100 s‘1 for CE whereas the same values of E yield kmax ~ 10,000 s'1 for CEC. The reason for this discrepancy is because bulk flow contributes linearly to Av in CEC whereas bulk flow has no contribution to Av in CE. Thus, increasing bulk flow via the electric field strength by employing shorter columns or chips has a greater influence on kmax than Au. 7.4.3 Effect of Equilibrium Constant on km, in CEC FIGURE 7.7 is a graph of normalized plate height as a funCtion of rate constant (10'2 s k s 10‘ s") and equilibrium constant (10'2 s K s 102). For this evaluation, it is assumed that surface diffusion is negligible (D3 = 0). It is apparent that the range of rate constants over which h 2 5 increases with K. It is also apparent that the effect of increasing K on km. decreases as K increases such that km. ~ 2000 s'1 is achieved by K = 100. These trends are the cumulative manifestation of the effects of K on velocity (Equation 7) and diffusion coefficient (Equation 3). In CEC, larger K values necessarily result in faster observed velocities for solutes with negative electrophoretic mobilities and slower observed velocities for solutes with positive electrophoretic mobilities. Moreover, larger K values also necessarily result in smaller D in CEC because the diffusion coefficient of a molecule on a surface (D3) is significantly less than that for a molecule in solution (DA). While some surfaces may permit molecules to diffuse within the phase, those diffusion coefficients are still quite small (D3 ~ 10’7 245 FIGURE 7.7: Effect of equilibrium constant on maximum calculable rate constants via capillary electrochromatography. Calculation conditions: E = 500 Vlcm, pea = 8.0 x 10" cm2Ns, HA = -1.0 x 10'5 cm2Ns, o. = 1.0 x 10'5 cmz/s, 10'2 < k < 104 s", and K = 0.01 (c), 0.1 (A), 1 (I), 10(0), 100 (A). h = 5 (- -). 246 FIGURE 7.7 oooow ooo _. be wezfiwzoo mead K 00.. —::- . h u or {nut - . p F , ..=.- p . p ...0 pp: pF- . - 5.0 _ _. o _. oo _. ooo _. oooo _. ooooo _. . ooooooe lHOIHH EIlV'Id CIEIZI'IVINEION 247 cm2/s). However, the dependence of km, on K for 0 < D3 < DA can be expected to have a similar effect as for CE, but with the inflection at K > 1 and km ~ 1000 — 2000 s". This means that surface phases that minimize surface diffusion will permit calculation of larger km. . 7.4.4 Effect of Diffusion Coefficient on km, in CEC FIGURE 7.8 is a graph of normalized plate height as a function of rate constant (10'2 s k s 104 s") and the diffusion coefficient. of species A (0.1 5 DA 5 1.0 x 10'5 cm2/s). It is apparent that the range of rate constants over which h 2 5 increases as DA decreases. This means that larger values of DA result in diffusion-limited plate heights at smaller rate constants, whereas smaller values of DA permit calculation of larger km”. It is also apparent that decreasing Diby a factor of 10 increases kmax proportionately, similar to the trends shown for CE. However, in CEC km. is extended to 10,000 s‘1 for the smallest values of DA whereas in CE km is extended only to ~ 100 s". 7.5 CONCLUSIONS This chapter discusses the physical and chemical contributions to the plate height model in terms of their individual effects on the maximum calculable rate constants in CE and CEC. It is shown for both CE and CEC, that the physical contributions from electric field strength, electrophoretic mobility difference, and diffusion coefficient have greater influence on km. than the chemical contributions from equilibrium constant. The greatest influence on km. 248 FIGURE 7.8: Effect of diffusion coefficient on maximum calculable rate constants via capillary electrochromatography. Calculation conditions: E = 500 Vlcm, 1130 = 8.0 x 10‘ cm2Ns, m = -1.0 x 10'5 cm2Ns, K =1, 10'2 < k <104 s", and 0. =1.0 x 10‘6 cm2/s (c), 2.0 x 10’5 cm2/s (A), 4.0 x 10" cm2/s (:1), 5.0 x 10‘6 cm2/s (e), 8.0 x 10‘ cm2/s (A), 1.0 x 10‘ cm2/s (I). h =5(--). 249 FIGURE 7.8 $3 hzfiwzoo ME». 0000—. coo ... 00.. or . _. _..o 5.0 huh-n. r ...-... . r p —.:.p- n F ub-pnPL _ . -.:.-L n p ::.. - . - _ e - or oer oooe ooooe oooooe m ooooooe 4.- lable rat Billion“ .0 x ii '1) 6"; ;(Il.h _ ooooooov .LHOIEH ELV‘Id OEIZI‘IVWHON 250 comes from E. This is fortuitous because it the parameter over which there is the greatest degree of experimental control. Whereas the relative effect of E on k...am is identical for CE and CEC, larger km,m can be achieved via CEC because of additional contributions from bulk flow to Av. The effect of Av on km is evaluated via Ap. for CE and via ILA for CEC. It is shown that the effect of Au in CE has a greater influence on kmax than the effect of m in CEC. However, CEC permits calculation of larger kmax because of the additional contribution to Av from bulk flow. Evaluation of the effect of K on kmax shows marked differences between CE and CEC. The largest kmax for CE is achieved at K = 1 and all other values of km. are distributed symmetrically as K deviates from unity. In contrast, the largest km for CEC is achieved at K > 1 and the value of K at which kmax is greatest approaches unity as D3 approaches DA. It is also important to note that CE permits investigations of second-order reactions via imposition of pseudo- first-order conditions by controlling concentration. In CEC, pseudo-first-order conditions are assumed because the number of surface sites is considered to be in excess. However, manipulation of pseudo-first-order equilibrium constants is more difficult for CEC than for CE. This is because it is more difficult to quantitatively vary of the number of surface sites in CEC than it is to quantitatively vary the concentration in CE. The effect of the observed “diffusion coefficient is evaluated via D for CE and via DA for CEC. It is shown that the relative effect of D on k"...x is identical for CE and CEC. However, larger km. can be achieved via CEC because only species A is the dominant contributor to diffusion. Thus, the plate height model may be used to investigate rate constants ~ 251 in CE and CEC. Smaller rate constants can be calculated via CE for first-order and pseudo-first-order reactions, whereas larger rate constants can be calculated via CEC. However, CE permits greater control over pseudo-first-order equilibrium constants than CEC. 252 CHAPTER 8: CONCLUSIONS AND FUTURE DIRECTIONS 8.1 INTRODUCTION In the 25 years since Jorgenson and Lukacs demonstrated the potential efficiency of capillary electrophoresis (CE) [1], it has become a staple for rapid separations of charged analytes. However, applications of CE for investigating reactions have only recently come into popularity. This delay is, in large part, the result of deficiencies in theoretical development more so than deficiencies in ‘ instrumentation. Whereas mathematical models for evaluating reactions in CE have been developed, much of this development has been limited in scope to evaluating thermodynamic quantities [2-4]. Rather, it is the lack of theoretical development for evaluating kinetic quantities that has been the true limitation. Moreover, the methods for evaluating kinetics in CE have historically relied on the presence of plateau profiles and the computational fitting of those profiles. Additionally, these methods and have little to no relationship with the models for evaluating thermodynamic quantities [5-8]. . The work presented in this dissertation attempts to bridge the existing gap between models for evaluating thermodynamics and models for evaluating kinetics. In doing so, physical and theoretical descriptions were presented of the thermodynamic and kinetic contributions to separation and broadening that are consistent with chromatographic theory. In Chapter 2, stochastic (Monte Carlo) simulations were used to illustrate the physical (electrophoretic mobility) and 253 chemical (equilibrium constant and rate constants) contributions to zone profiles, velocity, plate height, and skew. In Chapter 3, theoretical development of the plate height model was extended to reactive CE. It was through this theoretical development that zone velocity was related to equilibrium constant and zone plate height was related to rate constants. In Chapter 5, the plate height model was applied to the reversible, first-order isomerization of peptidyl-proline dipeptides. The equilibrium and rate constants calculated via the plate height model were compared to those calculated with ChromWin as well as to literature values and shown to be in good agreement. In Chapter 6, the plate height model was applied to the reversible, second-order reaction of dansyl phenylalanine inclusion by B—cyclodextrin. Equilibrium constants were calculated from velocity and shown to be in good agreement with literature values. In contrast, rate constants were calculated from plate height and are shown to be grossly underestimated. This unsuccessful application of the plate height model was attributed to physical limitations of the experimental system that are not expected . to be overcome under realistic CE operating conditions. In Chapter 7, these limitations of the plate height model were evaluated for capillary electrophoresis and capillary electrochromatography. This chapter discusses avenues of further theoretical and experimental development for investigating reactions via CE and potential difficulties of each. 8.2 EVOLUTION OF REACTIVE ZONE PROFILES The work in Chapter 1 provided a detailed physical description of how 254 zone profiles evolve during reaction. However, what is lacking is an accurate mathematical description of the process of separation and broadening prior to steady state. In order . to achieve this, it is first necessary to develop a mathematical description of the evolution of the profiles representative of the individual species. This latter task requires a description of how both the mean and variance change for the individual solute zones as a function of time. The crux of the matter will lie in developing a consistent mathematical treatment for the effects of kinetics on the mean and variance for the individual solute zone. Once this is achieved, it should be a relatively simple matter to apply the descriptions of the individual solute zones to the net zone. The end result of this is a complete and continuous mathematical model of how reaction influences the evolution of zone profiles. 8.3 THEORETICAL DEVELOPMENT In addition to further development of the physical and mathematical descriptions of how zone profiles evolve as a result of reaction, work is needed to extend and validate the plate height model to consider multiple equilibria. Although the theoretical development .in Chapter 2 advances the range of accessible rate constants over previous models, it too is limited in scope. This development of the plate height model is limited to systems at or near steady state that can be described by a single equilibrium expression. Steady state is a fundamental assumption of the plate height model; however, there are no fundamental assumptions regarding the number of equilibria. In order to be truly 255 general and more applicable to a broader range of chemical systems, the model should be extended to consider multiple equilibria. It is not a difficult task to describe zone velocity as a function of multiple equilibria. The crux will lie in accurately describing plate height from multiple equilibria and extracting kinetic rate constants from the experimental data. Experimentally, this should be possible by treating each equilibrium expression independently; however, application of such an approach will be predicated upon large electrophoretic mobility differences among species as well as large differences (2 10‘) between the individual equilibrium constants. The stochastic simulation described in Chapter 2 is the ideal candidate to investigate the effects of multiple equilibria on zone broadening and to validate the proposed extension of the plate height model. i 8.4 INSTRUMENT DESIGN The mathematical model developed in Chapter 3 illustrates that increasing the electric field strength increases the range of calculable rate constants. Thus, calculation of rate constants in excess of ~ 20 s'1 requires shorter capillaries, larger applied voltages, or both. This optimization of instrumental parameters is best realized with chip-based CE systems. Chip-based systems facilitate separations over distances a small as a few centimeters with electric field strengths up to ~ 1 kV/cm [9]. This combination of large electric field strengths over extremely short distances will extend the range of calculable rate constants to ~ 103 s". The crux with performing these measurements will lie in the 256 appropriate choice of chip substrate to ensure electroosmotic flow and effective heat dissipation, fabrication of uniform separation channels, and positioning of detection elements within extremely close proximity (~ 1 cm). 8.5 EXPERIMENTAL DESIGN The unsuccessful application of the plate height model in Chapter 6 illustrates an additional physical limitation that cannot be overcome with instrumentation alone. The upper limits of calculable rate constants are dictated not only by electric field strength and detector locations, but also by the inherent mobilities of solutes in the free and bound states. In order to control and maximize the effect of mobility on plate height it will be necessary to affix one of the reactants to a stationary surface and perform the experiments as capillary electrochromatography (CEC). The most ideal approach to these experiments would be to affix one reactant to the walls of open channels in chip CE, thereby extending the range of rate constants to ~ 104 s“. Chip CEC has been performed with packed channels and monoliths [10]; however, these stationary phase supports introduce additional contributions to broadening. These experiments will share the same potential difficulties associated with choice of chip substrate and positioning of detection elements discussed in the previous section. Moreover, an additional challenge in these experiments will arise from coating the channels with the stationary phase [11] and determining the mobilities of any non-neutral solutes under identical conditions a priori. 257 10 11 8.6 REFERENCES Jorgenson, J. W.; Lukacs, K. 0., Anal. Chem. 1981, 53, 1298 Busch, M. H. A.; Carels, L. B.; Boelens, H. F. M.; Kraak, J. C.; Poppe, H., J. Chromatogr. A 1997, 777, 311 Busch, H, M. A.; Kraak, J. C.; Poppe, H., J. Chromatogr. A 1997, 777, 329 Jia, 2., Curr. Pharm. Anal. 2005, 1, 41 Avilla, L. 2.; Chu, Y.-H.; Blossey, E. C.; Whitesides, G. M., J. Med. Chem. 1993, 36, 126 Okhonin, V.; Krylova, S. M.; Krylov, S. N., Anal. Chem. 2004, 76, 1507 Patterson, D. H.; Harmon, B. J.; Regnier, F. E., J. Chromatogr. A 1996, 732, 119 . Schoetz. Trapp. 0.; Schurig, V., Electorphoresis 2001, 22, 2409 Dolnik, V.; Liu, S., J. Sep. Sci. 2005, 28, 1994 Stachowiak, T. B.; Svec, F.; Frechet, J. M. J., J. Chromatogr. A 2004, 1044, 97 Dolnik, v., Electrophoresis 2004, 25, 3589 258 APPENDIX 259 APPENDIX: VALIDATION OF INSTRUMENT CONTROL PROGRAMS AND CAPILLARY ELECTROPHORESIS POWER SUPPLY This appendix includes descriptions of the function and validation of the programs written to operate the capillary electrophoresis with laser-induced fluorescence detection (CE-LIF) system described in Chapter 4. These programs were written with cOmmercially available software (Laerew, Version 5.1, Nation Instruments) and executed on computers running the Microsoft Windows 98 operating system. Additionally, this appendix includes a description of the validation of the high voltage power supply (HVPS) used to achieve separations with this system. FIGURES A.1A and A.1B show a typical screen view and program schematic of the capillary electrophoresis power supply control program (step- logic2). The program allows the user to specify the time, duration, polarity and magnitude of applied voltage or current. The program also allows the user to update the specified values in real-time as well as monitor the resultant current or vouage. FIGURES A.2A and A.2B ’show a typical screen view and program schematic of the Program for Unified Computer Control (PUnCC) used to control and monitor the separation progress. Signal from the detectors is monitored in real-time and written to a file with the PUnCC. Additionally, the PUnCC is used 260 FIGURE A.1A: Typical screen view of the-step-logic2 program. The toggle switch allows the user to choose either voltage or current control of the HVPS. The digital controls labeled “T1” - “T6” allow the user to specify the elapsed time at which voltage or current changes occur. The buttons labeled “set T4” - “set T6” allow the user to choose additional (> 3) voltage steps. The digital controls labeled “kV1” - “kV6” and “uA1” - “uA6' allow the user to specify the magnitude of the applied voltage or current. The digital control labeled “T max" allows the user to specify the maximum time for voltage or current application. The digital indicator labeled “Time” shows the elapsed experiment time. The digital indicator labeled “Output" shows the voltage output from the computer to the HVPS. 261 {if step FIGURE A.1A 262 FIGURE A.1 B: Schematic of the step-logic2 program. The schematic shows the data flow as a logical cascade from T1 through T6 to the output channels that interface with the HVPS. 263 FIGURE-A.1B l) "i :i lmiu ,' v. Ul 1:11.111. ' $~ throat” 264 FIGURE A.2A: Typical screen view of the PUnCC program. The dialog box labeled “Input Comments” allows the user to add comments regarding each experiment to the data file. The toggle switch allows the user to monitor either the current or voltage from the HVPS. The digital controls labeled “Buffer size”, “Scan rate”, and “Scans to read at time” allow the user to specify the number of scans stored in the data buffer, the number of scans recorded each second, and the number of scans written to a file, respectively. The button labeled “New file” allows the user to choose to store the data as a new file or to append the data to an existing file. The digital control labeled “Input limits” allows the user to specify the board gain. The digital indicators labeled “Scan backlog“ and “Time Elapsed” show the number of scans in the buffer and the elapsed time, respectively. The button labeled “STOP” allows the user to stop the experiment. The waveform charts and digital indicators show the real-time response from the detectors and the monitored current or voltage from the HVPS. 265 p Pllnl‘l‘. v- ' FIGURE A.2A 266 FIGURE A.28: Schematic of the PUnCC program. The schematic shows the data flow of specified inputs from the screen view to the data acquisition loop and the data file. The schematic also shows the interface to the step-logic2 program as a sub- routine (labeled “V or I Step”) within the data acquisition loop. 267 FIGURE A.28 iUIIIEo-oioi.‘ 533.333.282.338... 3583838286833..- 51533138295853... iosmggg‘oc atlc shirt newliii malls it asas'i' ES. Williint . c.3355 {Qua—E A 268 to control and monitor the HVPS via the program described above as an embedded sub-routine. The PUnCC allows the user to specify whether voltage or current is applied for the separation, the rate of data acquisition, the rate at‘ which data are written to the display and to a file, as well as the gain of the data acquisition board prior to each experiment. The PUnCC also allows the user to add pertinent comments to each data file. Output from the HVPS and its control program was validated with a 1000:1 high-voltage probe (Fluke) and a 495 (:1: 2%) Mn precision resistor. FIGURE A3 is a graph of the programmed voltage applied versus the output voltage measured by the high-voltage probe, together with the percent error. It is apparent that there is a linear relationship between the programmed voltage and output voltage with a slope of unity. It is also apparent that the output voltages are within the calculated upper and lower output limits specified for the power supply and that the output voltage error is greatest (> 2%) below 1 kV and least (< 2%) above 1 kV. ‘ FIGURE AA is a graph of the programmed voltage applied across the 495 - MD resistor versus the measured current, together with the percent error. It is apparent that there is a linear relationship between the programmed voltage and the measured current. The resistance calculated from the slope is 508 MO, within 2.1% of the expected value of 495 MO. It is also apparent that the ' measured current error is greatest at lower programmed voltages and decreases monotonically with increasing voltage. FIGURE A5 is a graph of the programmed current applied through the 269 FIGURE A.3: Programmed voltage versus the output voltage (—) and percent error of the output voltage (X). Dotted lines represent the specified output accuracy for the power supply. 270 imH“ Dotted it 1116705 FIGURE A.3 80883 .LNEIOEIEd l l l l l l l l l I) 4’ l) 4) 4) O O ‘p O ‘ n O 4} O O O 0 O O (D {D O O ‘. O 9 D C O O O O O O 0 ‘p C " O O .. O O O O O O O O O O O O O O Q I O O O O O O O O O Q ‘ ‘ 0 O \ O O O O 0 0 O O O 0 0 O O \ O O \ O O O O O O O O O O O O Q O O O O C O O O ‘.\ v H,"‘\ 9,.“ .3- .._:'3‘ \ A‘w ..=- A, ‘—-_.___ _ ..‘.—-_u: A“. I .:—_' LO V (‘0 N ‘- (NI) aovriorx indinO 271 3 4 5 2 PROGRAMMED VOLTAGE (kV) -—i—F FIGURE A.4: Programmed voltage versus the measured current (El) across the 495 MO precision resistor and the percent error (X) of the measured current calculated assuming a resistance of 495 MO. 272 FIGURE A.4 9o; mo 8.252800% m s c m e m N e — b — b — _ _ p 0) XP 3‘ H L‘ L‘ 1 LG I O ‘— I la ‘— i O N (Vii) tNaaano oaansvaw 273 _ ‘ FIGURE A.5: Programmed current versus the measured voltage ([3) across the 495 MO precision resistor the and percent error (X) between the measured voltage and that calculated. assuming a resistance of 495 M0. 274 FIGURE A.5 on n 2.3 Summso omzzéoomo o... 8 - on .3 2. .13 .3 N N ‘— ‘— (A>1) aovr10A oaansvaw fl 0 (O m . m .. 275 495 MO resistor versus the measured voltage, together with the percent error. It is apparent that there is a linear relationship between the programmed current and the measured voltage. The resistance calculated from the slope is 508 M0, ~within 2.1% of the expected value of 495 MO. It is also apparent that the measured voltage error is greatest when the programmed current is less than 5 [LA and remains relatively constant at < 5% when the programmed current is > 5 3A. The ability of the PUnCC to accurately record data (0 - 10 V input) was validated with a Dial-A-Volt voltage source (Model DAV-450, General Resistance Inc.) and a Fluke Digital Multimeter 77 (DMM). FIGURE A6 is a graph of the specified voltage versus the voltage measured by the PUnCC and the DMM, together with the percent error of the voltage measured by the PUnCC. It is apparent that there is negligible difference between the voltages measured by the PUnCC and the DMM and that the percent error of the voltage measured by the PUnCC is 1% or less. 276 FIGURE A.6: Input voltage versus the voltage measured by PUnCC (Cl), DMM (A), and the percent error (X) of the voltage measured by PUnCC. 277 ie Wit? FIGURE A.6 aCaaa .LNEIOHEch 8 O) Q I\ (0 l0 V (‘0 N O l l l l i J l l I a ;1 :1 :1 3 :1 3 3 . a :1 :1 :1 :1 l O C) ‘— (A) EIOVJJOA oaansvaw 278 9 10 8 INPUT VOLTAGE (V)