I 312%.? ' ‘II‘ I .V I2 *I‘ ' I .‘I‘Il-Z in ,I pnl'flp. .1. "I; , :I'J‘. " :‘4‘ ’l; I I. I. II. .IIIIII ..... I "'I I :. “315k; 4-: .‘ 'lI'" '.~ I” ." ’: hiilV.‘ I“ I'|'I\l" V I-" I '7 I " J II . II . I. IIIII - .\ IIIII I. f I . IVIII In I II II or. LI I . III .III I I I .vI I. III... .. 3. v. I... w. - . I n I III. I <0 I .F III. \II I III “IR-III: I .vIIIII II I. JI I I II<.I MN” -\ (VP: I. Md - I VnIIv IIIIIQ .lkv I v I 4. v4 LII-l h- Osv IL I . o I I . II. . o AulI 5‘... It I. I I III I a III I II I laI I OIII I. l I I I I I .0 I‘ll: r . V0. III, I; I II (War-I II I. . IV Ofi I: not .I J, I III: III C I. I. I‘ I I I ‘5'. . IIILII .000!!! In... I. III; III I. 1.! I .1. .IMI I....IJ IIII..II.I.I: II. I. . \. III. IrIPI. I 1295. .i. U” I .IIII. .. I. I {II I3”. V IIII v .I I I .II. 0 AI“ .- III. A I . I. I. .. I I. I 2"“. I '0.- ‘ I' ?.I.11LII II VIIVI I.‘ .. I I I. III II IVIII I II I-.. a. I. I.. . I I . I P .I... .III .I I... .II I. II I]. k'I I I I {IQ I II I}! . III. II I I. . .I III V I VI II. III I .I I II -.III ..I I I I . I I. I II I I . I!” III, I3 I II JIMMIIII iq}I I...“ I II II! I I” III! ..... JIIII I JII II .IJ II It, .I III, I0 I. ..... III—II .IIIIIPIHNII. I I II“ I 1...”! l- I I I / II I I I, I I I. I IIyIP z I I a I I.“ III. I 1' I II x I. I I Ia}. II I I I. III. I I J In I I I I~ 01. I I I II I I to A I .1 III I Iul t I» L.I I I al A \ II ‘0’ I ’ I I or I, I 0".» II: II N I IJIIII I I .7.- 1| I I I I I I I In I I II I I I? I I-III In I. It I ‘4 I J. . II -I o I. I. II . I. I . II I I I. I. . I III IVII I I IIVII. . IIIMII .. o1 - I K I I Inxr‘l . .I I .II. II... I I I III I III. I I... IIIIIIOT do .7 II- I III. I c INIIII I I I I (Iflj‘a III I INIIIHI . IIIIIIIIIIOIuIIJIIIIIII I III. NIILIIIII III-III III III: . I I rIIIrI. I. ”II. «I I II 94.1? .4! I III 3 II. A ~ .I D .. ‘OII I II' I III. IV'I.I|I I II I! II I I I. I I O I I I I. I ' IIIII I II ‘II III C . . I I I I I. l I 0.5.1 . I. C III I II III. I. I I I I H. I III II). I‘ II. III- II ’ O I 0‘ | Q. A g II I. .. CIVInII II I I I I I \ III. . II .II II IIJquhoIII 7 II. IIIIIIII LIIIIIIIIIIIIIIIIIHUI I .O‘V’I" On I‘ l I I f-I I I. I III ‘3 .VIO. . D I \ .I v v. I I PI» I v I . I I .I I. II III II I I . II I a I I I I III I. I I \ I ’ 0" Ihu' I I I I II I I III/x. I III, I’ll- I I. 1" (I 1 I. I I. I l I III! II II ..I In. I III. II I I I . IIIII IIII' ,I. I O l J I! I .I “ III I III III .I I IIIII I II I III II M I I .I I I . I I l I I II II I I I. 'I l ' .I . - lrII IIIIIINIIIIIIIIIIIIIII I. II IIIIIIIIIIIIIIIIII II II I . .I I I I . I I. I ' - '.l | . V I .I: I. I I ‘ I V I II I I! I I I I I II I 0 III . '. ’ .5. II." o l ' J I n I I I . 1 . III I III I H‘ .II I I I I '0 I I '1 'I II IIII III.I .I I . o \ III I I II o . I’I I II -II I I I I! ‘OIII IOII '. II 'IIQI |-III 0'- I '6 I II I II I I I AIIIIIIf III III III I II I I. II I'I'IIIIII III I I III I I"III'|" .I II. I u I I I, I|\.II I‘ II I III IIIII‘ I.’ I I ’I II I r I I IIII IV4I‘II'IIfIIIII InfloI‘J m” I“ ”II I“ ‘II II I7I‘I 'II‘ It'll I I I .IIIu III IIIIIIII “III! I III, I II I III III. (III. I III. I III! . III II I. , III II I] IWII... I .II" I I IIIIIIINI .II IIIo»- IIIIIIIIfl II . ‘I’IIOIIIIII’II’I".II- I ,I II 11" , t’I‘ II I ‘13- if I III II III; II I II.I.. IIIIIIIIJLIH.” IIIIII IIIIIII IL III I .IIIIIIflIIII IIIII I I? II III III) III“. .IIIII IIIIIIIIIIIIIIIIIIIII. III InInIIIII IIIHII I I III II II I I III III , I .I III (II |.II. .I HI H V. I III, III I--. III I’II . IIICW I I .o I. “I I? I II I. II I I ..IIII..IU III III! I. II. I. II‘IIIIv In. .I III II £1ng IIIIIHI I I .IIIIIIIIIIIIINTI I. I I III II I ' III. (III! II I (7.,II II 1 \MI II ‘ If . II .“II I l l ' I I. II I H I III. I III II . . . IIWHWI I I III I II I I. I I I I I III «IV I: - - I - - u III. III... I--I.. -I. IIIMII IIII I II, . I .I\ I. I. I II- I I I I I I l I I - J I .. I u “I, IIIIWIIIIIIIIIIIFII I [mLIIWIIIJHIIIII IIIII IIIGUIIgnIIIIIIIIIIIHNIIII’III IIIIIIIIHIIIIIIIIIIIIIIIIIIIIIIIII. II I III; I . I I III I I In“ I’ 'l- {L ”III... I. I .IIIIIIIdWIIIIIIIIMIIIIH IIIII gfldfijflrhrf III II “III III .I III III! IIIHI It I IIIIIIII Igfi .IIIIIIII. if I . .IIIII‘II’JIIIJHIINIIJ‘I’ III’IIIIII III,"- I‘IlIi 'IIIIIIII’IIIII IIIIIIIIIIIIIIIIIIIIIIIII‘I Ill-IIIIIIIIIIIIIIIIIIIIIIIIIIII . IIIII IHIUuIIIIIIIIIIIIIIIIIIIIII I!!! III II.II.I§I III: IAIIIIII I la»? V. III- II III III.II l . I”. I‘Llrl I. I II .I I II I - .. .4 A - \V W -Q.$$£ $339 2%» W W l ‘3 (if. ‘n _‘::‘T' \f'i‘ m2?- This is to certify that the dissertation entitled THE CHARACTERIZATION OF NYLON OPEN—TUBULAR IMMOBILIZED ENZYME REACTORS INCORPORATED IN STOPPED-FLOW AND CONTINUOUS FLOW SYSTEMS presented by Robert Q. Thompson has been accepted towards fulfillment of the requirements for PhoDo degree in ChemiStrX War/£4 Major fiofessor nae August 2, 1982 MS U is an Affirmative Action/Equal Opportunity Institution 0- 12771 MSU J LIBRARIES “- RETURNING MATERIALS: Place in book drop to remove this checkout from your record.~ FINES will be charged if book is returned after the date stamped below. THE CHARACTERIZATION OF NYLON OPEN-TUBULAR IMMOBILIZED ENZYME REACTORS INCORPORATED IN STOPPED-FLOW AND CONTINUOUS FLOW SYSTEMS By Robert Q. Thompson A DISSERTATION Submitted to Michigan State University in partial fulfillment of the requirements for the degree of DOCTOR OF PHILOSOPHY Department of Chemistry 1982 ABSTRACT THE CHARACTERIZATION OF NYLON OPEN-TUBULAR IMMOBILIZED ENZYME REACTORS INCORPORATED IN STOPPED—FLOW AND CONTINUOUS FLOW SYSTEMS By Robert Q. Thompson Open-tubular immobilized enzyme reactors are becoming widely useful in the clinical laboratory for specific and sensitive substrate determinations. However, the dependence of apparent enzyme activity on reactor design and mass transport effects has not been well documented. Several open-tubular immobilized enzyme reactors, prepared from 0.1 cm i.d. O-alkylated nylon tubing, were incorporated into flow systems and characterized. The apparent activity of an immobilized glucose oxidase reactor, used in a miniature, 1.0 mm i.d., continuous flow system to deter- mine glucose, was found to be independent of coiling diameter and ionic strength, but was affected by tempera- ture, segmentation gas composition, and the amount of co—immobilized catalase. Sodium azide was found to inhibit catalase, thereby improving the assay sensitivity, and to inhibit glucose oxidase at higher concentrations. To investigate the influence of mass transport on immobilized enzyme kinetics, a novel stopped-flow instru- ment was developed. The enzyme reactor was employed as a photometric observation cell to allow direct and continuous monitoring of the enzyme reaction. The kinetics, as mea- sured under static conditions, are simpler than in other flow systems because only molecular diffusion and/or coulombic forces contribute to mass transport. Kinetics constants for immobilized lactate dehydrogenase are shown to be influenced by radial diffusion in the stopped- flow instrument. The method is also applicable to clinical substrate determinations. Glucose, in the range of 1—10 mM, and lactate, in the range of 5-50 uM, were determined by reaction-rate methods with the instrument. A new enzyme reactor for use in flow-injection analysis was also developed. Glucose oxidase was immobilized on non-porous glass beads and the beads, 0.5 mm in diameter, were packed into a 0.8 mm i.d. Teflon tube. The resulting reactor was incorporated into a flow-injection manifold for the determination of glucose. Calibration curves were very linear over the range of 0.0—2.5 mM glucose, and at least 60 samples/hour could be processed. To Mom and Dad and Jan and all my other teachers ii ACKNOWLEDGMENTS I would first like to thank Dr. Stanley R. Crouch for his helpful advice and help with this and other manuscripts. He also has good taste in baseball teams. Special thanks also go to Dr. Gerald Babcock for serving as my Second Reader, and Dr. Enke andDr. Wood for being members of my Guidance Committee. Charlie Patton deserves a lot of credit. The ”Bubble King" taught me about continuous flow analysis and shared with me a lot of ideas. He was a friend and helpmate in the laboratory. Thanks Chas, I couldn't have done it without you. Thanks also go to Frank Curran for his friendship in and out of the chemistry building (inspite of his poor taste in baseball teams), and Gene Ratzlaff for loaned equipment and the gift of helpful advice. All the Crouch group members who lent a helping hand deserve thanks. The times of play were also very valuable. I thank all my basketball buddies, my golf partners—- Steve M., Steve B., and Pete, my tennis nemesis-— John Leckey, and my good friends-- Rich and Jane Farmer. Several support people deserve hearty thanks: Tom Atkinson for MULPLOT, Dick Menke and Deak Waters, and the graphics artists-- Katherine, Bev, and Jo. Thank you. Finally, I thank my wife, Jan, for everything else. iii TABLE OF CONTENTS INTRODUCTION .... ................... .............. 1 CHAPTER 1 Introduction to Immobilized Enzymes .............. 3 A. Immobilized Enzyme Characteristics ...... 3 B. The Uses of Immobilized Enzymes in Analytical Chemistry ......... ..... ...... 6 CHAPTER 2 Chemical Bonding of Glucose Oxidase to Nylon Tubing 14 A. Nylon Activation ........................ 16 B. Addition of Spacer Molecules ..... ..... .. 25 C. Attachment of Glucose Oxidase ........... 31 CHAPTER 3 Kinetics Of OpGH-TUbUlar Reactors o o o o o o o I o o o o o o o o 37 A. Kinetics Properties of Open-Tubular Reactors ............. ..... .............. 41 1. Electrostatic and environmental effects ... ......... ....... .......... 41 2. Mass transport effects .............. 43 B. Experimental Considerations ............. 51 C. The Determination of Michaelis Constants. 57 1. Introduction ........ ..... .... ....... 57 2. Reagent preparation ................. 59 3. Immobilization of lactate dehydro- genase ........ ...... ..... ..... ...... 60 4. The instrumentation ..... ....... ..... 60 5. Results and discussion .............. 62 D. Analytical Determinations ............... 7O 1. Introduction ........................ 7O 2. Reagent preparation ................. 71 iv E. CHAPTER 4 3. 4. 5. 6. Enzyme immobilization ......... . ..... The instrumentation .... ..... ........ The results of the glucose determination ............ ..... ...... The results of the lactate determination ......OOOOOOOOOOOOOOOOO A New Cell / Reactor Design ............. Colorimetric Determination of Glucose Based on Immobilized Glucose Oxidase ...................... A. The 1. 2. 3. 4. Method and Instrumentation .......... The glucose oxidase reaction ........ The Trinder/peroxidase reaction ..... The continuous flow manifold ........ The continuous flow instrument ...... Experimental Studies of the Trinder Reaction ......OOOOOOOOOOO.....OOIOOOOOO. Reaction Characteristics ... ........ ..... 1. Segmentation gas composition ........ 2. The effect of ionic strength ........ 3. The specificity of glucose oxidase .. 4. Mutarotation kinetics ............... Reactor Design Characteristics ..... ..... 1. Coiling diameter .... ............ .... 2. Coil temperature .................... The Characterization and Elimination of the Effect Of catalase ......OOOOOOOOOOO. 1. The nature of the interference ...... 2. The reagents and continuous flow instrument I...OOOOOOOOOOIOOOOOOOOOOO 3. Enzyme immobilization ............... 4. Effect of azide on aqueous peroxidase 5. Effect of azide on immobilized catalase ......OOOOOOOOOOOOOOOO0.0... 6. Effect of azide on immobilized glucose oxidase ..................... 7. Effect of azide on different preparations of glucose oxidase ..... 8. A method to determine the presence of catalase ..... ....... ...... ....... 9. Conclusions .............. ........... 73 73 77 81 90 97 97 98 99 105 107 114 114 115 118 121 124 124 125 127 127 130 131 131 133 137 140 141 143 CHAPTER 5 Enzyme Immobilization on Non-porous Glass ........ 147 A. Immobilization of Glucose Oxidase on Glass Tubing ......................... 148 B. Glucose Oxidase Bead String Reactor ..... 150 CHAPTER 6 Future Plans .......... ..................... . ..... 161 APPENDICES Appendix A. Derivation of Diffusion-control Rate Equation ..... ............. 164 Appendix B. Derivation of the Absorbance of a Solution with a Radial Concentration Gradient ......... 165 LIST OF REFERENCES .......... ........ . ....... ..... 166 vi LIST OF TABLES Table 1.1. Recent Examples of Enzyme Electrodes............................. ........... 8 Table 1.2. Recent Examples of Open- Tubular Reactors............... ................... 11 Table 1.3. Packed—Bed and Other ReactorSOOOOOO0.00.00.00.000............OOOOOOOOOO 12 Table 2.1. The Results of the Spacer Molecule Studies................... ......... . ..... 26 Table 2.2. The Activities of Nylon Tubular Reactors Resulting From the Incorporation of Various Spacer Molecules....................... ....... ........... 29 Table 2.3. Immobilization Procedure ............. . 35 Table 2.4. Various Enzymes Bonded to Nylon Tubing and Their Approximate Half-liveSOOOOOOOOO....OOOOOOOOOO ....... o ..... 0.0. 36 Table 3.1. Reviews of Immobilized Enzyme Kinetics...................... ............. 38 Table 3.2. Kinetics Values for Aqueous and Immobilized Lactate Dehydrogenase... .......... 68 Table 3.3. Results of Serum Assays for Glucose....................................... 76 Table 3.4. Results of Serum Assays for Lactate....................... .......... . ..... 80 Table 4.1. Chromogens for Reaction with Hydrogen Peroxide ..... ......... .............. 95 Table 4.2. Specificity of Glucose Oxidase.. ....... . ........... . ..................... 120 vii Table 4.3. Solutions Used to Prepare the Immobilized Enzyme Reactors.... ............... 132 Table 4.4. Glucose Concentrations Found in Control Sera With and Without Added Azide....................................... 146 Table 5.1. Immobilization Procedure for Glass Beads................. .......... ........ 152 viii LIST OF FIGURES Figure 1.1. Schematic of an instrument employing an open-tubular reactor.................. Figure 2.1. The immobilization scheme............. Figure 2.2. Reactor activity versus TOTFB reaction time.................. .............. Figure 2.3. Reactor activity versus reactor length for different TOTFB concentrations. The concentrations were 0.1 M (A), 0.05 M (B), 0.01 M (C), and 0.005 M (D)............... ......... Figure 2.4. The precision of the immobilization procedure.......... .............. ... Figure 2.5. The storage stability of reactors prepared with different spacer molecules........... Figure 2.6. Fluorescamine determination of 1’6-diamin0hexane........0...0.0......0. ..... 0.0... Figure 2.7. The effect of initial glucose oxidase concentration on the reactor actiVity.......0...0....................0....00.... Figure 3.1. Immobilized enzyme kinetics........... Figure 3.2. Influence of slow mass transport on enzyme kinetics......0 ........ 0.0....0.. ........ Figure 3.3. Concentration gradients in an Open-tUbUIar reactoroooooooo ooooooooo 000000000 ooooo Figure 3.4. Bolus flow pattern .................... ix 15 20 22 24 27 30 33 39 45 47 50 Figure 3.5. Cross-sectional View of the GCA McPherson stopped-flow mixer/observa- tion cell unit. The outer aluminum housing (a), quartz windows (b), and the main KEL-F body (0) are press fit with three bolts. Mixing occurs at (e), where the two reagent flow streams meet at 90° to each other -- one stream is in the plane of the figure and the other is perpendicular to it. The immobilized enzyme reactor is placed inside the observation cell (d). The dashed arrow represents the light path inside the cell...................... ......... .... 55 Figure 3.6. Stopped-flow colorimeter... .......... 56 Figure 3.7. Typical progress curves ob- tained with the instrument. The dotted lines represent the initial reaction rates calculated from these curves. The reactions represented were of 2.00 mM lactate with 0.425 mM (A), 0.850 mM (B), and 2.12 mM (C) B-NADOC00.00.000.000....0....0...00............... 63 Figure 3.8. Lineweaver-Burk plot of l/Rate versus 1/[Lactate] at four differ— ent B—NAD concentrations: 0.26mM (A), 0.43 mM (B), 0.85 mM (C), and 2.12 mM (D)............................................... 64 Figure 3.9. Lineweaver—Burk plot of 1/Rate versus 1/[B-NAD] at three differ- ent lactate concentrations: 0.80 mM (A), 2.00 mM (B), and 4.00 mM (C)........... ..... ...... 65 Figure 3.10. Plot of y-intercepts from Figures 3.8 and 3.9 versus 1/[Substrate] for different B—NAD and lactate concen- trations. The y-intercept equals 1/Vm, and the x-intercepts yield the two Km values.................................. ...... .... 67 Figure 3.11. Calibration curve for glucose..................0...... 0000000000 . ..... .0 75 Figure 3.12. Calibration curve for lactate........ ..... . ........ ........ ............. 78 Figure 3.13. Design of the new cell/ reactor. Two pneumatically actuated pistons (A) were used to push the slider (B) back and forth between the fill posi- tion (P1) and the Viewing position (P2). Quartz windows were used (C) to contain the solution in the viewing position............... Figure 3.14. Reproducibility of cell positioning. Distilled water was pumped through the cell at 0.46 mL/min. Each trial represents movement of the cell from the viewing position to the fill position and back again....................... Figure 3.15. The results of five pro- gress curve determinations with 0.44 mM glucose as the sample. The 1 mm i.d. reactor was used.................... ..... .......... Figure 3.16. Progress curves obtained with the 1 mm i.d. reactor. The glucose concentrations (mg/L) which were used are written at the end of each curve. ....... ....... Figure 3.17. Progress curves obtained with the 2 mm i.d. reactor. The glucose concentrations (mg/L) which were used are written at the end of each curve............... Figure 4.1. The three forms of D-glucose.......... Figure 4.2. The reactions involved in the enzymatic assays for glucose. .............. .... Figure 4.3. The continuous flow reaction manifold. The pump speed setting which gave nominal flow rates was 42......... ..... . ...... Figure 4.4. The pH dependence of the glucose assay. A 0.1 M phosphate buffer was used to control the pH. The dotted line indicates the approximate relationship in the absence of many data points........................ Figure 4.5. Diagram of the continuous flow instrument.0 .......... ......0........0..0..... xi 82 84 86 87 88 91 93 103 106 Figure 4.5. Progress curves for the reaction of hydrogen peroxide and the Trinder reagent. The peroxide concen— trations were 50 uM (A), 25 uM (B), 10 uM (C), 5 uM (D)...................... ..... .... 109 Figure 4.7. The effect of ascorbate concentration on the assay sensitivity ...... ...... 112 Figure 4.8. First order plots of data for the hydrogen peroxide-Trinder reaction in the presence of ascorbate. The ascorbate concentrations were 10 HM (A), 20 MM (B), 40 HM (C). A was the absorbance at equilibrium, afid A was the absorbance at any time, t........................... .......... ...... 113 Figure 4.9. Calibration curves for a series of glucose standards using three different segmentation gases: NZ, Air, 020.00.00.000000000000......0000000000.000....0... 116 Figure 4.10. The output from the con- tinuous flow instrument using three different segmentation gases. The glucose concentrations were 0.44 mM, 1.44 mM, 2.22 mM, and 4.44 mM. Only the first three concentrations were sampled using 02.................................. 117 Figure 4.11. The effect of ionic strength on the glucose oxidase reaction000.....00000.00.00.000.....0.....0......0119 Figure 4.12 First order plot of data for the mutarotation of 4.44 mM B-D-glucose solution at 22°C. The results of two independent experiments are presented. [8] is the concentration of the B-form of D—glucose at any time, t, while [B] is the initial B-form concentratiog....................... ............ .. 123 Figure 4.13. The effect of coiling diameter on the calibration slope (0) and wash ([3).......... ............ .. ..... ........ 126 xii Figure 4.14. The dependence of the assay sensitivity on temperature....... ..... . ........... 128 Figure 4.15. The effect of azide on the amount of hydrogen peroxide that was con- verted to product. A 79 nM H20 solution was sampled without an enzyme rgactor ([J) and with a 30 cm catalase reactor (0) in the manif01d00000000.00.00.00.0000000000000 134 Figure 4.16. Lineweaver-Burk plot of the data which exhibited azide inhibition of immobilized catalase. The azide concen- trations ranged from 0.0 (A) to 1.5 mM (B)........ 136 Figure 4.17. The effect of azide on three glucose oxidase reactors, type B (A), type C (0), and type D (EJ), containing various amounts of catalase (refer to Table 4.3)........................................ 138 Figure 4.18. Dixon plot of data from Figure 4.17, curve B. Three glucose concentrations were used: 0.44 mM (1), 0.22 mM (2), and 0.11 mM (3)...................... 139 Figure 4.19. The effect of azide on three commercial preparations of glucose oxidase, reactor types G, E, and F (refer to Table 4.3)............... ......... ...... 142 Figure 4.20. The change in calibration slope with different lengths of reactors B, C, and D (refer to Table 4.3). No azide was added....... ........ ............. ....... 144 Figure 4.21. The change in calibration slope with different lengths of reactors B, C, and D (refer to Table 4.3). The reactions occurred in the presence of 0.8 mM azide...................................... 145 Figure 5.1. View of the immobilized enzyme single bead string reactor..... ............ 153 Figure 5.2. Schematic diagram of the flow-injection instrument......................... 155 xiii Figure 5.3. The degree of incomplete mixing at various sample volumes. Four pump speeds were used: 8.3, 12.3, 25.0 and 33.3 uL/sec........................ Figure 5.4. Flow-injection instrument response. The four glucose concentra- tions were: 0.44 (A), 0.88 (B), 1.33 (C), and 2.22 (D) mM................... Figure 5.5. Calibration curve for the flow-injection determination of glucose. xiv ..... ...... 159 .......... 160 LIST OF SYMBOLS AU - Absorbance units AO - initial absorbance (AU) Aco - equilibrium absorbance (AU) - wash value (seconds) - ratio of mean ionic activity coefficients (unitless) — a constant (unitless) reactor diameter (cm) - substrate diffusion coefficient (cmZ/sec) (DUO-000‘ I - charge on the electron (coulombs) [E]o - total enzyme concentration (Molar) 1 h - mass transport rate constant (second ) H(ss) - height of CF peak at flat portion (AU) H(t) — height of CF peak at any time, t (AU) js - Bessel functions (unitless) k - Boltzmann constant (joules °K-1) ke - an enzyme rate constant (second-1) Ki - enzyme interaction constant (Molar) KI — enzyme inhibition constant (Molar) Km - Michaelis constant (Molar) L — reactor length (cm) Nu - Nusselt number [P]b - bulk product concentration (Molar) [P]w - product concentration at the reactor wall (Molar) r - reactor radius [S]b - bulk substrate concentration (Molar) [S]W - substrate concentration at reactor wall (Molar) t - time (seconds) T - temperature (°K) v — initial reaction rate (Molar second-1) V — volume of a single liquid segment (cm3) \hn - maximum rate of an enzyme reaction (Molar second-1) XV electrovalency of the substrate (unitless) effective concentration of charges on the carrier (M) . . —1 —1 molar absorpt1v1ty (AU Molar cm ) ratio of the mass transport rate constants times the ratio of Michaelis constants of the two substrates for the same enzyme (unitless) liquid viscosity (poise) surface tension of a liquid (dyne cm-l) electrostatic potential between the carrier and the substrate (volts) diffusion layer thickness (centimeters) ratio of the enzyme rate constant (V /K ) to the mass transport rate constant (h) (ugitTess) liquid flow rate (cm sec-1) xvi INTRODUCTION The concept of passing a reagent solution through an open tube to which an enzyme is attached has led to specific, sensitive, and fast determinations of clinically important compounds. The use of immobilized enzymes with continuous flow instruments is an important and growing area. This research is a part of the effort to improve such systems. The goals were to optimize the procedure for binding enzymes to nylon tubing and then to characterize the resulting reactors in terms of their physical and chemical properties. Glucose oxidase was chosen as the major enzyme of study, because it is relatively inexpensive, highly active, and bonds well to nylon. Lactate dehydrogenase, catalase, and other enzymes were used in immobilized form to a lesser extent. The enzyme activities were measured by incorporating the open-tubular reactors into a continuous flow or stopped-flow system. The continuous flow technique involves a flowing liquid stream in which the sample is injected, mixed with reagents, and detected. When gas bubbles are added to the liquid stream the technique is termed gas— segmented, continuous flow analysis (CF) [1-3]. Without {gas segments, the technique is called flow-injection analysis (FI) [4-6]. The major interest in this research has been with CF systems. The stopped-flow (SF) technique [7—9] involves the rapid mixing of reagents and sample, followed by the rapid cessation of the flow. The reaction mixture is most often monitored by spectrophotometry. These flow techniques are not discussed in detail; only important points pertinent to this research are mentioned. Informative discussions can be found in the references given above. After a brief introduction to immobilized enzymes (Chapter 1), the optimization of the nylon immobilization procedure is presented (Chapter 2). The kinetics proper- ties of enzyme reactors are discussed in Chapter 3. A novel stopped—flow instrument is described which should help in learning more about the inherent properties of immobilized enzymes. In Chapter 4, a glucose oxidase reactor, incorporated into a CF manifold, is character- ized. The influence of catalase impurities on the deter- mination of glucose was studied. In Chapter 5, the possibility of replacing nylon with non-porous glass is explored, and finally, suggestions for future projects are given (Chapter 6). This research project has provided several surprises and many interesting results. The work is mostly experi- mental and practical. Further work should provide a more mathematical description of the open-tubular immobilized enzyme reactors (OTIMERS). CHAPTER 1 Introduction to Immobilized Enzymes The interest and activity in the field of immobilized enzymes has been high for the past fifteen years. Hundreds of papers have appeared in the literature extolling the use of enzymes attached to an insoluble support. Many excellent reviews of the theory and practice of immobilized enzyme technology are available. At least nine books have been published in this area [10-18], with two of them appearing in 1980; several other books have chapters dealing with immobilized enzymes [19,20]. Furthermore, many reviews on the uses of immobilized enzymes have appeared in the scientific literature [21-25]. Since the research in this area has been well—docu— mented, only the fundamental concepts are presented here. Section A introduces bound enzymes and the factors which may alter the properties of the enzyme. Section B reviews the uses of the immobilized enzymes in analytical chemistry and introduces the concept of open-tubular reactors in flow systems. A. Immobilized Enzyme Characteristics An enzyme is a protein which can catalyze one or a small set of reactions. Thus, the enzyme is specific and very sensitive toward a few compounds and is useful as an analytical probe. Confining an enzyme to an insoluble 3 support provides the further advantages of longer life- time and easier manipulation. The enzyme reactor can be used again and again. Thus, immobilized enzymes lower the time and costs of employing biological catalysts. The enzyme can be confined or attached to a carrier in four basic ways. (1) Adsorption [13] occurs to some extent in all immobilization procedures. This method includes attachment by coulombic attraction and other short range forces. The procedure is the mildest and results in the highest percentage of active bound enzyme. However, the bond is weak, and the enzyme is gradually lost from the carrier with use. This method is not frequently used for analytical applications. (2) Encapsulation is really an indirect method of immobilization. A solution or slurry containing the enzyme is simply contained inside a semipermeable membrane. The technique is widely used to prepare enzyme electrodes [22] (See Section B). Substrate molecules must diffuse through the membrane in order for the reaction to occur. Thus, the membrane barrier limits the size of molecules that can react and limits the rate at which the enzyme can function. (3) Entrapment occurs when a gel is made to polymerize in the presence of an enzyme [13]. The enzyme molecule is trapped inside the empty spaces of the gel. Similarly, a :fiber stretched in a solution of enzyme will retain some (pf the enzyme in between the layers of fiber. Like encapsulated enzymes, enzymes trapped in a matrix are limited by substrate transport. (4) Covalent bonding of the enzyme to a carrier is the most widely used immobilization technique in analytical chemistry [12]. A chemical bond is formed between the carrier surface and the enzyme molecule. The reactor that results shows good temperature and time stability. The enzyme is not easily stripped from the carrier by a flowing liquid, so the technique is chosen in preparing open—tubular reactors. However, the percentage of active, bound enzyme is low. The reason for this is that no technique has yet been developed to single out a site on the enzyme for attachment. Thus, enzyme bound at the active site becomes inactive. Likewise, multiple bonds to the same enzyme molecule can change its tertiary structure, affecting its activity. Covalent bonding of enzymes is still rather empir- ical and not yet based on solid fundamental grounds. An enzyme covalently bound to a carrier experiences much of the carrier environment in addition to that of the bulk solution. The hydrophilicity or hydrophobicity, charge, and chemical reactivity of the support all influence the enzyme properties. The substrate affinity for the carrier and the enzyme is also affected. The enzyme is not always in intimate contact with the substrate as it is in solution; 'thus, substrate transport from the bulk solution to the <3arrier surface influences the reaction rate. Changes in pH optima and enzyme kinetics are discussed in more detail in Chapter 3. Thus, it is expected that the attached enzyme would display different properties from those of an enzyme in aqueous solution. B. The Use of Immobilized Enzymes in Analytical Chemistry Enzymes in aqueous solution have been used for thousands of years to produce beer, cheese, and other products. Similarly, industrial processes were among the first to use immobilized enzymes. Several reviews of the use of enzymes in industry have been written [26-29]. The lower costs that result are making bound enzymes popular. In analytical chemistry, enzymes have been transformed from aqueous reagents to integral parts of chemical instru- mentation. The immobilized enzyme is usually a part of the transducer or manifold of the instrument. The chemist does not need to prepare enzyme solutions each day, but instead must maintain the activity of the enzyme. Storage of the enzyme at reduced temperature, in the dark, and immersed in a buffer at neutral pH is usually all that is required. The job of the chemist is simplified by using immobilized enzymes. Immobilized enzyme instrumentation can be divided into three major areas: enzyme electrodes, open-tubular reactors, and miscellaneous methods. Enzyme electrodes are prepared by confining an enzyme solution or suspension to an area adjacent to the surface of an ion-selective electrode. A semipermeable membrane, stretched over the enzyme layer and the electrode, separates the enzyme from the reagent solution. Substrate molecules diffuse into the catalytic layer, the reaction occurs, and the product concentration is sensed by the electrode. Potentiometric and amperometric sensing are the usual means of detection. Enzyme electrodes do not include systems in which the electrochemical detector is separated from and subsequent to the immobilized enzyme reaction. The general reviews mentioned previously and several specific reviews [30-32] show the wide use of enzyme electrodes. Recent reports, listed in Table 1.1, are further evidence of the importance of these methods. Open-tubular reactors are even more widely used. A general schematic of a typical instrument is displayed in Figure 1.1. Open tubes allow low pressure pumping and preserve the integrity of the flow profile which are required in the analytical flow methods. Delivery of the reagents is usually performed by a peristaltic pump (CF and FI) or a pneumatically-driven syringe (SF). The mixed reagents are pushed into the reactor, where they are allowed to react for a short period of time. The reaction time is a compromise between sample throughput and acceptable instrumental response. The reaction products are usually measured by absorption spectroscopy, but can also be monitored electrochemically, thermometrically, or lay luminescence spectroscopy. mm mm mm vm mm .mmm cappoEOMoQE< oflgpmsofipcopoa ofihpoeouode< cappmeonmae¢ QHHpoEOAmQE< AOpoopoQ canspaaam mowom ocfiswlq Hosanna me< Hosanna mpwfidc< moUOMpomHm capafimo mochpooao opfinmmhw Op @202 usofia>ov How encamcdom How moHEdazgoahaom :oflpsaom mzoosv< xflgpmz msmNcm mszucm mo moHQmem mmmadpao mmmoflxOMoQ mmaofixo mmOUSHw I®E>Ncmflge .m U omapflxo How ocfls< .v ommcowonozcoo HOSOOH¢ .m mmacwxoxmn omwoflxo mmoosaw IoE>Nc®Hm .N omwsowouomnoo pcooom Hosoofi< .H vamEmNcm .H.H mange Rquent Dehvery > Mixer Systenw fr,» Enzyme Reactor [ DetectorH Woste Figure 1.1. Schematic of an instrument employing an open-tubular reactor. 10 In most cases, nylon tubing has been used as the carrier. It is inert, sturdy, easily molded into a helix, and readily activated for subsequent enzyme immobilization. The only disadvantages are that it has a small surface area (few sites for enzyme attachment), and it tends to adsorb organic compounds, making sample carryover a problem. Despite these problems, no other carrier has replaced nylon for analytical use. Table 1.2 lists some of the more recent uses of nylon open-tubular reactors. Several other analytical reactors have been reported in the literature. A few of these are listed in Table 1.3. The most common type is a packed bed reactor. The enzyme is covalently bound to porous glass beads, and the beads are packed into a tube. Thermometric detection is common with these systems. More unusual instrumentation includes a nylon tubular reactor connected to a pipette tip and a stirring bar coated with an enzyme. Immobilized enzymes have been termed a "solution in search of a problem”. Indeed, the potential usefulness of immobilized enzymes has not been reached. Only a few enzymes of high inherent activity have been used success- fully. In order for more and better enzyme reactors to be developed, the following needs must be met. The carrier environment needs to be modified so that it is there conducive to enzyme attachment. Also, the chemist zaeeds to be able to choose the site of attachment on the 11 Ne mocmnpomnw >3 av cosmopomnw >D ow moonpomao wfisoEE¢ mm mocmnpomnw >3 mm oGOAHQoao cowmxo mocogommm mammwmwm mAOHOHom swasnseucoao H0pmozao moflgmozawwhe ocflcflpmopo eaom can: Hosanpm mpwflaa< coapmfisxfia mmeoe eoflpafisxflm mmeoe ceapafisxfie mmeoe mHm%HOHU>: seawa>apo< Ho mmHQEMxm pcmomm omncomoucmnoo Honmo>ao .m omwcowOAozsoc Honookao .v ommcficflpmmpo .m .eszme masseuse ommOHps I®E>Ncoflm .N ommoflxo HOQOUH< .H Amvosmmcm .N.H manwe 12 man magmafia a Op conomppa om OHHpQEOHOQQoupoon omwoua many 20H>z .o av ofihposoaonQOMHOon omwcom0pc>noo Honooa< Han mafiaaflpm .m wv QHQHoEoHpcmpom omauozpop mamupflz .H mm oEmm .w omacmw0hozsmo mpwpomq be oflppoEOAoQE¢ omwcowoupznmp Hosoofi< .H mm osmm .m ow ofippoEonoo ommohb .H mm oEmm .m owmofixo Houmpmoaozo mmmpflxo mmoosfio :ESHoo m as coxoma mvlmv capmeOEnoze enema: mmwam machom .H mosopmmom acuoopom oewncm yoflnpwo myopoemm nmspo new nomucmxomm .m.H manna 13 enzyme. Furthermore, if diffusional limitations and electrostatic effects could be eliminated, the use of immobilized enzymes would be simpler since the large body of soluble enzyme data would be directly applicable. CHAPTER 2 Chemical Bonding of Glucose Oxidase to Nylon Tubing The goal of immobilization is to produce an enzyme reactor with the highest possible activity. This means that many sites on the carrier must be available for attachment, and that the enzyme must be bonded in its most active form. With nylon, four chemical steps are employed to produce the desired results, as shown in Figure 2.1. First, the inert, amide polymer is treated so that a reactive group is formed. Three different activators have been used, and these are discussed in Section A. Next, a bifunctional compound is added. This "spacer” molecule increases the distance between the nylon surface and the enzyme. Therefore, the enzyme experiences less of the carrier environment and more of the bulk solution properties. Diffusional and steric limitations are reduced to some extent by adding the spacer. In Section B, the use of diamines as spacers is described. Third, enzyme attachment is accomplished, usually via free amino groups (lysine residues or N-terminus) or free carboxyl groups (aspartic acid, glutamic acid, or C-terminus). In most cases, bonding through amino groups yields higher activities than bonding to other sites. This may be due to participation of carboxyl groups in the active site. Thus, glutaraldehyde is often used to link the amine group 14 15 .msonom :ofipaNHHHQOEEH one : Amasmmoq' \\ > Figure 3. 45 J“. ‘— 0.0.1 1 __ 2 5 __ 10 J. I I I I I L, I I I I I I I I I O 1 2 3 4 5 6 7 [Substrate]/Km 2. Influence of slow mass transport on enzyme kinetics. 46 lowering the constant of the other substrate [72,73]. This see-saw effect has been confirmed in several immobil- ized enzyme studies [74—76]. Under other conditions, the apparent Km values can both be higher or lower (see Engasser's paper for details [72]). The type of mass transport in open-tubular reactors depends on the system being considered. These reactors are used almost entirely in three flow system types: stopped-flow, flow-injection, and continuous flow. The mass transport processes and kinetics of each of these are discussed in turn. Only external mass transport needs to be examined, since the enzyme is bound essentially to a non-porous wall in a very thin layer. The reaction in the stopped-flow system occurs under static conditions. In the simplest mass transfer model, only molecular diffusion is considered. Once the enzyme reaction begins, concentration gradients in both substrate and product are established as shown in Figure 3.3. The mass transport rate constant under static conditions may be given by the following equation [70]. h = NuD/(2ra) (3.9) Here Nu is the Nusselt number, D is the molecular diffusion coefficient for the substrate, and r is the reactor radius. Under conditions of high enzyme binding and activity, molecular diffusion may totally control the overall reaction 47 Relative Concentration Distance along Reactor Diameter Figure 3.3. Concentration gradients in an open—tubular reactor. 48 rate. A mathematical description has been derived and is given below. Details of the model are described in Appendix A. 3 [p1b = [s1b<1 — 4 Eexp(-Dj:t/r2)/j:) (3.10) s—l where [P]b is the product concentration in the bulk solution at time t, r is the reactor radius, jS is a Bessel function, and t is the time inseconds. When a flowing liquid passes through an OTIMER, convection becomes the primary means of transport. Enzyme reactors used in flow-injection analysis [4-6] are examples of this system. Laidler and coworkers have been the major contributors to the theory of mass transport in such systems [77-80]. Equation 3.11 is a mathematical descrip— tion of the transport rate constant [77]. 3 h = 1.29 0-1(D20/rL)1/ (3.11) where v is the linear (cm/s) flow rate, and a is the diffusion layer thickness. The inherent Michaelis constant is transformed into the following [77]. _ . 2 1/3 Km(apparent)—Km(1nherent)+0.39ake[E]O(rL/D v) (3.12) Since mass transport is faster in flowing systems than under static conditions, the enzyme reaction rate should have a greater influence on the overall kinetics in flowing systems. 49 Similar interplay between the enzyme and mass trans- port is expected in air-segmented continuous flow systems [21-23], but the mass transport rate has not been described mathematically. A thin film of liquid which moves only slowly along the reactor wall and secondary, bolus, flow which is established in each liquid segment complicate the fluid dynamics. Therefore, primary flow, secondary flow, and mass transport in the thin film all contribute to the mass transfer of the system. Snyder has described the effect of incomplete mixing in the liquid segments on the sample dispersion [82,83]. One should be able to relate the extent of dispersion to the rate of radial mass trans- port, since it is known that mixing along the direction of flow is fast while mixing across adjacent streamlines is the limiting factor. Figure 3.4 shows the pattern of bolus flow. One can assume that the rate of mass transport is inversely proportional to the sample dispersion due to incomplete mixing. From Snyder's theory, the following equation may be a good estimate of mixing in OTIMERs in segmented—flow systems. 2/3 1/2 h = (1-25Y VD )/(v7/6n2/3L1/2d3 ) (3.13) Here Y is the surface tension of the liquid, V is the volume of a single liquid segment, n is the solution vis— cosity, v is the linear flow rate, L is the reactor 50 1/ ‘\ 1’ ‘\ 1’ \\ Air Liquid Air Segment Segment Segment 1’ ‘\ 1’ ‘\ ,/ l \ Liquid I Figure 3.4. Bolus flow pattern. 51 length, and d is the reactor diameter. Further theoretical and experimental work on this system is needed. B. Experimental Considerations The kinetics of open—tubular immobilized enzyme reactors are difficult to monitor. In all systems, the reagents are pushed into the reactor, allowed to react for a time, and then the substrates or products are monitored as they exit the tube. Thus, the reaction rate obtained is the result of a fixed-time measurement, with the initial time (t1) being zero and the final time (t2) being the moment that the substrate solution exits the reactor. While t2 can be varied, tl is usually restricted to a value of zero. Enzyme lag phases, therefore, are included in the rate measurements, which introduces error. Continuous flow methods are used widely in conjunc— tion with immobilized enzymes. Yet few kinetics experi- ments have been performed because data at early reaction times are impossible to obtain and entire progress curves would require many experiments under varying conditions. For example, a change in pump speed to change the reaction time might also alter the mass transport rate. Bi-direc— tional flow, continuous flow methods [84] and other designs are possible ways of reducing the number of required experiments, but the instrumentation and procedures are complicated. A stopped—flow method using OTIMERs has been described [85], but again many experiments are required in 52 order to draw a complete progress curve of the enzyme reaction. Ideally, one would like to monitor the reaction directly and continuously. We have developed an instrument that is ideal in its concept, and only a little less so in its function. A tubular reactor, containing an immobi- lized enzyme, is fitted inside the observation cell of a stopped-flow instrument. With the reagents mixed and pushed into the observation cell, the reaction begins. The light path of the photometer passes through the reactor and reacting solution; thus, the formation of product or de- pletion of substrate can be monitored directly and contin- uously by absorption spectroscopy. The advantages of the instrument are numerous. Early reaction times are obtainable since the injection of the mixed reagents is performed quickly by the stopped-flow. The time window can be varied so as to exclude lag periods and other error-prone regions. Also, initial reaction rates can be calculated more accurately by the linear regression slope method rather than by the fixed-time method. Only diffusion and/or electrostatic effects con- tribute to mass transport in the reactor so that funda- mental studies of the kinetics can be done more easily. The only disadvantage is that one of the substrates or products must have a unique absorbance maximum in the range of the monochromator. If not, a follow-up reaction can be 53 used, but only if its reaction rate is much faster than that of the analytical reaction rate. The requirements for the instrument described above are somewhat different from those of a conventional stopped- flow instrument. First, fast mixing is not essential since the reaction does not begin until the reagents enter the observation cell. Even manual mixing is possible. However, the injection time, the time taken to push the solution through the cell, must be small so that the reaction extent prior to the first data point is negligible. These re- quirements can be provided by simply employing a T-mixer and a strong syringe push. Second, downstream stopping is not required since typical enzyme reactions are moni— tored on the order of seconds rather than milliseconds. Third, mixing between reagents and the reaction volume inside the observation cell should be minimized. Unlike standard stopped—flow experiments, products are only formed in the observation cell/reactor; diffusion of product into the reagent solution at the entrance and exit of the cell will cause errors in the rate measurement. Any siphoning or solution leakage will lead to drastic errors. The design of an instrument which minimizes this mixing is a difficult problem. A prototype instrument which elimi- nates the reaction volume/reagent solution interface was built, and preliminary test results are given in Section E. A modified GCA McPherson stopped-flow module was used 54 in most of the kinetics studies. A 1.75 cm immobilized enzyme reactor, 1 mm i.d., was fit tightly inside the observation cell of the module. Figure 3.5 shows the location of the reactor in the cell housing. A T-mixer was employed, along with upstream stopping. The remainder of the instrument was constructed from commercial com- ponents as shown in Figure 3.6. Data acquisition was performed by a microcomputer [86], linked to a PDP-8/e minicomputer (Digital Equipment Corporation). Further details of the instrument and its use are given in Section C. One other important consideration in the use of this method is that the absorbance measurement is atypical. Because a product (and substrate) concentration gradient exists radially across the tube, the relationship between absorbance and concentration does not depend only on the molar absorptivity, c, and the reactor length, but also upon reactor radius and the ”steepness” of the concentra- tion gradient. A mathematical model of the time-indepen- dent absorbance versus reactor design relationship is described in Appendix B. The time dependent case is com- plicated by the rate of product formation and mass transport and has not been considered. The final result of the mathematical treatment is given in equation 3.14. Absorbance = eL([P]b+a([P]w—[P]b)/2r) (3.14) Here [P]b is the concentration of the product in the bulk TEL— Figure 3.5. l l Solufion Flow Cross-sectional View of the GCA McPherson stopped—flow mixer/observation cell unit. The outer aluminum housing (a), quartz win- dows (b), and the main KEL-F body (0) are press fit with three bolts. Mixing occurs at (e), where the two reagent flow streams meet at 90° to each other-- one stream is in the plane of the figure and the other is perpen- dicular to it. The immobilized enzyme reactor is placed inside the observation cell (d). The dashed arrow represents the light path inside the cell. 56 GCA McPherson GCA McPherson :UOb-se-rotn'onfi O , I GCA McPherson Model EU—701-5o -C°,,/;"'c,°ra FP‘M Model EU-701-30 Light Source Ilter Photomultiplier Stopped-flow Module Keithley Microcom uter Model 427 [3 Current to Voltage Amplifier PDP B/e Strip Chart Minicomputer Recorder Figure 3.6. Stopped-flow calorimeter. 57 solution, and [P]w is the concentration of product at the reactor wall. The absorbance can vary with reactor diameter. However, during a single experiment, the reactor diameter is a constant and the absorbance depends on the diffusion layer thickness, 0. Changes in this factor during an experiment are expected to occur until steady-state between the enzyme reaction rate and mass transport is established. If steady—state conditions are established quickly, then most of the data will be free of mass transport influence on the absorbance measurement. More definitive experiments are needed to establish when steady-state is attained and to verify equation 3.12. C. The Determination of Michaelis Constants 1. Introduction We report here a study of the kinetics of L(+)-lactate dehydrogenase, LDH (E.C. 1.1.1.27). This enzyme converts L-lactic acid and B-nicotinamide adenine dinucleotide (B-NAD) to pyruvic acid and B-NADH, the reduced form of B-NAD. The reaction goes by a compulsory order bi-bi mechanism [87]: + E1 + B-NAD + 2 E + E2 + Lactate + E3 E3 + E1 + Pyruvate + B-NADH + H+ Here El represents the free enzyme, and E2 and E3 represent 58 the bimolecular and termolecular enzyme—substrate complexes respectively. The initial reaction rate for such a mecha— nism is given in equation 3.15. v=k [E] /(1+KN/[NAD]+KL/[Lac]+K KL/([NAD][Lac])) <3 15) e o m m i m ° The term, Ki’ represents the degree of interaction between the two substrates on the enzyme; the value is zero for a ping-pong type mechanism. Lactate dehydrogenase is a tetrameric molecule, com— posed of heart (H) or skeletal muscle (M) type monomers [87]. MH M H M H, 4’ 3’ 2 2’ 3 and their relative proportion varies among different Thus, five different isozymes exist—— H and M4, species and tissues. The preparation used in this study was obtained from beef muscle, and so it contained mostly the M4 and M3H tetramers. The molecular weight of the enzyme is about 140,000. Each subunit contains a B-NAD binding site, but these sites are independent of one another. High concentrations of pyruvate and lactate inhibit the enzyme. The KI value, the inhibition constant, for lactate varies from 26 mM for the H isozyme to 130 mM for the 4 M4 tetramer. In this section, the three kinetics constants and Vm’ which equals ke[E]O, are reported for aqueous and nylon- immobilized LDH, and the values are compared. Partial diffusion control of the reaction rates is evident. 59 2. Reagent preparation Tris Buffer. The following quantities of materials were dissolved in distilled, deionized water (DDW): 10.0 g trishydroxyaminomethane, 13.0 g hydrazine sulfate, and 2.0 g ethylenediaminetetracetic acid. The solution was diluted to one liter with DDW and then the pH of the solution was adjusted to 9.6 with NaOH. The buffer was stable for one week. B-NAD Solutions. The desired amount of B-NAD (Grade III, molecular weight=706, Sigma Chemical Company) was dissolved in buffer and diluted to 50.0 mL with buffer to give a stock solution. The working solutions were prepared by appropriate dilution of this stock with buffer. The stock solution and working standards were prepared daily. Lactate Solutions. Purified L—lactic acid (0.960 g, Grade L-X, Sigma) and 0.125 mL of concentrated sulfuric acid were added to 500 mL of DDW. The resulting solution was diluted to 1.00 L with DDW. This stock was diluted appropriately with DDW to give the desired standard solu- tions. The stock solution was stable for two weeks, and the standards were prepared daily. Soluble Enzyme Reagent. Eighty to one hundred micro- liters of a suspension of beef muscle LDH (Type X, 8800 Units/mL suspension, Sigma) were dissolved in Tris buffer, and then diluted to 0.200 L with buffer. This reagent was then substituted for the Tris buffer in the B—NAD solution 60 preparations described above. The B-NAD/enzyme solutions were used within two hours. 3. Immobilization of lactate dehydrogenase Lactate dehydrogenase was covalently bonded to straight, 20 cm lengths of nylon tubing (Portex Ltd., 1 mm i.d.). The immobilization procedure is given in Chapter 2, Table 2.3. The lactate dehydrogenase suspension obtained from Sigma was used directly in the immobilization. 4. The instrumentation The stopped-flow system was assembled from commercially available components. The stopped—flow module (GCA Mc- Pherson) was modified for upstream stopping by removal of the stop syringe. The two push syringes had volumes of 370 uL. A filter photometer was employed for absorbance measurements. The radiation source was a deuterium lamp housed in a commercial light source module (GCA McPherson). A 340 nm interference filter (bandwidth = 12 nm) was mounted between the stopped-flow unit and a photomultiplier module (GCA McPherson). The photocurrents from the photomul- tiplier were converted to voltages (Kiethley model 427 current amplifier), digitized, and the digital voltages stored in RAM with a microcomputer data acquisition system [86]. Data were acquired at 1.0 Hz. At the end of each trial, the data were shipped serially to a PDP-8/e mini- computer (Digital Equipment Corporation) for processing and analysis. The initial reaction rates were computed as 61 the linear regression slopes over the time period from 10 to 20 seconds after the final reagent push. The instrument was operated in the following manner. The two syringes were filled with DDW, the water was pushed through the observation cell, and the flow was stopped. This procedure was repeated six times to wash out the reac- tor completely. Next, the syringes were filled with the reagents, the reagents were pushed through the mixer and cell, and the flow was stopped. This second procedure was repeated five times in order to ensure quantitative 1+1 mixing of the B-NAD and lactate solutions. After the final reagent push, the microcomputer data—taking routine was started. These steps were repeated for each experiment. The blank consisted of the B—NAD working solution mixed with distilled deionized water. For the soluble LDH studies, the instrument was used as described above. After the soluble enzyme studies were completed, the observation cell of the stopped-flow module was disassembled, and a 1.75 cm section of the immobilized LDH tube was fitted tightly inside the cell. The cell was reassembled, and the study was repeated. Between trials, the IME tube was filled with pH 6.85 phosphate buffer, .0.05 M. A room temperature water bath kept the reagents at a constant temperature, between 18 and 20 °C, during all experiments. (.3‘ H: (‘11 62 5. Results and discussion The precision of the rate measurement was determined by repetitively montoring the immobilized LDH catalyzed reaction of 0.50 mM lactate and 1.35 mM B-NAD. The rela- tive standard deviation was found to be 4%. The enzyme activity declined rapidly during the first two days of use and, thereafter, only slowly. Even after seven days at room temperature and about 175 experiments, 60% of the original enzyme activity remained. Typical progress curves obtained with our instrument are shown in Figure 3.7. The immobilized LDH displays a sigmoid-shaped curve, that is similar to the one produced by the aqueous LDH. No odd reaction kinetics are apparent. Also, a short lag phase is noticeable. These facts are quickly and easily shown, in contrast to other indirect kinetics methods. The Km values were determined by the method of Florini and Vestling [88]. The mean values of the initial reaction rates for each pair of substrate concentrations were fit to equation 3.15 by a computer regression program. The experimental and regression results are shown as Lineweaver-Burk plots in Figures 3.8 and 3.9. Because the lines in the double reciprocal plots intersect at a common point below the x-axis, the formation of a ternary complex is confirmed. A small concave curvature exists in the experimental 63 Absorbance Time (seconds) Figure 3.7. Typical progress curves obtained with the instrument. The dotted lines represent the initial reaction rates calculated from these curves. The reactions represented were of 2.00 mM lactate with 0.425 mM (A), 0.850 mM (B), and 2.12 mM (C) B-NAD. 64 6.0-m- 1/R0te (x102 sec) '-2.0 T ! ' g ; t . I ' i -45.0 —30.0 -15.0 030 15.0 30.0 1/{Lc0‘tate} (x302 M-i) Figure 3.8. Lineweaver—Burk plot of 1/Rate versus 1/[Lactate] at four different B-NAD concentrations: 0.26 mM (A), 0.43 mM (B), 0.85 mM (C), and 2.12 mM (D). 65 6.0'—r A ’5‘ 4.0—— o 0 B m . j_ C a: 0 ° . ° 3'. 20—— v o e o __ . ' o 0 I: '\~ 0 O..- l I . I . I I I -2‘0 ' I ‘ I l 1 -10.0 '-5.0 0.0 5.0 ‘10.0 1/[NA03 (x103 20“) Figure 3.9. Lineweaver-Burk plot of 1/Rate versus 1/[B—NAD] at three different lactate concentrations: 0.80 mM (A), 2.00 mM (B), and 4.00 mM (C). 66 values. Such curvature is not apparent in the aqueous enzyme data. Sundaram and Igloi also reported such curved plots in a multi-enzyme system bound inside nylon tubing [56]. Diffusional resistance or the presence of several iso- zymes may be the cause of this effect, but no realistic mathematical model has been found. The y-intercepts of the Lineweaver-Burk plots represent the reaction rate at infinite lactate or B-NAD concentrations. A second Lineweaver-Burk plot of these rates yields the desired Km and Vm values, as shown in Figure 3.10. The results of five independent trials were weighted and averaged to give final values for the kinetics constants. The values found for soluble and immobilized LDH by this method are given in Table 3.2. The Km of B-NAD increases and the Km of lactate decreases upon immobilization. Intuitively, Km values of bound enzymes should not be lower than those of soluble enzymes, unless a conformational change occurs upon immo- bilization. However, Engasser, £3.31. [72] recently ex— plained that if the Km values and diffusion rate constants (hs) of the enzyme substrates differ greatly, then the tendency is to increase the Km of the higher affinity substrate and decrease the Km of the lower affinity sub— strate. This is due to diffusional limitations on the reaction rate. Engasser, §£.al. restricted the model to ping-pong '1. o 67 I I l I. L / a. I! (D 1’ I N , 1’ o 2.0—— I, 1’ I / I ’l X . l’ , V l p ...-I T. a, ,’ Q l- [I Q I I o I I a; " ” I *0 100 —_ I /d C ’ I .- a , ' ’sz >‘ D, l/ 4' ”l I” i / I’ i ’ I ..’ ’ I I l l l l 0.0 I I i j -2 0 2 4 6 3 10 Figure 3.10. Plot of y-intercepts from Figures 3.8 and 3.9 versus 1/[Substrate] for different B-NAD and lactate concentrations. The y-intercept equals 1/Vm, and the x-intercepts yield the two Km values. 68 Table 3.2. Kinetics Values for Aqueous and Immobilized Lactate Dehydrogenase Agueous Immobilized vm (10'6 Ms‘l) 2.0 i 0.1 2.1 i 0.1 KS (10‘4 M) 2 5 1 0.3 4.3 i 0 5 K; (10'4 M) 19 i 4 9 i 1 K (10'4 M) 5 8 i 1.6 1.9 1 0.5 69 type reactions, but we believe that the same effect should occur with compulsory order mechanisms. In the case of LDH, the important dimensionless factor, 5 =(hNK:)/(hLKg), equals approximately 0.07. Since this is very much different from 1, it may be expected that the see-saw effect on the Km values will occur. Using Figures 1 a,b, e, and f of reference 72 and interpolating to approximate the conditions of this experiment, weestimated that the enzyme reaction rate was about five times faster than the diffusion rate. A second explanation of the KIn changes is based an electrostatic effects. The immobilization procedure im- parts a positive charge to the nylon surface. Since lactate has a negative charge and B-NAD has a positive charge, the apparent Km values could decrease and increase respectively, because the surface concentrations of the two substrates differ from those in the bulk solution. The kinetics values found here are relevant to other methods in which a reagent solution is allowed to react inside an OTIMER under static conditions. Sundaram's impette method [50] and stopped- flow, flow—injection methods employing OTIMERs should benefit directly. The method is very useful in determining the apparent Michaelis constant values for enzymes immobil- ized inside open tubes. A variety of enzymes can be tested in this manner. 70 D. Analytical Determinations 1. Introduction Direct reaction rate methods for two clinically important substrates, glucose and lactate, are reported here. The glucose determination was performed under fixed- time conditions, while the lactate results were obtained from rates calculated by linear regression analysis of the absorbance—time data. The instrument allows calculation of the initial reaction rate by the fixed—time, variable- time, or the slope approach [89]. The Trinder method [90,91] for the determination of glucose with glucose oxidase was adopted here. The two reactions involved are given below. B-D—glucose + 02 + H20 + 0—gluconolactone + H202 2 H202 + DCPS + AAP + Dye + 3 H20 + st04 Glucose oxidase (GO) catalyzes the first reaction, while peroxidase (P0) does the same for the second reaction. DCPS is 2,4-dichlorophenolsulfonate and AAP is 4-amino— antipyrine. The absorbance of the dye product at 505 nm was monitored. Lactate was determined by measuring the absorbance of B-NADH produced in the following reaction. Lactate + B-NAD + Pyruvate + B—NADH Lactate dehydrogenase (LDH) catalyzes this reaction. 71 The spectroscopic measurement was done at 340 nm. A basic solution and hydrazine were used to drive the reaction toward completion. The results of several control serum assays for glucose and lactate are also reported. 2. Reagent preparation Glucose Standards. The glucose stock solution was prepared by dissolving 2.000 g of anhydrous a-D-glucose and 0.50 g of benzoic acid in about 0.75 L of distilled deionized water (DDW). The solution was then diluted to exactly 1.00 L with DDW. The standard solutions were pre- pared by appropriate dilution of the stock with DDW. The stock was stable for a month at 4 °C. Trinder/Peroxidase Reagent. The method of Barham and Trinder [91] was used to prepare a solution of 2,4-dichloro- phenolsulfonate. Two mL of this solution (about 0.2 mmoles), 110 mg (1,670 Units) of peroxidase (Type II, Sigma), and ILO mg (about 0.05 mmoles) of 4-aminoantipyrine were dis- ssolved in 75 mL of pH 6.4 phosphate buffer, 0.1 M. The ssolution was then diluted to exactly 0.100 L with the pH €5.4 buffer. This reagent was prepared daily. Lactate Standards. Purified L-lactic acid (0.960 g, Cirade L—X, Sigma) and 0.125 mL of concentrated sulfuric aJZid were added to 500 mL of DDW. The resulting solution ‘VEts diluted to exactly 1.00 L with DDW. This stock was (ii.luted appropriately with DDW to give the desired standard SOlutions. The stock was stable for two weeks; standards 72 were prepared daily. B-NAD Reagent. The following quantities of materials were dissolved in DDW: 10.0 g of trishydroxyaminomethane, 13.0 g hydrazine sulfate, and 2.0 g ethylenediaminetetra- acetic acid. The pH of the solution was adjusted to 9.6 with NaOH after dilution to exactly 1.0 L with DDW. The buffer was stable for one week. In 40 mL of this buffer, 0.300 g of B-NAD (Grade III, Sigma) were dissolved. The resulting solution was diluted to exactly 50.0 mL to give a 8.5 mM B-NAD solution. This reagent was prepared daily. Serum Samples. The control sera (Monitrol and Patho- trol, Dade Division, American Hospital Supply) were re— constituted according to the manufacturer's instructions. For the glucose determination, 1.00 mL of the serum, 1.50 mL of a 20 g/L solution of Ba(OH)2, and 1.40 mL of a 20 g/L solution of ZnSO in that order, were added to a 4, centrifuge tube. The solution was thoroughly mixed and then centrifuged for three minutes. The supernatant fluid was used directly for the determination. No protein pre- cipitation was involved in sample preparation for the lactate determination. One millilter of the control serum was diluted to exactly 0.100 L with DDW, and the resulting solution was used in the analysis. 3. Enzyme immobilization The enzymes, glucose oxidase (Type II, Sigma) and beef muscle lactate dehydrogenase (Type X, Sigma), were 73 immobilized on Type '6' nylon tubing (0.1 cm i.d., Portex Ltd.) by a procedure described in Chapter 2, Table 2.3. A 10 mg/mL solution of glucose oxidase was prepared for use in step 4 of the procedure, while the lactate dehy- drogenase suspension was used directly as sold by Sigma. The completed reactor was filled with buffer (pH 6.4, phosphate) and stored at 4 °C when not in use. Short, 1.75 cm, segments of the tubing were cut for use in the stopped—flow observation cell when required. 4. The instrumentation The instrument has been described in detail in Section C. For the fixed-time measurements, a monostable timer circuit was built to cause a sample-and—hold device to hold light intensity values obtained thirty seconds (11) and sixty seconds (12) after the reaction began. The change in ab- sorbance was taken to be log(11/I2). This assumes no drift in the 100% T setting over the 30 second time interval. The light source was very stable, so this was probably a good (< 2% error) assumption. The time range was chosen as an Optimum between sample throughput and errors due to mixing effects and enzyme lag periods. The lactate data were taken and processed by computer as described in Section C. 5. The results of the glucose determination Glucose standards and the Trinder/peroxidase reagent were mixed rapidly and pushed into the observation cell/ immobilized enzyme reactor by the stopped-flow instrument. 74 The decrease in light intensity reaching the detector was monitored and the change in absorbance was calculated as described above. A blank was tested between each sam- ple, because the Trinder reagents and product adsorbed strongly to the nylon, producing some carryover. Thus, the sample throughput was limited to a maximum of thirty samples per hour. The results of five separate experiments over a four day period using the same glucose oxidase reactor are shown in Figure 3.11. A very linear calibration plot with a slope of 8.3 i 0.2 x10"2 and a y—intercept of -3.4 i 0.1 x10_3 was obtained. The error in the absorbance measurements was mostly due to a decline in the enzyme activity with time. The reactor, stored at room temperatrue for five days, dropped in activity by only 8%. The linear range of the method, 1-10 mM, covered the normal levels found in human blood serum (3.8—5.9 mM [92]). Three control serum samples were assayed with the instrument. During the protein precipitation treatment, the samples were diluted 1:3.9; the resulting concentrations were still within the linear range. Table 3.3 compares the concentrations found in our laboratory with those given on the data sheets accompanying the samples. The accuracy of the method was good, with the greatest deviations occurring with samples of low glucose content. This 75 0.6 —- 0.5-1— ‘1’ _ g) 0.4 — C a .o L. O 03* U) .a <3: <1 0.2 —r 0.1 ~— 00. I l I l I [Glucose] UnM) Figure 3.11. Calibration curve for glucose. 76 Table 3.3. Results of Serum Assays for Glucose Serum No. Detns. Expected(mM) Found(mM) Monitrol I 5 5.67 4 i 1 Pathotrol 6 12.8 13.3 i 0.3 Monitrol II 9 13.4 13.2 i 0.5 77 deviation was probably due to the presence of ascorbic acid, gentisic acid, and other reducing agents, which can interfere with peroxidase-catalyzed reactions [93,94]. More details of this problem are given in Chapter 4, Section B. 6. The results of the lactate determination Initial reaction rates for the reaction of lactate and B-NAD were determined between 10 and 20 seconds after the final reagent push, with a computer linear regression program. Sets of standards were tested with a blank run at the end of each set. Sample carryover was small with this system, apparently because little adsorption of reagents or products on the nylon occurred. Thus, the processing of 90 samples per hour was possible. Reaction rates obtained from five experiments over a four day period are plotted versus lactate concentrations in Figure 3.12. The same lactate dehydrogenase reactor was used in all of the experiments. The linearity is excellent, but a small positive intercept is apparent. This inter- cept may have been caused by a slow non—enzymatic reaction between the reagents. The blank, which consisted of DDW and the B-NAD reagent, would not correct for this. The slope of the calibration curve was 4.3 i 0.3 x10-3, and the y-intercept was 1.0 i 0.1 xio'lo. The errors in the reaction rate data were mostly due to fluctuations in the reactor activity. Sixty percent /'\ ,— I (n .2 |\ I O X v 1 (1) +1 0 01 1 C O .4: 0 O 0) 0: Figure 78 J1 1 l 1 l 1 l O J T I l I I r 0 2O 4O 60 [Loctote] (MND 3.12. Calibration curve for lactate. 79 of the original enzyme activity remained after seven days of heavy use at room temperature. The linear range of the method, 5—50 uM, is limited at the high end by the following requirement: [Lactate] < 0.05 Kg. The K; of immobilized LDH was determined to be 9 mM in Section C of this chapter. Thus, the calibration curve should bend off near 45 pM and it does. Serum assays were performed on 1:100 dilutions of the reconstituted controls with DDW in order to adjust the sample concentrations to the linear range of the method. This was quite advantageous, since the sample preparation was simplified and interfering substances should have a much smaller effect. Table 3.4 shows that the lac- tate concentrations in all three samples were accurately determined. However, the average values were slightly higher than those expected. The reason for this was that a small amount of aqueous lactate dehydrogenase was present in the sera, and the soluble enzyme reaction also occurred inside the reactor. Nevertheless, the method is very sensitive and the accuracy is within the limits of error. Since the sensitivity and accuracy are good, but the sample throughput is low, this stopped-flow instrument should be suitable for method development and substrate determinations in small clinical laboratories. Many other enzymes can be immobilized inside nylon tubing and used in the instrument to determine a wide variety of compounds. 80 Table 3.4. Results of Serum Assays for Lactate Serum No. Detns. Expected(mM) Found(mM) Monitrol II 6 1.8 2.1 i 0.2 Monitrol I 5 2.2 2.2 i 0.4 H- Monitrol I 6 3.1 3.2 0.6 81 The calculation of reaction rates is direct, and lag periods can be avoided. With greater automation and the use of more inert supports, such as glass or Teflon, higher through- put is likely, and the method should become more advanta- geous. The instrument is versatile in that it can function as a device which determines kinetics values for assay development and also performs the actual determination. E. A New Cell / Reactor Design In order to eliminate any error due to mixing between reacting and unreacted solutions in the stopped-flow instrument, the cell/reactor was redesigned. Figure 3.13 is a diagram of the new instrument, built around two Altex slider valves. The cell was moved between positions 1 and 2 with a pneumatic valve and applied gas pressure. Typically, the cell was filled while in position 1 and then was moved quickly into position 2 for spectrophoto- metric monitoring. In this way, the interface between the reaction volume and the mixed reagents was eliminated. Errors due to undesirable mixing were prevented. The slider, shown in Figure 3.13, was constructed of nylon to allow direct immobilization of enzymes to the walls of the observation cell. In the previous design, slight movements of the nylon tube at early reaction times often caused large fluctuations in the instrumental out- put. This problem was eliminated. Another advantage of 82 P1 P2 Fhfid Li ht Figure 3.13. Design of the new cell/reactor. Two pneumatically actuated pistons (A) were used to push the slider (B) back and forth between the fill position (P1) and the viewing position (P2). Quartz windows were used (C) to contain the solution in the viewing position. 83 using nylon sliders was that various reactor diameters could easily be implemented by simply changing sliders. One obvious disadvantage of this new method was the increase in cost of materials and time. While nylon tubing is fairly cheap, it is expensive to machine a new nylon block for each experiment. Also, more diffi- culty and time is involved in immobilizing enzymes on these blocks with high precision. The greatest disadvan— tage was found to be that the movement of the slider some- times caused bubbles to form in the observation cell, because the solution was lost from the ends of the cell during changes in position. The development of bubbles was infrequent and was detected by a large drop in instrumental output. Several stopped-flow pushes usually restored the original output. The results of a series of experiments with distilled water are displayed in Figure 3.14, and they show fairly good reproducibility in the cell positioning. The use of the instrument was demonstrated by determining glucose with immobilized glucose oxidase. The enzyme was bonded on two sliders-— one with a 1 mm i.d. cell and one with a 2 mm i.d. cell. The immobiliza— tion procedure was identical to that described above (Chapter 2). A stopped-flow unit employing a T-mixer and upstream stopping was used with the cell assembly. Each syringe held 0.16 mL. Four reagent pushes and 84 o o o o o o 4" oo oo 3 —- o o 0 o <1) cm 6 ‘— C +1 0 —11— > o 5—1... _ oo 0 o 4 I I I I I I I I I I O 5 1O 15 20 25 Figure 3.14. Triol Number Reproducibility of cell positioning. Distilled water was pumped through the cell at 0.46 mL/min. Each trial represents movement of the cell from the viewing position to the fill position and back again. 85 four wash pushes were employed to reduce sample carry- over. One reagent was the glucose standards, while the Trinder/peroxidase reagent filled the other syringe. The Trinder reagent was composed of the following: 4 mg peroxidase, 2 mM AAP, and 2 mM DCPS made up to 25.0 mL with a pH 6.85 phosphate buffer, 0.05 M. The precision of the progress curve determination is shown in Figure 3.15. Both reactors displayed very good precision. The determination of initial reaction rates from repetitive experiments with three different glucose concentrations gave an average relative standard deviation of 7%. A preliminary study of the effect of reactor diameter on overall reaction progress displayed some interesting results. Figure 3.16 displays the data for the 1 mm i.d. reactor and Figure 3.17 for the 2 mm i.d. reactor. The enzyme reactors were not prepared at the same time, nor were the absorbances corrected for the effect of reactor diameter (equation 3.14). Despite these possible errors, it appears that the reaction was much slower in the larger diameter reactor. The reaction in the 2 mm i.d. reactor is about five times slower than that in the other reactor. Equation 3.8 predicts about a three-fold change in the reaction rate with this change in reactor diameter. However, the diffusion—control model does not predict the absolute reaction rate very well. 86 0.6 —— —I—- 0) 0.4 U C o / n 11 L. O (n .O < 0.2 —— o.o — ~ I I I I I I o 1 2 :5 Time (min) Figure 3.15. The results of five progress curve determinations with 0.44 mM glucose as the sample. The 1 mm i.d. reactor was used. 1.2 Absorbance 87 (.40 320 leo 8'0 Figure 3.16. Progress curves obtained with the 1 mm i.d. reactor. The glucose concentrations (mg/L) which were used are written at the end of each curve. 88 um 1.0 -T- Absorbance Tinie (swim) Figure 3.17. Progress curves obtained with the 2 mm i.d. reactor. The glucose concentrations (mg/L) which were used are written at the end of each curve. 89 The enzyme concentration also changes with the reactor radius. The concentration is directly dependent on the surface-to—volume ratio or 1/r. The increase in reactor radius from 0.5 mm to 1.0 mm could decrease the reaction rate and also change the ratio of the enzyme rate constant to the diffusion rate constant. In these ways, variation in reactor radius can change the kinetics indirectly. To keep the enzyme concentration constant with changes in reactor radius, additional enzyme should be bound to tubes of larger diameter. The stopped-flow instrument should provide a fairly direct route to the determination of inherent enzyme properties. By varying the reactor diameter and enzyme concentration, information about the enzyme reaction and diffusional influences should be obtainable. Reactors from these studies under static conditions can then be incorporated into a continuous flow manifold. Differences in kinetics between the two systems can be attributed to the change in mass transport, from diffusion to primary and secondary flow. Thus, definitive information con- cerning continuous flow systems with integrated open- tubular immobilized enzyme reactors should be forthcoming. Much more research on the kinetics of such reactors is needed. CHAPTER 4 Colorimetric Determination of Glucose Based on Immobilized Glucose Oxidase The determination of blood glucose is the most frequently performed test in clinical laboratories. Because glucose is the major energy source for living cells, any variation in the concentration of this sugar has a large effect on cell function. The most common disease related to carbohydrate metabolism is diabetes mellitus; nearly two percent of the U.S. population has this disease [95]. Glucose assays can reveal such abnor- malities, and then treatment of the illness can begin. Glucose is a monosaccharide, containing six carbon atoms. The structure of D—glucose, the form that is found in nature, is shown in Figure 4.1. The D—form has the hydroxyl group on carbon #5 on the right side of the carbon chain. The aldehyde group and the hydroxyl group on carbon #5 react to form a hemiacetal and a six-membered ring. The carbon #1 atom becomes asymmetric and, thus, can exist in two anomeric forms, a and B, shown in Figure 4.1. Mutarotation of glucose occurs slowly in aqueous solution [96,97]. At equilibrium, the D—glucose exists as 64% a—form, 36% B-form, and traces of the free aldehyde form [95,98,99]. The normal range of blood glucose concentrations is 70—105 mg/dL or 3.9-5.8 mM [95]. 90 D-glucose CHO H-C-OH Ina—EH. 1.42-01. HJC-OH £11,011 Aldehyde Form a-D-glucose B—D—glucose Figure 4.1. The three forms of D—glucose. 92 Because the determination of glucose has been so important for so many years, a large number of methods have been used. Non—enzymatic methods are based on the reducing nature of D—glucose and include reactions with ferricyanide ion, cupric ion, and o—toluidine [95]. These methods are sensitive but relatively non-specific. The enzymatic methods are much more selective. A list of the three common enzymatic systems is given in Figure 4.2. These systems have been used in immobilized forms. The hexokinase/glucose-6-phosphate dehydrogenase method has been adopted for use in the Technicon SMAC instrument, the major clinical analyzer in hospital laboratories. In this study, the glucose oxidase method was used for several reasons. First, the enzyme has a high Km value for glucose. This property ensures a large analytical range of glucose concentrations. Second, the enzyme and other reagents necessary for the assay are relatively inexpen— sive. The hexokinase and glucose dehydrogenase methods require B-NAD, which is expensive. Furthermore, commercial preparations of the enzymes are more than fifty times as expensive as glucose oxidase [103]. The reason that these other methods are used is that they are somewhat less susceptible to chemical interferences. Another reason glucose oxidase was chosen is its high specific activity (up to 150,000 Units/g) and the fact that it bonds well to nylon tubing. Therefore, glucose oxidase is well—suited 93 GLDH l) B-D-glucose + B-NAD ----- 9~ é—D—gluconolactone + B-NADH pH optimum = 7.6 [100] Measure B-NADH absorbance at 340 nm 3- HK 2) D-glucose + P04 ----- 9. D—glucose-6-phosphate D-glucose-6-phosphate+B-NAD 993§329g9>6-phospho-D-glu- conate + B-NADH pH optimum = 8.5 [101] Measure B—NADH absorbance at 340 nm 3) B-D-glucose + H20 + 02 __§Q_9. 6-D-gluconolactone + H202 pH optimum = 6.0 [102] Measure hydrogen peroxide conc. GLDH: glucose dehydrogenase (E.C. 1.1.1.47) HK: hexokinase (E.C. 2.7.1.1) GL-6-PDH: glucose-6-phosphate dehydrogenase (E.C. 1.1.1.49) GO: glucose oxidase (E.C. 1.1.3.4) Figure 4.2. The reactions involved in the enzymatic assays for glucose. 94 for use in immobilized form. Most assays using glucose oxidase are based on the increase in concentration of peroxide or the depletion of dissolved oxygen. Electrochemical methods have involved an alternate oxidizing agent, p—benzoquinone [102,104], the measurement of the decrease in oxygen concentration with an oxygen electrode [105,106], and amperometric sen- sing of the production of hydrogen peroxide [107-109]. Luminescence methods include the reaction of peroxide to give chemiluminescence [110], the reaction of peroxide and homovanillic acid to yield a fluorescent compound [111], and the measurement of oxygen via its ability to quench the fluorescence of pyrene [112]. A novel method based on the absorbance or fluorescence of B—NADH using additional en- zymes, catalase and aldehyde dehydrogenase, has also been described [113]. Most often, colorimetric methods are used. Hydrogen peroxide, in the presence of peroxidase, can oxidize many aromatic compounds to give colored products. Under the proper conditions, the absorbance of these molecules is proportional to the initial D-glucose concentration. A variety of color-forming reactions have been used; some are listed in Table 4.1. Note that two of the methods do not require peroxidase, making then less costly and less suscep— tible to chemical interferences. A recent study presented several more water—soluble reagents for reaction with 95 Table 4.1. Chromogens for Reaction with Hydrogen Peroxide 1 Reference Chemical System :flfii €(AU M- cm_ ) AAP—DCPS 510 5 x 104 AAP-HE 500 - DPAS 465 ~3 x 101 MBTH-CTA 572 - MBTH-DMA 590 - o-Dianisidine 410 - o—Toluidine 630 — KI/Moo42'* 365 2 x 104 Xylenol orange* 582 2.5 x 104 94,114,115 116 117 118 101,119,120 121 93 85 122 AAP: 4-aminoantipyrine CTA: Chromotropic acid DCPS: 2,4-dichlorophenol sulfonate DMA: N-N-dimethylaniline DPAS: p-diphenylamine sulfonate HB: p-hydroxybenzoate MBTH: 3-methyl-2—benzothiazolinone hydrazone * - peroxidase is not required 96 hydrogen peroxide [123]. The multitude of available chromogenic agents reveals the fact that no one method is significantly better than the others. The 4-aminoantipyrine (AAP) - 2,4-dichlorophenol sulfonate (DCPS) system was chosen for this study. The product has a high molar absorptivity, the reaction occurs rapidly, and the reagents are readily available and water soluble. A comparative study of eight chromogens showed that systems involving AAP were most suitable [124]. Also, a variety of phenolic compounds are acceptable coupling agents for AAP [125]. Trinder and coworkers first reported the use of AAP-DCPS in the determination of glucose [90,91]. Since that time, the reagent was been termed the Trinder reagent and has enjoyed wide use. This chapter describes the characterization of an immobilized glucose oxidase/Trinder determination of glu— cose. The chemical reactions and continuous flow instru- mentation involved in this study are reported in Section A. The Trinder color—forming reaction is explored in more detail in Section B. Next, experiments concerning the effect of reactor design on the assay sensitivity (Section D) and the characterization of the glucose oxidase reaction (Section C) are discussed. Finally, the sensitivity of the determination was improved by adding sodium azide to the system. The reasons for doing this are given in Section E. 97 A. The Method and Instrumentation 1. The glucose oxidase reaction The reaction which is catalyzed by immobilized glucose oxidase, the analytical reaction, is given below. B-D-glucose + 02 + H20 + 6-D-gluconolactone + H202 (4.1) The enzyme is specific for the B-form of the substrate. Glucose oxidase reacts much more slowly with the a-form and other sugars [126]. The enzyme considered here is extracted from Aspergillus niger and contains two flavine adenine dinucleotide (FAD) groups per molecule [127]. The reaction involves only the fully oxidized and fully reduced forms of the enzyme in the following manner [128]. E-FAD + B-D-glucose I E-FAD—glucose E-FAD-glucose + H20 2 E-FADH2 + 6-D-gluconolactone E-FADH2 + 02 + E-FAD + H202 The mechanism is a ping-pong bi-bi type, and, consequently, the rate equation is of the following form. v = k [E] /(1 + KG/[Glucose] + KO/[Oxygen]) (4.2) e o m m The reported kinetics constants for glucose oxidase have recently been reviewed [129]. The mean values for Kg, K3, and ke are 9 xlo’2 M, 6 1110'4 M, and 1.2 x102 s"1 respectively. Several compounds have been found to be inhibitors: 8-hydroxyquinoline, sodium nitrate, 98 semicarbazide [128], a-D—glucose [130], and hydrogen peroxide [131], among others. The glucose oxidase preparations used here were purchased from Sigma Chemical Company. Type II enzyme, obtained several years ago, was used for the studies reported in Sections A-D of this chapter. The enzymes used in the experiments described in Section E were all purchased in 1982. This distinction is important because the two lots of enzymes displayed different properties. 2. The Trinder/peroxidase reaction Because neither the substrates or the products of the analytical reaction absorb light in the visible region, a second, color-forming reaction is required. Hydrogen peroxide is a good oxidizing agent (H202 + 2 H+ + 2 e- + 2 H20, E° = 1.77 V), and is usually chosen for further reaction. The reaction of hydrogen peroxide and a hydrogen donor is slow, however, so a soluble enzyme, horseradish peroxidase, is used as a catalyst. The reaction mechanism is given below [132,133]. E + H O I Complex I 2 2 Complex I + AH2 1 Complex II + AH Complex II + AH + E + A + 2 H20 Here AH2 represents the hydrogen donor. Excess peroxide can combine with Complex II to produce an inactive enzyme species. The substrate specificity of the free enzyme is 99 high; only peroxide, formic acid, and acetic acid yield complexes that are active [133]. However, the speci- ficity of the enzyme-substrate complexes is poor. Many hydrogen donors can participate, including phenols, aromatic amines, ascorbic acid, and leucodyes. The overall reaction in the case of the Trinder reagents is given below. Cl (C) c' 0 CH3 2 H202 + c.@.o.. + N\ :99 GN£N@+ 3 H20 + H250, 503" (IQ-c”: CI "3° \CH3 (4 ' 3) H DCPS AAP 2H CH: The exact coupling mechanism of the two organic compounds is not known. A possible mechanism is that each reagent first loses two electrons and two hydrogen ions to the hydrogen peroxide, forming water and a reactive intermediate. Then the intermediates instantaneously react to form the dye product. Therefore, two hydrogen peroxide molecules and, consequently, two glucose molecules are required for the production of one molecule of dye. This fact is impor- tant when the molar absorptivity of the product or the percent glucose conversion is calculated. The peroxidase used in the following experiments was purchased in 1982 from Sigma (Type II) and contained about 200 Units/mg of solid. 3. The continuous flow manifold The analytical and color-forming reactions occur in the manifold of a continuous flow instrument. The reagents 100 are added, mixed, and reacted in a flowing, gas-segmented, liquid stream. Figure 4.3 is a diagram of the reaction manifold, showing the order and rate in which the reagents are added. A peristaltic pump meters the reagents. The composition of the reagents is considered below in the order in which they are added to the flow stream. Sample and Wash Reagents. In continuous flow methods, the samples and another liquid are alternately injected into the flowing liquid stream. This other liquid is called the "wash” solution, and in the following studies was distilled, deionized water (DDW). The wash solution mini- mizes sample carryover by rinsing the manifold tubing and flow cell between consecutive samples. The sample:wash timing was usually 40 seconds:20 seconds. Thus, the sample throughput was 60 samples/hour. The samples were either hydrogen peroxide or glucose standards. A hydrogen peroxide stock solution was pre- pared by diluting 0.50 mL of 30% peroxide solution to 0.500 liters with DDW. This solution was standardized by reaction with acidic KI and subsequent titration with standard thio- sulfate [134]. Working standards were made daily by appro- priate dilution of the stock solution with DDW. The primary glucose standard was prepared by dissolving 2.000 g of anhydrous, granular a-D-glucose and 0.5 g of benzoic acid in about 750 mL of DDW contained in a 1.000 L volumetric flask. The resulting soultion was then diluted to the 101 NOMINAL REAGENT FLOW RATE (mL/nfin) 0.10 0.05 G“ s T 0mm: 20 1 was“ Sample 0-23 0909 9000 0090 f Buffer A I Trinder/P0 0-10 510 nm Figure 4.3. The continuous flow reaction manifold. The pump speed setting which gave nominal flow rates was 42. 102 mark with DDW and allowed to stand at room temperature for 24 hours to ensure complete mutarotation. Working glucose standards were prepared by appropriate dilution of the primary standard with DDW. The hydrogen peroxide and glucose stock solutions were stored in a refrigerator at 4 °C. Gas Segments. The sample and wash slugs were seg- mented with gas bubbles using dual pump tubes. The "helper" tube at a nominal flow rate of 0.10 mL/min builds up pressure in the second, smaller i.d. tube, and the bubbles are forced rapidly into the liquid stream. This procedure helps phase the segmentation with the pump roller lift-off. Phasing of the gas bubbles improves the system characteristics. Room air was used unless otherwise noted. The segmentation rate was 1.67 bubbles per second at a pump speed setting of 42. Buffer Reagent. The next reagent added to the mani- fold is a buffer which provides the optimal hydrogen ion concentration for the enzymatic reactions. Potassium phosphate buffers (0.1 M) were used in these studies. The pH optimum of the chemical system was found to be 6.85. The experimental evidence is presented in Figure 4.4. Several drops of Brij—35 wetting agent were always added to the pH 6.85 buffer. This agent promoted pulseless flow through the manifold and preserved the integrity of the gas bubbles. 103 Figure 4.4. The pH dependence of the glucose assay. A 0.1 M phosphate buffer was used to control the pH. The dotted line indicates the approximate relationship in the absence of many data points. 104 Glucose Oxidase Reactor. After mixing in a five-turn glass coil, the sample and buffer were directed into the immobilized enzyme reactor. The preparation of the glucose oxidase reactor was detailed in Chapter 2, Table 2.3. The analytical reaction occurred in the presence of the bound glucose oxidase. The percentage of glucose that was reacted in the enzyme loop was never more than 10%. The product, hydrogen peroxide, was swept out of the reactor by the flowing liquid stream. Trinder/Peroxidase Reagent. Finally, the Trinder/ peroxidase reagent was added. The reagent was prepared as follows. A 10 mM stock solution of 4-aminoantipyrine (AAP) was pepared by dissolving 1.02 g of the solid (Sigma) in 0.500 L of DDW. A 10 mM stock solution of 2-hydroxy- 3,5-dichlorobenzenesu1fonate was prepared by dissolving 1.33 g of the solid (Research Organics) in 0.500 L of DDW. The working reagent was made by dispensing 5.00 mL of each of these stock solutions and 5 mg of peroxidase into a 50.0 mL volumetric flask and then diluting to the mark with the pH 6.85 phosphate buffer. This composite reagent was prepared daily. A 20-turn mixing coil (about 100 seconds residence time at nominal flow rates) was used at the exit end of the manifold to ensure a complete reaction between the peroxide and the Trinder reagents. The pink dye was then passed into the flow cell. 105 4. The continuous flow instrument A modular single-channel continuous flow instrument, diagrammed in Figure 4.5, was used in all of the experiments. A Brinkmann IP-12 peristaltic pump, modified to accomodate sixteen rollers, was used to proportion the reagents. The pump speed setting that provided nominal flow rates was 42. This speed was used unless otherwise mentioned. The mani- fold was constructed from 1.0 mm i.d. glass mixing coils, injector fittings, and a 1.0 cm flow cell (0.5 mm i.d., Technicon). Sampling was performed manually. The colorimeter was designed and built in this labor- atory. The light source was a miniature tungsten—halogen lamp (Model 03000, Welch-Allyn), and the light was trans- mitted to the flow cell via a glass fiber optic bundle (Model EK15-12, Dolan-Jenner). Wavelength selection was accomplished with a 510 nm interference filter (Ditric Optics, bandpass = 8 nm) mounted between the exit window of the flow cell and the detector. The detection system consisted of a photodiode (Princeton Applied Research) and an operational amplifier current-to-voltage converter. A bubble-gate removed the air segments electronically [135]. The data were collected and stored on a strip-chart recorder or in RAM by a microcomputer [86]. The data stored in the microcomputer were shipped to a PDP-8/e (Digital Equipment Corporation) minicomputer following each experiment for long-term storage and computation of the 106 Peristaltic — Reacfion Calorimeter _ Manifold UWW Reagents Sampling Pulse >0 O——>Waste Bubble Gate Microcomputer Data Lines Chart Recorder] Figure 4.5. Diagram of the continuous flow instrument. 107 steady-state absorbance values. The data were initial sample concentrations and their associated absorbance values. In most cases, a calibration curve of absorbance versus analyte concentration was constructed for a series of standards. The slope of the calibration curve was used as a measure of the immobilized enzyme activity. A larger slope meant that the assay conditions had improved. B. Experimental Studies of the Trinder Reaction With coupled enzyme reactions, the rate of one of the reactions can be measured accurately (< 2% error) if the two reaction rates differ by a factor of 50 or more. Consequently, the Trinder reaction must occur at a high velocity so that the measured rate depends strictly on the glucose oxidase reaction. Another way to prevent errors is to allow the indicator reaction to go to completion. This is only possible, however, when the two reactions are separated in time, as they are in this determination. Thus, conditions were adjusted so that the color-forming reac- tion went to completion. The concentrations of the Trinder reagents and the peroxidase were set as high as possible without unduly increasing the cost of the assay. An indicator reagent containing 6 mM AAP, 5 mM DCPS, and 80 Units/mL of per- oxidase was prepared and used in the continuous flow manifold. Likewise, 1:5 and 1:10 dilutions of this reagent 108 were prepared and used. The least concentrated reagent only produced a 3% smaller instrumental response than did the most concentrated. Thus, concentrations of 1 mM AAP, 1 mM DCPS, and 20 Units/mL of peroxidase were chosen as optimal; the reaction approached completion without wasting reagents. The optimal reaction time for the Trinder reaction was determined. With peroxide standards as the sample and the glucose oxidase reactor removed, the effect of reaction time on the instrumental response was small but noticeable. Assuming that the reaction was complete after 200 seconds, 96% of the 50 uM H202 sample was reacted within the first 100 seconds. Even at a pump speed setting of 56 and a 20- turn glass mixing coil in place (about a 70 second reaction time), the reaction was more than 90% complete. Results of stopped-flow studies of the Trinder reaction, shown in Figure 4.6, revealed that reactions with peroxide solutions of 50 uM or less were more than 90% complete in about thirty seconds. The Trinder reagent and peroxidase concentrations were the same as described in Section A. The hydrogen peroxide, 0 - 50 uM, reacted according to pseudo-first order reaction kinetics, with a rate constant of 0.09 sec—1. This confirmed that the Trinder reagents were in large excess and that the mechanism involving a sin- gle peroxide molecule, given above, is correct. In all the studies presented here, the Trinder reaction was made to 0.9 - <1) 0 C. O .C) 0.6- x. O (I) .0 <1 0.3 - 0.0 Figure 4.6. 109 A B C D . l 1 I . L, 1 l . l ' I ' I ' I ' I I I 7 14 21 28 35 Thne (seconds) Progress curves for the reaction of hydrogen peroxide and the Trinder reagent. The peroxide concentrations were 50 uM (A), 25 uM (B), 10 uM (C), and 5 uM (D). 110 approach completion with the use of excess reagents and long reaction times. The molar absorptivity of the Trinder product was determined. A series of hydrogen peroxide solutions of known concentration were tested, and the resulting steady- state absorbance values were recorded. To account for dilution in the manifold, the actual flow rates were cal- culated by weighing the amount of water delivered by the various pump tubes over a known period of time. The molar absorptivity at 510 nm was calculated to be 5 i 1 x104 AU M_1cm-1. This does not compare well with the value of 1.2 x104 AU M-lcm"1 reported in the literature [136]. The reason for this discrepancy is not known. In almost all reports concerning the GO/Trinder assay, several substances are mentioned as interferences. Uric acid, ascorbic acid, gentisic acid, and L-DOPA are most often cited [94,115,136,137]. These interferences occur primarily because the peroxidase-peroxide complex does not react specifically with a single reducing agent. The severity of the problem can be lessened by performing a protein precipitation on the serum sample. A common precipitating agent for glucose assays is the Somogyi reagent [138]. A stepwise treatment with Ba(0H)2 and then ZnSO4 precipitates the proteins along with uric acid and other metabolites, but ascorbic acid remains in the super- natant. 111 Ascorbic acid, added to the CF manifold, lowered the instrumental response. Figure 4.7 reveals the effect of ascorbate on the apparent peroxidase activity. The glucose oxidase reactor was not used, and a series of peroxide standards were sampled. The loss in assay sensitivity may be due to inhibition of peroxidase, loss of peroxide, or other factors. Stopped—flow studies were performed to determine the exact nature of the ascorbate interference. Ascorbic acid absorbs light at 261 nm; its oxidized form does not [93]. By monitoring the decrease in the ascor— bate concentration, it was found that hydrogen peroxide and ascorbic acid react rapidly, but only in the presence of peroxidase. Thus, ascorbate lowers the concentration of peroxide in the glucose assay. Further evidence was pro- vided by studying the formation of the Trinder product in the presence of ascorbate. A lag phase which increases with increasing ascorbate concentration is evident in Figure 4.8. The rate constants, measured after the lag period, were nearly independent of the ascorbate concentration. These facts suggest that ascorbate acts as an alternate substrate for peroxidase and, in this way, inhibits the enzyme. The problem of removing the ascorbate interference is a difficult one. Prior oxidation to dehydroascorbic acid is possible, but this risks oxidation of the glucose. Also, remaining oxidizing agents may react with the Trinder solu— tion. One researcher has used ascorbate oxidase to remove 112 29-— A I E D < .0 6 O :2 v a) u—lI- O. O (f) c: 3—— O 1.: C L. .0 —I— B O I l I I O I I I I O 10 20 30 4O [Ascorbote] (uM) Figure 4.7. The effect of ascorbate concentration on the assay sensitivity. 4.0 — 113 00 3.0 - (‘4- \\ .4 L_ 2.0 — :5 C 1.0 — 0.0 Figure 4.8. -II- I- 1- —I— -l. .3 -I- b Time (seconds) First order plots of data for the hydrogen peroxide-Trinder reaction in the presence of ascorbate. The ascorbate concentrations were 10 uM (A), 20 uM (B), and 40 uM (C). A was the absorbance at equilibrium, and A was the absorbance at any time, t. 114 ascorbic acid [136], but this approach is very expensive. A satisfactory solution has not been found. C. Reaction Characteristics 1. Segmentation gas composition In analytical reaction rate determinations, a linear relationship between the rate and the analyte concentration must be established. In the case of glucose, the Trinder reaction has a negligible effect, and, thus, the reaction rate is described by equation 4.2. In order to attain a rate that is only proportional to the glucose concentration, the oxygen concentration must be larger than 20 K: or 1 x10.2 M, and the glucose concentration must be smaller than Kg/ZO or 4.5 x10.3 M. The concentration of glucose after dilution or dialysis of the sample is usually well below 4.5 mM. The concentra— tion of oxygen in 0.1 M phosphate buffer saturated with air is 0.26 mM at 22 °C [139]. This value is less than the K3 value, and, therefore, linear calibration curves should not be possible. This is not the case, however. A large source of oxygen is provided by the air bubbles which segment the liquid stream. To prove this, the effect of segmentation gas composition on the determination was studied. The manifold and reagents were used as described in Section A. The gases, N and 0 were bled from pressur- 2 2’ ized tanks and allowed to fill a large bottle. The 115 appropriate pump tube was placed in the bottle. Room air was the source of air. A series of glucose standards was sampled, and the resulting calibration curves are shown in Figure 4.9. When the assay is entirely dependent on dis- solved oxygen, the analytical range of glucose concentra- tions is limited to 0.5 mM or less. The linearity improves with increasing oxygen concentration. With air segmentation, samples containing less than about 3 mM glucose can be tested; this is well within the range for blood glucose analysis. The gas composition also affected the continuous flow peak shapes. When nitrogen gas was used, large Spikes occurred at the falling edges of the peaks. See Figure 4.10. These spikes were due to air bubbles which entered the sam- pling system as the probe was moved between the sample and wash solutions. With the extra oxygen, the reaction could occur to a greater extent. Small spikes also occur with the use of air segments. 2. The effect of ionic strength The ionic strength of the solution passing through the glucose oxidase reactor had little effect on the enzyme activity. The usual continuous flow manifold was used. The buffer concentration was lowered to 1.0 mM, so that the ionic strength was controlled by adding KCl to the sample and wash solutions. Chloride ion is not known to affect either of the enzymatic reactions. The ionic strength 116 1.250 —- 0 2 O O 1.000 —— AI r 0) U ...— : 0.750 N C D 2 L O m __ D 0.500 < 0.250 -—-— 0.000 I I I I I I I I a I 0 1 2 3 4 5 [Glucose] (mM) Figure 4.9. Calibration curves for a series of glucose standards using three different segmentation gases: NZ, Air, 02. ABSORBANCE 117 N2 AIR 02 Figure 4.10. TlME The output from the continuous flow instrument using three different seg- mentation gases. The glucose concentra- tions were 0.44 mM, 1.44 mM, 2.22 mM, and 4.44 mM. Only the first three concentrations were sampled using 02. 118 through the final mixing coil, in which the color- forming reaction occurred, was held nearly constant by preparing the Trinder reagent in 0.1 M buffer. Conse- quently, the final ionic strength only varied from 0.010 M to 0.015 M. Figure 4.11 shows an initial increase and then a slow decline in instrumental response as the ionic strength is increased. Only a small change was expected since the substrates are uncharged. However, since the carrier was positively charged, enzyme-carrier interactions may have been altered. According to equation 3.3, an increase in ionic strength should increase the Km value and lower the reaction rate. This study shows the same general trend. 3. The specificity of glucose oxidase The oxidation of most sugars, except B-D-glucose, is catalyzed very slowly by glucose oxidase. Nine different sugars, obtained commercially in the highest purity forms, were sampled, first with the immobilized glucose oxidase reactor in place, and then with soluble glucose oxidase added to the buffer reagent (0.2 mg/mL) and the enzyme reactor removed. The continuous flow manifold and the other reagents were the same in both experiments. Table 4.2 is a list of the results. No sugar other than B—D— glucose showed high reactivity toward glucose oxidase. The aqueous and immobilized enzyme results were remarkably similar. This indicates that the conformation of the enzyme 119 0.3 m—- l __ E I) < M r!) 0.2 "" *3 O 3? a) -r- 0. _° (.0 c; 0.1 ""_ .2 ‘6 L. :9 l_ B U l l l I 1 4| 1 l I J 0.0 I l T I l l I l I 1 O 10 20 30 4O 50 Figure 4.11. Ionic Strength (an) The effect of ionic strength on the glucose oxidase reaction. 120 Table 4.2. Specificity of Glucose Oxidase Relative Reactivity Sugar Immobilized Agueous Aqueous* B-D-glucose 100 100 100 2—deoxy—D-glucose 2.5 2.4 25 Maltose 1.1 0.95 0.19 a—D-glucose 0.6 - 0.64 Xylose 0.31 0.26 0.98 Mannose 0.27 0.23 0.98 Galactose 0.03 0.05 0.14 Fructose 0.03 0.02 - Glucosamine 0.03 0.02 <0.05 Sucrose 0.00 0.00 - * - Values reported in the literature [126] 121 was not significantly altered during immobilization. Correlation of the results with an aqueous enzyme study, performed in 1952, was fair. The literature values may be inflated by impure reagents; the amount of B—D-glucose in various sugar preparations is probably much lower today. 4. Mutarotation kinetics To replace the B-D—glucose which is converted to hydrogen peroxide, the a-form of D—glucose is converted spontaneously to the B—form. This process is termed muta- rotation. The chemical equation and associated rate equation are shown below. B-D-glucose I a-D—glucose (4.4) -d[B]/dt = k1[B] — k2[d] (4.5) If [B]O-[B] is substituted for [a] ([B]O is the initial concentration of the B—form) and the equation integrated, the rate equation appears in the following form. -1n(1.565[B]/[B]O - 1) = (k +k2)t (4.6) 1 The term, k +k is identified as the ”mutarotation 1 2’ constant" [96,97]. A 4.44 mM B-D-glucose solution was prepared, and immediately the solution was sampled with the continuous flow instrument. Steady-state absorbance values were obtained for the glucose solution at intervals over a two hour period. Assuming that the a-form had a negligible 122 reaction rate with the immobilized glucose oxidase, first order data were calculated and plotted in Figure 4.12. The mean rate constant was determined to be 0.017 r 0.002 min , in good agreement with the literature value of 0.012 min‘1 [96,97]. The specificity value for d-D-glucose was also obtained. Initially, only the B-form existed in the sample solution, and, thus, the sensitivity toward the B-form was calculated. Knowing this and that 64% of the glucose is of the B-form at equilibrium, reaction rate measurements at equilibrium allowed calculation of the a—form reactivity. The following equation was derived. S /S = 2.78(A /A ) - 1.77 (4.7) a B m 0 Here 80 and S are the sensitivities of glucose oxidase 8 toward the a and 8 forms of glucose, and Am and A0 are the equilibrium and initial absorbances. The reactivity of the d-form was only 0.6% of that for the B—form. Glucose oxidase shows very high specificity indeed. The effect of glucose mutarotation on the overall reaction kinetics is small because the mutarotation rate constant is about ten times smaller than the glucose oxidase reaction rate constant. If one considers a 30 second residence time in the enzyme reactor, up to 10% of the B-D—glucose is consumed by the enzyme, while only about 0.5% of the a—form can be converted to the B-form in -|n(1.565 [Bl/[1810 -1) Figure 4.12. 123 l 1 l I ' l 1 l ' I 30 60 90 ‘I’ Thne (nfinutes) First order plot of data for the muta- rotation of a 4.44 mM B-D-glucose solution at 22 °C. The results of two independent experiments are presented. [8] is the concentration of the B-form of D-glucose at any time, t, while [B]O is the initial B-form concentration. 124 the same time period. Consequently, once the standard glucose solutions have come to equilibrium, the muta- roatation process can be neglected. D. Reactor Design Characteristics 1. Coiling diameter Immobilized enzyme reactors are usually molded into helical coils to make them easier to incorporate into a continuous flow manifold. The coiling of the reactor should also improve the radial mixing in each liquid seg- ment [83]. The reaction rate should increase with better radial mass transport, and, thus, with more tightly wound reactors. Horvath and coworkers found that the apparent activity of trypsin immobilized inside a nylon tube improved as the coiling diameter decreased [140]. However, this effect only occurred at flow rates greater than 3 mL/min. By improving the radial mixing, coil formation de- creases sample dispersion, or the ”wash" of the system. Ideally, the rising portion of a continuous flow peak can be described by the following equation [1]. H(t) = H(ss)(1 - exp(-t/b)) (4.8) Here H(t) is the absorbance at any time along the peak, H(ss) is the steady-state absorbance (flat portion of the peak), t is the time, and b is a measure of the magnitude of the sample dispersion. The term, b, is called the wash 125 value of the system and is defined as the time required for the absorbance to change from H(t) to 0.37 H(t). Smaller b values are desirable and can result from tighter coiling of the mixing loops and enzyme reactor. A 25 cm glucose oxidase reactor, coiled to various diameters, was inserted into the manifold, and a series of glucose standards, 0.4 to 4.4 mM, was sampled. The manifold and reagents were the same as those described in Section A. The total flow rate through the reactor was only 0.38 mL/min. The results are presented in Figure 4.13. The reaction rate was unaffected by the changes in the reactor design. This indicates that either the reaction rate is limited by the enzyme kinetics or that radial substrate transport was not improved significantly at the low flow rate. The latter explanation is in agreement with Horvath's work [140]. The wash, however, was improved by about 30%. Thus, small coiling diameters are desirable, but the coiling process does not have to be very precise to obtain good analytical results. 2. Coil temperature An experiment was performed to determine the effect of temperature on the apparent activity of glucose oxidase. The manifold and reagents were the same as those described in Section A. The enzyme reactor was heated by gluing the reactor into a plastic drying tube, sealing the drying tube, and then passing heated water through the tube. The water, 126 1.2 Er- WA 7— 9.0 U —6 l __ _r_ ‘2 2:) 4 N 0.8 -h -- 6.0 O 32 a) " 1' o. 2 (f) c 0.4 m—' ‘—'3.0 2 B L... :9 .. _- B O L l l I l l 0.0 I l 1 I 1 l 0.0 O 20 4O 60 Figure 4.13. CoiHng [Morneter (nun) The effect of coiling diameter on the calibration slope (O) and wash ([3). (spuoaas) usoM 127 contained in a 2 liter beaker, was thermostatted using a Haake Model E52 heating bath/circulator unit. Since the color-forming reaction went nearly to completion, tempera- ture had little effect on the Trinder reaction. The temperature dependence of the analytical reaction is shown in Figure 4.14. The results do not conform to the Arrhenius equation. It is also interesting to note that at room temperature (22 °C) a small temperature shift can change the reaction rate by several percent. This fact could explain the fluctuations of the activity of enzyme reactors in long—term stability studies (e.g. Figure 2.5). The stability of the enzyme reactor at high temperatures was tested by heating the reactor to 41 °C, and sampling glucose standards over a one hour interval. The initial and final results were identical, which indicates that immobilized glucose oxidase is very stable with respect to temperature. E. The Characterization and Elimination of the Effect of Catalase 1. The nature of the interference In the glucose determination described here, the hydrogen peroxide produced in the immobilized enzyme tube is reacted downstream with the Trinder/peroxidase reagent. The analytical and color-forming reactions are separated in time for two reasons. First, separation allows the reactions to be optimized individually in terms of reaction 128 4.00 -— 3.00 2.00 1.00 —— CaHbrofion Slope (x102 AU hi—1) Mb I... 0.00 ‘— "ll" _L 15.0 25.0 35.0 45.0 Tenwperature (°C) Figure 4.14. The dependence of the assay sensitivity on temperature. 129 time, reagent concentrations, pH, etc. In soluble enzyme methods, a compromise between the optimal values for each reaction is necessary. Second, the separation improves the wash characteristics of the system. The Trinder compounds adsorb strongly to the nylon tubing, which degrades the wash. These two factors prohibit the simultaneous reaction of glucose and peroxide inside the immobilized enzyme reactor. Because of the separation of the two reactions, the sensitivity of the method was found to be reduced due to contamination of the glucose oxidase by catalase (E.C. 1.11.1.6) [141]. Catalase destroys hydrogen peroxide via the following mechanism [142]: + E + H202 + Compound I Compound I + H202 + E + 02 + 2 H20 where E is free catalase. Some of the hydrogen peroxide that otherwise would react to form product is destroyed inside the enzyme reactor by co-immobilized catalase. Since catalase has a specific activity of forty million units per gram [103], the small amounts found in most glucose oxidase preparations will have a large effect. If the color-forming reaction also occurred inside the reactor, then the effect of the catalase would be small since the peroxide would quickly be converted to the colored product by aqueous peroxidase. This is not the case in CF systems. 130 In a method proposed recently, the catalase was selectively inhibited by adding sodium azide to the system [141]. Hydrazoic acid is well known as an inhibi- tor of catalase [142-144]. Although the exact nature of the inhibition is still in question, it is known that hydrazoic acid reacts with Compound I to yield peroxide or other products. Thus, the addition of azide to the reaction system inhibits catalase by forming a compound with the enzyme-substrate complex. A study of the effect of sodium azide on the co- immobilized catalase is presented here. The improvement in the assay sensitivity with the addition of azide and the variation in catalase content of several commercial glucose oxidase preparations are described in detail. Finally, suggestions for the detection and elimination of the catalase interference are made. 2. The reagents and continuous flow instrument The reagents were prepared as outlined in Section A. Sodium azide solutions were prepared in the following manner. A 250 mM stock solution was made by dissolving 3.25 g of NaN in 0.200 L of pH 6.85 phosphate buffer (0.1 M). 3 This solution was stable for several weeks at room tempera— ture. Working azide solutions were prepared by appropriate dilution of the stock with buffer. Five drops of Brij-35 surfactant were added to these solutions prior to use. The azide solutions were substituted for the buffer reagent 131 in the continuous flow manifold. The term, "azide” is used to mean sodium azide. The concentrations of azide, hydrogen peroxide, and glucose reported here are the concentrations actually present inside the immobilized enzyme reactor, after dilu- tion in the continuous flow manifold. The instrument was the same as outlined in Section A. The data were acquired by the computer in all the experiments, and the pump was operated at a speed setting of 56. 3. Enzyme immobilization The catalase (Type C-10) and glucose oxidase (Types II, IX, and X) were purchased in 1982 from Sigma Chemical Co. and used without further purification. The enzymes were immobilized as outlined in Chapter 2, Table 2.3. The seven enzyme reactors used in this study were prepared with the aqueous enzyme solutions listed in Table 4.3. The listed catalase concentrations were derived from the catalase found in the glucose oxidase and from the amount of pure catalase added to the solutions -- reactors A, C, and D. The different reactor types will be referred to by letter names as listed in Table 4.3. When not in use, the reactors were filled with pH 6.85 phosphate buffer and stored at 4 °C. 4. Effect of azide on aqueous peroxidase A series of experiments was performed to investigate the effect of azide on the color—forming reaction. In the first experiment, the immobilized enzyme reactor was removed 132 Table 4.3. Solutions Used to Prepare the Immobilized Enzyme Reactors Reactor Type GO (U/mL) CAT (U/mL) Sigma Type GO A 0 3765 - B 89 0.4 II C 89 380 II D 89 1880 II E 125 0.58 II F 125 1200 IX G 125 1.7 X 133 from the manifold, and peak heights were obtained for a 79 0M peroxide standard. Then the experiment was re- peated with increasing amounts of azide added to the buffer reagent. The results, shown in the upper curve of Figure 4.15, revealed that for azide concentrations of up to 30.3 mM the sensitivity of the assay remained con- stant within the limits of experimental error (about 3%). Thus, azide does not appear to inhibit peroxidase activity or interfere with the color-forming reaction. 5. Effect of azide on immobilized catalase In the next series of experiments, the ability of azide to inhibit catalase activity was investigated. A- 30 cm immobilized catalase reactor (reactor A in Table 4.3) was inserted into the manifold, and peak heights were obatined for a 79 0M peroxide standard with increasing concentrations of azide in the buffer. The results of this experiment are also shown in Figure 4.15. Without azide in the buffer, more than 99% of the peroxide was destroyed as it passed through the catalase reactor, but as the azide concentration increased, peroxide recovery improved drama- tically. Even at an azide concentration of 15 mM, however, 20% of the peroxide was destroyed. The manner in which azide inhibited catalase was investigated in the next set of experiments. Shorter catalase reactors (type A) were used in these studies so that the mean residence time of peroxide in the enzyme 134 100M 1'] ,0 80—— U (I) L. a > o o __ a) 60 a: o E x 40- E Q) a. N 20— 1 1 1 l 1 I or”1 T I I l I l O 10 20 30 Afide Concentnfljon OnM) Figure 4.15. The effect of azide on the amount of hydrogen peroxide that was converted to product. A 79 uM H 0 solution was sampled without an enzyme reactor ([3) and with a 30 cm catalase reactor (0) in the manifold. 135 reactor was about 3.5 seconds. Under these conditions, Michaelis-Menten kinetics were applicable. A new 5 cm segment was cut from a fresh 50 cm catalase reactor and inserted into the manifold for each experiment. The dif- ferences between peak heights obtained with and without the catalase reactor in the manifold were determined for a series of hydrogen peroxide standards in the range of 20 0M — 80 0M. The decrease in steady-state absorbance with the catalase reactor in place was proportional to the catalase-induced rate of peroxide decomposition. Figure 4.16 shows Lineweaver-Burk (L-B) plots of the results of three separate experiments. All lines inter- sected in a small region just below the x-axis. Dixon plots [145] of the data were hyperbolic, and replots of the L-B intercepts and slopes were non-linear. These results indicate a mixed-type noncompetitive inhibition [146]; in addition to interaction with Compound I, azide may also bind at two sites on the free enzyme. The kinetics were complicated, and more experiments would be required to completely elucidate the action of azide on immobilized catalase. We next investigated the extent of catalase interfer- ence under non-separated reaction conditions. The Trinder/ peroxidase reagent was passed through the 5 cm catalase reactor (type A). The results revealed that peroxidase successfully captured 80% of the available peroxide in the 136 \\ g‘ l I G ‘\ \\ A l I K‘ ‘0 o > O l 1 I I \ 1x1. >0 o\\~ a a 0‘ Q. l/Rate (x105 M—Is) -4 I I l —4 0 4 1/[H202] (x104 M“) .0- m_ll_ Figure 4.16. Lineweaver-Burk plot of the data which exhibited azide inhibition of immobilized catalase. The azide concentrations ranged from 0.0 (A) to 1.5 mM (B). 137 presence of catalase. Thus, for the small amount of catalase impurity normally found in commercial glucose oxidase preparations, the loss of peroxide from the catalase reaction in the presence of peroxidase should be negligible. It should be noted that azide did not have an imme- diate effect on the immobilized catalase. The azide/buffer solution had to be passed through the reactor for about three minutes before its full effect was attained. To re- move the inhibitory effect of azide, the reactor was flushed with pure buffer for five minutes. This process, however, only returned the catalase to 90% of its original activity. Thus, azide inhibition of catalase appears to be only weakly reversible, in agreement with a previous study [142]. 6. Effect of azide on immobilized glucose oxidase A preparation of glucose oxidase containing a very small amount of catalase was chosen for this study. A 50 cm glucose oxidase (GO) reactor (type B) was inserted into the CF manifold, and glucose standards in the range of 0.1 mM - 0.5 mM were used as the sample. Curve B of Figure 4.17 shows the effect of azide on the GO reactor. Azide solutions of more than 0.5 mM had a large, negative effect on glucose oxidase activity. Dixon plots of these data, Figure 4.18, show that the inhibition was purely non- competitive and that the KI value was 18 mM. 138 1.2 5— Calibration Slope (x103 AU M“) O O) -II- I I 0 4 8 12 “JE Azide Concentrohon (nnM) Figure 4.17. The effect of azide on three glucose oxidase reactors, type B (A), type C (0), and type D (EJ), containing various amounts of catalase (refer to Table 4.3). 139 16n— -I-- '1 3 m I cu— IE '6 L0 8“ ' o 9 2 x - V 4 1. '95 m ' e 1 +1 O -11— ‘9: m . \\ r- O_.._. —4 I I I I I I I I I —20 —1O 0 10 20 Azide Concentration (mM) Figure 4.18. Dixon plot of data from Figure 4.17, curve B. Three glucose concentrations were used: 0.44 mM (1), 0.22 mM (2), and 0.11 mM (3). 140 Two other reactors were prepared from Type II glucose oxidase spiked with small (type C) and large (type D) amounts of catalase. The different concentration ranges at which azide first inhibits catalase and then glucose oxidase in these reactors can be seen in Figure 4.17. The optimal azide concentration increased as the amount of catalase in the reactor increased. The rapid rise in assay sensitivity observed as the concentration of azide in the buffer approached 1 mM is due to catalase inhibition. The more gradual decrease in assay sensitivity at higher azide concentrations is due to glucose oxidase inhibition. We were surprised that the addition of a second enzyme to the immobilization solution had only a small effect on the bonding of the primary enzyme. At an azide concentration of 15.2 mM, at which the action of catalase is very small, the calibration slopes of all three reac— tors (types B,C, and D) were similar (relative standard deviation between reactors = 4%), which indicates that the amount of active glucose oxidase bonded in all three reactors was nearly the same. 7. Effect of azide on different preparations of glucose oxidase Reactor types E, F, and G (40 cm in length) repre— sented immobilized forms of Sigma Types II, IX, and X glucose oxidase respectively. The catalase activities of 141 of these preparations (Table 4.3) were very different, and the differences can be seen if one compares the points at zero azide concentration for the three curves in Figure 4.19. The calibration curves were prepared by sampling peroxide standards in the range of 10 uM - 40 uM. Reac- tor F exhibited very little activity which reflects the higher catalase activity. Addition of azide improved the sensitivity only a small amount for reactor types G and E, while the charac- teristics of reactor F changed noticeably. At an azide concentration of 15.2 mM, the assay sensitivity with reactor type F in place approached that of the other two reactors. Again, the presence of a second enzyme had only a small effect on the bonding of the primary enzyme. 8. A method to determine the presence of catalase The properties of the same enzyme preparation can vary from lot to lot. For example, some Sigma Type II glucose oxidase purchased five years ago exhibited high catalase activity [141], while a batch purchased several months ago showed negligible catalase activity. Thus, it is important to know the catalase activity in the glucose oxidase preparations to be used. To determine whether catalase activity in a given lot of glucose oxidase is sufficient to cause problems in an immobilized enzyme glucose assay, a simple test is suggested. First, immobilize the glucose oxidase in a 50 cm or longer 142 A IE 0 6’0 1.0 C G E I) F <( 1 d- x 1-11— v CL 2 Hi- (I) 0.4 ‘I' C O -U- 1; S 0.2 ~— .0 ES - C) 0.0 I I I I I I I I O 4 8 12 16 Figure 4.19. AJide Concentrofion (nflw) The effect of azide on three commercial preparations of glucose oxidase, reactor types 0, E, and F (refer to Table 4.3). 143 reactor, and then cut the tube into 10 and/or 20 cm segments. Determine, with glucose as the sample, the calibration slopes with various lengths of reactor in the CF manifold. If the catalase content of the reactor is significant, a curved plot of the calibration slope versus reactor length will result. Figure 4.20 shows the results of such an experiment. As the catalase activity in the reactors increases (types B+D), the plot becomes more hyperbolic. To verify the presence of catalase, azide is added to the system so that the concentration inside the enzyme reactor is about 1 mM. If straight lines are produced, as shown in Figure 4.21, then catalase inter- ference is confirmed. In this way, one can determine if the use of sodium azide is warranted. 9. Conclusions The results presented here should be applicable to systems that use chromogenic agents other than the Trinder reagent. The analytical and color-forming reactions, how- ever, must be separated in time for the effect of azide to be evident. Azide does not affect the accuracy of control serum assays. Table 4.4 shows that no significant differ— ence was found between the results obtained with and with- out added azide. The sensitivity of the glucose determina- tion is improved by the addition of sodium azide when glucose oxidase immobilized enzyme reactors, containing appreciable amounts of catalase, are used. 144 1.0 *- l 22 :3 0.8 -— a 8 <1 [(7 O ; CL6-— v e C a 0. O O —I— 5; 0.4 . 9 c o o +) —n-— L. .o e ' B 0 0.0 I I I I I I O 20 40 60 Reactor Length (cnfl Figure 4.20. The change in calibration slope with different lengths of reactors B, C, and D (refer to Table 4.3). No azide was added. 145 1.0 fl— O.8 ~— Calibration Slope (x103 AU M“) 0.0 {I I I I O 20 4O 60 «I- «L- - Reactor Length (cm) Figure 4.21. The change in calibration slope with different lengths of reactors B, C, and D (refer to Table 4.3). The reactions occurred in the presence of 0.8 mM azide. 146 Table 4.4. Glucose Concentrations Found in Control Sera With and Without Added Azide Glucose Concentrations (mg/dL) Found Found Control Serum Expected (no azide) (2 mM azide) Calibrate 165 143 153 Monitrol 95 91 95 Sigma 92 74 75 The 95% confidence limits for the experimentally determined values are i 5 mg/dL. CHAPTER 5 Enzyme Immobilization on Non—porous Glass As shown in previous chapters, nylon serves well as a carrier for immobilized enzymes. Yet, nylon is far from perfect in many cases. Aromatic compounds tend to adsorb to the nylon, and nylon has a small surface area-to- volume ratio. Also, a charge develops an the surface during immobilization; this charge can alter the enzyme properties. Since these problems are not easily solved, the use of other carriers has been an active area of research. Glass offers the advantages of being inert, sturdy, and readily available in a variety of forms; glass shows good wash characteristics in continuous flow manifolds. 0n the other hand, non-porous glass has a small surface area-to—volume ratio and has no reactive groups for direct enzyme attachment. These problems can be overcome, however, by growing “whiskers” on the glass to increase the surface area and reacting the glass with a silynating reagent to activate the surface. Coiled glass tubing for use in continuous flow instruments is somewhat expensive, but the use of straight tubing in the stopped-flow instrument presented in Chapter 3 or the use of beads packed into a tube is ideal. The preparation and use of immobilized enzyme glass reactors are discussed in this chapter. Glucose oxidase was 147 148 again chosen as the enzyme; the reactions were discussed in Chapter 4. First, glass tubing was used in a seg— mented CF system. The results are presented in Section A. Then glucose oxidase was attached to non-porous glass beads for use in a flow-injection instrument. Section B includes analytical results from this novel system. A. Immobilization of Glucose Oxidase on Glass Tubing Borosilicate glass tubing of 1.25 mm i.d. and 30 cm in length was bent in a U-shape for use in the segmented continuous flow instrument. The glucose oxidase was attached to the glass by procedures adapted from those described by Iob and Mottola [147]. Four major steps are involved: tube cleaning, ”whisker” formation, activation, and enzyme bonding. The details are given below. The glass was cleaned first by filling the tube with concentrated HCl and heating it to 60 °C for 24 hours. Next, the tube was washed with 100 mL each of distilled water and acetone. The glass was dried by passing nitrogen gas over the surface. In the second step, the glass was etched and silica needles were formed on the inner surface of the tube. The procedure was taken from that reported by Onuska, §£.al. [148]. The tube was filled with a 5% w/v (saturated) methanol solution of ammonium bifluoride, NH F-HF for about 4 one hour at room temperature. The excess solution was 149 removed by passing a stream (1 mL/min) of nitrogen gas through the tube. The tube became milky white in colora— tion as the methanol was removed. A constant, but slow, stream of gas was crucial for uniform coverage of the tube with the reagent. This was the most difficult portion of the treatment. The ends of the coated tube were then sealed, and the tube was placed in a muffle furnace at 425 °C for three hours. Under these conditions, the bi- fluoride is decomposed to ammonia and hydrogen fluoride gas. The HF removes glass at one point and then deposits silica in the form of ”whiskers” at other points on the tube. The surface area of the reactor should be increased greatly. After cooling, the tube was opened and flushed with nitrogen gas. Activation was effected by reaction with 3-aminopropyltriethoxysilane (Sigma). The original pro- cedure was reported for activation of porous glass beads [149,150]. A neat solution was reacted at 60 °C for 1.5 hours or a 5% v/v solution in acetone was reacted at 60 °C for 24 hours. Both procedures were hampered by the fact that the solution slowly evaporated, leaving behind a yellow gum, which often restricted solution flow through the tube. The formation of the residue needs to be reme— died before this procedure can be used reliably for the activation of glass tubing. Perhaps lower reaction tem- peratures and times will have to be used. The activated tube was then washed with acetone and 150 water and filled with a 5% v/v solution of glutaraldehyde in pH 8.0 phosphate buffer. The reaction occurred for 1.5 hours at room temperature. No change in the colora- tion of the tube was evident. The glucose oxidase, 10 mg/mL in pH 6.85 phosphate buffer, was next added to the tube and I allowed to react for 18 hours at 4 °C. After these pro- cedures were performed, the tube was washed with distilled, deionized water and filled with the phosphate buffer when not in use. The results of the experiment were somewhat encoura- ging. The activity of the reactor was tested as described in Chapter 4. The best glass reactors had activities that were about 25% of those prepared from nylon tubing. The wash characteristics were somewhat better. Considering the small amount of time spent on this venture, the results were very good. Further testing is warranted. B. Glucose Oxidase Bead String Reactor In flow—injection (FI) analysis, sample dispersion and mixing between the reagents and sample are coupled. In order for complete mixing to occur, the sample dispersion must be fairly large. Consequently, sample throughput and sensitivity are sacrificed as longer residence times are employed. In 1981, van der Linden, et.al. [151,152] suggested the use of tubing packed with impervious glass beads, which have a diameter that is 60—80% of the inner 151 diameter of the tube; the single bead string reactor (SBSR) was born. The liquid must travel a tortuous path through the reactor. Mixing is enhanced, while dispersion is about ten times smaller than in open tubing of the same dimensions. Wide use of such packed tubes is expected in future FI research. Since tiny, porous glass beads had been used as carriers for enzymes in the past, it was expected that larger, non-porous glass beads would also prove to be good carriers. Once immobilized on the beads, the enzyme could be packed into a tube and used in a flow-injection instrument. The advantages of immobilized enzymes and SBSRs would be combined to yield a powerful technique. Jim Litch tested this method during an undergraduate re- search project. The immobilization procedure is listed in Table 5.1 and follows the same concepts as described for glass tubing. The treatment was easier to perform, since most of the reactions occurred in a beaker, and the beads were easier to handle than the tubing. The beads acquired a red color due to the reaction between the glutaraldehyde and the amine group on the glass; this is a sign of a successful activation step. A diagram of the completed reactor is shown in Figure 5.1. A large overall reaction rate is promoted by the rapid mixing and the intimate contact between the enzyme and the bulk solution. Table 5.1. Immobilization Procedure for Glass Beads Step 1. Cleaning 2. Whisker formation 3. Silanate 4. Enzyme 152 Reagent Chromic Acid 6 M HCl Distilled water Acetone Dry with N2 5% w/v solution of NHuF-HF in methanol Dry with N2 Apply heat 2% acetone solution of 3-aminopropyl- triethoxysilane -Decant excess Apply heat 1% glutaraldehyde in pH 8.0 buffer 5 mg/mL solution of glucose oxidase in pH 6.85 buffer Time 15 min. 10 min. 10 min. 10 min. 1 hour 3 hours 30 min. 24 hours 30 min. 24 hours Temp. (°C) 23 23 23 23 23 450 23 90 23 153 0.8 mm i.d. 0.5 mm diam. Teflon Enzyme Tubwng Coated Bead Figure 5.1. View of the immobilized enzyme single bead string reactor. 154 The continuous flow instrument described in Chapter 4 was modified to include a different manifold and to omit the bubble—gate. The manifold that was used is shown in Figure 5.2. A four-way slider valve (Altex), pneumatically actuated, and controlled by a ZX81 microcomputer (Sinclair) interfaced with a relay board (Byte-Back), was at the heart of the system. In position 1, the carrier stream (Trinder/ peroxidase reagent) was pumped through the valve and SBSR. In position 2, the sample (glucose) was pumped into the system. The volume of glucose injected was determined by the pump speed and the time the valve spent in position 2. Sample volume = (Pump speed)(Valve time) (5.1) Each experiment included one cycle of the valve. The valve times were controlled by ZX81 software so that the sample volume could be varied reproducibly over a wide range. In such systems, mixing only occurs at the leading edge of the sample zone until the valve is returned to position 1. Thus, mixing between reagents and sample in this design may differ from more common FI instruments which employ a 6-way valve and a sample 100p. The analytical and color-forming reactions occurred simultaneously inside the reactor. Results indicated that the indicator reaction was much faster than the glucose oxidase reaction. This is necessary for a useful deter— mination. All the reagents were prepared as described in 155 Glucose 4 1,Waste IIME- Waste \ SBS fl THnder/PO Colorhneter ZX81 510 nm Figure 5.2. Schematic diagram of the flow-injection instrument. 156 Chapter 4. The reactor was 21 cm in length and had an internal diameter of 0.8 mm. If the beads take up 30% of the total reactor volume, the void volume of the reactor was about 75 uL. The sample volume was varied at four different flow rates. When large sample volumes were used, two peaks often resulted from a single sample injection. The reason for this was incomplete mixing; the reaction occurred at both ends of the sample slug, but the reagents never reached the center of the slug. The two peaks varied as to which was the higher, but in all cases the heights were within 25% of each other. Large volumes were de- sirable because increased instrumental output resulted. However, if the sample volume was too large double peaks occurred. The optimal sample volume for the system was found by plotting the volume between the double peaks versus the sample volume. The x-intercept, the point where only a single peak is produced, should be the op- timal volume. Figure 5.3 shows the interesting results. The curves for the four pump speed settings all follow nearly the same line. This is evidence that mixing in a SBSR is independent of flow rate and residence time. The Optimal volume was about 45 uL. In practice, however, sample volumes up to 65 uL did not give perceptible double peaks. Four glucose standards were each sampled for five 157 ,\ 600 -- ..l 3 m a. x O 0) a 400 ~— C Q) m 0 3 __ Q) a) g 200 -~ 3 B >' _- 4 I l | L l I O T l I l I I I 0 200 400 600 Soaufle Vothe (0L) Figure 5.3. The degree of incomplete mixing at various sample volumes. Four pump speeds were used: 8.3, 12.3, 25.0, and 33.3 uL/sec. 158 seconds at a flow rate of 12.3 uL/sec. The residence time inthe enzyme reactor was about 6 seconds. The instrumental output is shown in Figure 5.4. The average relative standard deviation of the results was less than 2%. A calibration curve is shown in Figure 5.5. The results were very good; the correlation coefficient was 0.99994. The use of this method for the determination of glucose looks promising. With further optimization, the sample throughput should be increased from the current value of 60 samples/hour. Many other substrates could be determined in this way, using other enzymes. 159 0.3 -T' Absorbance fi- Ill Thne UnM) Figure 5.4. Flow-injection instrumental response. The four glucose concentrations were: 0.44 (A), 0.88 (B), 1.33 (C), and 2.22 (D) mM. 160 Mean Absorbance [Glucose] OnM) Figure 5.5. Calibration curve for the flow—injection determination of glucose. CHAPTER 6 Future Plans The use of open—tubular immobilized enzyme reactors in flowing systems continues to grow, and more research is needed to characterize such systems. A few of my ideas for future research in this area are presented below. I plan to continue some portions of my doctoral research, and, hopefully, others at Michigan State will spend their time working with immobilized enzymes. 1) Fast immobilization procedure for nylon. The current procedure for bonding enzymes to nylon takes about 1.5 days. The duration of the procedure might be shortened without adversely affecting the enzyme atachment. A three-hour or less immobilization treatment would be advantageous for undergraduate research and teaching laboratory experiments. It is known that most of the enzyme bonds to the activiated nylon within the first ten minutes of reaction. Thus, the reduction of the procedure in terms of time is feasible. 2) Properties of bi-enzyme systems. Many determinations involve two or more enzyme-catalyzed reactions. It would be helpful to know how each enzyme affects the attachment of the other. The studies presented in Chapter 4, Section E, suggested that the secondary enzyme, catalase, did not significantly disturb the bonding of the primary enzyme, 161 162 glucose oxidase. The properties of bi-enzyme systems should be studied in more detail. 3) Properties of immobilized enzymes in continuous flow systems. Immobilized glucose oxidase reactors have been fully characterized for use in a continuous flow instru- ment. Other enzymes could be incorporated into the CF manifold, and their properties could be studied. After several enzyme systems have been tested in this fashion, general conclusions about the characteristics of OTIMERs in continuous flow systems can be drawn. 4) Fundamental studies of the influence of molecular diffusion on immobilized enzyme kinetics. The stopped— flow instrument described in Chapter 3 may prove to be a very powerful tool in probing the interaction of diffusion and immobilized enzyme kinetics. By changing the reactor parameters and measuring the reaction rates, a mathematical description of such systems should be possible to derive. Subsequently, the inherent enzyme properties could be calculated from the measured apparent properties. This project deserves a lot of attention. 5) Immobilized enzyme bead string reactor. Very fast and sensitive clinical determinations may be possible with the use of enzyme-coated glass beads packed into a Teflon tube. Flow-injection determinations using such reactors show 163 great promise and can be expanded for use with highly active enzymes, other than glucose oxidase. The experi- mental design would be similar to that described in Chapter 5. APPENDICES 164 APPENDIX A Derivation of Diffusion-control Rate Equation Model: 1) Circular tube of radius, r 2) Substrate reacts at the reactor wall very fast to form product. 3) A radial concentration gradient in substrate and product is established. 4) Substrate diffuses toward reactor wall. 5) Product diffusing away from reactor wall. Initial condition: Boundary condition: Diffusion equation: Final result: [S] = [S] for 00. 3[S]/3t = D(32[S]/3x2 + (3[S]/8x)/x) 3 [P] = [S]b(1-4Z (eXp(-Dtj:/r2))/j:) S=1 165 APPENDIX B Derivation of the Absorbance of a Solution with a Radial Concentration Gradient Model: 1) Reaction solution is axially uniform. 2) The concentration gradient is radially symmetric. 3) The concentration gradient does not change with time (steady-state conditions). 4) The gradient is linear (d[P]/dx = constant). 5) The product is the absorbing species. 6) Light traverses the reactor parallel to the axis. Considering a slice of solution from the center of the tube to the reactor wall: Absorbance = eL[P] o r = - — + [P] (I ([P]w ([P]W [P]b)x/o)dx é [P]bdx)/r where x is the distance from the center of the reactor, 0 is the diffusion layer thickness, and where r is the reactor radius. Integrating the above equation and combing terms: [P] = o/2r([P]w-[P]b) + [P]b Thus, Absorbance = eL([P]b + o/2r([P]w-[P]b)) Case 1: No enzyme reaction (common absorbance measurement) [P]w=[P]b and Absorbance = eL[P]b Case 2: Start of enzyme reaction [P]b = 0 and Absorbance = eL[P]w(o/2r) L I ST OF REFERENCES 166 LIST OF REFERENCES 1. W.B. Furman, editor, Continuous Flow Analysis: Theory and Practice, Marcel Dekker, NY (1976). 2. L.R. Snyder, J. Levine, R. Stay, and A. Conetta, Anal. Chem. 48, 942A (1976). 3. L.R. Snyder, J. Chromatog. 125, 287 (1976). 4. J. Ruzicka and E.H. Hansen, Chemtech 9, 756 (1979). 5. J. Ruzicka and E.H. Hansen, Flow Injection Analysis, John Wiley and Sons, NY (1981). 6. D. Betteridge, Anal. Chem. 59, 832A (1978). 7. B. Chance, J. Franklin Inst. 229, 455 (1940). 8. S.R. Crouch, F.J. Holler, P.K. Notz, and P.M. Beck- with, Appl. Spec. Rev. 13, 165 (1977). 9. E.T. Gray and H.J. Workman, J. Chem. Ed. 51, 752 (1980). 10. P.W. Carr and L.D. Bowers, Immobilized Enzymes in Analytical and Clinical Chemistry, John Wiley and Sons, NY (1980). 11. M.D. Trevan, Immobilized Enzymes, John Wiley and Sons, NY (1980). 12. I. Chibata, editor, Immobilized Enzymes, Research and Development, John Wiley and Sons, NY (1980). 13. T.M.S. Chang, Immobilized Enzymes and Proteins, Volumes 1 and 2, Plenum Press, NY (1977) 14. L.B. Wingard, E. Katchalski-Katzir, and L. Goldstein, editors, Applied Biochemistry and Bioengineering, Volume 1, Academic Press, NY (1976). 15. H.H. Weetall, editor, Immobilized Enzymes, Antigens, Antibodies, and Peptides, Marcel Dekker, NY (1975). 167 16. R.A. Messing, editor, Immobilized Enzymes for Indus- trial Reactors, Academic Press, NY (1975). 17. M. Salmona, C. Saronio, and S. Garattini, Insolu- bilized Enzymes, Raven Press, NY (1974). 18. O. Zaborsky, Immobilized Enzymes, CRC Press, Cleveland (1973). 19. S. Aiba, A.E. Humphrey, and N.F. Millis, Biochemical Engineering, Academic Press, NY (1973), pp 393-416. 20. L.B. Wingard, Enzyme Engineering, John Wiley and Sons, NY (1972). Pp 177-400. 21. L.D. Bowers, Trends Anal. Chem. 1, 191 (1982). 22. L.D. Bowers and P.W. Carr, Adv. Biochem. Eng. 15, 89 (1980). 23. L.D. Bowers and P.W. Carr, Anal. Chem. 45, 544A (1976). 24. H.H. Weetall, Anal. Chem. 45,602A (1974). 25. L.B. Wingard, Adv. Biochem. Eng. g, 1 (1972). 26. G.G. Birch, N. Blakebrough, and K.J. Parker, Enzymes and Food Processing, Applied Science Publishers, London (1981). 27. L.B. Wingard, E. Katchalski—Katzir, and L. Goldstein, Applied Biochemistry and Bioengineering, Volume 2, Academic Press, NY (1979). 28. S. Kindel, Technology 1, 62 (1981). 29. K.J. Skinner, Chem. Eng. News, August 18,1975, 22. 30. C.R. Martin, Trends Anal. Chem. 1, 175 (1982). 31. G.G. Guilbault and M.H. Sadar, Acc. Chem. Res. 15, 344 (1979). 32. G.G. Guilbault and M.H. Sadar, Proc. Analyt. Div. Chem. Soc., October 1977, 302. 33. W.J. Blaedel and R.C. Engstrom, Anal. Chem. 55, 1691 (1980). 168 34. F. Scheller and D. Pfeiffer, Anal. Chim. Acta 117, 383 (1980). 35. W.J. Blaedel and J. Wang, Anal. Chem. 55, 1426 (1980). 36. R.M. Ianniello and A.M. Yacynych, Anal. Chim. Acta 131, 123 (1981). 37. R. Renneberg, D. Pfeiffer, F. Scheller, and M. Janchen, Anal. Chim. Acta 134, 359 (1982). 38. E.L. Gulberg and G.D. Christian, Anal. Chim. Acta 123, 125 (1981). 39. W. Hinsch and P.V. Sundaram, Fres. Z. Anal. Chem. 309, 25 (1981). 40. M. Mascini and G. Palleschi, Anal. Chim. Acta 136, 69 (1982). 41. W. Hinsch, W.—D. Edersbach, and P.V. Sundaram, Clin. Chim. Acta 104, 95 (1980). 42. W. Hinsch and P.V. Sundaram, Clin. Chim. Acta 104, 87 (1980). 43. B. Danielsson, B. Mattiasson, and K. Mosbach, Pure Appl. Chem. 54, 1443 (1979). 44. C. Spink, CRC Critical Rev. Anal. Chem. 5, 1 (1980). 45. R.S. Schifreen, D.A. Hanna, L.D. Bowers, and P.W. Carr, Anal. Chem. 45, 1929 (1977). 46. A.O. Lindberg, Anal. Chim. Acta 131, 133 (1981). 47. A.S. Attiyat and G.D. Christian 105, 154 (1980). 48. C.-H. Kiang, S.S. Kuan, and G.G. Guilbault, Anal. Chem. 55, 1319 (1978). 49. J.-C. Kaun, S.S. Kuan, and G.G. Guilbault, Anal. Chim. Acta 100, 229 (1978). 50. P.V. Sundaram and S. Jarayaman, Clin. Chim. Acta 54, 309 (1979). 51. O.R. Zaborsky and J. Ogletree, Biochem. Biophys. Res. Comm. 5;, 210 (1974). 169 52. D.J. Inmann and W.E. Hornby, Biochem. J. 129, 255 (1972). 53. D.L. Morris, J. Campbell, and W.E. Hornby, Biochem. J. 147, 539 (1975). 54. P.V. Sundaram, Biochem. J. 183, 445 (1979). 55. P.V. Sundaram and D.K. Apps, Biochem. J. 161, 441 (1977). 56. P.V. Sundaram and M.P. Igloi, Clin. Chim. Acta 54, 295 (1979). 57. F.N. Onyezili and A.C. Onitiri, Anal. Biochem. 113, 203 (1981). 58. G.A. Noy, A. Buckle, and K. Alberti, Clin. Chim. Acta 55, 135 (1978). 59. F.N. Onyezili and A.C. Onitiri, Anal. Biochem. 117, 121 (1981). 60. J. Campbell, W.E. Hornby, and D.L. Morris, Biochim. Biophys. Acta 384, 307 (1975). 61. V. Toome, S. DeBernardo, K. Manhart, and M. Weigele, Anal. Lett. 1, 437 (1974). 62. W.—T. Law and S.R. Crouch, Anal. Lett. 45, 1115 (1980). 63. O. Warburg and W. Christian, Biochem. Z. 310, 384 (1941). 64. P.D. Boyer, editor, The Enzymes, Academic Press, NY. 65. M. Dixon and E.C. Webb, Enzymes, Academic Press, NY (1964). 66. J.B. Sumner and G.F. Somers, Chemistyy and Methods 55 Enzymes, Academic Press, NY (1953). 67. K.J. Laidler and P.S. Bunting, The Chemical Kinetics 5: Enzyme Action, Clarendon Press, Oxford, England (1973), p 398. 68. E.W. Wharton, E.M. Crook, and K. Brocklehurst, Eur. J. Biochem. 5, 572 (1968). 170 69. W.E. Hornby, M.D. Lilly, and E.M. Crook, Biochem. J. 55, 420 (1966). 70. J.-M. Engasser and C. Horvath, Applied Biochemistry and Bioengineering, Volume 1, Academic Press, NY (1976), pp 146-149. 71. G.P. Royer, Immobilized Enzymesy Antigens, Antibodies, and Peptides, Marcel Dekker, NY (1975), pp 66-68. 72. J.-M. Engasser and P.Hisland, Biochem. J. 173, 314 (1978). 73. J.-M. Engasser, P.R. Coulet, and D.C. Gautheron, J. Biol. Chem. 252, 7919 (1977). 74. I.C. Cho and H. Swaisgood, Biochim. Biophys. Acta 334, 243 (1974). 75. P. Nitisewojo and H.O. Hultin, Eur. J. Biochem. 51, 87 (1976). 76. N.L. Smith and H.M. Lenhoff, Anal. Biochem. 54, 392 (1974). 77. K.J. Laidler and P.S. Bunting, Meth. Enzymol. 54, 227 (1980). 78. T. Kobayashi and K.J. Laidler, Biotech. Bioeng. 45, 99 (1974). 79. N.J. Daka and K.J. Laidler, Can. J. Biochem. 55, 774 (1978). 80. N.J. Daka and K.J. Laidler, Biochim. Biophys. Acta 612, 305 (1980). 81. Advances 45 Automated Analysis, Technicon International Congress 1976, Volume 1, Mediad Inc., Tarrytown, NY (1977). 82. L.R. Snyder and H.J. Adler, Anal. Chem. 45, 1017 (1976). 83. L.R. Snyder and H.J. Adler, Anal. Chem. 45, 1022 (1976). 84. S. Morgenstern, Advances 45 Automated Analysis, Volume 1, Therman Assoc., Miami (1971), p 85. 171 85. M.D. Joseph, D.J. Kasprazak, and S.R. Crouch, Clin. Chem. 55, 1033 (1977). 86. D.R. Christmann, S.R. Crouch, and A. Timnick, Anal. Chem. 55, 276 (1981). 87. P.D. Boyer, editor, The Enzymes, Volume XI, Academic Press, NY (1975), pp 289- 292. 88. J.R. Florini and C.S. Vestling, Biochim. Biophys. Acta 251 575 (1957). 89. H.L. Pardue, Advances 55 Analytical Chemistry and Instrumentation, Volume 7, John Wiley and Sons, NY (1968), pp 141-207. 90. P. Trinder, Ann. Clin. Biochem. 5, 24 (1969). 91. D. Barham and P. Trinder, Analyst 55, 142 (1972). 92. N. W. Tietz, Fundamentals of Clinical Chemistry, W. B. Saunders, Philadelphia (1976),— p 1213. 93. R.H. White-Stevens, Clin. Chem. 55, 578 (1982). 94. J.A. Lott and K. Turner, Clin. Chem. ‘55, 1754 (1975). 95. N. W. Tietz, editor, Fundamentals of Clinical Chemistry, W.B. Saunders, Philadelphia (1976), pp 234- 263. 96. F. Gram, J.A. Hveding, and A. Reine, Acta Chem. Scand. 55, 616 (1973). 97. E.N. Drake and C.E. Brown, J. Chem. Ed. 54, 124 (1977). 98. J. Okuda and I. Miwa, Anal. Biochem. 55, 387 (1971). 99. A.S. Kaston and R. Brandt, Anal. Biochem. 5, 461 (1963). 100. P.V. Sundaram, B. Blumemberg, and W. Hinsch, Clin. Chem. 55 1436 (1979). 101. L.P. Leon, M. Sansur, L.R. Snyder, and C. Horvath, Clin. Chem. 55, 1556 (1977). 102. L. Gorton and K.A. Bhatti, Anal. Chim. Acta 105, 43 (1979). 172 103. Sigma Chemical Company Catalog, 1982 edition. 104. D.L. Williams, A.R. Doig, and A. Korosi, Anal. Chem. 45, 118 (1970). 105. M. Koyama, Y. Sato, M. Aizawa, and S. Suzuki, Anal. Chim. Acta 116, 307 (1980). 106. T. Osaki, Biotech. Bioeng. 55, 1287 (1980). 107. G.J. Lubrano and G.G. Guilbault, Anal. Chim. Acta 55, 229 (1978). 108. B. Watson, D.N. Stifel, and F.E. Semersky, Anal. Chim. Acta 106, 233 (1979). 109. G. Sittampalam and G.S. Wilson, J. Chem. Ed. 55, 70 (1982). 110. D.C. Williams, G.F. Huff, and W.R. Seitz, Clin. Chem. gg, 372 (1976). 111. J.N. Miller, B.F. Rocks, and D.T. Burns, Anal. Chim. Acta 55, 353 (1977). 112. L.B. McGowan, B.J. Kirst, and G.L. LaRowe, Anal. Chim. Acta 117, 363 (1980). - 113. F. Heinz and T.W. Beuhausen, J. Clin. Chem. Clin. Biochem. ‘55, 977 (1981). 114. R. Chirillo, G. Caenaro, B. Pavan, and A. Pin, Clin. Chem. 55, 1744 (1979). 115. J. Ziegenhorn, U. Neumann, A. Hagen, W. Bablok, and K. Stinshoff, J. Clin. Chem. Clin. Biochem. 45, 13 (1977). 116. American Research Products Company, bulletin. 117. L.G. Morin and J. Prox, Clin. Chem. 55, 959 (1973). 118. R.H. White-Stevens and L.R. Stover, Clin. Chem. 55, 589 (1982). 119. N. Gochman and J.M. Schmitz, Clin. Chem. 45, 943 (1972). 173 120. L.P. Leon, S. Narayanan, R. Dellenbach, and C. Horvath, Clin. Chem. 55, 1017 (1976). 121. A.T. Romano, Clin. Chem. 45, 1152 (1973). 122. C. Matsubura, K. Ishii, and K. Takamura, Microchem. J. 55, 242 (1981). 123. K. Tamaoku, Y. Murao, K. Akiura, and Y. Ohkura, Anal. Chim. Acta 136, 121 (1982). 124. I. Seigo, Eisei Kensa 55, 760 (1979). 125. J.D. Artiss, R.J. Thibert, J.M. McIntosh, and B. Zak, Microchem. J. 55, 487 (1981). 126. D. Keilin and E.F. Hartree, Biochem. 55, 331 (1952). 127. D. Keilin and E.F. Hartree, Biochem. 45, 221 (1948). 128. R. Bentley, Meth. Enzymol. l, 340 (1955). 129. V. Linek, P. Benes, J. Sinkule, O. Holecek, and V. Maly, Biotech. Bioeng. ‘55, 2515 (1980). 130. S. Gondo, T. Osaki, and M. Morishita, J. Mol. Cat. 45, 365 (1981). 131. P.F. Greenfield, J.R. Kittrell, and R.L. Laurence, Anal. Biochem. 55, 109 (1975). 132. B.C. Saunders, A.G. Holmes-Siedle, and B.P. Stark, Peroxidase, Butterworths, Washington, D.C. (1964). 133. A.C. Maehly, Meth. Enzymol. 5, 801 (1955). 134. I.M. Kolthoff, E.B. Sandell, E.J. Meehan, and S. Bruchenstein, Quantitative Chemical Analysis, The MacMillan Co., NY (1969), p 854. 135. C.J. Patton, M. Rabb, and S.R. Crouch, Anal. Chem. 54, 1113 (1982). 136. S. Klose, M. Sholtz, E. Munz, and R. Portenhauser, Clin. Chem. 54, 250 (1978). 137. F. Zoppi and D. Fenili, Clin. Chem. 55, 1229 (1980). 174 138. N.W. Teitz, editor, Fundamentals 54 Clinical Chemistry, W.B. Saunders, Philadelphia (1976) p 254. 139. Y. Malpiece, M. Sharan, J.-N. Barbotin, P. Personne, and D. Thomas, J. Biol. Chem. 225, 6883 (1980). 140. C. Horvath, B.A. Solomon, and J.-M. Engasser, Ind. Eng. Chem. Fundam. 45, 431 (1973). 141. R.Q. Thompson, C.J. Patton, and S.R. Crouch, Clin. Chem. 55, 556 (1982). 142. G.R. Schonbaum and B. Chance, The Enzymes, Volume 13, Academic Press, NY (1976), pp 363-408. 143. D. Keilin and P. Nicholls, Biochim. Biophys. Acta 55, 302 (1958). 144. P. Nicholls and G.R. Schonbaum, The Enzymes, Volume 8, Academic Press, NY (1963), pp 147-225. 145. M. Dixon, Biochem. J. 55, 170 (1953). 146. I.H. Segel, Enzyme Kinetics, John Wiley and Sons, NY (1975). PP 170-203. 147. A. lab and H.A. Motolla, Clin. Chem. 55, 195 (1981). 148. F.I. Onuska, M.E. Comba, T. Bistricki, and R.J. Wilkinson, J. Chrom. 142, 117 (1977). 149. H.H. Weetall, Science 166, 616 (1969). 150. P.J. Robinson, P. Dunnill, and M.D. Lilly, Biochim. Biophys. Acta 242, 659 (1971). 151. W.E. van der Linden, Trends Anal. Chem. 4, 188 (1982). 152. J.M. Reijn, W.E. van der Linden, and H. Poppe, Anal. Chim. Acta 123, 229 (1981).