Summary
Hydrolytic reactions of oligopeptide 4-nitroanilides catalyzed by human-α-thrombin, human activated protein C and human factor Xa were studied at pH 8.0–8.4 and 25.0 ± 0.1 °C by the progress curve method and individual rate constants were calculated mostly within 10% internal error using DYNAFITV. A systematic strategy has been developed for fitting a three-step consecutive mechanism to eighteen hundred to six thousand time-course data points polled from two to four independent kinetic experiments. Enzyme and substrate concentrations were also calculated. Individual rate constants well reproduce published values obtained under comparable conditions and the Michaelis-Menten kinetic parameters calculated from these elementary rate constants are also within reasonable limits of published values. For comparison, the integrated Michaelis-Menten equation was also fitted to data from twelve sets. Both the kcat and kcat/Km values are within 15% agreement with those calculated using the elementary rate constants obtained with DYNAFITV. Rate constants for the second and third consecutive steps are within 3–4 fold indicating that both determine the overall rate. The Factor Xa-catalyzed hydrolysis of Nα-Z-D-Arg-Gly-Arg-pNA·2HCl at pH 8.4 in a series of buffers containing increasing fractions of deuterium at 25.0 ± 0.1 °C shows a very strong dependence of k3 and a moderate dependence of k2 on D content in the buffer: the fractionation factors are: 0.49 ± 0.03 for K1, 0.70 ± 0.05 for k2, and (0.32 ± 0.03)2 for k3.
Keywords: Enzyme kinetics, progress curve, elementary rate constants, blood cascade enzymes, solvent isotope effects, proton inventories
Introduction
Enzyme kinetic studies in the last decades have increasingly taken advantage of the revolution in computations. Software packages empower the user to fit a set of equations describing a reaction mechanism of multiple steps to a very large number of data points, even from a single time-dependent change in concentration. One of the most advanced and well-tested software for a range of enzyme kinetic calculations is DYNAFIT[1], incorporating the pioneering work of Garfinkel[2] and the experiences from programs KINSIM[3] and FITSIM.[4] These programs accept a given reaction scheme with elementary steps as input, and solve the requisite differential equations, without specific assumptions, by numerical integration. Two of the four novel features of DYNAFIT over its predecessors are that 1) it accepts initial concentrations of enzyme and substrate as fitting parameters after an initial guess, and 2) it can also treat as adjustable parameters molar absorptivity and baseline signal.
The two major gains of using enzyme kinetic software are: 1) the opportunity to account for increasingly complex multiple-step reactions at the level of elementary steps because all information is obtained from the entire reaction carried to completion and 2) much reduced need for material with respect to the traditional initial rate method pioneered by Michaelis and Menten. This application of the progress curve method is different in its goal from those used for determining Michaelis-Menten parameters by fitting the integrated Michaelis-Menten equation to a single and complete kinetic run.[5;6] This paper aims at developing methods and the conditions for obtaining individual rate constants, as accurate as possible, that can then be used to calculate Michaelis-Menten parameters, which agree with those determined by other methods.
Our interest in using DYNAFIT has been the elucidation of elementary rate constants in two and three-step consecutive reactions as exemplified by the serine hydrolase-catalyzed hydrolysis of chromogenic substrates, peptides and proteins. An accurate knowledge of the mechanism of an enzyme-catalyzed reaction must be based on competent kinetic information at the level of individual rate constants reporting on specific elementary steps. The slowest elementary step(s) determines the rate of a multi-step process.[7] Once the elementary rate constants are calculated, the isotopic sensitivity of each step can be studied directly. Such direct determination of intrinsic isotope effects is indeed a great gain in enzyme mechanistic studies. Characterization of the transition state(s) of the rate-determining step(s) is at the heart of enzyme mechanistic studies and it is a prerequisite to rational drug design of transition state analog inhibitors.
It is generally expected that the rate-determining step in enzyme-catalyzed ester hydrolysis is deacylation, while that of amide hydrolysis is predominantly acylation.[7;8] The latter premise is often on soft grounds.[8] An elucidation of some of the elementary steps, possibly acylation and deacylation, often requires extensive kinetic studies including trapping of the acyl enzyme or inducing a change in the rate-determining step with changing solvent viscosity,[9–13] salt or temperature.[14;15] The extensive experimental work followed by multidimensional analysis of the results can be abrogated if a reliable computational method can be used for the analysis of a well-designed single determination of the entire reaction progress. However, the uniqueness of calculated rate constants and the robustness and limitations of such a method needs to be proven: Fersht[7] called attention to a caveat that different sets of rate constants may satisfy multi-parameter kinetic equations. The validation of uniqueness of calculated rate constants is thus essential.
Our initial trials using DYNAFIT for fitting a three-step reaction scheme indicated the power of the program and held premise for obtaining reliable individual rate constants of a consecutive reaction scheme, provided a circumspect strategy for analysis is developed. In this paper we report the development of such a strategy. We validated its results against experimentally-determined enzyme kinetic rate constants in the thrombin-catalyzed hydrolysis of H-D-Phe-Pip-Arg-4-nitroanilide.HCl (pNA) (S-2238).[11] The elementary rate constants obtained are in good agreement with the published values. We determined elementary rate constants for three other serine protease-catalyzed reactions that have not before been determined. We probed the robustness of elementary rate constants by varying concentrations of thrombin and activated protein C (APC) and at varying levels of saturation of thrombin and APC, with a chromogenic substrate of medium efficiency for each. The value of the rate constants for the second and third step of the reactions were generally within 3–4-fold. The Michaelis-Menten constants obtained from a few progress curves reproduced the Literature values reasonably well. We then applied the validated method for exploring full and partial solvent isotope effects on the individual rate constants for the Factor Xa-catalyzed hydrolysis of N-α-Z-D-Arg-Gly-Arg-pNA·2HCl (S-2765). The kinetic rate constants calculated using DYNAFIT are nearly equal for the second and third consecutive steps, corresponding either to the formation and breakdown of the tetrahedral intermediate or to acylation and deacylation, in buffered water. A gradual increase in D2O content in the buffer resulted in reduced rate constants, mostly for the third consecutive step, moderately for the binding step and only slightly affected the second consecutive step.
Materials and Methods
Materials
Anhydrous dimethyl sulfoxide (DMSO), heavy water with 99.9 % deuterium content and anhydrous methanol, were purchased from Aldrich Chemical Co. All buffer salts were reagent grade and were purchased from either Aldrich, Fisher, or Sigma Chemical Co. H-D-Phe-Pip-Arg-4-nitroanilide.HCl (pNA) (S-2238) 99% (TLC), and Pyro-Glu-L-Pro-L-Arg-pNA.HCl (S-2366) 99% (TLC), were purchased from Diapharma Group Inc.. Human α-thrombin, MM 36,500 d, 3010 NIH u/mg activity in pH 6.5, 0.05 M sodium citrate buffer, 0.2 M NaCl, 0.1% PEG-8000 and Human activated protein C, (APC) MM 56,000 d, 1.19 mg/mL in pH 7.4, 0.02 M Tris-HCl buffer, 0.1 M NaCl M were purchased from Enzyme Research Laboratories. Rabbit thrombomodulin (TM), MM 74,000, 4.7 mg/ml in TBS buffer, 0.05%PDOC, 0.02% NaN3 was purchased from Haematologic Technologies Inc..
Instruments
Spectroscopic measurements were performed with a Perkin-Elmer Lambda 6 UV-Vis Spectrophotometer connected to a PC. The temperature was monitored using a temperature probe connected to a digital readout device. Either a Neslab RTE-4 or a Lauda 20 circulating water bath was used for temperature control. Positive displacement Gilson and Rainin Microman pipettes with plastic tips were used for the delivery of enzyme solution, substrate solution and organic solvent.
Solutions
Buffers were prepared by weight from Tris-base and Tris-HCl at 0.02 M or 0.005 M Tris, 0.3 M NaCl, 5 mM CaCl2, 0.1% PEG4000 at pH 8.0 for the experiments with thrombin and APC. The experimental details of the proton inventory studies with of the FXa-catalyzed hydrolysis of N-α-Z-D-Arg-Gly-Arg-pNA·2HCl have been described previously.[16]
Stock solutions for enzymes were prepared by dilution with Tris buffer to 2.90 × 10−7 M thrombin, 7.70 × 10−7 M and 20.8 × 10−7 M APC. Substrate solutions were prepared in DMSO at 2.5 × 10−3 M S-2238 and 0.01 M S-2366.
Enzyme active-site assays
Thrombin concentrations were determined from 20 μL of a fivefold dilution of the thrombin stock solution, which was added to 960 μL buffer in a cuvette. The solution was incubated in pH 8.0, 0.02 M Tris buffer, 0.3 M NaCl at 25.0 ± 0.1 °C for 15 minutes. Twenty microliters of 2.5 ×10−3 M S-2238 was added to the solution. Initial rates were measured at 405 nm after quick mixing of the solution. The active-site concentrations were calculated from the velocity measured for a minute and divided by kcat = 95 s−1.[17;18] The kcat value was measured under the conditions of these experiments and was based on active-site titration of thrombin with 4-methyl umbelliferyl guanidinobenzoate HCl (MUGB). APC activity was assayed using S-2366 and kcat = 160 s−1.[19] There was no background hydrolysis of the substrates in the absence of enzyme and the slopes were proportional to enzyme concentration within 90 % of the value given by the provider.
Kinetic methods
Fifteen microliters of a 1.49 ×10−7 M thrombin solution was added to 965 μL buffer, pH 8.0, 0.005 M Tris, NaCl 0.2 M, 0.1% PEG4000, and the mixture was incubated for 15 min. Twenty microliters of 2.5 × 10−3 M S-2238 stock solution in DMSO, to give 5 × 10−5 M in the cell, was added to initiate the reaction. Completion of a full curve was in 8 minutes. The protocol was nearly identical for the hydrolytic reactions of S-2366 catalyzed by thrombin and APC (Results).
For the proton-inventory study,[16] nine buffers pH 8.4 Tris (and equivalent pL), 0.02 M Tris, 0.3 M NaCl, 5 × 10−3 M CaCl2, 0.1% PEG-4000 were prepared with deuterium oxide composition ranging from atom fraction of D, n, 0 to 99. Ten microliters of 7.23 × 10−7 M FXa was added to 480 μL of buffer in a cuvette, to give a final concentration of 1.45 × 10−8 M. The cuvette was incubated at 25.0 ± 0.1 °C for 15 min. Ten microliters of 0.05 M S-2765 in DMSO was added to achieve 10−3 M in the incubated FXa solution. The measurement was conducted at 445 nm for about 40 min.
Determination of the molar absorptivity of 4-nitroaniline at 405 nm, 445 and 458nm
The molar absorptivity of 4-nitroaniline was determined for the calculation of precise rate constants. The stock solution of 4-nitroaniline was prepared at 9.05 × 10−3 M in DMSO for the measurement at 445 nm and 458 nm, and at 4.53 × 10−4 M for 405 nm. pH 8.4 Tris buffer in 460 μL volume was incubated in a cuvette at 25.0 ± 0.1 °C for 15 min. 4-Nitroaniline stock solution was added to the buffer in 10 μL increments of 60 μL in total. The absorbance was recorded after each addition. Calibration curves of absorbance versus concentration were constructed.
Data Analysis
The kinetic data were processed using DYNAFITV, which integrates the requisite simultaneous first-order, nonlinear, ordinary differential equations numerically, using a modification of the multidimensional Newton-Raphson method.[1] Regression analysis in DYNAFIT is performed by Reich's variation of the Levenberg-Marquardt least-squares fitting algorithm. Convergence criteria are multiple in DYNAFIT: The Marquardt parameter for individual parameters should be met first. This was typically met within 100 iterations and 50 subiterations. Standard errors of the individual parameters were computed from the square roots of diagonal elements of the final variance-covariance matrix. The protocol that gave individual rate constants and Michaelis-Menten parameters calculated from them, in agreement with experimentally determined individual rate constants and experimentally determined (or published) Michaelis-Menten parameters, was as follows: Fitting started with a simple two-step consecutive mechanism:
The initial guesses were based on published values of kcat, which was used for the estimation of k2 and the concentration of enzymes and substrates were as determined independently. Alternatively, an initial kcat was calculated from the first initial slope of the progress curve (while, [S] > 4Km), Vmax, and the enzyme concentration, [E], and then used as an estimate of k2. The rate constants obtained and offset values, which incur from initial imbalance in absorbance values between reference and sample cells, were then the input for a refined calculation in which the enzyme and substrate concentrations were also calculated. The calculated values of [E] and [S] were within a few percent of the values determined independently in nearly all cases, which is a good indication of program performance when fitting six parameters. In the third round, k1 k−1 (K1 = k1/k−1), k2 and k3 were calculated for individual runs of 900–3,000 absorption-time data pairs based on a three-step consecutive mechanism:
Two, three or four parallel runs were then combined in a "global fit" using individual [E] and [S] values calculated and the values obtained from fits of individual kinetic runs used as initial estimates in the final calculation of the rate constants within the 95% confidence limit. From the three-step reaction scheme, kcat = k2k3 / (k2 + k3) and kcat/Km = k1k2 / (k-1 + k2) were calculated for verification purposes with measured or published values.
Michaelis-Menten parameters were also calculated by fitting the integrated Michaelis-Menten equation:[16]
| (1) |
where A, Ao and Ainf are the absorbance values at time t, 0 and infinity, respectively; ε is the molar absorbtivity, [E] is the enzyme concentration, and kcat and Km are the Michaelis-Menten parameters; to the data from the thrombin-catalyzed hydrolysis of S-2238 and the Factor Xa-catalyzed hydrolysis of S-2675 obtained in nine mixtures of water and heavy water buffered at pH 8.2. Good initial estimates of the three parameters are required in the input for the nonlinear fitting here as well. The kcat values agreed within 12 % and the kcat/Km values agreed within 15 % between the two methods. These and other data reductions using predefined and custom-defined equations were performed using the GraFit 3.0 or GraFit 5.0 software.[20]
Results and Discussion
Molar absorptivities
The molar absorptivities for 4-nitroaniline at pH 8.4, 25.0 ± 0.1 °C and above the 0.998 confidence level are 9,420 ± 100, 1,245 ± 15 and 346 ± 8 OD cm−1 M−1 at 405, 445 and 458 nm, respectively. A very careful measurement[21] at 405 nm resulted 10,040 OD cm−1 M−1 in pH 7.8 Tris buffer without PEG4000, which is 6% higher than our value.
Reproducibility of elementary rate constants and Michaelis-Menten parameters calculated from them
As with other nonlinear least squares algorithms, there is a need for good estimates for the parameters to guide the calculation toward convergence at the global minimum, i.e. at correct values. [22;23] This of course is not a serious flaw because one always has access to good estimates as described above and below in this paper. The reproducibility of individual rate constants from individual runs under the same condition, as shown in Tables 1, 2 and 3, is generally very good. The least-squares surface is relatively flat in these calculations, which requires a few repeated restarts of the algorithm after it converged to the Levenberg-Marquardt criteria, in order to bring down the errors. However, the nominal values of the parameters stay stable during these refinements. The calculations obtained from combining three parallel runs provide elementary rate constants with standard deviations below 10% for most cases, and in many instances even below a percent. In Table 1, the standard error for each parameter is reported for each individual run. The benefit of using all data in a global fit is clearly discernable when comparing the standard errors of rate constants obtained from individual runs with the corresponding data from the global fit for any of the elementary rate constants. The same trends are observable for the data presented in Tables 2–5, although the standard errors are not all shown. Typically, the errors in individual rate constants are typically 10–20%,while the global fit gives up to 10% error in the elementary rate constants.
Table 1.
Rate constants calculated with the two- and three-step model for individual and combined runs for the 1.49 × 10−9 human thrombin-catalyzed hydrolysis of 4 × 10−5 MS-2238 at pH 8.0, 0.005 M Tris buffer, 0.2 M NaCl, 0.1% PEG4000 and 25.0 ± 0.1 °C
| Parameter | Individual trials, three steps | Global fit 1+2+3 | Ref.[11] | Diapharma a | |||
|---|---|---|---|---|---|---|---|
| 1 | 2 | 3 | Two steps | Three steps | |||
| k1, μM−1s−1 | 66.0 ± 8.7 | 67.3 ± 4.0 | 64.4. ± 7 9 | 62.0 ± 5.0 | 64.2 ± 1.8 | 103 ± 28 | |
| k−1, s−1 | 448 ± 73 | 423 ± 77 | 417 ± 81 | 249 ± 33 | 420 ± 26 | 280 ± 29 | |
| k2, s−1 | 187 ± 14 | 183 ± 24 | 183 ± 17 | 150.00 ± 0.02 | 185 ± 9 | 101 ± 10 | |
| k3, s−1 | 400 ± 65 | 409 ± 120 | 414 ± 90 | - | 404 ± 43 | 315 ± 63 | |
| kcat, s−1 | 127 ± 23 | 126 ± 40 | 127 ± 30 | 150.00 ± 0.02 | 127.0 ± 16.0 | 77 ± 3 | 180, 95b |
| Km, μM | 6.5 ± 1.9 | 6.2 ± 2.4 | 6.4 ± 2.1 | 6.43 ± 0.84 | 6.47 ± 1.0 | 2.8 ± 0.3 | 7.0 |
| kcat/Km, M−1s−1 | 19.4 ± 4.1 | 20.0 ± 4.4 | 19.7 ± 4.5 | 23.3 ± 3.1 | 19.6 ± 1.6 | 27.5 | 26.7 |
Table 2.
Rate constants calculated from individual and combined runs for the human thrombin-catalyzed S-2366 hydrolysis (see Figure 2) in the presence of thrombomodulin in pH 8.0, 0.02 M Tris buffer, 0.3 M NaCl, 0.1% PG4000 and 25.0 ± 0.1 °C
| Parameter | Individual trials, three steps | Global fit 1+2+3 | DiaPharmaa | ||||
|---|---|---|---|---|---|---|---|
| 1 | 2 | 3 | 4* | Two steps | Three steps | ||
| k1, μM−1s−1 | 8.01 | 7.55 | 8.13 | 8.17 | 8.33 ± 0.44 | 7.92 ± 0.6 | - |
| k−1, s−1 | 6970 | 7340 | 6670 | 6630 | 3010 ± 160 | 6910 ± 680 | - |
| k2, s−1 | 2070 | 1950 | 2110 | 2140 | 800 ± 43 | 2043 ± 130 | - |
| k3, s−1 | 1760 | 1130 | 1410 | 4760 | - | 1077 ± 92 | - |
| kcat, s−1 | 952 | 715 | 846 | 1477 | 800 ± 43 | 705 ± 83 | 330 |
| Km, μM | 519 | 451 | 433 | 740 | 457 ± 31 | 390 ± 67 | 150 |
| kcat/Km, μM−1s−1 | 1.84 | 1.58 | 1.95 | 2.00 | 1.76 ± 0.15 | 1.81 ± 0.23 | 2.2 |
Outlier
H 8.3, 0.05 M Tris, 0.13 M NaCl at 37 °C.
Table 5.
Calculated rate constants from individual and combined runs for human APC-catalyzed hydrolysis of S-2366 at 1.2 Km as indicated in Table 3, but in the presence of 10% DMSO
| Individual trials, three steps
|
Global fit Two step | Global fit Three step | |||
|---|---|---|---|---|---|
| Parameter | 1 | 2 | 3 | ||
| k1, μM−1s−1 | 3.20 | 3.19 | 3.19 | 2.65 ± 0.05 | 3.20 ±0.04 |
| k−1, s−1 | 9440 | 9450 | 9480 | 4200 ± 92 | 9488 ± 420 |
| k2, s−1 | 560 | 555 | 554 | 290 ± 3 | 556 ± 22 |
| k3, s−1 | 914 | 606 | 584 | - | 651 ± 29 |
| kcat, s−1 | 347 | 290 | 284 | 290 ± 3 | 300 ± 20 |
| Km, μM | 1940 | 1640 | 1620 | 1690 ± 49 | 169 ± 151 |
| kcat/Km, M−1s−1 | 0.179 | 0.177 | 0.176 | 0.171 ± 0.005 | 0.177 ± 0.01 |
As stated above, experimentally-determined elementary rate constants are not readily available for most complex enzyme-catalyzed reactions. A good agreement between Michaelis-Menten parameters that were calculated from initial rates and those calculated from individual rate constants may be one criterion for the quality of the individual rate constants. The Michaelis-Menten parameters calculated from the value of the optimized elementary rate constants in Table 1, whether with the assumption that k2 and k3 are comparable or that k2 < k3, are reasonably close to the values reported for human thrombin-catalyzed hydrolysis of S-2238.[11;17;18] The parameters for the three other cases also compare well with those reported earlier [16;18;19] and/or the parameters in the Diapharma catalogue measured at a pH near 8.3, at the plateau of the pH profile for these reactions at 37 °C. However, a good agreement between Michaelis-Menten parameters obtained from initial rate measurements and those calculated from progress curves as described here, is not enough guarantee for the accuracy of individual rate constants from which they were calculated, because of the potential distortions due to interdependence of the individual rate constants. The critical question is how well can the concentration of transients be defined in the calculation for the determination of individual rate constants. The greatest uncertainty is in the value of k1 and k−1, while their ratios are well reproduced. The partitioning of EP to ES versus E + P is the other key element of the calculation.
Validation of elementary rate constants
For this purpose, we chose to study the thrombin-catalyzed hydrolysis of S-2238 at pH 8.0, 0.005 M Tris, 0.2 M NaCl, 0.1% PEG4000 and 25.0 ± 0.1 °C as described in a benchmark study of Wells and Di Cera[11] and at [S] = 7.5 Km. Figure 1 presents the progress curves. Wells and Di Cera[11] used temperature and salt perturbation to change the rate-limiting step and thereby unraveled the individual rate constants for this system. In our progress curve studies, thrombin was highly saturated with S-2238 under the exact same conditions as in the Wells and Di Cera study. As Table 1 illustrates, the value of k1, k−1, k2 and k3 calculated with DYNAFIT agree within 60% with the values measured by a completely different method. The agreement is even better between the Michaelis-Menten parameters: the best is for kcat/Km, while both our calculated kcat and Km values are higher by 60 and 120%, respectively, than the values published by Wells and Di Cera (but in very good agreement with the values listed in the Diapharma catalogue). This deviation is not at all uncommon between two laboratories even when using identical methods. Both experiments predict the second step (k2) determining the overall rate of the reaction at 73–76%. We consider this a significant result lending confidence to the individual rate constants which were calculated with DYNAFIT using a circumspect protocol.
Figure 1.

Three progress curves for 1.49 × 10−9 M human thrombin-catalyzed hydrolysis of 4× 10−5 M S-2388 at pH 8.0, 0.005 M Tris, 0.2 M NaCl, 0.1% PEG4000 and 25.0 ± 0.1 °C, monitored at 405 nm. Symbols indicate selected calculated points.
Testing the effects of enzyme concentration and levels of enzyme saturation by substrate on the elementary rate constants
We included two enzyme-catalyzed reactions of S-2366, a frequently used substrate of APC and a non-ideal substrate of thrombin, for which individual rate constants have not yet been reported. Enzyme concentrations were varied in some of the repeats, which gave different curves displayed on Figures 2 and 3, but had essentially no effect on the calculated rate constants (Table 2, trials 1 and 2, and Table 3, trials 1–3). These experiments probe then the sensitivity of the calculation to the concentrations of transient intermediates, which vary with the concentration of the enzyme.
Figure 2.

Progress curves of human thrombin-catalyzed (IIa) hydrolysis of S-2366 (S) under non-ideal conditions ([S] ~ Km) in the presence of thrombomodulin in pH 8.0, 0.02 M Tris, 0.3 M NaCl, 0.1% PG4000 and 25.0 ± 0.1 °C: 1) [IIa] =11.6 × 10−9 M, [S] = 2.00 × 10−4 M ; 2) [IIa] = 3.9 × 10−9 M, [S] = 2.00 × 10−4 M; 3) and 4) [IIa] = 3.9 × 10−9 M, [S] = 1.32 × 10−4 M, monitored at 405 nm. The calculated and measured data points are indistinguishable on this scale.
Figure 3.

APC-catalyzed hydrolysis of 1.00 × 10−4 M S-2366 ([S] = 0.125 Km) in pH 8.0, 0.02 M Tris, 0.3 M NaCl, and 0.1% PEG4000, at 25.0 ± 0.1 °C, monitored at 405 nm: [APC] 1) 1.54 ×10−8 M, 2) 3.08 ×10−8 M and 3) 4.62 ×10−8 M.
Table 3.
Rate constants calculated from individual runs and combined data for the human APC-catalyzed hydrolysis of S-2366, in pH 8.0, 0.02 M Tris buffer, 0.3 M NaCl, and 0.1% PEG4000, at 25. 0 ± 0.1 °C (Figure 3)
| Parameter | Individual trials, three steps | Global fit 1+2+3 | Ref [19] * | |||
|---|---|---|---|---|---|---|
| 1 | 2 | 3 | Two steps | Three steps | ||
| 108 [APC], M | 1.54 | 3.08 | 4.62 | |||
|
| ||||||
| k1, μM−1s−1 | 3.61 | 3.50 | 3.58 | 4.17 ± 0.30 | 3.54 ± 0.20 | - |
| k−1, s−1 | 8560 | 8650 | 8540 | 2850 ± 230 | 8580 ± 390 | - |
| k2, s−1 | 593 | 586 | 594 | 159 ± 13 | 590 ± 33 | - |
| k3, s−1 | 170 | 297 | 227 | - | 229 ± 11 | - |
| kcat, s−1 | 132 | 197 | 164 | 159 ± 13 | 165 ± 14 | 160 |
| Km, μM | 571 | 883 | 708 | 723 ± 76 | 723 ± 90 | 800 |
| kcat/Km, μM−1s−1 | 0.231 | 0.223 | 0.232 | 0.220 ± 0.029 | 0.228 ± 0.021 | 0.2 |
pH 8.0, 0.05 M Tris buffer, 0.13 NaCl and 0.01 M CaCl2 at 25.0 °C.
Because the Km values for these reactions are near millimolar, full saturation of the enzymes is not attainable unless co-solvents are used at > 1%. The absorption values at 405 nm, the absorption maximum of pNA, are also prohibitively high at these values of [S]. Measurements were carried out at [S] ~ Km for thrombin and at [S] ~ 0.125 Km for APC as shown in Figures 2 and 3. A small reduction in the concentration of S-2366 in Table 2, trials 3 and 4, effected the calculated values of all rate constants. The values of kcat and Km calculated from these individual rate constants are two-fold greater than those reported under very comparable conditions in the Diapharma catalogue, whereas the ratio, kcat/Km, remain reasonably reproducible. The observation that it is the kcat/Km ratio rather than individual values of kcat and Km that is well reproduced at such low concentrations of substrate confirms previous experience in enzyme kinetics. In case of the APC-catalyzed hydrolysis of S-2366 at low substrate concentrations, the Michaelis Menten parameters are well reproduced.
We studied these two systems in the presence of 10% DMSO, which allowed for raising the concentration of S-2366 to [S] ~6 Km for thrombin and above Km for APC. The release of pNA was monitored at a shoulder at 445 nm. Figures 4 and 5 show the data obtained and Tables 4 and 5 list the individual rate constants calculated. As might be expected, all rate constants have been affected by the presence of 10% DMSO. In this case, the calculated kcat value for the thrombin-catalyzed reaction agrees well with the one reported earlier, but the Km value increased, which might have been predicted. The APC-catalyzed hydrolysis of S-2366 in the presence of 10% DMSO again gives both an increased kcat and Km value, while their ratio agrees with the one measured earlier at 1% DMSO and with the Literature value
Figure 4.

Three progress curves for the 7.9 × 10−9 M thrombin-catalyzed hydrolysis of S-2366, at [S] = 6 Km, in pH 8.0, 0.02 M Tris, 0.3 M NaCl, 0.1% PG4000 , 25.0 ± 0.1 °C, but in the presence of 10% DMSO, and monitored at 445 nm. Symbols indicate selected calculated points.
Figure 5.

The 4.15 × 10−8 M APC-catalyzed hydrolysis of S-2366 at [S] = 1.2 Km, in pH 8.0, 0.02 M Tris, 0.3 M NaCl, 0.1% PG4000 , 25.0 ± 0.1 °C but in the presence of 10% DMSO, and monitored at 445 nm. Symbols indicate calculated points.
Table 4.
Rate constants calculated from individual and combined data for the 7.90 × 10−9 M human thrombin-catalyzed hydrolysis of S-2366 at [S] = 6 Km, as described in Table 2 but in the presence of 10% DMSO
| Individual trials, three steps
|
Global fit Two step | Global fit Three step | |||
|---|---|---|---|---|---|
| Parameter | 1 | 2 | 3 | ||
| k1, μM−1s−1 | 1.56 | 1.65 | 1.73 | 2.42 ± 0.09 | 1.87 ± 0.1 |
| k−1, s−1 | 2680 | 2660 | 2530 | 1775 ± 58 | 2294 ± 160 |
| k2, s−1 | 868 | 883 | 892 | 372 ± 17 | 717 ± 56 |
| k3, s−1 | 719 | 677 | 543 | - | 877 ± 78 |
| kcat, s−1 | 393 | 383 | 338 | 372 ± 17 | 395 ± 52 |
| Km, μM | 1028 | 933 | 749 | 887 ± 40 | 886 ± 153 |
| kcat/Km, μM−1s−1 | 0.382 | 0.411 | 0.450 | 0.420 ± 0.027 | 0.445 ± 0.05 |
Noteworthy is the consistency of the individual rate constants for the thrombin-catalyzed reactions. The bimolecular rate constant, k1, is greater for the hydrolysis of S-2238 than for S-2366, which is consistent with the well-documented differences in their affinities for thrombin. Binding is much weaker with S-2366 than with S-2238, which is further supported by the greater rate constants (k−1 & k2) for the decomposition of the Michaelis-Menten complex formed with S-2366. However, both k2 and k3 are greater for S-2366 than for S-2238, which is consistent with the relative magnitude in the measured kcat values. Calculated and experimentally measured k2 and k3 for S-2238 agree well (Table 1).
Isotopic sensitivity of individual rate constants: intrinsic deuterium isotope effects
One of our major interests is in finding a reliable method for the determination of intrinsic deuterium kinetic isotope effects from proton inventory measurements[16;17;24–33] on the rate-determining step involving proton transfer(s) in hydrolase-catalyzed reactions. In the proton inventory method, kinetic measurements are conducted in mixtures of buffered light and heavy water and the obtained rate constants of such measurements are then plotted against the atom fraction of deuterium in the solution. A proton inventory study of the 1.45 × 10−8 M FXa-catalyzed hydrolysis of 1 × 10−3 M S-2765 in pH 8.4 and equivalent pL (L = H, D and their mixtures), 0.02 M Tris, 0.3 M NaCl, 5 × 10−3 M CaCl2, and 0.1% PEG4000, at 25.0 ± 0.1 °C, resulted in the progress curves in Figure 6 and Table 6 has the calculated rate constants. The initial estimate for k2 was from kcat, which was calculated from Vmax given by the slope using the first 10 data points. In the first round of calculations using DYNAFIT, we fitted the two-step model to the data. The calculated rate constants returned from these fits were the estimate for the next round of calculations using the three-step model. The initial estimate of k3 was varied around the value of k2 within a range of ten and the final values of all rate constants are in Table 6. The kcat values calculated with the refined values of the elementary rate constants obtained with DYNAFIT are ~20% higher (Table 6) than the initial estimate for all nine cases, which is as expected when enzyme saturation with [S] is extrapolated to infinity. The Michaelis-Menten parameters in aqueous buffer, calculated from the data in Table 6 are; kcat = 276 ± 19 s−1 and kcat/Km = (3.9 ± 0.02) × 105 M−1 s−1, which compare well with the values obtained from fitting the integrated Michaelis-Menten equation to the time course of two thousand absorption readings: kcat = 247.0 ± 6.2 s−1 and kcat/Km = (4.49 ± 0.32) × 105 M−1 s−1.[16] Michaelis-Menten parameters were also calculated form the elementary rate constants for the other eight data sets obtained in varying isotopic buffer mixtures and are compared in Table 7 with the corresponding values calculated with the integrated Michaelis-Menten equation. The agreement is good, as it is good with kcat = 240 s−1, Km = 2.6 × 10−4 M, and kcat/Km = 9.2 × 105 M−1 s−1 in pH 8.3, 0.05 M Tris, 0.13 M NaCl and 0.5% BSA at 37 °C measured in the Diapharma lab.
Figure 6.

Progress curves for the 1.45 × 10−8 M FXa-catalyzed hydrolysis of 1 × 10−3 M S-2765 in pH 8.4 and equivalent pL (L = H, D and their mixtures), 0.02 M Tris buffer, 0.3 M NaCl, 5 × 10−3 M CaCl2, 0.1% PEG4000, at 25.0 ± 0.1 °C and monitored at 445 nm. All runs were followed to completion (not all shown on the figure).
Table 6.
Rate constants calculated using DYNAFIT for proton inventory study of the 1.45 × 10−8M human FXa-catalyzed hydrolysis of S-2765 = 1 × 10−3 M in pH 8.4 and equivalent pL (L = H, D and their mixtures), 0.02 M Tris buffer, 0.3 M NaCl, 5 × 10−3 M CaCl2, and 0.1% PEG4000, at 25.0 ± 0.1 °C
| n | 106 k1, M−1s−1 | k−1, s−1 | k2, s−1 | k3, s−1 |
|---|---|---|---|---|
| 0.00 | 4.74 ± 0.07 | 5750 ± 250 | 521 ± 21 | 586 ± 27 |
| 0.12 | 4.75 ± 0.16 | 5789 ± 340 | 514 ± 40 | 474 ± 33 |
| 0.24 | 3.98 ± 0.20 | 6066 ± 470 | 456 ± 31 | 364 ± 20 |
| 0.37 | 4.00 ± 0.10 | 6040 ± 170 | 457 ± 22 | 296 ± 9 |
| 0.49 | 4.88 ± 0.13 | 7619 ± 67 | 518 ± 15 | 220 ± 3 |
| 0.61 | 4.24 ± 0.16 | 6738 ± 420 | 471 ± 44 | 175 ± 6 |
| 0.73 | 3.57 ± 0.09 | 7042 ± 150 | 387 ± 7 | 162 ± 1 |
| 0.86 | 3.49 ± 0.07 | 7603 ± 130 | 378 ± 9 | 116 ± 1 |
| 0.98 | 3.29 ± 0.21 | 7846 ± 110 | 363 ± 26 | 94 ± 2 |
Table 7.
A comparison of Michaelis Menten parameters calculated from the individual rate constants obtained with DYNAFIT in Table 6 with those calculated using eq 1 for the integrated Michaelis-Menten equation. The nine sets of 2000 data pairs were obtained for proton inventory study[16] of the 1.45 × 10−8 M human FXa-catalyzed hydrolysis of S-2765 = 1 × 10−3 M in pH 8.4 and equivalent pL (L = H, D and their mixtures), 0.02 M Tris buffer, 0.3 M NaCl, 5 × 10−3 M CaCl2, and 0.1% PEG4000, at 25.0 ± 0.1 °C
| kcat, s−1 | 10−5 kcat/Km, M−1s−1 | |||
|---|---|---|---|---|
| n | Dynafit | Equation 1 | Dynafit | Equation 1 |
| 0 | 276 ± 19 | 247.0 ± 6.2 | 3.9 ± 0.2 | 4.49 ± 0.32 |
| 0.12 | 247 ± 29 | 212 ± 13 | 3.9 ± 0.4 | 3.79 ± 0.33 |
| 0.24 | 202 ± 20 | 181.3 ± 5.5 | 2.8 ± 0.3 | 2.81 ± 0.15 |
| 0.37 | 179 ± 12 | 165.9 ± 1.8 | 2.8 ± 0.2 | 2.71 ± 0.18 |
| 0.49 | 154 ± 6 | 133.2 ± 5.1 | 3.1 ± 0.1 | 3.73 ± 0.39 |
| 0.61 | 128 ± 16 | 115.8 ± 5.1 | 2.8 ± 0.3 | 2.85 ± 0.16 |
| 0.73 | 114 ± 3 | 93.6 ± 1.5 | 1.9 ± 0.1 | 2.21 ± 0.13 |
| 0.86 | 88 ± 3 | 75.9 ± 0.6 | 1.7 ± 0.1 | 1.60 ± 0.21 |
| 0.98 | 74 ± 7 | 66.1 ± 0.1 | 1.5 ± 0.1 | 1.52 ± 0.21 |
Note: Maximal velocities were calculated from the slope given by the first ten data points in each set and an initial value of kcat was calculated from these and [E]. It was assumed that kcat = k2 for the first round of DYNAFIT calculations. The k2 value returned from the fit was then used as initial estimate for the next round of fit with the three step model. The estimate for k3 was varied within ten fold around the value of k2. The kcat values calculated from the individual rate constants in the first column are ~ 20% higher than the initial estimates, which is as expected when enzyme saturation with [S] is extrapolated to infinity.
Gross and Butler described the equation relating rates or rate constants to deuterium content in solution as follows:[24;27;28;34;35]
| (2) |
where Vn and Vo are velocity (or rate constant) in a binary solvent and in water, respectively, n = atom fraction of deuterium, RS = reactant state, TS = transition state, φR = RS fractionation factor and φT = TS fractionation factor. The fractionation factors are obtained from inverted equilibrium isotope effects, KD/KH, for exchange between a bulk water site and a particular structural site of the RS or the TS. Justifiable simplifications of this equation involve the assumption of a unit fractionation factor of RSs for catalytic residues with NH and OH functional groups and the assumption that one or two active-site units contribute in most hydrolytic enzymes.
Figure 7 shows the proton inventory curves: The equations of the lines for fitting the data are modifications,[16;17;33] of the Gross-Butler equation;[24;27;28;34;35] kn/kH = (1−n + n/(2.0)) for K1, kn/kH = (1−n + n/(1.4)) for k2, and kn/kH = (1−n + n/(3.1)2) for k3, respectively.
Figure 7.

Proton Inventory for K1 = k1/k−1, triangles, k2 closed circles, and k3 open circles. The equations of the lines for fitting the data were modifications,[17;31] of the Gross-Butler equation;[22;25;26;32;33] kn/kH = (1−n + n/(2.03)); kn/kH = (1−n + n/(1.43)); and kn/kH = (1−n + n/(3.11)2), respectively.
Rate-determining step
Hedstrom[8] warns against the fallacy in the generally held belief that acylation is slower for amide hydrolysis than deacylation. She arraigns a number of cases where this expectation fails. Many other hydrolase-catalyzed reactions are known for which the rate is determined partly by acylation and partly by deacylation.[36] Although the true identity of k2 and k3 has not been revealed in these studies, they were assigned to acylation and deacylation in the thrombin-catalyzed reactions of S-2238[11] and in reactions of other substrates when studied with the perturbation techniques cited in the Introduction.[10;12–15] By analogy, we may assign the k2 and k3 values for the thrombin-catalyzed hydrolysis of S-2238 also to acylation and deacylation, respectively. We then extend this choice to the other cases too: k2 and k3 measure acylation and deacylation, respectively. The DYNAFIT-calculated values of k2 and k3 are never distinctly different, in fact, they are nearly equal for the reactions of S-2366 and for the last case, the FXa-catalyzed hydrolysis of S-2675. As discussed above, this is often the case experimentally.
Of great importance is that the value of k2 is insensitive to the content of D, while k3 decreases dramatically as the D content is increased in the buffer system, everything else being the same. The third step, quite likely the hydrolysis of the acyl enzyme, becomes increasingly rate determining with increasing D content in the medium. The proton inventory curve (Figure 7) is deeply concave and is most consistent with two protonic bridges participating in deacylation with fractionation factors 0.32 ± 0.02 each and corresponding to an intrinsic isotope effect of 3.1 ± 0.2. The protonic bridges are most likely to be between the catalytic Ser and His and between His and Asp at the active site.[17;33] The proton inventory for the binding step and acylation step are nearly linear within experimental errors and the fractionation factors associated with them are smaller in value and less certain in origin.
This is a very significant result, because the method allows for the determination of the intrinsic solvent isotope effects or partial solvent isotope effects of individual steps of a complex reaction. An access to intrinsic deuterium isotope effects in the rate-determining step of enzyme-catalyzed acyl transfer reactions will be a major advance in the interpretation of the role of proton transfer in these mechanisms.
In conclusion, the use of DYNAFIT holds the premise for obtaining individual rate constants from a relatively small number of, albeit meticulously precise, progress curve measurements for irreversible one substrate reactions without substrate and product inhibition. Even poor substrates that may not be used in concentrations that saturate the active sites give reasonable results.
Acknowledgments
This work was supported in part by the US National Institute of Health, Grant No 1 R15 HL067754-01. We are grateful to Dr. Laszlo Szilagyi in the Department of Biochemistry at Eotvos Lorand University, Budapest Hungary and Dr. Petr Kuzmic BioKin Ltd. 1652 S. Grand Ave., Suite 337, Pullman, WA 99163, for helpful consultations in using DYNAFIT.
Footnotes
Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Reference List
- 1.Kuzmic P. Program DYNAFIT for the Analysis of Enzyme Kinetic Data: Application to HIV Proteinase. Anal Biochem. 1996;237:260–273. doi: 10.1006/abio.1996.0238. [DOI] [PubMed] [Google Scholar]
- 2.Garfinkel D, Ching SW, Adelman M, Clark P. Techniques and problems in the construction of computer models of biochemical systems including real enzymes. Ann NY Acad Sci. 1966;128:1054–1068. doi: 10.1111/j.1749-6632.1965.tb11716.x. [DOI] [PubMed] [Google Scholar]
- 3.Barshop BA, Wrenn RF, Frieden C. Analysis of Numerical Methods for Computer Simulation of Kinetic Processes: Development of KINSIM--A Flexible, Portable System. Analytical Biochemistry. 1983;130:134–145. doi: 10.1016/0003-2697(83)90660-7. [DOI] [PubMed] [Google Scholar]
- 4.Zimmerle CT, Frieden C. Analysis of progress curves by simulations generated by numerical integration. Biochem J. 1989;258:381–387. doi: 10.1042/bj2580381. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Duggleby RG. Quantitative analysis of enzyme progress curves by non-linear regression. Meth Enzymol. 1995;249:61–90. doi: 10.1016/0076-6879(95)49031-0. [DOI] [PubMed] [Google Scholar]
- 6.Goudar CT, Sonnad JR, Duggleby RG. Parameter estimation using a direct solution of the integrated Michaelis-Menten equation. Biochimica et Biophysica Acta. 1999;1429:377–383. doi: 10.1016/s0167-4838(98)00247-7. [DOI] [PubMed] [Google Scholar]
- 7.Fersht A. Structure and Mechanism in Protein Science. W. H. Freeman and Co.; NY, NY.: 1999. pp. 1–631. [Google Scholar]
- 8.Hedstrom L. Serine protease mechanism and specificity. Chem Rev. 2002;102:4501–4524. doi: 10.1021/cr000033x. [DOI] [PubMed] [Google Scholar]
- 9.Brouwer AC, Kirsch JF. Investigation of diffusion-limited rates of chymotrypsin reactions by viscosity variation. Biochemistry. 1982;21:1302–1307. doi: 10.1021/bi00535a030. [DOI] [PubMed] [Google Scholar]
- 10.Stone SR, Betz A, Hofsteenge J. Mechanistic Studies on Thrombin Catalysis. Biochemistry. 1991;30:9841–9848. doi: 10.1021/bi00105a005. [DOI] [PubMed] [Google Scholar]
- 11.Wells CM, Di Cera E. Thrombin is a Na+-Activated Enzyme. Biochemistry. 1992;31:11721–11730. doi: 10.1021/bi00162a008. [DOI] [PubMed] [Google Scholar]
- 12.Picozzi M, Landolfi R, Decristofaro R. Effects of Protons on the Thrombin-Fibrinogen Interaction. Eur J Biochem. 1994;219:1013–1021. doi: 10.1111/j.1432-1033.1994.tb18584.x. [DOI] [PubMed] [Google Scholar]
- 13.Hopfner KP, Di Cera E. Energetics of Thrombin-Fibrinogen Interaction. Biochemistry. 1992;31:11567–11571. doi: 10.1021/bi00161a040. [DOI] [PubMed] [Google Scholar]
- 14.Vindigni A, Di Cera E. Release of Fibrinopeptides by the Slow and Fast Forms of Thrombin. Biochemistry. 1996;35:4417–4426. doi: 10.1021/bi952834d. [DOI] [PubMed] [Google Scholar]
- 15.Dicera E, De Cristofaro R, Albright DJ, Fenton JW. Linkage Between Proton Binding and Amidase Activity in Human α-Thrombin - Effect of Ions and Temperature. Biochemistry. 1991;30:7913–7924. doi: 10.1021/bi00246a007. [DOI] [PubMed] [Google Scholar]
- 16.Zhang D, Kovach IM. Deuterium solvent isotope effect and proton-inventory studies of factor Xa-catalyzed reactions. Biochemistry. 2006;45:14175–14182. doi: 10.1021/bi061218m. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Enyedy EJ, Kovach IM. Proton Inventory Studies of Thrombin-Catalyzed Reactions of Substrates with Selected P and P' Sites. J Am Chem Soc. 2004;126:6017–6024. doi: 10.1021/ja0320166. [DOI] [PubMed] [Google Scholar]
- 18.Lottenberg R, Christensen U, Jackson CM, Coleman PL. Assay of Coagulation Proteases Using Peptide Chromogenic and Fluorogenic Substrates. Meth Enzymology. 1981;28:341–361. doi: 10.1016/s0076-6879(81)80030-4. [DOI] [PubMed] [Google Scholar]
- 19.Sala N, Owen GW, Collen D. A functional assay of Protein C in human plasma. Blood. 1984;63:671–675. [PubMed] [Google Scholar]
- 20.Leatherbarrow RJ. GraFit User's Guide. Ertihacus Sofware Ltd.; Staines, U. K.: 1992. [Google Scholar]
- 21.Lottenberg R, Jackson CM. Solution Composition Dependent Variation in Extinction Coefficients for p-Nitroaniline. Biochim Biophys Acta. 1983;742:558–564. doi: 10.1016/0167-4838(83)90273-x. [DOI] [PubMed] [Google Scholar]
- 22.Johnson ML, Faunt LM. Parameter estimation by least-squares methods. Methods Enzymol. 1994;210:1–37. doi: 10.1016/0076-6879(92)10003-v. [DOI] [PubMed] [Google Scholar]
- 23.Johnson ML. Why, when and how biochemists should use least squares. Anal Biochem. 1992;206:215–225. doi: 10.1016/0003-2697(92)90356-c. [DOI] [PubMed] [Google Scholar]
- 24.Venkatasubban KS, Schowen RL. The Proton Inventory Technique. CRC Crit Rev Biochem. 1985;17:1–44. doi: 10.3109/10409238409110268. [DOI] [PubMed] [Google Scholar]
- 25.Stein RL, Elrod JP, Schowen RL. Correlate Variations in Enzyme - Derived and Substrate - Derived Structures of Catalytic Transition States. Implications for the Catalytic Strategy of Acyl - Transfer Enzymes. J Am Chem Soc. 1983;105:2446–2452. [Google Scholar]
- 26.Stein RL, Strimpler AM, Hori H, Powers JC. Catalysis by Human Leukocyte Elastase: Proton Inventory as a Mechanistic Probe. Biochemistry. 1987;26:1305–1314. doi: 10.1021/bi00379a016. [DOI] [PubMed] [Google Scholar]
- 27.Alvarez FJ, Schowen RL. Mechanistic Deductions from Solvent Isotope Effects. In: Buncel E, Lee CC, editors. Isotopes in Organic Chemistry. Elsevier; Amsterdam: 1987. pp. 1–60. [Google Scholar]
- 28.Kresge AJ, More O, Powell MF. Solvent Isotopes Effects, Fractionation Factors and Mechanisms of Proton Transfer Reactions. In: Buncel E, Lee CC, editors. Isotopes in Organic Chemistry. Elsevier; Amsterdam: 1987. pp. 177–273. [Google Scholar]
- 29.Scholten JD, Hogg JL, Raushel FM. Methyl Chymotrypsin Catalyzed Hydrolysis of Specific Substrate Esters Indicate Multiple Proton Catalysis Is Possible with a Modified Charge Relay Triad. J Am Chem Soc. 1988;110:8246–8247. [Google Scholar]
- 30.Schowen RL. Structural and Energetic Aspects of Protolytic Catalysis by Enzymes: Charge - Relay Catalysis in the Function of Serine Proteases. In: Liebman JF, Greenberg A, editors. Mechanistic Principles of Enzyme Activity. VCH Publishers; New York: 1988. pp. 119–168. [Google Scholar]
- 31.Schowen KB, Limbach HH, Denisov GS, Schowen RL. Hydrogen bonds and proton transfer in general-catalytic transition state stabilization in enzyme catalysis. Biochem Biophys Acta. 2000;1458:43–62. doi: 10.1016/s0005-2728(00)00059-1. [DOI] [PubMed] [Google Scholar]
- 32.Quinn DM, Sutton LD. Theoretical Basis and Mechanistic Utility of Solvent Isotope Effects. In: Cook PF, editor. Enzyme Mechanism from Isotope Effects. CRC Press; Boston, MA: 1991. pp. 73–126. [Google Scholar]
- 33.Zhang D, Kovach IM. Full and Partial Deuterium Solvent Isotope Effect Studies of α-Thrombin-catalyzed Reactions of Natural Substrates. J Am Chem Soc. 2005:3760–3766. doi: 10.1021/ja043258o. [DOI] [PubMed] [Google Scholar]
- 34.Gold V. Protolytic processes in H2O-D2O mixtures. Adv Phys Org Chem. 1969;7:259–331. [Google Scholar]
- 35.Kresge AJ. Solvent isotope effects in H2O-D2O Mixtures. Pure Appl Chem. 1964;8:243–258. [Google Scholar]
- 36.Quinn DM. Acetylcholinesterase: Enzyme Structure, Reaction Dynamics, and Virtual Transition States. Chem Rev. 1987;87:955–975. [Google Scholar]
