Abstract
Despite growing evidence suggesting the importance of enzyme conformational dynamics (ECD) in catalysis, a consensus on how precisely ECD influences the chemical step and reaction rates is yet to be reached. Here, we characterize ECD in Cyclophilin A, a well-studied peptidyl-prolyl cis-trans isomerase, using normal and accelerated, atomistic molecular dynamics simulations. Kinetics and free energy landscape of the isomerization reaction in solution and enzyme are explored in unconstrained simulations by allowing significantly lower torsional barriers, but in no way compromising the atomistic description of the system or the explicit solvent. We reveal that the reaction dynamics is intricately coupled to enzymatic motions that span multiple timescales and the enzyme modes are selected based on the energy barrier of the chemical step. We show that Kramers’ rate theory can be used to present a clear rationale of how ECD affects the reaction dynamics and catalytic rates. The effects of ECD can be incorporated into the effective diffusion coefficient, which we estimate to be about ten times slower in enzyme than in solution. ECD thereby alters the preexponential factor, effectively impeding the rate enhancement. From our analyses, the trend observed for lower torsional barriers can be extrapolated to actual isomerization barriers, allowing successful prediction of the speedup in rates in the presence of CypA, which is in notable agreement with experimental estimates. Our results further reaffirm transition state stabilization as the main effect in enhancing chemical rates and provide a unified view of ECD’s role in catalysis from an atomistic perspective.
Keywords: cis-trans isomerization, Cyclophilin A, enzyme catalysis, enzyme dynamics, Kramers’ rate theory
Enzymes accelerate reaction rates by several orders of magnitude, allowing them to occur at timescales relevant for cellular functions (1). One of the long-standing issues in biochemistry is how enzymes achieve this remarkable speedup. It is commonly accepted that the most dominating effect arises from significant reduction in the free energy barrier compared to the corresponding noncatalyzed reaction in solution. It is also well established that this predominant effect is mainly electrostatic in nature (2, 3), which is more favorable for the transition state than the reactant or the product (1). However, to what degree and how other factors such as desolvation, steric strain, and enzyme dynamics contribute to catalysis remains disputable. Of particular interest is the role of enzyme dynamics in catalysis that has stirred considerable debate (4–11) partly because it has not been clearly defined, leading to a semantic issue. Also, the link between enzyme dynamics and catalysis is difficult to address both experimentally and theoretically. Currently, the implications of enzyme dynamics are from ensemble- and time-averaged experiments, as the temporal behavior of every atom cannot be observed directly. Although standard molecular dynamics (MD) simulations can provide an atomistic picture of enzyme dynamics, they are still not amenable to study catalytic reactions that usually occur in milliseconds. Computational approaches that have investigated the effects of millisecond-timescale enzyme dynamics on the chemical reaction have been possible only with the use of coarse-grained models (4, 10). NMR relaxation dispersion experiments that can probe microsecond-millisecond timescale motions have detected backbone and side-chain motions in and around the active site that occur on the same millisecond-timescale as the chemical step (12). It has been further shown that such slow motions are already present in the free enzyme (13). Furthermore, loss of conformational fluctuations occurring in milliseconds in the active site of mutant enzyme has been observed with concomitant reduction in activity (11). Single molecule studies on enzymes have also revealed that catalytic rates can fluctuate over five orders of magnitude—from milliseconds to hundreds of seconds, similar to the range of timescale for conformational fluctuations (14). These observations are not surprising given that protein dynamics comprise motions that span multiple timescales and occur in either a more localized or collective manner (15, 16). Nevertheless, protein dynamics has been suggested to directly contribute to catalytic function and rate enhancement. The exact nature of this dynamical contribution cannot be understood, unless specific questions regarding whether dynamical motions of enzymes help in lowering the activation barrier (i.e., barrier effects) or aid the substrate to surmount the barrier (i.e., prefactor effects) are addressed.
The energy landscape of proteins is characterized by several energy minima that represent conformational substates separated by barriers of varying heights (17, 18). Simultaneous motions of many degrees of freedom constitute protein dynamics and bring about equilibrium interconversions (16, 19). We sought to understand the role of enzyme conformational dynamics (ECD) in catalytic functions by employing a combination of normal MD (nMD) and accelerated MD (aMD) (20) approach that provides atomistic detail with extended timescale. We chose to study Cyclophilin A (CypA) (Fig. 1A), an extensively studied peptidyl-prolyl cis-trans isomerase, that catalyzes isomerization of the peptide (ω) bond preceding proline residues in proteins. Such system is ideal to study using classical molecular mechanics because no bond breaking or formation is involved in the catalytic process.
Fig. 1.
Structure of CypA and the influence of its dynamics on the kinetics of prolyl isomerization in its substrate. (A) CypA (gray) with its substrate Ace-Ala-Ala-Pro-Phe-Nme in the binding pocket. The binding cavity (VDW surface) is defined by ten residues (nonpolar (white): Phe60, Met61, Ala101, Ala103, Phe113, Leu122 and polar (cyan): Arg55, Gln63, Asn102, and His126) that are within 4 Å of the substrate’s Ala-Pro motif. Decay of probability of survival in the trans well as a function of time obtained from nMD simulations of (B) RO and (C) RC with V2 set to 11.0 (cyan), 9.0 (red), 7.0 (blue), 5.0 (yellow), 4.0 (green), and 0.0 (magenta) kcal/mol. Continuous lines are fits to single exponential in (B) and multiexponential functions in (C). (Inset to B) Plots of survival probability functions on an extended timescale when V2 = 7.0, 9.0 and 11.0 kcal/mol. Parameters from exponential fits of isomerization kinetics, i.e. the amplitudes and the time constants, τ, are plotted for (D) RO and (E) RC with the same color code as in (B) and (C).
Probing the influence of CypA dynamics on the chemical step
The uncatalyzed isomerization reaction (RO) is an extremely slow process with an activation barrier of approximately 20 kcal/mol and occurs readily in hundreds of seconds in solution (21, 22). Cyclophilins are known to accelerate prolyl isomerization by 105–106 times, reducing the timescale to around milliseconds (23, 24). It is not feasible to simulate even the catalyzed reaction (RC) with nMD, since it is currently limited to only hundreds of nanoseconds. Therefore, to probe the effects of ECD in catalysis, we used several lowered torsional energy barriers around the -Ala-Pro- ω bond of a well-studied substrate analogue, Ace-Ala-Ala-Pro-Phe-Nme (Fig. 1A). We then took advantage of the linearity of the behavior to extrapolate to the desired barriers. This approach is similar to carrying out experiments in an optimum temperature range and extrapolating to a temperature outside that range. The lower barriers allowed us to track the dwell times in the trans well before going over the barrier with sufficiently good statistics. Using nMD we investigated the kinetics of prolyl isomerization in the free and the enzyme-bound substrate with the same value of the AMBER force field parameter V2 (see Methods and SI Text, section 1.4), which is the main determinant of torsional barrier. We found that the decay of the survival probability function, S(t), of dwell times for the reference RO in the free substrate unambiguously exhibited single exponential behavior (Fig. 1B). Progressively slower kinetics resulted as V2 was systematically increased from 0 to 11 kcal/mol. Interestingly, for RC, the resulting kinetics exhibited multiexponential decays (Fig. 1C).
In view of the notion that protein dynamics is not independent of its environment (19), we observed that peptide isomerization dynamics was influenced by the environment, whether it be the fluctuations in the solvent or the active site of the environment. From Fig. 1 B and D, it became evident that in terms of timescale, the dynamics of the aqueous environment was relatively uniform with the solvent motions occurring on a single or a very narrow timescale. In contrast, enzymatic motions were dispersed over an extensive timescale and were coupled to substrate dynamics. Consequently, the different enzyme dynamic modes became apparent in the isomerization kinetics, yielding multiexponential behavior (Fig. 1 C and E). As the torsional barrier became progressively greater (with the increase in V2), the distributions of timescales showed the following trend: the relative amplitudes of the faster phases diminished and those of the slower phases showed a gradual increase. These results suggested that, depending on the barrier, and hence characteristic timescale, of the chemical step, the reaction dynamics got coupled to the slightly faster and slower enzyme motions, resulting in multiexponential decays. Overall, the time constant for RO in solution was always larger than the average lifetime of the chemical step in the enzyme-bound substrate, RC (Table I). It was evident from our results that the chemical step was coupled to, and would be affected by, the enzyme motions.
Table 1.
Free energy barriers, diffusion and time constants for the uncatalyzed and catalyzed prolyl isomerization*
|
V2 (kcal/mol) |
Ro | Rc | |||||
| ΔG# (kcal/mol) |
(ns) |
Deff(deg2/s) |
ΔG#(kcal/mol) |
(ns) |
Deff(deg2/s) |
||
| nMD | 0.0 | 0.81 | 0.05 | 18.1 × 1014 | - | 0.01 | 1.37 × 1014 |
| nMD | 4.0 | 2.3 | 0.28 | 18.1 × 1014 | - | 0.07 | 1.37 × 1014 |
| nMD | 5.0 | 3.16 | 0.54 | 18.1 × 1014 | - | 0.16 | 1.37 × 1014 |
| nMD | 7.0 | 3.93 | 2.48 | 18.1 × 1014 | 0.99 | 0.50 | 1.37 × 1014 |
| nMD | 9.0 | 5.57 | 16.11 | 18.1 × 1014 | 2.31 | 1.35 | 1.37 × 1014 |
| nMD | 11.0 | 7.42 | 61.65 | 18.1 × 1014 | 3.49 | 6.24 | 1.37 × 1014 |
| aMD | 28.0 | 20.0a | 1.22 × 1011b | 10.74a | 2.75 × 105b | ||
| aMD I | 7.0 | 2.09 | 0.26 | 5.443 × 1014 | |||
| aMD II | 7.0 | 2.37 | 0.14 | 13.362 × 1014 | |||
| aMD III | 7.0 | 2.7 | 0.09 | 31.229 × 1014 | |||
*ΔG#'s were calculated from potentials of mean force obtained from umbrella sampling except for (a) where they were obtained after reweighting free energy profiles resulting from aMD. (b) Time constants extrapolated from linear fits (red and green lines) in figure 3c. AMD levels I, II and III are the lowest, intermediate and the highest extents of acceleration (same as in Fig. 2D) subjected only on CypA. Deff = 2πkBT. exp(ln(k/wowb)ΔG#=0).
Accelerating CypA Dynamics and Its Effects on the Chemical Step
Another question of interest is whether the contribution from ECD can significantly enhance the rate of the chemical step as compared to that in solution. In answering this question, we subjected only CypA to increasing levels of aMD with the substrate still stimulated with nMD (SI Text, section 1.3). In order to confirm that aMD indeed resulted in faster ECD in CypA, we characterized the fluctuations in free CypA(Co) from independent nMD and aMD (SI Text, section 1.2 and 1.3). Accelerated MD brought about an increase in not only conformational plasticity; i.e., greater amplitudes of fluctuations as depicted from the shift of backbone (Fig. 2 A and B) and side chain (Fig. S1) order parameters (S2) to lower values, but also conformational heterogeneity (Fig. S2) at the active site of Co. Similar to our results, aMD has been shown to successfully increase the rate of conformational sampling, thereby characterizing millisecond-timescale protein/peptide dynamical motions and achieving notable agreement with experimental data (25–29). Our simulations further confirmed recent experimental observations (13, 14) that ECD in CypA takes place over a broad range of timescales even in the substrate-free state (Fig. S1). Accelerating CypA dynamics clearly affected the kinetics of prolyl isomerization in the bound substrate, resulting in faster decay of the survival probability (Fig. 2C). The decays fitted to multiexponential functions with only three phases as opposed to five phases in the nMD of CypA (Fig. 2D). Since the enzyme modes sped up, the relative contribution of the faster phases increased as slower phases became faster (Fig. 2D). The net result was a gradual speed up in the average lifetimes as the extent of acceleration of CypA motions was increased (Table I).
Fig. 2.
Effects of accelerating CypA dynamics on prolyl isomerization in the substrate. Distribution of order parameters (S2) (SI Text, section 1.8) obtained from (A) nMD and (B) aMD of free CypA using the highest level of acceleration. CypA structure is color-coded based on the S2 values of each backbone N-H bond vector (see color scale). The active site residues are shown with stick representation. Fluctuational motions with the largest amplitudes are indicated by the smallest S2 (red) while those with the smallest amplitudes are depicted with the largest S2 (blue). (C) Decay of probability of survival in the trans well as a function of time when V2 = 7 kcal/mol. (D) Parameters of exponential fits in C; i.e., amplitudes and time constants, τ, are plotted with the same color code as in (C). Shown are the isomerization kinetics in the free substrate (cyan, S) when subjected to nMD and the enzyme-bound substrate when CypA was subjected to nMD (blue, ES) as well as aMD at the lowest (orange, A1), intermediate (dark red, A2), and the highest (dark green, A3) level of acceleration. Continuous black lines are mono- or multiexponential fits.
Using Kramers’ Rate Theory to Explain the Effects of CypA Dynamics
Although the usage of traditional rate theories to explain enzyme kinetics has been a contentious matter (8, 30, 31), our results could be rationalized within the framework of Kramers’ theory (32, 33) in the high friction regime:
, where k is the rate of escape from the trans well with curvature ωo over the free energy barrier ΔG# with curvature ωb, kB is the Boltzmann constant and T is the temperature. Deff is the effective diffusion coefficient on a one-dimensional free energy profile and assumed to be independent of the reaction coordinate. Deff incorporates the effects of the environment, as well as those inherent in proteins, for example, frictional, dynamical effects and energetic roughness. Frictional effects arise from solvent viscosity and internal friction that impede protein motions. Dynamical effects originate from enzyme with inhomogeneous diffusivity, arising from ECD occurring on a wide continuum of timescales, or an aqueous medium that offers a more homogeneous environment with essentially single (or very narrowly distributed) diffusion coefficient. The substrate undergoes desolvation while moving into the active site of the enzyme from aqueous solvent, as a result the energetic roughness may reduce, leading to a less hindered substrate (34). The above Kramers’ relation can be rearranged in the log form; i.e.,
. From the plots of
vs.ΔG#,v2 (where rate constants, curvatures, and free energy barrier heights correspond to various values of V2) with a well-defined slope of 1/kBT, the relative contributions from the preexponential factor and the barrier effects were estimated for RO and RC. (Fig. 3). Clearly, there was speedup in the isomerization rate in each case of the enzyme-bound substrate as compared to the corresponding RO (Fig. 3A). The speedup was the consequence of two opposing effects: Increase in rate as a result of the reduction in free energy barrier, for example, 237 times from a barrier reduction of 3.26 kcal/mol in case of V2 = 9.0 kcal/mol, which was offset by approximately 2.6 times due to the modification in the curvatures of the free energy profiles and by approximately 13 times due to the differences in the Deff in solution and in enzyme-bound substrate, bringing the net rate enhancement to only about seven times. It can be seen from Fig. 3A that the y-intercept, from which the effective diffusion coefficient can be estimated, is smaller for RC than RO (Table I). Therefore, ECD does not enhance but rather hinders the rate enhancement. When the dynamics of CypA was accelerated, we clearly noticed an increase in the isomerization rates (Table I). Enzymatic CD does not directly modify the properties of the free energy landscape, as that scenario would violate Boltzmann statistics. However, for each level of acceleration subjected on CypA, the solvated CypA-substrate complex should be considered a distinct system associated with its Hamiltonian and characteristic free energy profile. And indeed, the curvatures and barrier heights were modified (Table I) when ECD was accelerated. As the levels of acceleration were raised, both ΔG# and Deff showed an increase (Fig. 3B) relative to the case in which ECD was not sped up (i.e., RC). Noticeably, even with the highest level of acceleration, Deff was not significantly faster (i.e., by only approximately 1.7 times) than the one in aqueous solution, given the errors associated with the calculation of quantities from logarithmic scale and MD simulations. Further acceleration of CypA dynamics could reach a limiting case where the integrity of the active site might be lost as a result of the unfolding of the enzyme brought about by very fast dynamical motions. Our analysis therefore suggested that altering motions associated with ECD made the enzyme’s active site environment more aqueous-like and directly affected the preexponential factor. At the same time, the favorable interactions between CypA and its substrate were possibly perturbed such that the free energy height of isomerization is increased relative to RC but still remains lower than that for RO.
Fig. 3.
Comparison of prolyl isomerization kinetics in the free and the enzyme-bound substrate. Kramers’ plots are shown in the form of ln (k/ωoωb) vs. ΔG#. (A) nMD data points for RO (open blue circles) and RC (filled blue circles) when V2 = 0,4, 5, 7, 9, and 11 kcal/mol. (B) Same plot as in (A). For clarity, only data points from nMD corresponding to V2 = 7.0 kcal/mol are shown. Also plotted are the data points for RC from aMD when the lowest (violet), intermediate (magenta), and the highest (red) level of acceleration are applied on CypA. All continuous lines are linear fits with slope = 1/kBT. (C) Same plot as in (A) with linear fits that are extrapolated to higher free energy barriers. Data points for RO (open orange circle) and RC (filled orange circle) are shown when the reoptimized V2 = 28 kcal/mol was used in aMD and assumed to follow the corresponding linear trends (i.e., red and green lines, respectively). Gray dashed lines above the red and below the green lines (with the same slope = 1/kBT) represent illustrative kinetic trends for RC in CypA mutants with faster and slower dynamics, respectively than the wild-type enzyme. Horizontal and vertical dashed lines with arrows depict reduction in free energy barriers and speedup in isomerization rates (along with sharper curvatures), respectively.
In an independent aMD study, setting V2 to the reoptimized value of 28 kcal/mol (35), we calculated the free energy profiles with the expected actual barriers for RO and RC (SI Text, section 1.3, and Fig. S4). We estimated the rate constants (Table 1) corresponding to the actual RO and RC from their respective linear fits in Kramers’ plots (Fig. 3C). RC showed a speed up of approximately 4.5 × 105 times over RO, which was strikingly similar to experimental estimates (23). The above result therefore validated Kramers’ theory in the analysis of prolyl isomerization kinetics and its catalysis. We would like to note from Fig. 3C. that in the case of CypA, as one goes to the higher barrier regime, the relative and dominant contribution from the reduction in barrier heights to the rate enhancement will continue to increase while the difference in Deff between RC and RC will remain at approximately 13 times.
Recently, ambient temperature X-ray and relaxation NMR studies on CypA have shown that impediment of motions that help in the interconversion of conformational substates in a mutant enzyme is accompanied with the reduction in catalytic rate (36). It was therefore concluded that protein dynamic motions contribute directly to the catalytic power of the enzyme. Such results can now be explained with Kramers’ theory that allows us to understand the nature of the dynamical contribution. The CypA-mutant with slower dynamics implies that the isomerization reaction would take place on a distinct free-energy profile with barrier heights and curvatures that are different from the wild-type CypA and with an effective diffusion coefficient that is perhaps slower than the wild-type enzyme (Fig. 3C). As we show above, it is equally important to investigate the dynamical effects as well as the free-energy barrier effects of the mutant, which, in most cases, is missing from experimental analyses. The recent studies linking enzyme dynamics to catalysis have focused on mutants with either slower or total absence of fluctuations in the active site as compared to the wild-type enzyme (11, 36). It would be interesting to investigate mutants with faster dynamics (Fig. 3C) and observe whether the catalytic rates are enhanced or not and how they compare with wild-type- and nonenzymatic rates.
Characterizing ECD in the Free and Bound CypA
We compared the dynamic motions in the active site of CypA in the absence and the presence of the substrate. Three separate nMD were carried out for the CypA-bound complex (Fig. 1A), when the -Ala-Pro- peptidyl-prolyl ω-bond of the substrate, Ace-Ala-Ala-Pro-Phe-Nme, was allowed to fluctuate in the trans (Ctrans), transition (CTS) and cis (Ccis) states (SI Text, section 1.2). No significant differences were observed in either the distribution of backbone dihedral angles (Fig. S5) or amplitudes of backbone amide bonds in the Co versus bound-CypA or between Ctrans, CTS, and Ccis (Fig. S7). However, the distribution of side-chain dihedrals of the active site residues (Fig. S6) and the fluctuations of selected side-chain bonds (Fig. S8) showed notable differences in the absence and the presence of the substrate. Prominently, we found that the side-chain rotamers that were preferred in the bound state are already sampled in Co and there is simply a redistribution of rotameric population in the bound-CypA as compared to Co. Conforming to earlier studies (37, 38), the catalytically important Arg55, which have been shown to form stabilizing electrostatic interactions with the substrate in the transition state, exhibited side-chain motions that were of smaller amplitudes and much more restricted in CTS than the end states (Figs. S6 and S8). The trajectories of Co, Ctrans, CTS, and Ccis were analyzed using principal component analysis (PCA) that decomposes the fluctuations in the atomic coordinates into modes ranked according to their relative contribution to the overall protein motion (SI Text, section 1.6). Projection of the trajectories onto the first three modes that accounted for 90% of the total fluctuations resulted in two-dimensional representation of the multidimensional phase space (Fig. 4 A to C). The conformational space sampled by the active site residues in Co was not only much larger than that in Ctrans, CTS, and Ccis, but also showed considerable overlap with them, indicating that certain fluctuations observed in the enzyme in the presence of the substrate are already preserved in the free enzyme. CTS occupied a much more restricted conformational space as compared to Ctrans and Ccis, in agreement with the reduction in conformational heterogeneity observed in side chains of active site residues in CTS; i.e., mainly a single side-chain rotamer is populated for all active site residues in CTS (Fig. S6). Also, as compared to Co, Ctrans, or Ccis, the side-chain fluctuations in CTS showed decreased amplitudes; i.e., higher S2 values (Fig. S8). As seen in Fig. 4, the smaller ensemble of CTS consisted of conformations that were a subset of both Ctrans and Ccis ensembles, each of which also had its own unique set of conformations. Such connectivity between the conformations of the ensembles is important for multiple numbers of possible catalytic pathways. The above results clearly affirmed the picture of conformational selection (39) wherein the substrate prefers to bind a subset of enzyme conformations and upon catalysis there is a shift toward the population of enzyme conformations bound to the product.
Fig. 4.
Conformational selection and transition state stabilization by CypA. (A, B, C) Projection of the multidimensional Cartesian space onto the first three principal components. Shown are the configurations of the binding site residues (see SI Text, section 1.6) resulting from the snapshots of a 300-ns long trajectory of free CypA (black) and 50-ns long trajectory each of Ctrans (blue), CTS (green), Ccis, (red) complexes. (D) Probability distribution of binding free energies, ΔGbind, of Ctrans, CTS, and Ccis with average values of -20.8, -34.1, and -24.3 kcal/mol, respectively. The color code is the same as in A–C (Inset) Binding site of CypA (gray VDW) in CTS where the substrate makes hydrogen bonds (dashed green lines) with active site residues.
Estimating Binding Free Energies of the CypA-substrate Complexes
We estimated the binding free energies from each snapshot of Ctrans, CTS, and Ccis (Fig. 4D and SI Text, section 1.7). In support of earlier studies, Fig. 4D revealed that CypA has maximum affinity for the substrate in the transition state and binds the cis isomer more favorably than the trans isomer (23, 38). The ability of CypA to bind transition states better than the ground states resulted in transition state stabilization by approximately 10-13 kcal/mol greater than the cis and the trans isomers, respectively (Fig. 4D). The estimates of binding free energy were, however, approximate and did not include contributions from the changes in configurational entropy (SI Text, section 1.7). As seen in Fig. S4 and Table 1, we indeed found that the trans to cis free-energy barrier height for the catalyzed isomerization of -Ala-Pro- ω bond was reduced by approximately 9.3 kcal/mol. Thus, the transition state was stabilized considerably greater as compared to that of the cis or the trans isomers and this alone accounted for a speed up of approximately 6 × 106 in isomerization rate if the same preexponential factor was assumed for RO and RC. However, we showed Deff for RC to be approximately 13 times smaller than that of RO. Our findings therefore also reinforce the idea of selective binding and transition state stabilization as the major barrier effect in enhancing the rate of catalysis.
Conclusions
Although a growing body of experimental data suggests that enzyme motions play an important role in its catalytic function, the exact nature of this dynamical contribution has never been explained earlier. Here, we show that substrate dynamics involved in the chemical step, is coupled to the dynamics of the surrounding medium which can either be the solvent or the active site of the enzyme. If the environment relaxes much faster than the chemical step such that there is a clear separation of timescales, then the motions are not coupled; e.g., second-timescale cis-trans isomerization in aqueous solution. Provided the timescale of the chemical step falls within that of the fluctuations in the environment, not only the motions on the same timescale but also those that are slightly faster and slower than the chemical step get coupled to substrate dynamics and result in multiexponential kinetic decays. Since our studies were carried out in the regime of lower free-energy barriers in which the timescale of substrate dynamics was shifted to nanoseconds, we could observe the coupling with nanosecond-motions of the enzyme. But for the actual cis-trans isomerization that involves larger barriers and takes place in milliseconds in CypA, increasingly slower modes of the enzyme clustered in the millisecond-timescale will be selected for coupling with substrate dynamics. We further show that one-dimensional Kramers’ theory in the high friction regime, which has proved sufficiently valid in analyzing and explaining the kinetics of various problems of biological interest—which includes protein folding (40–43) and cis-trans isomerization in peptides (25, 29), can be equally useful to interpret the kinetics of enzyme-catalyzed chemical step and understand the role of enzyme dynamics. The enzymatic motions do not and cannot modify the free energy landscape of the enzyme-substrate complex, but rather reduce the effective diffusion coefficient as compared to the reference nonenzymatic reaction in solution. These effects that are incorporated in the preexponential factor of the chemical reaction reduce the speedup possible just from barrier effects and, as we show, modify the diffusion coefficient by more than an order of magnitude. Therefore, as often implied, enzyme dynamics do not accelerate the chemical step. Enzymatic motions are important for reorganization of the active site so that the transition state is better stabilized. However, enzyme dynamics does not have to enhance the catalytic rates to be important to the chemical step. Perturbation of conformational motions caused due to mutations or acceleration affects the catalytic rates resulting from modification of both—the free energy profile and the effective diffusion coefficient. In any case, it is the alteration of free-energy barriers that remains the dominant effect in either impeding or enhancing catalytic rates. Our results provide the missing link between assertions made from theoretical studies and observations from experimental studies, thereby unifying apparently disparate views about the role of enzyme dynamics in catalysis.
Methods
We carried out an extensive characterization of CypA dynamics (including backbone of all residues and side chains of active site residues) in the absence and presence of its substrate using all-atom nMD (SI Text). In order to access long-timescale motions, we further subjected CypA to aMD. For details on aMD methodology, see SI Text. We then characterized the isomerization reaction in solution and in enzyme both kinetically and thermodynamically. The parameter (V2) in the AMBER force field is the force constant in the dihedral energy functional
where n, ω, and γ are the periodicity, dihedral angle, and phase angle, respectively. V2 (for which n = 2 and γ = 180°) predominantly controls the rotational barrier only around the peptidyl-prolyl bonds with almost isoenergetic cis and trans states. Changing V2 modifies the total potential and hence the free-energy barriers of isomerization. The default or reoptimized (35) value of V2 would yield such high barriers that sufficient number of trans-cis transitions will not be observed in a standard MD trajectory of even several hundred nanoseconds. Therefore, to make the simulation of kinetics of cis-trans isomerization feasible and obtain reasonable statistics, we shifted the isomerization timescale from seconds to nanoseconds by reducing V2 to lower values of 11.0, 9.0, 7.0, 5.0, 4.0, and 0 kcal/mol. It should be noted that V2 = 0 kcal/mol does not imply zero torsional barrier. With each value of V2, we performed individual nMD simulations on the free solvated substrate to model the uncatalyzed reaction in solution as well as on the CypA-substrate complex to model the catalyzed isomerization in the active site of the enzyme. In a different set of simulations, the CypA-bound substrate was simulated with nMD with V2 set to 7.0 kcal/mol while the enzyme was subjected to three increasing levels of acceleration (SI Text, section 1.3). For each case, free energy profiles projected onto the ω dihedral were calculated from umbrella sampling (SI Text, section 1.5). ΔG#, ωo, and ωb were estimated from such one-dimensional free energy profiles (Fig. S3). The probability of survival in the trans well for time t or longer was then calculated from the distribution of dwell times, p(τ), as follows:
. We analyzed our results using the DISCRETE (44, 45) software program that provides nonlinear least square solution of multiexponential decays without any a priori guesses for the number of exponentials or initial parameters. The survival probability function, S(t), was fitted to a sum of discrete multiexponential decays; i.e.,
. The average rate constant
where the average lifetime
. Ai and τi are amplitudes and time constants, respectively, of phase i in the above function. Using the MM/PBSA (SI Text, section 1.7) method, we further calculated the binding free energies of Ctrans, CTS, and Ccis to estimate the contribution of transition state stabilization to the speedup in the isomerization rate.
Supplementary Material
ACKNOWLEDGMENTS.
This work was supported in part by the National Science Foundation CAREER Grant MCB- 0953061, the Georgia Cancer Coalition (GCC) scholar award and Georgia State University. This work was also supported by Georgia State’s IBM System p5 supercomputer, acquired through a partnership of the Southeastern Universities Research Association and IBM supporting the SURAgrid initiative. We thank Drs W. David Wilson, Stuart Allison, and Tongye Shen for helpful discussions during the preparation of this manuscript.
Footnotes
The authors declare no conflict of interest.
*This Direct Submission article had a prearranged editor.
This article contains supporting information online at www.pnas.org/lookup/suppl/doi:10.1073/pnas.1117060109/-/DCSupplemental.
References
- 1.Wolfenden R, Snider MJ. The depth of chemical time and the power of enzymes as catalysts. Acc Chem Res. 2001;34:938–945. doi: 10.1021/ar000058i. [DOI] [PubMed] [Google Scholar]
- 2.Benkovic SJ, Hammes-Schiffer S. A perspective on enzyme catalysis. Science. 2003;301:1196–1202. doi: 10.1126/science.1085515. [DOI] [PubMed] [Google Scholar]
- 3.Warshel A, et al. Electrostatic basis for enzyme catalysis. Chem Rev. 2006;106:3210–3235. doi: 10.1021/cr0503106. [DOI] [PubMed] [Google Scholar]
- 4.Pisliakov AV, Cao J, Kamerlin SC, Warshel A. Enzyme millisecond conformational dynamics do not catalyze the chemical step. Proc Natl Acad Sci USA. 2009;106:17359–17364. doi: 10.1073/pnas.0909150106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 5.Kamerlin SCL, Warshel A. At the dawn of the 21st century: Is dynamics the missing link for understanding enzyme catalysis? Proteins. 2010;78:1339–1375. doi: 10.1002/prot.22654. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Karplus M. The role of conformation transitions in adenylate kinase. Proc Natl Acad Sci USA. 2010;107:E71. doi: 10.1073/pnas.1002180107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Warshel A. Reply to Karplus: Conformational dynamics have no role in the chemical step. Proc Natl Acad Sci USA. 2010;107:E72. [Google Scholar]
- 8.Garcia-Viloca M, Gao J, Karplus M, Truhlar DG. How enzymes work: Analysis by modern rate theory and computer simulations. Science. 2004;303:186–195. doi: 10.1126/science.1088172. [DOI] [PubMed] [Google Scholar]
- 9.Agarwal PK, Geist A, Gorin A. Protein dynamics and enzymatic catalysis: Investigating the peptidyl-prolyl cis-trans isomerization activity of cyclophilin A. Biochemistry. 2004;43:10605–10618. doi: 10.1021/bi0495228. [DOI] [PubMed] [Google Scholar]
- 10.Adamczyk AJ, Cao J, Kamerlin SC, Warshel A. Catalysis by dihydrofolate reductase and other enzymes arises from electrostatic preorganization, not conformational motions. Proc Natl Acad Sci USA. 2011;108:14115–14120. doi: 10.1073/pnas.1111252108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Bhabha G, et al. A dynamic knockout reveals that conformational fluctuations influence the chemical step of enzyme catalysis. Science. 2011;332:234–238. doi: 10.1126/science.1198542. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 12.Eisenmesser EZ, Bosco DA, Akke M, Kern D. Enzyme dynamics during catalysis. Science. 2002;295:1520–1523. doi: 10.1126/science.1066176. [DOI] [PubMed] [Google Scholar]
- 13.Eisenmesser EZ, et al. Intrinsic dynamics of an enzyme underlies catalysis. Nature. 2005;438:117–121. doi: 10.1038/nature04105. [DOI] [PubMed] [Google Scholar]
- 14.Min W, et al. Fluctuating enzymes: Lessons from single-molecule studies. Acc Chem Res. 2005;38:923–931. doi: 10.1021/ar040133f. [DOI] [PubMed] [Google Scholar]
- 15.Hammes-Schiffer S, Benkovic SJ. Relating protein motion to catalysis. Annu Rev Biochem. 2006;75:519–541. doi: 10.1146/annurev.biochem.75.103004.142800. [DOI] [PubMed] [Google Scholar]
- 16.Henzler-Wildman K, Kern D. Dynamic personalities of proteins. Nature. 2007;450:964–972. doi: 10.1038/nature06522. [DOI] [PubMed] [Google Scholar]
- 17.Frauenfelder H, Sligar SG, Wolynes PG. The energy landscapes and motions of proteins. Science. 1991;254:1598–1603. doi: 10.1126/science.1749933. [DOI] [PubMed] [Google Scholar]
- 18.Wales DJ. Energy Landscapes. Cambridge, UK: Cambridge Univ Press; 2003. [Google Scholar]
- 19.Frauenfelder H, et al. A unified model of protein dynamics. Proc Natl Acad Sci USA. 2009;106:5129–5134. doi: 10.1073/pnas.0900336106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 20.Hamelberg D, Mongan J, McCammon JA. Accelerated molecular dynamics: A promising and efficient simulation method for biomolecules. J Chem Phys. 2004;120:11919–11929. doi: 10.1063/1.1755656. [DOI] [PubMed] [Google Scholar]
- 21.Grathwohl C, Wuthrich K. NMR studies of the rates of Proline Cis-Trans Isomerization in Oligopeptides. Biopolymers. 1981;20:2623–2633. [Google Scholar]
- 22.Beausoleil E, Lubell WD. Steric effects on the amide isomer equilibrium of prolyl peptides. Synthesis and conformational analysis of N-acetyl-5-tert-butylproline N’-methylamides. J Am Chem Soc. 1996;118:12902–12908. [Google Scholar]
- 23.Kern D, Kern G, Scherer G, Fischer G, Drakenberg T. Kinetic analysis of cyclophilin-catalyzed prolyl cis/trans isomerization by dynamic NMR spectroscopy. Biochemistry. 1995;34:13594–13602. doi: 10.1021/bi00041a039. [DOI] [PubMed] [Google Scholar]
- 24.Harrison RK, Stein RL. Mechanistic studies of enzymic and nonenzymic prolyl cis-trans isomerization. J Am Chem Soc. 1992;114:3464–3471. [Google Scholar]
- 25.Doshi U, Hamelberg D. Extracting realistic kinetics of rare activated processes from accelerated molecular dynamics using Kramers’ theory. J Chem Theo Comp. 2011;7:575–581. doi: 10.1021/ct1005399. [DOI] [PubMed] [Google Scholar]
- 26.Markwick PR, Bouvignies G, Blackledge M. Exploring multiple timescale motions in protein GB3 using accelerated molecular dynamics and NMR spectroscopy. J Am Chem Soc. 2007;129:4724–4730. doi: 10.1021/ja0687668. [DOI] [PubMed] [Google Scholar]
- 27.Markwick PR, et al. Toward a unified representation of protein structural dynamics in solution. J Am Chem Soc. 2009;131:16968–16975. doi: 10.1021/ja907476w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Markwick PR, et al. Enhanced conformational space sampling improves the prediction of chemical shifts in proteins. J Am Chem Soc. 2010;132:1220–1221. doi: 10.1021/ja9093692. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.Xin Y, Doshi U, Hamelberg D. Examining the limits of time reweighting and Kramers’ rate theory to obtain correct kinetics from accelerated molecular dynamics. J Chem Phys. 2010;132:224101. doi: 10.1063/1.3432761. [DOI] [PubMed] [Google Scholar]
- 30.Karplus M, Kuriyan J. Molecular dynamics and protein function. Proc Natl Acad Sci USA. 2005;102:6679–6685. doi: 10.1073/pnas.0408930102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Min W, Xie XS, Bagchi B. Two-dimensional reaction free energy surfaces of catalytic reaction: Effects of protein conformational dynamics on enzyme catalysis. J Phys Chem B. 2008;112:454–466. doi: 10.1021/jp076533c. [DOI] [PubMed] [Google Scholar]
- 32.Kramers HA. Brownnian motion in a field of force and diffusion model of chemical reactions. Physica (Utrecht) 1940;7:284–304. [Google Scholar]
- 33.Hanggi P, Talkner P, Borkovec M. Reaction-rate theory: Fifty years after Kramers. Rev Mod Phys. 1990;62:251–342. [Google Scholar]
- 34.Johnson Q, Doshi U, Shen T, Hamelberg D. Water’s contribution to the energetic roughness from peptide dynamics. J Chem Theo Comp. 2010;6:2591–2597. doi: 10.1021/ct100183s. [DOI] [PubMed] [Google Scholar]
- 35.Doshi U, Hamelberg D. Re-optimization of the AMBER force field parameters for peptide bond (Omega) torsions using accelerated molecular dynamics. J Phys Chem B. 2009;113:16590–16595. doi: 10.1021/jp907388m. [DOI] [PubMed] [Google Scholar]
- 36.Fraser JS, et al. Hidden alternative structures of proline isomerase essential for catalysis. Nature. 2009;462:669–673. doi: 10.1038/nature08615. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Cui Q, Li G. What is so special about Arg 55 in the catalysis of cyclophilin A? insights from hybrid QM/MM simulations. J Am Chem Soc. 2003;125:15028–15038. doi: 10.1021/ja0367851. [DOI] [PubMed] [Google Scholar]
- 38.Hamelberg D, McCammon JA. Mechanistic insight into the role of transition-state stabilization in cyclophilin A. J Am Chem Soc. 2009;131:147–152. doi: 10.1021/ja806146g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Ma B, Nussinov R. Enzyme dynamics point to stepwise conformational selection in catalysis. Curr Opin Chem Biol. 2010;14:652–659. doi: 10.1016/j.cbpa.2010.08.012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Doshi U, Munoz V. Kinetics of alpha-helix formation as diffusion on a one-dimensional free energy surface. Chem Phys. 2004;307:129–136. [Google Scholar]
- 41.Naganathan AN, Doshi U, Munoz V. Protein folding kinetics: Barrier effects in chemical and thermal denaturation experiments. J Am Chem Soc. 2007;129:5673–5682. doi: 10.1021/ja0689740. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Plaxco KW, Baker D. Limited internal friction in the rate-limiting step of a two-state protein folding reaction. Proc Natl Acad Sci USA. 1998;95:13591–13596. doi: 10.1073/pnas.95.23.13591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Plotkin SS. Determination of barrier heights and prefactors from protein folding rate data. Biophys J. 2005;88:3762–3769. doi: 10.1529/biophysj.104.052548. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Provencher SW. A Fourier method for the analysis of exponential decay curves. Biophys J. 1976;16:27–41. doi: 10.1016/S0006-3495(76)85660-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Provencher SW. An eigenfunction expansion method for the analysis of exponential decay curves. J Chem Phys. 1976;64:2772–2777. doi: 10.1016/S0006-3495(76)85660-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.






