The kinetics of protein folding is often remarkably simple. For

Similar documents
Master equation approach to finding the rate-limiting steps in biopolymer folding

arxiv:cond-mat/ v1 [cond-mat.soft] 19 Mar 2001

Elucidation of the RNA-folding mechanism at the level of both

Many proteins spontaneously refold into native form in vitro with high fidelity and high speed.

Temperature dependence of reactions with multiple pathways

arxiv:cond-mat/ v1 [cond-mat.soft] 16 Nov 2002

A simple model for calculating the kinetics of protein folding from three-dimensional structures

Identifying the Protein Folding Nucleus Using Molecular Dynamics

Prediction of protein-folding mechanisms from free-energy landscapes derived from native structures

Stretching lattice models of protein folding

Local Interactions Dominate Folding in a Simple Protein Model

Visualizing folding of proteins (1 3) and RNA (2) in terms of

Effective stochastic dynamics on a protein folding energy landscape

arxiv:cond-mat/ v1 2 Feb 94

Folding of small proteins using a single continuous potential

arxiv:cond-mat/ v1 [cond-mat.soft] 5 May 1998

Protein Folding In Vitro*

Short Announcements. 1 st Quiz today: 15 minutes. Homework 3: Due next Wednesday.

Scattered Hammond plots reveal second level of site-specific information in protein folding: ( )

PHYSICAL REVIEW LETTERS

Protein Folding. I. Characteristics of proteins. C α

To understand pathways of protein folding, experimentalists

Outline. The ensemble folding kinetics of protein G from an all-atom Monte Carlo simulation. Unfolded Folded. What is protein folding?

letters Transition states and the meaning of Φ-values in protein folding kinetics S. Banu Ozkan 1,2, Ivet Bahar 3 and Ken A.

Entropic barriers, transition states, funnels, and exponential protein folding kinetics: A simple model

Protein Folding Prof. Eugene Shakhnovich

The role of secondary structure in protein structure selection

The starting point for folding proteins on a computer is to

Lecture 34 Protein Unfolding Thermodynamics

Pathways for protein folding: is a new view needed?

Simulating disorder order transitions in molecular recognition of unstructured proteins: Where folding meets binding

Protein folding is a fundamental problem in modern structural

Universal correlation between energy gap and foldability for the random energy model and lattice proteins

Reliable Protein Folding on Complex Energy Landscapes: The Free Energy Reaction Path

PROTEIN FOLDING THEORY: From Lattice to All-Atom Models

Simulation of mutation: Influence of a side group on global minimum structure and dynamics of a protein model

Nature of the transition state ensemble for protein folding

CHRIS J. BOND*, KAM-BO WONG*, JANE CLARKE, ALAN R. FERSHT, AND VALERIE DAGGETT* METHODS

Solvation in protein folding analysis: Combination of theoretical and experimental approaches

Exploring protein-folding ensembles: A variable-barrier model for the analysis of equilibrium unfolding experiments

Modeling protein folding: the beauty and power of simplicity Eugene I Shakhnovich

It is not yet possible to simulate the formation of proteins

Protein Folding experiments and theory

The folding mechanism of larger model proteins: Role of native structure

Computer simulation of polypeptides in a confinement

Molecular Interactions F14NMI. Lecture 4: worked answers to practice questions

Protein Folding Pathways and Kinetics: Molecular Dynamics Simulations of -Strand Motifs

Clustering of low-energy conformations near the native structures of small proteins

PROTEIN EVOLUTION AND PROTEIN FOLDING: NON-FUNCTIONAL CONSERVED RESIDUES AND THEIR PROBABLE ROLE

Determination of Barrier Heights and Prefactors from Protein Folding Rate Data

arxiv:cond-mat/ v1 [cond-mat.soft] 23 Mar 2007

Hairpin loops are ubiquitous in single-stranded DNA and

Protein folding in the cell occurs in a crowded, heterogeneous

arxiv:chem-ph/ v1 11 Nov 1994

Convergence of replica exchange molecular dynamics

Engineering of stable and fast-folding sequences of model proteins

Intermediates and the folding of proteins L and G

Toward an outline of the topography of a realistic proteinfolding funnel

Folding pathway of a lattice model for protein folding

Thermodynamics. Entropy and its Applications. Lecture 11. NC State University

Research Paper 577. Correspondence: Nikolay V Dokholyan Key words: Go model, molecular dynamics, protein folding

Random heteropolymer adsorption on disordered multifunctional surfaces: Effect of specific intersegment interactions

1 of 31. Nucleation and the transition state of the SH3 domain. Isaac A. Hubner, Katherine A. Edmonds, and Eugene I. Shakhnovich *

Energy landscapes, global optimization and dynamics of the polyalanine Ac ala 8 NHMe

Quiz 2 Morphology of Complex Materials

How protein thermodynamics and folding mechanisms are altered by the chaperonin cage: Molecular simulations

Universality and diversity of folding mechanics for three-helix bundle proteins

Energy Landscape Distortions and the Mechanical Unfolding of Proteins

Two-State Folding over a Weak Free-Energy Barrier

Single-molecule spectroscopy has recently emerged as a powerful. Effects of surface tethering on protein folding mechanisms

Frontiers in Physics 27-29, Sept Self Avoiding Growth Walks and Protein Folding

arxiv: v1 [cond-mat.stat-mech] 10 Jul 2018

Isothermal experiments characterize time-dependent aggregation and unfolding

Nucleation and the Transition State of the SH3 Domain

FREQUENCY selected sequences Z. No.

Guessing the upper bound free-energy difference between native-like structures. Jorge A. Vila

Paul Sigler et al, 1998.

Evolution of functionality in lattice proteins

Effects of Crowding and Confinement on the Structures of the Transition State Ensemble in Proteins

FOCUS: HYDROGEN EXCHANGE AND COVALENT MODIFICATION

Protein folding can be described by using a free energy

arxiv:q-bio/ v1 [q-bio.bm] 11 Jan 2004

Sampling activated mechanisms in proteins with the activation relaxation technique

Small organic molecules in aqueous solution can have profound

Recovering Kinetics from a Simplified Protein Folding Model Using Replica Exchange Simulations: A Kinetic Network and Effective Stochastic Dynamics

Supplementary Figures:

Effect of Sequences on the Shape of Protein Energy Landscapes Yue Li Department of Computer Science Florida State University Tallahassee, FL 32306

The protein folding problem consists of two parts:

Transport of single molecules along the periodic parallel lattices with coupling

It is now well established that proteins are minimally frustrated

The chemical environment exerts a fundamental influence on the

The effort to understand how proteins fold has consumed the

Can a continuum solvent model reproduce the free energy landscape of a β-hairpin folding in water?

Protein Folding & Stability. Lecture 11: Margaret A. Daugherty. Fall How do we go from an unfolded polypeptide chain to a

Protein molecules rapidly fold into a unique, native structure

Transient UV Raman Spectroscopy Finds No Crossing Barrier between the Peptide R-Helix and Fully Random Coil Conformation

Biological Thermodynamics

Monte Carlo Study of Substrate-Induced Folding and Refolding of Lattice Proteins

Protein folding. Today s Outline

The strength of the hydrogen bond in the linking of protein

Transcription:

Fast protein folding kinetics Jack Schonbrun* and Ken A. Dill *Graduate Group in Biophysics and Department of Pharmaceutical Chemistry, University of California, San Francisco, CA 94118 Communicated by George H. Lorimer, University of Maryland, College Park, MD, August 22, 2003 (received for review December 9, 2002) Proteins are complex molecules, yet their folding kinetics is often fast (microseconds) and simple, involving only a single exponential function of time (called two-state kinetics). The main model for two-state kinetics has been transition-state theory, where an energy barrier defines a slow step to reach an improbable structure. But how can barriers explain fast processes, such as folding? We study a simple model with rigorous kinetics that explains the high speed instead as a result of the microscopic parallelization of folding trajectories. The single exponential results from a separation of timescales; the parallelization of routes is high at the start of folding and low thereafter. The ensemble of rate-limiting chain conformations is different from in transition-state theory; it is broad, overlaps with the denatured state, is not aligned along a single reaction coordinate, and involves well populated, rather than improbable, structures. The kinetics of protein folding is often remarkably simple. For many proteins, both folding (from the denatured state D to the native state N) and unfolding processes are singleexponential functions of time (1 5). Combined with the observation that the ratio of the forward to reverse rate constants equals the equilibrium constant, folding is often described in terms of a two-state mass-action model, D 7 N, and a corresponding Arrhenius diagram (Fig. 1). The main ideas underlying such Arrhenius diagrams are the following: (i) a single reaction coordinate exists from D to N; (ii) relatively distinct molecular structures appear in sequential order along the reaction coordinate, e.g., D 3 TS 3 N; and (iii) the single-exponential behavior arises from a bottleneck called the transition state (TS) (6), which also appears in Kramers theory for reactions in solution (7), resulting from a slow step to an improbable structure. Barriers originated to explain why chemical and physical processes are slower than a speed limit such as k B T h 0.16 psec, where k B is Boltzmann s constant, T is temperature, and h is Planck s constant. But for protein folding, this macroscopic description does not give microscopic insight. Mass-action symbols such as D and TS define macrostates, or ensembles of chain conformations. What chain conformations, i.e., microstates, constitute those ensembles? Two main issues exist. First, barriers explain slowing down relative to an upper speed limit, but the challenge for proteins is to explain why folding is so fast, not why it is slow. Hundreds to thousands of bonds must rotate for a typical denatured conformation to become native, fine details of packing must be satisfied, and each rotation probably requires 1 100 psec, yet some proteins can fold in as little as 10 sec (2, 4). The key to understanding the mechanism of protein-folding kinetics is in explaining how this happens so quickly. Second, because the many denatured conformations are so different from each other, a single microscopic reaction coordinate, or trajectory, cannot lead to the folded state. Folding kinetics has been studied by microscopic theory and simulations (8 18). This type of modeling shows that protein folding can be described in terms of funnel-shaped energy landscapes (19 21) (see Fig. 1b). The depth represents the interaction free energy of the chain in fixed configurations, and the width represents the chain entropy. At deeper levels, the landscape represents lowerenergy, more native-like structures; the funnel shape results because the more native-like the energy, the fewer the conformations (22). The funnel concept shows how a single native structure can emerge rapidly from an astronomical number of different denatured starting conformations, but funnel pictures are often shown without a barrier. It has been suggested that if all the steps were energetically downhill, folding kinetics would not be two-state (8, 17). In short, TS theory rationalizes two-state kinetics but not the speed, whereas funnel cartoons rationalize the speed but not the two-state kinetics. How can we understand both behaviors? Experiments are not yet able to observe the microscopic folding processes. Experiments see only ensemble-averaged properties. Atomically detailed computer simulations give valuable insights (23 25), but they explore only a relatively small sampling of the conformational space. Currently, questions of the full microscopic folding kinetics can only be addressed by using low-resolution models of the type used here. Here, we generate and solve the master equation for the folding of the 2D lattice model. We use a Go potential (26). Go models are widely used for studying two-state protein folding (27 30), because such models have two-state kinetics for folding and unfolding, unique lowest-energy native states, and exponentially growing conformational searches with chain length. Restricting the conformations to a 2D lattice reduces the conformational space so that the complete solution is tractable, without compromising any of the general properties of proteins described above (see Methods). To obtain the kinetics, we follow the eigenvector method of Banavar and colleagues (31). We find the exact solution of the master equation for folding in terms of its eigenvalues and eigenvectors to reveal the details of the rate-limiting transitions. From this model, we reach three conclusions about the microscopic basis for fast-folding kinetics. Methods We use a 2D lattice formulation of the Go model. The conformation-to-conformation transition rate is given by k ij exp( B r), where is the number of bond angles that differ between structures i and j, and r is the sum of the squares of the displacements of each bead of the chain. The parameters B and correspond to the energies for dihedral angle rotation and diffusion through the solvent, respectively. Values of B 0.5 and 0.1 give transition probabilities similar to those in traditional Monte Carlo dynamics move sets. The transition rates are multiplied by a Metropolis factor: the lesser of exp( H ij )or1, where H ij is the difference in energy between the first and second conformations. In a Go model this energy is determined solely by the number of native contacts. The evolution of the ensemble with time is governed by a master equation (31, 32): dp(t) dt Kp(t), where p is a vector that contains the population of each conformation at each instant, and K is a matrix that contains the transition probability (per unit time) between each conformation. The elements of this Abbreviation: TS, transition state. Present address: Department of Biochemistry, University of Washington, Seattle, WA 98195. To whom correspondence should be addressed. E-mail: dill@maxwell.compbio.ucsf.edu. 2003 by The National Academy of Sciences of the USA 12678 12682 PNAS October 28, 2003 vol. 100 no. 22 www.pnas.org cgi doi 10.1073 pnas.1735417100

Fig. 1. (a) Transition-state explanation of single-exponential processes, such as protein folding, involves a rate-limiting step, shown as an obligatory thermodynamic barrier. (b) Theory and simulations show that energy landscapes for protein folding are funnel-shaped and have no apparent microscopic energetic or entropic barriers. matrix are the weighted k ij values described above. The dynamical evolution of this entire ensemble can be solved exactly in terms of the eigenvectors and eigenvalues of the matrix K by using readily available numerical algorithms (33). (The results in this article are based on a chain of nine residues, which leads to a tractable 750-by-750 matrix for the master equation.) The smallest nonzero eigenvalue corresponds to the rate constant for the onset of the native state. Its corresponding eigenvector represents all the relative contributions of the microscopic transitions that contribute to the slowest rate-limiting step. Hence, this eigenvector defines the TS macrostate. Results Fast Folding Results from Parallel Microscopic Routes. Fig. 2 compares series vs. parallel kinetic processes. Fig. 2 Right shows the bottleneck principle: if microscopic steps are in series, one step is rate-limiting. An example would be a high-energy protein conformation. In this model, the folding flux (number of proteins folded per unit time) would be slower than the slowest microscopic step, the flux through the bottleneck step. In contrast, Fig. 2 Left shows that in a parallel process, the folding flux is faster than the fastest microscopic step. Fig. 3 shows that in the fast early steps, our simulations most closely resemble the parallel model. If the rate of a microscopic step is k ij from conformation i to j and p i is the population of i, then the flux from i to j is p i k ij. But if multiple routes go to j, the flux to j is higher, i p i k ij, because of the parallelization of routes. A previous argument [the Levinthal paradox (34)] suggested that the folding search for the native structure should be slowest at the earliest stages, because the chain has so many high-energy (open, denatured) conformations to explore. Our model shows the opposite: the chain wastes the least amount of its folding time in the high-energy states. The large numbers conspire to speed up folding, not to slow it down. Fig. 3 shows this also in another way. The energy level-to-level rates are faster than the conformation-to-conformation rates, because the levelto-level transitions include many parallel routes. Hence, we believe that early stages of folding are not well described by the term search, for the same reasons that water moving down a hillside is more akin to a highly directed flow, and less like a search. The Separation of Timescales Results from Diminishing Multiplicities of Routes Down the Landscape. Zwanzig (16) has noted that single-exponential behavior can result from a separation of timescales. In our model, the single-exponential dynamics does not result from a microscopic barrier (i.e., a barrier along only one specific trajectory). Rather, we observe that the separation of timescales results from the funnel-shaped route structure. This route structure is a property of the whole landscape, not of a single trajectory. In funnel landscapes, the bottleneck is relatively low where the funnel becomes narrow. Consider two different conformations: one is high-energy, Fig. 2. Series vs. parallel models of kinetics. (Right) When microscopic steps are in series, the macroscopic flux is limited by the microscopic bottleneck rate. The hypothetical numbers illustrate the fluxes through different microscopic steps. (Left) But if parallel microscopic routes exist, the macroscopic flux is greater than the largest microscopic flux. BIOPHYSICS Schonbrun and Dill PNAS October 28, 2003 vol. 100 no. 22 12679

native contact, because the vast majority of conformations accessible to it are uphill and more unfolded. It has few routes downhill. Zwanzig has called such near-native conformations gateway structures (16). When folding is initiated, fast rearrangements take place to reach the gateway conformations, then slow steps from there to the native state. Gateway conformations act like transition states in the sense that they are obligatory and rate-limiting, but they differ in a fundamental way from transition states: they are not rare, improbable, or high in free energy. Rather, the gateway conformations are among the most heavily populated of the nonnative conformations because they are the funnelnecks for the microroutes. Hence two-state kinetics need not be caused by a structure of low population that is reached slowly. It can also be caused by a state having few exit routes that is reached quickly. Fig. 3. The rate of folding is faster than almost all the microscopic transitions between conformations, supporting the parallel model. The large arrow shows the overall folding rate. The histogram to the right shows the distribution of microscopic transition fluxes. The small arrows on the left show the level-to-level transition rates between energy levels: 0 1 indicates transitions at the top of the energy funnel, 1 2 is the next level down, etc. The level-tolevel fluxes are high because many microroutes contribute to them. Transition-State Conformations Overlap with Denatured Conformations. What are the rate-limiting conformations in the funnel model? Four macrostates in the model are relevant: N, the single native conformation; D u, the Boltzmann-weighted denatured ensemble under the unfolding conditions from which folding is initiated; D f, the Boltzmann-weighted nonnative ensemble under nonnative and the other is low-energy, nearly native. The unfolded conformation has a much higher probability per unit time of making a new native contact to proceed downhill, simply because it can happen so many different ways. In contrast, the near-native conformation is slower to make a new Fig. 4. (Left) The transition-state model involves macrostates (R, reactant; TS, transition state; P, product) that are small, localized ensembles of microstates. Each microscopic structure is a member of only one macrostate. (Right) The present model shows that the macrostates for two-state protein folding (D), the denatured state, or TS, the transition state under folding conditions) are broad ensembles that encompass all the nonnative chain conformations. A given chain conformation is not a unique member of only one macrostate. Fig. 5. (a) Chevron plot showing predicted relaxation rate (k k f k u ) versus strength of native contacts. The histograms show the distribution of conformations (most native-like to the left) under different folding conditions for the denatured ensemble D f (b) and the apparent transition state TS f (c). Under increasingly native-like conditions, both ensembles D f and TS f become more native-like. Hence, the left (folding branch) of the chevron curve shows rollover, a deviation from linearity. Under strong native conditions, the D f ensemble is more native-like than the TS f ensemble. 12680 www.pnas.org cgi doi 10.1073 pnas.1735417100 Schonbrun and Dill

particular folding conditions; and TS f (the apparent transition state under folding conditions), the eigenvector of conformations that fold to the native state having the lowest eigenvalue relaxation time. In transition-state theories, the macrostates, such as reactants, intermediates, transition states, or products, are all identifiable as distinct ensembles of microstates (specific atomic structures) that occur along a single reaction coordinate. Along any such 1D coordinate, a state A is either before or after some other state B. Applied to proteins, it implies the transition state is between, and distinct from, the denatured state (D) and the native state (N): D 3 TS 3 N. That description is macroscopic, however. In chemical rate theory, the observed kinetics (macroscopic behavior) has a microscopic explanation in terms of structure: the bottlenecks are specific atomic structures along a sequence of structures from reactant to product. For chemical reactions, macro- and microdescriptions coincide. But for folding, the present model microdescription of folding bears little resemblance to the macrodescription. No single trajectory, reaction coordinate or specific sequence of the microstates exist. Fig. 4 Right is a cartoon showing how the TS f ensemble could be so broad that it includes many of the same conformations that are in D u and D f. Indeed, this finding is observed in our simulations (Fig. 5). No individual chain conformation is exclusively in one state D u, D f,orts f, without being in the others at the same time. These ensembles differ only in the statistical weights of the individual conformations. The heterogeneity of the rate-limiting steps has also been observed in other simulations (23, 35, 36). The folding process is the evolution of a single ensemble, not a time sequence of distinct ensembles, such as the pouring of D into TS followed by the pouring of TS into N. In TS theory models, the transition state is between the reactant and the product. For folding, this placement would mean D 3 TS 3 N, but this is not what we observe here. The ensembles D u, D f, and TS f change with external conditions. If we rank-order these macrostates in terms of their structural similarity to the native state, Fig. 5 shows that, under strong native conditions, the order would be TS f 3 D f 3 N. This nonsensical result is a consequence of using a single reaction coordinate at the microscopic level and a series model to represent a process that is intrinsically parallel. We conclude that the TS is broad, but caveats exist. First, even in this model, different conformations have different weights. Ways exist to cluster the most important conformations into macrostates (37) that resemble more traditional ways of viewing transition states. Second, the breadth depends on conditions; weak folding conditions lead to a broader transition state. Our conclusion that the TS is late in folding may be limited by the short chain lengths accessible in our model. Indeed, the present model is perhaps best regarded as representing the early fast stages of secondary-structure formation and some assembly, which may represent nucleation events, because larger complexes are not explored here. A principal test of a two-state folding model is whether it predicts the chevron (V-shaped) dependence of the folding rate on the driving force (denaturant or temperature) that is observed in experiments (38 41). Fig. 5 shows that this model does so. Moreover, some experiments show small downward curvature on the left (folding) branch of the chevron plot, called rollover (42, 43). This downward curvature is also predicted by the model. What is the basis for this curvature? Classical theories expect no curvature, based on assuming that the TS doesn t change with external conditions. In that light, the curvature has been explained as resulting from intermediate states (43) or broad transition states (44). Fig. 5 shows the microscopic interpretation that the gateway conformations change with denaturant or temperature. TS f reaches a limiting native-like distribution under strong native conditions. Discussion Some proteins fold and unfold rapidly with single-exponential (two-state) kinetics. Some explanations have been based on transition-state assumptions that (i) macrostates (D, TS, and N) are distinct nonoverlapping collections of identifiable microstates, (ii) there is a single reaction coordinate, so the terms forward and backward have meaning, and (iii) the rate of folding does not exceed the rate of some microscopic (energetic or entropic) bottleneck step. The present work offers a different interpretation at the microscopic level. (i) The macrostates, the denatured and apparent transition states, encompass all the nonnative conformations, but each macrostate is a different distribution of microscopic populations. (ii)multiple microscopic routes exist in the high-dimensional conformational space, so a single microscopic reaction coordinate is not sufficient. (iii) Folding is faster than an intrinsic slowest rate, not a slowing down relative to an intrinsic fastest rate. (iv) The folding flux can be greater than individual microscopic fluxes. In this model, the apparent rate limit that gives single-exponential behavior arises because the multiplicity of parallel routes diminishes at the bottom of the folding funnel, not as the result of a microscopic barrier. J.S. was a Howard Hughes Medical Institute Predoctoral Fellow. K.D. thanks the National Institutes of Health for support (Grant GM34993). BIOPHYSICS 1. Jackson, S. E. (1998) Folding Des. 3, R81 R91. 2. Burton, R. E., Myers, J. K. & Oas, T. G. (1998) Biochemistry 37, 5337 5343. 3. Kim, P. S. & Baldwin, R. L. (1990) Annu. Rev. Biochem. 59, 631 660. 4. Fersht, A. R. (1999) Structure and Mechanism in Protein Science (Freeman, New York). 5. Englander, W. (2000) Annu. Rev. Biophys. Biomol. Struct. 29, 213 238. 6. Eyring, H. (1935) J. Chem. Phys. 3, 107 115. 7. Kramers, H. A. (1940) Physica 7, 284 304. 8. Bryngelson, J. D., Onuchic, J. N., Socci, N. D. & Wolynes, P. G. (1995) Proteins 21, 167 195. 9. Pande, V. S., Grosberg, A. Y. & Tanaka, T. (1997) Folding Des. 2, 109 114. 10. Chan, H. S. & Dill, K. A. (1994) J. Chem. Phys. 100, 9238 9257. 11. Socci, N. D., Onuchic, J. N. & Wolynes, P. G. (1998) Proteins 32, 136 158. 12. Klimov, D. K. & Thirumalai, D. (1998) J. Mol. Biol. 282, 471 492. 13. Shakhnovich, E. I., Abkevich, V. I. & Ptitsyn, O. B. (1996) Nature 379, 96 98. 14. Zhou, Y. & Karplus, M. (1999) Nature 401, 400 402. 15. Dill, K. A. (1999) Protein Sci. 8, 1166 1180. 16. Zwanzig, R. (1997) Proc. Natl. Acad. Sci. USA 94, 148 150. 17. Bicout, D. J. & Szabo, A. (2000) Protein Sci. 9, 452 465. 18. Zangi, R., Kovacs, H., van Gunsteren, W. F., Johansson, J. & Mark, A. E. (2001) Proteins 43, 395 402. 19. Leopold, P. E., Montal, M. & Onuchic, J. N. (1992) Proc. Natl. Acad. Sci. USA 89, 8721 8725. 20. Wolynes, P. G., Onuchic, J. N. & Thirumalai, D. (1995) Science 267, 1619 1620. 21. Dill, K. A. & Chan, H. S. (1997) Nat. Struct. Biol. 4, 10 19. 22. Chan, H. S. & Dill, K. A. (1989) Macromolecules 22, 4559 4573. 23. Lazaridis, T. & Karplus, M. (1997) Science 278, 1928 1930. 24. Duan, Y. & Kollman, P. A. (1998) Science 282, 740 744. 25. Kazmirski, S. L., Wong, K.-B., Freund, S. M. V., Tan, Y.-J., Fersht, A. R. & Daggett, V. (2001) Proc. Natl. Acad. Sci. USA 98, 4349 4354. 26. Taketomi, H., Ueda, Y. & Go, N. (1975) Int. J. Pept. Protein Res. 7, 445 459. 27. Takada, S. (1999) Proc. Natl. Acad. Sci. USA 96, 11698 11700. 28. Galzitskaya, O. V. & Finkelstein, A. V. (1999) Proc. Natl. Acad. Sci. USA 96, 11299 11304. 29. Alm, E. & Baker, D. (1999) Proc. Natl. Acad. Sci. USA 96, 11305 11310. 30. Muñoz, V., Henry, E. R., Hofrichter, J. & Eaton, W. A. (1999) Proc. Natl. Acad. Sci. USA 96, 11311 11316. 31. Cieplak, M., Karbowski, J., Henkel, M. & Banavar, J. R. (1998) Phys. Rev. Lett. 80, 3654 3657. Schonbrun and Dill PNAS October 28, 2003 vol. 100 no. 22 12681

32. Chan, H. S. (1998) in Monte Carlo Approach to Biopolymers and Protein Folding, eds. Grassberger, P., Barkema, G. T. & Nadler, W. (World Scientific, Teaneck, NJ), pp. 29 44. 33. Press, W. H., Teukolsky, S. A., Vetterling, W. T. & Flannery, B. P. (1992) Numerical Recipes in C (Cambridge Univ. Press, New York), Chap. 11. 34. Levinthal, C. (1968) J. Chim. Phys. 65, 44 45. 35. Klimov, D. K. & Thirumalai, D. (2001) Proteins 43, 465 475. 36. Sali, A., Shakhnovich, E. & Karplus, M. (1994) Nature 369, 248 252. 37. Ozkahn, S. B., Bahar, I. & Dill, K. A. (2001) Nat. Struct. Biol. 8, 765 769. 38. Jackson, S. E. & Fersht, A. R. (1991) Biochemistry 30, 10428 10435. 39. Otzen, D. E., Kristensen, O., Proctor, M. & Oliveberg, M. (1999) Biochemistry 38, 6499 6511. 40. Burton, R. E., Huang, G. S., Daugherty, M. A., Calderone, T. L. & Oas, T. G. (1997) Nat. Struct. Biol. 4, 305 310. 41. Matthews, R. (1993) Annu. Rev. Biochem. 62, 653 683. 42. Khorasanizadeh, S., Petrs, I. D. & Roder, H. (1996) Nat. Struct. Biol. 3, 193 205. 43. Matouschek, A., Kellis, J.T., Jr., Serrano, L., Bycroft, M. & Fersht, A. R. (1990) Nature 346, 440 445. 44. Oliveberg, M. (1997) Acc. Chem. Res. 31, 765 772. 12682 www.pnas.org cgi doi 10.1073 pnas.1735417100 Schonbrun and Dill