Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2019 Mar 8.
Published in final edited form as: Curr Opin Chem Biol. 2010 Sep 6;14(5):652–659. doi: 10.1016/j.cbpa.2010.08.012

Enzyme dynamics point to stepwise conformational selection in catalysis

Buyong Ma 1, Ruth Nussinov 1,2
PMCID: PMC6407632  NIHMSID: NIHMS1007578  PMID: 20822947

Abstract

Recent data increasingly reveal that conformational dynamics are indispensable to enzyme function throughout the catalytic cycle, in substrate recruiting, chemical transformation, and product release. Conformational transitions may involve conformational selection and induced fit, which can be viewed as a special case in the catalytic network. NMR, X-ray crystallography, single-molecule FRET, and simulations clearly demonstrate that the free enzyme dynamics already encompass all the conformations necessary for substrate binding, preorganization, transition-state stabilization, and product release. Conformational selection and substate population shift at each step of the catalytic turnover can accommodate enzyme specificity and efficiency. Within such a framework, entropy can have a larger role in conformational dynamics than in direct energy transfer in dynamically promoted catalysis.

Introduction

Evolution has optimized enzymes to selectively and efficiently catalyze biochemical reactions. It is generally accepted that the catalytic power of enzymes largely derives from their ability to stabilize transition states, lowering the barrier that reactants have to pass to reach productive states. However, the clear and attractive transition-state theory does not account for the reaction story in its entirety mainly because of the enzymes’ conformational ensembles and dynamics. On the basis of theoretical arguments, we suggested that the entire transition-state region, rather than a single saddle point, contributes to reaction kinetic; thus instead of optimizing the binding to a well-defined transition-state structure, enzymes are optimized by evolution to bind efficiently with a transition-state ensemble [1]. Further, we proposed that preexisting substates at the bottom of folding funnels of the free enzyme and population shifts can usefully explain enzyme pathways [2], and that there is no basic difference between funnels depicting ensembles of conformers of single molecules with loop, or domain motions, as compared to allosteric movements of subunits in multimeric quaternary structures. All states pre-exist, and evolution makes use of these for enzyme function [2]. Experimental evidence has strongly supported the role of conformational dynamics in catalysis [3]. Yet, the question of whether protein dynamics contribute to enzyme catalysis and if so to what extent [3,4●●,5] has been a debated issue. Since it is very difficult to establish a clear-cut linkage between protein flexibility and enzymatic function by experiment, it has largely been addressed by integrating experiments with computations.

Enzyme single-molecule studies have clearly shown fluctuations in conformations and in catalytic activity over a broad timescale ranging from 1 ms to 100 s [6]. Barrier crossing is over a much shorter, femtosecond to picosecond timescale. Before and following the barrier crossing leading to products, enzymes apparently spend most of their time assisting the reactants to reach the transition state, and escorting exiting products. Here, we show that at every step enzymes select their appropriate conformation optimized for substrate binding, chemical reaction, and product release, reviewing the literature published since 2007 in this light. We also note that conformational selection is general, and was observed for protein–protein, protein–DNA, protein–RNA, and protein–small molecule interactions [7,8].

Conformational selection and population shift constitute the catalytic network

The classical Michaelis–Menten equation (1) (Figure 1) is based on the assumption that an enzyme has only one (or ensemble-averaged) conformation, in which enzyme E quickly binds with substrate S to form Michaelis complex ES, followed by a reaction to EP and release of Product P.

Figure 1.

Figure 1

Conformational fluctuations of enzymes lead to ‘catalytic networks’. Eqn (1) is the classical Michaelis–Menten equation. Eqn (2) considers that the conformational ensemble of enzymes allows multiple catalytic reactions to occur in parallel, forming catalytic networks. Eqn (3) illustrates a narrowly defined conformational selection pathway, where only unbound enzyme has multiple conformers while the Michaelis ES and the Product EP complexes are not affected by conformational fluctuations. Eqn (4) is an induced fit mechanism where substrate binding leads to an ES complex with different enzyme conformation.

The conformational fluctuations of enzymes [6,9] lead to ensembles of the Michaelis ES and Product EP complexes and their related transition states [1,6]. Therefore, a more complete description of the overall catalytic process has to consider the interconverting conformers at each step, as in Eqn (2) (Figure 1), via ‘catalytic networks’ [6,10]. Catalytic networks can be defined as parallel reactions where each reaction step has multiple enzyme conformers in equilibrium. Even though each ‘single molecule’ does not necessarily exhibit Michaelis–Menten (MM) steady-state kinetics [11], in most cases, a single fluctuating enzyme can either follow or be reconciled with ensemble-averaged MM kinetics [6,11]. Eqn (2) indicates that many enzyme conformations pre-exist allowing the substrate to bind different conformers [12]. The populations of the Michaelis complex ES and Product complex EP will shift upon binding, depending on time and environment.Collectively, the conformational dynamics of the Michaelis ES and the Product EP complexes affect the overall turn-over rate, which corresponds to the most important transition-state barrier crossing (K2) and the product release steps (K3). Single-molecule studies indicate that the catalytic velocity of a single enzyme molecule fluctuates among the catalytic networks over a broad range of time-scales, from 10−3 to 10 s [6] among the catalytic networks (Eqn (2), Figure 1) indicating a strong dynamic influence; however, the question as to whether the dynamics of Michaelis complexes ES can contribute to barrier crossing is highly debated, and the dynamics of the EP complexes are usually less noticed.

Eqn (2) (Figure 1) can suggest two simplified scenarios for the relative contribution of ‘induced fit’ and ‘conformation selection’ to the overall catalytic mechanism. If one neglects the conformational fluctuations of the Michaelis ES and the Product EP complexes, conformational selection can be ‘narrowly’ defined by Eqn (3) (Figure 1). At the other extreme, if one ignores the conformational substates of the apo enzyme and only considers the conformational changes of Michaelis complexes ES, an induced fit mechanism can be described by Eqn (4) (Figure 1). The relative contributions of induced fit or conformational selection may depend on on-rate K1 and off-rate K−1 [13] and can be examined using the flux through conformational selection (Eqn (3)) and induced fit (Eqn (4), Figure 1) pathways [14]. In the case of NADPH binding to dihydrofolate reductase (DHFR), at a very low ligand concentration about 80% of the flux goes through the conformational selection pathway, but induced fit rapidly dominates as ligand concentration increases. For the Desulfovibrio desulfuricans flavodoxin folding and cofactor binding, which is an example of an apparently exclusive induced fit mechanism, the flux goes through each path as a function of the holoprotein concentration [14]. This work indicates that both conformational selection and induced fit pathways are often used, essentially supporting the complete description in Eqn (2) (Figure 1).

Sullivan and Holyoak [15] used phosphoenolpyruvate carboxykinases (PEPCK) as a representative enzyme in which transition between the open and closed conformations occludes the active site from the solvent. The crystal structures of the PEPCK revealed that the active site is lid-gated. Since a closed form does not allow substrate binding, it was argued that induced fit rather than conformation selection must operate [15]. However, crystal structure snapshots do not provide the entire conformational dynamic picture, particularly of loop movements. If loop dynamics are fast, its structure cannot be solved by X-ray crystallography. This suggests that snapshots often cannot directly obtain adequate conformational sampling. As conformational sampling indicated for bacterial phosphotriesterase [16], maltose-binding protein [17], and choline oxidase [18], many lid-gated enzymes have both closed and open conformations in equilibrium before ligand binding, suggesting that the binding mechanism of lid-gated enzymes could also follow a conformational selection pathway [14]. The opening–closing binding site dynamics may couple with large domain motions, as in the case of the cytochrome bc1 complex (Figure 2) [19]. In summary, the catalytic network in Eqn (2) (Figure 1) with conformational selection and population shift could provide a more comprehensive view of enzyme action. The range of binding constants (K1) and conformational flexibility of ES/EP complexes can be tuned to enzymatic function. One example is the versatile cytochrome p450, which recognizes a broad range of substrates. Tight binding ligands can restrict the conformational substates, while weak binders can broaden the distribution [20].

Figure 2.

Figure 2

Conformational dynamics of the binding site for the extrinsic domain of the iron-sulfur protein (ISP-ED) subunit in the cytochrome b (cyt b) subunit of the cyt bc1 complex is coupled with large motion in ISP-ED (Ref. [18]). MD simulations of the bc1 complex from the R. sphaeroides (Rsbc1) with the bound inhibitor stigmatellin reveal that if no inhibitor exists, the binding pocket opens (B Ma, L Esser, R Nussinov, D Xia, unpublished). The Rsbc1 catalyzes the electron transfer from ubiquinol to cytochrome c and simultaneously pumps protons across the membrane. In the right panel, the simulated nonsymmetrical dimer complex shows that the inhibitor (blue balls) is bound to only to one cyt b subunit (depicted by a red ribbon), while the other cyt b subunit (green ribbon) has an apo binding site. The yellow ribbons are cyt c1 subunits. The ISP subunit with the 2Fe–2S clusters are shown in pink ribbons. The ISP-ED can have a large motion accompanying the electron transfer reaction. In the left panel, two b subunits are superimposed to compare the binding site changes. While the bound subunit (red ribbon) keeps the closed form, the apo subunit (green ribbon) opened up during simulation.

Enzyme specificity can be controlled through multiple, consecutive conformational selection steps

To allow active site substrate selectivity, transition-state stabilization, and product release, enzymes have to balance specificity and efficacy. A ‘lock-and-key’ mechanism could provide ‘perfect’ selectivity. However, enzymes classified as following a lock-and-key mechanism also present a dynamic conformer selection process. Serine protease is conventionally regarded as fitting the rigid lock-and-key model. However, nanosecond timescale binding loop movement was observed to select an inhibitor conformation [21]. The flaviviral nonstructural 3 protease (NS3pro), a chymotrypsin-like serine protease also presents conformational selection to facilitate substrate binding and product release, promoting the formation of the catalytically competent oxyanion hole [22].

Evolution appears to have optimized the enzyme’s conformational substates toward its successive tasks, substrate binding, transition-state stabilization, and product release. As shown in Scheme 1 and Eqn (2) (Figure 1), conformational fluctuations take place at all catalytic stages. Thus, compared with the lock-and-key rigid specificity for substrate binding or transition-state stabilization, the multistep conformational selection mechanism could similarly provide a rigorous selection. For example, even though each stage yields only a 50% correct selection ratio, over all three stages it can reach 87.5%.

Scheme 1.

Scheme 1

Multiple conformational selection steps can boost enzyme specificity. This hypothetical scheme illustrates that if there is only 50% specificity for each of the three steps, only 12.5% lead to product release, assuming that the rates of other pathways are slow.

A first step toward validation of such a mechanism would require a demonstration that the enzyme can have all prior conformations with conversion speed comparable to enzyme turnover rate. Such evidence has emerged recently [23,24,25●●,26●●,27]. In a study combining experimental and computational approaches, Henzler-Wildman et al. identified conformational substates along the reaction trajectory of adenylate kinase, and determined the timescales for the transitions among them [23]. They observed that the hierarchy of timescales in protein dynamics is linked to enzyme catalysis, with picosecond to nanosecond timescale fluctuation in the hinge regions of the adenylate kinase facilitating the large-scale, slower lid motions that produce a catalytically competent state. Using NMR relaxation experiments, the dynamics of the prolyl cistrans isomerase cyclophilin A (CypA) were studied in the enzyme’s substrate-free state and during catalysis. Characteristic enzyme motions detected during catalysis were observed in the free enzyme state with frequencies corresponding to the catalytic turnover rates. This correlation could suggest that protein motions necessary for catalysis are an intrinsic property of the enzyme and might even limit the overall turnover rate [24]. The direct link between the intrinsic motions and the catalytic turnover rate was revealed using ambient-temperature X-ray crystallographic data collection and automated electron-density sampling of interconverting substates of the human pro-line isomerase CypA [25●●].

DNA polymerases may have the highest specificity pressure to ensure fidelity. Conformational transitions which are involved in the series of steps could assist in safe-guarding against errors. Before substrate binding, DNA polymerase I samples the open and closed conformations in millisecond timescale [26●●]. DNA polymerase μ (polμ) has a rate-limiting ‘precatalytic translocation step’ to ensure accuracy and retain efficiency [27]. The conformational dynamics of the Y-family DNA polymerase Dpo4 was also shown to control its selectivity [28,29]. Of particular note is the flexible region surrounding the H-helix of the thumb domain, which selects the correct Watson–Crick base pair [28]. In the DNA polymerase Klentaq1, there is a fast equilibration between the fingers subdomain in the open and closed states. Even though the difference in the ground-state binding affinity to each of the dNTPs is very small, the enzyme’s selectivity between Watson–Crick and non-Watson–Crick base pairs amplifies small differences [30]. A similar control exists for ribonucleases [31,32].

Why can product release be a rate-limiting step and conformational dynamics of the EP complex control the enzymatic reaction? Intuitively, if the product cannot be stabilized, it could go back to become a reactant and the reaction is reversed. In ribonuclease A (RNase A), the product release-related conformational change limits the overall catalytic rate [33]. Formamidopyrimidine-DNA N-glycosylase provides an example of control by product release [34]. In addition, nonequilibrium dynamics may also be involved in retaining a favorable enzyme configuration if product release follows a slow conformational relaxation, resulting in dynamic acceleration of enzyme turnover at high substrate concentration [35].

Conformational substates may help transition-state barrier crossing

With enzyme conformations changing constantly over varied timescales and catalytic steps, one can wonder if the energy transfer accompanying these changes can be used to help decrease the transition-state barrier. Kamerlin and Warshel asked ‘At the dawn of the 21st century: Is dynamics the missing link for understanding enzyme catalysis?’ [4●●]. Pisliakov et al. have shown that enzyme millisecond conformational dynamics do not couple with fast motions, since the specific memory of fast motions dissipates during the slow conformation fluctuations [36].

Coupled motions spanning femtoseconds to milliseconds in DHFR catalysis were proposed to promote catalysis already years ago [37]. These coupled motions explain nonadditivity effects in the hydride transfer rates for mutations distal to the active site [38]. NMR studies found that each intermediate in the catalytic cycle of DHFR samples low-lying excited states whose conformations resemble the ground-state structures preceding and following the intermediates. The maximum hydride transfer and steady-state turnover rates are governed by the dynamics of the transitions between the ground and excited states of the intermediates [39]. Computational analysis revealed that protein dynamics can be changed by ligand binding via network-bridging effects which modify the residue interaction network in the protein [40]. Such effect was recently confirmed by NMR relaxation dispersion measurement, showing that the dynamics of the substrate and product binary complexes are governed by quite different kinetic and thermodynamic parameters [41]. However, even though these works revealed that structural fluctuations in DHFR are exquisitely optimized for every intermediate in the catalytic cycle [41], the exact motion that helps the hydride transfer remains to be unveiled. Evidence in support of promoting vibration modes was reported for a similar system (pentaerythritol tetranitrate reductase) with a hydrogen tunneling mechanism [42].

The transition-state region is often energetically flat, with a range of structures which are very close in energy [1,43]. Substates with a large population may lead to transitionstate ensemble [1,44,45]. Thermodynamically, a barrier could decrease due to an entropy compensation effect [1,46]. Therefore, entropy can have a larger role in the conformational change dynamics than in direct energy transfer [47]. The entropy contribution is also directly related to the complex scenarios described in Eqn (2) (Figure 1).

Conformational dynamics may greatly facilitate the positioning of the substrate toward barrier crossing or product release through a preorganization mechanism [4●●]. RNA polymerase hasa trigger loop (TL) which can fold into α-helices (TH) during catalysis. TH formation helps in the geometrical alignment of reactants in the enzyme’s active site, allowing rapid nucleotidyl transfer [48]. Conformational dynamics-controlled substrate alignments which in turn control catalysis are also observed in other systems [49,50]. The bacterial phosphotriesterase has both closed and open conformations in the apo state. The closed conformation is ideally preorganized to lower the reaction barrier, but it is not compatible with product release. In contrast, the open conformation is better organized for product release but not for chemical reaction [16].

Enzyme conformational dynamics raise an interesting question related to the ‘internal friction’ [51] incurred during the barrier crossing, and the movements of the substrate or product in the protein. Around the barrier crossing area, the coupling between the protein conformational kinetics and the reaction kinetics can increase the transmission coefficient by a small ratio and Kramers’ model may not be a good choice [52]. However, Kramers’ model is very effective in measuring how viscosity will affect substrate or product diffusion and thus catalysis [51]. The generalized Kramers’ model can explain the conformational dynamics of enzymatic reaction [53], predicting that intermediates along the reaction pathways can also promote barrier crossing [54]. Therefore, collectively, conformational transitions among the steps (as illustrated in Eqn (2)) will promote enzyme reaction.

Conclusions

Even though it is still highly debated whether the enzyme dynamic motions contribute to decrease the chemical reaction barrier, the consensus picture emerging from experimental and computational studies indicates that enzyme conformational transitions are highly organized increasing enzyme specificity and efficiency. It is striking that the enzyme structure has evolved to have all functionally relevant conformational substates [2], and that these are optimized for substrate binding, chemical reaction, and product release. Even enzymes catalyzing multiple chemical steps can still accommodate the conformational reorganization required to stabilize the transition states [55] via pre-existing substates. We further note that the ensemble continuum between extreme conformations (open and closed) can also be controlled allosterically [56,57], since allosteric substates similarly pre-exist and undergo population shift [2]. Broader substrate specificities as in the multiply promiscuous hydrolase [58] or via allosteric mutations as in the directed evolution of an enantioselective Baeyer–Villiger monooxygenase [59] further highlight the conformational ensemble heterogeneity and the ruggedness of the free energy landscape. Such conformational dynamics suggest a description of enzyme turnover by a ‘catalytic network’, which initiates with conformational selection of substrate binding and proceeds via conformational substates population shifts which selectively stabilize transition states and facilitate product release. Induced fit can be viewed as a specific pathway within this broad spectrum. Multiple conformational steps can boost enzyme specificity and reaction rates regardless of whether the rate-limiting step is substrate binding, pre-organization, chemical reaction, or product release.

Ideally, substrate-binding energy can be used to promote both ground-state association and transition-state stabilization, or be translated into barrier crossing promoting motion. However, even though experimental evidence implicates vibration assistance in hydrogen tunneling, additional experimental data are needed to reveal the complex energetic benefit of the promoting motions. Thermodynamically, entropy can have a larger role in the conformational change dynamics than in direct energy transfer in catalysis. Enzymes can also use multimers to perform complex tasks. In the case of the homing endo-nuclease I-AniI, the two monomers in the dimer have different specificities: one for substrate binding, the other for transition-state stabilization [60]. Since the dimer is symmetric, conformational changes must exist to accommodate the two different tasks. A possible mutual conformational selection between DNA and a dimeric protein can be best illustrated by the glucocorticoid receptor (GR), in which DNA tailors the activity of the receptor toward specific target genes via a sequence-specific allosteric mechanism [8,61].

Acknowledgements

This project has been funded in whole or in part with Federal funds from the National Cancer Institute, National Institutes of Health, under contract number HHSN261200800001E. The content of this publication does not necessarily reflect the views or policies of the Department of Health and Human Services, nor does mention of trade names, commercial products, or organizations imply endorsement by the U.S. Government. This research was supported (in part) by the Intramural Research Program of the NIH, NCI, Center for Cancer Research.

Footnotes

Papers of particular interest, published within the period of review, have been highlighted as:

● of special interest

●● of outstanding interest

References and recommended reading

  • 1.Ma B, Kumar S, Tsai CJ, Hu Z, Nussinov R: Transition-state ensemble in enzyme catalysis: possibility, reality, or necessity? J Theor Biol 2000, 203:383–397. [DOI] [PubMed] [Google Scholar]
  • 2.Kumar S, Ma B, Tsai CJ, Sinha N, Nussinov R: Folding and binding cascades: dynamic landscapes and population shifts. Protein Sci 2000, 9:10–19. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Henzler-Wildman K, Kern D: Dynamic personalities of proteins. Nature 2007, 450:964–972. [DOI] [PubMed] [Google Scholar]
  • 4.●●.Kamerlin SC, Warshel A: At the dawn of the 21st century: is dynamics the missing link for understanding enzyme catalysis? Proteins 2009, 78:1339–1375.A critical assessment of the contribution of dynamic motions to transition-state barrier crossing through various mechanisms. While the paper agrees that the information obtained from the experimentally observed dynamics is crucial to the understanding of enzyme action, it argues that further experiments are essential in order to delineate the dynamic effects on enzymatic function.
  • 5.●.Schwartz SD, Schramm VL: Enzymatic transition states and dynamic motion in barrier crossing. Nat Chem Biol 2009, 5:551–558.A review of computational studies linking dynamic motions to transition-state formation, mostly based on QM/MM simulations of transition-state path sampling.
  • 6.Min W, English BP, Luo G, Cherayil BJ, Kou SC, Xie XS: Fluctuating enzymes: lessons from single-molecule studies. Acc Chem Res 2005, 38:923–931. [DOI] [PubMed] [Google Scholar]
  • 7.●.Boehr DD, Nussinov R, Wright PE: The role of dynamic conformational ensembles in biomolecular recognition. Nat Chem Biol 2009, 5:789–796.A general review of conformational selection and population shift for protein–ligand, protein–protein, protein–DNA, protein–RNA and RNA– ligand interactions.
  • 8.Ma B, Tsai CJ, Pan Y, Nussinov R: Why does binding of proteins to DNA or proteins to proteins not necessarily spell function? ACS Chem Biol 2010, 5:265–272. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 9.Ma B, Kumar S, Tsai CJ, Nussinov R: Folding funnels and binding mechanisms. Protein Eng 1999, 12:713–720. [DOI] [PubMed] [Google Scholar]
  • 10.●.Benkovic SJ, Hammes GG, Hammes-Schiffer S: Free-energy landscape of enzyme catalysis. Biochemistry 2008, 47:3317–3321.A review discussing the concept of catalytic networks in enzyme catalysis.
  • 11.Min W, Gopich IV, English BP, Kou SC, Xie XS, Szabo A: When does the Michaelis–Menten equation hold for fluctuating enzymes? J Phys Chem B 2006, 110:20093–20097. [DOI] [PubMed] [Google Scholar]
  • 12.Ma B, Shatsky M, Wolfson HJ, Nussinov R: Multiple diverse ligands binding at a single protein site: a matter of pre-existing populations. Protein Sci 2002, 11:184–197. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.●.Weikl TR, von Deuster C: Selected-fit versus induced-fit protein binding: kinetic differences and mutational analysis. Proteins 2009, 75:104–110.The paper presents a kinetic model of conformational selection and induced fit, focusing on the influence of the on-rate and off-rate substrate binding.
  • 14.●.Hammes GG, Chang YC, Oas TG: Conformational selection or induced fit: a flux description of reaction mechanism. Proc Natl Acad Sci U S A 2009, 106:13737–13741.This paper introduces a flux description of the relative contributions of conformational selection and induced fit in enzyme catalysis and found that both can contribute simultaneously.
  • 15.Sullivan SM, Holyoak T: Enzymes with lid-gated active sites must operate by an induced fit mechanism instead of conformational selection. Proc Natl Acad Sci U S A 2008, 105:13829–13834. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16.●.Jackson CJ, Foo JL, Tokuriki N, Afriat L, Carr PD, Kim HK, Schenk G, Tawfik DS, Ollis DL: Conformational sampling, catalysis, and evolution of the bacterial phosphotriesterase. Proc Natl Acad Sci U S A 2009, 106:21631–21636.An experimental report of crystal structures and kinetic analysis of an enzyme presenting open and closed conformations in substrate binding and product release.
  • 17.Tang C, Schwieters CD, Clore GM: Open-to-closed transition in apo maltose-binding protein observed by paramagnetic NMR. Nature 2007, 449:1078–1082. [DOI] [PubMed] [Google Scholar]
  • 18.Xin Y, Gadda G, Hamelberg D: The cluster of hydrophobic residues controls the entrance to the active site of choline oxidase. Biochemistry 2009, 48:9599–9605. [DOI] [PubMed] [Google Scholar]
  • 19.Esser L, Elberry M, Zhou F, Yu CA, Yu L, Xia D: Inhibitor-complexed structures of the cytochrome bc1 from the photosynthetic bacterium Rhodobacter sphaeroides. J Biol Chem 2008, 283:2846–2857. [DOI] [PubMed] [Google Scholar]
  • 20.Lampe JN, Brandman R, Sivaramakrishnan S, de Montellano PR: Two-dimensional NMR and all-atom molecular dynamics of cytochrome P450 CYP119 reveal hidden conformational substates. J Biol Chem 2010, 285:9594–9603. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21.Gaspari Z, Varnai P, Szappanos B, Perczel A: Reconciling the lock-and-key and dynamic views of canonical serine protease inhibitor action. FEBS Lett 2010, 584:203–206. [DOI] [PubMed] [Google Scholar]
  • 22.Ekonomiuk D, Caflisch A: Activation of the West Nile virus NS3 protease: molecular dynamics evidence for a conformational selection mechanism. Protein Sci 2009, 18:1003–1011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Henzler-Wildman KA, Thai V, Lei M, Ott M, Wolf-Watz M, Fenn T, Pozharski E, Wilson MA, Petsko GA, Karplus M et al. : Intrinsic motions along an enzymatic reaction trajectory. Nature 2007, 450:838–844. [DOI] [PubMed] [Google Scholar]
  • 24.Eisenmesser EZ, Millet O, Labeikovsky W, Korzhnev DM, Wolf-Watz M, Bosco DA, Skalicky JJ, Kay LE, Kern D: Intrinsic dynamics of an enzyme underlies catalysis. Nature 2005, 438:117–121. [DOI] [PubMed] [Google Scholar]
  • 25.●●.Fraser JS, Clarkson MW, Degnan SC, Erion R, Kern D, Alber T: Hidden alternative structures of proline isomerase essential for catalysis. Nature 2009, 462:669–673.This paper describes a new experimental approach allowing to simulta-neously reveal detailed atomic structures of rare states and catalysis. It links the intrinsic motions in the free enzyme to catalysis.
  • 26.●●.Santoso Y, Joyce CM, Potapova O, Le Reste L, Hohlbein J, Torella JP, Grindley ND, Kapanidis AN: Conformational transitions in DNA polymerase I revealed by single-molecule FRET. Proc Natl Acad Sci U S A 2010, 107:715–720.The real-time observation of conformational dynamics described in this paper illustrated that most of the unliganded polymerase molecules sample the open and closed conformations in the millisecond timescale. This work links the fidelity of DNA polymerases to conformational selection.
  • 27.Andrade P, Martin MJ, Juarez R, Lopez de Saro F, Blanco L: Limited terminal transferase in human DNA polymerase mu defines the required balance between accuracy and efficiency in NHEJ. Proc Natl Acad Sci U S A 2009, 106:16203–16208. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Eoff RL, Sanchez-Ponce R, Guengerich FP: Conformational changes during nucleotide selection by Sulfolobus solfataricus DNA polymerase Dpo4. J Biol Chem 2009, 284:21090–21099. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29.Xu C, Maxwell BA, Brown JA, Zhang L, Suo Z: Global conformational dynamics of a Y-family DNA polymerase during catalysis. PLoS Biol 2009, 7:e1000225. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.Rothwell PJ, Waksman G: A pre-equilibrium before nucleotide binding limits fingers subdomain closure by Klentaq1. J Biol Chem 2007, 282:28884–28892. [DOI] [PubMed] [Google Scholar]
  • 31.Chimnaronk S, Forouhar F, Sakai J, Yao M, Tron CM, Atta M, Fontecave M, Hunt JF, Tanaka I: Snapshots of dynamics in synthesizing N(6)-isopentenyladenosine at the tRNA anticodon. Biochemistry 2009, 48:5057–5065. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Hsieh J, Fierke CA: Conformational change in the Bacillus subtilis RNase P holoenzyme–pre-tRNA complex enhances substrate affinity and limits cleavage rate. RNA 2009, 15:1565–1577. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 33.Doucet N, Watt ED, Loria JP: The flexibility of a distant loop modulates active site motion and product release in ribonuclease A. Biochemistry 2009, 48:7160–7168. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Kuznetsov NA, Zharkov DO, Koval VV, Buckle M, Fedorova OS: Reversible chemical step and rate-limiting enzyme regeneration in the reaction catalyzed by formamidopyrimidine-DNA glycosylase. Biochemistry 2009, 48:11335–11343. [DOI] [PubMed] [Google Scholar]
  • 35.●.Min W, Xie XS, Bagchi B: Role of conformational dynamics in kinetics of an enzymatic cycle in a nonequilibrium steady state. J Chem Phys 2009, 131:065104.A theoretical analysis which considers nonequilibrium dynamics in enzyme catalysis. This work represents a new way to evaluate protein dynamics when equilibrated transition states are invalid. The authors observed that a slow conformational relaxation during product release could accelerate catalysis.
  • 36.Pisliakov AV, Cao J, Kamerlin SC, Warshel A: Enzyme millisecond conformational dynamics do not catalyze the chemical step. Proc Natl Acad Sci U S A 2009, 106:17359–17364. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Agarwal PK, Billeter SR, Rajagopalan PT, Benkovic SJ, Hammes-Schiffer S: Network of coupled promoting motions in enzyme catalysis. Proc Natl Acad Sci U S A 2002, 99:2794–2799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Wong KF, Selzer T, Benkovic SJ, Hammes-Schiffer S: Impact of distal mutations on the network of coupled motions correlated to hydride transfer in dihydrofolate reductase. Proc Natl Acad Sci U S A 2005, 102:6807–6812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Boehr DD, McElheny D, Dyson HJ, Wright PE: The dynamic energy landscape of dihydrofolate reductase catalysis. Science 2006, 313:1638–1642. [DOI] [PubMed] [Google Scholar]
  • 40.Hu Z, Bowen D, Southerland WM, del Sol A, Pan Y, Nussinov R, Ma B: Ligand binding and circular permutation modify residue interaction network in DHFR. PLoS Comput Biol 2007, 3:e117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Boehr DD, McElheny D, Dyson HJ, Wright PE: Millisecond timescale fluctuations in dihydrofolate reductase are exquisitely sensitive to the bound ligands. Proc Natl Acad Sci U S A 2010, 107:1373–1378. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42.●.Pudney CR, Hay S, Levy C, Pang J, Sutcliffe MJ, Leys D, Scrutton NS: Evidence to support the hypothesis that promoting vibrations enhance the rate of an enzyme catalyzed H-tunneling reaction. J Am Chem Soc 2009, 131:17072–17073.This work represents experimental evidence supporting vibrational pro-motion of hydrogen tunneling by showing temperature and pressure dependence of the primary kinetic isotope effect.
  • 43.Yang Y, Cui Q: The hydrolysis activity of adenosine triphosphate in myosin: a theoretical analysis of anomeric effects and the nature of the transition state. J Phys Chem A 2009, 113:12439–12446. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Lu Q, Wang J: Single molecule conformational dynamics of adenylate kinase: energy landscape, structural correlations, and transition state ensembles. J Am Chem Soc 2008, 130:4772–4783. [DOI] [PubMed] [Google Scholar]
  • 45.Saen-Oon S, Quaytman-Machleder S, Schramm VL, Schwartz SD: Atomic detail of chemical transformation at the transition state of an enzymatic reaction. Proc Natl Acad Sci U S A 2008, 105:16543–16548. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Edwards AA, Mason JM, Clinch K, Tyler PC, Evans GB, Schramm VL: Altered enthalpy–entropy compensation in picomolar transition state analogues of human purine nucleoside phosphorylase. Biochemistry 2009, 48:5226–5238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.●.Edwards AA, Tipton JD, Brenowitz MD, Emmett MR, Marshall AG, Evans GB, Tyler PC, Schramm VL: Conformational states of human purine nucleoside phosphorylase at rest, at work, and with transition state analogues. Biochemistry 2010, 49:2058–2067.Using hydrogen–deuterium exchange and sedimentation analysis, the authors show that protein conformations (dynamic motion) correlate more closely with binding entropy than with affinity in enzyme catalysis.
  • 48.Zhang J, Palangat M, Landick R: Role of the RNA polymerase trigger loop in catalysis and pausing. Nat Struct Mol Biol 2010, 17:99–104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Heyes DJ, Menon BR, Sakuma M, Scrutton NS: Conformational events during ternary enzyme–substrate complex formation are rate limiting in the catalytic cycle of the light-driven enzyme protochlorophyllide oxidoreductase. Biochemistry 2008, 47:10991–10998. [DOI] [PubMed] [Google Scholar]
  • 50.Davulcu O, Flynn PF, Chapman MS, Skalicky JJ: Intrinsic domain and loop dynamics commensurate with catalytic turnover in an induced-fit enzyme. Structure 2009, 17:1356–1367. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Raber ML, Freeman MF, Townsend CA: Dissection of the stepwise mechanism to beta-lactam formation and elucidation of a rate-determining conformational change in beta-lactam synthetase. J Biol Chem 2009, 284:207–217. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 52.Ruiz-Pernia JJ, Tunon I, Moliner V, Hynes JT, Roca M: Dynamic effects on reaction rates in a Michael addition catalyzed by chalcone isomerase. Beyond the frozen environment approach. J Am Chem Soc 2008, 130:7477–7488. [DOI] [PubMed] [Google Scholar]
  • 53.Min W, Xie XS: Kramers model with a power-law friction kernel: dispersed kinetics and dynamic disorder of biochemical reactions. Phys Rev E Stat Nonlin Soft Matter Phys 2006, 73:010902. [DOI] [PubMed] [Google Scholar]
  • 54.Ma B, Nussinov R: The stability of monomeric intermediates controls amyloid formation: Abeta25–35 and its N27Q mutant. Biophys J 2006, 90:3365–3374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55.●.Smith AJ, Muller R, Toscano MD, Kast P, Hellinga HW, Hilvert D, Houk KN: Structural reorganization and preorganization in enzyme active sites: comparisons of experimental and theoretically ideal active site geometries in the multistep serine esterase reaction cycle. J Am Chem Soc 2008, 130:15361–15373.This paper describes quantum mechanical studies of an enzyme which catalyzes multiple chemical steps. The paper shows that the active site can be reorganized for stepwise reactions.
  • 56.Osmulski PA, Hochstrasser M, Gaczynska M: A tetrahedral transition state at the active sites of the 20S proteasome is coupled to opening of the alpha-ring channel. Structure 2009, 17:1137–1147. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 57.Hindie V, Stroba A, Zhang H, Lopez-Garcia LA, Idrissova L, Zeuzem S, Hirschberg D, Schaeffer F, Jorgensen TJ, Engel M et al. : Structure and allosteric effects of low-molecular-weight activators on the protein kinase PDK1. Nat Chem Biol 2009, 5:758–764. [DOI] [PubMed] [Google Scholar]
  • 58.van Loo B, Jonas S, Babtie AC, Benjdia A, Berteau O, Hyvonen M, Hollfelder F: An efficient, multiply promiscuous hydrolase in the alkaline phosphatase superfamily. Proc Natl Acad Sci U S A 2010, 107:2740–2745. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Wu S, Acevedo JP, Reetz MT: Induced allostery in the directed evolution of an enantioselective Baeyer–Villiger monooxygenase. Proc Natl Acad Sci U S A 2010, 107:2775–2780. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Thyme SB, Jarjour J, Takeuchi R, Havranek JJ, Ashworth J, Scharenberg AM, Stoddard BL, Baker D: Exploitation of binding energy for catalysis and design. Nature 2009, 461:1300–1304. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Meijsing SH, Pufall MA, So AY, Bates DL, Chen L, Yamamoto KR: DNA binding site sequence directs glucocorticoid receptor structure and activity. Science 2009, 324:407–410. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES