Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2015 Apr 1.
Published in final edited form as: Curr Opin Struct Biol. 2014 Mar 22;0:98–103. doi: 10.1016/j.sbi.2014.02.004

Computational approaches to mapping allosteric pathways

Victoria A Feher 1, Jacob Durrant 1, Adam T Van Wart 1, Rommie E Amaro 1
PMCID: PMC4040315  NIHMSID: NIHMS572451  PMID: 24667124

Abstract

Allosteric signaling occurs when chemical and/or physical changes at an allosteric site alter the activity of a primary orthosteric site often many Angstroms distant. A number of recently developed computational techniques, including dynamical network analysis, novel topological and molecular dynamics methods, and hybrids of these methods are useful for elucidating allosteric signaling pathways at the atomistic level. No single method prevails as best to identify allosteric signal propagation path(s), rather each has particular strengths in characterizing signals that occur over a specific timescale range and magnitude of conformational fluctuation. With continued improvement of accuracy and predictive power, these computational techniques aim to become useful drug discovery tools that will allow researchers to identify allostery critical residues for subsequent pharmacological targeting.

Introduction

The recently broadened paradigm of allosteric regulation, which is inclusive of monomeric single-domain [13] as well as multimeric proteins [4], can be thought of as any mechanism by which an effector perturbation at one site propagates a signal to an orthosteric site and shifts the population of a preexisting conformational sub-state to become more predominant [5,6]. Effector perturbations can result from a wide range of biological and physical phenomena, including the binding of a small effector molecule, post-translational modifications, protein binding, temperature changes, and pH changes. Until recently, an atomistic understanding of allosteric propagation could only be inferred by projecting experimental results onto static structural models provided by x-ray crystallography. However, over the last decade advances in computational simulation have also provided qualitative and quantitative descriptions of allosteric phenomena that have been experimentally validated and are increasingly predictive of experimental observables. These computational techniques can be broadly divided into two categories: those which seek to predict the overall conformational impact of an allosteric event on the protein state (ie. apo vs. holo), and those which aim to elucidate specific atomic-level allosteric pathways through which effector perturbations are transmitted. As a number of excellent reviews focusing on the former methodologies already exist [7••,8], the focus of this opinion will be on the latter.

Mapping Networks by Topology Analyses

Many early computational approaches for elucidating allosteric pathways are based on protein topological analyses such as graph theory, statistical coupling analysis, and perturbation algorithms [911]. These approaches, which assume that allosteric networks can be defined solely through inter-residue non-bonding interactions, have generally been successful in identifying residues involved in functional and correlated motions.

A recent topology-based algorithm, CONTACT [12•], relies on a single crystal structure to identify residues that have side-chain contact heterogeneity. CONTACT defines “contact networks” by tracing paths of residues that can adopt alternate conformations. This method accurately predicts the identity of residues involved in DHFR allostery and residues whose mutation affects allostery, as previously characterized using NMR and biochemical experiments [13]. The remarkable agreement between topological methods like CONTACT and experimental mutagenesis studies is impressive given that the topological characterizations include only the steric inter-residue van der Waals interactions. This suggests that such interactions are a major component of allosteric signal propagation and is consistent with previous studies that utilize a measure of residue steric incompatibilities such as residue “frustration” [14,15].

Jenik et. al [16•] has formalized a previous residue frustration algorithm [14] as a webserver for elucidating specific pathways of allosteric transmission named the “protein frustratometer” (http://lfp.qb.fcen.uba.ar/embnet). In its most straightforward implementation, it systematically “mutates” each set of interacting residues in silico and predicts the energetic consequences of each mutation. Each interaction is thus associated with a spectrum of energies. If the energy of wild-type residues is favorable relative to the energies of this spectrum, the interaction is said to be minimally frustrated. If it is unfavorable, the interaction is highly frustrated. Typically, ~10–15% of residue contacts are judged “frustrated” by this metric [15]. By tracing out pathways of highly frustrated interactions, it is possible to predict sets of residues with alternative conformations that are potential pathways for propagating an allosteric signal.

Mapping Networks by Simulation Analyses

Simulation-based methods have long been used to study allosteric processes. Increasingly, molecular dynamics trajectories, whether conventional, accelerated, steered, coarse-grained, or umbrella sampled, have been used to generate conformational ensembles for subsequent allosteric network analysis (reviewed elsewhere [7••]). Multiple methods such as normal mode analysis [1719], correlation matrices [2022], community-network analysis [23], mutual information [24] and dynamical network analysis [2527] have been applied to evaluate and identify pathways or regions of correlated residue motions presumably related to allostery. Recent advances in some of these methods are highlighted below.

ENM and Structural Perturbation Method

An elastic network model (ENM) represents a biomolecular structure as a coarse-grained network of nodes connected by a series of harmonic springs. ENMs are typically analyzed by normal mode analysis (NMA) [28]. In recent implementations electrostatic forces are incorporated into the model and selected spring-force constants can be modified to better account for water hydration effects at ionized interfacial surfaces [29•]. Once such a model has been constructed, techniques like the “structural perturbation method” can be used to identify potential allosteric pathways [30,31]. In brief, for each network node, the force constants of all connecting springs are modified, essentially mimicking a point mutation. The impact of this subtle modification on the overall motions of the system (as judged by normal-mode analysis) is assessed for each perturbed node. The nodes with the greatest impact are thought to be those that are most likely to participate in the allosteric mechanism. When applied to large protein systems such as GroEL and GroES, these methods have successfully identified correlated motions within and between the rings that link GroEL substrate binding site residues to the GroES ATPase activity [31].

Modified MD Simulations and Fourier Transform Analysis

Pump-probe molecular dynamics simulations apply an oscillating force to selected atoms or residues during the course of a molecular dynamics simulation [32]. Conformations are extracted from the simulation then a fluctuation power spectrum for distant atoms and/or atom groups is generated by applying a Fourier transform to their atomic Cartesian coordinates. This spectrum is then compared to similar spectra obtained when no oscillating force is applied, or when varied pump force magnitudes, periods and atom selections are used. In theory, atoms connected to the excited protein region by well defined pathways of transmitted momentum will be more affected by the generating oscillating force than other residues. These same residues are more likely to participate in allosteric pathways connecting the excited region of the protein to other biologically relevant domains (e.g., an orthosteric site).

MD and the Perturbation-Response Scanning Method

The perturbation-response scanning method [33,34] can also be used to gain insight into specific allosteric pathways. As a first step, multiple random forces are sequentially applied to each node to capture the linear response of the whole system. The resulting changes in the whole-system conformation are compared to allostery-induced structural changes observed experimentally (e.g., as captured by crystallographic or NMR structures resolved in the presence and absence of an allosteric effector). When applied to the bacterial ferric binding protein [33], this method suggested that residues distant from the primary active site are nevertheless associated with experimentally validated structural changes. These residues are most likely to belong to distant allosteric sites.

MD and Dynamical Network Analysis

Dynamical network analysis is a cross-correlation method that draws upon molecular dynamics simulations and is constructed by considering all pairwise interactions between protein residues [2527]. Individual protein residues are each represented by a network node and all node pairs are connected by edges with lengths that are inversely proportional to the correlation between their motions (i.e., the distance in network space between highly correlated residues is shorter). Pathways between “source” and “sink” nodes, typically associated with the allosteric and orthosteric sites, respectively, are represented by simple node lists the length of a given pathway in network space is calculated by summing the lengths of the edges between adjacent list entries. The number of paths between sink and source nodes increases combinatorially with the total number of interconnected system nodes. While efficient algorithms exist for finding the single optimal (shortest) path between two nodes, the identification of near optimal pathways quickly becomes computationally intractable as the number of nodes increases. Several techniques are therefore used to make subsequent path finding efforts more feasible.

A recently developed algorithm [35•] is capable of identifying nodes that cannot possibly participate in pathways that are shorter than a user defined cutoff length. These nodes can be effectively ignored, further simplifying the network and thus speeding up subsequent path-finding efforts. Following network simplification, a recursive algorithm is used to identify paths between source and sink nodes. Dynamical network analysis is greatly facilitated by publicly available programs like WISP (http://nbcr.ucsd.edu/wisp), Figure 1, among others.

Figure 1.

Figure 1

A dynamical network analysis [35•] of the protein dimer HisH-HisF with bound effector molecule, PRFAR (effector not shown). The residues predicted to assist in propagating the allosteric signal between two residues of interest, Glu180:HisH (top) and Leu50:HisF (bottom), are depicted as white spheres along suboptimal signaling pathways, depicted as cyan lines. Expanding signaling pathways with lengths in network space near that of the optimal pathway are believed to facilitate an allosteric signal and may provide insight into differences between allosterically active and inactive forms.

Different methods sample different allosteric timescales, motions and pathway mechanisms

Comparisons of recent work utilizing these different methodologies demonstrate that there is no single method that best captures all the allosteric pathways for a given system. Presumably this is because of limitations in the conformational timescales and fluctuations sampled by each approach. These limitations become most apparent when NMR derived measures of protein dynamics are compared to those back calculated from computationally generated conformational ensembles.

Fuglestad and colleagues [36••] used a number of methods to characterize thrombin allosteric networks. They find that different methods correlate to fluctuations occurring on differing timescales. For example, an ensemble of conformations generated using accelerated molecular dynamics produced the highest degree of correlation between back-calculated residual dipolar couplings (RDCs) and NMR-derived RDCs for the thrombin-PPACK complex (R2 = 0.92). In contrast, conventional molecular dynamics and a single crystal structure did not yield the same degree of correlation (R2 = 0.8 and 0.72, respectively). RDCs are most sensitive to longer-timescale fluctuations, typically in the μs to ms regime. Comparison of NMR backbone order parameter values (S2), a measure of ps to ns timescale fluctuations, to S2 values back-calculated from the conventional molecular dynamics ensemble demonstrated the closest correlation with protein fluctuations on this timescale; results also seen in other work [37,38]. Interestingly, longer timescale motions are better correlated with the results of the frustratometer than are shorter-timescale events, suggesting topological methods are particularly well suited to characterizing conformational rearrangements that occur on these longer timescales.

Ever longer molecular dynamics simulations permit a direct comparison between simplified and atomistic techniques for studying allostery. In a recent study comparing soft modes derived from network models (such as ENM and ANM) to principal components derived from a millisecond-long Anton molecular dynamics simulation [39], the three lowest-frequency ANM modes captured the backbone conformations and motions observed in the first five principal components of the MD simulation [40••. In other words, the network method successfully identified the first three sub-states of the MD-generated energy landscape, but not the remaining five. The authors point out that the network method does not sample high energy, less populated sub-states, perhaps explaining why some were missed. Nevertheless, the network method did identify the most significant (i.e., most highly populated) sub-states at considerably less computational expense to the Anton simulations.

Fenwick and colleagues [41•] report that using a molecular dynamics simulation method highly restrained by NMR derived RDC data, termed ERNST, is more sensitive to identifying pathways correlated through hydrogen bond networks than using unrestrained MD. Application of the ERNST method and subsequent correlation-coefficient matrix analysis based on residue backbone torsion angles map weak, long-range motions through the β-sheet hydrogen-bond network that propagates communication between distant residues and the UBD binding sites [41•,42•]. Such motion also allows for propagation of an allosteric signal through a pathway without breaking the secondary-structure hydrogen bond.

Finally, DuBay et al. [43•] sample very fast timescale motion that do not result in large conformational changes upon an allosteric event but rather identify allosteric pathways that utilize population shifts in side-chain entropies. Rather than using molecular dynamics to develop a conformational ensemble of structures, the authors use Monte Carlo sampling of side-chain torsions on a static protein backbone. They find long-range correlations, up to 60Å. The authors point out these long-range correlations allow for allosteric signal to propagate through a variety of ways; van der Waals, hydrogen bond, salt bridge or implicit solvent. This both expands the types of inter-residue interactions typically considered for signal transmission and incorporates new ways of thinking about population shifts without large conformational changes but through entropic, thermodynamic mechanisms [44,45].

A couple of recent publications have combined topology-based methods with one or more simulations in order to increase conformational sampling, in what can be considered hybrid methods, and therefore capture longer-timescale correlated motions [4648]. For example, Weinkam and colleagues recently compared the structures of effector-bound and unbound crystal structures for three protein-effector systems [48]. They applied molecular dynamics to sample the conformations of the allosteric transition in order to identify residues affected by effector binding [48]. A similarity metric (QIdiff) thought to reflect residue involvement in the allosteric network was then calculated for each amino acid by comparing the inter-atomic contacts of each residue center of mass over the course of the trajectory to those of the apo and holo crystal structures. The residues predicted to participate in the allosteric signal were highly correlated with those that, when mutated, are known to disrupt allostery [48,49], thus validating the technique. Subsequent use of a machine-learning algorithm to predict allosteric-pathway residues based on QIdiff, as well as the inclusion of additional features into the model, led to even more robust predictions[49]. The Allosmod webserver, which provides tools for developing energy landscapes and analyzing simulations, can be found at http://modbase.compbio.ucsf.edu/allosmod/.

A hybrid method combining multiple simulation methodologies have also been reported and applied to the study of adenylate kinase [47].

Future Directions and Conclusions

Computational methods for identifying specific allosteric pathways provide useful atomic-level information that complements existing experimental techniques. While impressive progress has been made, the methods developed thus far are still foundational and each shows some gap in being fully predictive of experimental observations. As with other areas of computational chemistry, methods for prediction of allosteric pathways will benefit from improved and/or polarizable force field development. These methods, hinted at by Martin et al. [29•], will improve on modeling pathway components dependent upon electrostatics and may reveal a larger role in hydrogen bonding interactions than previously appreciated. Also, despite the recognized importance of water in protein folding and molecular recognition [50], the observation that certain waters can have very long residence times in protein interiors [51], and recent experimental studies exploring water’s role (reviewed in [45]), there is a surprising omission of water from covariance analysis or most of the computational allosteric pathway methods.

Nevertheless, as the accuracy and predictive power of computational methods improve, we expect them to become powerful tools for drug discovery. Current therapeutics focus principally on orthosteric sites, though the number targeting allosteric sites is on the rise [52,53]. Computational methods for identifying entire allosteric pathways, rather than just the endpoints of those pathways (e.g., allosteric and orthosteric binding pockets), may provide additional opportunities for the pharmacological modulation of pathogenic allosteric signaling [54•,55•].

Highlights.

  • Advances in computational methods for allosteric pathways are reviewed.

  • New methods are able to replicate many but not all experimental results.

  • Each method accesses specific dynamic timescales or conformational amplitudes.

  • Multiple webservers are available that automate allosteric pathway elucidation.

  • Atomistic level allosteric pathway identification will be of use in drug discovery.

Acknowledgments

This work was funded in part by the Director’s New Innovator Award Program (1-DP2-OD007237 to REA), the National Science Foundation through XSEDE Supercomputer resources provided via TG-CHE060073N to REA. Funding and support from the National Biomedical Computation Resource is provided through NIH P41 GM103426. Additional resources from the Center for Theoretical Biological Physics and the Keck II Computing and Visualization facility are gratefully acknowledged.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

Contributor Information

Victoria A. Feher, Email: vfeher@ucsd.edu.

Jacob Durrant, Email: jdurrant@ucsd.edu.

Adam T. Van Wart, Email: avanwart@ucsd.edu.

References and recommended reading

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

• of special interest

•• of outstanding interest

  • 1.Feher VA, Tzeng YL, Hoch JA, Cavanagh J. Identification of communication networks in Spo0F: a model for phosphorylation-induced conformational change and implications for activation of multiple domain bacterial response regulators. FEBS Lett. 1998;425:1–6. doi: 10.1016/s0014-5793(98)00182-3. [DOI] [PubMed] [Google Scholar]
  • 2.Feher VA, Cavanagh J. Millisecond-timescale motions contribute to the function of the bacterial response regulator protein Spo0F. Nature. 1999;400:289–293. doi: 10.1038/22357. [DOI] [PubMed] [Google Scholar]
  • 3.Volkman BF, Lipson D, Wemmer DE, Kern D. Two-state allosteric behavior in a single-domain signaling protein. Science. 2001;291:2429–2433. doi: 10.1126/science.291.5512.2429. [DOI] [PubMed] [Google Scholar]
  • 4.Monod J, Wyman J, Changeux JP. On the nature of allosteric transitions: A plausible model. J Mol Biol. 1965;12:88–118. doi: 10.1016/s0022-2836(65)80285-6. [DOI] [PubMed] [Google Scholar]
  • 5.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: 10.1110/ps.9.1.10. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 6.Pozzi N, Vogt AD, Gohara DW, Di Cera E. Conformational selection in trypsin-like proteases. Curr Opin Struct Biol. 2012;22:421–431. doi: 10.1016/j.sbi.2012.05.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 7••.Collier G, Ortiz V. Emerging computational approaches for the study of protein allostery. Arch Biochem Biophys. 2013;538:6–15. doi: 10.1016/j.abb.2013.07.025. An excellent review describing recent progress in experimental and computational methods for the study of protein allostery. [DOI] [PubMed] [Google Scholar]
  • 8.Spyrakis F, BidonChanal A, Barril X, Luque FJ. Protein flexibility and ligand recognition: challenges for molecular modeling. Curr Top Med Chem. 2011;11:192–210. doi: 10.2174/156802611794863571. [DOI] [PubMed] [Google Scholar]
  • 9.Lockless SW, Ranganathan R. Evolutionarily conserved pathways of energetic connectivity in protein families. Science. 1999;286:295–299. doi: 10.1126/science.286.5438.295. [DOI] [PubMed] [Google Scholar]
  • 10.Daily MD, Upadhyaya TJ, Gray JJ. Contact rearrangements form coupled networks from local motions in allosteric proteins. Proteins. 2008;71:455–466. doi: 10.1002/prot.21800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 11.Kannan N, Vishveshwara S. Identification of side-chain clusters in protein structures by a graph spectral method. J Mol Biol. 1999;292:441–464. doi: 10.1006/jmbi.1999.3058. [DOI] [PubMed] [Google Scholar]
  • 12•.van den Bedem H, Bhabha G, Yang K, Wright PE, Fraser JS. Automated identification of functional dynamic contact networks from X-ray crystallography. Nature Methods. 2013;10:896–U110. doi: 10.1038/nmeth.2592. This paper describes a new algorithm that evaluates alternate side-chain conformers while constrained by crystallographic electron density. Remarkable agreement with experimental mutagenesis data is found. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13.Boehr DD, Schnell JR, McElheny D, Bae SH, Duggan BM, Benkovic SJ, Dyson HJ, Wright PE. A distal mutation perturbs dynamic amino acid networks in dihydrofolate reductase. Biochemistry. 2013;52:4605–4619. doi: 10.1021/bi400563c. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Ferreiro DU, Hegler JA, Komives EA, Wolynes PG. Localizing frustration in native proteins and protein assemblies. Proceedings of the National Academy of Sciences of the United States of America. 2007;104:19819–19824. doi: 10.1073/pnas.0709915104. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Ferreiro DU, Hegler JA, Komives EA, Wolynes PG. On the role of frustration in the energy landscapes of allosteric proteins. Proceedings of the National Academy of Sciences of the United States of America. 2011;108:3499–3503. doi: 10.1073/pnas.1018980108. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 16•.Jenik M, Parra RG, Radusky LG, Turjanski A, Wolynes PG, Ferreiro DU. Protein frustratometer: a tool to localize energetic frustration in protein molecules. Nucleic Acids Research. 2012;40:W348–W351. doi: 10.1093/nar/gks447. This paper describes the new publicly available webserver for generating a frustration index for each residue and various useful graphical outputs. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 17.Brooks B, Karplus M. Harmonic dynamics of proteins: normal modes and fluctuations in bovine pancreatic trypsin inhibitor. Proc Natl Acad Sci U S A. 1983;80:6571–6575. doi: 10.1073/pnas.80.21.6571. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 18.Go N, Noguti T, Nishikawa T. Dynamics of a small globular protein in terms of low-frequency vibrational modes. Proc Natl Acad Sci U S A. 1983;80:3696–3700. doi: 10.1073/pnas.80.12.3696. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19.Levitt M, Sander C, Stern PS. The normal-modes of a protein-native bovine pancreatic trypsin-inhibitor. Intl J Quantum Chem. 1983;(Suppl 10):181–199. [Google Scholar]
  • 20.Garcia AE. Large-amplitude nonlinear motions in proteins. Phys Rev Lett. 1992;68:2696–2699. doi: 10.1103/PhysRevLett.68.2696. [DOI] [PubMed] [Google Scholar]
  • 21.Amadei A, Linssen AB, Berendsen HJ. Essential dynamics of proteins. Proteins. 1993;17:412–425. doi: 10.1002/prot.340170408. [DOI] [PubMed] [Google Scholar]
  • 22.Lange OF, Grubmuller H. Full correlation analysis of conformational protein dynamics. Proteins. 2008;70:1294–1312. doi: 10.1002/prot.21618. [DOI] [PubMed] [Google Scholar]
  • 23.Girvan M, Newman ME. Community structure in social and biological networks. Proc Natl Acad Sci U S A. 2002;99:7821–7826. doi: 10.1073/pnas.122653799. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.McClendon CL, Friedland G, Mobley DL, Amirkhani H, Jacobson MP. Quantifying correlations between allosteric sites in thermodynamic ensembles. J Chem Theory Comput. 2009;5:2486–2502. doi: 10.1021/ct9001812. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Sethi A, Eargle J, Black AA, Luthey-Schulten Z. Dynamical networks in tRNA:protein complexes. Proc Natl Acad Sci U S A. 2009;106:6620–6625. doi: 10.1073/pnas.0810961106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Eargle J, Luthey-Schulten Z. NetworkView: 3D display and analysis of protein RNA interaction networks. Bioinformatics. 2012;28:3000–3001. doi: 10.1093/bioinformatics/bts546. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 27.VanWart AT, Eargle J, Luthey-Schulten Z, Amaro RE. Exploring residue component contributions to dynamical network models of allostery. Journal of Chemical Theory and Computation. 2012;8:2949–2961. doi: 10.1021/ct300377a. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Bahar I, Lezon TR, Yang LW, Eyal E. Global dynamics of proteins: Bridging between structure and function. Annual Review of Biophysics. 2010;39:23–42. doi: 10.1146/annurev.biophys.093008.131258. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 29•.Martin DR, Matyushov DV. Solvated dissipative electro-elastic network model of hydrated proteins. Journal of Chemical Physics. 2012;137:e165101. doi: 10.1063/1.4759105. This work represents a rare example where protein electrostatics and hydration are included in the allosteric network elucidation. [DOI] [PubMed] [Google Scholar]
  • 30.Tehver R, Chen J, Thirumalai D. Allostery wiring diagrams in the transitions that drive the GroEL reaction cycle. J Mol Biol. 2009;387:390–406. doi: 10.1016/j.jmb.2008.12.032. [DOI] [PubMed] [Google Scholar]
  • 31.Zheng W, Brooks BR, Thirumalai D. Allosteric transitions in the chaperonin GroEL are captured by a dominant normal mode that is most robust to sequence variations. Biophysical Journal. 2007;93:2289–2299. doi: 10.1529/biophysj.107.105270. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 32.Sharp K, Skinner JJ. Pump-probe molecular dynamics as a tool for studying protein motion and long range coupling. Proteins. 2006;65:347–361. doi: 10.1002/prot.21146. [DOI] [PubMed] [Google Scholar]
  • 33.Atilgan C, Atilgan AR. Perturbation-response scanning reveals ligand entry-exit mechanisms of ferric binding protein. PLoS Computational Biology. 2009;5:e1000544. doi: 10.1371/journal.pcbi.1000544. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 34.Gerek ZN, Ozkan SB. Change in allosteric network affects binding affinities of PDZ domains: analysis through perturbation response scanning. PLoS Comput Biol. 2011;7:e1002154. doi: 10.1371/journal.pcbi.1002154. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 35•.VanWart AT, Durrant J, Votapka L, Amaro RE. Weighted Implementation of Suboptimal Paths (WISP): An optimized algorithm and tool for dynamical network analysis. Journal of Chemical Theory and Computation. 2014;10:511–517. doi: 10.1021/ct4008603. This paper describes a novel method for elucidating multiple allosteric propagation paths and for reducing the computational time required to determine dynamical networks through functionalized correlation matrices. A public webserver is provided. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36••.Fuglestad B, Gasper PM, McCammon JA, Markwick PR, Komives EA. Correlated motions and residual frustration in thrombin. J Phys Chem B. 2013;117:12857–12863. doi: 10.1021/jp402107u. This paper provides a rare example where a range of computational methods for defining allosteric pathways are applied to an experimentally well characterized protein. The results are systematically compared to experimental observables and conclusions generalized for the application of these methods to other systems. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 37.Kasinath V, Sharp KA, Wand AJ. Microscopic insights into the NMR relaxation-based protein conformational entropy meter. J Am Chem Soc. 2013;135:15092–15100. doi: 10.1021/ja405200u. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Chen Y, Campbell SL, Dokholyan NV. Deciphering protein dynamics from NMR data using explicit structure sampling and selection. Biophys J. 2007;93:2300–2306. doi: 10.1529/biophysj.107.104174. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Shaw DE, Maragakis P, Lindorff-Larsen K, Piana S, Dror RO, Eastwood MP, Bank JA, Jumper JM, Salmon JK, Shan Y, et al. Atomic-level characterization of the structural dynamics of proteins. Science. 2010;330:341–346. doi: 10.1126/science.1187409. [DOI] [PubMed] [Google Scholar]
  • 40••.Gur M, Zomot E, Bahar I. Global motions exhibited by proteins in micro- to milliseconds simulations concur with anisotropic network model predictions. Journal of Chemical Physics. 2013;139:e121912. doi: 10.1063/1.4816375. By drawing comparisons between results for computational coarse-grained methods such as anisotropic network models and all atom Anton MD simulations for the same BPTI protein system, these authors demonstrate the relative strengths and limitations of each method. Coarse-grained methods, typically used for large scale conformational change simulation, does identify the most populated BPTI conformations observed by long time-scale MD simulation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41•.Fenwick RB, Esteban-Martin S, Richter B, Lee D, Walter KFA, Milovanovic D, Becker S, Lakomek NA, Griesinger C, Salvatella X. Weak long-range correlated motions in a surface patch of ubiquitin involved in molecular recognition. Journal of the American Chemical Society. 2011;133:10336–10339. doi: 10.1021/ja200461n. This paper describes a restrained molecular dynamics methodology, coined ERNST, that utilizes NMR experimental parameters as restraints. Their focus on backbone torsion angle correlation analyses identify residues involved in concerted pathways that signal in part through β-sheet hydrogen bonds. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 42•.Fenwick RB, Esteban-Martin S, Salvatella X. Understanding biomolecular motion, recognition, and allostery by use of conformational ensembles. European Biophysics Journal with Biophysics Letters. 2011;40:1339–1355. doi: 10.1007/s00249-011-0754-8. This is a comprehensive review on molecular dynamics simulations methods used for generating conformational ensembles and their utility in understanding protein ligand recognition and allostery. The authors also provide an excellent overview of various reweighting methods, used both during and following simulation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43•.Dubay KH, Bothma JP, Geissler PL. Long-range intra-protein communication can be transmitted by correlated side-chain fluctuations alone. PLoS Comput Biol. 2011;7:e1002168. doi: 10.1371/journal.pcbi.1002168. This paper uses Monte Carlo simulation exclusively on protein side-chain to explore allosteric networks. It expands the type of side-chain interactions typically considered beyond hydrophobic to include hydrogen bonds and salt bridges. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 44.Cooper A, Dryden DT. Allostery without conformational change. A plausible model. Eur Biophys J. 1984;11:103–109. doi: 10.1007/BF00276625. [DOI] [PubMed] [Google Scholar]
  • 45.Wand AJ. The dark energy of proteins comes to light: conformational entropy and its role in protein function revealed by NMR relaxation. Curr Opin Struct Biol. 2013;23:75–81. doi: 10.1016/j.sbi.2012.11.005. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Armenta-Medina D, Perez-Rueda E, Segovia L. Identification of functional motions in the adenylate kinase (ADK) protein family by computational hybrid approaches. Proteins-Structure Function and Bioinformatics. 2011;79:1662–1671. doi: 10.1002/prot.22995. [DOI] [PubMed] [Google Scholar]
  • 47.Gur M, Madura JD, Bahar I. Global transitions of proteins explored by a multiscale hybrid methodology: Application to adenylate kinase. Biophysical Journal. 2013;105:1643–1652. doi: 10.1016/j.bpj.2013.07.058. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Weinkam P, Pons J, Sali A. Structure-based model of allostery predicts coupling between distant sites. Proceedings of the National Academy of Sciences of the United States of America. 2012;109:4875–4880. doi: 10.1073/pnas.1116274109. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 49.Weinkam P, Chen YC, Pons J, Sali A. Impact of mutations on the allosteric conformational equilibrium. Journal of Molecular Biology. 2013;425:647–661. doi: 10.1016/j.jmb.2012.11.041. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 50.Levy Y, Onuchic JN. Water mediation in protein folding and molecular recognition. Annu Rev Biophys Biomol Struct. 2006;35:389–415. doi: 10.1146/annurev.biophys.35.040405.102134. [DOI] [PubMed] [Google Scholar]
  • 51.Otting G, Liepinsh E, Wuthrich K. Protein hydration in aqueous solution. Science. 1991;254:974–980. doi: 10.1126/science.1948083. [DOI] [PubMed] [Google Scholar]
  • 52.Zorn JA, Wells JA. Turning enzymes ON with small molecules. Nat Chem Biol. 2010;6:179–188. doi: 10.1038/nchembio.318. [DOI] [PubMed] [Google Scholar]
  • 53.Conn PJ, Christopoulos A, Lindsley CW. Allosteric modulators of GPCRs: a novel approach for the treatment of CNS disorders. Nat Rev Drug Discov. 2009;8:41–54. doi: 10.1038/nrd2760. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 54•.Kar G, Keskin O, Gursoy A, Nussinov R. Allostery and population shift in drug discovery. Curr Opin Pharmacol. 2010;10:715–722. doi: 10.1016/j.coph.2010.09.002. This review explores the impact of allostery in the disease state and cellular function. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55•.Szilagyi A, Nussinov R, Csermely P. Allo-network drugs: extension of the allosteric drug concept to protein- protein interaction and signaling networks. Curr Top Med Chem. 2013;13:64–77. doi: 10.2174/1568026611313010007. A comprehensive review on the development of computational methodologies for allosteric network elucidation and enumeration of strategies for the use of these networks in drug discovery. Interesting proposals for allo-drug finding at the systems-level are suggested. [DOI] [PubMed] [Google Scholar]

RESOURCES