Abstract
The coupling between the mechanical properties of enzymes and their biological activity is a well-established feature that has been the object of numerous experimental and theoretical works. In particular, recent experiments show that enzymatic function can be modulated anisotropically by mechanical stress. We study such phenomena using a method for investigating local flexibility on the residue scale that combines a reduced protein representation with Brownian dynamics simulations. We performed calculations on the enzyme guanylate kinase to study its mechanical response when submitted to anisotropic deformations. The resulting modifications of the protein's rigidity profile can be related to the changes in substrate binding affinity observed experimentally. Further analysis of the principal components of motion of the trajectories shows how the application of a mechanical constraint on the protein can disrupt its dynamics, thus leading to a decrease of the enzyme's catalytic rate. Eventually, a systematic probe of the protein surface led to the prediction of potential hotspots where the application of an external constraint would produce a large functional response both from the mechanical and dynamical points of view. Such enzyme-engineering approaches open the possibility to tune catalytic function by varying selected external forces.
Introduction
The importance of protein flexibility and dynamics for the understanding of protein function has now been clearly established (1–4). During the execution of their biological function, proteins can be subjected to forces and their mechanical properties evolved in response to fit this selection pressure. Experimentally, many techniques, such as optical and magnetic tweezers (5–7) or atomic force microscopy (8–10), make it possible to probe biomolecular mechanics directly on the single-molecule level (11). In particular, experiments with linkages other than the usual N-to-C-terminal have shown how these mechanical properties strongly depend on the loading geometry (12–15).
Although the first experiments mostly investigated the mechanical response of proteins and the sequence of unfolding events that would result from the application of a force (8), recent setups have focused more on their functional response, thus leading to the field of mechanoenzymatics (16,17). In this perspective, the mechanism of allosteric control (18,19) of an enzyme, which plays a crucial part in signaling pathways in the cell (20), can now be studied experimentally via the building of protein-DNA chimeras where a DNA molecular spring is coupled to the protein at specific locations on its surface (21–23). Through this allosteric spring probe (ASP), one can affect the static and dynamic conformations of the protein and follow its functional response to the application of an external stress (24).
From the theoretical point of view, atomic coordinate-based methods, such as constrained molecular dynamics simulations, can mimic force-extension experiments but operate on much shorter timescales and remain computationally expensive (25–29). Therefore, lower-resolution models have been widely used in recent years to study protein dynamics (30–35). These coarse-grained representations comprise the elastic network model (ENM) (36,37), which reduces the protein to a set of pseudoatoms with pairs below a given cutoff distance being linked by Gaussian springs. Despite their simplicity, these models led to many results concerning protein mechanics and dynamics (38–44). Recently, coarse-grained approaches were used to successfully model the anisotropy of the mechanical response of proteins subjected to an external force (45–47).
In this work, we used a method combining a coarse-grained protein representation and Brownian dynamics simulations. This approach was previously successfully applied to model the mechanical response of the green fluorescent protein to understand the single-molecule experiments carried out by Dietz et al. (14,47). Here, we investigated the enzyme guanylate kinase (GK), which was studied by Tseng et al. via the ASP approach (48). GK is an essential enzyme that catalyzes the transfer of a phosphate group from adenosine triphosphate (ATP) to guanosine monophosphate (GMP) (49). Upon substrate binding, GK undergoes a structural transition from the open to the closed state through a movement of the two lobes formed by the LID and GMP domains (see Fig. 1), leading to an ∼1 nm conformational change (50,51). With DNA springs anchored on three different locations on the protein surface, Tseng et al. determined for each case the changes in substrate binding affinities and catalytic rate constant resulting from the directional stress exerted on the protein. They showed that the functional response strongly depends on the direction of load. Using our molecular modeling approach, we investigated the mechanics and dynamics of GK when subjected to an external constraint and related our results to the variations of the enzymatic activity observed experimentally.
Figure 1.
A cartoon representation of GK with the pulling directions tested experimentally by Tseng et al. (48) The color coding of the protein is according to the domain definitions of Hible et al. (67) The GMP and ATP binding sites are located at the GMP/CORE and CORE/LID interfaces, respectively. The arrow indicates the direction of the opening/closing transition, which constitutes the first mode of motion of the protein. The images in this figure and in the upper part of Fig. 5 were prepared using visual molecular dynamics (71).
Computational Details
Brownian Dynamics simulations
Rigidity profile of a protein
Coarse-grained Brownian Dynamics (BD) simulations were run using a modified version of the ProPHet (probing protein heterogeneity) program (41,42), where an external mechanical constraint can be applied between two residues. In this approach, the protein is represented using an ENM. Diverging from most common coarse-grained models, where each residue is described by a single pseudoatom (52), we chose a more detailed representation (53) that involves up to three pseudoatoms per residue and enables different amino acids to be distinguished. Pseudoatoms closer than the cutoff parameter, Rc = 9 Å, are joined by Gaussian springs that all have identical spring constants of γ = 0.42 N m−1 (0.6 kcal mol−1 Å−2). The springs are taken to be relaxed for the experimentally observed conformation of the protein, in this case the crystallographic structure of guanylate kinase from Mycobacterium tuberculosis in its open conformation available in the protein data bank with the code 1S4Q.
Mechanical properties are obtained from 200,000 BD steps at 300 K. The simulations are analyzed in terms of the fluctuations of the mean distance between each pseudoatom belonging to a given amino acid and the pseudoatoms belonging to the remaining residues of the protein. The inverse of these fluctuations yields an effective force constant ki that describes the ease of moving a pseudoatom with respect to the overall protein structure:
where 〈〉 denotes an average taken over the whole simulation and di = 〈dij〉j∗ is the average distance from particle i to the other particles j in the protein (the sum over j∗ implies the exclusion of the pseudoatoms belonging to residue i). The distances between the Cα pseudoatom of residue i and the Cα pseudoatoms of the adjacent residues i − 1 and i + 1 are excluded, since the corresponding distances are virtually constant. The force constant for each residue k is the average of the force constants for all its constituent pseudoatoms i. We will use the term rigidity profile to describe the ordered set of force constants for all the residues of the protein.
Applying an external constraint on the protein
Whereas in our previous work on the green fluorescent protein (47), the mechanical stress was simply modeled by applying a constant force between the Cα pseudoatoms of the corresponding residues, in this study, we chose to model the external constraint by adding to the ENM representation a supplementary spring termed the constraint spring in opposition to the structural springs resulting from the original conformation of the protein,. This way we could model more accurately the experiment of Tseng et al. (48), where DNA molecular springs of identical length (60 bp) were used at three different locations on the surface of the protein to apply a controlled mechanical stress (see Fig. 1). From the available experimental data regarding the contour length of the DNA spring (200 Å) and the amount of elastic energy that resides in the protein (54) (∼1 kT), we derived for our constraint spring the parameters equilibrium length (LC = 150 Å) and spring constant (γC = 0.84 N m−1) (1.2 kcal mol−1 Å−2). This constraint spring was added between the Cα pseudoatoms of the anchor residues of the three locations tested in the experiment, Thr75/Arg171, Cys40/Arg171, and Cys40/Lys130 (see Fig. 1).
Principal component analysis of the coarse-grained trajectories
The BD trajectories for the protein without (relaxed protein) and with the application of an external force (protein under stress) were investigated using principal component analysis (PCA) (55–58) with tools from the Gromacs (59–61) software package. In particular, we calculated the inner product matrices of the ten first eigenvectors, which always cover >89% of the total variance of the protein (see Fig. S1, Fig. S2, and Table S1 in the Supporting Material) of each constrained trajectory with the 10 first eigenvectors of the relaxed trajectory to assess how the mechanical constraint affects the leading modes of motion of the protein.
Systematic scan of the protein surface
To determine whether the deformations that were studied experimentally are representative of the full heterogeneity of the GK structure, we performed a more systematic study of residue-pair deformations. The selection of representative pairs is first narrowed by limiting our choice to surface residues, that is, residues with at least 5% solvent accessibility (62) (as calculated by the NACCESS program (63)), which are thus amenable to experimental study. Second, we chose residue pairs separated by at least 20 Å and 30 amino acids in the primary sequence. Last, we eliminated residue pairs that differed from already-selected pairs by fewer than five residues in the primary sequence in either of the constituent residues. This method led to the selection of 236 residue pairs, which were all tested with the constraint spring described earlier (see Fig. S3).
Results
Mechanical properties of guanylate kinase
The rigidity profile of GK is represented in Fig. 2 a. It is worthy of remark that most of the force-constant peaks from the first half of the protein sequence correspond to residues belonging to ligand-binding sites, such as Lys34 and Glu119 for the ATP/Mg2+-binding site, or Ser53, Glu88, and Thr101 for the GMP-binding site. In particular, note the peaks corresponding to Ser27 and Lys34, two residues surrounding the flexible P-loop, a highly conserved motif that binds the β-phosphate of the ATP donor in nucleoside monophosphate kinases (64,65); and the highly rigid area on the β7-sheet, around Glu119, which corresponds to residues interacting with GMP in the closed conformations of GK from Saccharomyces cerevisiae (50) and from Mus musculus (66).
Figure 2.
(a) Rigidity profile of GK when no mechanical stress is applied on the protein. (b) Variation of the force-constant profile upon applying an external constraint on the protein in pulling directions 75/171 (upper), 40/171 (middle), and 40/130 (lower). The black horizontal bars at the upper left of Fig. 2, a and b,indicate the position of the P-loop and the β7-sheet along the sequence.
The variation of the force-constant profile of the protein upon mechanical stress is represented in Fig. 2 b. The P-loop base and the β7-sheet are the protein segments whose mechanical properties are the most sensitive to the application of an external stress, independent of the pulling direction. The mechanical response of GK is nevertheless markedly anisotropic. Although the 40/171 and 40/130 pulling directions only lead to weak (<20 kcal mol−1 Å−2) variations in the force constant of the residues, stressing the protein along the 75/171 direction results in a strong rigidity decrease of Ser27 (−57 kcal mol−1 Å−2), Lys34 (−31 kcal mol−1 Å−2), Glu119 (−64 kcal mol−1 Å−2), and Val120 (-78 kcal mol−1 Å-2). The initial distances between the Cα atoms from the 75/171, 40/171, and 40/130 residue pairs are 35.7 Å, 27 Å, and 28 Å, respectively. This indicates that the stress exerted by the constraint spring, which has an equilibrium length of 150 Å, should intrinsically be less pronounced for the 75/171 direction than for the 40/171 and 40/130 directions of load. However, the highly anisotropic architecture of the protein leads to the opposite effect, with the 75/171 direction inducing the most important changes in the enzyme's mechanics.
Dynamics of the constrained protein
In a second step, we used the PCA approach to compute the inner product of the 10 first eigenvectors of a constrained trajectory with the eigenvectors of the relaxed trajectory. The resulting matrices for the 75/171, 40/171, and 40/130 pulling directions are plotted in Fig. 3, a–c, respectively. Table 1 summarizes the overlap between the covariance matrices (CMOs) of the relaxed and constrained trajectories. We can see how the application of an external stress along the 40/130 direction feebly modifies the modes of motion of the protein (CMO of 0.83), thus resulting in an almost diagonal matrix. On the other hand, applying an external constraint along the 75/171 and 40/171 directions induces some important disruptions of the protein's dynamics, but in different ways. Although pulling the protein along the 75/171 direction leads to a generally more important perturbation of GK movements compared with the 40/171 direction, with CMOs of 0.65 and 0.80, respectively, it turns out that the main functional mode of motion of the enzyme, corresponding to its opening and closing around the GMP and ATP binding sites, is more preserved for the 75/171 than for the 40/171 direction. This is shown by the projections of the first eigenvector of the constrained trajectory on the first eigenvector of the relaxed trajectory, which amount to overlaps of 0.83 and 0.77, respectively. This variation in the disruption of the enzyme dynamics is easily understandable, since the 75/171 direction actually coincides with opening and closing motions of GK, whereas pulling the protein via a constraint spring anchored on Cys40 and Arg171 introduces a new direction of motion with a component orthogonal to the main enzymatic movement.
Figure 3.
Inner-product matrices of the 10 first eigenvectors of the constrained trajectories over the relaxed trajectory. The pulling directions are (a) 75/171, (b) 40/171, (c) 40/130, and (d) 65/122. The color scale ranges from 0 to 1.
Table 1.
Overlap of constrained trajectories with relaxed trajectory along the 10 first modes of motions of the protein.
Pulling direction | Covariance matrix overlap (CMO) | Overlap of first eigenvectors |
---|---|---|
Thr75-Arg171∗ | 0.65 | 0.83 |
Cys40-Arg171∗ | 0.80 | 0.77 |
Cys40-Lys130∗ | 0.83 | 0.97 |
Asp65-Leu122 | 0.66 | 0.55 |
Asterisks identify the pulling directions experimentally tested by Tseng et al. (48)
Prediction of hotspots on the protein surface
We performed a systematic search of the protein mechanical response to the application of an external constraint by probing 236 new nonredundant pulling directions via residues anchored all over the surface of the protein. The resulting variations of the force constant of each residue are plotted in Fig. 4. From the qualitative point of view, most directions of load result in variations of the rigidity profile that are similar to those previously observed for the experimentally tested directions. Once again, the most important changes in the force constants of the residues occur in the P-loop and β7-sheet areas, which usually undergo a strong increase in flexibility.
Figure 4.
Distribution of the force-constant variation over all the pulling directions that have been modeled. Δk is expressed in kcal mol−1 Å−2. The black horizontal bars above the figure indicate the position of the P-loop and the β7-sheet along the sequence. The extended pairs have been ordered starting with the direction leading to the largest average perturbation, 〈abs(Δk)〉, of the GK rigidity profile. A negative/positive, value of Δk denotes a decrease/increase, of the rigidity of the residue.
The pulling direction that led to the most important perturbation of the rigidity profile (in terms of both the maximum force constant variation and the average perturbation of the profile) was formed by Asp65 and Leu122 (see Fig. 5). As can be seen in Fig. 5 (lower), the application of a constraint spring between these two residues induces an important disruption of the GMP binding site, with force-constant decreases beyond 100 kcal mol−1 Å−2 for Glu119 and Val120, which is not surprising since the anchor residues actually surround the catalytic site. It is of interest that this new pulling direction also strongly disturbs the enzyme's dynamics. As we can see from Fig. 3 d and Table 1, the 65/122 direction yields a CMO between the constrained and the relaxed trajectories of 0.66, whereas the projection of the first eigenvector is now reduced to 0.55, a much lower value than previously obtained for the experimental directions of load.
Figure 5.
(Upper) Cartoon representation of GK with the 65/122 pulling direction. The ellipse indicates the location of the GMP-binding site. (Lower) Variations in force constant (kcal mol−1 Å−2) as the protein moves along the 65/122 pulling direction. The black horizontal bars above the figure indicate the position of the P-loop and the β7-sheet along the sequence.
Discussion and Conclusion
In the ASP experiment, protein-DNA chimeras are used to strain the conformation of a protein, thus potentially providing further insight into the mechanism of allosteric control of biological function. In their work, Tseng et al. (48) applied a mechanical constraint at three different locations on the surface of GK, Thr75/Arg171, Cys40/Arg171, and Cys40/Lys130. These experiments yielded different results in terms of changes of the enzymatic activity: The 75/171 pulling direction induced a decrease in binding affinity for GMP, thus increasing the Michaelis-Menten constant, KG, which was measured by GMP titration experiments, but having little or no effect on KA (the binding affinity for ATP) and kcat (the catalytic rate of the enzymatic reaction). For the 40/171 direction, Tseng et al. (48) observed a decrease of kcat, whereas here, KA and KG were not affected. Finally, for the 40/130 pulling direction, no noticeable effect was observed for KG, KA, or kcat. In this study, we combined a coarse-grained protein representation and BD simulations to investigate the mechanical and dynamical response of GK when an external stress is applied on the protein. During the simulations, the spring network oscillates around its equilibrium state within a limited range, with the deformations amounting to an ∼1-Å root-mean-square deviation from the average conformation of the protein. Our model is therefore well adapted to describe the experiment of Tseng et al., where the protein's structure undergoes very few changes.
The force-constant profiles, which were obtained for trajectories with and without the application of a mechanical constraint, present rigidity peaks that correspond to residues belonging to the ligand-binding sites, thus stressing once more the importance of the catalytic site's stiffness for enzymatic activity (41,42). From the qualitative point of view, the force-constant variations observed for the protein under stress were mainly located around the P-loop and the β7-sheet. Quantitatively, however, only the 75/171 pulling direction led to an important decrease of the rigidity of residues Ile118 to Asp121, which belong to the GMP binding site and do normally form interactions with the ligand in the closed conformation of the protein (67). Since such catalytic residues require an enhanced rigidity for the execution of their biological function (68–70), this disruption of the mechanical properties of the GMP binding site provides a first explanation for the decrease of the GMP binding affinity observed experimentally with the 75/171 chimera.
We then used PCA to study the variation in protein dynamics induced by the external stress. The resulting inner-product matrices obtained for the first 10 eigenvectors indicate that the 75/171 pulling direction leads to the most important perturbations of the general enzyme dynamics. However, if we focus only on the first mode of motion of the protein, which corresponds to the opening and closing movement of the LID and GMP domains over the CORE domain, and which is essential for GK to perform its catalytic function, it turns out that this mode is most perturbed when load is applied along the 40/171 direction. Once again this result is in agreement with the experimental data, where the 40/171 mutant alone presented a decrease of its catalytic rate, kcat. All in all, the variations in enzymatic activity observed via the ASP experiments can either be related to some local mechanical perturbation of a substrate binding site (in the case of the GMP binding affinity), or to more global changes in the protein large-amplitude movements (for the catalytic rate constant).
Eventually, since Tseng et al. raised the question of the prediction of hotspots at the surface of the protein where a mechanical perturbation would produce a large functional response, we scanned 236 nonredundant locations on the protein surface. From a mechanical point of view, all pairs led to similar changes in the protein properties, with a rigidity decrease in the P-loop and β7-sheet regions. We were yet able to single out a specific residue pair that would be interesting to study experimentally, Asp65-Leu122. Since these two residues surround the GMP binding site, pulling in the 65/122 direction should yield an important disruption of its rigidity. It is also noteworthy that the neighboring Asp121 residue is implicated in GMP binding and initiation of the enzyme's closure (67). From a dynamical point of view, it appears that this direction of load also induces a strong perturbation of the protein's first mode of motion. This means that, were experiments performed on a 65/122 mutant, one should observe a decrease in both the enzyme's binding affinity for GMP and its catalytic rate. We furthermore tried to assess from the results of the systematic scan whether a particular pulling direction could specifically disrupt the ATP binding site while leaving the GMP binding site intact. This would result in a protein-DNA chimera where KA, but not KG, would be affected. In terms of mechanics, this meant finding a direction of load leading to a decrease in rigidity of the protein around the P-loop but not in the β7-sheet area. However, we could not find any residue pairs that would satisfy such a criterion, and it seems that the mechanical properties of these two elements of GK are tightly coupled and cannot be modulated independent of each other. To mechanically separate these subunits, one would probably have to disrupt the set of interactions that bind together the α1-helix and the β1- and β7-sheets via site-directed mutagenesis. In this perspective, residues Leu26, Lys34, Val38, Leu117, and Glu119, whose side chains are directed toward the center of the CORE domain, appear to be potential candidates. It is interesting that all these residues also present important force constants (>35 kcal mol−1 Å−2) in the rigidity profile of GK, which supports the idea that they might play an important part in the enzyme's structural stability. However, since Lys34 and Glu119 are also involved in ATP- and GMP-binding, respectively, it is unlikely that one could perturb the protein's intrinsic mechanics without disturbing its biological activity as well.
Altogether, we showed how a simple protein representation combined with BD simulations can yield a molecular level picture of the way the application of an external constraint on the GK enzyme perturbs its mechanical properties and dynamics. These perturbations can then be related to the changes observed experimentally in parameters KG, KA, and kcat of the enzymatic reaction, helping us to understand the origin of the anisotropic functional response of the protein to various pulling directions. It is interesting to note that the apparent origin of the variations in thermodynamic parameter KG is a disturbance of the protein's mechanical properties around its GMP binding site, whereas the changes in kinetic parameter kcat arise from a perturbation of the protein dynamics, more precisely from the disruption of its first mode of motion, which defines the opening and closing movement performed by the protein when it undergoes a catalytic cycle. Our model can also be used in a predictive outlook. From a general search on the protein surface, we could suggest a new direction of load that should lead to the simultaneous perturbation of the two parameters KG and kcat, an effect not observed with the protein-DNA chimeras produced thus far.
Acknowledgments
This work was funded by the French Agency for Research (grants ANR-06-PCVI-0025 and ANR-07-CIS7-003).
Contributor Information
Sophie Sacquin-Mora, Email: sacquin@ibpc.fr.
Marc Baaden, Email: baaden@smplinux.de.
Supporting Material
References
- 1.Daniel R.M., Dunn R.V., Smith J.C. The role of dynamics in enzyme activity. Annu. Rev. Biophys. Biomol. Struct. 2003;32:69–92. doi: 10.1146/annurev.biophys.32.110601.142445. [DOI] [PubMed] [Google Scholar]
- 2.Parak F.G. Physical aspects of protein dynamics. Rep. Prog. Phys. 2003;66:103–129. [Google Scholar]
- 3.Bustamante C., Chemla Y.R., Izhaky D. Mechanical processes in biochemistry. Annu. Rev. Biochem. 2004;73:705–748. doi: 10.1146/annurev.biochem.72.121801.161542. [DOI] [PubMed] [Google Scholar]
- 4.Hammes-Schiffer S., Benkovic S.J. Relating protein motion to catalysis. Annu. Rev. Biochem. 2006;75:519–541. doi: 10.1146/annurev.biochem.75.103004.142800. [DOI] [PubMed] [Google Scholar]
- 5.Charvin G., Allemand J.F., Croquette V. Twisting DNA: single molecule studies. Contemp. Phys. 2004;45:383–403. [Google Scholar]
- 6.Koster D.A., Croquette V., Dekker N.H. Friction and torque govern the relaxation of DNA supercoils by eukaryotic topoisomerase IB. Nature. 2005;434:671–674. doi: 10.1038/nature03395. [DOI] [PubMed] [Google Scholar]
- 7.Moffitt J.R., Chemla Y.R., Bustamante C. Recent advances in optical tweezers. Annu. Rev. Biochem. 2008;77:205–228. doi: 10.1146/annurev.biochem.77.043007.090225. [DOI] [PubMed] [Google Scholar]
- 8.Brockwell D.J. Probing the mechanical stability of proteins using the atomic force microscope. Biochem. Soc. Trans. 2007;35:1564–1568. doi: 10.1042/BST0351564. [DOI] [PubMed] [Google Scholar]
- 9.Puchner E.M., Gaub H.E. Force and function: probing proteins with AFM-based force spectroscopy. Curr. Opin. Struct. Biol. 2009;19:605–614. doi: 10.1016/j.sbi.2009.09.005. [DOI] [PubMed] [Google Scholar]
- 10.Galera-Prat A., Gómez-Sicilia A., Carrión-Vázquez M. Understanding biology by stretching proteins: recent progress. Curr. Opin. Struct. Biol. 2010;20:63–69. doi: 10.1016/j.sbi.2010.01.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 11.Kumar S., Li M.S. Biomolecules under mechanical force. Phys. Rep. 2010;486:1–74. [Google Scholar]
- 12.Brockwell D.J., Paci E., Radford S.E. Pulling geometry defines the mechanical resistance of a β-sheet protein. Nat. Struct. Biol. 2003;10:731–737. doi: 10.1038/nsb968. [DOI] [PubMed] [Google Scholar]
- 13.Carrion-Vazquez M., Li H., Fernandez J.M. The mechanical stability of ubiquitin is linkage dependent. Nat. Struct. Biol. 2003;10:738–743. doi: 10.1038/nsb965. [DOI] [PubMed] [Google Scholar]
- 14.Dietz H., Berkemeier F., Rief M. Anisotropic deformation response of single protein molecules. Proc. Natl. Acad. Sci. USA. 2006;103:12724–12728. doi: 10.1073/pnas.0602995103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 15.Nome R.A., Zhao J.M., Scherer N.F. Axis-dependent anisotropy in protein unfolding from integrated nonequilibrium single-molecule experiments, analysis, and simulation. Proc. Natl. Acad. Sci. USA. 2007;104:20799–20804. doi: 10.1073/pnas.0701281105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Puchner E.M., Alexandrovich A., Gautel M. Mechanoenzymatics of titin kinase. Proc. Natl. Acad. Sci. USA. 2008;105:13385–13390. doi: 10.1073/pnas.0805034105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Oberhauser A.F., Carrión-Vázquez M. Mechanical biochemistry of proteins one molecule at a time. J. Biol. Chem. 2008;283:6617–6621. doi: 10.1074/jbc.R700050200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 18.Monod J., Changeux J.P., Jacob F. Allosteric proteins and cellular control systems. J. Mol. Biol. 1963;6:306–329. doi: 10.1016/s0022-2836(63)80091-1. [DOI] [PubMed] [Google Scholar]
- 19.Goodey N.M., Benkovic S.J. Allosteric regulation and catalysis emerge via a common route. Nat. Chem. Biol. 2008;4:474–482. doi: 10.1038/nchembio.98. [DOI] [PubMed] [Google Scholar]
- 20.Alberts B., Bray D., Watson J. Garland Science; New York: 1994. Molecular Biology of the Cell. [Google Scholar]
- 21.Choi B., Zocchi G., Perry L.J. Artificial allosteric control of maltose binding protein. Phys. Rev. Lett. 2005;94:038103. doi: 10.1103/PhysRevLett.94.038103. [DOI] [PubMed] [Google Scholar]
- 22.Choi B., Zocchi G., Jeanne Perry L. Allosteric control through mechanical tension. Phys. Rev. Lett. 2005;95:078102. doi: 10.1103/PhysRevLett.95.078102. [DOI] [PubMed] [Google Scholar]
- 23.Zocchi G. Controlling Proteins Through Molecular Springs. Annu. Rev. Biophys. 2009;38:75–88. doi: 10.1146/annurev.biophys.050708.133637. [DOI] [PubMed] [Google Scholar]
- 24.Choi B., Zocchi G. Guanylate kinase, induced fit, and the allosteric spring probe. Biophys. J. 2007;92:1651–1658. doi: 10.1529/biophysj.106.092866. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Paci E., Karplus M. Unfolding proteins by external forces and temperature: the importance of topology and energetics. Proc. Natl. Acad. Sci. USA. 2000;97:6521–6526. doi: 10.1073/pnas.100124597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 26.Lu H., Schulten K. The key event in force-induced unfolding of Titin's immunoglobulin domains. Biophys. J. 2000;79:51–65. doi: 10.1016/S0006-3495(00)76273-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 27.Gräter F., Shen J.H., Grubmüller H. Mechanically induced titin kinase activation studied by force-probe molecular dynamics simulations. Biophys. J. 2005;88:790–804. doi: 10.1529/biophysj.104.052423. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Ortiz V., Nielsen S.O., Discher D.E. Unfolding a linker between helical repeats. J. Mol. Biol. 2005;349:638–647. doi: 10.1016/j.jmb.2005.03.086. [DOI] [PubMed] [Google Scholar]
- 29.Lee E.H., Hsin J., Schulten K. Discovery through the computational microscope. Structure. 2009;17:1295–1306. doi: 10.1016/j.str.2009.09.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Nielsen S.O., Lopez C.F., Klein M.L. Coarse grain models and the computer simulation of soft materials. J. Phys. Cond. Mat. 2004;16:R481–R512. [Google Scholar]
- 31.Tozzini V., Trylska J., McCammon J.A. Flap opening dynamics in HIV-1 protease explored with a coarse-grained model. J. Struct. Biol. 2007;157:606–615. doi: 10.1016/j.jsb.2006.08.005. [DOI] [PubMed] [Google Scholar]
- 32.Monticelli L., Kandasamy S.K., Marrink S.J. The MARTINI coarse-grained force field: extension to proteins. J. Chem. Theory Comput. 2008;4:819–834. doi: 10.1021/ct700324x. [DOI] [PubMed] [Google Scholar]
- 33.Noid W.G., Chu J.W., Andersen H.C. The multiscale coarse-graining method. I. A rigorous bridge between atomistic and coarse-grained models. J. Chem. Phys. 2008;128:244114. doi: 10.1063/1.2938860. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 34.Noid W.G., Liu P., Voth G.A. The multiscale coarse-graining method. II. Numerical implementation for coarse-grained molecular models. J. Chem. Phys. 2008;128:244115. doi: 10.1063/1.2938857. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Tozzini V. Multiscale modeling of proteins. Acc. Chem. Res. 2010;43:220–230. doi: 10.1021/ar9001476. [DOI] [PubMed] [Google Scholar]
- 36.Tirion M.M. Large amplitude elastic motions in proteins from a single-parameter, atomic analysis. Phys. Rev. Lett. 1996;77:1905–1908. doi: 10.1103/PhysRevLett.77.1905. [DOI] [PubMed] [Google Scholar]
- 37.Bahar I., Atilgan A.R., Erman B. Direct evaluation of thermal fluctuations in proteins using a single-parameter harmonic potential. Fold. Des. 1997;2:173–181. doi: 10.1016/S1359-0278(97)00024-2. [DOI] [PubMed] [Google Scholar]
- 38.Hinsen K., Thomas A., Field M.J. Analysis of domain motions in large proteins. Proteins. 1999;34:369–382. [PubMed] [Google Scholar]
- 39.Atilgan A.R., Durell S.R., Bahar I. Anisotropy of fluctuation dynamics of proteins with an elastic network model. Biophys. J. 2001;80:505–515. doi: 10.1016/S0006-3495(01)76033-X. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 40.Tama F., Wriggers W., Brooks C.L., 3rd Exploring global distortions of biological macromolecules and assemblies from low-resolution structural information and elastic network theory. J. Mol. Biol. 2002;321:297–305. doi: 10.1016/s0022-2836(02)00627-7. [DOI] [PubMed] [Google Scholar]
- 41.Sacquin-Mora S., Lavery R. Investigating the local flexibility of functional residues in hemoproteins. Biophys. J. 2006;90:2706–2717. doi: 10.1529/biophysj.105.074997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Sacquin-Mora S., Laforet E., Lavery R. Locating the active sites of enzymes using mechanical properties. Proteins. 2007;67:350–359. doi: 10.1002/prot.21353. [DOI] [PubMed] [Google Scholar]
- 43.Yang L., Song G., Jernigan R.L. Protein elastic network models and the ranges of cooperativity. Proc. Natl. Acad. Sci. USA. 2009;106:12347–12352. doi: 10.1073/pnas.0902159106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 44.Yang L., Song G., Jernigan R.L. Close correspondence between the motions from principal component analysis of multiple HIV-1 protease structures and elastic network modes. Structure. 2008;16:321–330. doi: 10.1016/j.str.2007.12.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 45.Dietz H., Rief M. Elastic bond network model for protein unfolding mechanics. Phys. Rev. Lett. 2008;100:098101. doi: 10.1103/PhysRevLett.100.098101. [DOI] [PubMed] [Google Scholar]
- 46.Eyal E., Bahar I. Toward a molecular understanding of the anisotropic response of proteins to external forces: insights from elastic network models. Biophys. J. 2008;94:3424–3435. doi: 10.1529/biophysj.107.120733. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 47.Sacquin-Mora S., Lavery R. Modeling the mechanical response of proteins to anisotropic deformation. ChemPhysChem. 2009;10:115–118. doi: 10.1002/cphc.200800480. [DOI] [PubMed] [Google Scholar]
- 48.Tseng C.Y., Wang A., Zocchi G. Mechano-chemistry of the enzyme Guanylate Kinase. Eur. Phys. Lett. 2010;91:18005. [Google Scholar]
- 49.Oeschger M.P., Bessman M.J. Purification and properties of guanylate kinase from Escherichia coli. J. Biol. Chem. 1966;241:5452–5460. [PubMed] [Google Scholar]
- 50.Blaszczyk J., Li Y., Ji X. Crystal structure of unligated guanylate kinase from yeast reveals GMP-induced conformational changes. J. Mol. Biol. 2001;307:247–257. doi: 10.1006/jmbi.2000.4427. [DOI] [PubMed] [Google Scholar]
- 51.Sekulic N., Shuvalova L., Lavie A. Structural characterization of the closed conformation of mouse guanylate kinase. J. Biol. Chem. 2002;277:30236–30243. doi: 10.1074/jbc.M204668200. [DOI] [PubMed] [Google Scholar]
- 52.Tozzini V. Coarse-grained models for proteins. Curr. Opin. Struct. Biol. 2005;15:144–150. doi: 10.1016/j.sbi.2005.02.005. [DOI] [PubMed] [Google Scholar]
- 53.Zacharias M. Protein-protein docking with a reduced protein model accounting for side-chain flexibility. Protein Sci. 2003;12:1271–1282. doi: 10.1110/ps.0239303. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 54.Tseng C.Y., Wang A., Levine A.J. Elastic energy of protein-DNA chimeras. Phys. Rev. E Stat. Nonlin. Soft Matter Phys. 2009;80:061912. doi: 10.1103/PhysRevE.80.061912. [DOI] [PubMed] [Google Scholar]
- 55.García A.E. Large-amplitude nonlinear motions in proteins. Phys. Rev. Lett. 1992;68:2696–2699. doi: 10.1103/PhysRevLett.68.2696. [DOI] [PubMed] [Google Scholar]
- 56.Amadei A., Linssen A.B.M., Berendsen H.J.C. Essential dynamics of proteins. Proteins. 1993;17:412–425. doi: 10.1002/prot.340170408. [DOI] [PubMed] [Google Scholar]
- 57.Case D.A. Normal-mode analysis of protein dynamics. Curr. Opin. Struct. Biol. 1994;4:285–290. [Google Scholar]
- 58.Amadei A., Linssen A.B.M., Berendsen H.J. An efficient method for sampling the essential subspace of proteins. J. Biomol. Struct. Dyn. 1996;13:615–625. doi: 10.1080/07391102.1996.10508874. [DOI] [PubMed] [Google Scholar]
- 59.Berendsen H.J.C., Vanderspoel D., Vandrunen R. GROMACS: a message-passing parallel molecular-dynamics implementation. Comput. Phys. Commun. 1995;91:43–56. [Google Scholar]
- 60.Lindahl E., Hess B., van der Spoel D. GROMACS 3.0: a package for molecular simulation and trajectory analysis. J. Mol. Model. 2001;7:306–317. [Google Scholar]
- 61.Van Der Spoel D., Lindahl E., Berendsen H.J. GROMACS: fast, flexible, and free. J. Comput. Chem. 2005;26:1701–1718. doi: 10.1002/jcc.20291. [DOI] [PubMed] [Google Scholar]
- 62.Miller S., Janin J., Chothia C. Interior and surface of monomeric proteins. J. Mol. Biol. 1987;196:641–656. doi: 10.1016/0022-2836(87)90038-6. [DOI] [PubMed] [Google Scholar]
- 63.Hubbard S.J. Department of Biochemistry and Molecular Biology, University College of London; London: 1992. ACCESS: A Program for Calculating Accessibilities. [Google Scholar]
- 64.Dreusicke D., Schulz G.E. The glycine-rich loop of adenylate kinase forms a giant anion hole. FEBS Lett. 1986;208:301–304. doi: 10.1016/0014-5793(86)81037-7. [DOI] [PubMed] [Google Scholar]
- 65.Leipe D.D., Koonin E.V., Aravind L. Evolution and classification of P-loop kinases and related proteins. J. Mol. Biol. 2003;333:781–815. doi: 10.1016/j.jmb.2003.08.040. [DOI] [PubMed] [Google Scholar]
- 66.Stehle T., Schulz G.E. Refined structure of the complex between guanylate kinase and its substrate GMP at 2.0 Å resolution. J. Mol. Biol. 1992;224:1127–1141. doi: 10.1016/0022-2836(92)90474-x. [DOI] [PubMed] [Google Scholar]
- 67.Hible G., Christova P., Cherfils J. Unique GMP-binding site in Mycobacterium tuberculosis guanosine monophosphate kinase. Proteins. 2006;62:489–500. doi: 10.1002/prot.20662. [DOI] [PubMed] [Google Scholar]
- 68.Bartlett G.J., Porter C.T., Thornton J.M. Analysis of catalytic residues in enzyme active sites. J. Mol. Biol. 2002;324:105–121. doi: 10.1016/s0022-2836(02)01036-7. [DOI] [PubMed] [Google Scholar]
- 69.Yuan Z., Zhao J., Wang Z.X. Flexibility analysis of enzyme active sites by crystallographic temperature factors. Protein Eng. 2003;16:109–114. doi: 10.1093/proeng/gzg014. [DOI] [PubMed] [Google Scholar]
- 70.Yang L.W., Bahar I. Coupling between catalytic site and collective dynamics: a requirement for mechanochemical activity of enzymes. Structure. 2005;13:893–904. doi: 10.1016/j.str.2005.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 71.Humphrey W., Dalke A., Schulten K. VMD: visual molecular dynamics. J. Mol. Graph. 1996;14:33–38. doi: 10.1016/0263-7855(96)00018-5. 27–38. [DOI] [PubMed] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.