Current Physical Chemistry, 2012, 2, 79-91

79

Protein Folding and Ligand-Enzyme Binding from Bias-Exchange Metadynamics Simulations Fahimeh Baftizadeh, Pilar Cossio, Fabio Pietrucci and Alessandro Laio* SISSA- Scuola Internazionale Superiore di Studi Avanzati, via Bonomea 265, I-34136 Trieste, Italy; Centre Européen de Calcul Atomique et Moléculaire (CECAM), Ecole Polytechnique Fédérale de Lausanne (EPFL), CH-1015 Lausanne, Switzerland Abstract: Bias-Exchange Metadynamics is a powerful technique that can be used for reconstructing the free energy and for enhancing the conformational search in complex biological systems. In this method, a large set of collective variables (CVs) is chosen and several metadynamics simulations are performed on different replicas of the system, each replica biasing a different CV. Exchanges between the bias potentials are periodically attempted according to a replica exchange scheme, and this process is repeated until convergence of the free energy profiles is obtained. Bias-Exchange Metadynamics has been used to understand several different biological phenomena. In particular, due to the efficaciously multidimensional nature of the bias, it is useful to study the folding process of small-to-medium size proteins, and ligandenzyme binding. This review intends to provide a comprehensive description of the algorithm and the approach used to analyze its output. We focus on the practical aspects that need to be addressed when one attempts to apply the method to study protein systems: choice of the appropriate set of parameters and CVs, proper treatment of boundary conditions, convergence criteria, and derivation of a thermodynamic and kinetic model of the system from the simulation results.

Keywords: Enhanced sampling, Free energy calculations, Protein folding, Ligand-enzyme binding, Bias-exchange metadynamics. INTRODUCTION Atomistic simulations of complex systems are becoming increasingly useful, as they have the potential to investigate processes with a high resolution in time and space. However in several important cases simulations are still not competitive if compared with more empirical methods. This is mainly due to the fact that realistic potential energy functions are intrinsically complex and their evaluation is computationally demanding. Moreover, the dynamics of realistic systems spans a wide range of characteristic times and the integration time step that is used for evolving the system has to commensurate with the fastest dynamics, which is of the order of 1 fs. As a result, the time scales that can be currently simulated are in the range of microseconds. Using purpose built machines [1, 2] or distributed computing [3] it is nowadays possible to reach time scales of milliseconds. However, these exceptional resources are available only to a few research groups. As an alternative, one can exploit a methodology aimed at accelerating rare events using the available computer time with improved efficiency. Several methods have been developed to this aim in the last years and important successes have been achieved in various fields, ranging from solid state physics to quantum chemistry [4-7]. Most of these techniques are unfortunately only partially useful for biophysical applications. The efficiency of the methods based on a search over the potential energy surface [6, 8, 9] is hindered by the enormous amount of degrees of freedom involved. In the *Address correspondence to this author at the SISSA- Scuola Internazionale Superiore di Studi Avanzati, via Bonomea 265, I-34136 Trieste, Italy; Tel: +390403787426; Fax: +390403787599; E-mail: [email protected] 1877-9468/12 $58.00+.00

simulation of a normal size protein with explicit water, one has to treat explicitly 20000 atoms, thus a 10 4 dimensional configuration space. What is commonly done is to reduce the dimensionality of the system into a few collective variables, that are assumed to provide a coarse-grained description, and to explore the free energy surface as a function of these variables. The methods based on the exploration of a single or a few reaction coordinates, like WHAM [10, 11] and thermodynamic integration [5, 12] are of limited use in protein folding, since phenomena in a biological process usually involve the concerted or sequential motion of several independent degrees of freedom. If an important degree of freedom is not considered explicitly, the simulation can show hysteresis effects and poor reproducibility of the results. Also history-dependent search methods, such as local elevation [7], Wang-Landau sampling [13] and metadynamics [14], allow a free energy reconstruction only as a function of a few variables. This is due to the fact that their performance deteriorates rapidly with dimensionality. The only methodology that seems to offer a general route for studying complex rearrangements, such as protein folding, is replica exchange molecular dynamics (REMD, also known as parallel tempering, or multicanonical ensemble method [4, 15]). Applications to large biological systems can, however, be computationally very intense. When a biomolecule is immersed in solvent, the structure of the density of states imposes use of a great amount of replicas even for small systems. This has so far limited the scope of this, otherwise extremely powerful, methodology. In a recent paper [16] another technique for studying protein folding was introduced, called bias-exchange © 2012 Bentham Science Publishers

80 Current Physical Chemistry, 2012, Vol. 2, No. 1

Baftizadeh et al.

metadynamics. This method is based on the combined use of replica exchange [15] and metadynamics [14] that allows reconstructing the free energy in a virtually unlimited number of variables. In this method, a large set of collective variables is chosen and several metadynamics simulations are performed in parallel, biasing each replica with a timedependent potential acting on just one or two of the collective variables. Exchanges between the bias potentials in the different variables are periodically allowed according to a replica exchange scheme. Due to the effective multidimensional nature of the bias, the method allows exploring complex free-energy landscapes with great efficiency. Moreover, since all the replicas are simulated at the same temperature, it is not necessary to use a large number of replicas for systems described with explicit solvent, as it is instead compulsory in standard REMD. In fact, the force field energy contribution cancels out in the acceptance, and the number of replicas that have to be used for a given problem depends only on the number of collective variables that one decides to explore. This can be large at will if enough computing resources are available. This review is dedicated to a detailed description of this approach. In Section I, the algorithm is described. Further, the manner in which the CVs have to be chosen is described and the most useful CVs used in the literature are listed. Moreover, the optimal choice of the parameters for efficiently using the algorithm is presented. In Section II, the convergence criteria, and the methods for analyzing data to obtain the thermodynamics and kinetics of the system are summarized. Finally, in Section III some recent applications of the method are presented. I. METHODS Standard metadynamics is based on a dimensional reduction, in the spirit of the work by Kevrekidis[17]. The bias acts in the space defined by a few collective variables s(x) , which are assumed to provide a coarse-grained description of the system (e.g. gyration radius of a protein or amount of helical secondary structure), and are explicit function of the Cartesian coordinates x . The dynamics is driven by the free energy F(s) and is biased by a historydependent potential constructed as a sum of Gaussians centered along the trajectory of the CVs:

VG (s(x),t) = w



exp(

t' = G ,2  G ,...

(s(x)  s(t ' ))2 ), 2 s 2

(1)

where s(t) = s(x(t)) is the value taken by the CV at time t . Three parameters enter into the definition of VG : (i) The Gaussian height w (ii) The Gaussian width  s (iii) The frequency  G1 at which the Gaussians are deposited. These parameters influence the accuracy and efficiency of the free energy reconstruction. If the Gaussians are large, the free energy surface will be explored at a fast pace, but the

reconstructed profile will be affected by large errors. Instead if the Gaussians are small or are placed infrequently the reconstruction will be accurate, but it will take a longer time. Typically the width ( s) is chosen of the order of the standard deviation of the CV in an unbiased simulation in which the system explores a local minimum in the free energy surface [18]. The bias potential, in time, fills the minima in the free energy surface, allowing the system to efficiently explore the space defined by the CVs. For long times VG (s,t)  F(s) [19]. Qualitatively, as long as the CVs are uncorrelated, the time required to reconstruct a free energy surface at a fixed accuracy scales exponentially with the number of CVs. Therefore, the performance of the algorithm rapidly deteriorates as the dimensionality of the CV space increases. This makes an accurate calculation of the free energy prohibitive, when the dimensionality of the space is large. Unfortunately, this is often the case for very complex reactions such as protein folding, in which it is very difficult, if not impossible, to select a priori a limited number of variables that describe the process, if the structure of the native state is not known in advance. To overcome these difficulties a new method called biasexchange metadynamics (BE-META) was proposed by A. Laio and S. Piana [16]. BE-META is a combination of replica exchange [15] and metadynamics [14], in which multiple metadynamics simulations of the system at the same temperature are performed. Each replica is biased with a time-dependent potential acting on a different collective variable. Exchanges between the bias potentials in the different variables are periodically allowed according to a replica exchange scheme. If the exchange move is accepted, the trajectory that was previously biased in the direction of the first variable continues its evolution based by the second and viceversa. In this manner, a large number of different variables can be biased, and a high-dimensional space can be explored after a sufficient number of exchanges. The result of the simulation is not however a free-energy hyper-surface in several dimensions, but several (less informative) lowdimensional projections of the free energy surface along each of the CVs. The high-dimensional hyper-surface can still be reconstructed using the approach in Section II.B. In more details, let us consider N R non-interacting replicas of the system, all at the same temperature T , and each biased in a different CV, s (x) with  = 1,..., N R . Each replica accumulates a history-dependent metadynamics potential as VG (x,t) = VG (s (x),t) . In BE-META, the replicas are allowed to exchange their configurations, like in conventional REMD and in the approach introduced in Ref. [19]. The exchange move consists on swapping the atomic coordinates x a and x b of two replicas a and b (evolved under the action of two different history-dependent potentials), selected at random among the N R available. The move is accepted with a probability:

Pab = min(1, e

(

b ( xb ,t )V a ( xb ,t )V b ( x a ,t )  VGa ( x a ,t )+VG G G

)).

(2)

Protein Folding and Ligand-Enzyme Binding from Bias-Exchange

The normal potential energy of the system cancels out exactly for this kind of move. If the move is accepted, the CVs of replica a perform a jump from s a (x a ) to s a (x b ), and for replica b from s b (x b ) to s b (x a ). The exchange moves introduce a jump process on top of the ordinary metadynamics evolution. This greatly increases the diffusivity of each replica in the whole CV space. The working principle of the approach is better illustrated in a simple example. Consider a dynamics on a two-dimensional potential like the one in Fig. (1-A). If one performs simple metadynamics biasing x, one obtains an estimate of the free energy affected by large errors: indeed, the system jumps between the two wells at the bottom and the two wells at the top only due to (rare) thermal fluctuations, and obtaining the correct free energy would require to take the average over several transitions along y. In practice, the free energy profile will not converge (see Fig. 1-B).

Current Physical Chemistry, 2012, Vol. 2, No. 1

81

to Eq. 2. Even if the computational cost has only doubled with respect to the simulation above, one observes a very significant reduction of hysteresis. Now the metadynamics potential almost exactly compensates the free energy, both as a function of x and y : indeed, the profiles are practically flat lines. Hysteresis is much reduced, and the metadynamics potential provides a very good estimate of the free energy. This example shows that, like in ordinary metadynamics, in BE-META the Gaussian potential converges to the negative of the free energy (see also Section II.A and III.B). Both metadynamics and BE-META are implemented in the freelyavailable plug-in PLUMED, which allows to use a number of molecular dynamics engines including gromacs, amber, namd, lammps, etc [20]. A)

y

A)

B)

x

Free energy

B)

x Fig. (1). A) An example of a two dimensional free energy potential. B) An ordinary metadynamics reconstruction performed biasing the coordinate x in a system with the potential depicted in panel A. The correct free energy as a function of x is represented as a thick black line. The thin lines represent F ( x ) + VG ( x,t ) at different times. Clearly, the sum of the two quantity does not converge to a constant. The large jumps between the two different regimes are due to transitions of the system along y . These transitions are triggered by thermal fluctuations. Thus are rare on the time scale of the simulation, making equilibration along y extremely slow and determining the aforementioned convergence problems.

In Fig. (2) we show the result of a simulation consisting of two metadynamics on two replicas, one biasing x, the other y. From time to time, we allow the two replicas to exchange configurations, accepting the exchange according

Fig. (2). An example of a bias exchange metadynamics reconstruction performed in a system with the potential depicted in panel A of Fig. (1). The simulation is performed with two replicas, biasing with metadynamics x and y respectively, with metadynamics parameters identical to those used for Fig. (1). F ( x ) + VG ( x,t ) and F ( y ) + VG ( y,t ) , represented respectively in panel A and B, are now approximately flat at all times, indicating that the CV jumps introduced by the exchange moves have greatly improved the convergence.

A. Choice of the Collective Variables Similar to the other methods that reconstruct the free energy as a function of a set of generalized coordinates, in BE-META the choice of the CVs, plays an essential role in determining the convergence and efficiency of the freeenergy calculation. If the chosen set of CVs does not distinguish different metastable states of the system, the simulation will be affected by hysteresis as not all of the important regions of the conformational space will be

82 Current Physical Chemistry, 2012, Vol. 2, No. 1

Baftizadeh et al.

explored. Unfortunately, there is no a priori recipe for finding the correct set of CVs, and in many cases it is necessary to proceed by trial and error. However, in BEMETA the variables can be as many as necessary, making the selection less critical. To choose the proper set, one needs to exploit some basic knowledge on the topological, chemical, and physical properties of the system. In the case of proteins, that are chains of amino acids with well-defined topological features, commonly used CVs (available in the PLUMED plug-in [20]) are: •

Coordination number: this is probably the most used CV. It is defined as:

(

{ })

S = n  RMSD { Ri }i , R 0  , 

 

n ( RMSD ) =

1 ( RMSD / 0.1)

1 ( RMSD / 0.1)

(6)

8

12

,

(7)

where n is a function switching smoothly between 0 and 1, the RMSD is measured in nm, and { Ri }i are the atomic 

coordinates of a set  of six residues of the protein, while R 0 are the corresponding atomic positions of the ideal

{ }

beta conformation. By changing the template

{R } 0

the CV

CN =  Cij ,

can be employed to count the content of alpha secondary structure, instead of beta, or of other structural patterns.

with

The first three types of CVs were successfully used to fold with BE-META some small proteins such as Advillin [22], Tryptophan Cage [23], Insulin [24]. But in the case of larger and more complex proteins the alpha/betarmsd CVs may prove more efficient, as it has been shown by simulations on GB1 and SH3, which are about 60 amino acids with a large content of beta secondary structure [21].

i, j

rij 1  ( )n r0 Cij = , rij m 1 ( ) r0

(3)

(or an analogous smooth function of rij ) where rij is the distance between atoms or groups i and j , and m , n are exponents that allow to tune the smoothness of the function. This CV can be used to count the number of contacts, e.g. chemical bonds, hydrogen bonds, or hydrophobic contacts between selected sets of atoms or groups. •

Dihedral distance: it measures the number of dihedrals  i (involving the backbone atoms C-N-C  -C) that are similar to a reference  0 :

1 N D =  [1+ cos( i   0 )]. i 2

(4)

It can be efficiently used to count how many residues belong to an  -helix. In principle, it can also count the number of residues belonging to a  -sheet. However, in this case, controlling the dihedral is normally not sufficient to drive the system towards the correct configuration, a betabridge, that requires the formation of the correct hydrogen bonds between residues that are at a large sequence separation. •

Dihedral correlation: it measures the similarity between adjacent dihedral angles of the protein backbone:

1 DC =  [1+ cos( i   i1 )]. i 2

(5)

Since secondary structure elements  -helices or  sheets have a correlation between successive dihedrals, also this CV can be used to detect the presence of secondary structure elements. •

Betarmsd: it counts how many fragments of 3+3 residues in the protein chain belong to a beta-sheet secondary structure, by computing their RMSD with respect to an ideal beta conformation [21]:

B. Choice of the BE-META Parameters In this Section we discuss how to choose the simulation parameters that are specific to a BE-META calculation, like the frequency of exchanges of the bias potentials, temperature of the replicas, etc. Part of the analysis presented here is based on Ref. [25], where extensive benchmarks were performed using a coarse-grained force field, UNRES, comparing BE-META with standard REMD. The authors considered 1E0G, a 48-residue-long alpha-beta protein that folds with UNRES into the experimental native state. 1. Frequency of Exchange Moves While in REMD it is known that it is marginally better to exchange as often as possible, as exchanges always enhance decorrelation, extensive folding simulations of 1E0G indicate that very frequent or very rare exchanges make BEMETA marginally less efficient (see Fig. 3). The optimal exchange time for this system is  exch = 120 ps, which gives the system enough time to explore the local free-energy wells before a change in the direction of bias. In all cases considered BE-META leads to a significant improvement in the speed of exploration of the configuration space compared to REMD. 2. Effect of Temperature In standard REMD the replicas span a wide temperature range and their number has to be increased if the system is large [15]. Instead, in BE-META all the replicas can have the same T , because the enhanced sampling does not rely on temperature jumps but on the effect of the bias. This is an advantage if the protein is modeled with an explicit solvent Hamiltonian, since the large increase in the number of degrees of freedom due to the addition of solvent molecules does not require an increase in the number of replicas. As shown for 1E0G in Fig. (4), with increasing temperature the

Protein Folding and Ligand-Enzyme Binding from Bias-Exchange

system is able to find faster the folded state of the protein, with best performance between T = 315 and 350 K. Comparing this with the results reported by Ref. [26] for the specific heat of the system, the temperature which optimizes the performance of BE-META is close to the peak of the specific heat. This is an expected result due to the temperature dependence of the Boltzmann probability of  E/k T

 : sampling is enhanced by conformational transitions e a high T , but not higher than the critical temperature, because in that case the system will mostly explore unfolded structures reducing the efficiency in localizing the folded state.

TOTAL AVERAGE FOLDING TIME (ns)

111111111111 000000000000 000000000000 111111111111 000000000000 111111111111

1000

500

250

0

100

50

000 111 0000000000000 1111111111111 000 111 0000000000000 1111111111111 000 111 0000000000000 1111111111111 000 111 0000000000000 1111111111111 000000000000000 111111111111111 000 111 0000000000000 1111111111111

000 111 0000000000000 No exchange 1111111111111 150 000 111 0000000000000 1111111111111 0000000000000 1111111111111 EXCHANGE TIME (ps)

Fig. (3). BE-META folding simulations on 1E0G at T = 280 K with different  exch . The Gaussian height is held fixed w = 0.012 kcal/mol, and a set of 8 CVs has been used (see Ref. [25] for details). The average folding times are shown.

TOTAL AVERAGE FOLDING TIMES (ns)

500

400

300 T=400K 200

T=280K

100

0

83

more time to fill with Gaussians a high dimensional space than a low-dimensional one. On the other hand, if the freeenergy landscape of the system is intrinsically high dimensional and one uses only a one-dimensional bias, the metadynamics estimate of the free energy is affected by large errors. BE-META provides, at least in principal, a solution to this problem, as one can explore high dimensional free energy landscape by using several low-dimensional biases. Benchmarks indicate that using one-dimensional Gaussians increases the efficiency of the method by a factor of ~ 2 compared with two-dimensional Gaussians [25]. The difference in performance might be due to the fact that the gain in the speed by which wells are filled with onedimensional hills overwhelms the larger flexibility in finding complicated transition pathways with two-dimensional hills. The general validity of this assertion may however require to be checked on different systems. C. Proper Treatment of the Boundaries

750

0

Current Physical Chemistry, 2012, Vol. 2, No. 1

T=315K

300

T=350K

350 TEMPERATURES (K)

400

Fig. (4). Effect of temperature on BE-META folding simulations of 1E0G.  exch = 20 ps, w = 0.012 kcal/mol, and a set of 4 CVs has been used (see Ref. [25] for details). The average folding times are shown.

3. Dimensionality of the Bias Potentials The efficiency of standard metadynamics degrades with increasing dimensionality of the Gaussian hills, as it takes

As it has been already discussed, when a metadynamics simulation reaches convergence, the history-dependent potential fluctuates around the negative of the free energy of the system, namely VG (s(x),t)~  F(s) . The lack of stationary fluctuations may indicate serious convergence problems (see Section II.A). Depending on the boundary of the CV space, however, there may be a technical problem which has to be solved to allow convergence. Metadynamics simulations typically take advantage of a finite and not too small width of Gaussian hills to "fill" the free-energy surface quickly. On the other hand, finite width Gaussians can induce systematics errors at the boundaries  of CV space [18] (whether natural or artificially imposed ones): these errors are due to the fact that a sum of finite-width Gaussianshaped hills can not accurately reproduce discontinuities on the free-energy profile. The presence of the discontinuities is common for several types of CVs that are intrinsically limited e.g. to be non-negative, like coordination numbers, alpha/betarmsd, etc. At the beginning of the simulation these errors are small and are usually overlooked, but at long times they can become important, preventing the system from reaching stationary condition. In Ref. [19] it was shown that if the component of the free energy gradient in the direction normal to  vanishes at the boundaries (  n F() = 0 ), the systematic errors can be eliminated by choosing a functional form for the history-dependent potential that satisfies the same condition ( nVG () = 0) . A generalization of this procedure, valid also if  n F()  0 , is introduced in Ref. [27]. We here introduce a remedy to the same problem that works only for one-dimensional biases, but that has the great advantage of being very robust and parameter-free. The idea is to change the metadynamics algorithm setting the bias force equal to zero beyond the boundary. If, for example, one performs metadynamics on a CV s and one is interested only to the free energy for s > sw , one still updates the history dependent potential according to Eq. 1, but then sets

84 Current Physical Chemistry, 2012, Vol. 2, No. 1

dVG(s,t)/ds = 0, s < sw

Baftizadeh et al.

(8)

a)

Notice that updating VG implies that Gaussians are added

-10

also if s < sw , as the tails of these Gaussians influence VG in the region where it is relevant, s > sw . In this way, the force

s > sw

-20

comes from both

metadynamics and the force field, in the region s < sw only from the latter. This approach allows to reconstruct correctly the free energy. In Fig. (5), we show VG ( s,t ) at different times for a dialanine molecule in vacuum at 300 K, simulated performing metadynamics with s =  (the backbone dihedral angle), restraining s in the interval [ 1, 2 ] with two harmonic potentials, centered at the limits of the interval. Panel a) shows that normal metadynamics is affected by large systematic errors close to the boundaries, and that the errors become larger and larger. In panel b) we show VG ( s,t ) in exactly the same conditions, except that we imposed dVG(s,t)/ds = 0 for s < 1 and s > 2 . Clearly, the history-dependent potential reaches a stationary state. In panel c) we compare the time average of VG ( s,t ) with the fully converged result, obtained by performing metadynamics on the same system without the restraining potentials. One can see that the two estimates differ only on a tiny region close to the boundaries. One can show that dF systematic errors disappear also if = 0 , namely if the ds s w

free energy is flat at the boundary: by imposing this condition Eq. 8 we are forcing the history dependent potential to converge to a state in which its derivative at the boundary is zero, and not equal to dF / ds like in normal metadynamics. However, the region where systematic deviations are observed is really small, of the order of the Gaussian width  s . Thus, at a distance from sw larger than

~2 s , one can very safely use the average value of VG to estimate the free energy. More importantly, the systematic error close to the boundary does not grow with time, but remains stationary, allowing for a meaningful time average of the history dependent potential. II. ANALYSIS A. Convergence Criteria For standard metadynamics, in Ref. [19] a formalism was introduced which allows to map the history-dependent evolution into a Markovian process in the original variable and in an auxiliary field that keeps track of the configurations visited. With this formalism, it has been shown that the average value of VG ( s,t ) over several

independent metadynamics runs is exactly equal to F ( s ) , and an explicit expression for the error has been derived for Langevin dynamics. It was found that VG ( s,t ) estimates

F ( s ) with an accuracy that is determined by the Gaussian width  s , the Gaussian height w and the Gaussians

-30

b)

-VG(kcal/mol)

on the system in region

0

0

-10

-20

-30

c)

0

-2

-4

-6

-1

-0.5

0

0.5

1

1.5

2

Dihedral Angle (Rad) Fig. (5). Time evolution of VG for a metadynamics performed on a

dialanine, using the dihedral 

as a collective variable. The

metadynamics parameters are w = 0.1 ,  s = 0.2 ,  G = 2 ps. The blue line is the exact free energy of the system. a) Metadynamics is performed using parabolic wall with KAPPA = 100000 at  = 1 and  = 2 rad. Significant errors in the reconstruction appear close to the boundaries. b) Metadynamics is performed by setting the force equal to zero beyond  = 1 and  = 2 rad, according to the approach described in Section I.C. The errors close to the boundaries are much reduced and do not become larger with simulation time. c) The average value of VG in panel b is compared with the exact result.

deposition time  G . In Ref. [28] an alternative expression for the error has been deduced by performing extensive numerical simulations of a stochastic differential langevin equation. It was shown that the error on the reconstructed profile is approximately determined by the ratio w /  G . In Ref. [27] metadynamics is numerically shown to converge to the correct free-energy even in absence of adiabatic separation between the slow CVs and the fast degrees of freedom. Moreover the error on the reconstructed profile is shown to decay with time as for an optimal umbrella sampling simulation, namely one in which the dynamics is biased with an external potential exactly equal to F ( s ) . In BE-META the convergence of the bias potential VG ( s,t ) is monitored like in standard metadynamics [18]: after a transient time t eq , VG ( s,t ) reaches a stationary state

Protein Folding and Ligand-Enzyme Binding from Bias-Exchange

Current Physical Chemistry, 2012, Vol. 2, No. 1

in which it grows evenly fluctuating around an average. The free energy as a function of s is then estimated by the time average of VG ( s,t )

F ( s ) ~VG ( s ) =

1 t sim  t eq



t sim

t eq

dt VG ( s, t  ) ,

(9)

where t sim is the total simulation time. Convergence is evaluated independently over the profile reconstructed by each replica. The convergence of the method has been illustrated e.g. on ALA 3 by direct comparison of the bias potentials with the logarithm of the histograms obtained from very long equilibrium simulations (Ref. [23] and Section III.B). A lack of stationary fluctuations in the bias profile, typically due to the neglect of some important slow degree of freedom, is a powerful and convenient feature that allows to detect hysteresis and to improve the simulation setup. In Fig. (1-B) we show a prototypical example of how the history-dependent potential evolves with time if an important degree of freedom is not explicitly biased (the y coordinate in the case of the example): VG does not converge to a stationary shape, and taking a time average of the potential of the form in Eq. 9 is not meaningful. In this case, one should analyze the trajectory and find the “hidden” variable responsible for the large fluctuations of VG , and add it to the set of CVs. B. Cluster Analysis and Thermodynamic Model BE-META allows computing simultaneously the one- or two-dimensional free energy as a function of a large number of CVs (ten or more in practical applications). However, low-dimensional free-energy projections of a highdimensional hyper-surface are often not very insightful, as in complicated processes, like protein conformational transitions, each minimum in a low dimensional profile may correspond to several different structures. In order to estimate the relative probability of the different states one should find a manner to reconstruct the free energy in a higher dimensional space. The idea is to exploit the lowdimensional free energies obtained from BE-META to estimate, by a weighted-histogram procedure, the free energy of a finite number of structures that are representative of all the configurations explored by the system. Using the approach of Ref. [23], the CV space is subdivided so that all the frames of the BE-META trajectories are grouped in sets (clusters) whose members are close to each other in CV space. Since the scope of the overall procedure is constructing a model that describes also the kinetic properties of the system, it is important that the clusters are defined in such a way that they satisfy three properties: i) The clusters must cover densely all the configuration space explored in BE-META, including the barrier regions. ii) The distance in CV space between nearest neighbor cluster centers must not be too large. iii) The population of each cluster in the BE-META trajectory has to be significant, otherwise its free energy estimate will be unreliable

85

(we recall that due to the bias also barrier regions are significantly populated in BE-META simulations). A set of clusters that satisfy these properties is defined dividing the CV space in small hypercubes forming a regular grid. The size of the hypercube is defined by its side in each direction: ds = (ds1 , ds2 ,..., dsn ) where n is the number of collective variables. This determines directly how far the cluster centers are in CV space. Each frame of the BE-META trajectory is assigned to the hypercube to which it belongs and the set of frames contained in a hypercube defines a cluster. The free energy F of each cluster  is estimated by a weighted-histogram analysis approach (WHAM) [29], following the procedure described in Ref. [23]. In WHAM, the effect of the bias is removed from the populations allowing calculating the free energy of a finite number of clusters that are representative of all the configurations explored by the system. The free energy of cluster  is given by

F = T log

n  je

i i  1 j f Vj T

(

,

)

(10)

where ni is the number of times cluster  is observed in the trajectory of walker i and Vi is the bias potential acting on cluster  in walker i . Vi is estimated as the time average of the history-dependent potential acting on walker i , evaluated in s , the center of cluster  :

Vi = VGi ( s ) =

1 t sim  t eq



t sim

teq

dt VGi ( s , t  ) ,

(see Eq. 9). The normalization constants f j appearing in Eq. 10 are determined self-consistently like in standard WHAM[23]. Corrections taking into account the variation of the bias over different structures assigned to the same cluster  are also reported in Ref. [23]. An important issue in this procedure is how many and which CVs to use for the clusterization. Of course, it is not necessary to use all the CVs that have been explicitly biased in one replica, as some of these CVs might prove to be a posteriori less relevant for the process, or strongly correlated with other variables. The variables used for cluster analysis must provide an accurate but handy description of the system. An accurate description entails a set of clusters where each member contains consistently similar structures (thus with very similar free energy compared with kBT ). If the variables are too few, a cluster will contain structures that are very different from each other. On the other hand, performing the analysis in a very high-dimensional CV space will lead to poor statistics. A graphical user interface for VMD is available which helps performing this choice by easily visualizing the structures assigned to each cluster for different CVs[30].

86 Current Physical Chemistry, 2012, Vol. 2, No. 1

Baftizadeh et al.

C. Kinetic Model In this section we describe a manner for constructing a kinetic model describing transitions between the clusters introduced in the previous section. Constructing the model requires estimating the rates k for a transition between every pair of neighboring clusters  and  . As BE-META trajectories are biased, the transition probabilities observed in the simulation cannot be taken as a direct measure of the true transition rates. The kinetic model is constructed assuming that the transitions between clusters are described by rates of the form introduced in Ref. [31,32], namely by diffusion on a non-flat free-energy landscape: 0 k = k e



(

)

1 F F /k T 2   B

,

(11)

0 0 = k are the rates associated to simple diffusion where k

on a flat free-energy surface. This form of the transition rates ensures that the limiting probability distribution of the dynamics is correct, namely that the probability to observe cluster  at long times is proportional to exp ( F / kBT ) . If the clusters form a hyper-cubic grid in CV space the rates 0 k can be exactly expressed as a function of the (possibly position-dependent) diffusion matrix hypercube side ds [31,32].

D

and of the

The diffusion matrix can be estimated using the approach of Ref. [23,32], which maximizes the likelihood that a given equilibrium MD trajectory is generated by a rate equation of the form Eq. 11. Computing D requires first generating at least one MD trajectory without any metadynamics bias. This trajectory is then mapped at a time lag t onto the clusters ( ( 0 ) ,  ( t ) ,  ( 2t ) ,) . Then several kinetic Monte Carlo [33] trajectories are run with an initial guess for D , and the conditional transition probabilities between clusters are computed as in Ref. [23]. Using these probabilities one evaluates the logarithm of the likelihood L ( D ) to observe the sequence of clusters obtained by MD, log (L ( D )) is then maximized as a function of D . This can be done by simulated annealing, starting from an initial guess of D and iterating until the likelihood reaches a plateau. As outlined in Ref. [34], the diffusion matrix found in this way depends in general by the chosen time lag. A common behavior is that by increasing the time lag t the elements of D converge to a well defined value. This means that after this t the dynamics between clusters is close to Markovian and is well approximated by the model proposed. As a consequence the model describes correctly only transitions that occur on a time scale larger than t . The metastable states (kinetic basins) of the network are found by diagonalizing the rate matrix k . The relaxation times  i of the system are given by the inverse of the eigenvalues i . One eigenvalue is by construction equal to zero, as the rate matrix satisfies detailed balance. If the spectrum has a gap below the m th eigenvalue (i.e. im  0

and i
Protein Folding and Ligand-Enzyme Binding from Bias-Exchange

validated by NMR and CD experiments. This study has given us the confidence that BE simulations have the capability, at least in small proteins, to predict the native fold starting from an extended conformation and with an explicit solvent description. This requires a relatively small amount of computer time in comparison with other methods that are currently available for folding proteins in explicit solvent. Moreover, the method is capable of predicting differences in folds resulting from single amino acid substitutions by capturing the subtle balance between entropic and enthalpic contributions.

Current Physical Chemistry, 2012, Vol. 2, No. 1

87

a)

B. Ala3: Benchmark System for the Thermodynamic and Kinetic Model In Ref. [23], the approach described in Section II.B has been carefully benchmarked on solvated Ace-Ala 3 -Nme. For this small system, it was possible to compare the predictions of the kinetic model with the results of a long ( ~2 μ s ) unbiased molecular dynamics trajectory. The BEMETA simulations were done using the six dihedral angles (  i , i , where i = 1, 2, 3 ) as CVs, one per replica. After ~ 5 ns, convergence in the free energy profiles was obtained. It was shown that the profiles extracted from three independent BE-META runs do not show sizable differences between each other, and agree with the MD results within the error bars (see Fig. 6-a). The estimate of the relative free energies between the different structures, was done by dividing the six dimensional space in small hypercubes ('clusters') of size 30 deg and using the procedure described in Section II.B. The free energies were used for constructing the kinetic model according to eq. 11, and D was calculated by maximum likelihood with the MD trajectory using an optimal time lag t = 16 ps . The mean first passage times (MFPT) have been calculated both from MD and KMC, for the transitions between the 8 clusters corresponding to the 8 free energy minima obtained assigning the three  dihedral angles in the  or  regions. The main result is that the overall correlation between MD and KMC is excellent both for free energy and for kinetic rates (see Fig. 6-b). Barrier estimates from equilibrium MD are affected by a large error bar due to a small Boltzmann population, whereas BEMETA is as accurate in the minima as on the barrier tops. C. Kinetics of TRP-cage Folding The Trp-cage is a designed small protein (20 amino acids) showing a stable compact native structure. BE-META folding simulations on this system were performed in Ref. [16] by biasing the following five CVs: number of C  contacts (CV 1 ), number of C  contacts (CV 2 ), number of

 dihedrals belonging to the  region in the Ramachandran plot (CV 4 ), and correlation between successive  dihedrals (CV 5 ). Eight replicas of the system were used: one for each variable plus two walkers reconstructing a free energy surface in two dimensions: (CV 3 -CV 4 ) and (CV 4 -CV 5 ). The last walker, backbone h-bonds (CV 3 ), fraction of

Fig. (6). Ala3 Benchmark system. a) Correlation between the cluster free energies calculated from the BE-META simulations (Eq. 10) with the correct values, obtained by a benchmark unbiased MD simulation. b) The mean first passage times between a few selected structures, obtained by MD simulations and by the kinetic model constructed from BE-META results as described in Section II.C.

the "neutral walker", was not biased by any metadynamics potential, but was allowed to exchange conformations with the others. The total length of the simulations was 50 ns, and convergence was obtained after 22 ns. The free energies were estimated by partitioning the five-dimensional CV space in small hypercubes of size dsi = 2 i , where  i is the Gaussian width of CV i , and applying the approach described in Section II.B. An excellent correlation was found between the latter free energies and those obtained by the unbiased neutral walker (which has Boltzmann population). The diffusion matrix was estimated using the maximum likelihood approach described in Section II.C on five MD trajectories for a total time of ~500 ns, with an optimal lag time t = 12 ns. Five significantly populated ( P > 1 %) basins were found (see Fig. 7): basin 1, with P  58 %, resembles very closely the NMR structure, while basins 2 and 3 retain at least part of the native  -helix. The other basins show only a small residual secondary content and can

88 Current Physical Chemistry, 2012, Vol. 2, No. 1

Baftizadeh et al.

Fig. (7). Metastable kinetic clusters of Trp-cage and the schematic representation of its folding dynamics. Times (inverse of rates) for the transitions between the relevant clusters are shown on the arrows. The uncertainty on each transition time due to both the error on the free energies and the position-dependence of D is at most 40%. Continuous arrows correspond to direct transitions between clusters that occur on a time smaller than 1 μ s . Dashed arrows correspond instead to transition that occur on a time larger than 1 μ s or taking place through other intermediate low-populated clusters, not represented in the Figure.

be generically referred to as "unfolded states". The rates for the transitions between the basins were extracted from a very long KMC simulation ( t tot = 1.5 s), with the slower timescale of the order of μ s in fair agreement with experimental data.

biased trajectory explores the ~300 experimental folds of similar length found in the PDB. And more importantly this study shows that the 300 known folds form a rather small subset of the total number of structures that is explored.

D. Exploring the Conformational Space of Val60

The binding mechanism of a peptide substrate (Thr-IleMet-Met-Gln-Arg, cleavage site p2-NC of the viral polyprotein) to wild-type HIV-1 protease was investigated in Ref. [36] by 1.6 μ s of all-atom BE-META simulations in explicit water. The configuration space was explored biasing seven reaction coordinates: the number of hydrophobic contacts or hydrogen bonds between ligand and cavity/ligand and flap tips (Eq. 3), the distance of the ligand from the catalytic dyad or from the enzyme's core, the distance between the flat tips, and the number of water molecules at the interface between the cavity and the ligand. Exchanges of the replicas were attempted every 2 ps. Within 35 ns, several simulations started with the substrate outside the enzyme ended up in the experimental Michaelis complex (within a backbone rmsd of 0.9 Å). After constructing a kinetic model as described in Section II.C, several states of the system were identified with the complex resulting the most stable one: the calculated free energy of binding was found to be -6 kcal/mol (to be compared with -8 kcal/mol for the similar inhibitor MVT-101), and the kinetic constants for association and dissociation were 1.310 6 M 1 s 1 and 57 s 1 , respectively, consistent with available experiments. Remarkably, in the main binding pathway the flaps of the protease do not open significantly: the substrate slides inside the enzyme cavity from the tight lateral channel. This may contrast with the natural polyprotein substrate which is

In Ref. [25], by using BE-META an exhaustive exploration of the conformational space of a 60-residue polyvaline chain is performed. This system was simulated in vacuum at 400K with an all-atom force field. Six CVs of the functional form described in Eq. 6 are used: 3 CVs counting the fraction of alpha helix in each third of the protein, one variable counting the fraction of residues forming parallel beta bridges, and two counting the fraction of antiparallel bridges. The Gaussians entering in the metadynamics potential are added every 10ps. Their height and width are 5 kj/mol and 0.3, respectively. Exchanges between the biasing potentials are allowed every 25ps . The total simulation time was ~50 μ s . BE-META allows exploring very efficiently the configuration space of the system. In this specific case, the simulation was not performed with the scope of studying the thermodynamics of the system but, rather, with the scope of finding all the possible structures that a homopeptide can take. A database of around 7,000 independent folds with at least 30% of secondary structure (see Fig. 8) corresponding to local minima of the potential energy was generated. This ensemble plausibly represents the universe of protein folds of similar length: indeed, all the known folds are represented in the set with good accuracy. This study offered a striking demonstration of the power of BE-META as an enhanced sampling method: in a relatively short simulation time, the

E. Substrate Binding Mechanism of HIV-1 Protease

Protein Folding and Ligand-Enzyme Binding from Bias-Exchange

Current Physical Chemistry, 2012, Vol. 2, No. 1

89

Fig. (8). Gallery of representative VAL60 structures generated by BE. a) A selection of 260 out of the 30,000 structures generated by BEMETA. b) Examples of successful alignments. The CATH structure is represented together with its VAL60 equivalent for three cases.

Fig. (9). Representative structures of the free-energy clusters involved in the binding/unbinding mechanism of peptide substrate with HIV-1 protease. (shaded box), plus the transition state (TS). The clusters B6-B9, which are not involved in the main binding pathway, are also displayed. The free energy of each state, in kcal/mol, is reported below the corresponding structure. Arrows are labeled with the corresponding transition rates ( ms 1 ) when larger than 100 ms 1 (solid line). Transitions with a smaller associated rate are depicted as dashed arrows.

90 Current Physical Chemistry, 2012, Vol. 2, No. 1

expected, due to its very large size, to bind by opening the flaps. Thus, mutations might influence differently the binding kinetics of peptidomimetic ligands and of the natural substrate. Other successful studies done with BE-META for protein-ligand binding [41], investigating trans/cis prolyl isomerization [42] and to characterize the conformational space of proteins [43] are also available. CONCLUSIONS In this review we described bias exchange metadynamics [16], a powerful enhanced sampling technique based on the combined use of replica exchange [15] and metadynamics [14]. In this approach a large set of collective variables is chosen and metadynamics simulations are performed by biasing in parallel different CVs on different replicas. Exchanges between the bias potentials are periodically attempted according to a replica exchange scheme. The approach allows a parallel reconstruction of the free energy as a function of many variables. The computational cost scales only linearly with the number of variables that are used, making the selection of the CVs less critical. As a general rule, the more variables one adds, the better the convergence is, as an extra and smartly chosen variable has always the chance of corresponding to an essential degree of freedom of the system. However, the method does not provide a solution to the problem of choosing an appropriate set of collective variables, which is ubiquitous in free energy methods: if one forgets a relevant variable, the method does not converge. Lack of convergence can be monitored as described in Section II.A.

Baftizadeh et al.

structure, with a significant secondary structure content (see Section III.D). However, in this case the approach is used with the only scope of enhancing the exploration, and no attempt is done of ``predicting'' the folded state, namely of computing the free energy of the states that are explored. Constructing kinetic models and computing free energies is much more computationally demanding than exploring the configuration space. In the case of HIV protease (Section III.E) deriving the thermodynamic model and computing K on and K off required almost 2 μ s of simulation. In the case of the TRP cage (Section III.C) and of Advillin (Section III.A), obtaining a reliable prediction of the folding kinetics required more than 1 μ s. Clearly, applying this technique to larger proteins will require very significant resources. Tentative simulations performed on SH3 and GB1 show that obtaining truly converged free energies for these systems will likely require tens of microseconds of simulations or developing novel collective variables specifically targeted for triggering tertiary structure rearrangements. REFERENCES [1] [2]

[3]

[4] [5]

The linear scaling of the computational cost in the number of variables comes at a price: if one uses NR variable, the result of the simulation is not a free energy as a function of NR variables, but NR one-dimensional projections of the free energy. Obtaining the free energy as a function of all the variables requires a post-processing of the results, following the approach described in Section II.B. This technique can be applied to obtain free energies of structures (clusters) embedded in a high dimensional CV space, and it has been successfully applied in several contexts, some of which are described in Section III.

[6]

Despite BE-META's great success in folding small helical proteins, exploring exhaustively protein conformational landscapes, and finding substrate binding mechanisms, still some issues have been encountered that require further research. First of all, it is not clear if it is possible using this technique for folding an average size protein, with a significant content of beta structure. A substantial step forward in this direction has been obtained using the so-called betarmsd CVs (see Section I.A), that allow to form extended secondary structure elements in a short simulation time. Using these variables it was possible to explore thousands of independent metastable states of a VAL60 protein, each state corresponding to a native-like

[11]

[7] [8]

[9] [10]

[12] [13] [14] [15] [16] [17]

D. Shaw, M. Deneroff, R. Dror, et al. Anton, a special-purpose machine for molecular dynamics simulation. Commun. ACM, 2008, 51, 91-97. F. Allen, G. Almasi, W. Andreoni, et al. Blue Gene: a vision for protein science using a petaflop supercomputer. IBM Syst. J., 2001, 40, 310-327. Larson, S.; Snow, C.; Shirts, M.; Pande, V. Folding@ Home and Genome@ Home: Using distributed computing to tackle previously intractable problems in computational biology. Arxiv preprint arXiv:0901.0866 2009. Berg, B.; Neuhaus, T. Multicanonical ensemble: A new approach to simulate first-order phase transitions. Phys. Rev. Lett., 1992, 68, 9. Carter, E.; Ciccotti, G.; Hynes, J.; Kapral, R. Constrained reaction coordinate dynamics for the simulation of rare events. Chem. Phys. Lett., 1989, 156, 472-477. Henkelman, G.; Jónsson, H. A dimer method for finding saddle points on high dimensional potential surfaces using only first derivatives. J. Chem. Phys., 1999, 111, 7010. Huber, T.; Torda, A.; Gunsteren, W. Local elevation: a method for improving the searching properties of molecular dynamics simulation. J. Comput-Aided Mol. Des., 1994, 8, 695-708. Henkelman, G.; Uberuaga, B.; Jónsson, H. A climbing image nudged elastic band method for finding saddle points and minimum energy paths. J. Chem. Phys., 2000, 113, 9901. Voter, A. Hyperdynamics: Accelerated molecular dynamics of infrequent events. Phys. Rev. Lett., 1997, 78, 3908-3911. Kumar, S.; Rosenberg, J.; Bouzida, D.; Swendsen, R.; Kollman, P. Multidimensional free-energy calculations using the weighted histogram analysis method. J. Comput. Chem., 1995, 16, 13391350. Roux, B. The calculation of the potential of mean force using computer simulations. Comput. Phys. Commun., 1995, 91, 275-282. Sprik, M.; Ciccotti, G. Free energy from constrained molecular dynamics. J. Chem. Phys., 1998, 109, 7737. Wang, F.; Landau, D. Efficient, multiple-range random walk algorithm to calculate the density of states. Phys. Rev. Lett., 2001, 86, 2050-2053. Laio, A.; Parrinello, M. Escaping free-energy minima. Proc. Natl. Acad. Sci. USA, 2002, 99, 12562-12566. Sugita, Y.; Okamoto, Y. Replica-exchange molecular dynamics method for protein folding. Chem. Phys. Lett., 1999, 314, 141-151. Piana, S.; Laio, A. A bias-exchange approach to protein folding. J. Phys. Chem. B., 2007, 111, 4553-4559. Kevrekidis, I.; Gear, C.; Hummer, G. Equation-free: The computeraided analysis of complex multiscale systems. AIChE J., 2004, 50, 1346-1355.

Protein Folding and Ligand-Enzyme Binding from Bias-Exchange [18]

[19] [20]

[21]

[22]

[23] [24]

[25]

[26]

[27]

[28]

[29]

Laio, A.; Gervasio, F. Metadynamics: a method to simulate rare events and reconstruct the free-energy in biophysics, chemistry and material science. Rep. Prog. Phys., 2008, 71, 126601. Bussi, G.; Laio, A.; Parrinello, M. Equilibrium free energies from nonequilibrium metadynamics. Phys. Rev. Lett., 2006, 96, 090601. Bonomi, M.; Branduardi, D.; Bussi, G.; Camilloni, C.; Provasi, D.; Raiteri, P.; Donadio, D.; Marinelli, F.; Pietrucci, F.; Broglia, R. A.; Parrinello, M. PLUMED: a portable plugin for free-energy calculations with molecular dynamics. Comput. Phys, Commun., 2009, 180, 1961. Pietrucci, F.; Laio, A. A collective variable for the efficient exploration of protein beta-sheet structures: application to SH3 and GB1. J. Chem. Theory Comput., 2009, 5, 2197-2201. Piana, S.; Laio, A.; Marinelli, F.; Van Troys, M.; Bourry, D.; Ampe, C.; Martins, J. Predicting the effect of a point mutation on a protein fold: the villin and advillin headpieces and their Pro62Ala mutants. J. Mol. Biol., 2008, 375, 460-470. Marinelli, F.; Pietrucci, F.; Laio, A.; Piana, S. A kinetic model of trp-cage folding from multiple biased molecular dynamics simulations. PLoS Comput. Biol., 2009, 5, e1000452. Todorova, N.; Marinelli, F.; Piana, S.; Yarovsky, I. Exploring the folding free energy landscape of insulin using bias exchange metadynamics. J. Phys. Chem. B., 2009, 113, 3556-3564. Cossio, P.; Marinelli, F.; Laio, A.; Pietrucci, F. Optimizing the performance of bias-exchange metadynamics: folding a 48-residue LysM domain using a coarse-grained model. J. Phys. Chem. B., 2010, 114, 3259-3265. Liwo, A.; Khalili, M.; Czaplewski, C.; Kalinowski, S.; Oldziej, S.; Wachucik, K.; Scheraga, H. Modification and optimization of the united-residue [UNRES] potential energy function for canonical simulations. I. Temperature dependence of the effective energy function and tests of the optimization method with single training proteins. J. Phys. Chem. B., 2007, 111, 260-285. Crespo, Y.; Marinelli, F.; Pietrucci, F.; Laio, A. Metadynamics convergence law in a multidimensional system. Phys. Rev. E., 2010, 81, 055701[R]. Laio, A.; Rodriguez-Fortea, A.; Gervasio, F.; Ceccarelli, M.; Parrinello, M. Assessing the accuracy of metadynamics. J. Phys. Chem. B., 2005, 109, 6714-6721. Kumar, S.; Rosenberg, J.; Bouzida, D.; Swendsen, R.; Kollman, P. The weighted histogram analysis method for free-energy calculations on biomolecules. I. The method. J. Comput. Chem., 1992, 13, 1011-1021.

Received: April 30, 2011

Revised: October 31, 2011

Accepted: November 02, 2011

Current Physical Chemistry, 2012, Vol. 2, No. 1 [30]

[31] [32]

[33] [34] [35]

[36] [37]

[38]

[39]

[40]

[41]

[42] [43]

91

Biarnés, X.; Pietrucci, F.; Marinelli, F.; Laio, A. METAGUI. A VMD interface for analyzing metadynamics and molecular dynamics simulations. Comput. Phys. Commun., 2012, 183, 203211. Bicout, D.; Szabo, A. Electron transfer reaction dynamics in nonDebye solvents. J. Chem. Phys., 1998, 109, 2325-2338. Hummer, G. Position-dependent diffusion coefficients and free energies from Bayesian analysis of equilibrium and replica molecular dynamics simulations. New J. Phys., 2005, 7, 34. Bortz, A.; Kalos, M.; Lebowitz, J. A new algorithm for Monte Carlo simulation of Ising spin systems. J. Comput. Phys., 1975, 17, 10-18. Buchete, N.; Hummer, G. Coarse master equations for peptide folding dynamics. J. Phys. Chem. B., 2008, 112, 6057-6069. Noé, F.; Horenko, I.; Schütte, Ch.; Smith, J. Hierarchical analysis of conformational dynamics in biomolecules: transition networks of metastable states. J. Chem. Phys., 2007, 126, 155102. Pietrucci, F.; Marinelli, F.; Carloni, P.; Laio, A. Substrate binding mechanism of HIV-1 protease from explicit-solvent atomistic simulations. J. Am. Chem. Soc., 2009, 131, 11811-11818. Cornell, W.; Cieplak, P.; Bayly, C.; Gould, I.; Merz, K.; Ferguson, D.; Spellmeyer, D.; Fox, T.; Caldwell, J.; Kollman, P. A second generation force field for the simulation of proteins, nucleic acids, and organic molecules. J. Am. Chem. Soc., 1995, 117, 5179-5197. Jorgensen, W.; Chandrasekhar, J.; Madura, J.; Impey, R.; Klein, M. Comparison of simple potential functions for simulating liquid water. J. Chem. Phys., 1983, 79, 926. 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. Berendsen, H.; van der Spoel, D.; Van Drunen, R. GROMACS: A message-passing parallel molecular dynamics implementation. Comput. Phys. Commun., 1995, 91, 43-56. Kranjc, A.; Bongarzone, S.; Rossetti, G.; Biarnés, X.; Cavalli, A.; Bolognesi, M.; Roberti, M.; Legname, G.; Carloni, P. Docking Ligands on Protein Surfaces: The Case Study of Prion Protein. J. Chem. Theory Comput., 2009, 5, 2565-2573. Leone, V.; Lattanzi, G.; Molteni, C.; Carloni, P. Mechanism of action of cyclophilin a explored by metadynamics simulations. PLoS Comput. Biol., 2009, 5, e1000309. Rossetti, G.; Cossio, P.; Laio, A.; Carloni, P. Conformations of the Huntingtin N-term in aqueous solution from atomistic simulations. FEBS Lett., 2011, 585, 3086-3089.

Protein Folding and Ligand-Enzyme Binding from Bias ...

In Section II, the convergence criteria, and the methods for analyzing data to ..... [19] a formalism was introduced which allows to map the history-dependent.

549KB Sizes 1 Downloads 225 Views

Recommend Documents

Protein Folding and Ligand-Enzyme Binding from Bias ...
Keywords: Enhanced sampling, Free energy calculations, Protein folding, Ligand-enzyme binding, ..... history dependent potential according to Eq. 1, but then sets. 0. 50. 100. 150 ...... LysM domain using a coarse-grained model. J. Phys.

protein folding pdf
There was a problem loading more pages. protein folding pdf. protein folding pdf. Open. Extract. Open with. Sign In. Main menu. Displaying protein folding pdf.

Protein folding by motion planning
Nov 9, 2005 - Online at stacks.iop.org/PhysBio/2/S148 .... from the roadmap using standard graph search techniques ... of our iterative perturbation sampling strategy shown imposed on a visualization of the potential energy landscape. 0. 5.

Protein Folding Paper - Ghosh.pdf
Download. Connect more apps... Try one of the apps below to open or edit this item. Protein Folding Paper - Ghosh.pdf. Protein Folding Paper - Ghosh.pdf. Open.

protein binding of drugs pdf
There was a problem previewing this document. Retrying... Download. Connect more apps... Try one of the apps below to open or edit this item. protein binding ...

ARE THERE PATHWAYS FOR PROTEIN FOLDING ?
A second approach involved the use of computer- ... display system, the molecule thus generated can be ... Finally, the computer system has been used in at-.

The Protein Folding Problem
In 1994, John Moult invented CASP (Critical. Assessment of Techniques ..... (Left) The density of states (DOS) cartoonized as an energy landscape for the three-helix bundle protein. F13W∗: DOS (x-axis) ... The peak free energy (here, where the DOS

Extracting Protein-Protein Interactions from ... - Semantic Scholar
statistical methods for mining knowledge from texts and biomedical data mining. ..... the Internet with the keyword “protein-protein interaction”. Corpuses I and II ...

Extracting Protein-Protein Interactions from ... - Semantic Scholar
Existing statistical approaches to this problem include sliding-window methods (Bakiri and Dietterich, 2002), hidden Markov models (Rabiner, 1989), maximum ..... MAP estimation methods investigated in speech recognition experiments (Iyer et al.,. 199

Social attributions from faces bias human choices
ders we select and the individuals we choose to trust. Nevertheless, our impressions of people are shaped by their facial appearances and, consequently, so too ...

Social attributions from faces bias human choices
1 Tepper School of Business, Carnegie Mellon University, Pittsburgh, PA, USA. 2 Department of ... and emotional state, yet face-based social inferences also.

-1cm Advertising Spending and Media Bias: Evidence from News ...
and participants in the 2016 Media Economics Workshop, the 10th Toulouse Internet ... this case illustrates the importance of media scrutiny and the potential social ... focusing on the top 100 recalls in terms of the number of potentially affected .

Protein Structure and Function: From Sequence to ...
Fax: +44 (0) 1235 465555. E-mail: ... For information on the online version of the text, please contact: Customer ... In the U.K., call free on 0800 389 8136. Fax: ...

Optimization and modeling of cellulase protein from Trichoderma ...
Jan 4, 2007 - Logistic kinetic model was the best model for the mixed substrates. A conceptual Artificial Neural. Network (ANN) model was well incorporated in the fermentative .... In reality the growth of cell was governed by a hyperbolic.

The fl-cell glibenclamide receptor is an ADP-binding protein - NCBI
were incubated for 2 h at room temperature with [3H]glibenclam- ide (25.3 d.p.m./fmol) in .... activity bound. Data are presented as Scatchard plots; each point is.

binding site
3: College of Computing, Georgia Institute of Technology, USA .... Systems Bioinformatics CSB2007, 13-17 August 2007, University of California, San Diego.

éBIAS/
Nov 13, 1995 - output signal of the photo-detector increases with an increas. U'S' P ...... digital measurement signal produced at the output 110 of the counter.

Bias Neglect
community is that there should be less aggression between ants that share a nest ... condition. Blind studies were much more likely to report aggression between.

Network random walk model of two-state protein folding
Jan 18, 2013 - View online: http://dx.doi.org/10.1063/1.4776215. View Table of ... Information Technology, National Institutes of Health, Bethesda, Maryland 20892, USA. 2School of ... the master equation, Eq. (1), describes stochastic dynam-.

Substrate Binding Mechanism of HIV-1 Protease from ...
Jul 31, 2009 - free energy of binding is -6 kcal/mol, and the kinetic constants for association and ..... SF counts the number of water molecules bridging between .... corresponding transition rates (ms-1) when larger than 100 ms-1 (solid line).

Bias Neglect
Experimenter bias occurs when scientists' hypotheses influence their results, even if involuntarily. Meta-analyses (e.g. ... This is true even when participants read descriptions of studies that have been shown to ... sometimes influence their result

éBIAS/
Nov 13, 1995 - source to cancel gain changes produced by changes in ambient ..... The analog output signal of the peak averager and memory circuit 90 is.