REVIEW The following article is Open access

Generalized-ensemble molecular dynamics and Monte Carlo algorithms beyond the limit of the multicanonical algorithm

Published 22 October 2010 2010 Vietnam Academy of Science & Technology
, , Citation Hisashi Okumura 2010 Adv. Nat. Sci: Nanosci. Nanotechnol. 1 033002 DOI 10.1088/2043-6254/1/3/033002

2043-6262/1/3/033002

Abstract

I review two new generalized-ensemble algorithms for molecular dynamics and Monte Carlo simulations of biomolecules, that is, the multibaric–multithermal algorithm and the partial multicanonical algorithm. In the multibaric–multithermal algorithm, two-dimensional random walks not only in the potential-energy space but also in the volume space are realized. One can discuss the temperature dependence and pressure dependence of biomolecules with this algorithm. The partial multicanonical simulation samples a wide range of only an important part of potential energy, so that one can concentrate the effort to determine a multicanonical weight factor only on the important energy terms. This algorithm has higher sampling efficiency than the multicanonical and canonical algorithms.

Export citation and abstract BibTeX RIS

Content from this work may be used under the terms of the Creative Commons Attribution-NonCommercial-ShareAlike 3.0 licence. Any further distribution of this work must maintain attribution to the author(s) and the title of the work, journal citation and DOI.

1. Introduction

Biomolecules such as proteins have complicated free energy surfaces with many local minima. Conventional molecular dynamics (MD) and Monte Carlo (MC) simulations in physical ensembles, such as the canonical [1–4] and isobaric–isothermal [56] ensemble, tend to get trapped in these local-minimum states. One of the powerful techniques to avoid this difficulty is generalized-ensemble algorithms [7–11], such as the multicanonical algorithm [12–15]. In the multicanonical ensemble, a free one-dimensional random walk is realized in the potential-energy space and a simulation does not get trapped in free-energy-minimum states.

However, because the multicanonical simulation is performed in a fixed volume, neither the pressure dependence nor the temperature dependence at certain pressure can be investigated as in experiments. In order to overcome this difficulty, I proposed recently the multibaric–multithermal algorithm [16–22]. In this ensemble, two-dimensional random walks are realized both in the potential-energy space and in the volume space, so that the temperature and pressure dependence can be discussed.

Another problem of the multicanonical algorithm is that the determination of the multicanonical weight factor to give a flat distribution becomes difficult in a large system. In order to alleviate this difficulty, I proposed the partial multicanonical algorithm [2324]. The partial multicanonical simulation samples a wide range of a part of the potential energy terms, which is necessary to sample the conformational space widely. Thus, one can concentrate the effort to determine a multicanonical weight factor only on the important energy terms, so that the partial multicanonical algorithm has a higher sampling efficiency than the multicanonical and canonical algorithms.

In this review article, I explain the multibaric–multithermal algorithm and partial multicanonical algorithm. Applications of these algorithms to an alanine dipeptide in explicit water are also presented.

The multicanonical algorithm is briefly reviewed in section 2. In section 3, the multibaric–multithermal algorithm is explained. The partial multicanonical algorithm is described in section 4. Section 5 is devoted to conclusions.

2. The multicanonical algorithm

In the canonical ensemble, the distribution of total potential energy E is given by

where n(E) is the density of states, β 0 is the inverse of the product of the Boltzmann constant k B and absolute temperature T0, at which simulations are performed. This ensemble has a bell-shaped distribution in the potential-energy space.

In the multicanonical ensemble [12–15], every state is sampled with a weight factor W muca (E) ≡ exp{−β 0 E muca (E)} so that a uniform distribution of E may be obtained,

where E muca (E) and W muca (E) are referred to as the multicanonical energy and the multicanonical weight factor, respectively. The difference between E muca and E is written as δE(E):E muca (E) ≡EE(E). The difference δE(E) is characteristic of the multicanonical simulation. The case of δE(E)=0 gives the regular canonical ensemble. The multicanonical simulation covers a wide range of E and escapes from local-minimum free-energy states. The expectation value of physical quantity A at temperature T is obtained by reweighting techniques [25] from

where 〈 ... 〉 muca is the multicanonical ensemble average. Because of the random walks in the potential-energy space, we can calculate physical quantities in a wide range of T. The multicanonical weight factor W muca (E) is not a priori known and has to be determined, for example, by the iterations of short simulations [26–28] and the Wang–Landau techniques [29].

3. The multibaric–multithermal algorithm

In the isobaric–isothermal ensemble [56], the distribution of potential energy E and volume V is given by

where n(E,V) is the density of states as a function of E and V and H is the 'enthalpy' (without the kinetic energy contributions): H=E+P0 V. Here, P0 is the pressure at which simulations are performed. This ensemble has bell-shaped distributions both in the potential-energy space and in the volume space, as shown in figure 1(a). In order to obtain the isobaric–isothermal ensemble, the combination of the Nosé thermostat [12] and the Andersen barostat [5] is frequently employed.

Figure 1

Figure 1 Distributions of potential energy E and volume V of a Lennard–Johnes fluid system (a) by isobaric–isothermal MD simulation and (b) by multibaric–multithermal MD simulation.

In the multibaric–multithermal ensemble [16–20], every state is sampled with a weight factor W mbt (E,V) ≡ exp {−β 0 H mbt (E,V)} so that a uniform distribution of both E and V, as shown in figure 1(b), may be obtained,

Here, W mbt (E,V) and H mbt are referred to as the multibaric–multithermal weight factor and the multibaric–multithermal enthalpy, respectively. The difference between H mbt and H is written as δH(E,V):H mbt (E,V)=HH(E,V). The difference δH(E,V) is therefore characteristic of the multibaric–multithermal simulation. The case of δH(E,V)=0 gives the regular isobaric–isothermal ensemble.

The isobaric–isothermal Hamiltonian for a system consisting of N atoms based on the Nosé thermostat [1, 2] and the Andersen barostat [5] is given by

The variable is the coordinate scaled by the length of the simulation box: ( r i is the real coordinate). We have introduced a simplified notation for the set of the scaled coordinates, . The variable is the conjugate momentum for . The real momentum p i is related to the virtual momentum by , where s is the additional degree of freedom for the Nosé thermostat. The variable PV and Ps are the conjugate momenta for V and s, respectively. The real-time interval Δt is associated with the virtual time interval Δt' by Δtt'/s. The constant mi is the mass of atom i. The constants W and Q are the artificial 'mass' related to V and s, respectively. In the case of a system consisting of Natoms, g equals 3N if the time development is performed in real time t, or g equals 3N+1 if the time development is performed in virtual time t'.

The equations of motion in real time are given from the Hamiltonian in equation (6) by

where the dot ( ⋅) stands for the real-time derivative d/d t and F i stands for the force acting on atom i. Performing the MD simulation by the equations of motion in equations (7)–(12), the multibaric–multithermal distribution in equation (5) is obtained.

In the case of δH(E,V)=∂ δH(E,V)/∂E=∂ δH(E,V)/∂V=0, the Hamiltonian in equation (6) and the equations of motion in equations (7)–(12) become those for the regular isobaric–isothermal MD simulation of the Nosé–Andersen formulation.

In order to perform a multibaric–multithermal MC simulation, we employ the Metropolis criterion on and V. The trial moves from to and is generated by uniform random numbers. The multibaric–multithermal enthalpy is consequently changed from to by the trial moves. The trial moves are now accepted with the probability

The multibaric–multithermal probability distribution is obtained by this scheme.

After an optimal weight factor W mbt (E, V) is determined—for example, by the iterations of short simulations [26–28] and by the Wang–Landau techniques [29]—a long production run is performed for data collection. The reweighting techniques [25] are used for the results of the production run to calculate the isobaric–isothermal-ensemble averages. The expectation value of a physical quantity A at T and P is then obtained by

where 〈 ... 〉 mbt is the multibaric–multithermal ensemble average. Because of the random walks both in the potential-energy space and in the volume space, we can calculate physical quantities in wide ranges of T and P.

I introduce here an application of the multibaric–multithermal MD algorithm to a system consisting of one alanine dipeptide molecule ((S)-2-(acetylamino)-N-methylpropanamide) and 63 water molecules [22]. The AMBER parm96 force field [30] was used for the alanine dipeptide molecule and the TIP3P rigid-body [31] model was employed for the water molecules. The electrostatic potential was calculated by the Ewald method. The time step was taken to be Δt=0.5 fs. The temperature was set as T0=600 K and the pressure was set as P0=300 MPa. The MD simulations were performed with the symplectic combination [32] of the Nosé–Poincaré thermostat [3334], the Andersen barostat [5] and the symplectic quaternion scheme [35]. In order to obtain an optimal multibaric–multithermal weight factor W mbt (E,V), two-dimensional versions of both the iterative method [26–28] and the Wang–Landau techniques [29] for multicanonical weight-factor determination were employed. An optimal weight factor was obtained after these preliminary simulations in total for 2.61 ns in the AMBER parm96 simulation. Two sets of the initial values of the alanine-dipeptide backbone dihedral angles were prepared for the data-production runs. The initial condition 1 was ϕ=ψ=180° and the initial condition 2 was ϕ=180° and ψ=0°. The initial conformation 1 is shown in figure 2. We then performed a long multibaric–multithermal MD simulation for data collection. The simulation time was 20 ns for each initial condition. The isobaric–isothermal MD simulations of the same system were also performed for comparisons.

Figure 2

Figure 2 The initial conformation of alanine dipetide for the MD simulations. The dihedral angles are ϕ=ψ=180°. The figure was created with RasMol [36].

Figure 3 shows the contour maps of the distributions  (ϕ, ψ) at T=240 K and P=0.1 MPa by the MD simulations started from the initial condition 1. The distribution  (ϕ, ψ) by the multibaric–multithermal MD were obtained using the reweighting techniques. The isobaric–isothermal distribution  (ϕ, ψ) has only four peaks of the α R P ,P II and C5 states, as shown in figure 3(a). On the other hand, the multibaric–multithermal MD simulation provides correct distributions  (ϕ, ψ) with all the six peaks of the α R P ,P II ,C5 L , and C7 ax states, as shown in figure 3(b). These results mean that the multibaric–multithermal simulation can get over the free energy barriers and has a high sampling efficiency.

Figure 3

Figure 3 Contour maps of the logarithms of the probability distributions (ϕ, ψ) of the backbone dihedral angles ϕ and ψ at T=240 K and P=0.1 MPa. (a) (ϕ, ψ) obtained from initial condition 1 by the conventional isobaric–isothermal MD simulation. (b) (ϕ, ψ) obtained from initial condition 1 by the reweighting techniques from the results of the multibaric–multithermal MD simulation.

The volume under the surface  (ϕ, ψ) around each peak corresponds to the population W of each state. The ratio of each state population W against the P II state population was calculated as a function of the inverse of temperature 1/T at the constant pressure of P=0.1 MPa, as shown in figure 4(a) and as a function of the inverse of pressure P at the constant temperature of T=298 K, as shown in figure 4(b). The difference in partial molar enthalpy ΔH and the difference in partial molar volume ΔV can be obtained from the gradient in figure 4(a) and that in figure 4(b), respectively. Partial molar enthalpy and partial molar volume are important quantities in solution chemistry, which describes the change in the ratio of each state population due to temperature and pressure changes. As the temperature increases, the relative population W C 5 /W P II of the C 5 state increases, as shown in figure 4(a). This indicates that the enthalpy for the C 5 state is higher than that of the P II state. The difference in partial molar enthalpy ΔH and the difference in partial molar volume ΔV of the C 5 state from that of the P II state, for example, are calculated from the derivative of W C 5 /W P II by

where R is the gas constant: R=8.3145 J (mol K)−1. The values of ΔH and ΔV are listed in tables 1 and 2, respectively. The differences ΔH and ΔV of the C 5 state and ΔH of the α R state are consistent in the multibaric–multithermal MD and the Raman spectroscopy experimental data [37]. The value of ΔV of the α R state, however, is smaller than that of the Raman spectroscopy.

Figure 4

Figure 4 The population ratios W/W P II against the P II (a), (a') as functions of the inverse of temperature 1/T at constant pressure of P=0.1 MPa and (b), (b') as functions of pressure P at constant temperature of T=298 K. These data were obtained by the reweighting techniques from the results of the multibaric–multithermal MD simulation of (a), (b) the C 5 state and (a'), (b') the α R state.

Table 1. Differences ΔH/(kJmol −1) in partial molar enthalpy of the C 5 and α R states from that of the P II state calculated by the multibaric–multithermal (MUBATH) MD simulations. Raman experimental data are taken from [37].

StateMUBATH MDRaman exp.
C 5 1.8±0.52.5±0.3
α R 4.4±1.14.4±1.5

Table 2. Differences ΔV/(cm 3mol −1) in partial molar volume of the C 5 and α R states from that of the P II state calculated by the multibaric–multithermal (MUBATH) MD simulations. Raman experimental data are taken from [37].

StateMUBATH MDRaman exp.
C 5 −0.3 ± 0.70.1±0.3
α R −2.3 ± 1.31.1±0.2

The multicanonical simulation can escape from local-minimum free-energy states but cannot change volume. The conventional isobaric–isothermal MD simulation can change volume but is trapped in a local free energy minimum. Therefore, ΔH and ΔV cannot be calculated by these simulation algorithms. The multibaric–multithermal algorithm has the advantages of both algorithms. This is the first simulational work to calculate ΔH and ΔV of biomolecules.

4. The partial multicanonical algorithm

Although the multicanonical algorithm is a powerful technique, the determination of the multicanonical weight factor is time consuming for a large system. In order to alleviate the effort to determine the weight factor, the total potential energy is divided into two terms in the partial multicanonical algorithm [2324],

where E1 is a part of the potential energy which is important to sample a wide range of the conformational space. The partial potential energy E1 can be taken in any manner, in principle. The electrostatic, Lennard Jones and torsion-angle energy terms between solute atoms may be included in a biomolecular system. The second term E2 is the rest of the potential energy. The partial multicanonical algorithm covers a wide range of E1. The partial multicanonical simulation is performed with a weight factor W pmuca (E1,E2) ≡ exp {−β 0 E pmuca (E1,E2)} so that a uniform distribution of E1 may be obtained,

where E pmuca (E1,E2) and W pmuca (E1,E2) are referred to as the partial multicanonical energy and the partial multicanonical weight factor, respectively. The difference between E pmuca and E depends only on E1 and is written as δE1(E1),

The difference δE1(E1) is characteristic of the partial multicanonical simulation. If we define here a distribution of E1 and the density of states n(E1) of E1 as

equation (17) is written as

The expectation value of a physical quantity A is obtained by the reweighting techniques [25] from

where 〈 ... 〉 pmuca is the partial multicanonical ensemble average. Note that 〈A NVT can be calculated correctly at a temperature T only in the vicinity of T0, although 〈A NVT can be calculated in a wide range of T in the multicanonical algorithm. This is because a limited range of E2 is sampled with the normal Boltzmann weight factor exp(−β 0 E2) in partial multicanonical algorithm.

The scheme of the partial multicanonical algorithm includes the canonical and the multicanonical algorithms. In the case of E1=0 and E2=E, this algorithm becomes the regular canonical algorithm. In the case of E1=E and E2=0, the multicanonical algorithm is obtained.

In order to perform a partial multicanonical MD simulation, a constant-temperature MD method is employed. The partial multicanonical Hamiltonian based on the Nosé thermostat [12] is given by

where p' i is the conjugate momentum for r i . The real momentum p i is related to the virtual momentum p' i by p i = p' i /s. In the Nosé–Hoover formalism [1–3], a variable is introduced and the equations of motion are given from the Hamiltonian in equation (23) by

Equation (25) is obtained from regular canonical-ensemble equations of motion by modifying force from

to

For a partial multicanonical MC simulation, we perform Metropolis sampling on the coordinates r . The trial moves from r i to r' i are generated by uniform random numbers. The partial multicanonical energy is consequently changed from E pmuca ( r ) to E' pmuca E pmuca ( r' ) by these trial moves. The trial moves are accepted by the Metropolis criterion [4] with the probability

The partial multicanonical probability distribution is obtained by this scheme.

The partial multicanonical MD algorithm was applied to a system consisting of one alanine dipeptide molecule and 67 water molecules [23]. The total potential energy is the sum of the potential energy between solute atoms E sltslt , that of the solute–solvent interaction E sltsvn , and that between solvent atoms E svnsvn , as follows:

Each term for the AMBER force field [30] is given by

where the subscripts bond, angle, torsion, imp, elec and LJ stand for the bond-length, bond-angle, torsion-angle, improper torsion-angle, electrostatic and Lennard Jones potential energy, respectively. Although these terms may vary for different force fields, the partial multicanonical algorithm can be applied in essentially the same way. Even if we sample wide ranges of bond-length, bond-angle and improper torsion-angle energy in E sltslt , the structure of the solute molecule is not expected to be changed very much. Sampling a wide range of the solvent energy E svnsvn would not change the solute structure very much, either. Sampling a wide range of the solute–solvent potential energy E sltsvn may sample a wide range of the interface area between the solute molecule and solvent molecules. The solute molecule of the alanine dipeptide in this simulation, however, is small and this effect was not taken into account in this simulation. Finally, the partial potential energy E1 was here taken to be the sum of the torsion-angle, electrostatic and Lennard Johnes energy between the solute molecules as follows:

and E2 was the rest of the potential energy,

Note that this choice of E1 is not always the best. The best choice of E1 should depend on the system and the force field.

In order to obtain an optimal partial multicanonical weight factor W pmuca (E1,E2), both the iterative method [26–28] and the Wang-Landau techniques [29] were employed. An optimal weight factor was obtained after these preliminary simulations for 2 ns in total. I then performed a long partial multicanonical MD simulation for data collection for 6 ns at T0=300 K.

The probability distributions of E and E1 are shown in figure 5. The distribution of E in the multicanonical simulation was much wider than those in the other two simulations. The partial multicanonical algorithm sampled a much wider range of E1. These generalized-ensemble MD simulations actually sampled a wide range of the potential energy, which and only which are predesignated to be sampled widely.

Figure 5

Figure 5 Logarithms of the probability distributions of (a) total potential energy E and (b) partial potential energy E1. The dotted lines are the results of the canonical MD simulation, the dashed lines are those of the multicanonical MD simulation and the solid lines are those of the partial multicanonical MD simulation.

The time series of ϕ is shown in figure 6. The canonical MD simulation did not make a complete rotation, as shown in figure 6(a). It covered only 65.6% of the whole range of ϕ. There seem to be potential energy barriers around ϕ=0° and ϕ=120°, which cannot be overcome by the canonical simulation (these energy barriers correspond to the boundaries between the local-minimum states). This means that the canonical simulation was trapped in local-minimum free-energy states. The multicanonical MD simulation also did not make a complete rotation and sampled 82.8% of the whole range of ϕ, as shown in figure 6(b). It overcame the potential energy barrier around ϕ=0° but did not overcome that around ϕ=120°. The partial multicanonical MD simulation, on the other hand, overcame both energy barriers and covered the whole range of ϕ, as shown in figure 6(c). The backbone dihedral angle ϕ underwent a complete rotation (covering 360°) six times (six turns) in this MD simulation for 6 ns.

Figure 6

Figure 6 The time series of the backbone dihedral angle ϕ obtained by (a) the canonical MD simulation, (b) the multicanonical MD simulation and (c) the partial multicanonical MD simulation.

Note that these results do not mean that the multicanonical algorithm cannot sample the whole range of ϕ. If a better weight factor W muca (E) were obtained with the longer preliminary simulations, the whole range of ϕ could be sampled also in the multicanonical ensemble. These results mean that the partial multicanonical algorithm shows higher sampling efficiency than the multicanonical algorithm if the weight factor is determined with the same effort.

The probability distributions (ϕ, ψ) of ϕ and ψ are shown in figure 7. In order to obtain the distribution (ϕ, ψ) in the canonical ensemble from the multicanonical simulation and that from the partial multicanonical simulation, the reweighting techniques in equation (3) and those in equation (22) were used, respectively. The probability distributions (ϕ, ψ) by the canonical simulation show four peaks of the P II , C 5, α R , and α P states, as shown in figure 7(a). However, they do not have any peaks in the range of 0° < ϕ < 120°. The distribution (ϕ, ψ) by the multicanonical MD simulation has five peaks of the P II , C 5, α R , α P and α L states, as shown in figure 7(b). The partial multicanonical MD simulation also sampled the C7 ax states in addition to the five peaks that were sampled in the multicanonical simulation, as shown in figure 7(c). In the present simulations, only the partial multicanonical simulation sampled all of the six states of the alanine dipeptide.

Figure 7

Figure 7 Contour maps of the logarithms of the probability distributions at T=300 K obtained (a) by the canonical MD simulation, (b) by the reweighting techniques from the results of the multicanonical MD simulation and (c) by the reweighting techniques from the results of the partial multicanonical MD simulation.

5. Conclusions

I have reviewed two new generalized-ensemble MD and MC algorithms: the multibaric–multithermal and partial multicanonical algorithms. These generalized-ensemble algorithms allow the simulation to escape from local-minimum free-energy states and sample many conformations.

The multibaric–multithermal simulation performs a two-dimensional random walk both in the potential-energy space and in the volume space so that one can obtain various isobaric–isothermal ensemble averages at different temperatures and pressures from only one simulation run. From the temperature and pressure dependences of an alanine dipeptide in explicit water, the differences in the partial molar enthalpy ΔH and the partial molar volume ΔV were calculated between the P II state and other states. This is the first work to calculate ΔH and ΔV by molecular dynamics simulations. The multibaric–multithermal algorithm will thus be a powerful simulation technique to study the temperature and pressure dependences of biomolecules like proteins.

In the partial multicanonical algorithm, the simulation covers a wide range of only important potential energy, which is necessary to investigate a wide range of the conformational space. Thus, one can concentrate the effort for the weight factor determination on the important energy terms. The MD simulation of an alanine dipeptide showed that the multicanonical simulation showed higher sampling efficiency than the canonical simulation. The partial multicanonical algorithm will thus be of great use for the protein folding problem. This idea, in which a simulation covers a wide range of only important potential energy terms, can be utilized not only for the multicanonical algorithm but also for multidimensional extensions of the multicanonical algorithm, such as the multibaric–multithermal [16–22] and multicanonical–multioverlap [38] algorithms.

Please wait… references are loading.
10.1088/2043-6254/1/3/033002