+ All Categories
Home > Documents > DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Date post: 01-Dec-2021
Category:
Upload: others
View: 2 times
Download: 0 times
Share this document with a friend
111
DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED FREE ENERGY METHODS By DANIAL SABRI DASHTI A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2013
Transcript
Page 1: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED FREE ENERGYMETHODS

By

DANIAL SABRI DASHTI

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOLOF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OFDOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2013

Page 2: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

c⃝ 2013 Danial Sabri Dashti

2

Page 3: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

I dedicate this dissertation to my wife and my family.

3

Page 4: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

ACKNOWLEDGMENTS

I primarily thank my advisor, Professor Adrian Roitberg for our scientific conversations

and discussions. I am also grateful for significant assistance of my lovely wife, Sahar,

during my Ph.D. period. I would never have been able to finish my dissertation without

the support of my family. I also like to thank Dr. Yilin Meng for his effective collaboration.

Moreover, thanks go out to all who supported me over my Ph.D study at the University of

Florida.

4

Page 5: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

TABLE OF CONTENTS

page

ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

LIST OF TABLES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

LIST OF FIGURES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9

ABSTRACT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

CHAPTER

1 INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

1.1 Prologue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121.2 Free Energies and Ensembles . . . . . . . . . . . . . . . . . . . . . . . . 12

1.2.1 Phase Space . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121.2.2 Partition Functions and Ensembles . . . . . . . . . . . . . . . . . . 131.2.3 Free Energy and Potential of Mean Force . . . . . . . . . . . . . . 14

1.2.3.1 Helmholtz and Gibbs free energies . . . . . . . . . . . . . 141.2.3.2 Potential of mean force . . . . . . . . . . . . . . . . . . . 15

1.3 Sampling in Biomolecular . . . . . . . . . . . . . . . . . . . . . . . . . . . 151.3.1 Sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 151.3.2 Standard Sampling Techniques in Biomolecular Simulations . . . . 16

1.3.2.1 Molecular dynamics . . . . . . . . . . . . . . . . . . . . . 161.3.2.2 Monte Carlo . . . . . . . . . . . . . . . . . . . . . . . . . 16

1.4 Advanced Sampling Methods . . . . . . . . . . . . . . . . . . . . . . . . . 181.4.1 Generalized Ensemble Methods . . . . . . . . . . . . . . . . . . . . 18

1.4.1.1 Simulated tempering . . . . . . . . . . . . . . . . . . . . . 181.4.1.2 Multicanonical algorithm . . . . . . . . . . . . . . . . . . . 191.4.1.3 Wang-Landau sampling . . . . . . . . . . . . . . . . . . . 191.4.1.4 Replica exchange . . . . . . . . . . . . . . . . . . . . . . 201.4.1.5 Umbrella sampling . . . . . . . . . . . . . . . . . . . . . . 22

1.4.2 Slow-Growth Methods . . . . . . . . . . . . . . . . . . . . . . . . . 221.4.2.1 Thermodynamic integration . . . . . . . . . . . . . . . . . 231.4.2.2 Free energy perturbation . . . . . . . . . . . . . . . . . . 24

1.5 Convergence, Error Estimation, and Sampling Quality in BiomolecularSimulations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 241.5.1 Error in Computational Methods . . . . . . . . . . . . . . . . . . . . 241.5.2 Ergodicity and Convergence . . . . . . . . . . . . . . . . . . . . . . 251.5.3 Sampling Quality Checks . . . . . . . . . . . . . . . . . . . . . . . 26

1.5.3.1 Root mean square deviation analysis . . . . . . . . . . . 261.5.3.2 Root mean square deviation clustering . . . . . . . . . . 261.5.3.3 Block averaging . . . . . . . . . . . . . . . . . . . . . . . 271.5.3.4 Principal component analysis . . . . . . . . . . . . . . . . 271.5.3.5 Kullback-Leibler divergence . . . . . . . . . . . . . . . . . 28

5

Page 6: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

1.6 Outline of My Research . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

2 COMPUTING ALCHEMICAL FREE ENERGY DIFFERENCES WITH HAMILTONIANREPLICA EXCHANGE MOLECULAR DYNAMICS (HREMD) Simulations . . . 31

2.1 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 312.2 Theory and Method . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33

2.2.1 Free Energy Perturbation(FEP) . . . . . . . . . . . . . . . . . . . . 332.2.2 Thermodynamic Integration . . . . . . . . . . . . . . . . . . . . . . 352.2.3 Hamiltonian Replica Exchange Molecular Dynamics (HREMD) . . 362.2.4 Simulation Details . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2.3 Results and Discussions . . . . . . . . . . . . . . . . . . . . . . . . . . . . 412.3.1 Acceptance Ratio of HREMD Simulations . . . . . . . . . . . . . . 412.3.2 Aspartic Acid Model Compound Study . . . . . . . . . . . . . . . . 412.3.3 Study on Asp26 in Thioredoxin . . . . . . . . . . . . . . . . . . . . 432.3.4 pKa Prediction for Asp26 in Thioredoxin . . . . . . . . . . . . . . . 44

2.4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3 LARGE PKa shifts FOR BURIED PROTONABLE RESIDUES. RATIONALIZINGTHE CASE OF GLUTAMATE 66 IN STAPHYLOCOCCAL NUCLEASE . . . . . 47

3.1 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 473.2 Simulation Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 483.3 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 503.4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

4 PH-REPLICA EXCHANGE MOLECULAR DYNAMICS IN PROTEINS USINGA DISCRETE PROTONATION METHOD . . . . . . . . . . . . . . . . . . . . . 62

4.1 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 624.2 Theoretical Method and Simulation Details . . . . . . . . . . . . . . . . . 65

4.2.1 Constant pH Molecular Dynamics and pH-Replica Exchange MolecularDynamics (pH-REMD) . . . . . . . . . . . . . . . . . . . . . . . . . 65

4.2.2 Titration Curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 664.2.3 pH-REMD . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 67

4.3 Simulation Details . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 684.4 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 69

4.4.1 Titratable Model Compounds . . . . . . . . . . . . . . . . . . . . . 694.4.2 ADFDA Model Compounds . . . . . . . . . . . . . . . . . . . . . . 694.4.3 Heptapeptide Derived from OMTKY3 . . . . . . . . . . . . . . . . . 71

4.5 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74

5 OPTIMIZATION OF UMBRELLA SAMPLING REPLICA EXCHANGE MOLECULARDYNAMICS BY REPLICA POSITIONING . . . . . . . . . . . . . . . . . . . . . 78

5.1 Literature Review . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 785.2 Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 81

6

Page 7: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

5.2.1 Umbrella Sampling Replica Exchange . . . . . . . . . . . . . . . . 815.2.2 Calculating Exchange Acceptance Ratio in Umbrella Sampling

Replica Exchange . . . . . . . . . . . . . . . . . . . . . . . . . . . 825.2.3 Umbrella Sampling Replica Exchange Optimization . . . . . . . . . 835.2.4 Umbrella Sampling Replica Exchange Optimization Workflows . . 845.2.5 Simulation Details . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

5.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . 875.3.1 Potential of Mean Force Along the Butane Dihedral Angle . . . . . 885.3.2 Potential of Mean Force for NH+

4 + OH− Salt Bridge in ExplicitSolvent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90

5.4 Concluding Remarks . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

6 CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93

APPENDIX

A CALCULATING THE PDF OF ∆ξ IN USRE . . . . . . . . . . . . . . . . . . . . 95

B ESTIMATING MEANS AND VARIANCES OF THE WINDOWS USING NEARESTNEIGHBOR WEIGHTED AVERAGING AND REWEIGTHING . . . . . . . . . . 96

B.1 Reweighting Fitted Gaussian Distribution on Windows . . . . . . . . . . . 98B.2 Nearest Neighbor Weighted Averaging (NNWA) . . . . . . . . . . . . . . . 99

C BARZILAI AND BORWEIN OPTIMIZATION (BB METHOD) . . . . . . . . . . . 100

REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 101

BIOGRAPHICAL SKETCH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 111

7

Page 8: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

LIST OF TABLES

Table page

2-1 Free energy difference (in Kcal/mol) between protonated and deprotonatedAspartic acids obtained from TI, REFEP, and FEP alchemical free energy simulations 43

4-1 pKas of the reference compounds computed by different methods . . . . . . . . 69

4-2 pKa prediction and Hill coefficient of fitted from the HH equation. . . . . . . . . 70

4-3 pKa values of the titratable residues in the heptapeptide derived from OMKTY3. 72

5-1 Position and EAR for four different settings of USRE calculations on the butanedihedral simulation in implicit water. . . . . . . . . . . . . . . . . . . . . . . . . 88

8

Page 9: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

LIST OF FIGURES

Figure page

2-1 Diagrams displaying the HREMD exchange algorithm and free energy calculation. 37

2-2 Thermodynamic cycle used to compute the pKa shift. . . . . . . . . . . . . . . 39

2-3 Cumulative average free energy differences between protonated and deprotonatedaspartic acid in the model compound (∆G(AH→ A−)) . . . . . . . . . . . . . . 42

2-4 Predicted pKa value of Asp26 in thioredoxin as a function of time . . . . . . . . 45

3-1 one turn of an α−helix exposes the side chain of GLU66. . . . . . . . . . . . . 50

3-2 Free energy convergence in REFEP for unrestrained V66E ∆+PHS (red) andGlu model compound (green). . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

3-3 Thermodynamic cycle for conformation protonation model of GLU66. . . . . . . 52

3-4 Cumulative Free energy VS Simulation time for Glu-66 restrained inside (Red)and restrained outside (Green). . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

3-5 Concentration of species vs. pH for two different value of Ke/b,D i.e., Ke/b,D;1 =2 and Ke/b,D;2 = 108. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

3-6 CSS, as defined in Equation 3–4, as a function of pH for Ke/b,D=10 . . . . . . . 56

3-7 Apparent pKa vs. Ke/b,D. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57

3-8 Carbon α RMSD distributions of residues 62 to 69 and the side chain of GLU66in unrestrained REFEP simulations . . . . . . . . . . . . . . . . . . . . . . . . . 58

3-9 Average secondary structure for residue 57-69 vs. reaction coordinate. . . . . 59

3-10 Ellipticity at 222 nm vs. reaction coordinate . . . . . . . . . . . . . . . . . . . . 60

4-1 Comparison of Hill plots between pH-REMD and CpH methods for Lys referencemodel. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70

4-2 The titration curves of Asp side chains in ADFDA computed by both constantpH MD (blue and purple) and pH-REMD (red and green) methods. . . . . . . . 71

4-3 Cumulative average protonation fractions of Asps side chains in ADFDA VSMC titration steps at pH=4.0. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 72

4-4 The titration curves of Asp3 in the heptapeptide derived from OMTKY3. . . . . 73

4-5 The titration curves of Lys5 and Tyr7 in the heptapeptide derived from OMTKY3. 74

4-6 Cumulative average protonation fraction for TYR7 versus MC titration steps atpH=8.0,9.0,10.0,11.0,12.0 and 13.0. . . . . . . . . . . . . . . . . . . . . . . . . 75

9

Page 10: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

4-7 The Kullback-Leibler divergence measure of RMSD distributions of CpH (Green)and pH-REMD (Red) respect to the final RMSD distribution in CpH. . . . . . . 76

4-8 RMSD Autocorrelation for CpH (Green) and pH-REMD (Red) at all pHs. . . . . 77

5-1 FBSF workflow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85

5-2 SBSF workflow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 86

5-3 ART probability density. The inset shows the PMF vs. reaction coordinate (dihedralangle) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 89

5-4 Window positions vs. EARs in the equi-distance (non-optimized) (red circles)and optimized (set 2) (green circles) simulations. . . . . . . . . . . . . . . . . . 90

5-5 The mean and RMSE of Kullback-Leibler divergence over 10 simulations foreach set. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 91

10

Page 11: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Abstract of Dissertation Presented to the Graduate Schoolof the University of Florida in Partial Fulfillment of theRequirements for the Degree of Doctor of Philosophy

DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED FREE ENERGYMETHODS

By

Danial Sabri Dashti

August 2013

Chair: Adrian RoitbergMajor: Physics

In recent decades, by increasing the power of computers’ hardware and developing

efficient algorithms, the study of large biomolecular systems has been facilitated. During

my PhD period, I attempted to improve the efficiency of sampling by developing new

algorithms and optimizing existing ones. My first project was about developing the

Replica Exchange Free Energy Perturbation (REFEP) method, which is a combination

of the Free Energy Perturbation (FEP) and the Hamiltonian Replica Exchange Molecular

Dynamics (HREMD) methods. We showed that the HREMD method not only improves

convergence in free energy calculations, but also can be used to estimate free energy

differences directly via the FEP algorithm. My next project was a demonstration of

the capabilities of REFEP in estimating the pKa of complicated proteins. According to

experimental measurements, the pKa value of Glutamate 66 (GLU66) in a hyperstable

mutant of staphylococcal nuclease displays a large shift, roughly 4.6 pH units, relative

to its normal value in water. In my third research project I developed and validated a

pH-Replica Exchange Molecular Dynamics (pH-REMD) method, which improves the

coupling between conformational and protonation sampling. Finally, in my last project,

I have focused on optimizing HREMD methods. The goal is to find the best position for

replicas in order to maximize the round trip between extremum positions on a replica

ladder.

11

Page 12: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

CHAPTER 1INTRODUCTION

1.1 Prologue

Molecular simulation is a branch of statistical physics. Since biophysical systems

are too complex and inhomogeneous to be treated theoretically, the role of numerical

simulations has been recognized. With recent advancements in computer hardware

and software, the simulations of more complex molecular systems become more

feasible. There are two basic problems in studying molecular systems using computer

simulations:

• Because of the large number of particles in those systems, the size of the phasespace is enormous.

• The accuracy of the molecular models is limited.

Many methods have been proposed to address those problems. This dissertation is

about the first problem, i.e. improving the efficiency of phase space sampling. More

specifically, I concentrate on the Replica Exchange sampling method[1–4], which is one

of the most frequently used techniques for efficient sampling of the phase space.

This chapter is dedicated to the fundamentals of statistical molecular mechanics,

sampling methods, and error analysis. I start with the concept of phase space and

partition functions and finish with introducing some of the frequently used sampling

analysis techniques. Also, I provide an overview on my research projects during my

Ph.D period, and in the next chapters, describe these projects in detail.

1.2 Free Energies and Ensembles

In this section I overview the key concepts in studying of a molecular system from

classical-statistical physics’ point of view.[5–9]

1.2.1 Phase Space

According to classical statistical mechanics, the state of a system can fully be

described by knowing the positions and momenta of its particles. There are six

coordinates associated with the configuration and the momentum of every particle

12

Page 13: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

in 3-dimensional space. Consequently, a system with N particles can be fully described

in a 6N-dimensional space, which is called phase space.

1.2.2 Partition Functions and Ensembles

An ensemble is a collection of all possible configurations of a system in phase

space. Each ensemble is concomitant with a partition function, which describes all

possible microstates of a system in equilibrium. All collective properties of the system

(e.g., average energy, entropy, and free energy) can be computed using the partition

function and its derivatives. Different types of ensembles can be defined based on the

types of communication between the system and the rest of universe. In the context of

statistical mechanics, the following are the most frequently used ensembles:

• Microcanonical ensemble (N,V, E): In a Microcanonical ensemble, volume(V), number of particles(N), and total energy (E) of a system are fixed, and thesystem has no energy communication with the universe.it can be represented asΩ(N,V, E) =

∫p

∫q

δ(H(q, p)− E) dp dq. Here, H, q, and p are the Hamiltonian, the

configuration, and the momentum of the system, respectively.

• Canonical ensemble (N,V,T): In this ensemble, V,N, and temperature (T) areconstant and the system is in thermal equilibrium with its environment. Thecanonical partition function is Q(N,V,T) =

∫p

∫q

exp(−βH(q, p))dp dq , where β =

1/kBT and kB is the Boltzmann constant. The majority of molecular simulationsare performed in this ensemble. In the case of conservative Hamiltonian, onecan separate the contributions of kinetic and potential energies. The kineticenergy contribution cancels out in most of the relative free energy calculations.Consequently it is appropriate to use configuration integral (or sum) instead ofthe full partition function, i.e., Z(N,V,T) =

∫q

exp(−βU(q)) instead of Q(N,V,T),

where U is the potential energy of the system. From now on, I will use the partitionfunction and the configuration integral interchangeably in this manuscript.

• Isothermal-isobaric ensemble (N,P,T): In this ensemble pressure (P), N ,and T are constant. The partition function can be written as Φ(N,P,T) =∫∞0

dVQ(N,V,T) exp(−βPV), which is the Laplace transform of the canonicalpartition function.

• Grand canonical ensemble(µ, V, T): In the Grand canonical ensemble, thechemical potential (µ), V, and T are fixed. The grand canonical partition functioncan be written as Ξ =

∑∞N=0 Z(N,V,T) exp(−βµN).

13

Page 14: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

The appropriate ensemble to use depends upon the physical circumstances, although

in the limit of large systems (i.e., thermodynamic limit), where the fluctuations are

insignificant, all descriptions of ensembles become indistinguishable. The fixed values

in each ensemble are called state variables. In the next section I describe the relation

between free energies, state variables, and partition functions.

1.2.3 Free Energy and Potential of Mean Force

Free energy is the maximum amount of work that a thermodynamic system can

perform. Free energy is a state function, i.e., in each ensemble it only depends upon

the state variables of that ensemble. From the statistical mechanics point of view, in

all ensembles (except the Microcanonical ensemble), free energies are related to the

logarithm of the partition functions. Here I explain Gibbs and Helmholtz free energies,

which are frequently used in biomolecular simulations.

1.2.3.1 Helmholtz and Gibbs free energies

Helmholtz free energy is the free energy associated with the canonical ensemble

and is equal to maximum extractable work from a closed thermodynamic system at

constant volume and temperature. It is related to the canonical partition function by:

A(N,V,T) =−1β

ln(Q(N,V,T)). (1–1)

On the other hand, the Gibbs free energy (also known as free enthalpy) measures

the useful work in the isothermal-isobaric ensemble and given by:

G(N,V,T) =−1β

ln(Φ(N, P,T)). (1–2)

Since the free energy is a state function, it takes a definite, non-fluctuating value

at equilibrium. It does not give us any information about the energy of the system as a

function of some specific reaction coordinates. In biomolecular simulations, instead of

free energy, it is conventional to use potential of mean force, which is a projection of free

energy along some chemical/alchemical/physical coordinates.

14

Page 15: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

1.2.3.2 Potential of mean force

The Potential of Mean Force (PMF) is a decomposition of the free energy along

some reaction coordinate parameters which measures how the energy of a system

changes along those coordinates. In order to calculate the PMF along some coordinates

in an ensemble, one needs to integrate the projection of the partition function over

all other degrees of freedom in configuration space. For example in the case of the

Canonical ensemble, the PMF along X can be calculated as:

PMF(X) =−1β

ln

(∫(exp(−βU(rN)δ(X− rN) drN

), (1–3)

where U(rN) is the potential energy.

1.3 Sampling in Biomolecular

Due to the complexity of potential energy and the large number of microstates,

analytical calculation of the partition function is not feasible except for very simple

systems with a few particles. Instead, in biomolecular simulations, we estimate the

partition function either by randomly sampling of the phase space or by propagating a

trajectory in the phase phase. In the next part I explain the concept of sampling as it is

presented in the statistics and statistical physics.

1.3.1 Sampling

In statistics, sampling is the process of selecting a subset of an original population.

A good sample is a statistical representation of the population properties, however as

the size of population increases appropriate sampling becomes more demanding. Due

to the size of phase space in biomolecular simulations, the importance of sampling has

been recognized. A good sample has two characteristics: it is unbiased, and it is large

enough to be precise. Later, I will explain more on those characteristics in the error

estimation section of this chapter.

15

Page 16: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

1.3.2 Standard Sampling Techniques in Biomolecular Simulations

Molecular Dynamics and Monte Carlo are the two fundamental sampling methods in

the field of molecular modeling. Almost all other sampling methods are either inherited

or branched from one or both of these methods.

1.3.2.1 Molecular dynamics

Molecular Dynamics (MD) [7, 10, 11] simulation propagates a system in the phase

space by numerically integrating Newton’s equations. According to MD, the trajectory of

each particle (in the case of the Microcanonical ensemble) can be obtained by iteratively

solving Equations 1–4: F(X) = −∇U(X) = mV(t)

V(t) = X(t), (1–4)

where X and V are the vectors of position and velocity of the particle respectively. Then,

based on the ergodic hypothesis (see section 1.5.2), the average of an observable A

can computed by:

⟨A⟩ = limt→∞

1

M

M∑i=1

A(ti). (1–5)

where A(ti) is the value of A at time ti, also M is the total number of tis. The computational

problem with regular MD is that it may be trapped in local minima of the potential energy

surface. Such a trapping prevents the simulation from sampling the whole phase space

efficiently.

1.3.2.2 Monte Carlo

The Monte Carlo (MC) [7, 12, 13] technique was used to generate the first computer

simulation of a molecular system. In a Monte Carlo simulation, a new configuration is

introduced by applying a random change to the positions of the current configuration

of the system. Unlike the MD simulations there is no momentum contribution in

MC simulations and consequently there is no dynamics involves in the simulation.

16

Page 17: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Estimating the partition function for a system of N atoms using regular Monte Carlo

comprises the following steps:

1. Creating a configuration by randomly generating 3N Cartesian coordinates.

2. Calculating the potential energy, U(rN), and Boltzmann factor, exp(−βU(rN)), forthat configuration.

3. Repeating steps 1 and 2, M times (M >> 1) until adding new configuration doesnot change the average of observable A in the system,

⟨A⟩ = limM→∞

M∑i=1

Ai exp(−βU(rN))

M∑i=1

exp(−βU(rN)). (1–6)

Since many of the generated configurations may have no significant contribution to the

partition function, this is not a practical approach in systems with a large number of

particles. One way to overcome this barrier is use importance sampling techniques,

i.e. to add configurations to the sample based on their associated probability in the

partition. This idea was introduced by Metropolis et al[14]. Unlike in regular Monte

Carlo, in Metropolis Monte Carlo (MMC), one chooses the configurations based on their

probabilities and weights them evenly instead of choosing configurations randomly and

weighting them with proper probability afterwards. According to the MMC technique,

one needs to add a new step between steps 2 and 3 of regular MC in order to accept or

reject a new configuration. The probability of acceptance is

Pacc = min (1, exp (−β∆U)) , (1–7)

where ∆U is the difference between the potential energy of old and new configurations.

Considering the computational cost of MMC, although many of the randomly produced

configurations may be rejected, still one has to calculate the associated energy of all the

configurations to decide about acceptance or rejection.

17

Page 18: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

1.4 Advanced Sampling Methods

In biomolecular systems with many degrees of freedom, there are large numbers

of local minima in the potential energy surface. Conventional sampling methods, such

as MC and MD, tend to be trapped in those minima, and consequently give inaccurate

thermodynamic averages. Advance sampling methods help to overcome this limitation.

There are many types of advanced sampling techniques. Here I explain two of the

more prominent categories of those: generalized ensemble methods and slow-growth

methods.

1.4.1 Generalized Ensemble Methods

Generalized Ensemble Methods[3, 15–17] overcome the problem of trapping in local

minima by changing the Boltzmann probability weights to non-Boltzmann weights such

that a random walk in potential energy space is realized.

1.4.1.1 Simulated tempering

In simulated tempering [15], temperature dynamically changes with time such

that the system accomplishes a random walk in temperature space. The probability of

visiting a microstate is proportional to the simulated tempering weight:

WST(E, T) = exp(−βE + a(T)), (1–8)

where a(T) is chosen such that the probability distribution of microstates in temperature

space becomes flat, i.e.:

PST(T) =

∫dEn(E)WST(E, T) = const, (1–9)

where n(E) is the density of energy states. In numerical simulations temperature axes is

limited to a list of M discrete values and the simulation can hop between those values.

The optimal values of a(T) should be determined by iteration of trial simulations. The

following is a summary of the ST algorithm:

1. Running a canonical MD or MC simulation at temperature Ti for a certain steps.

18

Page 19: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

2. Changing Ti to one of its earliest neighbors in the temperature list with transitionprobability of:

π(Ti → Ti±1) = min

(1,

WST(E, Ti±1)

WST(E, Ti),

), (1–10)

which is based on MMC criteia.

1.4.1.2 Multicanonical algorithm

In the Multicanonical Algorithm (MUCA) [17, 18], a flat potential energy distribution

is achieved by weighting each microstate using a non-Boltzmann weight factor, WMUCA,

which implies:

p(E) ∝ n(E)WMUCA(E) = const. (1–11)

This induces a random walk in the energy space of the system, which prevents a

simulation from trapping in local minima of the potential energy surface of the system.

Similar to the ST method, the density of states is not a priori known and the MUCA

weight factor should be determined by iteratively running short trial simulations[17, 19].

One can perform MUCA using both MD and MC methods. Moreover the average

of physical quantity A at temperature T can be calculated using single reweighting

techniques:

⟨A⟩T =

∑i

A(qi)W−1MUCA(E(qi)) exp (−βE(qi))∑

i

W−1MUCA(E(qi)) exp (−βE(qi))

(1–12)

where qi is the ith configuration.

1.4.1.3 Wang-Landau sampling

The Wang-Landau algorithm[16] is a MMC technique for estimating the density

of states. It can be used as a complementary tool for determining the weight factor in

MUCA. In this method, we initially set all n(E) = 1 and the probability of jumping from an

19

Page 20: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

energy state Ei to Ej is

p(Ei → Ej) = min

(1,

n(Ej)

n(Ei)

). (1–13)

which induced a random walk in enery space of the system. Each time the current

energy density of state updates as following:

n(E)← n(E)f, (1–14)

where f is the modification factor. Typically a simulation begins with f = e ≃ 2.71828.

During the random walk, we accumulate a histogram H(E), which is the number of

visits at each energy level E. When the histogram becomes flat, we reset the density of

states, update the modification factor f ←√f , and restart the whole process from the

beginning. We can continue this scheme until we reach desired accuracy in estimating

of the density of states.

1.4.1.4 Replica exchange

The Replica Exchange (RE) method, which originally was proposed by Swendsen[1,

20], is one of the most successful methods in the field of molecular simulations. This

method has two advantages: First, unlike ST and MUCA, in RE, the weight factor is

known a priori. Second, RE can intuitively be combined with many other enhanced

sampling methods[21–24]. In this technique, M independent copies (replicas) of the

system are treated by MD (i.e., REMD) or MC (i.e., REMC) simultaneously, and they

ask to exchange their configurations after every certain numbers of steps of simulation.

By exchanging configurations between replicas, a random walk in the replicas’ ladder

space is realized. Based on the type of RE, replicas can span temperature space

(Temperature RE), Hamiltonian space (Hamiltonian RE), or both (Multi Dimensional RE).

The core of the exchange criteria is based on imposing a detailed balance equation to

20

Page 21: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

the generalized ensemble:

w(X→ X)℘(X) = w(X→ X)℘(X), (1–15)

where ℘(X) and w(X → X) are the probability of being at state X and the probability of

transition from X to X, respectively. X and X are generalized ensembles before and after

the exchange, i.e.,

X =

T1 ... Ti Tj ... TM

H1 ... Hi Hj ... HM

q1 ... qi qj ... qM

, X =

T1 ... Tj Ti ... TM

H1 ... Hj Hi ... HM

q1 ... qi qj ... qM

(1–16)

Here Hi,Ti, and qi represent the Hamiltonian, temperature, and configuration of replica

i just before exchange. Because all replicas are independent, the probability of the

system being at the generalized state of X is ℘(X) =∏M

i=1 pi(Ti, Hi, qi), where

pi(Ti, Hi, qi) is the probability that replica i, with Hamiltonian Hi and temperature of

Ti is found with the conformation state of qi. In the case of canonical ensemble

pi(Ti, Hi, qi) ∝ exp(−βiHi(qi)), (1–17)

where βi =1

kBTi. Then the detailed balance equation can be written as:

w(X→ X)

w(X→ X)=

pi(Tj, Hj, qi)pj(Ti, Hi, qj)

pi(Ti, Hi, qi)pj(Tj, Hj, qj)=

exp(−βjHj(qi)) exp(−βiHi(qj))

exp(−βiHi(qi)) exp(−βjHj(qj))= exp(−∆),

(1–18)

where

∆ = [(βjHj(qi) + βiHi(qj))− (βiHi(qi) + βjHj(qj))] . (1–19)

Then using the MMC crieria, the probability of transition from X to X can be calculated

as:

w(X→ X) = min (1, exp(−∆)) . (1–20)

21

Page 22: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

In the case of TRE, where all Hamiltonians are the same, Equation 1–19 can be

simplified to:

∆ = [(βj − βi)(H(qj)− H(qi))] , (1–21)

and in the case of HRE, where all temperatures are equal, it can be written as:

∆ = β [Hj(qi) + Hi(qj)− Hi(qi)− Hj(qj)] . (1–22)

1.4.1.5 Umbrella sampling

The Umbrella Sampling method, which was introduced by Torrie and Valleau[25]

in 1977, is one of the major techniques for calculating the PMF along preset reaction

coordinates. In this method, we restrain the system at different part/parts of the reaction

coordinate by adding single/multiple bias potential/potentials. The bias is an additional

potential energy term, which restrains the system to the reaction coordinate. Usually, the

bias is a quadratic function of the reaction coordinate:

B(ξ) = k(ξ − ξ0)2, (1–23)

where k, ξ, and ξ0 are the bias strength, the order parameter and the bias center on

the reaction coordinate and configuration respectively. In order to extract an estimation

of the PMF as a function of a reaction coordinate, it is necessary to remove the effect

of bias using post-processing methods. Many methods have been proposed in the

literature; among those, the Weighted Histogram Analysis Method (WHAM)[26] and

Umbrella Integration (UI)[27, 28] are the most promising.

1.4.2 Slow-Growth Methods

Slow-growth methods are used to compute the PMF between two given states (e.g.,

A and B) of a system by summing the free energy differences between the intermediate

22

Page 23: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

steps. Here the potential energy for intermediate stages is defined as:

U(λ) = UA + λ(UB − UA), (1–24)

here λ is called the coupling parameter, U(0) = UA , and U(1) = UB. In slow-growth

techniques, intermediate states i.e. 0 < λ < 1, may not have any physical meaning.

Consequently, for the case of the canonical ensemble, the Helmholtz free energy at the

state of λ can be written as:

A(N,V,T,λ) =−1β

ln(Z(N,V,T,λ)), (1–25)

where Z(N,V,T,λ) =∑

i exp(−βUi(λ)) is the partition function for the state of λ. Here

I explain Thermodynamics Integration (TI) and Free Energy Perturbation (FEP), which

are among the prevalent slow-growth methods. One of the applications of this technique

is the calculation of pKas of different species in proteins[29]. In this type of calculation,

one of the two end states (i.e., λ = 0 or 1) characterizes the protonated state, and the

other represents the deprotonated state or vice versa. The intermediate λs correspond

to hybrid protonated and deprotonated states.

1.4.2.1 Thermodynamic integration

According to the TI technique, the free energy difference between states A and B is:

∆A(A→ B) =

∫ 1

0

dλ∂A

∂λ. (1–26)

Using Equation 1–25 this can be rewritten as:

∆A(A→ B) =

∫ 1

0

dλ−1βZ

∂Z

∂λ=

∫ 1

0

dλ1

Z

∑i

exp(−βUi(λ))∂Ui(λ)

∂λ=

∫ 1

0

dλ⟨∂U(λ)∂λ⟩λ,

(1–27)

where the bracket average represent an ensemble average generated at λ. In practice,

one has to break the integration to the sum of infinitesimally different intermediate

23

Page 24: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

states. Moreover, the integration path should be reversible, i.e., changes in λ should be

small enough, and the system should be relaxed at each λ value.

1.4.2.2 Free energy perturbation

The Free energy perturbation method, which was introduced by Zwanzig in

1954[30], is another frequently used free energy difference method. Considering

the canonical ensemble, the Helmholtz free energy difference between states A and B

can be expressed as:

∆A(A→ B) =−1β

lnZB

ZA

=−1β

ln∑ exp(−βUB)

ZA

=−1β

ln∑ exp(−βUA)

ZA

exp(−β(UB − UA))

=−1β

ln ⟨exp(−β(UB − UA))⟩A.

(1–28)

During the derivation of Equation 1–28, we implicitly assumed the number of energy

microstates in states A and B to be equal, but this is only true when two states are

infinitesimally close. One can use the Equation 1–24 to adapt the Equation 1–28 for

computing the free energy difference between distinct states in the system:

∆A(A→ B) =−1β

M−1∑i=1

ln ⟨exp(−β(Uλi+1− Uλi

))⟩λi, (1–29)

where M is the number of discrete steps which has been used for moving from state A

to state B.

1.5 Convergence, Error Estimation, and Sampling Quality in BiomolecularSimulations

1.5.1 Error in Computational Methods

There are two types of errors in estimating a variable using computational methods,

systematic and random. Random or statistical errors, which are associated with

precision of measurements, are fluctuations in the measured data due to imperfect

sampling. In order to estimate the random error, one needs to repeat the simulation

24

Page 25: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

many times and calculate the variance of the measured variable. In contrast, systematic

errors or biases are reproducible inaccuracies, which usually come from the deficiency

of energy and force computation methods. The bias can be defined as a difference

between the correct value and the measured value. The estimation of bias is complicated,

since the correct value of a variable is not a priori known. To overcome this problem,

typically the bias is measured with respect to a more sophisticated model (e.g., bias of

a force field method with respect to a quantum method). On the other hand, random

error is associated with the quality of sampling. In the next section I describe the ergodic

theory, which is the base of many sampling quality measurement methods.

1.5.2 Ergodicity and Convergence

Ergodic theory, which originated from Boltzmann’s work in statistical physics,

describes the behavior of a dynamical system when it runs for a long time. According

to this hypothesis, a system eventually visits all points in its phase space, if it is allowed

to run for enough time. One of the consequences of ergodic theory asserts that, under

certain conditions, the time average of a function over a dynamics trajectory is equal to

the space average of that function, i.e.:

⟨A⟩ =∫p

∫q

p(q, p))dp dq = limt→∞

1

t

∫ t0+t

t0

A(τ)dτ , (1–30)

where p(q, p) is the probability of the system having at configuration q and the

momentum of p, moreover t0 is an arbitrary origin on the time coordinate.

The ergodic theory is closely related to the concept of convergence. In the context

of molecular simulations, a simulation is considered converged when the collected

set (sampled set) is a precise representation of the original ensemble. Moreover, a

system is called semi-ergodic, when some states are not accessible during a reasonable

simulation time, so the system will not converge easily. Since the original ensemble is

not feasible, full convergence is not practicable either. In other words, one can approach

the ideal ensemble, but cannot achieve it. In the next parts, I discuss some tests that

25

Page 26: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

quantify the degree of confidence about a simulation convergence. We have to note that

those assessments only describe necessary assurances, but not the sufficient ones.

1.5.3 Sampling Quality Checks

One of the most important questions in sampling methods is this: when have

we collected enough samples? To answer this question, we need to quantify the

convergence rate and the sampling quality in a simulation. Here I describe the most

employed convergence analysis tools in computational biophysics and biochemistry.

1.5.3.1 Root mean square deviation analysis

Root Mean Square Deviation (RMSD) of the configurations measures the average

movements of particles in a snapshot from (usually) the initial snapshot. For a system

with N particles, It is defined as :

RMSDt =

√∑Ni=1 |Xi(t)− X0

i |2N

(1–31)

where |Xi(t)− X0i | is the spatial displacement of particle ith between time t and time 0.

One way to track the convergence rate is to calculate the cumulative RMSD average

as function of time:

⟨RMSD⟩t =∑

t RMSDt

n. (1–32)

where n is the number of snapshots from beginning to time t. This method is useful

for tracking the folding state of protein, since in a folded state the protein fluctuates

around the average configuration. The definition above is not unique and there are other

variations of RMSD definitions[31–35].

1.5.3.2 Root mean square deviation clustering

Clustering is the process of organizing a set of objects into groups whose members

are related in some way. Clustering is a part of many statistical analyses. However,

considering the long list of those methods is beyond the scope of this dissertation.

Among them, Root mean square deviation clustering is frequently used in the context

26

Page 27: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

of molecular simulations. This method can be used for dividing an ensemble into

sets of self-similar structures. Daura et al[36] adopted this method for assessing the

convergence rate in the simulation. According to this method, a significant decrease

in the rate of discovery of new clusters during the simulation is a sign of convergence.

However a closer investigation reveals that this condition is necessary but not sufficient.

For example, trapping in energy minima can significantly decrease the rate of cluster

discovery in a simulation, although the simulation is not converged.

1.5.3.3 Block averaging

Block Averaging (BA), as proposed by Flyvbjerg and Petersen[37], is a method

for examining the quality of sampling in a correlated simulation. This method can be

described as following: A trajectory with time length of τ = M.n snapshots is split in to M

different block of size n, starting with small n, e.g., n = 1. The mean of observable A is

calculated for each block. Then one can calculate the Blocked Standard Error (BSE) as:

BSE(A, n) = σn/√M, (1–33)

where σn is the variance among M blocked averages. For small values of n (i.e., small

blocks) blocks are more correlated, So BSE underestimates the standard error. As the

value of n increases (or M decreases) the correlation decreases and BSE converges

to a value. Plotting the BSE(A, n) vs. n comprises a signal of both statistical error

convergence and decorrelation of A.

1.5.3.4 Principal component analysis

Principal Component Analysis (PCA) [? ] is a technique for calculating the

large-scale characteristic motions from a simulation trajectory. In order to perform

PCA, one needs to create the 3N × 3N (where N is the number of atoms) covariance

matrix as:

Cij = ⟨(xi − xi) (xj − xj)⟩ (1–34)

27

Page 28: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

and compute the eigenvalues and eigenvectors of the matrix. Where xi represents

ith degree of freedom in the configuration space. Each eigenvalue represents the

mean square deviation along the corresponding eigenvector. The modes with higher

eigenvalues signify the main directions of motion in the system. Another application of

PCA is for measuring the degree of similarity in the variations of two trajectories[38].

Also, this can be used as a test of convergence in a single long trajectory by dividing the

long trajectory into trajectories with smaller lengths.

1.5.3.5 Kullback-Leibler divergence

Kullback-Leibler divergence[39, 40] or relative entropy is defined as:

DK−L(P||Q) =∑i

P(i) ln

(P(i)

Q(i)

), (1–35)

where P and Q are the probability distributions of a random variable. It is the average,

over the distribution P, of the logarithmic difference between the probabilities P and

Q.It measures the degree to which P is distinguishable from Q. DK−L is non-negative

quantity and a small value of it indicates that P and Q are highly overlapped. This metric

can be used as a measure of convergence rate, where Q is a target (or reference)

distribution and P is changing with time.

1.6 Outline of My Research

During my Ph.D studies, I have been mostly focused on developing and testing

enhanced sampling methods and mainly Hamiltonian Replica Exchange Molecular

Dynamics (HREMD).

My first project, which is described in the next chapter, was about developing

Replica Exchange Free Energy Perturbation (REFEP) method, which is a combination

of the Free Energy Perturbation (FEP) and HREMD[22]. We demonstrated that HREMD

method not only improves convergence in alchemical free energy calculations, but also

can be used to compute free energy differences directly via the FEP algorithm. We

showed a direct mapping between the HREMD and the usual FEP equations, which are

28

Page 29: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

then used directly to compute free energies. We tested REFEP on predicting the pKa

value of the buried Asp26 in thioredoxin. We compared the results of REFEP with TI and

regular FEP simulations. REFEP calculations converged faster than those from TI and

regular FEP simulations. The final pKa value predicted from HREMD simulation was only

0.4 pKa units above the experimental value. Briefly, we showed REFEP algorithm not

only improves conformational sampling, but also improves the convergence rate of free

energy simulations.

My next project, i.e., the third chapter, was a demonstration of the capabilities of

REFEP in estimating the pKa of complicated proteins[41]. According to the experimental

measurements, the pKa value of Glutamate 66 (GLU66) in a hyperstable mutant of

staphylococcal nuclease displays a large shift, roughly 4.6 pKa units, relative to its

normal value in water, as measured in the lab of Moreno et al [83, 84, 92–98]. In order

to reproduce the large experimental shift using a single structure, continuum solvent

and computational methods, an internal dielectric constant around 10 is necessary. The

physical reason for this is not yet understood, but hypotheses have been produced by

Moreno et al [91, 97, 99, 104] regarding solvent penetration, protein reorganization, etc.

We aimed to resolve this inconsistency between experimental and continuum methods

by introducing a four-state thermodynamic cycle that couples conformational states

with protonation states of GLU66. We proposed that the experimental methods (which

are mostly sensitive to configurational changes) measure the equilibrium constant

between the two configurational states instead of the two protonation states. We used

REFEP method in implicit solvent to calculate the pKa value of GLU66 for each of the

configurational states as well as the mixed configuration. The results are in almost

perfect agreement with the experiments of Moreno et al.

In my third research project I developed and validated a pH-Replica Exchange

Molecular Dynamics (pH-REMD) method[24]. This method improves the coupling

between conformational and protonation sampling. Under a HREMD setup, conformations

29

Page 30: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

are swapped between two neighboring replicas, each having different pHs. We applied

pH-REMD to a series of model compounds, a terminally charged ADFDA pentapeptide,

and a heptapeptide derived from the ovomucoid third domain (OMTKY3). In all of those

systems, the predicted pKa values by pH-REMD were very close to the experimental

values and almost identical to the ones obtained by constant pH molecular dynamics

(CpH MD).

In the last year of my PHD, I have focused on optimizing HREMD methods. The

goal is finding the best position for replicas in order to maximize the round trip between

extremum positions on replica ladder. We developed, validated, and tested a method

for estimating the probability of exchange between neighboring replicas in Umbrella

Sampling Replica Exchange MD (USRE)[42]. We use information from very short

umbrella runs, needing only a handful of windows. We designed a multi dimensional

scoring function to optimize the set of replicas (windows). By maximizing the scoring

function, we enforce the same exchange acceptance for all neighbor replica pairs. We

found having equal exchange acceptance between pairs increases the number of round

trips and improves the efficiency of sampling. A description of this method can be found

in chapter 5.

30

Page 31: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

CHAPTER 2COMPUTING ALCHEMICAL FREE ENERGY DIFFERENCES WITH HAMILTONIAN

REPLICA EXCHANGE MOLECULAR DYNAMICS (HREMD) SIMULATIONS

2.1 Literature Review

Free energy, especially the free energy difference between two states, is a crucial

quantity in the study of chemical and biological systems.[43] Knowledge of the free

energy differences can help us understand the behaviors of such systems. For example,

the free energy of binding is one of the criteria used to evaluate the performance of

drugs.[44] Therefore, one important aspect of molecular modeling is to yield accurate

free energy differences efficiently. Many free energy calculation methodologies (such as

free energy perturbation,[30] thermodynamic integration,[45] umbrella sampling,[25, 46,

47] and Jarzynski’s equality [48] as well as analysis techniques (such as the weighted

histogram analysis method[49] and Bennett acceptance ratio method [50, 51] have been

developed to achieve this goal. In general, free energy calculations could be divided

into alchemical free energy and conformational free energy calculations. The alchemical

free energy calculations are often employed when studying the free energy differences

of processes that involve changes in noncovalent interactions. In an alchemical free

energy simulation, a nonphysical reaction coordinate λ is generally adopted in to

connect the initial and final states. This reaction coordinate is usually expressed as

an interpolation of the initial and final states. Thus, an alchemical process is achieved

through a series of intermediate states having no direct physical meaning. Since the free

energy difference between two states is a state function, the actual choice of coordinate

cannot, in the limit of infinite sampling, affect the results. Free energy perturbation (FEP)

Reprinted with permission from Meng, Y.; Dashti, D. S.; Roitberg, A. E. J. Chem.Theory Comput. 2011, 7, 2721–2727.

31

Page 32: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

and thermodynamic integration (TI) are two common methodologies that are utilized in

alchemical free energy computations.

One important issue in alchemical free energy calculations is the convergence of

the free energy difference versus computational cost. Such convergence is particularly

difficult in systems involving slow structural transition or large environmental reorganization

as λ changes.[21, 52, 53] Therefore, conformational sampling is crucial in alchemical

free energy calculations. Enhanced sampling methods, such as replica exchange

molecular dynamics (REMD),[2] orthogonal space random walk (OSRW),[53] and

accelerated molecular dynamics (AMD)[54] have been applied to free energy simulations

in order to accelerate conformational sampling and, in turn, to yield accurate and

converged free energy differences. Among the enhanced sampling methodologies, the

REMD method is of particular interest because the weight of each state is a priori known

(Boltzmann factor).

Both temperature-based and Hamiltonian-based REMD have been applied to

alchemical free energy calculations. Woods et al.[21] and Rick[60] have combined the

temperature-based REMD with TI calculation. A temperature-based REMD simulation

is conducted at each state along the reaction coordinate. Woods et al.[21] have also

applied the HREMD methodology to FEP and TI calculations. Each replica in the

HREMD simulation represents a state along the reaction coordinate λ, and a periodic

swap in λ is attempted. Relative solvation free energy of water and methane as well

as the relative binding free energies of halides to calis pyrrole have been calculated

in this way[21]. The Yang group has developed a dual-topology alchemical HREMD

(DTA-HREM) method[61]. Their method was tested on the free energy of mutating

an asparagine amino acid (with two ends blocked) to leucine. More recently, the

Roux group coupled the FEP methodology with the distributed replica technique

(REPDSTR)[62, 63]. An additional acceleration in the sampling of the side-chain

dihedral angle was also incorporated when Jiang and Roux utilized the FEP/HREMD

32

Page 33: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

method to study the absolute binding free energy of p-xylene to the T4 lysozyme L99A

mutant [63]. In all of those studies, the conformational sampling and convergence of

free energy computations showed significant improvement when the REMD method

was applied. The protocol presented here accelerates convergence but, of course,

does not solve known problems in the field related to enhanced sampling of coordinates

orthogonal to λ space, which would hamper many of the current methods.

In this chapter, we will demonstrate that FEP is actually already incorporated in the

HREMD method in an elegant and formal way. The REFEP method is shown to be not

only an enhanced sampling method but also a free energy calculation algorithm. We will

apply the REFEP method to the pKa prediction of thioredoxin Asp26. The experimental

pKa value of 7.5 has been shown to be one of the largest shifted from the intrinsic pKa

value[64, 65] and, hence, makes it an interesting case to be studied theoretically. TI and

FEP (regular molecular dynamics for conformational sampling) alchemical free energy

simulations have been conducted in order to compare with REFEP simulations. A very

accurate theoretical pKa value is obtained from REFEP simulations. The convergence

of the free energy difference and pKa value is achieved in REFEP simulations much

faster than that in the FEP and TI simulations. The advantage and simplicity of using the

HREMD simulation to compute the alchemical free energy difference is clearly shown.

2.2 Theory and Method

2.2.1 Free Energy Perturbation(FEP)

The FEP method, which was initially introduced by Zwanzig in 1954[30], is a well

established method and is considered the most frequently employed methodology

in alchemical free energy calculations[52]. The details of the FEP, as well as the TI,

methodology and its applications have been extensively reviewed[52, 66–69]. Therefore,

only a very brief description of the FEP and TI methods will be given here. Consider two

states (1 and 2) of a system in the canonical (NVT) ensemble, and their corresponding

Helmholtz free energies A1 and A2. The Helmholtz free energy difference between two

33

Page 34: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

states can be expressed as

∆A1→2 = −kBT ln⟨exp([U2(q)− U1(q)]

kBT

)⟩1. (2–1)

Here, kB is the Boltzmann constant, T is the temperature, and q is the molecular

structure. U1 and U2 are the potential energies of states 1 and 2, respectively.

The bracket with subscript 1 stands for the average calculated over the structural

ensemble generated by state 1. In order to compute ∆A1→2, one simulation of state

1 is performed. Once a configuration q is taken, the potential energy difference at

configuration q is computed. The ensemble average, which is ⟨exp(

[U2(q)−U1(q)]kBT

)⟩1,

can be calculated easily, and hence, ∆A1→2 is obtained. Although the Helmholtz free

energies are utilized here, Equation 2–1 can be extended to an isothermal-isobaric

(NPT) ensemble and to the Gibbs free energy in the same manner.

When the fluctuations in ∆U in Equation 2–1 are too large, FEP calculations are

notoriously hard to converge. The convergence of the FEP calculation will be poor if

the overlap in phase space between the two states is small. In order to compute the

free energy difference between two states that are very different, intermediate states

mixing the two end points are adopted in such a way that the differences between

neighbors can be treated as perturbations. A frequently employed method to generate

intermediate states is to interpolate potential energy functions linearly, as shown in

Equation 2–2. In Equation 2–2, U1 and U2 are the potential energy functions of states

1 and 2, respectively. Free energy differences between neighboring states are then

computed. The sum of individual free energy differences will be the targeted free energy

difference between states 1 and 2 (Equation 2–3). There are many ways of executing

FEP calculations involving intermediate states. The double-ended, double-wide[67, 70],

and overlap sampling algorithms[71] are among the most popular ones. A thorough

description of different algorithms and their performance can be found in a recent review

34

Page 35: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

by Jorgensen and Thomas[67].

U(λ) = U1 + λ(U2 − U1), (2–2)

∆A1→2 = −kBT∑i

ln⟨exp([U(λi+1)− U(λi)]

kBT

)⟩i. (2–3)

In practice, computing ∆A1→2 (forward free energy difference) is equally easy (or hard)

as computing ∆A2→1 (backward free energy difference), and one is exactly the opposite

of the other in principle. Evaluation of forward and backward free energy differences

provides an indication of convergence. Furthermore, the potential energy differences

generated from both directions can be utilized to reduce statistical error. The Bennett

acceptance ratio (BAR) method is a frequently employed scheme to improve the

precision of a free energy estimator[50–52].

2.2.2 Thermodynamic Integration

Another way of writing the free energy difference between two states 1 and 2 is

∆A1→2 = −kBT∑i

ln⟨exp([U(λi+1)− U(λi)]

kBT

)⟩i. (2–4)

Here, λ is a reaction coordinate connecting states 1 and 2, and U is the potential energy

of a state along the reaction coordinate. The bracket represents an ensemble average

generated at a value of λ. The integration is often evaluated numerically via trapezoidal

rule or Gaussian quadrature. If U(λ) is constructed as in Equation 2–2, the derivative of

U(λ) with respect to λ is

∂U(λ)

∂λ= U2 − U1. (2–5)

And the free energy difference between states 1 and 2 can be expressed as

∆A1→2 =

1∫0

⟨U2 − U1⟩λdλ (2–6)

35

Page 36: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Hence, the ensemble average of the potential energy gap between states 1 and 2 at

each λ value is needed in a TI calculation. In this chapter, we use the term TI to refer to

constrained TI, in which the value of λ is not allowed to change at each window.

2.2.3 Hamiltonian Replica Exchange Molecular Dynamics (HREMD)

The original REMD method utilizes replicas having different temperatures (TREMD).

Replicas at high temperatures overcome potential energy barriers more easily than

those at low temperatures. Another way to overcome potential energy barriers is simply

to change the potential energy surface to reduce potential energy barriers. In the

HREMD algorithm, replicas differ in their Hamiltonians but have the same temperature.

Regular MD is performed, and an exchange of configurations between two neighboring

replicas is attempted periodically.

Figure 2-1 demonstrates the HREMD algorithm and the free energy computation in

an HREMD simulation. Let us consider two replicas 1 and 2 with corresponding potential

energies U1 and U2. By employing the detailed balance condition and Boltzmann weight

of each molecular structure, the transition probability can be written as

w(q1 → q2) = min 1, exp [−(U1(q2) + U2(q1)− U1(q1)− U2(q2))/kBT] (2–7)

where q1 and q2 are the molecular structures of replicas 1 and 2 before an exchange

attempt, respectively. A Monte Carlo Metropolis criterion[14] is used to evaluate whether

the attempted swap of structures between two replicas should be accepted or not.

Equation 2–7 can be regrouped as

w(q1 → q2) = min 1, exp− [(U2(q1)− U1(q1) + U1(q2)− U2(q2))/kBT (2–8)

When comparing the exponential terms in Equation 2–1 and 2–8, it is clear that

Equation 2–8 incorporates all information necessary for a FEP calculation. U2(q1) −

U1(q1) is the potential energy difference computed on the basis of the structural

ensemble generated by U1, while U1(q2) − U2(q2) is the potential energy difference

36

Page 37: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Figure 2-1. Diagrams displaying the HREMD exchange algorithm and free energycalculation. (A) Exchange attempt orders. Replicas connected by a curveare neighbors, and attempts are made to exchange molecular congurations(q). (B) Free energy calculations in the HREMD method. Each replica hastwo free energy differences: Aup and ∆Adown from its attempting neighborform a pair and are computed simultaneously, while ∆Adown and ∆Adown fromits attempting neighbor form the other pair. In exchange attempts (regardlessif the attempts are accepted or rejected), two pairs of free energy differencesare computed in an alternating fashion utilizing Equation 2–1.

computed on the basis of the structural ensembles generated by U2. Every time the

transition probability is computed, those potential energy differences can be utilized to

compute the ensemble average shown in Equation 2–1. Therefore, ∆A1→2 and ∆A2→1

can be computed on-the-fly utilizing the double-ended scheme. The ensemble average

in Equation 2–1 is computed regardless of whether an exchange attempt is accepted

or rejected. When employing the HREMD method to improve conformational sampling

in the study of alchemical changes, HREMD simulations are able not only to enhance

conformational sampling but also to yield the free energy difference directly. In fact,

37

Page 38: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

a regular FEP calculation can be thought of as an HREMD calculation in which no

exchanges are allowed between replicas.

In practice, as shown in Figure 2-1, there are two free energy difference calculations

(∆Aup and ∆Adown) continuously associated with each replica. Take replica 1 as an

example: ∆Aup = ∆A1→2 while ∆Adown = ∆A2→1. In principle, when converged, ∆A1,up

should be equal to the negative of ∆A2,down:

∆A1,up = −kBT ln⟨exp([U2 − U1]

kBT

)⟩1 = −∆A2,down

= kBT ln⟨exp([U1 − U2]

kBT

)⟩2

(2–9)

Any difference (except for the sign) between the two is an indication of error or lack of

convergence.

Convergence also was gauged by the time dependence of the predicted free energy

differences, computing ∆G versus simulation length. This provides an asymptotically

unbiased estimator for ∆G, thus all methods presented here must eventually reach the

same final value (within error bars). REFEP is presented in this chapter as showing

faster convergence toward the final value.

2.2.4 Simulation Details

Accurately determining the pKa values of ionizable residues, especially those

with large shifts from intrinsic pKa values, is of great interest both experimentally and

computationally[64, 65, 72]. Here, the pKa calculation of Asp26 in thioredoxin has been

selected as a test case in order to compare the performance of alchemical free energy

simulations. Asp26 has been found deeply buried in thioredoxin and possesses one of

the largest pKa shifts among protein carboxylic groups[64, 65]. Following the protocol

employed in the paper of Simonson et al [72], the thermodynamic cycle utilized to

compute the pKa value of an ionizable residue is given in Figure 2-2. As it can be seen

there, the use of a model compound as an auxiliary leg in the thermodynamic cycle

makes ∆G3 (proton to proton) equal to zero. Essentially, the pKa shift relative to the

38

Page 39: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

intrinsic value (pKa,model) is computed as

pKa(protein) = pKa(model) +1

2.303kBT

[∆G(proteinAH→ proteinA−)−∆G(AH→ A−)

](2–10)

where ∆G(proteinAH → proteinA−) and ∆G(AH → A−) are the free energy differences

between protonated and deprotonated aspartic acid in the protein environment and in

aqueous solution, respectively. Alchemical free energy simulations were performed in

order to yield those two terms. In Equation 2–10, the Gibbs free energy differences are

used because experiments determining pKa values generally are conducted under an

isobaric-isothermal condition.

Figure 2-2. Thermodynamic cycle used to compute the pKa shift. Both acid dissociationreactions occur in aqueous solution. The protein-AH represents the ionizableresidue in a protein environment. The AH represents the model compoundwhich is usually the same ionizable residue with capped terminii. In practice,a proton does not disappear but instead becomes a dummy atom. Theproton still has its position and velocity. The bonded interactions involvingthe proton are still effective. However, there are no nonbonded interactionsfor that proton. The change in the ionization state is reflected by changes ofpartial charges in the ionizable residue.

Aspartic acid dipeptide in implicit water solvent was taken as the model compound

with a pKa value taken as 4.0[73]. The oxidized form of thioredoxin (PDB code

2TRX)[74] in implicit water was used in our simulation. Changes in ionization were

represented by changes in the partial charges of the aspartic acid side chain (ASH→ASP

39

Page 40: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

in the AMBER terminology). Since the van der Waals radius of the proton in aspartic

acid is zero for both protonated and deprotonated species, the free energy difference

only contains the electrostatic interactions.

Three types of free energy simulations were performed for both the model

compound and the protein: TI (forward and backward), HREMD-FEP (REFEP), and

regular FEP simulations. Our regular FEP simulations were carried out via HREMD

simulations but with all exchange attempts rejected. Comparing the pKa prediction

and free energy convergence from FEP and REFEP simulations will directly indicate

the effect of the enhanced conformational sampling due to the exchanges. Linear

interpolation of point charges was carried out in order to assign side chain charges for

intermediate states. A seven-point Gaussian quadrature was selected to compute total

free energy difference for TI calculations. Therefore, eight λ values (one end point is

needed in either direction) were utilized in the TI simulation. Due to the implementation

of the TI algorithm in AMBER [75], 16 replicas were utilized to ensure the same amount

of simulation time for all free energy simulations. A simulation time of 5 ns was used

for each λ value and for each replica in the study of the model compound, while for

thioredoxin, we used 4 ns runs. Structural swaps between neighboring replicas were

attempted every 2 ps (1000 MD steps). No particular attempt was made in this work to

optimize the number or location of the replicas, nor the exchange attempt frequency.

All simulations were done using the AMBER 10 molecular simulation suite[75],

locally modified to add HREMD/REFEP capabilities. The AMBER ff99SB force field [76]

was utilized in all of the simulations. The SHAKE algorithm[77] was used to constrain

the bonds connecting hydrogen atoms with heavy atoms in all of the simulations,

which allowed the use of a 2 fs time step. The OBC (Onufriev, Bashford, and Case)

generalized Born implicit solvent model (igb = 5 in the AMBER terminology)[78] was

used to model the water environment in all of our calculations. The cutoff for nonbonded

interaction and the Born radii was set to 99 A. This value is larger than the dimension of

40

Page 41: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

both systems. Langevin dynamics was employed in order to maintain the temperature at

300 K, using a friction coefficient of 3.0 ps−1.

2.3 Results and Discussions

2.3.1 Acceptance Ratio of HREMD Simulations

The accuracy of FEP depends on the overlaps between phase spaces, which can

be measured as overlaps between potential energy difference distributions[52]. The

acceptance ratio in an HREMD simulation is an indication of the overlap between two

potential energy difference distributions[61]. Therefore, it could be utilized to monitor

the convergence of free energy calculation qualitatively. In our study, large acceptance

ratios were observed in both the model compound and protein HREMD simulation.

The acceptance ratio between two neighbors ranged from 0.7 to 0.9 in all HREMD

simulations. Those large acceptance ratios indicate that the overlap in phase space is

large.

2.3.2 Aspartic Acid Model Compound Study

The free energy differences on the right-hand side of Equation 2–10 were

calculated as described in the Theory and Method section. The cumulative average

free energy difference as a function of time is reported here. Figure 2-3A shows the

∆G(AH → A−) from TI, HREMD, and FEP simulations (as mentioned before, a FEP

simulation has been performed by rejecting all exchange attempts in an HREMD

simulation). The differences between forward and backward ∆G(AH → A−) are shown

in Figure 2-3B. A converged alchemical free energy simulation should generate the

same forward and backward free energy numerically (except for an opposite sign). Any

nonzero value is an indication of free energy not converged.

For a simple system such as aspartic acid in implicit water, 5 ns of simulation time

was long enough for ∆G(AH → A−) to stabilize in all three alchemical free energy

simulations, as shown in Figure 2-3A. The forward and backward ∆G(AH → A−) at

the end of each free energy calculation and the corresponding error bars are listed

41

Page 42: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Figure 2-3. (A) Cumulative average free energy differences between protonated anddeprotonated aspartic acid in the model compound (∆G(AH→ A−)). (B)The differences between forward and backward ∆G(AH→ A−). (C)Cumulative average free energy differences between protonated anddeprotonated Asp26 in thioredoxin (∆G(proteinAH→ proteinA−)). (D) Thedifferences between forward and backward(∆G(proteinAH→ proteinA−)).

42

Page 43: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Table 2-1. Free energy difference (in Kcal/mol) between protonated and deprotonatedAspartic acids obtained from TI, REFEP, and FEP alchemical free energysimulations

TI REFEP FEP

ASP modelforward -59.43± 0.06 -59.69 ± 0.05 -59.84 ± 0.06backward -59.56± 0.06 -59.66 ± 0.05 -59.72 ± 0.06average -59.50 ± 0.08 -59.68 ± 0.08 -59.78 ± 0.08

Asp26 in thioredoxinforward -54.35 ± 0.61 -54.29 ± 0.17 -54.23 ± 0.56backward -55.82 ± 0.39 -54.24 ± 0.14 -53.84 ± 0.56average -55.09 ± 0.72 -54.27 ± 0.22 -54.04 ± 0.79

∆G differenceforward 5.08 ± 0.61 5.40 ± 0.18 5.61 ± 0.56backward 3.74 ± 0.39 5.42 ± 0.15 5.88 ± 0.56average 4.41 ± 0.72 5.41 ± 0.23 5.74 ± 0.79

predicted pKa,protein

forward 7.7 ± 0.4 7.9 ± 0.1 8.1 ± 0.4backward 6.7 ± 0.3 7.9 ± 0.1 8.3 ± 0.4average 7.2 ± 0.5 7.9 ± 0.2 8.2 ± 0.6

in Table 2-1. The forward and backward free energy differences are the same (within

error bars) for both REFEP and FEP simulations. However, the TI simulations failed

to do that, although the difference was very small (the difference between forward

and backward ∆G(AH → A−) was only 0.13Kcal/mol). The average of forward and

backward ∆G(AH → A−) was taken as the final value of ∆G(AH → A−) for the model

compound and is also reported in Table 2-1. Clearly, as shown in Figure 2-3B, the

REFEP simulations have converged much faster than the FEP calculations did.

2.3.3 Study on Asp26 in Thioredoxin

The free energy difference between protonated and deprotonated Asp26 is shown

in Figure 2-3C and D. By analogy with the model compound plots, the cumulative

average as a function of time is reported. The cumulative average was clearly not

converged during the TI simulation, and neither was the difference between forward

and backward ∆G(proteinAH → proteinA−). According to Table 2-1, after 4 ns of TI

simulation, the difference between forward and backward free energy was 1.4 Kcal/mol,

while the uncertainty of the forward and backward free energy differences was 0.61

and 0.39 Kcal/mol, respectively. Data not presented here show that TI requires roughly

43

Page 44: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

40 ns of dynamics before converging to results comparable with FEP/REFP. It is worth

noting that this comparison is slightly unfair to TI and deserves further explanation.

First, we used eight intermediate states for TI versus 16 for FEP/REFEP. This setup,

when executed within Amber, uses the same CPU time since the TI implementation

is done with dual-topology methods. In fact, reusing the ensemble generated with

the FEP Hamiltonians and computing TI values on that ensemble produces very

rapidly-converging results.

For regular FEP free energy calculations, the cumulative averages stabilized

after roughly 2.2 ns of simulation, while the cumulative averages for the REFEP

simulation stabilized much more rapidly (shown in Figure 2-3C). Furthermore, Figure

2-3D illustrates that the difference between forward and backward ∆G(proteinAH →

proteinA−) in the REFEP reached a value very close to zero (-0.05 Kcal/mol) very

quickly. As described previously, the final value of ∆G(proteinAH → proteinA−) was

calculated as the average of forward and backward free energy differences. Although

the final free energy differences computed from 4 ns of simulation were the same for

REFEP and regular FEP, the calculations converged much faster in REFEP than in FEP

simulation. Since the HREMD and FEP calculations only differed in whether structures

were allowed to be exchanged or not, the improvement in alchemical free energy

convergence resulted from employing enhanced conformational sampling technique is

significant. Data not presented here show that the histograms of P1(∆U) exp(−β(∆U))

for the calculation of the free energy difference between replicas 1 and 2 for different

sampling times are slightly different for FEP and REFEP. The REFEP distributions

converge faster with time and sample the left side of the distribution better. This helps

rationalize the faster convergence of our technique.

2.3.4 pKa Prediction for Asp26 in Thioredoxin

The pKa value of Asp26 in thioredoxin can be computed from Equation 2–10.

The final value of ∆G(proteinAH → proteinA−) from the REFEP simulation was

44

Page 45: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0 1000 2000 3000 40006.0

6.5

7.0

7.5

8.0

8.5

9.0

9.5

10.0P

redi

cted

pK

a V

alue

Time (ps)

REFEP, forward REFEP, backward FEP, forward FEP, backward

Figure 2-4. Predicted pKa value of Asp26 in thioredoxin as a function of time. The(∆G(AH→ A−)) values utilized in Equation 2–10 were -59.68 and -59.78Kcal/mol for REFEP and FEP, respectively. The experimental value is 7.5.

-54.3 Kcal/mol, with a predicted pKa value of 7.9, which is only 0.4 pKa units above

the experimental value. The predicted pKa value with respect to time from REFEP

simulations was plotted in Figure 2-4 in order to demonstrate the convergence of

the pKa prediction. Figure 2-4 shows that REFEP simulations not only yielded an

accurate predicted pKa value but also achieved convergence very fast. The regular FEP

simulation predicted a pKa value of 8.2, which is 0.7 pKa units above the experimental

value. The convergence in the regular FEP simulation was also worse than that in the

REFEP simulation.

45

Page 46: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

2.4 Concluding Remarks

Conformational sampling is crucial in free energy calculations. In the case of

alchemical free energy calculations, HREMD is a useful and popular method to

enhance the accuracy and convergence of free energy simulations. In this chapter,

we have demonstrated that REFEP not only improves conformational sampling in

free energy calculations but also yields a free energy difference directly via the FEP

algorithm. The implementation of REFEP is trival, once a HREMD code is in place.

The REFEP alchemical free energy calculation was tested on predicting the pKa value

of Asp26 in thioredoxin and compared with TI and regular FEP simulations. Free

energy differences from the REFEP simulation converged faster than those from TI

and regular FEP simulations. The final predicted pKa value from the REFEP simulation

was very accurate, only 0.4 pKa unit above the experimental value. Utilizing the REFEP

algorithm significantly improves conformational sampling, and this in turn improves the

convergence of alchemical free energy simulations.

46

Page 47: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

CHAPTER 3LARGE PKa SHIFTS FOR BURIED PROTONABLE RESIDUES. RATIONALIZING THE

CASE OF GLUTAMATE 66 IN STAPHYLOCOCCAL NUCLEASE

3.1 Literature Review

The pKa of an ionizable species is related to the equilibrium constant between

its protonated and deprotonated states. Because the electrostatic potential and the

protonation equilibrium are coupled, protein electrostatics can be probed indirectly by

pKa measurements. Ionizable groups are usually solvent-exposed and situated near

the protein surface, where the pKa shift (as measured versus the same residue by itself,

in the same solution, usually called a model compound) of these groups is relatively

small and ascribed to the coulombic interactions with other ionizable residues and salt

bridges. On rare occasions, proteins have buried ionizable residues that behave very

differently from those located near the surface. These exceptional cases can be very

useful to understand protein electrostatics.

Buried titratable groups can play critical roles in a variety of situations[79–82]. The

charged forms of ionizable groups are more favorable in strongly polar environments,

while the neutral species of ionizable groups are dominant in hydrophobic environments.

This preference for the neutral species in the hydrophobic interior of proteins causes

a large shift in the pKa of buried titratable groups , acidic groups have pKa values

shifted significantly higher and basic groups have pKa values shifted significantly lower

[83, 84]. Measuring and rationalizing the pKa of internal ionizable groups is crucial to

understanding biochemical processes such as charge transfer [82, 85–87], molecular

recognition [88], and ion transport [89, 90]. Electrostatic interactions between titratable

sidechains govern all pH-dependent properties of proteins [91].

Recent studies show that the hydrophobic interior of some proteins can tolerate

charged residues without any significant structural adaptation. Garcia Moreno

et al. have studied the stability and pKa shift of E. coli’s staphylococcal nuclease

(SNase) mutants using fluorescence, circular dichroism spectroscopy (CD), and other

47

Page 48: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

experimental methods [83, 84, 92–98]. In one of those studies, a valine-to-glutamate

mutation at position 66 (V66E) has been introduced to ∆ + PHS, a hyper-stable mutant

of SNase. This substitution results in a highly shifted pKa value for glutamate (GLU) 66

compared to that of the GLU model compound. Different experimental methods have

reported this shifted value as 9.0 ± 0.2 (i.e., ∆pKa ≈ 4.6) [92, 94, 99, 100], making it

one of the largest pKa shifts ever reported for a titratable residue. On the other hand,

continuum solvent computational methods (e.g., the Generalized Born model) predict a

smaller pKa shift for GLU66 [97, 99].

According to the Born formulation [101, 102], this shift corresponds to the transfer

of a charged group from water to a medium of dielectric constant of 10-12. This value,

required for improved agreement between continuum solvent methods and experimental

results in this case, is much larger than the usual value of 2 to 4[101, 103]. There is no

obvious physical rationale behind this high value of dielectric constant [91, 97, 99, 100].

Therefore different reasons have been proposed to rationalize this discrepancy. Some

of the most significant ones are changes in the state of ionization of other residues,

interaction with buried water, protein relaxation[83, 84], and interactions with other

ionizable residues[91, 97, 99, 104]. Even after accounting for these effects, a dielectric

constant around 6 is still necessary to reproduce experimental data[91, 97, 99, 105].

In this chapter we show that to understand these systems properly, one needs to

explicitly treat the conformational and protonation equilibria as coupled, and there is no

single structure that can be seen as responsible for the pKa switch.

3.2 Simulation Details

We built the V66E mutant of the hyperstable form of staphylococcal nuclease

(SNase) known as ∆ + PHS (pdb code 3BDC17). As before, all simulations were

performed using the Amber FF99SB force field [76] as implemented in the Amber

10 molecular simulation package [75]. The SHAKE algorithm [77] was used to keep

bonds involving hydrogen atoms at their equilibrium length. Newton’s equations were

48

Page 49: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

integrated with a timestep of 2 fs. We minimized the system using 500 steps of the

steepest descent algorithm, followed by 500 steps of the conjugate gradient algorithm.

Next, we heated the system to 300 K in three steps, followed by a final equilibration in

the canonical ensemble. All calculations were performed using the Langevin Thermostat

[106] and distinct random seeds (ig=-1) in order to generate independent simulations

[107]. The generalized Born implicit solvent model [102, 108] was applied to model the

water environment in all calculations. The cutoff used for non-bonded interactions

and calculating Born radii was 999 A. We used the recently developed Replica

Exchange Free Energy Perturbation method (REFEP)[22] (Details in chapter 2) to

calculate the free energy difference between the protonated and deprotonated states

in each conformational state (i.e., either buried or exposed GLU66) and in a mix of

conformational states.

Three sets of REFEP simulations were performed on this system: (1) conformationally

unrestrained, (2) a 0.4 kcalmol−1A−2 harmonic restraint on all heavy atoms in the

conformation with GLU66 buried and (3) the same restraints in the conformation where

GLU is exposed to the solvent. In all three sets of simulations, 16 replicas were used to

perturb the protonated GLU66 (λ = 0) to the deprotonated state (λ = 0). The length of

the simulation was 5 ns for simulation(1) and 4 ns for simulations (2) and (3). Also a 5 ns

REFEP simulation was performed on a Glu reference compound (i.e., Ace−Glu− Nme)

using 16 replicas to perturb the protonation state of Glu to establish the deprotonation

free energy of a free Glutamic acid in solution.

Hereafter, we will refer to the pKa of GLU66 as the pKa of GLU66 in V66E ∆ + PHS

and to the pKa shift as the difference between this value and that of the reference

compound in water (4.4).

In this work, we used the CDPro software package [109] to calculate the Circular

Dichroism (CD) spectra as the protonation of the system is modified. CDPro computes

49

Page 50: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

the CD spectra as function of conformation, using parameterized dipole-dipole

interactions[110].

3.3 Discussion

According to experimental methods such as X-ray, TRP Florescence, and CD

Spectroscopy, ionizing the GLU66 side chain in V66E ∆ + PHS triggers the unwinding of

one turn of an α−helix, exposing the previously buried carboxylic group of GLU66 to the

bulk solvent, shown in Figure 3-1.

Figure 3-1. one turn of an α−helix exposes the side chain of GLU66.

As explained in the introduction, there is inconsistency between pKa shift calculated

by experimental methods and continuum solvent computational methods. In this part we

tried to rationalize this inconsistency.

We used REFEP to calculate the free energy difference between the two-protonation

states for both the GLU reference compound and GLU66 in V66E ∆ + PHS and turned

that data into a pKa shift due to the protonation. The REFEP data is shown in Figure

50

Page 51: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

3-2. Comparing the final free energies in both directions (Figure 3-2) shows that both

53 54 55 56 57 58 59 60 61 62

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

fre

e e

ne

rgy

(kc

al/

mo

l)

simulation time (ns)

A) Final backward free energy

V66E ∆+PHS, UnrestrainedReference compound

-62

-61

-60

-59

-58

-57

-56

-55

-54

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

fre

e e

ne

rgy

(kc

al/

mo

l)

simulation time (ns)

B) Final forward free energy

V66E ∆+PHS; UnrestrainedReference compound

Figure 3-2. Free energy convergence in REFEP for unrestrained V66E ∆ + PHS (red)and Glu model compound (green). The cumulative free energy differencesare between the highest (fully deprotonated) and the lowest (fullyprotonated) replica For A) Backward and B) Forward free energy differences.

REFEP calculations of unrestrained V66E ∆ + PHS and Glu compound are converged.

We found that the GLU66 pKa in unrestrained is shifted about 3.2 pKa units higher

relative to the model compound, roughly 1.5 pKa units less than the experimental

value[92, 104]. Analyzing this data, it becomes obvious that changing the protonation

coordinate during REFEP, also induced a local conformational change around the loop

containing residue 66 (See Figure 3-1). Accordingly, translating the computed free

energy difference for the protein into a pKa is not correct.

51

Page 52: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

At this point, we present the most important result of this chapter: so far many

macroscopic and microscopic models have been proposed for estimation and calculation

of pKa in proteins [22, 111–115]. The microscopic approach in this chapter is grounded

on calculation of protonation/deprotonation energies at a given conformation. Based

on that, the proper treatment of these systems requires explicit inclusion of coupled

protonation and conformational equilibria[116]. Basically, at low pH, GLU66 is both

buried and protonated, while at high pH it is solvent exposed and deprotonated. To

understand this system we introduce a conformation-protonation equilibrium model,

shown in Figure 3-3, to calculate the pKa of GLU66 in the SNase mutant. Using this

model, we distinguished between conformational states of the GLU66 sidechain (i.e.,

e=exposed/b=buried) and its protonation states (i.e., P=protonated/D=deprotonated).

Figure 3-3. Thermodynamic cycle for conformation protonation model of GLU66. Db andPb stand for deprotonated-buried and protonated-buried, respectively; andDe and Pe stand for deprotonated-exposed and protonated-exposed,respectively.

The horizontal line at the top corresponds to the protonation change at the exposed

conformation, the horizontal line at the bottom to the protonation change at the buried

conformation, while the two vertical lines are the free energies required to change

52

Page 53: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

conformations at fixed protonation states. This separation allows us to study the

conformational and protonation changes separately and coupled, to better understand

the results.

To calculate the pKas in the buried and exposed cases individually, we performed

two sets of restrained REFEP simulations. In the buried case, the GLU66 side chain was

restrained to the interior of the protein, and in the exposed case, it was restrained to stay

exposed to the solvent by applying a harmonic restraint on all heavy atoms.

Restrained REFEP calculations (shown in Figure 3-4) compute a pKa of 10.6 for the

buried situation and of 4.6 for the exposed GLU66 sidechain, respectively.

52

53

54

55

56

57

58

59

60

61

62

0 0.5 1 1.5 2 2.5 3 3.5 4

free e

nerg

y(k

cal/m

ol)

simulation time (ns)

A) Final backward free energy

GLU66 side chain restrained-buriedGLU66 side chain restrained-exposed

-62

-61

-60

-59

-58

-57

-56

-55

-54

-53

-52

0 0.5 1 1.5 2 2.5 3 3.5 4

free e

nerg

y(k

cal/m

ol)

simulation time (ns)

B) Final forward free energy

GLU66 side chain restrained-buriedGLU66 side chain restrained-exposed

Figure 3-4. Cumulative Free energy VS Simulation time for Glu-66 restrained inside(Red) and restrained outside (Green).

53

Page 54: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

The thermodynamic cycle in Figure 3-3 is written in such a way that knowledge

of the two microscopic pKas and one of the equilibrium constant (or free energy) for

the conformational change, fully fixed the value of the second equilibrium constant.

Using mass balance equations, we have a system that has a single free parameter,

corresponding to one of the two vertical lines in Figure 3-3.

Since most experimental probes are sensitive to conformations, we will compute an

equilibrium constant that depends on either buried of exposed GLU66, disregarding

protonation states. The equilibrium constant between the exposed and buried

conformations of the GLU66 side chain can be written as:

Ke/b =[De] + [Pe]

[Db] + [Pb]=

[De]

[Db]

1 + 10(pKa,e−pH)

1 + 10(pKa,b−pH)= Ke/b,D

1 + 10(pKa,e−pH)

1 + 10(pKa,b−pH)(3–1)

Here, Ke/b,D is the Ke/b at high pH (i.e., when GLU66 is deprotonated).

Using the cycle presented in Figure 3-3, analytical expressions of the concentration

of each species as a function of the equilibrium constants and pH can be derived.

[De] =(

Ke/b

1+Ke/b

)(1

1+10(pKa,e−pH)

)[Pe] =

(Ke/b

1+Ke/b

)(10(pKa,e−pH)

1+10(pKa,e−pH)

)[Db] =

(1

1+Ke/b

)(1

1+10(pKa,b−pH)

)[Pb] =

(1

1+Ke/b

)(10

(pKa,b−pH)

1+10(pKa,b−pH)

). (3–2)

These expressions can be used to calculate the concentration of these species for

different values of Ke/b,D (Figure 3-5). For high values of Ke/b,D, the exposed-deprotonated

specie (De) is dominant at High pH value. The overall pKa can be defined in terms of a

protonation equilibrium constant, i.e.,

KD/P =[De] + [Db]

[Pe] + [Pb](3–3)

However, most experimental methods are sensitive to conformations and do not directly

report on the protonation state of a specific residue. Based on this, we can define a

54

Page 55: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

-0.2

0

0.2

0.4

0.6

0.8

1

1.2

2 4 6 8 10 12

co

ncen

trati

on

pH

pKa;e = 4.6

pKa;b = 10.6

Ke/b,D;1 = 2

Ke/b,D;2 = 1e+08

[De]1[Pe]1[Db]1[Pb]1[De]2[Pe]2[Db]2[Pb]2

Figure 3-5. Concentration of species vs. pH for two different value of Ke/b,D i.e.,Ke/b,D;1 = 2 and Ke/b,D;2 = 108.

Conformation-Sensitive Signal (CSS) for the case of GLU66 in V66E ∆ + PHS as:

CSS = [De] + [Pe]− [Db]− [Pb] (3–4)

At a very high pH, GLU66 is predominantly deprotonated and only species where

GLU66 is deprotonated will contribute significantly to the observed CSS. On the other

hand, at a very low pH, only the protonated species will be significant, i.e., CSSpHhigh= [De]− [Db]

CSSpHlow= [Pe]− [Pb]

. (3–5)

55

Page 56: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Figure 3-6 shows the CSS, as defined in Equation 3–4, as a function of pH. The pH

at which the inflection point is observed corresponds to the experimentally-measured,

apparent pKa. If the GLU66 is fully exposed at high pH and fully buried at low pH, then

the observed inflection point will correspond to a zero value of the CSS. However, if

there is a mixture of buried and exposed states, then the inflection point will be shifted

to a negative signal value. Based on Equation 3–1 and Equation 3–2, it is possible to

calculate the experimental inflection point as a function of pKa,e, pKa,b, and Ke/b,D.

Figure 3-6. CSS, as defined in Equation 3–4, as a function of pH for Ke/b,D=10; theorange arrows point to the inflection point ( i.e., half way between the highpH and the low pH values).

Figure 3-7 shows the apparent pKa plotted as a function of Ke/b,D. Figure 3-7

demonstrates that the pKa, which is measured using experimental methods that are

56

Page 57: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

sensitive to conformations, depends on the ratio of two distinct conformational states at

high pH. This model predicts that the apparent pKa ranges between 4.6 (Ke/b,D ≫1,

fully solvent exposed) and 10.6 (fully buried, Ke/b,D ≪1). Ke/b,D modulates the apparent

value of pKa. As Ke/b,D increases, the apparent pKa decreases. As Ke/b,D approaches

4

5

6

7

8

9

10

11

0.0001 0.01 1 100 10000 1e+06 1e+08

ap

pare

nt

pK

a

Ke/b,D

Figure 3-7. Apparent pKa vs. Ke/b,D.

infinity, the apparent pKa tends to 4.6.

We estimated Ke/b,D based on Root Mean Square Deviation (RMSD) distributions

of the GLU66 side chain and part of its belonged α−helix (residue62-69), for all the

replicas in the unrestrained REFEP simulations (Figure 3-8). In Figure 3-8, for λ=0

(GLU66 is protonated) we can identify two peaks at 0.4 and 1.1 A, with the former being

the highest. As λ increases, a new population around 1.6 A will appear and the others

will disappear. At λ=1.0 the peak at 0.4 A has disappeared, but there is still a small peak

at 1.1 A. Because of overlap between protonated and deprotonated RMSD distributions

57

Page 58: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

λ= 0.0

RMSD(A⋅ )

Fre

quen

cy

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0λ= 0.067

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.133

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.2

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.267

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.333

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.4

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.467

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.533

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.6

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.667

0.0 0.5 1.0 1.5 2.0 2.50

100

250

λ= 0.733

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.8

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.867

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 0.933

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

λ= 1.0

0.0 0.5 1.0 1.5 2.0 2.5

010

025

0

Figure 3-8. Carbon α RMSD distributions of residues 62 to 69 and the side chain ofGLU66 in unrestrained REFEP simulations for all sixteen replicas withrespect to the average structure.

(λ=0 and 1.0), we can conclude that both protonation states are ensembles of both

buried and exposed configurations. This shows that the changes in protonation states

are not fully synchronized with the changes in conformation states.

For more rigorous investigation of the decoupling, we calculated the CD spectra and

average secondary structure for all λ values in the unrestrained simulations. Average

secondary structure shows that as λ increases, the α−helix content decreases and

310 helix content increases (Figure 3-9). Figure 3-10 shows the ellipticity at 222 nm,

58

Page 59: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

10

20

30

40

50

60

70

80

0 0.2 0.4 0.6 0.8 1

perc

en

tag

e o

f th

e s

eco

nd

ary

str

uctu

re

λ

3-10-Helixα helix

turn

Figure 3-9. Average secondary structure for residue 57-69 vs. reaction coordinate.

which is the wavelength at which the finger print of α−helix appears in CD, for the whole

protein and the region containing Glu66 (i.e., residues 57-69). The α−helix content

decreases as the system shifts toward the deprotonated Glu66. The trend observed for

the whole protein and the local region is the same and is consistent with the fact that the

deprotonation of GLU66 locally denaturates part of α−helix content of that region. The

results of Figure 3-10 can be rationalized by looking at Figure 3-8. In the latter, RMSD

histograms of λ ∼ 0.2-0.4 show the maximum frequency on the leftmost part of the

RMSD axis (less than 1.3 A). These are the λs that have the maximum absolute value

of ellipticity in Figure 3-10. Estimating RMSD ∼ 1.0-1.3 A as the boundary between

an exposed and a fully buried GLU66 side chain, when λ=1.0 less than 1.8-6.7 % of

59

Page 60: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

-7.15E+04

-7.10E+04

-7.05E+04

-7.00E+04

-6.95E+04

-6.90E+04

-6.85E+04

-6.80E+04

-6.75E+04

0 0.2 0.4 0.6 0.8 1 -2.05E+04

-2.00E+04

-1.95E+04

-1.90E+04

-1.85E+04

-1.80E+04

-1.75E+04

-1.70E+04

-1.65E+04

-1.60E+04ellip

ticit

y

λ

whole proteinresidues 57-69

Figure 3-10. Ellipticity at 222 nm vs. reaction coordinate ( λ=0, i.e., GLU66 protonatedand λ=1, i.e., GLU66 deprotonated ) for whole protein(red, left y axis) andfor the region containing Glu66 (green, right y axis)

the population is in a fully buried conformation. Based on these results, Ke/b,D can be

estimated to be ∼ 15-57, which yields an apparent pKa of 8.7-9.2 (Figure 3-7). This

agrees well with the results of structure sensitive experimental methods, which estimate

the pKa of the GLU66 side chain to be 9.0-9.1 [92].

3.4 Concluding Remarks

There are many challenges in determining pKa shifts of buried ionizable groups.

The mutation of position 66 in staphylococcal nuclease is an interesting example for two

reasons. First, according to experimental methods the shift in pKa of ionizable group

mutants range from 3.5 -5.0, depending on the ionizable group and their relative normal

60

Page 61: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

pKa. Second, there is an inconsistency between pKa shifts calculated by implicit solvent

methods and those of experimental methods. To solve this challenging problem, we

used REFEP to calculate the pKa shift of staphylococcal nuclease mutant (V66E∆ +

PHS).

Not only the conformational response of a protein to a change in protonation state

of an ionizable group may be beyond the sensitivity of experimental methods, but also

conformational changes may not be fully correlated to the changes in protonation state.

The fact that protonation changes are not necessarily accompanied with configurational

changes shows the robust threshold of a protein against local electrostatics variations.

Based on this fact, the inconsistency between implicit solvent simulations and

conformational sensitive experimental methods has been explained. A four-state

thermodynamic cycle has been introduced to decipher this inconsistency. Through the

thermodynamic cycle, four species have been defined, which correspond to all possible

combinations of conformational-protonation of the GLU66 sidechain. We have shown

that the inflection point of experimental titration curves is dependent on the ratio of

conformational states at high pH; by calculating the ratio of conformational states at high

pH, the experimental inflection point that we have determined is in good agreement with

the one obtained from the experimental results.

61

Page 62: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

CHAPTER 4PH-REPLICA EXCHANGE MOLECULAR DYNAMICS IN PROTEINS USING A

DISCRETE PROTONATION METHOD

4.1 Literature Review

Solution pH plays a very important role in protein function, dynamics and structure.

Many important biological phenomena function only at a certain range of pHs, including

protein folding/misfolding[117–119], enzyme catalysis[120–122] and ligand-substrate

docking[123, 124]. In those cases, a pH change leads to a change in the ratio of

protonation states of different titratable residues, which is usually coupled to a change

in the conformation and dynamics of the protein itself. The pKa of a titratable residue

is the pH value at which the ratio of deprotonated to protonated concentrations of that

residue is equal to one[72, 112, 125, 126]. The pKa of an ionizable residue in a protein

is highly dependent on its electrostatic environment, which is coupled to the protonation

states of other titratable groups and the conformation state of the protein. Most current

simulations of pH dependent properties do not, in fact, change protonation states

during molecular dynamics, but rather pick a certain set of protonation states using

educated guesses, or guided by simplified algorithms, and then keep it constant for the

remainder of the simulation. These constant protonation MD methods suffer from two

big disadvantages[127]. First, at a pH near the pKa of any of the titratable residues, any

possible choice of constant protonation would clearly be wrong. Second, the choice

of protonation state is coupled for neighboring residues, which itself couples to the

conformational space of the system.

Like concentration and temperature, the solution pH is a very useful external and

controllable variable in experimental methods. Hence, the importance of constant

Reprinted with permission from Sabri Dashti, D.; Meng, Y.; Roitberg, A. E. J. Phys.Chem. B 2012, 116, 8805–8811.

62

Page 63: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

pH MD (CpH MD) has been recognized. In the last two decades, many constant pH

MD methods have been developed[128, 129]. The goal of CpH MD is to describe

correctly the protonation equilibrium coupled by conformation equilibrium at a certain

pH. The majority of CpH MD methods can be divided into continuous[128–136]

and discrete[127, 137–146] protonation state methods. Continuous protonation

methods use a continuous protonation parameter to perturb ionizable residue between

protonated and deprotonated states. In 1994, Mertz and Pettitt[135] developed a

grand canonical method for simulating a simple chemical reaction. They applied the

method for exchanging a proton between water molecules and an ionizable side chain.

In 1997, Baptista et al.[128] introduced a continuous constant pH method in implicit

solvent based on a mean-field approximation. In 2001, Borjesson et al. [129] used

a weakly coupled proton bath to continuously adjust the protonation fraction of each

titratable group towards equilibrium. More recently, the Brooks group has developed

the continuous protonation state method further[130–134, 147]. In the case of highly

coupled titration groups, for which cooperativity effects are non-negligible, this model

leads to inappropriate estimation of physical variables. To alleviate this problem, Lee et

al.[147] added a biasing potential, centered at λ equal 0.5, to help drive the protonation

coordinate value to fully protonated/deprotonated states, and away from the mid-λ

unphysical states.

In contrast to the continuous protonation state methods, discrete protonation

models define the protonation state of the ionizable group as either zero or one during

the simulation, corresponding to protonated and deprotonated states only. These

models use a hybrid MD-MC scheme; the MD is used to sample conformational space

for a number of steps, after which a Metropolis MC[14] attempt is done for changing

the protonation state/states. A new set of MD steps is then done with the protonation

state chosen by the MC step, and the process is repeated. Many versions of the

constant pH/discrete protonation MD method have been developed[115, 127, 138–

63

Page 64: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

143, 145, 146, 148]. The Baptista group[137–140, 148] used explicit solvent for the

propagation of coordinates and the Poisson-Boltzmann (PB) method for calculating

the energy in the MC section of the algorithm. Walczak et al. [146] employed Langevin

Dynamics for MD and the PB method for both of the MC and MD steps. Burgi et al.[141]

applied Thermodynamic Integration (TI) to calculate the transition energy between

the protonated and deprotonated species in explicit solvent MD. The drawback for this

approach is the large computational cost of TI calculations, which limits the amount of

sampling in protonation space. In 2004, Mongan et al. developed a discrete protonation

MD method by using the GB implicit solvent method to both the MD (structure) and MC

(protonation state) sampling sections[127]. A more detailed explanation of the method

can be found in the methods section. This method is implemented in the AMBER

Molecular Dynamics Package[75].

It has become clear in recent years that accurate modeling of protonation space

also requires enhanced sampling of conformational space[149–151]. Accurate

sampling of conformational space of proteins remains a challenging area [2, 3, 18,

152, 153, 155–157]. Many theoretical methods have been proposed to overcome

the free energy barriers in conformational space (see chapter 1). To account for the

coupling of protonation and conformational sampling, Wallace et al. [158] recently

have combined the continuous protonation Constant pH MD with the REMD method

(REX-CPHMD)[131]. They applied it to the problems of pKa prediction in protein

folding and pH dependent conformation, among others[158, 159]. Recently, Meng and

Roitberg[160] utilized a hybrid method by combining the Temperature REMD (TREMD)

and discrete protonation Constant pH MD. In this chapter, we introduce a method for

pH-Exchange MD by combining the discrete protonation Constant pH MD(proposed

by Mongan et al.[127]) and Hamiltonian REMD (HREMD) [22]. We tested our method

by applying it to five model dipeptides, to an uncapped pentapeptide with sequence

+H3N−Ala−Asp− Phe−Asp−Ala− COO− (ADFDA), and to an heptapeptide derived

64

Page 65: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

from the ovomucoid third domain OMTKY3 protein. Because the two ends of ADFDA

are not capped, its two Asp residues have slightly different electrostatic environments

and their pKas deviate in different directions from the pKa of unperturbed Asp. We will

show that the pHREMD MD improves the sampling efficiency in both protonation and

conformation spaces.

In the rest of this chapter, Constant pH MD (CpH) refers to Mongan’s CpH MD

approach unless mentioned otherwise.

4.2 Theoretical Method and Simulation Details

4.2.1 Constant pH Molecular Dynamics and pH-Replica Exchange MolecularDynamics (pH-REMD)

The goal of CpH MD is to sample the equilibrium between protonated and

deprotonated state of titratable sites at a given pH. The free energy difference between

protonated and deprotonated states determines the ratio of their concentrations. This

free energy difference can not be calculated by Molecular Mechanics (MM), because

the change in protonation state involves a bond breaking/forming phenomenon, which

requires a series of highly accurate quantum calculations. To address this issue, a

method[115, 127, 161], which uses a pre-calculated pKa of reference compounds, has

been developed. Reference/model compounds, in AMBER terminology, are represented

by a capped dipeptide for each titratable residue (i.e., ACE-titratable residue-NME). The

free energy for the protonation change to be used in the MC criteria is described by

Mongan et al. [161] as:

∆Gprotein = ∆Gprotein,MM + kBT(pH− pKa,ref) ln 10−∆Gref,MM. (4–1)

Here, as before, T is the temperature, and kB is the Boltzmann constant. ∆Gprotein,MM is

the molecular mechanics part of the free energy of the titratable site in the protein,

and ∆Gref,MM is the pre-computed deprotonation free energy for the reference

compound, described above. Using Equation 4–1, there is no need to calculating

65

Page 66: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

the QM contribution to the free energy in the protein. This method is implemented in

the AMBER MD suite[75], using the GB implicit solvent model. Every few MD steps

a Metropolis MC[14] attempt is done to change the protonation state of the titratable

residue and ∆Gprotein is used to make a decision about accepting or rejecting the

proposed MC move. In other words, the MC moves sample protonation space, and the

MD steps sample configuration space. During the MD steps, the protonation state is

kept constant.

4.2.2 Titration Curve

The pKa of a titratable residue is related to the pH environment through the

Henderson-Hasselbalch (HH) equation,

pKa = pH− n log

([A−]

[HA]

), (4–2)

with [A−] and [HA] being the deprotonated and protonated concentrations respectively; n

is the Hill coefficient, which should approach 1 for non-interacting ionizable residues, but

deviate from one in the case of interacting titratable residues because of cooperativity[162].

Because of the ergodicity assumption underlying MD, the ratio of the time that a

titratable residue spend at deprotonated state to the time spend in protonated state can

be considered as a ratio of concentration of deprotonated state to that of protonated

state.

On the other hand, pH-REMD is a combination of CpH[161] and HREMD[22]. In

contrast to temperature replica exchange[2], in which each replica runs at a different

temperature, in HREMD, each replica runs in a distinct Hamiltonian but at the same

temperature. In pH-REMD each replica runs a constant pH MD at a unique pH, and

periodically an exchange of conformation between two adjacent replicas is attempted

(section 4.2.3).

66

Page 67: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

We note that a recent publication[23] presented a pH replica exchange method

using a different exchange criterion, that does not require recomputing energies for

different replicas. In the future, we will compare the two formulations.

4.2.3 pH-REMD

In the HREMD algorithm, each replica runs a distinct Hamiltonian. In pH-REMD

each replica runs a CpH MD simulation at a unique pH, with a swap of conformations

between two adjacent replicas is attempted regularly. Using the detailed balance

condition, we can write an equilibrium proposition for the ensemble before and after an

exchange is attempted,

w(X→ X)℘(X) = w(X→ X)℘(X), (4–3)

where ℘(X) is the probability of being at state X and w(X → X) is the probability of

transition from state X to that of X.

At the exchange moment, if we only swap the conformations between two adjacent

replicas and keep the protonation state unchanged, then the generalized states of X and

X can be written as

X =

n1 ... ni nj ... nM

q1 ... qi qj ... qM

, X =

n1 ... nj ni ... nM

q1 ... qi qj ... qM

, (4–4)

Here q represents a conformation and n represents protonation states. The ith column is

related to the ith replica.

Because all M replicas are independent, the probability of the system being in the

generalized state X can be written as ℘(X) =∏M

i=1 Pi(qi, ni), where Pi(qi, ni) is the

probability of replica i at conformation qi and protonation state of ni. Substituting those

probabilities in Equation 4–3 will result

w(X→ X

)w(X→ X

) =P(qj, ni)P(qi, nj)

P(qi, ni)P(qj, nj)= exp(−β∆), (4–5)

67

Page 68: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

In the canonical ensemble regime (NVT), and using the Boltzmann distribution as a

limiting ensemble, this can be written as

w(X→ X

)w(X→ X

) =exp(−βH(qi, nj)) exp(−βH(qj, ni))

exp(−βH(qi, ni)) exp(−βH(qj, nj))= exp(−∆), (4–6)

where β = 1kBT

and ∆ = [H(qi, nj +H(qj, ni)− H(qi, ni)− H(qj, nj)].

This setup can be realized by setting the exchange probability according to the

Metropolis MC criteria as

w(X→ X

)= w ((qi, ni); (qj, nj)→ (qj, ni); (qi, nj)) = min (1, exp(−∆)) . (4–7)

For computing ∆ only the potential energies are required. Since the two exchanging

replicas are running at the same temperature and have the same number and mass of

particles, their kinetic energy terms will cancel.

4.3 Simulation Details

To validate and test the method presented here, we chose and ran simulations

on three categories of systems. First, we studied the capped reference compounds

(described in the methods section), consisting of the ACE − titratable residue − NME.

Simulation times were 3 ns for both CpH MD and pH-REMD (for each replica) methods.

We used eight pH-replicas for all the model compounds in pH-REMD. We also tested

our simulation method on a terminally charged pentapeptide model, +H3N − Ala-

− Asp − Phe − Asp − Ala − COO− (ADFDA). Because the two ends of ADFDA are

oppositely charged, the two Asp experience different electrostatic environments and

have different pKas. The simulation times of ADFDA were 90 ns for CpH and 10 ns

for each replica in pH-REMD. Eight replicas, at pH=2.5-6.0 with increment of 0.5 were

used in pH-REMD method. The third system was a heptapeptide derived from OMTKY3

(ACE − Ser − Asp − Asn − Lys − Thr − Tyr − Gly − NME). We simulated 100 ns for

CpH and 10 ns for pH-REMD. Twelve replicas, at pH=2-13 with increment of 1 pH unit,

were used in pH-REMD method. All calculations in this chapter were done using the

68

Page 69: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

AMBER 10 molecular simulation suite. The AMBER ff99SB force field[76] and OBC

Generalized Born implicit solvent[77] (igb=2) were used in all simulations. The Shake

algorithm [101] was applied to constrain the bonds between heavy and hydrogen atoms,

which allowed the use 2 fs MD steps. A cutoff of 30 A for non-bonded interactions was

chosen in all calculations. In all pH-REMD simulations, except the heptapeptide derived

from OMTKY3, replicas attempted a pH exchange every 1000 MD steps. In the case of

the heptapeptide, replicas attempted a pH exchange every 500 MD step to accelerate

sampling[163]. The exchange acceptance ratios between replicas for all system were

between 0.2 and 1.0. For calculating the error bars and the uncertainty of pKas, every

4000 MC steps (in protonation space), the deprotonation fraction has been calculated.

4.4 Results and Discussion

4.4.1 Titratable Model Compounds

When we apply pH-REMD to the model compounds that were initially used

to parametrize the method, it is not surprising to find that it produces the correct

pKas. It is, however, an important calculation to perform to check the method and to

gauge its efficiency versus constant pH runs. The results are shown in Table 4-1,

where all pKa values have been calculated by a fit to the linearized version of the

Henderson-Hasselbalch Equation.

Table 4-1. pKas of the reference compounds computed by different methodsASP GLU HIS TYR LYS

Experimental value 4.0 4.4 6.3 9.6 10.5CpH MD 3.9±0.1 4.4 ±0.1 6.3±0.1 9.6±0.03 10.4±0.03pH-REMD 3.9±0.1 4.4 ±0.1 6.4±0.1 9.7±0.02 10.4 ±0.04

In Figure 4-1, the Hill plot for capped Lysine is shown for both pH-REMD and CpH

MD methods, with agreement between them over a large pH range.

4.4.2 ADFDA Model Compounds

We now apply the method to the model peptide, +H3N − Ala − Asp − Phe − Asp-

− Ala− COO− (ADFDA), that has charged ends. This system has an intrinsic but subtle

69

Page 70: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

-2

-1.5

-1

-0.5

0

0.5

1

1.5

9 9.5 10 10.5 11 11.5

log 1

0 ([

A- ]/

[AH

])

pH

pkaCpH= 10.4

pkapH-REMD= 10.4

HH fit CpH

HH fit pH-REMD

Figure 4-1. Comparison of Hill plots between pH-REMD and CpH methods for Lysreference model.

asymmetry in the electrostatic environment for the two Asp ionizable side chains. Asp2

is close to the NH+3 , which causes its pKa to be shifted slightly below 4.0. Asp4 is close

to the COO− terminal, which causes its pKa to be shifted above 4.0.

The plotted titration curves of Asp2 and Asp4 in Figure 4-2 have Asp2 shifted to the

left of the model compound, and Asp4 shifted to the right. There is a good agreement

between the pH-REMD and CpH curves. The predicted pKas and Hill coefficients for

both methods have been calculated by linear fitting on a Hill plot.

Table 4-2. pKa prediction and Hill coefficient of fitted from the HH equation.ASP2 ASP4

pH Hill coefficient pH Hill coefficientpH-REMD 3.8±0.1 0.89 4.5±0.1 0.88CpH MD 3.8±0.1 0.91 4.5±0.1 0.86

70

Page 71: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

0.2

0.4

0.6

0.8

1

2.5 3 3.5 4 4.5 5 5.5 6

Der

oton

atio

n F

ract

ion

pH

Asp2; CpH

Asp4; CpH

Asp2; pH-REMD

Asp4; pH-REMD

Asp; reference

Figure 4-2. The titration curves of Asp side chains in ADFDA computed by both constantpH MD (blue and purple) and pH-REMD (red and green) methods; Thetitration curve of Asp for the reference compound (light blue) is also shown.

Table 4-2 shows that the results of both methods are identical. It is worth noting that

the free energy difference associated with the two different pKas is 0.96 kcal/mol, which

highlights the sensitivity of the method to very small environmental changes.

To gauge the comparative efficiency of CpH and pH-REMD, in Figure 4-3, the

cumulative average protonation fractions of Asp2 and Asp4 for both methods at pH=4.0

have been plotted. The data clearly shows that for both titratable groups, pH-REMD

converges more rapidly in protonation space than the regular CpH method.

4.4.3 Heptapeptide Derived from OMTKY3

We applied the pH-REMD method to a capped heptapeptide derived from OMTKY3

(ACE−Ser−Asp−Asn−Lys−Thr−Tyr−Gly−NME). Dlugosz and Antosiewicz[142, 143]

studied this heptapeptide and predicted the pKa of 4.24 for Asp3 using their CpH MD

71

Page 72: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 100000 200000 300000 400000 500000 600000

Pro

tona

tion

Fra

ctio

n

MC Titration Steps

Asp2; pH-REMD

Asp4; pH-REMD

Asp2; CpH

Asp4; CpH

Figure 4-3. Cumulative average protonation fractions of Asps side chains in ADFDA VSMC titration steps at pH=4.0.

method. According to NMR experiments[144, 145], the pKa of Asp in this heptapeptide

is about 3.6.

Table 4-3. pKa values of the titratable residues in the heptapeptide derived fromOMKTY3.

ASP3 Lys5 Tyr7pH-REMD 3.6±0.2 10.6±0.1 10.1±0.1CpH MD 3.7±0.2 10.6±0.1 9.9±0.1

From a Hill plot, the pKa of the three titratable groups has been calculated with the

results presented in table 4-3. Our computed for values the Asp3 pKa of 3.7 (CpH) and

3.6 (pH-REMD) are in excellent agreement with the experimental value. The titration

curves are shown in Figure 4-4.

There are two more titratable groups in the heptapeptide, i.e., Lys5 and Tyr7, which

we also titrated. Figure ?? presents the titration curves of Lys5 and Tyr7, with the

72

Page 73: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

0.2

0.4

0.6

0.8

1

2 3 4 5 6 7

Dep

roto

natio

n F

ract

ion

pH

Asp3; CpH Asp3; pH-REMD

Figure 4-4. The titration curves of Asp3 in the heptapeptide derived from OMTKY3.

computed pKas listed in Table 4-3. To compare the convergence speed between CpH

and pH-REMD, we studied the cumulative average protonation/deprotonation fraction

as function of MC titration steps for all ionizable residues. Figure 4-5 shows the data for

Tyr7. It is evident that pH-REMD converges faster (and smoother) to the final protonation

fraction.

While the convergence of protonation equilibrium is crucial for the proper computation

of a pKa, the convergence of structural properties is also important. To consider

this issue, we calculated the RMSD of α-Carbons for all pHs (for both CpH and pH-

-REMD simulations) with respect to the average structure from a CpH simulation at

pH 10. For both CpH and pH-REMD methods, the conformational convergence was

studied by calculation of the Kullback-Leibler divergence (see section 1.5.3.5) of RMSD

cumulative distributions vs. time. This is a measure of the rate of convergence to the

final conformational ensemble. We presented results for both CpH and pH-REMD at pH

10 for every 100 ps (Figure 4-7). We used the final RSMD distribution of CpH simulation

at pH 10 as a reference for the plot. As the inset of Figure 4-7 shows, both simulations

73

Page 74: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

0.2

0.4

0.6

0.8

1

7 8 9 10 11 12 13

Dep

roto

natio

n F

ract

ion

pH

Lys5; CpH Lys5; pH-REMD

Tyr7; CpHTyr7; pH-REMD

Figure 4-5. The titration curves of Lys5 and Tyr7 in the heptapeptide derived fromOMTKY3.

converge to the same RMDS distribution. It is clear that pH-REMD converges more

rapidly and smoothly than CpH.

We also compared the rate of visiting distinct structures by calculating the RMSD

autocorrelation at all pH for both methods. According to Figure 4-8, the correlation

time of RMSDs in pH-REMD simulations are significantly shorter than that of CpH

simulations, which implies that pH-REMD visits distinct conformations more often than

CpH.

4.5 Concluding Remarks

In the present chapter, we have combined Hamiltonian REMD with Constant pH

MD to create what we call pH-REMD. The predicted pKa for a number of systems

using pH-REMD is in excellent agreement with experimental data. Compared to CpH

MD, pH-REMD converges faster in both conformational and protonation spaces. In

contrast to the Temperature Replica Exchange Molecular Dynamics methods (TREMD),

in pH-REMD the replica ladder (pH) is very limited, so the overlap of energy and

74

Page 75: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 10000 20000 30000 40000 50000 60000

Dep

roto

natio

n F

ract

ion

MC titration step

(A)

pH 8.0pH 9.0

pH 10.0pH 11.0pH 12.0pH 13.0

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 10000 20000 30000 40000 50000 60000

Dep

roto

natio

n F

ract

ion

MC titration step

(B)

pH 8.0pH 9.0

pH 10.0pH 11.0pH 12.0pH 13.0

Figure 4-6. Cumulative average protonation fraction for TYR7 versus MC titration stepsat pH=8.0,9.0,10.0,11.0,12.0 and 13.0, comparison between the CpH (A)and the pH-REMD (B) methods.

75

Page 76: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

10

20

30

40

50

60

70

0 2 4 6 8 10

Kul

lbac

k-Le

ible

r di

verg

ence

t (Simulation Time (ns))

CpH; pH=10pH-REMD; pH=10

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

3 3.5 4 4.5 5 5.5 6 6.5 7

RMSD (A)

Figure 4-7. The Kullback-Leibler divergence measure of RMSD distributions of CpH(Green) and pH-REMD (Red) with respect to the final RMSD distribution inCpH. The inset shows the RMSD distributions of CpH (Green) andpH-REMD (Red) respectively after 100 and 10 ns.

consequently the exchange ratio between neighboring replicas are always high.

This new method is expected to perform very well for biosystems with highly coupled

conformational and protonation states, like proteins with pH-dependent structure and

dynamics.

76

Page 77: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

0.2

0.4

0.6

0.8

1

0 0.05 0.1 0.15 0.2

Aut

ocor

rela

tion

Lag (ns)

pH 2.0 pH 3.0 pH 4.0 pH 5.0

pH 6.0 pH 7.0 pH 8.0 pH 9.0

pH 10.0 pH 11.0 pH 12.0 pH 13.0

Figure 4-8. RMSD Autocorrelation for CpH (Green) and pH-REMD (Red) at all pHs.

77

Page 78: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

CHAPTER 5OPTIMIZATION OF UMBRELLA SAMPLING REPLICA EXCHANGE MOLECULAR

DYNAMICS BY REPLICA POSITIONING

5.1 Literature Review

Among the advanced sampling methods (see section 1.4), the Umbrella Sampling

method (US)[25, 49, 170] is well known for improving the sampling of rare events along

the reaction coordinate. US attempts to solve the bad sampling problem by applying a

biasing potential to the system, to guarantee the effective sampling along the reaction

coordinates. This can be achieved either in one or several simulations (windows).

While the efficiency of US is clear, it needs careful tuning. For example, the positions of

windows and the strength of the bias potentials have to be such that they have enough

overlap.

REMD techniques [1–3, 22, 24] have become increasingly popular schemes. REMD

is not only a promising method for tackling the quasi-ergodicity[171] problem, but it

is also able to combine intuitively with other enhanced sampling techniques[2, 4, 21,

23, 24, 55, 168]. In REMD methods, N non-interacting copies (replicas) of the system

with different temperatures (Temperature REMD[2–4]) or different potential energy

parameters (Hamiltonian REMD[3, 22, 24, 56, 172]) run concurrently and each replica

attempts to exchange with its neighbors every few steps. In other words, the system

performs a random walk in the replicas’ ladder space.

Three types of efforts have been made to optimize REMD with most of those

being only applicable to TREMD. First, some techniques try to keep the number

of replicas limited. Second, there are methods that try to increase the efficiency of

REMD by increasing the probability of accepting an exchange. Third, are methods

that increase the efficiency by optimal selection of the positions of replicas on a

Temperature/Hamiltonian ladder. In TREMD, the number of needed replicas in a system

increases with the number degrees of freedom[56, 58, 173, 174]. This scaling confines

TREMD to small systems. A clear way to avoid this problem is to decrease the number

78

Page 79: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

of degrees of freedom[175, 176]. Approaches comprise the use of hybrid explicit/implicit

solvent models for the MD and exchange parts respectively[173], the use of separate

baths for solvent and solute[174], the use of a coarse grained model for a subset of the

system[58] etc. Also, some studies have tried to decrease the number of replicas by

coupling the system to a pre-sampled reservoir[177, 178].

A few efforts have been made to increase the probability of exchange. Rick[179]

used dynamically scaled replicas between conventional replicas at broad temperature

ranges. Li et al. [180] devised a new approach called Temperature Intervals with Global

Energy Reassignment (TIGER). In their method, short runs of heating-sampling-

-quenching sequences are done and at the end of each sequence the potential energies

of all replicas are compared using Metropolis MC sampling and finally reassigned to

the temperature levels. Kamberaj et al.[181] added appropriate bias potentials to the

system to flatten the Potential of Mean Force and to achieve a random walk along the

temperature ladder. Also, Ballard et al. [168] used non-equilibrium works simulations to

generate the attempted configuration swaps.

For the optimal choice of the ladder of replicas, it has been argued that the highest

efficiency is achieved when the Exchange Acceptance Ratio (EAR) between neighbor

replicas is about 20%.[182–185]. In 2002, Kofke[182, 183] showed that when the

specific heat of the system is independent of temperature, assigning temperatures in

a geometric progression leads to equal EAR for TREMD. But, if the specific heat is a

function of temperature, then the choice of temperatures is not system-independent.

Although setting the equi-EAR between the neighbors is the most accepted

criterion for optimizing REMD (especially in TREMD); it has been shown [186–

190] that for systems with phase transitions along replica ladders, this is not the

optimal arrangement. A few techniques have been proposed for selecting the set of

temperatures when the specific heat changes with temperature. For example, Rathore

et al. [187] used a small set of pre-simulated replicas at different temperatures to

79

Page 80: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

determine the mean and variance of energy at those temperatures and then applied

an iterative scheme to find the right positions of replicas (temperatures). Katzgraber

et al. [188] also devised an iterative feedback algorithm for optimizing TREMD, but

their method is computationally demanding. In 2007, Hirtz and Oostenbrink[189]

proposed a technique for optimizing TREMD and HREMD in systems with known

stable conformations. In their method, one first generates many short MD trajectories

using different Hamiltonians/temperatures and starting from different regions of

conformation space. Then one clusters the frames based on their conformational

states and Hamiltonians/temperatures at which those were simulated and finally applies

the transition probability between the pairs in different groups. Shenfeld et al. [191]

minimized the thermodynamic length [192] in order to find the optimal positions of

replicas.

The majority of proposed optimization methods for REMD are applicable only to

TREMD, with those that are applicable to HREMD methods being computationally

demanding. In this chapter we propose a method for optimizing Umbrella Sampling

Replica Exchange. We use three unique properties of USRE for optimizing the position

of replicas.

First, in the case of USRE, there is no phase transition along the reaction

coordinate, because the bias potentials constrain the sampled coordinate locally. As

described above, the optimal replica sets in the absence of phase transitions are those

that have equi-EAR between all neighbor replicas. Second, the exchange probability in

USRE is only a function of the difference between order parameters at the moment of

exchange. This Exchange Acceptance Ratio (EAR) is very easy to compute. Third, in

USRE we can safely approximate the order parameter distribution in each window as a

Gaussian distribution[27, 28]. In this chapter, we do not discuss or propose any optimal

value of EAR between the neighbors, but we provide a way to optimizes the efficiency of

USRE given either the number of replicas or the desired EAR between all neighbors.

80

Page 81: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

To test our method we applied it to optimize USRE simulations to sampling butane

dihedral rotation in implicit solvent and to compute the free energy of formation of a

NH+4 + OH− salt bridge in explicit solvent. The proposed positions of replicas (windows)

by our method not only resulted in better mixing and convergence with respect to that of

equal distance on the reaction coordinate but also the best possible arrangement with

given range and number of replicas. We show that any deviation from this arrangement

produces a lower rate of mixing. In the rest of the chapter we use replica and windows

interchangeably.

5.2 Theory

5.2.1 Umbrella Sampling Replica Exchange

USRE[4] is a combination of Umbrella Sampling and Hamiltonian Replica Exchange

MD (see chapter 1), i.e., in USRE N non-interacting replicas run simultaneously with

N different bias potentials. The bias potentials are harmonic functions with distinct

strengths and coordinate references.

The bias of the ith window has a quadratic form of

Bi(ξ) = ki(ξ(q)− ξ0i )2, (5–1)

where ki, ξ(q), ξ0i , and q are the bias strength, the order parameter, the bias center

on the reaction coordinate, and configuration respectively. We will present results

only for the case for which all our replicas have the same force constant but different

coordinate centers. However the method is easily expandable to the case of dissimilar

bias strengths.

The Hamiltonian for the ith window can be written as:

Hi = K+UUB + Bi, (5–2)

where UUB and K are respectively the unbiased potential energy and the kinetic energy.

For two exchanging replicas, i and j, with ki = kj = k , ∆ in Equation 1–22 can be

81

Page 82: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

simplified to:

∆ = β[k(ξi − ξ0j )

2 + k(ξj − ξ0i )2]−

[k(ξi − ξ0i )

2 + k(ξj − ξ0j )2]

= 2kβ(ξ0i − ξ0j )(ξi − ξj) = 2k∆ξ0ij∆ξij.

(5–3)

For a given choice of bias centers for each window, ∆ξ0ij is a constant and then ∆

becomes a linear function of ∆ξij and is not dependent on the potential energies of

exchanging replicas.

5.2.2 Calculating Exchange Acceptance Ratio in Umbrella Sampling ReplicaExchange

By inserting Equation 5–3 into Equation 1–20, we can write the probability of

accepting an exchange as:

Pacc = min(1, exp(−2βk ∆ξ0ij ∆ξij)

). (5–4)

Using the Probability Density Function (PDF) of ∆ξij (i.e., p(∆ξij), the EARij (or the

average probability of acceptance between replicas i and j) can be calculated as:

EARij = ⟨Pacc⟩p(∆ξij)=

∫Pacc(∆ξij)p(∆ξij)d∆ξij. (5–5)

We can estimate the p(∆ξij)as a normal Distribution (see appendix A for details)

p(∆ξij) =1√

2π(σ2ξi+ σ2

ξj

) exp

−(∆ξij − (⟨ξi⟩ − ⟨ξj⟩))2

2(σ2ξi+ σ2

ξj

) . (5–6)

here, ⟨ξi⟩ and σ2ξi

are the mean and the variance of ξi distribution. Note that ⟨ξi⟩ = ξ0i .

By substituting the Equation 5–4 and Equation 5–6 in the Equation 5–5, the average

probability of acceptance becomes

EARij =

∫ ∞

−∞min

(1, exp(−2βk ∆ξ0ij ∆ξij)

) 1√2π

(σ2ξi+ σ2

ξj

) exp

−(∆ξij − (⟨ξi⟩ − ⟨ξj⟩))2

2(σ2ξi+ σ2

ξj

) d∆ξij.

(5–7)

82

Page 83: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Without loss of generality, we can simplify the min by assuming ξ0i < ξ0j , and finally

EARij =1

2erfc

− (⟨ξi⟩ − ⟨ξj⟩)√2(σ2ξi+ σ2

ξj

)

+1

2erfc

(⟨ξi⟩ − ⟨ξj⟩)− 2βk(∆ξ0ij)(σ2ξi+ σ2

ξj

)√

2(σ2ξi+ σ2

ξj

)

exp[−2βk(∆ξ0ij)(⟨ξi⟩ − ⟨ξj⟩) + 2β2k2

(σ2ξi+ σ2

ξj

)(∆ξ0ij)

2].

(5–8)

In Equation 5–8, all other parameters except the ⟨ξi|j⟩ and σ2ξi|j

are known. In appendix

B we present a method for estimating the mean and variance of ξ at any point of the

reaction coordinate using very few short pre-simulated windows.

5.2.3 Umbrella Sampling Replica Exchange Optimization

As we discussed above, there is no phase transition along the replicas ladder in the

case of USRE. If there is no phase transition, the best solution for positions of replicas

is for the EAR between all neighbor pairs is equal. We present here two different

approaches for making a set of equi-EARs among neighbor pairs. We discuss fitness

functions for each case, that, when maximized, lead to equal EAR among neighbor

replicas.

In the first method, the number of replicas and the center of the first and last

replicas are fixed (the range for the sampling coordinate). In order to find the set of

equi-EAR replicas, we define a Full Batch Scoring Function (FBSF) as:

FBSF(ξ02 ... ξ

0N−1

)= −

N∑i=2

(EARi−1i(ξ

0i−1 − ξ0i )− ⟨EAR⟩

)2 (5–9)

and maximize it with respect to the center of biases in all replicas except the last

and first ones (i.e., ξ02 ... ξN−12 ). Here N is the number of replicas and ⟨EAR⟩ =

N∑i=2

EARi−1i(ξ0i−1−ξ0i )

N−1.

83

Page 84: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

In FBSF we set the position of the first and the last replicas. Consequently the total

number of degrees of freedom is equal to N − 2. Since EARi−1i is a strictly decreasing

function of ξ0i−1 − ξ0i , then it can be concluded that the scoring function (which is an N− 2

dimensional parabola) has a unique maximum. Maximizing the FBSF leads to equal

EAR between adjacent replicas.

In the second method, we choose a desired value for the EAR between replicas,

and we fix the center of bias for the first replica (lowest value on reaction coordinate)

and the range of the reaction coordinate. In this scheme, starting from the lowest value

on the reaction coordinate, we add a new replica at a time and maximize the Single

Batch Scoring Function (SBSF) respect to that center of bias in that replica. We continue

this scheme until the last replica passes the maximum range of the reaction coordinate.

In order to adjust the position ith replica, we try to maximize the following SBSF at every

replica accumulation:

SBSF(ξ0i)= −

(EARi−1i(ξ

0i−1 − ξ0i )− EARc

)2 (5–10)

where EARc is the desired value of EAR, which is bounded between 0 and 1. At the end

of this process the EAR between all neighbor pairs equals EARc. We used the Barzilai

and Borwein[194, 195] optimization method for maximizing both SBSF and FBSF (see

the Appendix C for more details).

5.2.4 Umbrella Sampling Replica Exchange Optimization Workflows

Based on the choice of optimization, two types of workflows exist. In both methods

we pre-simulate M windows. In the FBSF method, the number of replicas is set, but the

final average EAR is not known a priori (Figure 5-1). In the SBSF method the EARc is a

chosen parameter and the number of replicas in not known in advance (Figure 5-2).

We have written scripts (for both FBSF and SBSF) that implement the ideas

presented here. Both scripts are available at http://www.clas.ufl.edu/users/

roitberg/software.html.

84

Page 85: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

• pre-simulated data from M windows• convergence criterion (ǫ)• N (Number of windows)• range of reaction coordinate

calculate the meansand variancesof M windows

initilize N equi-distance center ofbiases along the

reaction coordinate

estimate the meanand varianceof N windows

estimate the EARsof N windows

|FBSF |<ǫ

stop

move all replicas alongthe FBSF gradient

yes

no

Figure 5-1. FBSF workflow

85

Page 86: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

• pre-simulated data from M windows• convergence criterion (ǫ)• Exchange Acceptance Ratio (EARc)• range of reaction coordiante

estimate the meansand variancesof M windows

adding the first replica

adding thenext replica

estimating the EARwith previous replica

|SBSF |<ǫ

is thepositionof the lastreplica isin thegivenrange?

stop

move the lastreplica along theSBSF gradient

no

yes

no

yes

Figure 5-2. SBSF workflow

5.2.5 Simulation Details

In order to validate our method, we performed USRE simulations on two systems.

First, we studied the PMF along the butane torsional angle in the range from -65 to

67 degrees using 12 replicas. We performed four sets of simulations for different

arrangements of replicas on the reaction coordinate. We applied the OBC Generalized

Born[101] implicit solvent model (igb=5) with a cutoff of 12 A for nonbonded interactions.

A spring constant of k = 30kcalmol−1radian−2 has been applied to all replicas.

The length of each simulation was about 50 ns in order to have enough sampling

86

Page 87: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

for estimating the Average Roundtrip Time between the highest and the lowest

replicas in each set. Second, we performed two sets of simulations using 24 replicas

on a NH+4 + OH− salt bridge immersed in a box of TIP3P solvent model[196] of

21.4×21.4×21.4 A3. We calculated the long range electrostatic interactions with the

particle mesh Ewald method, and used an 8A cutoff for the short range nonbonded

interactions. A spring constant of k = 30kcalmol−1A−2 has been applied to all replicas.

The lengths of all simulations were 40 ps, however we used a 5 ns USRE simulation as a

reference in calculation of the Kullback-Leibler divergence[39].

In all calculations in this chapter, we used the AMBER 12 molecular simulation

suite[197] and the general AMBER force field (GAFF)[198]. The SHAKE[77] algorithm

was employed to constrain the distance between hydrogens and heavy atoms, which

permitted the use of 2 fs MD steps. Replica were attempted to exchange every 50 MD

steps[199] and Langevin dynamics with friction coefficient of 2.0 ps−1 was employed

to sustain the systems at 300K. In order to produce independent simulations[107],

we used distinct random seeds (ig=-1 in AMBER) for the Langevin thermostats in all

simulations. We saved the order parameter every 10 MD steps (i.e., 20 fs) and we used

Grossfields[26] implementation of Weighted Histogram Analysis method (WHAM) for

calculating the Potential of Mean Forces versus reaction coordinate in all simulations.

5.3 Results and Discussion

To study the effect of replica optimization in USRE, we computed and compared the

Average Roundtrip Time for different positions of the replicas, including our optimized

one. Furthermore we measured the convergence speed for optimized and non-

-optimized USRE arrangement of NH+4 + OH− salt bridge in explicit solvent. In both

simulations, we choose the total number of replicas to be equal to that of pre-simulated

replicas (i.e., M = N).

87

Page 88: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

5.3.1 Potential of Mean Force Along the Butane Dihedral Angle

We performed four sets of USRE calculations on butane dihedral in implicit solvent.

In the first one, the centers of biases were equally distributed (every 12) on the reaction

coordinate. In the second set of calculations we moved two neighboring replicas in the

middle of the reaction coordinate and we kept all other replicas in place. For the third

set we optimized the replica positions by maximizing the FBSF. For the optimization,

we used the first 50 ps of the first set as pre-simulated data. Finally, in the fourth set of

simulations we perturbed the position of one replica in the third (optimized) set. The

positions of replicas and corresponding EAR in each set has been shown in Table 5-1.

Table 5-1. Position and EAR for four different settings of USRE calculations on thebutane dihedral simulation in implicit water. Position of each replicacorresponds to the center of bias for that replica. Also, the EAR of eachposition corresponds to the EAR between that replica position and its higherneighbor on the reaction coordinate.

Set 1 Set 2 Set 3 Set 4Position EAR Position EAR Position EAR Position EAR

-65.00 0.18 -65.00 0.19 -65.00 0.14 -65.00 0.14-53.00 0.18 -53.00 0.18 -51.72 0.14 -51.72 0.14-41.00 0.16 -41.00 0.16 -38.58 0.14 -38.58 0.14-29.00 0.13 -29.00 0.13 -26.01 0.14 -26.01 0.14-17.00 0.10 -17.00 0.13 -14.44 0.15 -14.44 0.15

-5.00 0.07 -6.00 0.03 -4.16 0.14 -4.16 0.107.00 0.10 9.00 0.18 5.73 0.15 7.00 0.21

19.00 0.14 19.00 0.14 16.10 0.14 16.10 0.1331.00 0.16 31.00 0.17 27.98 0.14 27.98 0.1443.00 0.18 43.00 0.18 40.60 0.14 40.60 0.1455.00 0.19 55.00 0.19 53.71 0.14 53.71 0.1467.00 0.00 67.00 0.00 67.00 0.00 67.00 0.00

For each set, the Average Roundtrip Time (ART) between extremum replicas was

computed. To gather sufficient statistics, each simulation was repeated 10 times. Since

there were 12 replicas in each simulation, the total number of measured ARTs in each

set was 120. We present below the probability density of the ARTs (Figure 5-3). Figure

5-3 can be analyzed based on the Central Limit Theorem. The CLT describes the

characteristics of the population of the means of an infinite number of random samples

88

Page 89: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

800 900 1000 1100 1200 1300 1400

0.000

0.005

0.010

0.015

ART(Exchange)

De

nsi

ty

set 1

set 2

set 3

set 4

−60 −40 −20 0 20 40 60

01

23

4

ξ

PM

F(k

cal/

mo

l)

Figure 5-3. ART probability density. The inset shows the PMF vs. reaction coordinate(dihedral angle)

of size N, drawn from an infinite parent population. According to the CLT, the distribution

of the mean converges to a Gaussian distribution, and the mean of the means (i.e.,

center of the Gaussian) is equal to the mean of the parent distribution.

Based on the CLT the centers of distributions are equal to the mean of the parent

pools. According to Figure 5-3, the Optimized set (set 3) has the lowest ART among

the all sets, moreover as the replica positions deviate more from optimized arrangement

(i.e., sets 4, 1 and 2 respectively) the ART increases. This shows the equi-EAR set

corresponds to the minimum ART and the mixing is the best in this arrangement.

Since the number of exchange attempts is equal in all sets, we expect the sets with

higher ARTs to have the smaller sample sizes (i.e., the number of roundtrips in each

simulation), which according to the CLT leads to wider ART distributions. Moreover,

Figure 5-3 also shows the naive arrangement of equal distance between the centers of

biases (set 1) is not the optimized arrangement of replicas.

89

Page 90: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

5.3.2 Potential of Mean Force for NH+4 +OH− Salt Bridge in Explicit Solvent

In order to show the effect of USRE optimization on the rate of the convergence for

free energy calculations, we applied it to a more computationally demanding system,

i.e., NH+4 + OH− salt bridge in explicit solvent. We have computed the salt bridge PMF

vs. N−O distance (inset of Figure 5-5).

We performed 2 sets of simulations on the salt bridge and for each set we repeated

the calculations 10 times. In set 1 we used the naive setting of equal distance between

the adjacent bias centers along the reaction coordinate and in the set 2 we optimized

the positions of replicas. For the optimization leading to set 2, we used dumped values

of the first 20 ps of one of the simulations in set 1 to maximize the FBSF.

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45equi-distance

0

0.05

0.1

0.15

0.2

0.25

0.3

0.35

0.4

0.45

2.5 3 3.5 4 4.5 5 5.5 6 6.5 7

EA

R

Replica Position (ξ)

optimized

Figure 5-4. Window positions vs. EARs in the equi-distance (non-optimized) (red circles)and optimized (set 2) (green circles) simulations. The EAR of each positioncorresponds to EAR between that replica position and its higher neighbor onthe reaction coordinate.

90

Page 91: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

0

0.002

0.004

0.006

0.008

0.01

0.012

0 5 10 15 20 25 30 35 40

< K

ullb

ack−

Leib

ler

dive

rgen

ce >

Time (ps)

set 1 (non−optimized)set 2 (optimized)

0

1

2

3

4

5

3 4 5 6 7

PM

F (

kcal

/mol

)

ξ (N−O distance, Angstrom)

reference PMF

Figure 5-5. The mean and RMSE of Kullback-Leibler divergence over 10 simulations foreach set. Set 1 (red) is the naive equi-distance arrangement and set 2(green) is our optimized arrangement. We used the PMF of 5 ns USREsimulation as a reference for both simulation sets (the inset). The insetshows PMF vs. reaction coordinate (N−O distance)

5.4 Concluding Remarks

In the present work, we used statistical mechanics techniques to estimate the

Exchange Acceptance Ratios between any two windows on the reaction coordinate for

USRE.

Using the proposed schemes we set the positions of replicas such that all replicas

have the same EAR with their immediate neighbors. Since in USRE there is no phase

transition along the reaction coordinate then this setup maximizes the number of

roundtrips between the lowest and highest replicas and consequently the efficiency of

USRE. We applied our optimization method to butane in implicit solvent and NH+4 +OH−

salt bridge in explicit solvent. From the results, it was evident that this optimization

substantially increased the mixing and convergence rate for both systems. We expect

91

Page 92: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

the optimization of USRE to significantly increase the speed of convergence in larger

systems.

92

Page 93: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

CHAPTER 6CONCLUSIONS

In spite of the excessive growth in the power of computer hardware in recent years,

still the role of efficient sampling in Computational biophysics and biochemistry is

crucial. My PhD research was mostly dedicated to developing and optimizing enhanced

sampling methods and predominantly Hamiltonian Replica Exchange Molecular

Dynamics (HREMD). In my first project, we demonstrated that the HREMD method

improves the convergence rate in alchemical free energy calculations. Moreover,

we showed that there is a direct mapping between the HREMD and Free Energy

Perturbation (FEP) methods, which can be used for both exchange acceptance and free

energy calculations.

In the second project, I used the Replica Exchange Free Energy Method to

estimate the pKa shift in Glutamate 66 in a hyperstable mutant of staphylococcal

nuclease. In addition, we aimed to resolve an inconsistency between pKa experimental

and continuum methods in estimation of pKa value of position 66 in that protein. We

proposed that the experimental methods, which are mostly sensitive to configurational

changes, measure the equilibrium constant between two configurational states

instead of two protonation states. The results are in almost perfect agreement with

the Morenoes lab experiments.

In the third project, we developed and validated a pH-Replica Exchange Molecular

Dynamics (pH-REMD) method. We applied this method to a series of model compounds,

a terminally charged ADFDA pentapeptide, and a heptapeptide derived from the

Ovomucoid third domain (OMTKY3). We showed that not only the predicted pKas in

pH-REMD are very similar to that of Constant pH MD (CpHMD), but also the sampling is

more effective than that of CpHMD methods.

Finally in the last project, we focused on optimizing the Umbrella Sampling Replica

Exchange (USRE) method. We invented, validated, and tested a method for estimating

93

Page 94: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

the probability of exchange between neighboring replicas. Using information from

short umbrella runs, we optimized the position of replicas (windows) along the reaction

coordinate, and we showed that the equal exchange acceptance between replica pairs

is the optimum setting for USRE.

94

Page 95: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

APPENDIX ACALCULATING THE PDF OF ∆ξ IN USRE

In US (and subsequently USRE), approximating the Probability Density Function

(PDF) of ξ by a normal distribution is valid assumption[169]. So e.g., for replica i we can

write:

pi(ξ) =1√2πσ2

ξi

exp

[−(ξ − ⟨ξi⟩)2

2 σ2ξi

]. (A–1)

Since the windows are independent, the joint PDF of ξi and ξj (i.e., p(ξi, ξj)) is equal to

multiplication of their PDF; then we can compute the PDF of ∆ξij = ξi − ξj by integrating

the joint PDF of ξi and ξj, i.e.,

p(∆ξij) =

∫ ∞

−∞dξi p(ξi, ξj) =

∫ ∞

−∞dξi p(ξi) p(ξj) =

∫ ∞

−∞dξi p(ξi) p(−∆ξij + ξi)

=

∫ ∞

−∞

1

2πσξiσξj

exp−1

2

[(ξi − ⟨ξi⟩)2

σ2ξi

+(−∆ξij + ξi − ⟨ξj⟩)2

σ2ξj

]d∆ξi

=1√

2π(σ2ξi+ σ2

ξj

) exp

−(∆ξij − (⟨ξi⟩ − ⟨ξj⟩))2

2(σ2ξi+ σ2

ξj

) .

(A–2)

which itself is a normal distribution.

95

Page 96: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

APPENDIX BESTIMATING MEANS AND VARIANCES OF THE WINDOWS USING NEAREST

NEIGHBOR WEIGHTED AVERAGING AND REWEIGTHING

In order to find the optimal positions of replicas on the reaction coordinate, we need

to estimate the mean and variance at the replicas’ new positions.

The most primitive way to estimate the mean and variance of ξ (i.e. ⟨ξ⟩ and σ2)

for each window is to run a short US simulation at that window. Since optimizing the

positions of replicas requires moving the positions (i.e., center of biases) of replicas at

each step of optimization, it is not very efficient to use this method for estimating the

mean and variance of ξ. Instead, we develop a method below, which approximates ⟨ξ⟩

and σ2 for any arbitrary replica on the reaction coordinate using the ⟨ξ⟩ and σ2 of a few,

very short time pre-simulated windows.

Suppose we know the center of bias, variance and mean for N pre-simulated

windows, i.e., ξ01 ... ξ0N

⟨ξ1⟩ ... ⟨ξN⟩

⟨ξ21⟩ ... ⟨ξ2N⟩

. (B–1)

If replica i is a pre-simulated replica, then theoretically we can estimate the first two

non-central moments of the ξ distribution (i.e., ⟨ξ⟩ and σ2) on any new window m on the

reaction coordinate using the reweighting technique (section B.1 )as:

⟨ξ⟩m;i =

∫∞−∞ ξ 1√

2πσ2ξi

exp

[− (ξ−⟨ξi⟩)2

2σ2ξi

]exp [βk ((ξ − ξ0i )

2 − (ξ − ξ0m)2)] dξ

∫∞−∞

1√2πσ2

ξi

exp

[− (ξ−⟨ξi⟩)2

2σ2ξi

]exp [βk ((ξ − ξ0i )

2 − (ξ − ξ0m)2)] dξ

(B–2)

and

⟨ξ2⟩m;i =

∫∞−∞ ξ2 1√

2πσ2ξi

exp

[− (ξ−⟨ξi⟩)2

2σ2ξi

]exp [βk ((ξ − ξ0i )

2 − (ξ − ξ0m)2)] dξ

∫∞−∞

1√2πσ2

ξi

exp

[− (ξ−⟨ξi⟩)2

2σ2ξi

]exp [βk ((ξ − ξ0i )

2 − (ξ − ξ0m)2)] dξ

(B–3)

96

Page 97: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

where ⟨f(ξ)⟩m;i is the average of f(ξ) over distribution of m using the distribution of i. We

can solve the above integrations by writing the power of exponential in a single square

form:

−(ξ−⟨ξi⟩)2

2σ2ξi

+ βk((ξ − ξ0i )

2 − (ξ − ξ0m)2)

=[−2⟨ξi⟩βk(ξ0i − ξ0m) + 2σ2

i (βk(ξ0i − ξ0m))

2]−

(ξ − (⟨ξi⟩ − 2σ2i βk(ξ

0i − ξ0m))

2σ2i

+ βk((ξ0i )

2 − (ξ0m)2) (B–4)

and using ∞∫

−∞

1√πC

ξ exp[− (ξ−B)2

C

]= B

∞∫−∞

1√πC

ξ2 exp[− (ξ−B)2

C

]= 1

2C + B2

, (B–5)

Equations (B–2) and B–3 simplifies to: ⟨ξ⟩m;i = ⟨ξi⟩ − 2σ2i βk(ξ

0i − ξ0m)

⟨ξ2⟩m;i = (⟨ξi⟩ − 2σ2i βk(ξ

0i − ξ0m))

2+ σ2

i

, (B–6)

If replicas i and i + 1 are the nearest pre-simulated neighbors of our imaginary replica m,

i.e., ξ0i < ξ0m < ξ0i+1, then using the nearest neighbor weighted averaging method (section

B.2) we can estimate the mean and variance of ξ on windows m as following:⟨ξ⟩m = wi (⟨ξi⟩ − 2σ2

i βk(ξ0i − ξ0m)) + wi+1

(⟨xii+1 − 2σ2

i+1βk(ξ0i+1 − ξ0m)

)⟨ξ2⟩m = wi

[(⟨ξi⟩ − 2σ2

i βk(ξ0i − ξ0m))

2+ σ2

i

]+ wi+1

[(⟨ξi+1⟩ − 2σ2

i+1βk(ξ0i+1 − ξ0m)

)2+ σ2

i+1

]σ2m = ⟨ξ2⟩m − ⟨ξ⟩2m

,

(B–7)

97

Page 98: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

where we choose the power parameter equal to 1.2 and number of involving nearest

neighbors equals to 2, so Equation B–14 in the section B.2 implies:

wi =

(|ξ0i −ξ0m|)

0.6

(|ξ0i −ξ0m|)0.6

+(|ξ0i+1−ξ0m|)0.6 ξ0m = ξ0i and ξ0m = ξ0i+1

1 ξ0m = ξ0i

0 ξ0m = ξ0i+1

. (B–8)

Using the above setup, we are able to estimate the mean and variance of any

arbitrary window on the reaction coordinate. Subsequently it is possible to approximate

the EAR between any two replicas using Equation 5–8.

B.1 Reweighting Fitted Gaussian Distribution on Windows

Here we describe the reweighting method that we used for estimation of the mean

and the variance of arbitrary windows on reaction coordinate. Consider a continuous

function f(ξ). We can calculate the average of f(ξ) over a distribution of p(ξ) as following:

⟨f(ξ)⟩p =

∫f(ξ) p(ξ) dξ∫p(ξ) dξ

. (B–9)

If we want to calculate the average of f(ξ) over another distribution q(ξ), where the

domain of q(ξ) is a subset of p(ξ), then we may write:

⟨f(ξ)⟩q;p =

∫f(ξ) q(ξ) dξ∫q(ξ) dξ

=

∫f(ξ) p(ξ)q(ξ)

p(ξ)dξ∫

p(ξ)q(ξ)p(ξ)

dξ=⟨q(ξ)p(ξ)

f(ξ)⟩p⟨q(ξ)p(ξ)⟩p

, (B–10)

where ⟨f(ξ)⟩q;p is the average of f(ξ) over distribution of q using the distribution of p.

In the case of Umbrella Sampling and by assuming Gaussian distributions for the

collective variables for every window (i.e., using Equation A–1), we can estimate the

average of f(ξ) over window i:

⟨f(ξ)⟩pi =

∫f(ξ) 1√

2πσ2ξi

exp

[− (ξ−⟨ξi⟩)2

2 σ2ξi

]dξ

∫1√

2πσ2ξi

exp

[− (ξ−⟨ξi⟩)2

2 σ2ξi

]dξ

. (B–11)

98

Page 99: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

Using the Equation 5–1 and Equation 5–2, we can compute the ratio of probability

distribution of replicas i and m as

pm(ξ)

pi(ξ)=

exp [−β(UUB + Bm)]

exp [−β(UUB + Bi)]= exp

[βk

((ξ − ξ0i )

2 − (ξ − ξ0m)2)]

(B–12)

Finally using Equation B–9 and Equation B–12, we can exploit

⟨f(ξ)⟩pm;pi = ⟨f(ξ)⟩m;i =

∫f(ξ) 1√

2πσ2ξi

exp

[− (ξ−⟨ξi⟩)2

2 σ2ξi

]exp [βk ((ξ − ξ0i )

2 − (ξ − ξ0m)2)] dξ

∫1√

2πσ2ξi

exp

[− (ξ−⟨ξi⟩)2

2 σ2ξi

]exp [βk ((ξ − ξ0i )

2 − (ξ − ξ0m)2)] dξ

.

(B–13)

Which is the average of f(ξ) over the window m using the window i. Using Equation

B–13 we calculate the values of ⟨ξ⟩ and ⟨ξ2⟩ for any arbitrary windows, using the pre-

-simulated windows (i.e., Equation B–6).

B.2 Nearest Neighbor Weighted Averaging (NNWA)

We used NNWA to make a better estimation of mean and variance of ξ for any

arbitrary window by pre-simulated windows. NNWA is a member of Locally Weighted

Learning methods[200]. Consider a function f(ξ) with a few known points (i.e., ξi);

furthermore suppose the value of f at any arbitrary ξ position can be estimated

separately by each of those known points; then using k known nearest neighbor of

ξ, we can approximate the value of f, as:

f(ξ) =k∑

i=1

wi(ξ) fξi(ξ)k∑

i=1

wi(ξ). (B–14)

where fξi(ξ) is the estimation of f(ξ) using ξi and wi(ξ) = 1d(ξ,ξi)

. d is so-called metric

operator, which in its simplest form is d(ξ, ξi) = [(ξ − ξi)2]. p is a positive real number

known as a power parameter. Depending on the problem, k can be changed from the

first nearest neighbors to the total number of known points. In this paper we set p = 1.2

and k = 2, which helped the optimizer to smoothly converge.

99

Page 100: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

APPENDIX CBARZILAI AND BORWEIN OPTIMIZATION (BB METHOD)

The Steepest Descent or Gradient Descent method is among the most important

first order optimization algorithms. It can be formulized as follows:

xk+1 = xk + λk gk, (C–1)

where gk = −∇f(xk) and λk is the step size at step k. There are many methods

for estimating the optimal step size[195, 201]. BB method[194] is one of the most

efficient routines among those, in which the step size is determined by minimizing either

∥∆x−λ∆g∥2 or its correspondent ∥λ−1∆x−∆g∥2 with respect to λ, where ∆x = xk−xk+1

and ∆g = gk − gk+1. Based on those, the optimal step size can be derived as:

λk =∆x⊺∆g

∆g⊺∆g. (C–2)

100

Page 101: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

REFERENCES

[1] Swendsen, R. H.; Wang, J.-S. Phys. Rev. Lett. 1986, 57, 2607–2609.

[2] Sugita, Y.; Okamoto, Y. Chem. Phys. Lett. 1999, 314, 141–145.

[3] Mitsutake, A.; Sugita, Y.; Okamoto, Y. Biopolymers. 2001, 60, 96–123.

[4] Sugita, Y.; Kitao, A.; Okamoto, Y. arXiv preprint cond-mat 2000, 113, 6042.

[5] McQuarrie, D. Statistical Mechanics; University Science Books, 2000.

[6] Chandler, D. Introduction to modern statistical mechanics; Oxford university press:New York, Oxford, 1987.

[7] Leach, A. Molecular Modelling: Principles and Applications (2nd Edition), 2nd ed.;Prentice Hall, 2001.

[8] Landau, R. H.; Paez Mejıa, M. J.; Bordeianu, C. A Survey of computationalphysics : introductory computational science, har/cdr ed.; Princeton UniversityPress, 1997.

[9] Cramer, C. J. Essentials of Computational Chemistry: Theories and Models, 2nded.; John Wiley Sons: West Sussex, England, 2005.

[10] van Gunsteren, W. F.; Berendsen, H. J. Angew. Chem. 1990, 29, 992–1023.

[11] Levitt, M.; Warshel, A. Nature 1975, 253, 694–698.

[12] Li, Z.; Scheraga, H. A. Proc. Natl. Acad. Sci. U. S. A. 1987, 84, 6611–6615.

[13] Landau, D. P.; Binder, K. A guide to Monte Carlo simulations in statistical physics;Cambridge university press, 2009.

[14] Metropolis, N.; Rosenbluth, A. W.; Rosenbluth, M. N.; Teller, A. H.; Teller, E. J.Chem. Phys. 1953, 21, 1087.

[15] Marinari, E.; Parisi, G. Europhys. Lett. 1992, 19, 451.

[16] Wang, F.; Landau, D. P. Phys. Rev. Lett. 2001, 86, 2050–2053.

[17] Okamoto, Y.; Hansmann, U. H. E. J. Phys. Chem. 1995, 99, 11276–11287.

[18] Berg, B. A.; Neuhaus, T. Phys. Rev. Lett. 1992, 68, 9–12.

[19] Berg, B. A.; Celik, T. Phys. Rev. Lett. 1992, 69, 2292–2295.

[20] Murata, K.; Sugita, Y.; Okamoto, Y. Chem. Phys. Lett. 2004, 385, 1–7.

[21] Woods, C. J.; Essex, J. W.; King, M. A. J. Phys. Chem. B 2003, 107,13703–13710.

101

Page 102: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[22] Meng, Y.; Dashti, D. S.; Roitberg, A. E. J. Chem. Theory Comput. 2011, 7,2721–2727.

[23] Itoh, S. G.; Damjanovic, A.; Brooks, B. R. Proteins: Struct., Funct., Bioinf. 2011,79, 3420–3436.

[24] Sabri Dashti, D.; Meng, Y.; Roitberg, A. E. J. Phys. Chem. B 2012, 116,8805–8811.

[25] Torrie, G.; Valleau, J. J. Comput. Phys. 1977, 23, 187–199.

[26] Grossfield, A. WHAM: an implementation of the weighted histogram analysismethod. 2010; http://membrane.urmc.rochester.edu/content/wham.

[27] Kastner, J.; Thiel, W. J. Chem. Phys. 2005, 123, 144104.

[28] Kastner, J. J. Chem. Phys. 2012, 136, 234102.

[29] Riccardi, D.; Schaefer, P.; Cui, Q. J. Phys. Chem. B 2005, 109, 17715–17733,PMID: 16853267.

[30] Zwanzig, R. W. J. Chem. Phys. 1954, 22, 1420–1426.

[31] Carugo, O.; Pongor, S. Protein Sci. 2001, 10, 1470–1473.

[32] Matthews, B. W.; Rossmann, M. G. In Diffraction Methods for Biological Macro-molecules Part B; Harold W. Wyckoff, S. N. T., C. H. W. Hirs, Ed.; Methods inEnzymology; Academic Press, 1985; Vol. 115; pp 397 – 420.

[33] Orengo, C. Curr. Opin. Struct. Biol. 1994, 4, 429 – 440.

[34] Holm, L.; Sander, C. J. Mol. Biol. 1993, 233, 123 – 138.

[35] Garc´, A. E. Phys. Rev. Lett. 1992, 68, 2696–2699.

[36] Smith, L. J.; Daura, X.; van Gunsteren, W. F. Proteins: Struct., Funct., Bioinf.2002, 48, 487–496.

[37] Flyvbjerg, H.; Petersen, H. G. J. Chem. Phys. 1989, 91, 461–466.

[38] Hess, B. Phys. Rev. E 2002, 65, 031910.

[39] Kullback, S.; Leibler, R. A. Ann. Math. Statist. 1951, 22, 79–86.

[40] McClendon, C. L.; Hua, L.; Barreiro, A.; Jacobson, M. P. J. Chem. Theory Comput.2012, 8, 2115–2126.

[41] Sabri Dashti, D.; Roitberg, A. submitted to J. Phys. Chem. B

[42] Sabri Dashti, D.; Roitberg, A. submitted to J. Chem. Theory Comput.

102

Page 103: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[43] Chipot, C.; Pohorille, A. Free Energy Calculations: Theory and Applications inChemistry and Biology ; Springer series in chemical physics; Springer-VerlagBerlin Heidelberg, 2007.

[44] Bash, P.; Singh, U.; Langridge, R.; Kollman, P.; et al., Science 1987, 236,564–568.

[45] Kirkwood, J. G. J. Chem. Phys. 1935, 3, 300.

[46] Mezei, M. J. Comput. Phys. 1987, 68, 237–248.

[47] Roux, B. Comput. Phys. Commun. 1995, 91, 275–282.

[48] Jarzynski, C. Phys. Rev. Lett. 1997, 78, 2690–2693.

[49] Kumar, S.; Rosenberg, J. M.; Bouzida, D.; Swendsen, R. H.; Kollman, P. A. J.Comput. Chem. 1992, 13, 1011–1021.

[50] Bennett, C. H. J. Comput. Phys. 1976, 22, 245–268.

[51] Shirts, M. R.; Chodera, J. D. J. Chem. Phys. 2008, 129, 124105–124110.

[52] Pohorille, A.; Jarzynski, C.; Chipot, C. J. Phys. Chem. B 2010, 114, 10235–10253.

[53] Zheng, L.; Chen, M.; Yang, W. Proc. Natl. Acad. Sci. U. S. A. 2008, 105, 20227–20232.

[54] Hamelberg, D.; Mongan, J.; McCammon, J. A. J. Chem. Phys. 2004, 120,11919–11929.

[55] Fukunishi, H.; Watanabe, O.; Takada, S. J. Chem. Phys. 2002, 116, 9058.

[56] Liu, P.; Kim, B.; Friesner, R. A.; Berne, B. J. Proc. Natl. Acad. Sci. 2005, 102,13749–13754.

[57] Liu, P.; Voth, G. A. J. Chem. Phys. 2007, 126, 045106.

[58] Lyman, E.; Ytreberg, F.; Zuckerman, D. Phys. Rev. Lett. 2006, 96, 028105.

[59] Lwin, T. Z.; Luo, R. J. Chem. Phys. 2005, 123, 194904.

[60] Rick, S. W. J. Chem. Theory Comput. 2006, 2, 939–946.

[61] Min, D.; Li, H.; Li, G.; Bitetti-Putzer, R.; Yang, W. J. Chem. Phys. 2007, 126,144109.

[62] Jiang, W.; Hodoscek, M. J. Chem. Theory Comput. 2009, 5, 2583–2588.

[63] Jiang, W.; Roux, B. J. Chem. Theory Comput. 2010, 6, 2559–2565.

[64] Dyson, H. J.; Tennant, L. L.; Holmgren, A. Biochemistry 1991, 30, 4262–4268.

103

Page 104: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[65] Langsetmo, K.; Fuchs, J. A.; Woodward, C. Biochemistry 1991, 30, 7603–7609.

[66] Christ, C. D.; Mark, A. E.; van Gunsteren, W. F. J. Comput. Chem. 2010, 31,1569–1582.

[67] Jorgensen, W. L.; Thomas, L. L. J. Chem. Theory Comput. 2008, 4, 869.

[68] Kollman, P. Chem. Rev. 1993, 93, 2395–2417.

[69] Straatsma, T.; McCammon, J. Annu. Rev. Phys. Chem. 1992, 43, 407–435.

[70] Jorgensen, W. L.; Ravimohan, C. J.Chem.Phys. 1985, 83, 3050–3054.

[71] Lu, N.; Kofke, D. A.; Woolf, T. B. J. Comput. Chem. 2004, 25, 28–40.

[72] Simonson, T.; Carlsson, J.; Case, D. A. J. Am. Chem. Soc. 2004, 126, 4167–4180.

[73] Bashford, D.; Case, D. A.; Dalvit, C.; Tennant, L.; Wright, P. E. Biochemistry 1993,32, 8045–8056.

[74] Katti, S. K.; LeMaster, D. M.; Eklund, H. J. Mol. Biol. 1990, 212, 167–184.

[75] Case, D. A.; Darden, T. A.; Cheatham, I.; Simmerling, C. L.; Wang, J.; Duke, R. E.;Luo, R.; Crowley, M.; Walker, R. C.; Zhang, W.; et al., AMBER 10. 2008.

[76] Hornak, V.; Abel, R.; Okur, A.; Strockbine, B.; Roitberg, A.; Simmerling, C.Proteins: Struct., Funct., Bioinf. 2006, 65, 712–725.

[77] Ryckaert, J. P.; Ciccotti, G.; Berendsen, H. J. C. J. Comput. Phys. 1977, 23,327–341.

[78] Onufriev, A.; Case, D. A.; Bashford, D. J. Comput. Chem. 2002, 23, 1297–1304.

[79] Perutz, M. F. Faraday Discuss. 1992, 1–11, PMID: 1290926.

[80] Varadarajan, R.; Zewert, T.; Gray, H.; Boxer, S. Science 1989, 243, 69–72.

[81] Churg, A. K.; Warshel, A. Biochemistry 1986, 25, 1675–1681, PMID: 3011070.

[82] Gennis, R. B. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 12747 –12749.

[83] Isom, D. G.; Cannon, B. R.; Castaneda, C. A.; Robinson, A.; et al., Proc. Natl.Acad. Sci. U. S. A. 2008, 105, 17784–17788.

[84] Isom, D. G.; Castaneda, C. A.; Cannon, B. R.; E, B. G.-m. Proc. Natl. Acad. Sci. U.S. A. 2011, 108, 5260–5265.

[85] Luecke, H.; Lanyi, J. K.; Rees, D. C. Membrane Proteins; Academic Press, 2003;Vol. 63; pp 111–130.

[86] Parson, W. W.; Chu, Z.-T.; Warshel, A. Biochim Biophys Acta Bioenerg 1990,1017, 251–272.

104

Page 105: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[87] Li, Y. K.; Kuliopulos, A.; Mildvan, A. S.; Talalay, P. Biochemistry 1993, 32,1816–1824, PMID: 8439542.

[88] Bogan, A. A.; Thorn, K. S. J. Mol. Biol. 1998, 280, 1–9.

[89] Adelroth, P.; Ek, M. S.; Mitchell, D. M.; Gennis, R. B.; Brzezinski, P. Biochemistry1997, 36, 13824–13829, PMID: 9374859.

[90] Luecke, H.; Richter, H.-T.; Lanyi, J. K. Science 1998, 280, 1934–1937.

[91] Denisov, V. P.; Schlessman, J. L.; Garcıa-Moreno E, B.; Halle, B. Biophys. J 2004,87, 3982–3994, PMID: 15377517.

[92] Karp, D. A.; Stahley, M. R.; Garcıa-Moreno, B. Biochemistry 2010, 49, 4138–4146,PMID: 20329780.

[93] Mehler, E. L.; Fuxreiter, M.; Simon, I.; GarciaMoreno E, B. Proteins: Struct, Funct,Bioinf 2002, 48, 283–292.

[94] Stites, W. E.; Gittis, A. G.; Lattman, E. E.; Shortle, D. J. Mol. Biol 1991, 221, 7–14,PMID: 1920420.

[95] Castaneda, C. A.; Fitch, C. A.; Majumdar, A.; Khangulov, V.; Schlessman, J. L.;Garcıa-Moreno, B. E. Proteins. 2009, 77, 570–588, PMID: 19533744.

[96] Isom, D. G.; Castaneda, C. A.; Cannon, B. R.; Velu, P. D.; Garcıa-Moreno E., B.Proc. Natl. Acad. Sci. U.S.A 2010, 107, 16096–16100, PMID: 20798341 PMCID:2941338.

[97] Karp, D. A.; Gittis, A. G.; Stahley, M. R.; Fitch, C. a.; Stites, W. E.;Garcıa-Moreno E, B.; Garcıa-Moreno E., B. Biophys. J. 2007, 92, 2041–53.

[98] Chimenti, M. S.; Khangulov, V. S.; Robinson, A. C.; Heroux, A.; Majumdar, A.;Schlessman, J. L.; Garcıa-Moreno, B. Structure 2012, 20, 1071–1085, PMID:22632835.

[99] Dwyer, J. Biophys. J. 2000, 79, 1610–1620.

[100] Garcıa-Moreno, B.; Dwyer, J. J.; Gittis, A. G.; Lattman, E. E.; Spencer, D. S.;Stites, W. E. Biophys. Chem 1997, 64, 211–224, PMID: 9127946.

[101] Onufriev, A.; Bashford, D.; Case, D. A. J. Phys. Chem. B 2000, 104, 3712–3720.

[102] Tsui, V.; Case, D. A. Biopolymers 2000, 56, 275–291, PMID: 11754341.

[103] Onufriev, A.; Bashford, D.; Case, D. A. Proteins 2004, 55, 383–394, PMID:15048829.

[104] Damjanovic, A.; Garcıa-Moreno, B.; Lattman, E. E.; Garcıa, A. E. Proteins 2005,60, 433–449, PMID: 15971206.

105

Page 106: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[105] Fitch, C. Biophys. J. 2002, 82, 3289–3304.

[106] Wu, X.; Brooks, B. R. Chem. Phys. Lett. 2003, 381, 512–518.

[107] Sindhikara, D. J.; Kim, S.; Voter, A. F.; Roitberg, A. E. J. Chem. Theory Comput.2009, 5, 1624–1631.

[108] Sosa, C. P.; Hewitt, T.; Lee, M. R.; Case, D. A. J. Mol. Struct. Theochem 2001,549, 193–201.

[109] Sreerama, N.; Woody, R. W. Anal. Biochem. 2000, 287, 252–260.

[110] Sreerama, N.; Woody, R. W. Anal. Biochem. 2004, 383, 318–351.

[111] Alexov, E.; Mehler, E. L.; Baker, N.; M. Baptista, A.; Huang, Y.; Milletti, F.;Erik Nielsen, J.; Farrell, D.; Carstensen, T.; Olsson, M. H. M.; Shen, J. K.;Warwicker, J.; Williams, S.; Word, J. M. Proteins: Struct., Funct., Bioinf. 2011,79, 3260–3275.

[112] Tanford, C.; Kirkwood, J. G. J. Am. Chem. Soc. 1957, 79, 5333–5339.

[113] Bashford, D.; Karplus, M. Biochemistry 1990, 29, 10219.

[114] Warshel, A.; Russell, S. T. Q. Rev. Biophys. 1984, 17, 283–422.

[115] Baptista, A. M. J. Chem. Phys. 2002, 116, 7766–7768.

[116] Di Russo, N. V.; Estrin, D. A.; Martı, M. A.; Roitberg, A. E. PLoS Comput Biol2012, 8, e1002761.

[117] Bierzynski, A.; Kim, P. S.; Baldwin, R. L. Proc. Natl. Acad. Sci. U.S.A. 1982, 79,2470–2474.

[118] Shoemaker, K. R.; Kim, P. S.; Brems, D. N.; Marqusee, S.; York, E. J.;Chaiken, I. M.; Stewart, J. M.; Baldwin, R. L. Proc. Natl. Acad. Sci. U.S.A. 1985,82, 2349–2353.

[119] Schaefer, M.; Van Vlijmen, H. W.; Karplus, M. Adv. Protein Chem. 1998, 51, 1–57.

[120] Demchuk, E.; Genick, U. K.; Woo, T. T.; Getzoff, E. D.; Bashford, D. Biochemistry2000, 39, 1100–1113.

[121] Dillet, V.; Dyson, H. J.; Bashford, D. Biochemistry 1998, 37, 10298–10306.

[122] Harris, T. K.; Turner, G. J. IUBMB life 2002, 53, 85–98.

[123] Antosiewicz, J.; Briggs, J. M.; McCammon, J. A. Eur. Biophys. J. 1996, 24,137–141.

[124] Hunenberger, P. H.; Helms, V.; Narayana, N.; Taylor, S. S.; McCammon, J. A.Biochemistry 1999, 38, 2358–2366.

106

Page 107: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[125] Hill, T. L. J. Am. Chem. Soc. 1956, 78, 1577–1580.

[126] Warshel, A. Nature 1987, 330, 15–16.

[127] Mongan, J.; Case, D. A.; McCammon, J. A. J. Comput. Chem. 2004, 25,2038–2048.

[128] Baptista, A. M.; Martel, P. J.; Petersen, S. B. Proteins 1997, 27, 523–544.

[129] Borjesson, U.; Hunenberger, P. H. J. Chem. Phys. 2001, 114, 9706.

[130] Khandogin, J.; Brooks, C. L. Biophys. J. 2005, 89, 141–157.

[131] Khandogin, J.; Brooks, C. L. Biochemistry 2006, 45, 9363–9373.

[132] Khandogin, J.; Brooks, C. L. Proc. Natl. Acad. Sci. U.S.A. 2007, 104,16880–16885.

[133] Khandogin, J.; Chen, J. H.; Brooks, C. L. Proc. Natl. Acad. Sci. U.S.A. 2006, 103,18546–18550.

[134] Khandogin, J.; Raleigh, D. P.; Brooks, C. L. J. Am. Chem. Soc. 2007, 129,3056–3057.

[135] Mertz, J. E.; Pettitt, B. M. Int. J. Supercomput. Ap. 1994, 8, 47–53.

[136] Borjesson, U.; Hunenberger, P. H. J. Phys. Chem. B 2004, 108, 13551–13559.

[137] Baptista, A. M.; Teixeira, V. H.; Soares, C. M. J. Chem. Phys. 2002, 117, 4184.

[138] Machuqueiro, M.; Baptista, A. M. Biophys. J. 2006, 92, 1836–1845.

[139] Machuqueiro, M.; Baptista, A. M. Proteins: Struct., Funct., Bioinf. 2008, 72, 289.

[140] Machuqueiro, M.; Baptista, A. M. J. Am. Chem. Soc. 2009, 131, 12586–12594.

[141] Burgi, R.; Kollman, P. a.; Van Gunsteren, W. F. Proteins 2002, 47, 469–80.

[142] Dlugosz, M.; Antosiewicz, J. M. Chem. Phys. 2004, 302, 161–170.

[143] Dlugosz, M.; Antosiewicz, J. M. J. Phys. Chem. B 2005, 109, 13777–13784.

[144] Dlugosz, M.; Antosiewicz, J. M. J. Phys.: Condens. Matter 2005, 17, S1607.

[145] Dlugosz, M.; Antosiewicz, J. M.; Robertson, A. D. Phys. Rev. E 2004, 69,21915–21924.

[146] Walczak, A. M.; Antosiewicz, J. M. Phys. Rev. E 2002, 66, 051911–051918.

[147] Lee, M. S.; Salsbury, F. R.; Brooks, C. L. Proteins: Struct., Funct., Bioinf. 2004, 56,738–752.

107

Page 108: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[148] Machuqueiro, M.; Baptista, A. M. J. Phys. Chem. B 2006, 110, 2927–2933.

[149] Johnson, M. L., Brand, L., Eds. Methods in Enzimology ; Academic Press: SanDiego, Burlington, London, 2009; Vol. 467.

[150] Warwicker, J. Protein Sci. 2004, 13, 2793–2805.

[151] Barth, P.; Alber, T.; Harbury, P. B. Proc. Natl. Acad. Sci. U.S.A. 2007, 104,4898–4903.

[152] Li, H.; Fajer, M.; Yang, W. J. Chem. Phys. 2007, 126, 024106.

[153] Zheng, L. Q.; Chen, M. G.; Yang, W. J. Chem. Phys. 2009, 130, 234105.

[154] Lyubartsev, A. P.; Martsinovski, A. A.; Shevkunov, S. V.;VorontsovVelyaminov, P. N. J. Chem. Phys. 1992, 96, 1776.

[155] Berg, B. A.; Neuhaus, T. Phys. Lett. B 1991, 267, 249–253.

[156] Laio, A.; Parrinello, M. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 12562–12566,PMID: 12271136 PMCID: 130499.

[157] Zheng, L.; Chen, M.; Yang, W. Proc. Natl. Acad. Sci. U. S. A 2008, 105,20227–20232.

[158] Wallace, J. A.; Shen, J. K. J. Chem. Theory Comput. 2011, 7, 2617–2629.

[159] Wallace, J. A.; Wang, Y.; Shi, C.; Pastoor, K. J.; Nguyen, B.; Xia, K.; Shen, J. K.Proteins: Struct., Funct., Bioinf. 2011, 79, 3364–3373.

[160] Meng, Y.; Roitberg, A. E. J. Chem. Theory Comput. 2010, 6, 1401–1412.

[161] Mongan, J.; Case, D. A. Curr. Opin. Struct. Biol. 2005, 15, 157–163.

[162] Williams, D. H.; Stephens, E.; O’Brien, D. P.; Zhou, M. J. Chem. Phys. 2004, 43,6596–616.

[163] Sindhikara, D.; Meng, Y.; Roitberg, A. E. J. Chem. Phys. 2008, 128, 24103.

[164] Earl, D. J.; Deem, M. W. Phys. Chem. Chem. Phys. 2005, 7, 3910–3916.

[165] Faller, R.; Yan, Q.; de Pablo, J. J. Biopolymers. 2002, 116, 5419.

[166] Fenwick, M. K.; Escobedo, F. a. J. Chem. Phys. 2003, 119, 11998.

[167] Chodera, J. D.; Shirts, M. R. J. Chem. Phys. 2011, 135, 194110.

[168] Ballard, A. J.; Jarzynski, C. Proc. Natl. Acad. Sci. U. S. A. 2009, 106,12224–12229.

[169] Kastner, J. WIREs. Comput. Mol. Sci. 2011, 1, 932–942.

108

Page 109: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[170] Bartels, C. Chem. Phys. Lett. 2000, 331, 446–454.

[171] Neumann, J. V. Proc. Natl. Acad. Sci. U. S. A. 1932, 18, 70–82.

[172] Hritz, J.; Oostenbrink, C. J. Chem. Phys. 2008, 128, 144121.

[173] Okur, A.; Wickstrom, L.; Layten, M.; Geney, R.; Song, K.; Hornak, V.;Simmerling, C. J. Chem. Theory Comput. 2006, 2, 420–433.

[174] Cheng, X.; Cui, G.; Hornak, V.; Simmerling, C. J. Phys. Chem. B 2005, 109,8220–8230.

[175] Zhou, R.; Berne, B. J. Proc. Natl. Acad. Sci. U. S. A 2002, 99, 12777–12782.

[176] Zhou, R. Proteins. 2003, 53, 148–161.

[177] Roitberg, A. E.; Okur, A.; Simmerling, C. J. Chem. Theory Comput. 2007, 111,2415–2418.

[178] Okur, A.; Roe, D. R.; Cui, G. L.; Hornak, V.; Simmerling, C. J. Chem. TheoryComput. 2007, 3, 557–568.

[179] Rick, S. W. J. Chem. Phys. 2007, 126, 054102.

[180] Li, X.; O’Brien, C. P.; Collier, G.; Vellore, N. A.; Wang, F.; Latour, R. A.;Bruce, D. A.; Stuart, S. J. J. Chem. Phys. 2007, 127, 164116.

[181] Kamberaj, H.; van der Vaart, A. J. Chem. Phys. 2009, 130, 074906.

[182] Kone, A.; Kofke, D. A. J. Chem. Phys. 2005, 122, 206101.

[183] Kofke, D. J. Chem. Phys. 2002, 117, 6911.

[184] Kofke, D. A. J. Chem. Phys. 2004, 121, 1167.

[185] Predescu, C.; Predescu, M.; Ciobanu, C. V. J. Chem. Phys. 2004, 120,4119–4128.

[186] Sanbonmatsu, K.; Garcia, A. Proteins. 2002, 46, 225–234.

[187] Rathore, N.; Chopra, M.; de Pablo, J. J. J. Chem. Phys. 2005, 122, 024111.

[188] Katzgraber, H. G.; Trebst, S.; Huse, D. a.; Troyer, M. J. Stat. Mech. Theor. Exp.2006, 2006, 03018.

[189] Hritz, J.; Oostenbrink, C. J. Chem. Phys. 2007, 127, 204104.

[190] Trebst, S.; Troyer, M.; Hansmann, U. H. E. J. Chem. Phys. 2006, 124, 174903.

[191] Shenfeld, D.; Xu, H.; Eastwood, M.; Dror, R.; Shaw, D. Phys. Rev. E 2009, 80,046705.

109

Page 110: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

[192] Crooks, G. Phys. Rev. Lett. 2007, 99, 100602.

[193] Murata, K.; Sugita, Y.; Okamoto, Y. J. Theor. Comput. Chem 2005, 04, 411–432.

[194] Barzilai, J.; Borwein, J. M. J. Inst. Math. Its App. 1988, 8, 141–148.

[195] Yuan, Y.-x. AMS/IP Stud. Adv. Math. 2008, 42, 785–796.

[196] Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. J.Chem. Phys. 1983, 79, 926.

[197] Case, D.; Darden, T.; Cheatham, T.; Simmerling, C.; Wang, J.; et al., AMBER 12.2012.

[198] Wang, J.; Wolf, R. M.; Caldwell, J. W.; Kollman, P. a.; Case, D. a. J. Comput.Chem. 2004, 25, 1157–1174.

[199] Sindhikara, D.; Emerson, D.; Roitberg, A. J. Chem. Theory Comput. 2010, 6,2804–2808.

[200] Atkeson, C.; Moore, A.; Schaal, S. Artif. Intell. Rev. 1997, 11–73.

[201] Burachik, R.; Grana Drummond, L. M.; Iusem, A.; Svaiter, B. F. Optimization 1995,32, 137–146.

110

Page 111: DEVELOPING AND OPTIMIZING THE REPLICA EXCHANGE-BASED …

BIOGRAPHICAL SKETCH

Danial Sabri Dashti was born in, Tehran, Iran. He went to the University of Kashan

majoring in physics and graduated with a bachelor’s degree in 2003. Then, he went to

Sharif University of Technology where, he received his master’s degree in physics in

2006. After finishing his master, he moved to Gainesville, Florida in 2007. He entered

the Department of Physics of the University of Florida to pursue a Ph.D degree in

computational biophysics. He received his Ph.D. from the University of Florida in the

summer of 2013.

111


Recommended