challenges in computational biochemistry: solvation and ...172052/fulltext01.pdf · acta...

66
ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations from the Faculty of Science and Technology 432 Challenges in Computational Biochemistry: Solvation and Ligand Binding JENS CARLSSON ISSN 1651-6214 ISBN 978-91-554-7200-9 urn:nbn:se:uu:diva-8738

Upload: others

Post on 31-Aug-2020

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

ACTA

UNIVERSITATIS

UPSALIENSIS

UPPSALA

2008

Digital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology 432

Challenges in ComputationalBiochemistry: Solvation and LigandBinding

JENS CARLSSON

ISSN 1651-6214ISBN 978-91-554-7200-9urn:nbn:se:uu:diva-8738

Page 2: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

���������� �������� �� ������ �������� � �� �������� ������� � ������� ������������� � ������� !������ ��� " � "��# �� � ��$ %� �&� ������ % ���� % '&����&�()&� �������� *��� �� ������� � +����&(

��������

������� ,( "��#( �&������� � ���������� ���&�������� -����� �� .���� �����(/��� ����������� ���������( ������� ��� � ���� ����� � � ������� ���� ��������� �� ������� � ��� �� ��� � ������ 0 "( 1" ��( ������( 2-�3 4�#54�5$$05�"��54(

/������� ���������� % %��� ������� %� �������� �������� �� ������ ��� �������%� �&� ���������� % ���&������ ��������� �� ��� ���%�� � ��� �&����������������������( 2 �&�� �&����� �������� ������� 6��7 ��������� ��� ���� � ����������&��������� ��������� %� ������ �� ����� �����()&� �&��������� �������� ���&�8�� �� ���� � ��������� ��

������� %� �&��� ��������

���� �������� � �* ��%%���� ������( �� ��������� ��� ������� �� � �������� ��9������:��5�� ������ �����( )&� ���������� ��

������� ��� � 8���������� ��������

*��& ��������� � ��& �����( / ������� % �� ��������� �� � ������������������� ���&� �� ������� � ������ ��

��&�%�� � *���5���� �� ����� ������

&�������( )&� ���������� ��������� ������ � ���� �&�� �� ������ ��� % �&� ��

������� ��������� % �&�� �:���(��������� % �&� ����� �������� ����� 6.2+7 ���&� %� ���������� ������ ��

����� %��� ������� �� ��������( / �* ���� %� ��������� �&� ������������ ���� � �&�.2+ ���&� �� ������� �� �� �&* � �������� ����������� %��� ������� % &������( /.2+ ���&� ����� � ������ ����� ����������� �� ��� �������� �� �� �� �&* ��������� ����� %��� ������� %� �&������ % � ������� �:���( )&� ���������� % ���� �������� % ��;��� �� �� �&� .2+ ���&� � ������� ����� �%%������ %� ������������� % ������ �� ��� �����������( 9� �������� *��& ��������� �� %�� %� � ��� %5�������� �&������ % �2<5� ������� ������������(/�����&�� %� �������� ������ �� ����� %��� ������� �� ��&����� ��

������ ������ ��� ��� �������( ���&�� %� ���������� �&� ���������� ��������� ����� ������� %� � ����� ��� ��������( )&� ���������� � ��������� �&������ %��� ������� �� ������� %� ��;��� ����� �� �� ���� ������ %��� ��������&�8��� �� ��� ����������� �� �&� ������� ����� *��� *��& ����������� ����(

� ������ ������� ���������� �������� �������� ������ %��� ������9������:��5��� '���5���:��� ����� ������ ����� %��� ������ ����� �������������� ����� ������ &������ �����

� �� �������� � ���� �� � � �� ��� � ����� !����� !" #$%� ������� ���� �������&'(#)*+ �������� �� � �

= ,�� ������ "��#

2--3 �1$�51"�02-�3 4�#54�5$$05�"��54���������������5#� # 6&����>>��(;�(��>������?��@���������������5#� #7

Page 3: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

“School’s out forever”

Alice Cooper

Page 4: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations
Page 5: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

Publications List of publications included in this thesis I Simonson, T., Carlsson, J., and Case, D. A. (2004) Proton binding

to proteins: pKa calculations with explicit and implicit solvent mod-els, J. Am. Chem. Soc. 126, 4167–4180.

II Thomaeus, A., Carlsson, J., Åqvist, J., and Widersten, M. (2007)

Active site of epoxide hydrolases revisited: A noncanonical residue in potato StEH1 promotes both formation and breakdown of the alkylenzyme intermediate, Biochemistry 46, 2466–2479.

III Carlsson, J., Andér, M., Nervall, M., and Åqvist, J. (2006) Contin-

uum solvation models in the linear interaction energy method, J. Phys. Chem. B 110, 12034–12041.

IV Carlsson, J., Almlöf, M., and Åqvist, J. (2007) Improving the accu-

racy of the linear interaction energy method for solvation free ener-gies, J. Chem. Theory Comput. 3, 2162–2175.

V Carlsson, J., Boukharta, L., and Åqvist, J. (2008) Combining dock-

ing, molecular dynamics and the linear interaction energy method to predict binding modes and affinities for non-nucleoside inhibitors to HIV-1 reverse transcriptase, J. Med. Chem. 51, 000–000.

VI Carlsson, J. and Åqvist, J. (2005) Absolute and relative entropies

from computer simulation with applications to ligand binding, J. Phys. Chem. B 109, 6448–6456.

VII Carlsson, J. and Åqvist, J. (2006) Calculations of solute and solvent

entropies from molecular dynamics simulations, Phys. Chem. Chem. Phys. 8, 5385–5395.

VIII Carlsson, J. and Åqvist, J. (2008) Absolute hydration entropies of

alkali metal ions from molecular dynamics simulations, manuscript.

Page 6: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

Additional publications i Henriksson, L. M., Unge, T., Carlsson, J., Åqvist, J., Mowbray, S.

L., and Jones, T. A. (2007) Structures of Mycobacterium tuberculo-sis 1-deoxy-D-xylulose-5-phosphate reductoisomerase provide new insights into catalysis, J. Biol. Chem. 282, 19905-19916.

Page 7: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

Contents

Introduction.....................................................................................................9

1. Computational methods ............................................................................10 1.1 Molecular mechanics models and force fields ...................................10 1.2 Molecular dynamics simulations........................................................11 1.3 Free energy perturbation and thermodynamic integration .................12 1.4 Molecular docking..............................................................................14

2. Explicit and implicit solvent models in free energy calculations .............15 2.1 Explicit and implicit solvent models in MM and MD calculations....15 2.2 Calculation of pKa values for ionizable residues in proteins..............17

2.2.1 Calculations of pKa values using explicit and implicit solvent MD simulations...........................................................................................18 2.2.2 Prediction of pKa shifts in the active site of epoxide hydrolase..21

2.3 Explicit and continuum solvent in binding free energy calculations..23

3. The linear interaction energy method .......................................................25 3.1 Hydration free energies ......................................................................26

3.1.1 The polar term.............................................................................27 3.1.2 The non-polar term .....................................................................30 3.1.3 The free energy of hydration ......................................................32

3.2 Binding free energies .........................................................................32 3.2.1 The polar term.............................................................................33 3.2.2 The non-polar term .....................................................................34 3.2.3 The free energy of binding .........................................................35 3.2.4 Continuum solvent models in the LIE method ...........................36 3.2.5 The LIE method in structure-based drug design.........................38

4. Estimating entropies from MD simulations ..............................................42 4.1 Solute entropies ..................................................................................43

4.1.1 Schlitter’s formula and quasiharmonic analysis .........................43 4.1.2 Estimation of binding entropies..................................................45

4.2 Solvent entropies ................................................................................48 4.2.1 Absolute hydration entropies of alkali metal ions from MD simulations...........................................................................................50

5. Future perspectives ...................................................................................52

Page 8: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

6. Summary in Swedish ................................................................................54

7. Acknowledgements...................................................................................56

8. References.................................................................................................57

Page 9: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

Abbreviations

MM Molecular mechanics MD Molecular dynamics SBC Spherical boundary conditions PBC Periodical boundary conditions FEP Free energy perturbation TI Thermodynamic integration GA Genetic algorithm PB Poisson-Boltzmann GB Generalized-Born SASA Solvent accessible surface area TSO Trans-stilbene oxide MCCE Multiconformation continuum electrostatics LIE Linear interaction energy LR Linear response rms Root mean square SBDD Structure-based drug design NNRTI Non-nucleoside reverse transcriptase inhibitor HIV Human immunodeficiency virus RT Reverse transcriptase DXR 1-deoxy-D-xylulose-5-phosphate reductoisomerase SF Schlitter’s formula QHA Quasiharmonic analysis CM Center of mass RR Restraint-release

Page 10: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations
Page 11: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

9

Introduction

Take a close look at something in your surroundings. Imagine being able to zoom in one billion times on this object – you have now entered the world explored in this thesis. At this level of detail, computer simulations can be used to study chemical processes involving single molecules. One of the most fascinating areas to which computer simulations can be applied is bio-chemistry, but modeling of biomolecules is a daunting challenge.

This thesis describes simulations based on molecular mechanics models, in which atoms are represented as spheres connected by springs. In spite of their simplicity, these models have successfully been applied to study pro-teins and nucleic acids. The availability of experimentally determined struc-tures of these macromolecules has made it possible to carry out atomistic simulations of biochemical processes, which can be used to explain the con-nection between the structure and function of biomolecules. This knowledge is not only of fundamental importance for the understanding of life itself, but can also be used in pharmaceutical applications.

In this thesis, two challenges in computational biochemistry are ad-dressed. The first is the representation of solvent in calculations on proteins. While these macromolecules often are the main focus of a simulation, the solvent in which they are immersed will significantly affect biochemical processes. For this reason, accurate representation of water in simulations of biomolecules is crucial to be able to make predictions for in vivo experi-ments. The second challenge is the calculation of thermodynamic properties for molecular association and solvation. Estimates of protein–ligand binding energies and solvation energies for small molecules are of significant interest in drug design, where accurate and efficient theoretical methods for predict-ing these properties would be extremely valuable.

The thesis is divided into three main parts. In the first, the accuracy of different types of solvent models is investigated for calculations of pKa val-ues of ionizable residues in proteins and protein–ligand binding free ener-gies. This is followed by a description of development and application of the linear interaction energy method for calculating solvation and protein–ligand binding free energies. In the last chapter, methods for decomposing free energies of binding and hydration into enthalpic and entropic components are presented.

Page 12: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

10

1. Computational methods

This chapter summarizes some of the theoretical and computational methods used in the work reported in this thesis.

1.1 Molecular mechanics models and force fields Molecular mechanics (MM) models are used to describe the potential energy for a system of interacting atoms. The typical form of a potential energy function ( potU ) is1,2

20

bonds

20

angles

( ) ( )

( )

pot

b

U

k r r i

k� � �

� �

� �

��

� �

20

impropers

dihedrals

nonbonded

( )

( ) ( )

1 cos( ) + ( )2

i j

ij

ii

k iii

kn iv

q qr

� �

� �

12 6nonbonded

( )

( )ij ij

ij ij

v

A Bvi

r r�

�� �� �� �

(1) where terms i–iv describe the bonded interactions in the system and v–vi represent the nonbonded interactions. The first three terms (i–iii) represent bond stretching, angle bending, and improper dihedral bending which are described by harmonic functions with force constants bk , k� , and k , re-spectively. The reference bond lengths, angles, and improper angles are 0r ,

0� , and 0 , respectively. Dihedral torsions (iv) are described using a series of periodic functions with barrier height k , periodicity n , and phase shift � .

Page 13: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

11

The nonbonded terms are evaluated for intermolecular interactions and intramolecular atom pairs that are separated by three or more bonds. The electrostatic interaction energy (v) for each atom pair is calculated from Cou-lomb’s law, where rij, qi and qj are the distance between the two atoms and their (partial) charges, respectively. The last sum represents the van der Waals interactions between atoms i and j, which are described by a Lennard-Jones potential with parameters Aij and Bij.

Before the MM model can be used in practice, the empirical parameters in eq. 1 ( bk , k� , k , k , 0r , 0� , 0 , n , � , iq , jq , ijA , and ijB ) must be determined for all atoms in the system. This calibration is, in most cases, based on experimental data and quantum chemistry calculations. Four exam-ples of popular biomolecular force fields are AMBER3, CHARMM4, OPLS-AA5 , and GROMOS6.

1.2 Molecular dynamics simulations To relate the potential energy of a system to experimentally measurable thermodynamic properties, thermally accessible configurations must be gen-erated. One approach is to simulate the system as a function of time using the molecular dynamics (MD) method.7

In the MD technique, the force on an atom ( iF ) at time t is calculated from the gradient ( i� ) of the MM potential function. Newton’s second law is then used to calculate the acceleration ( ia ) of the atom

1( ) i

i i poti i

t Um m

� � � �Fa (2)

where mi is the mass of the atom. From the acceleration, the velocity and position at time t t� � can be approximated from truncated Taylor expan-sions. A popular MD algorithm is the leap-frog Verlet method, in which positions and velocities are obtained from

12

1 12 2

( ) ( ) ( )

( ) ( ) ( )

i i i

i i i

t t t t t t

t t t t t t

� � � � � � �

� � � � � � �

r r v

v v a (3)

where ir , iv , and t� are the position, velocity, and time step, respectively. In MD simulations of biomolecules, the time step is usually set to 1–2 fs and initial velocities are assigned from the Maxwell-Boltzmann distribution.

Page 14: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

12

MD simulations are often used to calculate thermodynamic properties from simulations of relatively small systems. To be able to compare these to experiment, it is important to remove effects from the choice of boundaries for the system. Simulations of biomolecules in the aqueous phase are usually carried out using spherical (SBC) or periodical boundary conditions (PBC).

PBC simulations of biomolecules are often carried out in a box of water molecules. Using this technique, the system is surrounded by images of itself in all directions. To remove effects from the edges of the box, the molecules that are close to the boundary are allowed to interact with atoms in neighbor-ing systems. Furthermore, when a molecule crosses the boundary of the box, the same molecule enters from the image box on the other side, which con-serves the number of atoms in the system.

In SBC, the solutes are surrounded by a sphere of water molecules. This technique is particularly useful in simulations of biomolecules because the spherical system can be centered on the region of interest, e.g. a binding or an active site. Only atoms within a certain radius of the center of the sphere are explicitly included in the simulations. All atoms outside the sphere are restrained to their initial positions and are excluded from nonbonded interac-tions. This is significantly less computationally demanding than the use of PBC, where the entire biomolecule must be included in the system. How-ever, empirically calibrated restraints must be applied in SBC to reproduce the solvent density in the system and the polarization of water molecules close to the sphere boundary. In this thesis MD simulations were carried out with the programs Q8 and Amber79.

1.3 Free energy perturbation and thermodynamic integration The free energy perturbation (FEP) and thermodynamic integration (TI) techniques can be used to calculate free energy differences from molecular simulations.10,11

Using the FEP technique, the free energy difference between two states, A and B, is calculated from

( )( ) ln

B AU UkT

A B B AA

G U U G G kT e� �

� � � � � � (4)

where k is Boltzmann’s constant, T is the temperature and UA and UB are the potentials used to describe state A and B, respectively. The brackets, ...

A,

represent an ensemble average on potential A, which can be replaced by a

Page 15: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

13

time average according to the ergodic hypothesis. Thus, the free energy dif-ference can be calculated from an MD simulation on state A. However, for eq. 4 to give accurate results, states A and B must overlap in phase space. In other words, the MD simulation on state A must also sample relevant con-figurations for state B. When the FEP technique is applied to calculate free energies for chemical processes, this condition is seldom fulfilled. This prob-lem can be solved by taking advantage of the fact that the free energy is a state function, which means that it does not matter which path we use to calculate the change in free energy. The free energy can be calculated in several small steps by introducing a mixed potential

( ) (1 ) A BU U U� � �� � � (5) where � varies from 0 to 1. Using eq. 5 the transformation between state A and B can be divided into a number of discrete steps, n, and the total free energy difference can be calculated as the sum of these, i.e.

� � � �

1

11

( ) ( ) .n

A B m mm

G U U G U U� ��

��

� � � � �� (6) An alternative approach to calculate the free energy is given by the TI tech-nique, which is based on

1 1

0 0

( ) ( )( ) .A BG UG U U d d

� �� �� �

� �� � � �

� �� � (7) Using TI, the free energy difference between the two states A and B is calcu-lated by collecting the average value of ( ) /U � �� � for a number of � val-ues. Numerical methods are then used to approximate the integral in eq. 7.

The FEP and TI techniques have successfully been used to calculate relative binding and solvation free energies using thermodynamic cycles.10 The advantage of the FEP and TI techniques is that if sufficient sampling is achieved, the calculated result will be as accurate as the force field em-ployed. However, when FEP and TI are applied to biochemical systems, a large number of (unphysical) intermediate states have to be simulated, which often makes these methods very computationally demanding. In practice, this usually limits the application of the FEP and TI methods to perturbations involving only a few atoms. These problems have stimulated the develop-ment of a large number of approximate free energy methods, which will be further discussed in chapter 3.

Page 16: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

14

1.4 Molecular docking The goal of most molecular docking methods is to predict the binding mode of a small molecule to a macromolecular receptor.12 These methods are widely used to predict the structure of protein–ligand complexes. In this thesis, docking was applied to provide starting structures for MD simulations and free energy calculations.

The docking problem can be split into two different components. These are the generation of conformations of the receptor–ligand complex and scoring to discriminate between the correct and wrong binding poses. The ensemble of generated conformations must be sufficiently large to contain the true structure of the receptor–ligand complex and the scoring function should be able to assign the best score to this conformation. Examples of available docking programs are GOLD13,14, GLIDE15,16, DOCK17, and Auto-Dock18. In papers II and V, GOLD was used to perform small molecule docking to proteins.

In the generation of conformations, the receptor is often considered to be rigid. This reduces the search problem significantly, but it is still impossible to enumerate all possible binding poses for a ligand. Three efficient algo-rithms to explore conformations of the ligand in the biding site are based on genetic, Monte Carlo, and incremental construction search methods.19 The genetic algorithm (GA) used in GOLD is based on the concept of evolution. In the GA, the genes contain parameters describing translation, rotation, and the dihedral angles of the ligand in the binding site. In the first step a random population of individuals (ligand conformations) is generated and the fitness of each individual is calculated from a scoring function that is designed to identify favorable binding modes. The conformations with the best fitness are selected and a new population is generated based on the genes of these individuals. Some random mutations are also introduced in the next genera-tion to explore new potential solutions. The GA leads to an evolutionary pressure towards conformations with better fitness and after a number of generations the best individual is selected as the most likely binding mode.

Since the search algorithms generate a large number of possible recep-tor–ligand complexes, the scoring functions used in molecular docking are the result of a compromise between accuracy and speed. The three types of scoring functions used in molecular docking are physics based, empirical, and knowledge based.20 In the GOLD docking program, an empirical scoring function based on a sum of terms describing hydrogen bonds and van der Waals interactions is used.

Page 17: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

15

2. Explicit and implicit solvent models in free energy calculations

Water is necessary for all forms of life on earth. The human body consists of approximately 70 % water and a majority of all chemical processes in a cell occurs in the aqueous phase. For these reasons, accurate representation of the solvent properties of water is crucial in the modeling of biomolecules. Solva-tion models used in biomolecular modeling range from detailed descriptions, in which each water molecule is represented by an all-atom model, to more approximate approaches that reproduce the average properties of water.21 In papers I, II, and III explicit and implicit solvent models were applied to es-timate pKa values for ionizable residues in proteins and protein–ligand bind-ing free energies.

2.1 Explicit and implicit solvent models in MM and MD calculations The most rigorous approach to represent water in biomolecular simulations is to place the solute in a droplet or box filled with explicit water molecules, which are described using an all-atom model. In two widely used all-atom models, TIP3P22 and SPC23, a water molecule is represented by a van der Waals sphere centered on the oxygen and point charges located on each of the three atoms. The derivation of water models has been focused on repro-ducing some of the many important properties of liquid water such as ex-perimental energy, density, and radial distribution functions.1,24 A bio-molecular force field combined with explicit solvent can be used to simulate proteins in the aqueous phase. By sampling millions of different conforma-tions of a system, average energies can be calculated and used to estimate free energies for chemical processes.10

Explicit solvent representation requires that averaging over interactions with thousands of water molecules is carried out and long simulations may be necessary to obtain converged energies. In implicit solvent models the explicit solvent molecules are replaced by a continuous medium having the average properties of water.21,25-28 In this case, free energies in the aqueous

Page 18: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

16

phase can be calculated directly without sampling of the solvent degrees of freedom.

The energetic contribution from the solvent in implicit solvent models is often divided into a polar (electrostatic) and a non-polar (non-electrostatic) term. One description of the polar term was presented by Born, who derived a formula for the free energy of charging a spherical cavity in a continuous medium.29 Using the Born equation, the polar contribution to the hydration free energy of an ion can be estimated from

2 11

2polar

Borns

QGa �� �

� � � �� ��

(8) where Q is the charge of the ion, a is the ionic radius, and s� is the dielectric constant of water ( 80s� ! ). Remarkably, with only small adjustments of ionic radii, the Born equation can accurately predict ion hydration free ener-gies.30 Obviously, there is also a need to estimate hydration properties for molecules of any shape and charge distribution, e.g. proteins. This became possible through numerical solutions of the Poisson-Boltzmann (PB) equa-tion.31 For zero salt concentration the PB equation reduces to the Poisson equation

( ) ( ) 4 ( )� " #$� % � � �r r r (9) where ( )� r is the dielectric constant, ( )" r is the electrostatic potential, and

( )$ r is the charge density. The electrostatic potential can be calculated for a cubic lattice that is placed onto the solute. Pre-defined atomic radii, partial charges, and dielectric constants are first used to assign a charge density, dielectric constant, and an initial electrostatic potential to each grid point. The Poisson equation is then solved in each grid point until convergence is reached using finite difference methods.32,33 One of the major difficulties with using the PB method is that a dielectric constant for the solute has to be defined. Depending on the level of detail included in the modeling, internal dielectric constants between 1 and 8 are generally used for proteins.25,32 Al-though it is possible34, the PB method is generally not used in MD simula-tions of proteins because of the computational effort required to solve eq. 9. Instead, the technique is applied to a single or small ensemble of structures. The electrostatic contribution to a hydration free energy can be obtained from

� �80 11

2s spolar

PB i i ii

G q � �" "� �� � �� (10) where the sum is over all charges, iq , of the solute and 80s

i�" � and 1s

i�" � are

the electrostatic potential at atom i in water and vacuum, respectively. The

Page 19: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

17

PB model can also be used to calculate pKa values and to estimate electro-static contributions to molecular association.33

A less computationally demanding alternative to the PB method is based on the Generalized-Born (GB) equation35, which is an empirical combination of Coulomb’s law and the Born formula.26,27,36 The GB expression for the electrostatic contribution to the free energy of hydration is

2 /(4 )21 1

1 112 ij i j

i jpolarGB

ri jsij i j

q qG

r e & &� & & �� �

� �� � � �� �

� ��� (11)

where iq and i& are the (partial) charge and the effective Born radius of atom i, respectively. The effective Born radius can be interpreted as the dis-tance from an atom to the boundary of the dielectric medium. These are typically calculated using an analytical formula, which is often parameter-ized to reproduce results obtained using the PB equation.37-39 The analytical form of eq. 11 also makes it feasible to use the GB model as solvent in MD simulations.40,41

The non-polar term in implicit solvent descriptions represents free energy contributions arising from cavitation and solute–solvent van der Waals inter-actions. Since the free energy of hydration for hydrocarbons has been shown to depend linearly on the size of the solute42, the non-polar term is often es-timated from

non polar

SASA SASAG A b'�� � � (12)

where SASAA is the solvent accessible surface area (SASA) of the solute.43 The constants ' and b are parameterized to reproduce experimental hydra-tion free energies of non-polar molecules. Approximations of the non-polar term will be further discussed in chapter 3.

2.2 Calculation of pKa values for ionizable residues in proteins Ionizable residues in proteins play key roles in fundamental biochemical processes such as enzyme catalysis, substrate binding, and protein stability. For this reason, accurate estimates of pKa values are of great interest because they can give detailed information about the protonation state of an ionizable residue in a protein. Furthermore, because pKa values can be measured ex-perimentally, they also provide a valuable benchmark for evaluation of force field parameters and approximate models for calculating free energies.25

Page 20: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

18

The pKa value of an ionizable amino acid depends on the local environ-ment in the protein and it can be shifted by several units compared to the standard values in solution.44 For example, an ionizable residue on the sur-face of a protein will have a pKa value close to that of the free amino acid in solution while the same residue buried in the non-polar interior of a protein will have a large shift in pKa because it is unfavorable to create a charge in such an environment. Calculation of the free energy of deprotonation for an amino acid poses many challenges.45,46 The release of a proton from a resi-due involves large changes in electrostatic energy and may lead to both local and global conformational changes. The pKa of a residue can also be coupled to the protonation state of nearby ionizable residues, which has to be taken into account in the calculations. For example, the catalytic aspartic acids in the HIV-protease dimer, which are in close contact with each other, have pKa values of approximately 6.8 and 3.3.47 This reflects that deprotonation of one of the aspartic acids stabilizes the protonated form of the other. To accu-rately calculate the pKa value in such cases, one free energy calculation has to be carried out for each possible combination of protonation states. This may be computationally demanding in many cases. In papers I and II, ex-plicit and implicit solvent models were used together with MD simulations to calculate pKa values of ionizable residues in proteins.

2.2.1 Calculations of pKa values using explicit and implicit sol-vent MD simulations In paper I, explicit and GB solvent MD simulations in combination with the TI technique were used to calculate pKa values for three aspartic acid resi-dues. The pKa values were calculated from the free energies corresponding to the upper ( protG� ) and lower ( watG� ) horizontal legs of the thermody-namic cycle in Figure 1, which yields ,

1p p2.303a a watK K G

kT� � �� (13)

where ,p a watK is the pKa value of the amino acid in solution, k is Boltzmann’s constant, T is the temperature, and prot watG G G�� � � � � .48

The MD simulations were carried out using the AMBER3 and CHARMM4 force fields in both explicit (TIP3P22) and GB (GB/OBC39,41 and GB/ACE37,40) solvents. To provide a challenging test, two aspartic acid residues with extreme pKa shifts compared to the value in solution ( ,p a watK ! 4.0) were chosen. Asp26 of the oxidized form of Escherichia coli thioredoxin is positioned in a hydrophobic cavity and has a large pKa shift of

Page 21: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

19

+3.5 units compared to the free amino acid in solution49 while Asp14 of human pancreatic ribonuclease A has an extreme negative shift of �2.0 units50. The third residue, Asp20 of Escherichia coli thioredoxin, is solvent exposed51 and was expected to have a pKa similar to that of aspartic acid in solution.

Figure 1. Thermodynamic cycle for calculating the pKa value of an aspartic acid residue in a protein relative to the free residue in solution.

The results from the TI calculations in explicit and implicit solvent are

presented in Table 1. The calculated pKa values of Asp20 are close to that of aspartic acid in solution in both explicit and GB solvent. For the two more challenging cases, the directions of the pKa shifts were correctly predicted with the AMBER force field, but only for Asp26 with CHARMM. For Asp26, the calculated pKa values are several units too large with both force fields. A possible explanation to this result is that, because the structure of thioredoxin was determined at pH 4.1, there may be problems to accurately sample the ionized state of Asp26 in the simulations. For example, longer simulations may be required for an appropriate number of explicit water molecules to penetrate the cavity and relax the charge on the carboxylate group. This would lead to an underestimation of the stability of the ionized state and thereby increase the calculated pKa shift.

Page 22: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

20

The MD simulations in GB solvent were 1–2 orders of magnitude faster than those carried out in explicit solvent. Since the solvent is treated implic-itly in this case, there is no need for long equilibrations of the solvent and the protein will also explore conformational space more quickly. An encourag-ing improvement with experiment compared to the explicit solvent calcula-tions was obtained in these simulations for all three cases.

Table 1. Calculated pKa values using explicit and implicit solvent simula-tions in combination with the TI technique. Experimental pKa values are also given.

Explicit solvent Implicit solvent Expt. AMBER CHARMM AMBER CHARMM Asp26 7.5 10.7 13.1 9.2 12.1 Asp14 2.0 3.3 4.8 2.8a n.d.b Asp20 4.0 5.2 4.7 4.4 4.4

a The average pKa value for simulations carried out with all carboxylates (except Asp14) unprotonated or protonated. b Not determined

The MD simulations clearly showed the importance of including confor-mational flexibility in free energy calculations on proteins. For example, as a response to the ionization of Asp26, a hydrogen bond between the amine group of nearby Lys57 and the carboxylate of Asp26 was formed in the ex-plicit solvent simulations. Due to the small probability of this state at the experimental conditions, this salt bridge has not been observed in crystallo-graphic structures. Methods that estimate pKa values based on fixed protein structures have been shown to have difficulties to predict free energy changes associated with such conformational changes.45,52-54

One difficulty with the use of long MD simulations is that inaccuracies in the force field parameters and approximations made in the algorithms (e.g. cut-offs and time step) may cause the protein structure to slowly drift away from the experimental structure. On one hand, several nanoseconds of simu-lation may be required for relevant conformational changes to occur and to obtain converged free energies. On the other hand, this will not necessarily improve the results because irrelevant parts of conformational space may be explored. In the explicit solvent simulations, the protein remained close to the initial crystallographic structure throughout the simulations. The calcula-tions also gave detailed information about conformational changes and water reorganization in response to ionization of the aspartic acid residues. Intro-ducing the implicit solvent approximation led to significantly larger devia-tions from the crystallographic structures and important local interactions found in the explicit solvent simulations were missing in implicit solvent.

Page 23: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

21

For example, the salt bridge between ionized Asp26 and Lys57 was formed in all simulations except in those with the GB/ACE solvent model. These results agree with several other studies that have shown that GB solvent simulations can not always reproduce explicit solvent calculations or ex-perimentally determined structures.55,56 These observations suggest that fur-ther development of the implicit solvent MD technique is necessary in order to provide accurate and reliable conformational sampling.

MD simulations are computationally demanding and the presented calcu-lations required several weeks of computation time. Still, the free energies are not fully converged and the presented results deviate from experiment by more than one pKa unit in the challenging cases. For more accurate predic-tions of pKa values, continuum solvent approaches that include side chain flexibility and have been tuned to reproduce experiment may be a better choice. These methods often perform better than the techniques presented here and to a considerably lower computational cost.57,58 The strength of more rigorous free energy calculations is rather to provide a benchmark for development of accurate implicit solvent models, force fields, and approxi-mate free energy methods.

2.2.2 Prediction of pKa shifts in the active site of epoxide hy-drolase Epoxide hydrolases catalyze the hydrolysis of epoxides to diols. The active site of this enzyme is shown in Figure 2. In the first step of the catalytic mechanism, one of the carboxylate oxygens of Asp105 attacks a carbon in the epoxide ring which yields an anionic alkylenzyme intermediate. In the second step, the general base His300 activates a water molecule, which at-tacks the alkylenzyme intermediate and this leads to the creation of the diol product via a tetrahedral intermediate.59 The aim of paper II was to further investigate the mechanism of epoxide hydrolase and to examine a possible role of active site residue Glu35.

Experimental studies of epoxide hydrolase have shown that the Glu35Gln mutant displays an inverted pH dependence of kcat compared to the wild type enzyme for the substrate R,R-trans-stilbene oxide (R,R-TSO). At pH 7, where the wild type has its maximal catalytic rate, the measured rate for the Glu35Gln mutant instead reached its minimum. Using kinetic models, it was found that to explain the inverted pH behavior for R,R-TSO, the acidity of His300 must be increased in the mutant. The models also revealed that His300 may act both as a general base, by activating the catalytic water, and as the acid that protonates the leaving group diol. These hypotheses were examined by estimating the pKa of His300 using a combination of explicit solvent MD simulations and PB continuum electrostatics calculations.

Page 24: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

22

Figure 2. TSO and meso-hydrobenzoin.(left) Selected active site residues (PDBID:2CJP60) of potato epoxide hydrolase (StEH1) and the docked conformation of R,R-TSO.(right)

R,R-TSO was docked into a recently solved crystallographic structure of

the enzyme.60 Sampling of the different states of the reaction (free enzyme, enzyme–substrate complex, alkylenzyme, and tetrahedral intermediate) was carried out using explicit solvent MD simulations. To reduce structural bias towards any of the protonation states of His300, simulations of both the charged and neutral forms of the residue were carried out. A series of snap-shots from the simulations were then used to estimate pKa values using the multiconformation continuum electrostatics (MCCE) method.54,57 The MCCE method is based on the PB solvent model and includes partial side chain flexibility for polar residues. The use of side chain flexibility and en-sembles of structures in pKa calculations have been shown to significantly improve agreement with experiments.57,61,62

The calculated pKa values of His300 in wild type epoxide hydrolase for the hydrolysis of R,R-TSO are shown in Table 2. Compared to the pKa val-ues used in the kinetic simulations, the calculated values are several units too large. However, the shifts relative to the enzyme–substrate complex are qualitatively in agreement with the kinetic simulations. In the model based on the kinetic simulations, the pKa of His300 first decreases by 0.8 units for the alkylenzyme and then increases by the same amount in the tetrahedral intermediate. In the MCCE calculations, where the pKa of His300 is first reduced by 2.7 units and then increases by almost the same amount, the same trend is observed. This supports the hypothesis that His300 can act both as a base and an acid in the reaction.

Page 25: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

23

Table 2. pKa values of His300 in wild type epoxide hydrolase calculated with the MCCE method and from kinetic simulations for the enzyme–substrate complex ( E S% ), alkylenzyme ( E alkyl� ), and tetrahedral intermediate ( TetI ).

E S% E alkyl� TetI

MCCE 10.7 8.0 10.3 Kinetic simulations 6.3 5.5 6.3

Kinetic simulations of the pH dependence of the reaction predicted that to

reproduce the inverted pH dependence for R,R-TSO in the mutated form of epxide hydrolase, the pKa of His300 must be shifted by �2.0, �2.3, and �2.9 units for the enzyme–substrate complex, alkylenzyme, and tetrahedral inter-mediate, respectively. In agreement with these simulations, the MCCE cal-culations resulted in pKa shifts equal to �2.3, �3.1, and �2.9 for the same steps of the reaction. These calculations and MD simulations of the different states indicate that a likely role of Glu35 is to orient the catalytic water and to maintain optimal acid–base characteristics of His300.

2.3 Explicit and continuum solvent in binding free en-ergy calculations Continuum electrostatics methods have become widely used to estimate free energy changes in biochemical systems. In FEP as well as more approximate free energy methods, the use of implicit solvent instead of explicit solvent is becoming an increasingly popular approach to estimate solvent contributions to protein–ligand binding free energies.25-27,33

The ability of continuum solvent models to reproduce hydration free en-ergies for small molecules has been demonstrated repeatedly35,38,63, but to predict the electrostatic interaction energy with a handful of solvent mole-cules in a binding site is a considerably more challenging problem. Are the PB and GB solvent models sufficiently accurate for this type of application? This question was addressed in paper III by calculating the difference in ligand–water interaction energy between the bound and free states of ligands, el

l watU �� , using explicit, PB, and GB solvent models. Explicit sol-vent MD simulations were carried out for the bound and free state for a se-ries of eight inhibitors to the malarial aspartic protease plasmepsin II. An ensemble of structures from these simulations was saved for post-processing in PB and GB continuum solvent. The calculated energies were also used to estimate binding free energies with the linear interaction energy method, which will be presented in the following chapter.

The PB model reproduced the explicit solvent electrostatic desolvation energies very well and the results are shown in Figure 3A. Analysis of the

Page 26: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

24

inhibitor with the largest deviation from the explicit solvent interaction en-ergy showed that there were several conserved water molecules in the bind-ing pocket in this case. This result indicates that there are limitations to the accuracy of the continuum solvent approach in cases where specific water molecules may play an important role.64

The GROMOS876 force field charges and radii were used directly in the PB calculations. The use of a different radii set (PARSE63) showed that ad-justment of the radii to be compatible with the applied force field is neces-sary to predict absolute binding free energies. Such radii sets have already been presented for the CHARMM65 and AMBER66 force fields.

Figure 3. Electrostatic desolvation energies ( el

l watU �� ) for the ligands from the PB (A) and GB (B) solvation models compared to explicit solvent representation.

The GB method did not reproduce the explicit solvent calculations as well as the PB model. As shown in Figure 3B, the overall trend for the inhibitors is reproduced but because small differences in affinity often are sought in binding free energy calculations, these errors will have a large impact on the accuracy of the linear interaction energy method. The differences between the two continuum models were somewhat surprising considering that GB models are often benchmarked against hydration free energies calculated with the PB model. A more detailed comparison between the PB and GB hydration energies showed that these differences could partially be explained by the fact that the effective Born radii, which are the cornerstone of the GB approach, were underestimated.67 The effect of this problem also becomes more pronounced when interaction energies instead of total hydration ener-gies are calculated.68 More recently developed GB models have improved methods for predicting the effective Born radii and agree significantly better with the PB model.69 These methods are likely to improve the results and perhaps make it possible to use the GB method in binding free energy calcu-lations.

Page 27: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

25

3. The linear interaction energy method

In the eighties, when the FEP and TI techniques were used to calculate free energies for the first time, a new era in computational chemistry began.10 Several early applications indicated that hydration free energies, pKa values, and relative protein–ligand binding affinities could be calculated with an impressive accuracy and precision. However, it soon became clear that there were also several problems with using the FEP and TI techniques.11,70 In particular, significantly longer simulations than initially expected were re-quired to obtain reliable results. The annihilation and creation of atoms in simulations often lead to severe convergence errors in FEP and TI calcula-tions. For this reason, these techniques are often restricted to the study of free energy differences between relatively similar systems. The transforma-tion of one molecule to another must also be divided into a large number of steps, which makes the methods computationally demanding. Development of approaches to remedy these problems, e.g. the soft-core method71,72, con-tinues to increase the range of applications, in which the FEP and TI tech-niques can be used. However, for many types of problems these methods are still too computationally expensive to be useful.

Parallel to the advancements made for the FEP and TI techniques, there has also been an enormous interest for more approximate and less computa-tionally demanding methods for calculating free energies. In particular, a large variety of approaches for calculating hydration and receptor–ligand binding free energies have been developed.11,20,73 The most rapid methods predict free energies of binding or hydration from single structures using a combination of empirical terms, which have been parameterized to repro-duce experiments. Another class of methods is based on more advanced ap-proximations of the FEP formula and can be used to estimate free energies from only a few simulations, e.g. the MM-PB/SA74,75 and linear interaction energy (LIE)76 methods. In this chapter the derivation, development, and application of the LIE method will be described (papers III, IV, and V).

Page 28: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

26

3.1 Hydration free energies The LIE method was derived by Åqvist et al.76 and was first applied to esti-mate solvation free energies by Jorgensen and co-workers.77,78 The free en-ergy of solvation was derived by partitioning the free energy transfer from the gas phase to the solvent into a polar and non-polar term. The thermody-namic cycle in Figure 4 describes these contributions for calculation of hy-dration free energies: Starting from the solute in the gas phase, the electro-static solute–solvent interactions are first turned off ( polar

gG� ). The free en-ergy of this step is identical to zero because, in this case, there are no sur-rounding solvent molecules. In the resulting state, there are only van der Waals interactions between the solute and its surroundings. In the second step, the solute is transferred from the gas phase to the aqueous phase ( non polar

hydG �� ). The electrostatic solute–solvent interactions are then turned on again ( polar

wG� ), which completes the thermodynamic cycle. Hence, the hy-dration free energy can be estimated from .polar polar non polar polar non polar

hyd w g hyd hyd hydG G G G G G� �� � � � � � � � �� � � (14) The aim of the LIE method is to accurately estimate the polar and non-polar terms from as few simulations as possible.

Figure 4. The thermodynamic cycle used to calculate hydration free energies with the LR model based on eq. 16. The double-headed black and white arrows represent solute–solvent and solute–solute electrostatic interactions, respectively.

Page 29: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

27

3.1.1 The polar term The polar term is estimated using the linear response (LR) approximation, which predicts that polar

wG� can be calculated from the endpoints of the right vertical leg in Figure 4. According to the LR approximation76,79, the free energy associated with turning on the electrostatic solute–solvent interac-tions is

� �.w wpolar polar el el

hyd w r s r sA BG G U U( � ��� � � � � (15)

In eq. 15, ( is a scaling factor and

welr s B

U � represents the average electro-static solute–solvent (r–s) interaction energy from a simulation of the solute in water (state B).

welr s A

U � is extracted from a simulation in which the elec-trostatic solute–solvent interactions are turned off (state A), but the electro-static interaction energy is estimated as if the partial charges of the solute were present. Since the solvent dipoles in state A are, on average, oriented randomly around the solute, this term should be close to zero. Hence, the polar term can be estimated from a single simulation of the solute in water, which yields

.

wpolar elhyd r s B

G U( ��� � (16) The LR approximation predicts that 1

2( � . Åqvist and Hansson (ÅH) inves-tigated this approximation for molecules in polar solvents using the FEP technique.79 It was found that for neutral molecules, ( values lower than 1

2 were necessary to reproduce FEP calculations in water. They also proposed that the FEP derived coefficients could be used to improve the accuracy of simplified free energy calculations. In a subsequent paper, a method to pre-dict ( values for small molecules was presented (Table 3).80 In this scheme,

12( � is used for all ionic molecules and a ( value of 0.43 is assigned to

neutral compounds without any hydroxyl groups. For neutral molecules with one or more hydroxyl groups the ( value is set to 0.37 and 0.33, respec-tively.

In paper IV, an extensive study of the ( coefficient was carried out. Compared to ÅH, calculations on a larger set of different molecules were performed with the aim to develop an improved method to predict ( for neutral and ionic molecules. An alternative derivation of the LR approxima-tion was also proposed and this model is based on the thermodynamic cycle in Figure 5. In this case, both the intra- and intermolecular electrostatic in-teractions are turned on in the gas and aqueous phase. A LR approximation for this cycle yields

� �wpolar el el

hyd r s r rBG U U( � ��� � � � (17)

Page 30: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

28

where el

r rU �� represents the difference in solute–solute (r–r) electrostatic energy between the water and the gas phase (

w gel elr r r rB B

U U� �� ). This for-mula was expected to improve the accuracy of the LR approximation for flexible molecules, which was not investigated by ÅH.

Figure 5. The thermodynamic cycle used to estimate hydration free energies with the LR model based on eq. 17. The double-headed black and white arrows represent solute–solvent and solute–solute electrostatic interactions, respectively.

The FEP technique was used to calculate ( values according to eq. 17 for 211 small rigid molecules representing a series of chemical groups (alcohols, amides, amines, ketones, aldehydes, thiols, ethers, esters, carboxylic acids, nitriles, nitros, sulfides, anions, and cations). Since relatively rigid molecules were used in this training set, the calculated ( values should not differ sig-nificantly from those that would have been obtained using the thermody-namic cycle in Figure 4. In agreement with the observations of ÅH, devia-tions from the LR approximation were found for neutral molecules. Most neutral molecules had a ( value close to 0.43, but for many of the hydrogen bond donating chemical groups lower values of the coefficient were ob-tained. For ionic molecules, systematic deviations depending on the sign of their net charge were found ( 0 0.52Q( ) ! and 0 0.45Q( * ! ). These deviations do not originate from deviations from LR, but rather because the assumption made from eq. 15 to eq. 16, does not hold, i.e. 0

welr s A

U � + .79 From these calculations, ( coefficients for molecules containing only one of the inves-tigated chemical groups were derived. The results for this model are pre-

Page 31: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

29

sented in Table 3, where the ( value for a specific group, i, can be calcu-lated from 0i i( ( (� � � .

Table 3. Comparison between four different models for predicting the ( coefficient and the rms errors (kcal/mol) for the training and test sets (zwitterions excluded). Q denotes the net charge of the molecule.

Model rms (training)

rms (test)

Linear response

1

2( �

1.6

3.7

Carlson et al.77

0.48( �

1.5

3.7

Hansson et al. 80

aOH

aOH

aOH

0 0.5

0.43 00 0.37 1

0.33 2

Q

nQ n

n

(

(

(

(

) �

, � �-

� � �.- � /0

1.2

3.3

Paper IV (model E)

0

i ii

ii

w

w

(( (

�� �

��

where

0

1o o

2o

3

4

5

6

7

(alcohols)

(neutral 1 ,2 -amines)

(1 -amides)(COOH)(anions)(cations)(other)

((

(

(((((

��

0.430.060.040.020.03

0.020.090.00

�� �� �� �� ����

1

2

3

4

5

6

7

, 1, 1, 1, 1, 11, 11, 1

wwwwwww

���

��

0.3

1.3

a The number of hydroxyl groups in the molecule. ÅH did not specifically investigate how ( is influenced by the presence

of several chemical groups in a molecule. Furthermore, the scheme proposed by Hansson et al. uses a somewhat counter-intuitive approach to solve this problem. For example, a ( value equal to 0.37 is assigned to neutral mole-cules containing one hydroxyl group irrespective of any other chemical groups in the molecule. This is based on the observation that ethanol has a ( coefficient equal to 0.37.80 However, if the same molecule consists of

Page 32: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

30

several other polar chemical groups, the ( value would be affected by all of these and result in a different optimal coefficient. In paper IV this was taken into account by deriving a weighting scheme that included the effect of all chemical groups in a molecule

0

i ii

ii

w

w

(( (

�� �

��

(18)

where the 0( , wi , and i(� values for different chemical groups are given in Table 3. Four of the models investigated in paper IV for predicting ( values are presented in Table 3. The first two models are the original LR approxi-mation ( 1

2( � ) and the use of a single ( value parameterized on the train-ing set77. These are compared to the model developed by Hansson et al.80 and the scheme developed in paper IV (model E). The models were com-pared both for the parameterization set of 211 small molecules and a test set with 361 molecules, which are more flexible and contain several chemical groups. It is clear that the scheme developed in paper IV outperforms the three other models and significantly improves the accuracy of the LR ap-proximation. Flexible zwitterions are a problematic group of molecules for which eq. 18 can not predict ( values. For these molecules, the large free energy changes arising from strong intramolecular interactions are not accu-rately reproduced by the LR approximation.

The accuracy of the two different versions of the LR approximation pre-sented in paper IV (eq. 16 and eq. 17) was also investigated on a subset of the training and test sets. As expected, there were no differences between eqs. 16 and 17 for the training set because these molecules are relatively rigid. For the test set, which contains many flexible molecules, it was found that eq. 17 is significantly more accurate for estimating the polar contribu-tion to hydration free energies.

3.1.2 The non-polar term The non-polar term in Figure 4, non polar

hydG �� , represents the free energy con-tribution for transferring a molecule with no electrostatic intermolecular interactions from the gas phase to the aqueous phase. In the alternative cycle (Figure 5), it represents the corresponding free energy for a solute with both intra- and intermolecular electrostatic interactions turned off. For non-polar molecules, the free energy of hydration should originate almost exclusively from non polar

hydG �� . For such molecules, e.g. n-alkanes, there are also experi-mentally observed linear relationships between solute size and hydration free

Page 33: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

31

energies, which suggests that size-measures can be used to predict the non-polar term.42

In the original LIE method, the solute–solvent van der Waals energy of state B,

wvdWr s B

U � , was identified as a useful size-measure, which yields the following expression for the non-polar term

wnon polar vdW vdW vdWhyd w r s wB

G U& '��� � � (19)

where vdW

w& and vdWw' are empirical scaling factors. The constant vdW

w' can be interpreted as the contribution from inserting a point particle in the sol-vent. The first term then corresponds to the free energy of increasing the size of this particle.81 In paper IV, the vdW

w& and vdWw' coefficients were param-

eterized on 184 experimental hydration free energies. In this optimization, the total calculated free energies of hydration were estimated by combining eq. 17 and eq. 19 with ( values according to paper IV (model E). The opti-mal values of vdW

w& and vdWw' were found to be 0.01 and 1.2 kcal/mol, re-

spectively (root-mean-square (rms) error 1.1 kcal/mol). Surprisingly, the vdWw& coefficient is close to zero in this model. In fact, removing the first

term in eq. 19 did not significantly reduce the agreement with experiment. Similar results were obtained for other size-measures such as the surface areas of the solutes. These observations reflect that a majority of the com-pounds in the training set are similar in size and that errors in the polar term introduce noise in the optimization. To obtain more reliable coefficients, the parameterization could be based on experimental data for non-polar mole-cules or on non polar

hydG �� values calculated using the FEP technique. Another type of expression for the non-polar term combines the molecu-

lar surface area of the solute with the solute–solvent van der Waals energy of state A, which yields

wnon polar vdWhyd s MS r s A

G A U'��� � � (20)

where MSA is the molecular surface area of the solute and s' is an empirical scaling factor. A physical interpretation of this formula is that the free en-ergy of cavity formation can be estimated from the surface area of the solute scaled by the surface tension of the solvent. After the cavity has been formed, the solute–solvent van der Waals interactions are turned on and, given that the process does not perturb the solvent structure, this yields a free energy contribution of

wvdWr s A

U � .82,83 Optimization of eq. 20, as described above, yielded 0.09s' � kcal/mol Å2 (rms error 0.8 kcal/mol). Since the derived value of this coefficient is quite close to the experimental macro-scopic surface tension for water (0.105 kcal/mol Å2), this equation can be used without any empirical parameters.

Page 34: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

32

3.1.3 The free energy of hydration Two different expressions for predicting hydration free energies can be formed by combining eq. 17 with either eq. 19 or eq. 20, which yields � �w wvdW vdW el el vdW

hyd w r s r s r r wB BG U U U& ( '� � �� � � � � � (21)

and � � .

w wel el vdWhyd r s r r s MS r sB A

G U U A U( '� � �� � � � � � (22) By using these equations, absolute hydration free energies for neutral mole-cules can be reproduced with an average unsigned error of below 1 kcal/mol. The use of eq. 21 requires only that simulations of the solute in the gas and aqueous phase are carried out. For rigid molecules, the hydration free energy can even be calculated from a single simulation ( 0el

r rU �� ! ). Eq. 22 was the most accurate expression for absolute hydration free energies investi-gated in paper IV, but an additional simulation of the uncharged solute in water (state A ) has to be carried out in this case.

Continuum solvent models have been shown to reproduce hydration free energies for small molecules with similar accuracy as the method presented here.35,38,63 However, it is likely that eqs. 21 and 22 are more accurate for flexible molecules, for which conformational sampling can be important. Furthermore, if a method can reproduce solvation free energies in several different media, it is likely that the same approach can be used to study more complex biochemical problems. In the following section, the methodology for solvation energies described in paper IV will be used to estimate protein–ligand binding free energies.

3.2 Binding free energies The LIE method is based on the idea that receptor–ligand binding free ener-gies can be estimated as a free energy of transfer between water and a recep-tor-like environment.76 The thermodynamic cycle used to derive an expres-sion for the binding free energy is shown in Figure 6, which yields

.polar polar non polar

bind b f bindG G G G �� � � � � � �� (23) Hence, receptor–ligand binding free energies can, in analogy with solvation free energies, be calculated from a polar ( polar polar polar

bind b fG G G�� � � � � ) and a non-polar ( non polar

bindG ��� ) term. To estimate binding free energies with the

Page 35: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

33

FEP technique, a large number of simulations of intermediate (unphysical) states between the endpoints in Figure 6 have to be carried out. In contrast, the LIE method attempts to estimate the terms in eq. 23 only from the upper corners of the thermodynamic cycle, i.e. from simulations of the free ligand in solution and bound to the receptor.

Figure 6. The thermodynamic cycle used to estimate binding free energies with the LIE model based on eq. 23. The double-headed black and white arrows represent ligand–surrounding and ligand–ligand electrostatic interactions, respectively.

3.2.1 The polar term

The left vertical leg of the thermodynamic cycle, polarfG� , is identical to the

polar term estimated for hydration free energies (eq. 16). The same approxi-mation is assumed to be valid for the right vertical leg, which corresponds to the free energy of turning on the electrostatic ligand–surrounding interac-tions in the receptor ( polar

bG� ). This yields � �

b fpolar polar polar el el elbind b f l s l s l sB B

G G G U U U( (� � ��� � � � � � � � � (24) where

bell s B

U � and fel

l s BU � are the average ligand–surrounding (l–s) electro-

static energies in the bound (b) and free (f) states, respectively. Also note that the electrostatic interactions with both the receptor and the solvent are

Page 36: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

34

included in bel

l s BU � . In the first version of the LIE method, the LR approxi-

mation was assumed to be exactly valid and 12( � was used for both the

receptor and water environment. In a refinement of the method, it was shown that the ( values derived from FEP calculations by ÅH yielded signifi-cantly better agreement with experimental binding affinities and these are now used in the standard parameterization.80 How accurate the LR approxi-mation is in the receptor environment is not completely understood and de-serves further investigation.

3.2.2 The non-polar term In analogy to hydration free energies, the non-polar contribution to binding is estimated from the intermolecular van der Waals energy of the ligand. Solvation free energies of non-polar molecules depend linearly on the size of the solute in many different solvents42,84, which indicates that this approxi-mation might also be applicable to a receptor-like environment. Assuming that both the non-polar term and the change in ligand–surrounding van der Waals energy depend linearly on a size-measure (1 ) yields

non polar

bind

b fvdW vdW vdWl s l s l sB B

G a b

U U U c d

1

1

� � �

,�� � �-.� � � � �-0

(25) where a, b, c, and d are empirical scaling factors and

bvdWl s B

U � and fvdW

l s BU �

are extracted from simulations of the bound and the free state of the ligand, respectively. Combining the two expressions in eq. 25 yields non polar vdW vdW

bind l s l sa adG U b Uc c

& '�� �

� ��� � � � � � � �� ��

(26) where /a c& � and /b ad c' � � . If a and b were derived from experimen-tal solvation free energies in water and in a receptor-like environment, & and ' could be determined from these relations. The LIE parameters would then not rely on any experimental free energies of binding. However, this has been avoided for several reasons. The assumption that non polar

bindG ��� is equal to the free energy of transfer between two solvents is not completely valid. For example, the translational and rotational entropy loss for the ligand will cancel for free energies of transfer, but not for binding affinities. It would also be rather difficult, if not impossible, to find a suitable solvent to represent a receptor binding site. Instead, the non-polar contribution was parameterized on experimental free energies of binding. For this reason, the value of & simply reflects the contributions to the binding affinity that cor-

Page 37: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

35

relate with vdWl sU �� and � represents the terms that can be considered to be

constant.

3.2.3 The free energy of binding The total free energy of binding, which is obtained by combining eqs. 24 and 26, is vdW el

bind l s l sG U U& ( '� �� � � � � � (27) where ( is determined according to Hansson et al., 0.18& � , and 0' � kcal/mol in the standard parameterization.80 Although there has been some debate about the system transferability of the standard parameterization of the LIE method85, this model has been applied to several protein–ligand complexes with impressive results.11,80,86,87

Almlöf et al. investigated the force field dependence of & , ( and ' for a set of nine ligands to P450cam.88 This study showed that the same set of coefficients reproduced experiment with three different biomolecular force fields. The standard values of & and ( predicted the relative binding affini-ties very well, but a ' value equal to about �4 kcal/mol was necessary to reproduce absolute binding free energies. This led to the conclusion that ' is system dependent. Hence, the standard parameterization of the LIE method can only be used to estimate relative binding free energies. A more detailed analysis of different receptor–ligand complexes showed that the value of ' is correlated with the hydrophobicity of the binding site, which will be useful in the derivation of a method to predict this constant.

In some studies, the value of & has also been found to vary for different systems, but this does not necessarily mean that the coefficient is system dependent.85 Since the value of & reflects a number of contributions to the binding free energy and the parameterizations are often based on just a few experimental binding free energies, it is difficult to interpret and identify the sources of the differences. In some cases, ' has been set to zero in the opti-mization of eq. 27, which affects the value of & if this assumption is not valid. Differences may actually also originate from inaccuracies in the LR approximation for binding sites. Another source of error is that important interactions in the receptor–ligand complex may be missing in the modeling in cases where the binding modes are not known. Sometimes it has been possible to identify the reason for a system specific & coefficient. For ex-ample, in a study of protein–protein complexes, significantly higher values of & were required to reproduce experimental data.89 This effect could par-tially be attributed to the neglect of the

bell s A

U � term. It is clear that to re-solve the question about the value and system dependence of & , calculations

Page 38: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

36

on large datasets with high resolution structures of protein–ligand complexes in combination with reliable experimental binding free energies are neces-sary. If the & coefficient turns out to be system dependent, which seems likely, the LIE equation will still be a useful method to predict binding free energies, but it will have to be reparameterized for each receptor.90-92

From the derivation of the polar terms for binding and hydration free en-ergies, a series of alternative forms of the LIE equation can be identified. Two of these were investigated in paper V. The ( values used in the stan-dard parameterization of the LIE method are based on FEP calculations in water, but there is no reason to expect that the same coefficients should be optimal for the receptor environment. In fact, it can be argued that the ( value for the bound state should be larger than for the free state in some cases.93 Introducing separate ( values in eq. 27 yields

,

b fvdW el elbind l s b l s f l sB B

G U U U& ( ( '� � �� � � � � � (28)

where b( and f( are the scaling factors for the receptor and water environ-ments, respectively. This model has been considered previously, but it was not found to significantly improve agreement with experiment compared to the standard form of the LIE equation.80 Furthermore, it is clear that the force field independence of eq. 27 relies on cancellation of errors in

ell sU �� . Thus, if different ( values are assigned to the receptor and water

environments, this would inevitably lead to a force field dependence of the LIE method.

In paper IV it was found that introducing intramolecular electrostatic en-ergies significantly improved the accuracy of the LR approximation for cal-culation of hydration free energies. Implementing the same idea into eq. 27 yields � �vdW el el

bind l s l s l lG U U U& ( '� � �� � � � � � � � (29) where el

l lU �� is the difference in ligand–ligand (l–l) electrostatic energy between the bound and the free state of the inhibitor.

3.2.4 Continuum solvent models in the LIE method As described in the previous chapter, there are several advantages with the use of continuum solvent methods in free energy calculations. The main reason for introducing these models in the LIE method is that the continuum solvent energies converge faster than the ligand–water interaction energies from explicit solvent. This gives the possibility to reduce computation times or to study larger and more complex systems. A continuum version of the

Page 39: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

37

LIE method can be derived by dividing the polar term into solvent and re-ceptor contributions, which yields � �bel el vdW

bind l receptor l wat l sBG U U U( & '� � �� � � � � � � (30)

where

bell receptor B

U � and b fel el el

l wat l wat l watB BU U U� � �� � � are the electrostatic

ligand–receptor and the change in electrostatic ligand–water interaction en-ergies, respectively. All terms in eq. 30 except el

l watU �� , which is estimated using a continuum solvent method, are calculated from the force field poten-tial function. The choice of sampling method and continuum solvent model to evaluate el

l watU �� will determine the speed and accuracy of methods based on eq. 30. The least computationally demanding sampling method is to simply use a single minimized structure of the receptor–ligand complex. To obtain more reliable results, ensembles of conformations can be extracted from implicit or explicit solvent simulations. In the evaluation of el

l watU �� from the sampled conformations, a PB or GB solvent model can be used. Several physics-based methods for calculating binding free energies have been developed by combining these alternatives. In the LIECE method, minimization of the receptor–ligand complex is followed by post-processing with the PB method.94 In the MM-PB/SA method, sampling in explicit sol-vent is combined with calculations of hydration free energies using the PB model.95 Finally, Zhou et al. proposed the SGB-LIE method, in which a GB model is used both for sampling and evaluation of the electrostatic solvent contribution to the binding free energy.96

In paper III, the LIE method combined with explicit, GB, and PB solvent models were used to estimate binding free energies according to eq. 30 for a set of eight malarial plasmepsin II inhibitors.97 To provide the best possible sampling and focus on evaluating the accuracy of the continuum models, snapshots from explicit solvent simulations of the ligands and protein–ligand complexes were used to generate ensembles of conformations. These were then post-processed in PB and GB continuum solvent.

Figure 7. Calculated ( LIE

bindG� ) and experimental binding ( obsbindG� ) free energies for

eight malarial plasmepsin II inhibitors using (A) LIE (explicit solvent), (B) LIE-PB and (C) LIE-GB.

Page 40: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

38

The results for the LIE-PB method are shown in Figure 7B and for this model an average unsigned error from experiment equal to 0.7 kcal/mol was obtained using the standard parameterization of the LIE method. The LIE-GB method resulted in significantly worse results. To obtain agreement with experimental binding free energies a constant offset, 3.6' � � kcal/mol, had to be added to each binding free energy, which yielded an average unsigned error of 1.2 kcal/mol (Figure 7C). This shows that an improved GB model has to be used to accurately predict binding free energies with this method. The LIE-PB method, on the other hand, can be used to estimate binding free energy with accuracy comparable to explicit solvent LIE calculations (Figure 7A) given that an ensemble of relevant structures is available.

The extent to which sampling length and accuracy can be reduced was not examined in paper III and should be further investigated to elucidate the potential of the combined LIE and continuum model approach.

3.2.5 The LIE method in structure-based drug design One of the most interesting areas to which the LIE method can be applied is computational structure-based drug design (SBDD) against proteins, e.g. enzymes. The aim of SBDD is to design small molecules (i.e. drugs), that can inhibit the function of an enzyme from the information available in an atomic structure of this macromolecule.20,73 Reliable computational SBDD methods would provide an effective approach to filter out possible inhibitors from large databases of molecules. This area is also of considerable interest for the pharmaceutical industry because of its potential to become an inex-pensive technique to identify lead compounds.

In the first step of SBDD, a drug target has to be identified. The structure of the target protein can either be obtained from experimental techniques or, in certain cases, from homology modeling. A database of molecules is then docked into the binding site of the protein. In this step fast docking algo-rithms are used to generate and score putative binding poses in the binding site. Since the binding site is assumed to be rigid in most docking calcula-tions, the quality of the protein structure and the context in which it was determined will have a great impact on the success of this step. Docking to experimentally determined structures generally give better results than to a homology modeled structure and docking to the holo state of the protein is more likely to generate good predictions than to the apo form.98 To be able to extract compounds, which are likely to inhibit the enzyme, for experimen-tal evaluation, a score that reflects how strong the molecules bind to their target must be calculated for the predicted binding modes. While the scoring functions used in docking programs are often successful in predicting bind-ing modes, their ability to rank molecules by affinity has been shown to be

Page 41: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

39

poor in many cases.99 These results and a large amount of available compu-tational power have increased the interest for more advanced scoring func-tions that have the potential to make more reliable predictions. One approach to further refine the results from the docking calculation is to use MD simu-lations in combination with the LIE method to rescore a subset of the data-base. This approach was applied in two different projects, which are summa-rized below.

3.2.5.1 Combining docking, MD and the LIE method In paper V, a combined docking, MD and LIE method approach was tested in a retrospective study to predict binding modes and binding affinities for a series of non-nucleoside inhibitors (NNRTIs) of HIV-1 reverse transcriptase (HIV-RT). HIV-RT is one of the main targets in the development of drugs against AIDS and several NNRTI drugs are currently used in clinic. How-ever, because of the fast emergence of drug resistant HIV mutants, there is a continuous need for new inhibitors to this target.100

Figure 8. Experimental ( obs

bindG� ) and calculated ( calcbindG� ) LIE binding free energies

for 39 of the 43 studied NNRTIs.

GOLD was first used to dock a series of 43 known inhibitors to a HIV-RT structure, which had been determined in complex with one of the studied molecules.101,102 Inhibitor starting structures to be used in explicit solvent MD simulations were extracted from the docking calculations. It was found that the predicted binding modes were similar to experimentally determined structures, but that the docking score did not correlate with measured bind-ing affinities. The calculated LIE binding free energies (standard parameteri-zation) agreed very well with the experimental binding free energies for a majority of the inhibitors (Figure 8). Two alternative forms of the LIE equa-tion, eqs. 28 and 29, were also investigated but none of these significantly improved agreement with experiment compared to the original form of the LIE method.

Page 42: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

40

3.2.5.2 Virtual screening against a tuberculosis target

In an ongoing cross-disciplinary collaboration project to design drugs against tuberculosis, a prospective virtual screen against an enzyme of the Myobac-terium tuberculosis bacterium was carried out. The target enzyme, 1-deoxy-D-xylulose-5-phosphate reductoisomerase (DXR) is part of the non-mevalonate pathway, which produces isopentenyl diphosphate, and is essen-tial to the bacterium. Furthermore, this pathway does not exist in humans, which makes it an interesting drug target. Recently, a structure of DXR in complex with the known inhibitor fosmidomycin (FOM), NADPH, and Mn2+ was determined103 (Figure 9), which gave the possibility to use computa-tional SBDD to find inhibitors of the enzyme.

Figure 9. Active site of DXR with FOM, NADPH, and metal ion (Mn2+).103

The goal of the project was to find an inhibitor that replaced FOM and parts of the co-factor NADPH. The different steps of the virtual screen are sum-marized in Figure 10. From a database of 2.8 million compounds, a collabo-rating medicinal chemistry group reduced the number of molecules to ap-proximately 1000 potential binders. In the first steps of the screen, drug-like molecules which had similarities to NADPH and FOM were extracted. The remaining 30 000 molecules of the database were then reduced to about 1000 using docking, pharmacophore filtering, and a scoring function. In the

Page 43: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

41

next step the LIE method was used to reduce these to around 60 molecules of which 30 were tested for experimental activity.

Figure 10. Summary of the different steps in the virtual screen against DXR.

Activity measurements first indicated that two of the 30 compounds inhib-ited the enzyme, but NMR experiments carried out in parallel found the same molecules to have very low solubility. The NMR experiments have instead showed that two other molecules in the set bind to the protein. Some of these compounds have been set up for co-crystallization with DXR, but no structures with a bound ligand have been solved so far.

Page 44: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

42

4. Estimating entropies from MD simulations

To this point, the focus of this thesis has been on calculating free energy, which is the most important thermodynamic property for studying biochemi-cal processes. The change in free energy predicts the direction of a chemical process and can be expressed G H T S� � � � � (31) where T is the temperature and G� , H� , and S� are the changes in free energy, enthalpy, and entropy, respectively. The enthalpy and entropy pro-vide a more detailed description of the driving forces acting in chemical processes. The change in enthalpy represents changes in the strength of the interactions between the molecules while the entropy contribution can be described as a measure of the change in the order of the system.

The possibility to estimate entropies and enthalpies for solvation and molecular association is of considerable interest because this partitioning can often simplify the interpretation of a free energy. For example, the hydration of an ion is often characterized by both a negative enthalpy and entropy change. The enthalpy term originates from favorable interactions between the ion and the solvent dipoles ( 0hydH� * ). These strong electrostatic inter-actions also restrict the motions of water molecules, which increase the order in the system and leads to a negative change in entropy ( 0hydT S� � ) ). For molecular association it is often found that the balance between enthalpy and entropy can be very different for ligands with similar binding free energies. In drug design, the possibility to analyze enthalpies and entropies can be used to optimize the affinity of a ligand for a receptor.104

In papers VI, VII, and VIII, two different types of methods for calculat-ing entropies from MD simulations were investigated. The first type is used to calculate entropies for the solute degrees of freedom. In papers VI and VII, such methods were evaluated and particular focus was put on estimating the loss of entropy for a ligand that binds to a receptor. The other type of method is based on rigorous free energy methods, i.e. the FEP and TI tech-niques. In papers VII and VIII, the FEP technique was used to estimate hy-dration entropies for alkali metal ions.

Page 45: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

43

4.1 Solute entropies Entropy is often described as a measure of the “disorder” or “dispersal of energy” in a system. To be able to calculate entropies from MD simulations, these words have to be more clearly defined and translated to equations.

Consider a system with constant energy E composed of N molecules. Each molecule in this system can occupy energy levels 0� , 1� , 2� , … with the constraint that the total energy of the system must be equal to E. This leads to a large number of possible configurations for the system with a spe-cific number of molecules in in each energy state i� . The number of ways, W, that the molecules can be arranged to achieve a certain configuration is equal to ! .

!ii

NWn

�2

(32) Statistical thermodynamics shows that the configuration that can be formed in the largest number of ways will completely dominate the macroscopic properties of the system. The entropy is defined as lnS k W� (33) where k is Boltzmann’s constant. This means that the more arrangements that are available to a system, i.e. the more “disordered” it is, the larger the entropy becomes. An alternative way to express the entropy, which can be derived from eq. 33, is lni i

i

S k p p� � � (34) where ip is the probability of state i ( /i ip n N� ).105 In papers VI and VII, methods based on this equation were used to calculate translational, rota-tional, and vibrational entropies of small molecules.

4.1.1 Schlitter’s formula and quasiharmonic analysis In 1993, Schlitter presented a new method for calculating entropies from molecular simulations.106 Previously, methods for calculating solute entro-pies required that a transformation from Cartesian to internal coordinates was carried out.107,108 Schlitter derived an ad hoc formula for calculating the absolute entropy of a system directly from the covariance matrix of atomic fluctuations in Cartesian coordinates, which can easily be obtained from an MD simulation. It was also shown that the calculated entropy represented an

Page 46: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

44

upper bound to the true entropy of a system. Schlitter’s formula (SF) for the entropy is 1 1

2 2

2

2

1 ln det2SF

kTeS k� ��

� �� �� �� �� 1 M �M

� (35)

where � is the covariance matrix of atomic fluctations in Cartesian coordi-nates, M is the mass matrix, 1 is the unity matrix, � is Planck’s constant divided by 2# , and e is Euler’s number.

Starting around year 2000, SF was extensively applied to calculate en-tropies involved in the dynamics and association of biomolecules.109-115 It seemed to be very difficult to obtain converged estimates for macromole-cules, but SF appeared suitable for calculating entropies for small molecules. Interestingly, this gave the possibility to quantify the change in translational, rotational, and conformational/vibrational entropy for a ligand that binds to a receptor. However, it was not clear if SF could be used to predict ligand binding entropies because it had not been validated for this type of calcula-tions. Furthermore, Andricioaei and Karplus showed that quasiharmonic analysis (QHA) could also be used to calculate entropies from the covariance matrix in Cartesian coordinates.116 This method is based on the assumption that the fluctuations of the system can be described by a Gaussian probabil-ity distribution.107,116-118 In 1981, when Karplus and Kushick first introduced QHA107, it was used to calculate differences in conformational and vibra-tional entropies from internal coordinates.107 Andricioaei and Karplus showed that the quasiharmonic frequencies for a system could be calculated from the covariance matrix of atomic fluctuations in Cartesian coordinates. From the frequencies, i3 , the entropy can be estimated from

� �//

/ ln 1 .1

i

i

kTiQH kT

i

kTS k ee

33

3 �� � ��� �

� (36) Furthermore, they showed that QHA represented a tighter upper bound to the true entropy than SF. This also made it interesting to evaluate if QHA was suitable for calculations of small molecule entropies.

In paper VI, QHA and SF were compared and validated for calculations of solute entropies on a test set of relatively rigid organic molecules, e.g. methane and benzene. For these molecules, the entropy can approximately be decomposed as

trans rot vibS S S S� � � (37)

Page 47: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

45

where transS , rotS , and vibS are the translational, rotational, and vibrational entropies, respectively. Reference values for these entropies were derived from statistical thermodynamics and experiments.

For translational entropies, it was found that QHA and SF yielded very similar results, which were in good agreement with the reference values. The entropies were consistently somewhat larger than the theoretical values. This was expected since these methods were derived to represent upper bounds to the true entropy. The Gaussian approximation resulted in a systematic error of about 1 cal/mol K for each molecule. Thus, relative entropies were in very good agreement with theory for both methods. Both QHA and SF were found to significantly overestimate the absolute rotational entropies com-pared to reference values. It was clear from these calculations that QHA and SF could likely not be used to accurately estimate the change in rotational entropy for a ligand that binds to a receptor. Finally, QHA yielded a slightly better correlation with reference values than SF for vibrational entropies. Since all molecules in the test set were relatively rigid, the accuracy of SF and QHA for calculating conformational entropies was not investigated. This was addressed by Chang et al.119 and, in agreement with the results for rota-tional entropies, the absolute entropies for flexible molecules were signifi-cantly overestimated. In particular, very large errors were found for mole-cules that occupy several energy minima.

4.1.2 Estimation of binding entropies Binding entropies have been the focus of many theoretical calcula-tions.112,114,120-130 In paper VI and VII, approximate methods to estimate ligand binding entropies were derived from the classical expression for the entropy ( ) ln ( )S k p p d� � � r r r (38) where ( )p r is the positional probability distribution. The momentum contri-bution has been excluded in eq. 38 because it will not contribute to the change in binding entropy. The ligand binding entropy, lig

bindS� , is the differ-ence between the entropy for the ligand bound to a receptor and free in solu-tion. Partitioning lig

bindS� into three terms that represent different modes of motion yields

/lig trans rot conf vib

bind bound free bind bind bindS S S S S S� � � � � � � � � (39) where transl

bindS� , rotbindS� , and /conf vib

bindS� are the translational, rotational, and vibrational/conformational contributions to the binding entropy, respec-tively. It should be noted that lig

bindS� does not include differences in confor-

Page 48: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

46

mational/vibrational entropy for the receptor or the change in solvent en-tropy. These terms also contribute significantly to the binding entropy and will be briefly discussed in the end of this section.

Figure 11. Illustration of the available volumes for a ligand free in solution and bound to a receptor.

The description of the loss of translational and rotational entropy for a ligand that binds to a receptor is often oversimplified in approximate binding free energy calculations. For example, these terms are sometimes completely neglected131 or assumed to contribute only with a constant value132. In paper VI, it was shown that if the receptor confines the ligand in an infinite square well potential, the change in translational entropy can be directly calculated from the volumes available to the ligand in the bound and free states. This is illustrated in Figure 11, where boundV and freeV represent the available vol-umes for the ligand bound to the receptor and free in solution, respectively. By integrating eq. 38, the loss of translational entropy can be expressed

ln .trans bound

bindfree

VS kV

� �� � � �� �

� (40)

The free volume is determined by the chosen standard state and at 1 M stan-dard concentration freeV =1660 Å3. If the distribution for the motion of the center of mass (CM) of the ligand ( ( )transp r ) was known for the bound state,

boundV could be calculated by integrating eq. 38. For example, the probability distribution could be estimated by analyzing MD simulation trajecto-ries.124,129 However, using simulations to obtain distributions is difficult be-cause the tails of these are seldom sampled. An effective alternative is to assume a specific distribution for the motions of the CM for the ligand. If a Gaussian distribution is assumed, the loss of translational entropy can be estimated from the root-mean-square (rms) fluctuations for the CM of the ligand ( x1 , y1 , and z1 ). In this case, boundV should be set equal to

Page 49: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

47

� �3/ 22 x y ze# 1 1 1 in eq. 40. In paper VI, this expression was shown to give results similar to SF and QHA for free translation in the gas phase. Further-more, eq. 40 was directly compared to rigorous FEP calculations of the en-tropy of confining a single particle in water in paper VII. Eq. 40 was in ex-cellent agreement with the calculated results and also converged more rap-idly than the FEP method.

In analogy with eq. 40, the loss of rotational entropy can be expressed as lnrot bound

bindfree

S k� �4

� � � �� �4� (41)

where bound4 and free4 are the rotational “volumes” in the bound and free states, respectively. In the reference state, the ligand rotates freely in solu-tion, which yields 28free #4 � . For the bound state, Euler angles ( " , 5 , and � ) can be used to describe rotation of the ligand.124 Assuming a Gaus-sian distribution for the Euler angles yields that bound4 is approximately equal to � �3/ 22 sine " 5 �# 1 1 1 � . SF and QHA had been found to be inap-propriate for estimating rotational entropies of molecules in the gas phase. In papers VI and VII, eq. 41 was shown to be a significantly better approxima-tion for this case and the method also agreed well with FEP calculations on a small test system.

In paper VI, eqs. 40 and 41 were used to calculate the loss of transla-tional and rotational entropy when benzene binds to a mutant of T4-lysozyme. Since the total binding entropy was not calculated, it was not pos-sible to compare the results to experimental values. Instead, the calculated entropies were compared to those estimated by Hermans and Wang.123 In their approach, a rigorous restraint-release (RR) method was applied to esti-mate translational and rotational contributions to the binding free energy. First, restraints were applied to the rotational and translational degrees of freedom for the ligand. By gradually removing restraints on the ligand in the binding site, the translational and rotational binding free energy contribution could be determined using the TI technique. The RR method should, in prin-ciple, give as accurate results as the employed force field, but long simula-tions may be required to achieve convergence. The approximate expressions for trans

bindS� and rotbindS� , on the other hand, can be calculated from a single

simulation of the ligand in the binding site and the results generally converge quite rapidly. For the change in translational entropy, eq. 40 yielded a free energy contribution of trans

bindT S� � ! 5.2 kcal/mol at 298 K, which was in very good agreement with the RR results (5.6 kcal/mol). For the rotational contri-bution eq. 41 yielded rot

bindT S� � ! 2.0 kcal/mol, which was fairly close to value calculated by Hermans and Wang (3.4 kcal/mol). In total, the transla-tional and rotational entropy contribution to the free energy of binding is 7.2 kcal/mol. This clearly shows that translational and rotational entropy contri-

Page 50: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

48

butions can not be neglected in calculations of binding free energies. Fur-thermore, this term can not be considered constant for different types of ligands and receptors as assumed by many scoring functions.127

The last term in eq. 39 represents the changes in vibrational and confor-mational entropy upon binding for the ligand. A rough estimate of the loss of conformational entropy for a ligand can be estimated directly from eq. 34. Assuming that a rotable bond has three minima, which are completely frozen upon binding, results in a contribution to the binding free energy of +0.7 kcal/mol ( 1 1

3 3ln ln3confbindT S RT RT� � � � �� ). According to this approxima-

tion, which is used in some scoring functions, the change in conformational entropy can contribute significantly to the total binding free energy. How-ever, all possible minima of a molecule are likely not accessible in the free state and the rotable bonds of the ligand are not completely frozen in the binding site. Thus, this approximation will overestimate the conformational entropy loss of the ligand. Attempts to quantify the change in conforma-tional/vibrational entropy for a ligand using SF have also been made.112 However, since SF and QHA based on Cartesian coordinates significantly overestimates the entropy for molecules that occupy several minima, the accuracy of such estimates is questionable.119

For the semi-empirical LIE method, the entropy contributions to binding are implicitly taken into account by the & , ( , and ' coefficients. The change in receptor and solvent entropies for charging the ligand in water and in the binding site is taken into account by using the LR approximation. Non-polar solvation entropy contributions are known to correlate well with size-measures and are parameterized into the & coefficient. Furthermore, it was shown in paper VII that transl

bindS� correlates with the ligand–surrounding van der Waals energies. This basically reflects that a large ligand in a bind-ing site would not have the same rotational and translational freedom as a small ligand. While it is easy to envision cases where specific interactions rather than the size of a ligand determine rotational and translational entro-pies a binding site, the vdW

l sU& �� term likely takes some of the change en-tropy into account. However, to further improve the accuracy of approximate methods to calculate binding free energies, ligand binding entropies should be explicitly included.

4.2 Solvent entropies Solvent entropies play an important role in ligand binding, enzyme catalysis, and protein folding. Binding affinities observed for hydrophobic ligands are sometimes dominated by a positive entropy contribution, which reflects that the hydration entropy of the receptor–ligand complex is larger than for the

Page 51: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

49

isolated binding partners ( complexhydS� ) ligand

hydS� � receptorhydS� ).104 Microscopically,

this increase in entropy mainly originates from the release of ordered water molecules bound to the surface of the ligand and the binding site of the re-ceptor. Calculation of hydration free energies for amino acid side chain ana-logues has become a standard means of validating biomolecular force fields3,133-137, but the balance between entropy and enthalpy can be very dif-ferent and still reproduce the same free energy.138,139 For this reason, hydra-tion entropies and enthalpies should also be included in the parameterization of partial charges and van der Waals parameters.

To estimate hydration enthalpies and entropies from MD simulations is considerably more challenging than to calculate hydration free energies.140 Solvation free energies can be calculated with high precision with the FEP technique because only the solute–solvent interaction energies are used in the exponential average in the FEP formula. The solvent–solvent terms, which are associated with large fluctuations in simulations, cancel in eq. 4 because they are the same in both potentials. On the contrary, the standard approach to estimate the hydration enthalpy and entropy requires that a dif-ference between total potential energies is calculated. This leads to large statistical errors when entropies and enthalpies are computed.

In paper VII and VIII, ion hydration entropies were calculated using two different methods. The standard approach, in which the entropy is obtained by subtracting the free energy from the enthalpy, yields

� �1 0

1hyd hydS U U G

T� � � � � �� � (42)

where

1U and

0U are the total potential energies for the ion/water and

pure water system, respectively.141,142 The P V� term has been neglected in eq. 42 because it is typically very small. An alternative approach is to calcu-late entropies from the temperature dependence of the free energy.138,142,143 For example, the hydration entropy and enthalpy can be estimated from a van't Hoff plot using free energies calculated at several different tempera-tures. The van’t Hoff equation is given by rearranging eq. 31 into

1 .hydhyd hyd

GH S

T T�� � � �� � � �� � � �

� � (43)

If hydH� and hydS� are assumed to be independent of temperature over a temperature range, the hydration entropy can be estimated from the intercept of eq. 43 ( 1 0T � � ). Interestingly, this method avoids the calculation of total potential energies, but requires that the hydration free energies are calculated with high precision.

Page 52: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

50

4.2.1 Absolute hydration entropies of alkali metal ions from MD simulations Metal ions often play an important role in enzyme active sites and are also crucial for regulating different chemical processes in the cell. Due to the long-range nature of electrostatic interactions, estimating thermodynamic properties for ions is also one of the major challenges in computational chemistry. Åqvist derived the first set of van der Waals parameters for alkali metal ions that were tuned to reproduce experimental free energies of hydra-tion.134 In papers VII and VIII, the hydration free energies and entropies for five alkali metal ions (Li+, Na+, K+, Rb+, and Cs+) were calculated from MD simulations in combination with the FEP technique.

Figure 12. FEP calculated (unfilled boxes) and observed144 (filled boxes) absolute hydration entropies ( T S� � , kcal/mol) for five alkali metal ions (Li+, Na+, K+, Rb+, and Cs+) at 298 K. Uncertainties in calculated and observed values are shown using error bars.

In paper VII, the difference in hydration entropy between Na+ and Cs+ was calculated using both eq. 42 and 43. The calculated entropic contribution to the free energy of hydration ( hydT S� �� ) was �2.3 kcal/mol using eq. 42 while a van’t Hoff plot yielded �3.6 kcal/mol. The latter result was in good agreement with the experimental value of �3.7 kcal/mol.145 In paper VIII, the absolute entropies were calculated for five alkali metal ions (Li+, Na+, K+, Rb+, and Cs+) using van't Hoff plots based on hydration free energies at eight different temperatures. As expected, the absolute free energies of hydration were in good agreement with experimental values. Interestingly, the hydra-tion entropies were also found to be close to experimentally determined val-ues (Figure 12). This is quite remarkable considering that the van der Waals parameters were only adjusted to reproduce the hydration free energies and the first peaks in the ion–water radial distribution functions. An interesting

Page 53: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

51

future application of the FEP technique in combination with van’t Hoff plots is to calculate relative binding free energies, enthalpies and entropies for host-guest systems, e.g. ion binding to crown ethers.

The possibility to use approximate methods to predict hydration entro-pies was also investigated in paper VIII. In this analysis the hydration en-tropy was partitioned into two contributions

0 00hyd I I I

S S S �� �� � � � � (44)

where 00 I

S�

� and 0I IS ��

� describe creation of the uncharged van der Waals sphere in water and the charging of this sphere, respectively. The first term in eq. 44 was found to correlate well with the solvent accessible surface area (SASA) of the ions. 0I I

S ��� was estimated from the temperature depend-

ence of free energies calculated with the Born model (eq. 8) and the LR ap-proximation (eq. 15). The Born equation combined with an accurate empiri-cal relationship for the dielectric constant of liquid water146 could not be used to reproduce 0I I

S ��� . A similar result was obtained using the LR ap-

proximation. In this case, it was found that the ( coefficient in eq. 15 de-pends linearly on the temperature, which limits the possibility to use this approach in predictions of hydration entropies.

Page 54: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

52

5. Future perspectives

The calculation of thermodynamic properties from molecular simulations is an exciting and challenging field of research. The continuously increasing amount of computational power has made it possible to rigorously calculate solvation and receptor–ligand binding free energies with high precision. However, it is important to keep in mind that the results will only be as accu-rate as the employed force field. For a long time, inadequate force field pa-rameters have probably been hidden behind sampling problems. In the fu-ture, it will be possible to make large scale comparisons between simulations and experiments, which will give the opportunity to further improve the ac-curacy of force fields. In particular, there is a need for rapid and reliable assignment of force field parameters for small organic molecules. This is one of the keys to improve the accuracy of free energy calculations.

The representation of water in free energy calculations is one of the problems addressed in this thesis. Some of the results indicate that implicit solvent models can successfully be used to obtain results of similar quality as explicit solvent calculations. However, there are still problems with the accuracy of current explicit solvent models in free energy calculations and, for this reason, it would be bold to claim that implicit solvent is suitable for modeling of proteins in general. Instead, the choice of water model should be determined by the speed and accuracy required to solve a specific prob-lem. Explicit solvent is the most accurate representation of water in MM based MD simulations, but it is too computationally demanding for studying processes that take place on long time scales. In these cases, GB models give the opportunity to include solvent to a reasonable computational cost. In SBDD, where large databases are screened for inhibitors of enzymes, ex-plicit solvent representation may also be too computationally expensive. Even if current GB models can not provide as relevant structures as explicit solvent, performing the simulations in implicit solvent is certainly better than not carrying out any sampling at all. Furthermore, a combined approach of explicit and implicit solvent representation may be a suitable alternative in many cases. New GB models are continuously being developed and are of-ten shown to give impressive agreement with the PB method. However, the PB model is also a continuum description of solvent. For this reason, it would be useful to focus more on reproducing experimental data and explicit solvent calculations in the parameterization of GB models.

Page 55: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

53

More than 13 years have passed since the first LIE calculations of bind-ing free energies were presented. Since then, the method and variants thereof have been successfully used to predict binding affinities for many receptor–ligand complexes. For example, an LIE method was very recently shown to identify new inhibitors of two different kinases.147 In the work reported in this thesis, a continuum version of the LIE method, which can be used to improve convergence of binding free energy calculations, was presented. Furthermore, the combination of docking and the LIE method was shown to be a promising approach to identify drug candidates even for relatively large datasets. However, there are still several problems that need to be further investigated to improve the method. Issues that only concern the LIE method are the accuracy of the LR approximation in the protein environment, the system transferability of & , the prediction of ' , and the treatment of in-tramolecular ligand energies. To address these problems, it will be useful to carry out calculations on series of small ligands that bind to well-behaved receptors. In these cases, problems associated with force field parameters and conformational sampling can be avoided, which gives the opportunity to focus solely on the above questions. Some of the problems that concern all approximate free energy methods are conformational sampling and model validation. For flexible ligands, it is very difficult to obtain converged elec-trostatic interaction energies in the simulation of the free molecule in water. Starting from an extended structure of the ligand, it can take many nanosec-onds before the molecule reaches relevant parts of conformational space. In this case, algorithms that identify low energy conformations of small mole-cules and continuum models can be used to reduce these problems. Finally, it is important to note that the standard parameterization of the LIE method is only based on 18 protein–ligand complexes. It is now possible to apply the method to larger and more diverse datasets. This may reveal more weak-nesses of the current model and give directions to how a more general LIE method can be developed.

The possibility to calculate the enthalpic and entropic components of a free energy will provide a new dimension for interpreting the driving forces in biochemical processes. For example, the microscopic origin of the balance between enthalpy and entropy in solvation and ligand binding is not com-pletely understood. The development of physics based scoring functions for calculating protein–ligand binding affinities has also mainly been focused on estimating enthalpic contributions to the binding free energy. The ap-proaches outlined in this thesis can be used to explicitly calculate the transla-tional and rotational entropy loss for a ligand, which can improve the accu-racy of approximate methods for estimating binding affinites.

Page 56: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

54

6. Summary in Swedish

I denna avhandling presenteras metoder för att studera molekyler med hjälp av datormodeller. I dessa modeller beskrivs atomer och deras interaktioner med matematiska funktioner som, i kombination med kraftfulla datorer, kan användas för att simulera molekylers rörelser. Att kunna göra simuleringar av proteiner, som är en av naturens grundläggande byggstenar, är av stort intresse. Datorsimuleringarna kan ge en detaljerad bild av proteiner och de matematiska modeller som beskriver deras interaktioner kan användas för att uppskatta energiförändringar för processer som dessa molekyler är involve-rade i. En särskilt intressant grupp av proteiner är s.k. enzymer, vilka kan beskrivas som biologiska katalysatorer. Att kartlägga sambandet mellan struktur och funktion för enzymer kan dels användas för att förstå hur olika biokemiska processer fungerar, men också för att utveckla nya läkemedel. Avhandlingens huvudresultat presenteras i tre olika avsnitt. I det första stu-deras tekniker för att beskriva vatten i datorsimuleringar. I det andra och tredje avnittet presenteras metoder för att uppskatta energier för små mole-kyler i vatten och hur de interagerar med proteiner.

I princip så finns det två typer av vattenmodeller för datorsimuleringar av proteiner. Den ena beskriver varje molekyl i detalj och denna modell kan därigenom representera vattnets egenskaper som lösningsmedel på ett bra sätt, men kräver samtidigt extremt mycket datorkraft. Den andra typen är inte lika dyr beräkningsmässigt, men kan endast beskriva vattnets effekt i stora drag. I den första artikeln i avhandlingen jämförs simuleringar med de två olika typerna av vattenmodeller för beräkningar av syra-bas egenskaper för två olika proteiner. Proteiner är uppbyggda av aminosyror och vissa av dessa kan binda vätejoner, vilket ofta bestämmer enzymers egenskaper vid olika pH värden. I artikeln uppskattas tre asparaginsyrors förmåga att binda vätejoner med simuleringar baserade på två olika typer av vattenmodeller. Beräkningarna visar att simuleringar kvalitativt kan förutsäga aminosyrornas syra-bas egenskaper. I jämförelsen mellan de två olika vattenmodellerna visar det sig att den förenklade typen av vatten, i viss utsträckning, kan ersät-ta den detaljerade vattenmodellen. I avhandligens andra artikel studeras en-zymet epoxidhydrolas egenskaper vid olika pH värden. I detta projekt illu-streras hur de två olika vattenmodellernas egenskaper, på bästa sätt, kan kombineras för att försöka förklara hur detta enzym fungerar. De beräknade syra-bas egenskaperna för epoxidhydrolas stödjer några nya teorier om detta enzym och kan även förklara experimentella resultat. I den tredje artikeln

Page 57: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

55

jämförs olika vattenmodeller för beräkningar av hur starkt små molekyler binder till proteiner. Även i detta fall verkar det som att det finns bra förenk-lade typer av modeller som kan vara användbara i studier av proteiner.

I avhandlingens tredje kapitel beskrivs linear interaction energy (LIE) metoden, som kan användas för att uppskatta hur starkt små molekyler, s.k. ligander, binder till proteiner. I LIE metoden beräknas detta genom att utföra två simuleringar; en av liganden bunden till proteinet och en av liganden i vattenlösning. Skillnaden i energi mellan dessa två beräkningar kan ge en uppskattning av bindningsenergin, som indikerar just hur starkt liganden binder till proteinet. I artikel fyra undersöks hur bra LIE metoden fungerar för att beräkna ligandens energi i vatten. En ny modell för att uppskatta den-na energi presenteras och den stämmer väl överens med experimentella mät-ningar på små molekyler i vattenlösning. Dessutom illustreras i detta avsnitt hur de förenklade vattenmodellerna, som undersöktes i avhandlingens första del, kan användas i LIE metoden. Beräkningar på en serie ligander som bin-der till ett enzym visar att bra resultat kan erhållas även med en förenklad typ av vattenmodell. Eftersom evaluering av vattenenergier idag upptar stör-re delen av datorernas beräkningskraft, så kan de förenklade modellerna göra LIE metoden mer effektiv. Ett av de främsta användningsområdena för LIE metoden är datorbaserad läkemedelsdesign. I datorbaserad läkemedelsdesign försöker man identifiera och optimera små molekyler som kan användas som läkemedel. I artikel två och fem används LIE metoden för att studera möjliga läkemedelskandidater mot malaria och HIV. I ett pågående projekt för att finna nya läkemedel mot tuberkulos så har LIE metoden också bidragit till att försöka identifiera möjliga läkemedelskandidater utifrån en stor databas av små molekyler. Denna metodik kan bidra till att effektivisera den mycket kostnadskrävande processen att utveckla läkemedel.

I avhandligens sjätte och sjunde artikel presenteras metoder för att stude-ra ett speciell typ av bidrag till bindningsenergin. Då små molekyler binder till proteiner förlorar de sin rörelsefrihet, vilket resulterar i ett energibidrag som motverkar den totala bindningsenergin. Detta bidrag, som kan kallas bindningsentropi, har man ofta bortsett ifrån i förenklade metoder för att beräkna bindningsstyrkan för ligander till proteiner. Flera olika metoder för att uppskatta denna energi presenteras och visar att det är viktigt att ta hän-syn till detta bidrag till bindningsenergin. Dessa metoder kan komma att förbättra möjligheten att göra förutsägelser av liganders bindningsenergier till proteiner. I avhandligens sista artikel så utvärderas hur joner, som är vik-tiga i många biokemiska processer, kan representeras i vattenlösning. Resul-taten visar att de parametrar som tidigare utvecklats för fem olika joner re-producerar en rad olika experimentellt bestämda egenskaper på ett bra sätt.

Page 58: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

56

7. Acknowledgements

First of all, I would like to thank my supervisor Johan Åqvist for giving me the opportunity to explore any problem that I have found interesting during these four years. I have really enjoyed working in your group. I would also like to thank everyone outside the Åqvist group that has con-tributed to my papers: David Case, Thomas Simonson, Mikael Widersten & Co, and Lena Henriksson & Co. My time in the Åqvist lab would not have been the same without the other members of the group: Martin, Martin, Martin, Sinisa, Lars, Björn, Hugo, Stefan, Chris, Fredrik, Johan, Göran, and Viktor. Thank you for good times in and outside the lab, good scientific advice, and for proofread-ing my thesis. Extra special thanks to: Dr. Almlöf for being amazingly clever, for always taking the time to explain why most of my theories are completely wrong, and for making me under-stand that science should always be fun, but not necessarily useful. Dr. Bjelic for never hesitating to battle integrals with pen and paper! The mysterious Mr. Boukharta for brilliant work on HIV-RT and because you always make me smile. Mr. Ander for the GB code and for introducing me to MM and AFM. Dr. Nervall for the PlmII inhibitors and for being the only one in the group who can beat me at frisbee golf. Everyone at xray for nice coffee breaks and to Lotta for running the beer club. Till min familj och vänner - Tack för allt stöd under mina doktorandstudier. Speciellt tack till mamma för dina försök att korrekturläsa och förstå avhand-ligen! Nina, Nina, Nina! För att du kan allt som jag inte kan. Tack för att du hitta-de mig och förvandlade mig till den jag ville vara. bElla – I löv you baby!

Page 59: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

57

8. References

1. Jorgensen, W. L.; Tirado-Rives, J. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 6665. 2. Mackerell, A. D. J. Comput. Chem. 2004, 25, 1584. 3. Cornell, W. D.; Cieplak, P.; Bayly, C. I.; Gould, I. R.; Merz, K. M.; Fer-guson, D. M.; Spellmeyer, D. C.; Fox, T.; Caldwell, J. W.; Kollman, P. A. J. Am. Chem. Soc. 1995, 117, 5179. 4. MacKerell, A. D.; Bashford, D.; Bellott, M.; Dunbrack, R. L.; Evanseck, J. D.; Field, M. J.; Fischer, S.; Gao, J.; Guo, H.; Ha, S.; Joseph-McCarthy, D.; Kuchnir, L.; Kuczera, K.; Lau, F. T. K.; Mattos, C.; Michnick, S.; Ngo, T.; Nguyen, D. T.; Prodhom, B.; Reiher, W. E.; Roux, B.; Schlenkrich, M.; Smith, J. C.; Stote, R.; Straub, J.; Watanabe, M.; Wiorkiewicz-Kuczera, J.; Yin, D.; Karplus, M. J. Phys. Chem. B 1998, 102, 3586. 5. Jorgensen, W. L.; Maxwell, D. S.; TiradoRives, J. J. Am. Chem. Soc. 1996, 118, 11225. 6. van Gunsteren, W. F.; Berendsen, H. J. C. Groningen Molecular Simula-tion (GROMOS) Library Manual; Biomos B. V.: Nijenborgh, Groningen, The Netherlands, 1987. 7. Leach, A. R. Molecular Modelling: Principles and Applications, 2nd ed.; Prentice Hall, 2001. 8. Marelius, J.; Kolmodin, K.; Feierberg, I.; Åqvist, J. J. Mol. Graph. Model. 1998, 16, 213. 9. Case, D. A.; Pearlman, D. A.; Caldwell, J. W.; Cheatham III, T. E.; Wang, J.; Ross, W. S.; Simmerling, C. L.; Darden, T. A.; Merz, K. M.; Stanton, R. V.; Cheng, A. L.; Vincent, J. J.; Crowley, M.; Tsui, V.; Gohlke, H.; Radmer, R. J.; Duan, Y.; Pitera, J. W.; Massova, I.; Seibel, G. L.; Singh, U. C.; Weiner, P. K.; Kollman, P. A. AMBER 7; University of California, San Francisco, 2002. 10. Kollman, P. Chem. Rev. 1993, 93, 2395. 11. Brandsdal, B. O.; Österberg, F.; Almlöf, M.; Feierberg, I.; Luzhkov, V. B.; Åqvist, J. Adv. Protein Chem. 2003, 66, 123. 12. Taylor, R. D.; Jewsbury, P. J.; Essex, J. W. J. Comput. Aided Mol. Des. 2002, 16, 151. 13. Jones, G.; Willett, P.; Glen, R. C. J. Mol. Biol. 1995, 245, 43. 14. Jones, G.; Willett, P.; Glen, R. C.; Leach, A. R.; Taylor, R. J. Mol. Biol. 1997, 267, 727.

Page 60: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

58

15. Friesner, R. A.; Banks, J. L.; Murphy, R. B.; Halgren, T. A.; Klicic, J. J.; Mainz, D. T.; Repasky, M. P.; Knoll, E. H.; Shelley, M.; Perry, J. K.; Shaw, D. E.; Francis, P.; Shenkin, P. S. J. Med. Chem. 2004, 47, 1739. 16. Halgren, T. A.; Murphy, R. B.; Friesner, R. A.; Beard, H. S.; Frye, L. L.; Pollard, W. T.; Banks, J. L. J. Med. Chem. 2004, 47, 1750. 17. Ewing, T. J. A.; Makino, S.; Skillman, A. G.; Kuntz, I. D. J. Comput. Aided Mol. Des. 2001, 15, 411. 18. Morris, G. M.; Goodsell, D. S.; Halliday, R. S.; Huey, R.; Hart, W. E.; Belew, R. K.; Olson, A. J. J. Comput. Chem. 1998, 19, 1639. 19. Halperin, I.; Ma, B. Y.; Wolfson, H.; Nussinov, R. Proteins: Struct., Func., Genet. 2002, 47, 409. 20. Gohlke, H.; Klebe, G. Angew. Chem. Int. Ed. 2002, 41, 2645. 21. Orozco, M.; Luque, F. J. Chem. Rev. 2000, 100, 4187. 22. Jorgensen, W. L.; Chandrasekhar, J.; Madura, J. D.; Impey, R. W.; Klein, M. L. J. Chem. Phys. 1983, 79, 926. 23. Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; Hermans, J. Intermolecular Forces; Pullman, B., Ed. Reidel: Dordrecht, 1981; pp 331. 24. Guillot, B. J. Mol. Liq. 2002, 101, 219. 25. Warshel, A.; Sharma, P. K.; Kato, M.; Parson, W. W. Biochim. Biophys. Acta 2006, 1764, 1647. 26. Simonson, T. Curr. Opin. Struct. Biol. 2001, 11, 243. 27. Feig, M.; Brooks, C. L. Curr. Opin. Struct. Biol. 2004, 14, 217. 28. Baker, N. A. Curr. Opin. Struct. Biol. 2005, 15, 137. 29. Born, M. Z. Phys. 1920, 1, 45. 30. Rosseinsky, D. R. Chem. Rev. 1965, 65, 467. 31. Warwicker, J.; Watson, H. C. J. Mol. Biol. 1982, 157, 671. 32. Sharp, K. A.; Honig, B. Annu. Rev. Biophys. Biophys. Chem. 1990, 19, 301. 33. Fogolari, F.; Brigo, A.; Molinari, H. J. Mol. Recognit. 2002, 15, 377. 34. David, L.; Luo, R.; Gilson, M. K. J. Comput. Chem. 2000, 21, 295. 35. Still, W. C.; Tempczyk, A.; Hawley, R. C.; Hendrickson, T. J. Am. Chem. Soc. 1990, 112, 6127. 36. Bashford, D.; Case, D. A. Annu. Rev. Phys. Chem. 2000, 51, 129. 37. Schaefer, M.; Karplus, M. J. Phys. Chem. 1996, 100, 1578. 38. Qiu, D.; Shenkin, P. S.; Hollinger, F. P.; Still, W. C. J. Phys. Chem. A 1997, 101, 3005. 39. Onufriev, A.; Bashford, D.; Case, D. A. J. Phys. Chem. B 2000, 104, 3712. 40. Calimet, N.; Schaefer, M.; Simonson, T. Proteins: Struct., Func., Genet. 2001, 45, 144. 41. Onufriev, A.; Bashford, D.; Case, D. A. Proteins: Struct., Funct., Bioinf. 2004, 55, 383. 42. Ben-Naim, A.; Marcus, Y. J. Chem. Phys. 1984, 81, 2016. 43. Eisenberg, D.; Mclachlan, A. D. Nature 1986, 319, 199. 44. Forsyth, W. R.; Antosiewiez, J. M.; Robertson, A. D. Proteins: Struct., Func., Genet. 2002, 48, 388.

Page 61: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

59

45. Schutz, C. N.; Warshel, A. Proteins: Struct., Func., Genet. 2001, 44, 400. 46. Beroza, P.; Case, D. A. Methods Enzymol. 1998, 295, 170. 47. Ido, E.; Han, H. P.; Kezdy, F. J.; Tang, J. J. Biol. Chem. 1991, 266, 24359. 48. Warshel, A.; Sussman, F.; King, G. Biochemistry 1986, 25, 8368. 49. Langsetmo, K.; Fuchs, J. A.; Woodward, C. Biochemistry 1991, 30, 7603. 50. Baker, W. R.; Kintanar, A. Arch. Biochem. Biophys. 1996, 327, 189. 51. Katti, S. K.; Lemaster, D. M.; Eklund, H. J. Mol. Biol. 1990, 212, 167. 52. Langsetmo, K.; Fuchs, J. A.; Woodward, C.; Sharp, K. A. Biochemistry 1991, 30, 7609. 53. Dillet, V.; Dyson, H. J.; Bashford, D. Biochemistry 1998, 37, 10298. 54. Georgescu, R. E.; Alexov, E. G.; Gunner, M. R. Biophys. J. 2002, 83, 1731. 55. Zhou, R. H.; Berne, B. J. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 12777. 56. Krol, M. J. Comput. Chem. 2003, 24, 531. 57. Alexov, E. G.; Gunner, M. R. Biophys. J. 1997, 72, 2075. 58. Sham, Y. Y.; Chu, Z. T.; Warshel, A. J. Phys. Chem. B 1997, 101, 4458. 59. Armstrong, R. N.; Cassidy, C. S. Drug Metab. Rev. 2000, 32, 327. 60. Mowbray, S. L.; Elfstrom, L. T.; Ahlgren, K. M.; Andersson, C. E.; Widersten, M. Protein Sci. 2006, 15, 1628. 61. Beroza, P.; Case, D. A. J. Phys. Chem. 1996, 100, 20156. 62. Gorfe, A. A.; Ferrara, P.; Caflisch, A.; Marti, D. N.; Bosshard, H. R.; Jelesarov, I. Proteins: Struct., Func., Genet. 2002, 46, 41. 63. Sitkoff, D.; Sharp, K. A.; Honig, B. J. Phys. Chem. 1994, 98, 1978. 64. Gouda, H.; Kuntz, I. D.; Case, D. A.; Kollman, P. A. Biopolymers 2003, 68, 16. 65. Nina, M.; Beglov, D.; Roux, B. J. Phys. Chem. B 1997, 101, 5239. 66. Swanson, J. M. J.; Adcock, S. A.; McCammon, J. A. J. Chem. Theory Comput. 2005, 1, 484. 67. Zhu, J.; Alexov, E.; Honig, B. J. Phys. Chem. B 2005, 109, 3008. 68. Onufriev, A.; Case, D. A.; Bashford, D. J. Comput. Chem. 2002, 23, 1297. 69. Feig, M.; Onufriev, A.; Lee, M. S.; Im, W.; Case, D. A.; Brooks, C. L. J. Comput. Chem. 2004, 25, 265. 70. Chipot, C.; Pearlman, D. A. Mol. Simulation 2002, 28, 1. 71. Beutler, T. C.; Mark, A. E.; Vanschaik, R. C.; Gerber, P. R.; Vangun-steren, W. F. Chem. Phys. Lett. 1994, 222, 529. 72. Zacharias, M.; Straatsma, T. P.; Mccammon, J. A. J. Chem. Phys. 1994, 100, 9025. 73. Jorgensen, W. L. Science 2004, 303, 1813. 74. Kollman, P. A.; Massova, I.; Reyes, C.; Kuhn, B.; Huo, S. H.; Chong, L.; Lee, M.; Lee, T.; Duan, Y.; Wang, W.; Donini, O.; Cieplak, P.; Srinivasan, J.; Case, D. A.; Cheatham, T. E. Acc. Chem. Res. 2000, 33, 889. 75. Srinivasan, J.; Cheatham, T. E.; Cieplak, P.; Kollman, P. A.; Case, D. A. J. Am. Chem. Soc. 1998, 120, 9401.

Page 62: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

60

76. Åqvist, J.; Medina, C.; Samuelsson, J. E. Protein Eng. 1994, 7, 385. 77. Carlson, H. A.; Jorgensen, W. L. J. Phys. Chem. 1995, 99, 10667. 78. McDonald, N. A.; Carlson, H. A.; Jorgensen, W. L. J. Phys. Org. Chem. 1997, 10, 563. 79. Åqvist, J.; Hansson, T. J. Phys. Chem. 1996, 100, 9512. 80. Hansson, T.; Marelius, J.; Åqvist, J. J. Comput. Aided Mol. Des. 1998, 12, 27. 81. Pierotti, R. A. Chem. Rev. 1976, 76, 717. 82. Westergren, J.; Lindfors, L.; Hoglund, T.; Luder, K.; Nordholm, S.; Kjel-lander, R. J. Phys. Chem. B 2007, 111, 1872. 83. Su, Y.; Gallicchio, E.; Das, K.; Arnold, E.; Levy, R. M. J. Chem. Theory Comput. 2007, 3, 256. 84. Abraham, M. H. J. Am. Chem. Soc. 1982, 104, 2085. 85. Åqvist, J.; Marelius, J. Comb. Chem. High. T. Scr. 2001, 4, 613. 86. Marelius, J.; Hansson, T.; Åqvist, J. Int. J. Quantum Chem. 1998, 69, 77. 87. Bjelic, S.; Nervall, M.; Gutierrez-de-Teran, H.; Ersmark, K.; Hallberg, A.; Åqvist, J. Cell. Mol. Life Sci. 2007, 64, 2285. 88. Almlöf, M.; Brandsdal, B. O.; Åqvist, J. J. Comput. Chem. 2004, 25, 1242. 89. Almlöf, M.; Åqvist, J.; Smalås, A. O.; Brandsdal, B. O. Biophys. J. 2006, 90, 433. 90. Lamb, M. L.; Tirado-Rives, J.; Jorgensen, W. L. Bioorgan. Med. Chem. 1999, 7, 851. 91. Stjernschantz, E.; Marelius, J.; Medina, C.; Jacobsson, M.; Vermeulen, N. P. E.; Oostenbrink, C. J. Chem. Inf. Model. 2006, 46, 1972. 92. Rizzo, R. C.; Udier-Blagovic, M.; Wang, D. P.; Watkins, E. K.; Smith, M. B. K.; Smith, R. H.; Tirado-Rives, J.; Jorgensen, W. L. J. Med. Chem. 2002, 45, 2970. 93. Sham, Y. Y.; Chu, Z. T.; Tao, H.; Warshel, A. Proteins: Struct., Funct., Bioinf. 2000, 39, 393. 94. Huang, D.; Caflisch, A. J. Med. Chem. 2004, 47, 5791. 95. Kuhn, B.; Kollman, P. A. J. Med. Chem. 2000, 43, 3786. 96. Zhou, R. H.; Friesner, R. A.; Ghosh, A.; Rizzo, R. C.; Jorgensen, W. L.; Levy, R. M. J. Phys. Chem. B 2001, 105, 10388. 97. Ersmark, K.; Nervall, M.; Hamelink, E.; Janka, L. K.; Clemente, J. C.; Dunn, B. M.; Blackman, M. J.; Samuelsson, B.; Åqvist, J.; Hallberg, A. J. Med. Chem. 2005, 48, 6090. 98. McGovern, S. L.; Shoichet, B. K. J. Med. Chem. 2003, 46, 2895. 99. Warren, G. L.; Andrews, C. W.; Capelli, A. M.; Clarke, B.; LaLonde, J.; Lambert, M. H.; Lindvall, M.; Nevins, N.; Semus, S. F.; Senger, S.; Tedesco, G.; Wall, I. D.; Woolven, J. M.; Peishoff, C. E.; Head, M. S. J. Med. Chem. 2006, 49, 5912. 100. De Corte, B. L. J. Med. Chem. 2005, 48, 1689. 101. Benjahad, A.; Croisy, M.; Monneret, C.; Bisagni, E.; Mabire, D.; Coupa, S.; Poncelet, A.; Csoka, I.; Guillemont, J.; Meyer, C.; Andries, K.; Pauwels, R.; de Bethune, M. P.; Himmel, D. M.; Das, K.; Arnold, E.; Nguyen, C. H.; Grierson, D. S. J. Med. Chem. 2005, 48, 1948.

Page 63: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

61

102. Himmel, D. M.; Das, K.; Clark, A. D.; Hughes, S. H.; Benjahad, A.; Oumouch, S.; Guillemont, J.; Coupa, S.; Poncelet, A.; Csoka, I.; Meyer, C.; Andries, K.; Nguyen, C. H.; Grierson, D. S.; Arnold, E. J. Med. Chem. 2005, 48, 7582. 103. Henriksson, L. M.; Unge, T.; Carlsson, J.; Åqvist, J.; Mowbray, S. L.; Jones, T. A. J. Biol. Chem. 2007, 282, 19905. 104. Velazquez-Campoy, A.; Kiso, Y.; Freire, E. Arch. Biochem. Biophys. 2001, 390, 169. 105. McQuarrie, D. A. Statistical Mechanics; University Science Books: Sausalito, CA, 2000. 106. Schlitter, J. Chem. Phys. Lett. 1993, 215, 617. 107. Karplus, M.; Kushick, J. N. Macromolecules 1981, 14, 325. 108. Dinola, A.; Berendsen, H. J. C.; Edholm, O. Macromolecules 1984, 17, 2044. 109. Schafer, H.; Mark, A. E.; van Gunsteren, W. F. J. Chem. Phys. 2000, 113, 7809. 110. Schafer, H.; Daura, X.; Mark, A. E.; van Gunsteren, W. F. Proteins: Struct., Func., Genet. 2001, 43, 45. 111. Schafer, H.; Smith, L. J.; Mark, A. E.; van Gunsteren, W. F. Proteins: Struct., Func., Genet. 2002, 46, 215. 112. Bakowies, D.; van Gunsteren, W. F. J. Mol. Biol. 2002, 315, 713. 113. Rueda, M.; Kalko, S. G.; Luque, F. J.; Orozco, M. J. Am. Chem. Soc. 2003, 125, 8007. 114. Hsu, S. T. D.; Peter, C.; van Gunsteren, W. F.; Bonvin, A. M. J. J. Bi-ophys. J. 2005, 88, 15. 115. Dolenc, J.; Baron, R.; Oostenbrink, C.; Koller, J.; van Gunsteren, W. F. Biophys. J. 2006, 91, 1460. 116. Andricioaei, I.; Karplus, M. J. Chem. Phys. 2001, 115, 6289. 117. Edholm, O.; Berendsen, H. J. C. Mol. Phys. 1984, 51, 1011. 118. Levy, R. M.; Karplus, M.; Kushick, J.; Perahia, D. Macromolecules 1984, 17, 1370. 119. Chang, C. E.; Chen, W.; Gilson, M. K. J. Chem. Theory Comput. 2005, 1, 1017. 120. Steinberg, I. Z.; Scheraga, H. A. J. Biol. Chem. 1963, 238, 172. 121. Finkelstein, A. V.; Janin, J. Protein Eng. 1989, 3, 1. 122. Gilson, M. K.; Given, J. A.; Bush, B. L.; McCammon, J. A. Biophys. J. 1997, 72, 1047. 123. Hermans, J.; Wang, L. J. Am. Chem. Soc. 1997, 119, 2707. 124. Lazaridis, T.; Masunov, A.; Gandolfo, F. Proteins: Struct., Func., Genet. 2002, 47, 194. 125. Luo, H. B.; Sharp, K. Proc. Natl. Acad. Sci. U.S.A. 2002, 99, 10399. 126. Gohlke, H.; Case, D. A. J. Comput. Chem. 2004, 25, 238. 127. Chen, W.; Chang, C. E.; Gilson, M. K. Biophys. J. 2004, 87, 3035. 128. Swanson, J. M. J.; Henchman, R. H.; McCammon, J. A. Biophys. J. 2004, 86, 67. 129. Minh, D. D. L.; Bui, J. M.; Chang, C. E.; Jain, T.; Swanson, J. M. J.; McCammon, J. A. Biophys. J. 2005, 89, L25.

Page 64: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

62

130. Chang, C. E. A.; Chen, W.; Gilson, M. K. Proc. Natl. Acad. Sci. U.S.A. 2007, 104, 1534. 131. Shoichet, B. K.; Leach, A. R.; Kuntz, I. D. Proteins: Struct., Func., Genet. 1999, 34, 4. 132. Böhm, H. J. J. Comput. Aided Mol. Des. 1994, 8, 243. 133. Jorgensen, W. L.; Ravimohan, C. J. Chem. Phys. 1985, 83, 3050. 134. Åqvist, J. J. Phys. Chem. 1990, 94, 8021. 135. Lee, F. S.; Chu, Z. T.; Warshel, A. J. Comput. Chem. 1993, 14, 161. 136. Oostenbrink, C.; Villa, A.; Mark, A. E.; Van Gunsteren, W. F. J. Com-put. Chem. 2004, 25, 1656. 137. Shirts, M. R.; Pitera, J. W.; Swope, W. C.; Pande, V. S. J. Chem. Phys. 2003, 119, 5740. 138. Kubo, M. M.; Gallicchio, E.; Levy, R. M. J. Phys. Chem. B 1997, 101, 10527. 139. Hess, B.; van der Vegt, N. F. A. J. Phys. Chem. B 2006, 110, 17616. 140. Levy, R. M.; Gallicchio, E. Annu. Rev. Phys. Chem. 1998, 49, 531. 141. Gallicchio, E.; Kubo, M. M.; Levy, R. M. J. Phys. Chem. B 2000, 104, 6271. 142. Peter, C.; Oostenbrink, C.; van Dorp, A.; van Gunsteren, W. F. J. Chem. Phys. 2004, 120, 2652. 143. Fleischman, S. H.; Brooks, C. L. J. Chem. Phys. 1987, 87, 3029. 144. Florian, J.; Warshel, A. J. Phys. Chem. B 1999, 103, 10282. 145. Burgess, M. A. Metal Ions in Solution; Ellis Horwood: Chichester, Eng-land, 1978. 146. Moelwyn-Hughes, E. A. Physical Chemistry, 2nd ed.; Pergamon Press: Oxford, England, 1961. 147. Kolb, P.; Huang, D.; Dey, F.; Caflisch, A. J. Med. Chem. 2008, 51, 1179.

Page 65: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations
Page 66: Challenges in Computational Biochemistry: Solvation and ...172052/FULLTEXT01.pdf · ACTA UNIVERSITATIS UPSALIENSIS UPPSALA 2008 Digital Comprehensive Summaries of Uppsala Dissertations

Acta Universitatis UpsaliensisDigital Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology 432

Editor: The Dean of the Faculty of Science and Technology

A doctoral dissertation from the Faculty of Science andTechnology, Uppsala University, is usually a summary of anumber of papers. A few copies of the complete dissertationare kept at major Swedish research libraries, while thesummary alone is distributed internationally through theseries Digital Comprehensive Summaries of UppsalaDissertations from the Faculty of Science and Technology.(Prior to January, 2005, the series was published under thetitle “Comprehensive Summaries of Uppsala Dissertationsfrom the Faculty of Science and Technology”.)

Distribution: publications.uu.seurn:nbn:se:uu:diva-8738

ACTA

UNIVERSITATIS

UPSALIENSIS

UPPSALA

2008