Calculations of Protein-Ligand Binding Affinities

download Calculations of Protein-Ligand Binding Affinities

of 24

Transcript of Calculations of Protein-Ligand Binding Affinities

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    1/24

    Calculation ofProtein-Ligand BindingAffinities

    Michael K. Gilson1 and Huan-Xiang Zhou2

    1Center for Advanced Research in Biotechnology, University of MarylandBiotechnology Institute, Rockville, Maryland 20850; email: [email protected]

    2Department of Physics and Institute of Molecular Biophysics and School ofComputational Science, Florida State University, Tallahassee, Florida 32306;email: [email protected]

    Annu. Rev. Biophys. Biomol. Struct 2007. 36:2142

    First published online as a Review in Advance onJanuary 2, 2007

    The Annual Review of Biophysics and BiomolecularStructure is online at biophys.annualreviews.org

    This articles doi:10.1146/annurev.biophys.36.040306.132550

    Copyright c 2007 by Annual Reviews.All rights reserved

    1056-8700/07/0609-0021$20.00

    MKG dedicates this article to the memory of hisfather, Saul Bernard Gilson, MD: scholar, poet,and physician, 19222006.

    Key Words

    computation, drug-design, entropy, free energy, modeling

    Abstract

    Accurate methods of computing the affinity of a small molecule with

    a protein are needed to speed the discovery of new medications andbiological probes. This paper reviews physics-based models of bind-

    ing, beginning with a summary of the changes in potential energysolvation energy, and configurational entropy that influence affinity

    and a theoretical overview to frame the discussion of specific computational approaches. Important advances are reported in modeling

    protein-ligand energetics, such as the incorporation of electronicpolarization and the use of quantum mechanical methods. Recen

    calculations suggest that changes in configurational entropy stronglyoppose binding and must be included if accurate affinities are to be

    obtained. The linear interaction energy (LIE) and molecular me-

    chanics Poisson-Boltzmann surface area (MM-PBSA) methods areanalyzed, as are free energy pathway methods, which show promise

    and may be ready for more extensive testing. Ultimately, major im-provements in modeling accuracy will likely require advances on

    multiple fronts, as well as continued validation against experiment.

    21

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    2/24

    Ligand: in general,

    any molecule thatbinds another; usedhere for a drug-likeorganic molecule,typically ofmolecular weightless than 600 Da

    Contents

    INTRODUCTION . . . . . . . . . . . . . . . . . 22MODELING THE PHYSICAL

    CHEMISTRY OF BINDING:FORCES AND FLEXIBILITY . . . 23

    Potential Energy . . . . . . . . . . . . . . . . . 23

    Effects of Water, DissolvedElectrolytes, and pH. . . . . . . . . . . 24

    Theoretical Framework for

    Calculating Affinities . . . . . . . . . . 26Strain, Configurational Entropy,

    and Preorganization . . . . . . . . . . . 27

    APPROACHES TOCALCULATING AFFINITIES . . 28

    Docking and Scoring . . . . . . . . . . . . . 28Free Energy Methods . . . . . . . . . . . . 30

    PROSPECTS AND

    CHALLENGES . . . . . . . . . . . . . . . . . 33

    INTRODUCTION

    Molecular recognition is central to biology,

    forming the foundation for precise hormonalcontrol of distant organs, immune targeting

    of nonself proteins, and the specificity of en-zymatic catalysis, for example. It also offers a

    powerful mechanism for exogenous control ofbiological systems, and many medications and

    biological probes act by binding and usually

    inhibiting a specific macromolecular target.Indeed, the discovery of a ligand that binds

    a targeted protein with high affinity is a ma-jor preoccupation of early-stage drug discov-

    ery and also of chemical genomics projectsseeking small-molecule inhibitors to eluci-

    date gene function. However, discovering asmall molecule that binds a specific protein

    tightly, while retaining favorable pharmaco-logical properties, can be a major and costly

    challenge.Computation can speed this process (62)

    by artificial intelligenceand machine-learning

    technologies and through simulation andmodeling. In particular, structure-based mod-

    eling uses the three-dimensional atomic co-

    ordinates of the targeted protein to calculatethe binding free energy of a proposed lig-

    and, or at least to rank candidate ligands ac-cording to their predicted affinities for the

    target. This information is then available toguide molecular design and synthesis. Ide-

    ally, computational methods would be as ac-curate as experiment, with binding affini-ties correct to within 1 kcal mol1, but this

    goal has not been achieved. Indeed, cal-culating protein-ligand affinities is difficult

    One reason is that the free energy of bind-ing is typically a small difference between

    large numbers, namely the interactions of theligand and protein with each other versus

    their interactions with water and counteri-ons in their unbound state. Also, the inter-

    atomic forces are strong and short ranged,making for an energy function sharply de-pendent on the details of molecular confor-

    mation. Finally, the protein and the ligandare flexible and have many degrees of free-

    dom; therefore exploring all potentially rele-vant conformations is a large computationa

    task.Despite these difficulties, computer mod-

    eling has enormous potential to make accu-rate predictions and provide insights that can

    guide molecular design. In this paper, we re-view recent progress and discuss prospects for

    further advances. We begin with a discussionof the forces involved in molecular recogni-

    tion and how they are modeled. In partic-

    ular, many approaches to calculating bind-ing affinities can be analyzed by identifying

    how they treat the potential energy of theprotein and ligand as a function of confor-

    mation, the influence of the aqueous solventand the conformational flexibility of the two

    molecules. This discussion provides the basisfor a review of approaches to modeling bind-

    ing, which are broadly categorized as dock-ing and scoring and free energy methods

    A short conclusion touches on the prospectsfor further progress. Space limitations pre-

    cludeaddressing manyinterestingtopics, such

    22 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    3/24

    as the treatment of protonation equilibria,

    the possible consequences of macromolecularcrowding on binding, the automated design

    of ligands, and the special challenges posedby nucleic acid targets. For the same rea-

    son, the literature is summarized with a broadbrush, using representative, rather than ex-

    haustive citations, in hopes of conveying ina few pages the structure of the field and pro-viding a framework that helps the reader make

    sense of a range of papers.

    MODELING THE PHYSICALCHEMISTRY OF BINDING:FORCES AND FLEXIBILITY

    Perhaps someday there will be computers fastenough to compute binding affinities from

    first-principles quantum mechanical simula-tions of all the nuclei and electrons in a fem-toliter solvent droplet, with dissolved protein

    and ligand molecules. For now, however, onemust rely on a theoretical framework that

    yields the affinity from calculations on justone protein and one ligand, and on approx-

    imations that allow energies and forces to becomputed more rapidly than does an ab ini-

    tio quantum mechanical treatment. In gen-eral, such approximations entail reducing the

    number of degrees of freedom that are treatedexplicitly and replacing them with new forces

    that account for the implicit degrees of free-

    dom. For example, one can model the wa-ter around a protein by explicitly simulating

    the motions of thousands of water molecules,allowing the properties of water to emerge

    automatically from the simulation. Alterna-tively, one can discard the explicit water

    molecules and instead use an implicit watermodel consisting of new energy terms for the

    hydrophobic effect and dielectric screening.Such models can provide a useful intuitive

    understanding of the system, but it is alsoimportant to remain alert to their potential

    weaknesses.

    Below we outline current thinking on keyenergycomponents that drive or opposebind-

    Strain: the energycost of adopting thebound conformation

    ing, breaking them into potential energy (as

    if in vacuo) and the effects of solvent. Thena theoretical framework for computing affini-

    ties based on these forces is described, fol-lowed by some of the implications of the

    theory for binding thermodynamics, such asstrain and entropy.

    Potential Energy

    The potential energy of an isolated moleculeor complex in a given conformation can be

    computed by quantum mechanical methodsthat explicitlyaccount for the electronicstruc-

    ture. Such quantum calculations can be donein various ways, but a central division is

    that between more time-consuming ab ini-tio methods and faster semiempirical meth-

    ods, which substitute fitted parameters forquantities that would otherwise need to becalculated. The details of these approaches

    are beyond the scope of this review. Quan-tum methods in principle are broadly ap-

    plicable, but they are time-consuming rel-ative to the force-field models discussed in

    the next paragraph, typically requiring min-utes to months of computer time, depending

    upon the size of the system and the level ofthe calculation. Therefore, it is difficult to

    use them in methods that calculate the en-ergy of many different conformations. It is

    also important to be aware that many ab initio

    methods markedly underestimate the attrac-tive forces due to electron-electron correla-

    tions between atoms, i.e., dispersion forces,the attractive part of the van der Waals inter-

    action.Empirical force fields (8) represent a

    computationally fast alternative to quantummechanical calculations of the potential en-

    ergy, because they encapsulate the chief con-sequences of electronic structure in a few

    energy terms that are easily calculated, of-ten in well under 1 s for one conforma-

    tion of a system. Current examples includeCHARMM (15), AMBER (30), and GRO-

    MOS (121). A typical functional form is given

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 23

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    4/24

    Electronicpolarization:inducedredistribution ofelectronic charge inan atom or molecule

    by Equation 1:

    U =Nbonds

    i=1

    kb,i(b i b0,i)2

    +

    Nanglesi=1

    k,i(i 0,i)2

    +

    Ndihedralsi=1

    k,i(cos(nii 0,i))

    +

    Nimprpdhdi=1

    k,i(i 0,i)2

    +

    Natomsi

    Natomsj=i

    Ai j

    r12i j

    Bi j

    r6i j

    +

    Natomsi

    Natomsj>i

    qiqjDri j

    . 1.

    The first two terms treat bond stretch-

    ing and angle bending by harmonic springswith large force constants. The third termaccounts for bond torsions (dihedrals), and

    the fourth, improper dihedral, term is usedto keep planar groups planar. The fifth term

    accounts for the van der Waals interactionsamong atoms with a short-ranged hard re-

    pulsion that prevents atoms from overlap-ping and a slightly longer-ranged weak at-

    traction attributable to correlated motionsof the electrons (dispersion forces). The fi-

    nal term accounts for Coulombic interactionsbetween atoms, which result chiefly from thefact that electrons tend to concentrate around

    atoms with large electronegativities and be-come depleted elsewhere, leading to negative

    and positive partial atomic charges; those withcharges of large magnitude are polar atoms.

    It is evident from Equation 1 that an empir-ical energy function contains many parame-

    ters, such as the spring force constants k, thevan der Waals parameters A and B, and the

    atomic charges q. These parameters can of-ten be obtained by fitting to high-level quan-

    tum calculations, but it is best if they canbe tested and adjusted by computing physical

    properties, such as solvation energies of model

    compounds,and comparing them with experi-mental data. When suitable experimental data

    are not available, it can be difficult to generate

    trustworthy parameters.Newer force fields are becoming more

    complex in order to achieve greater accuracyMore detailed models of the charge distri-

    bution can be employed (87, 102, 115), astouched on below, andthere is alsogreatinter-

    est in accounting for electronic polarization,the redistribution of atomic charges induced

    by changes in the electrical fields to which the

    atoms are subjected (66, 77).

    Effects of Water, DissolvedElectrolytes, and pH

    Protein-ligand binding normally occurs in a

    salt-water environment, which has a strongeffect on energetics. Water has a dielectric

    constant (D in Equation 1) of about 80whereas the dielectric constant of vacuum is

    1. This leads to a new, one-body solvation en-

    ergy for each atomiccharge, which arises fromthe favorable interaction between the charge

    andthe high-dielectric environment (12). As aconsequence,therecanbeasubstantialenergy

    penalty for moving the polar part of a ligandout of water and into the binding site. In ad-

    dition, water screens the charge-charge inter-actions of fully hydrated atoms by80-fold

    as indicated in the Coulombic term in Equa-tion 1. However, atoms in a protein-ligand in-

    terface are sequestered from the solvent andtherefore interact with an effective dielectric

    constant less than 80. In general, atoms that

    are further apart are more likely to interactthrough solvent, and this idea led to the intro-

    duction of a crude screening model consistingof a distance-dependent dielectric. For atoms

    i and j in Equation 1, Dij = Crij, where C isa constant often set to 4 and rij is the inter-

    atomic distance. This model allows one of thechief effects of the solvent to be accounted for

    in a computationally efficient manner, andit isusedinanumberofligand-proteindockingal-

    gorithms. However, it fails to account for theone-body solvation energy of each atom. In

    addition, the electrostatic interaction of two

    24 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    5/24

    atoms actually depends not only on their mu-

    tual distance, but also on the positions of allthe other protein and ligand atoms, because

    these determine wherethe high-dielectricsol-vent can penetrate.

    These additional complexities can be cap-tured by continuum electrostatics models,

    which treat the protein and ligand as low-dielectric regions with embedded atomiccharges, surrounded by the high-dielectric

    aqueous solvent (55). In particular, numeri-cal solutions of the Poisson-Boltzmann (PB)

    equation offer a reasonably efficient methodof accounting for the subtleties of molec-

    ular electrostatics, including the additionalelectrostatic screening produced by dissolved

    salts. The PB equation usually can be solvedfor a protein in seconds to minutes of com-

    puter time for a biomolecular system, and ityields the interaction energy of the protein,

    ligand, or complex with the water in which

    it is immersed. This electrostatic part of thesolvation energy can be added to the poten-

    tial energy obtained by a force field or othermethod to provide a more complete account-

    ing of conformational energetics. Whenmanyconformations need to be processed, faster

    generalized Born (GB) models are valuable asapproximations to the full PB model (34, 51,

    58, 59, 80, 97). These are based on a recog-nition that dielectric screening of a charge-

    charge interaction correlates with the degreeto which each charge interacts with surround-

    ing water (43).

    Water also gives rise to the hydropho-bic effect, the tendency of water molecules

    to drive nonpolar solutes together (68). Thispromotes the association of nonpolar sur-

    faces of the ligand and the protein. The hy-drophobic effect is often accounted for by an

    additional solvation energy term that is pro-portional to molecular surface area, with a

    positive coefficient. The effect is to add a pos-itive (unfavorable) solvation energy to con-

    formations with more surface area and thusto favor binding, which reduces surface area.

    Combining the PB or GB electrostatics mod-

    Poisson-Boltzmann(PB) equation: apartial differentialequation solved toobtain the molecularelectrostatic

    potential, whichdepends on thelocations and valuesof atomic charges,the dielectricconstant as afunction of position,and the ionicstrength

    Generalized Born(GB) model: a fastapproximation to thePoisson-Boltzmannequation

    PBSA:Poisson-Boltzmannsurface area

    GBSA: generalizedBorn surface area

    Implicit solventmodel: a model thaaccounts for solventwithout explicitlymodeling individualwater molecules

    MD: moleculardynamics

    MC: Monte Carlo

    els with such a surface area term yields the

    PBSA (112) and GBSA (97) solvation mod-els, respectively. These are called implicit sol-

    vent models because they do not treat anywater molecules explicitly. Once parameter-

    ized, the PBSA and GBSA models providerather good agreement with experimental sol-

    vation energies of small model compounds(10, 112), but they may be less accurate formore complex molecules, such as proteins,

    that can bind or completely sequester individ-ual water molecules. Unfortunately, it is not

    straightforward to generate experimental datathat directly address this issue, but computa-

    tional studies are beginning to be applied to it(50, 83, 91, 118).

    The influence of solvent on binding canalso be treated with molecular dynamics (MD)

    or Monte Carlo (MC) simulations that in-clude thousands of explicit water molecules

    modeled with an empirical force field (6, 64,

    102). Dielectric screening, the solvation ofpolar groups, and the hydrophobic effect all

    emerge automatically with this approach. Inaddition, it should provide a better treatment

    of bound and sequestered water molecules, atleast in principle. However, an explicit treat-

    ment of solvent is substantially more costlycomputationally than an implicit model, by

    perhaps an order of magnitude, depending onthe specifics of the comparison.

    Physiological solutions also buffer protonsat a specific pH and thus control the elec-

    trical charges of acidic and basic groups of

    both the protein and the ligand. As a re-sult, the pH can strongly influence confor-

    mational preferences and binding affinities.Indeed, protonation and binding are thermo-

    dynamically linked, so that the pH affectsthe apparent binding affinity and binding in

    turn affects protonation states. A number ofresearch groups are working on methods of

    predicting protonation states, which will un-doubtedly be important for computing accu-

    rate affinities. The potential for many drug-like molecules to exist in multiple tautomeric

    states poses similar challenges.

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 25

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    6/24

    Theoretical Framework forCalculating Affinities

    Given a method of computing the potential

    and solvation energies of a protein, a ligand,and their complex, a theory of binding is still

    needed to compute their affinity. One widelyused approach (42) is based on recognition

    that the free energy of binding is the changein free energy when one protein and one lig-

    and react to form one complex. The change infree energy of a solution when one complex is

    added to the system is defined as the complexschemical potential, P L; similarly, the changein free energy when one protein and one lig-

    and are removed is P L. Thus, the freeenergy of binding is G = P L P L.

    Statistical thermodynamics provides an ex-

    pression for the chemical potential of the pro-

    tein, for example, as

    P = RTln

    QNP+1,NS

    QNP,NS

    = RTln

    82

    CP

    e (U(rP)+W(rP))/RTdrp

    ,

    2.

    where R is the gas constant; T is the abso-

    lute temperature; QNP+1,NS and QNP,NS arethe canonical partition functions of solutions

    ofNP + 1 proteins and NS solvent moleculesandNP proteins andNS solvent molecules, re-

    spectively; the factor of 82 results from thefact that overall rotation of the protein does

    not affect the integral over the other coordi-

    nates of the system; CP is the concentration

    of the protein; U(rP) is the potential energy of

    the protein as a function of its internal coor-dinates rP and thus of its conformation; and

    W(rP) is the corresponding solvation energy.

    (A mass-dependent prefactor that cancels in

    the final expression for the binding affinity hasbeen omitted, as has a small pressure-volume

    term.) The Boltzmann factor in the integralis larger for more stable conformations of the

    protein, so that the more low-energy confor-mations the protein can adopt, the greater the

    integral and the more negative its standardchemical potential, implying greater stabilityand less reactivity. Also, the higher the con-

    centration CP, the more positive the standardchemical potential, implying greater reactiv-

    ity, consistent with the law of mass action.Supplying analogous formulae for the

    chemical potentials of the ligand and thecomplex allows the free energy of binding to

    be written as

    G = RTln

    182

    CPCLCP L

    e (U(rP L )+W(rP L ))/RTdrP L

    e (U(rP)+W(rP))/RTdrP

    e (U(rL )+W(rL ))/RTdrL . 3

    For the most part, the integrals in this expres-

    sion are considered to range over all confor-mations, but only conformations that are low

    in energy actually require attention; for ex-ample, conformations with highly stretched

    bonds can be neglected. However, a sub-tlety arises for the protein-ligand complex

    because the integral should include only con-

    figurations in which the ligand is consideredboundtotheprotein.Fortightcomplexes,any

    integration domain that includes the lowest-energy configurations and not too much more

    suffices (42); for weak binding, a more in-volved definition is needed (88).

    Replacing each concentration in Equation3 by the standard concentration C (1 M)

    yields the standard free energy of binding:

    Go

    = RTln Co

    82e (U(rP L )+W(rP L ))/RTdrP L

    e (U(rP)+W(rP))/RTdrP

    e (U(rL )+W(rL ))/RTdrL

    . 4

    The factor of C makes it apparent that the

    higher the standard concentration, the morenegative the standard free energy of binding

    again consistent with the law of mass actionThe relationship ofGo to the equilibrium

    26 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    7/24

    constantK is established by recognizing thatG= 0 when the system is at equilibrium.Equation 3 then gives the equilibrium ratio

    of concentrations as:

    CP LCP CL

    eq

    =1

    82e

    (U(rP L )+W(rP L ))/RTdrP Le (U(rP)+W(rP))/RTdrP

    e (U(rL )+W(rL ))/RTdrL

    5.

    Multiplying both sides byCo and using Equa-tion 4 yields that

    K = e Go/RT =

    CP LCo

    CPCL

    eq

    6.

    Equation 4 can be used to derive variousmethods of computing binding affinities, in-

    cluding Double Decoupling andMining Min-ima (23, 52). Furthermore, Equation 4 can be

    manipulated into the form

    Go = UP L UP UL + WPL

    WP WL TSoconfig, 7.

    where UP L and WP L, for example, arethe Boltzmann-averaged potential and sol-

    vation energies of the complex; and Soconfig,the change in configurational entropy, is the

    change in the entropy associated with ligandand protein motions, including translational

    motion, for binding at standard concentra-tion. The solvation terms are actually solva-

    tion free energies and include both energetic

    and entropic components. Equation 7 is rec-ognizable as the basis of most physics-based

    scoring functions as well as the molecularmechanics Poisson-Boltzmann surface area

    (MM-PBSA) method (46, 114).Some methods of computing binding

    affinities yield predicted values ofGo thatdo not depend on the standard concentra-

    tion. This is problematic because the stan-dard concentration is arbitrary. Indeed, if it

    were changed from 1 M to 1 nM, for example,then every experimental standard free energy

    of binding would be reported as RTln109 =

    12kcalmol1 more positive. Therefore, a casecan be made that the success of a model is

    fortuitous if the models predictions do not

    Configurationalentropy: theentropy associatedwith motions of theligand and/or protein

    depend on standard concentration. On the

    other hand, a model that omits standard con-centration can legitimately provide the differ-

    ence between two binding affinities, because

    the standard concentration cancels in suchdifferences.

    Strain, Configurational Entropy,and Preorganization

    Equations 4 and 7 not only provide a basis forcomputational models, but also have impor-

    tant implications that can be understood intu-itively. In particular, they bear directly on theconcepts of strain, configurational entropy,

    and preorganization. If the ligand adopts astrained conformation in the bound state, the

    rise of its internal energy on binding will makea positive contribution to UP L UP UL

    and thus oppose binding. Alternatively, if theligand accesses a wide range of equally sta-

    ble conformations in the free state, but only anarrow range of conformations in the bound

    state, its configurational entropy will drop andthus oppose binding. Finally, if the bound

    conformation of the ligand corresponds to an

    especially stable conformation of the free lig-and, then the ligand is preorganized for bind-

    ing, and both the strain and entropic penaltieswill be low. The same considerations apply to

    the protein.Changes in configurational entropy upon

    binding are the subject of a literature too com-plex and extensive to be summarized here; a

    key problem has been the lack of a methodto compute these quantities. The Mining

    Minima technique does allow such calcula-tions, and applications to host-guest model

    systems (23, 26) and a protein-ligand sys-

    tem (22) indicate that the entropic penalty ofEquation 7 is large, opposing binding nearly

    as strongly as the energy changes favor it.

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 27

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    8/24

    Thus, the loss in configurational entropy can-

    cels most of the binding energy, and the netbinding free energy is a small difference be-

    tween larger numbers. Importantly, leavingout the configurational entropy, and thus fo-

    cusing only on energy, leads to significantlyworse correlation with experimental free en-

    ergies of binding. Nonetheless, there is signif-icant entropy-energy correlation, and scalingthe energy terms by a factor of perhaps 0.1 to

    0.3 yields numbers in the experimental range.Similar results should hold for protein-ligand

    binding.

    APPROACHES TOCALCULATING AFFINITIES

    Docking and Scoring

    Docking methods try to identify the moststable conformation of the ligand-protein

    complex using a simplified energy model, fre-quently an empirical force field with a sim-

    ple solvent model for the sake of computa-tionalspeed.Theconformationisthenusedto

    assign a binding energy or score to the ligand,sometimes using the same energy model that

    was used during docking and sometimes witha more sophisticated and time-consuming

    model. Reliance on a single bound conforma-tion in effect assumes that this conformation

    is the only one that is significantly occupied,

    so that it dominates the averages in Equation7. Pioneering initial approaches (32, 45, 76)

    have been followed by further advances fromthe same groups and many others, as recently

    reviewed (14, 113), and a number of meth-ods are actively used for computer-aided drug

    discovery (110).The proliferation of docking and scor-

    ing methods also has inspired evaluations andcomparisons (25, 69, 125, 127), which are of

    particular interest to drug companies mak-ing purchasing decisions. Some of the most

    commonly used criteria are a methods abil-

    ity to reproduce the correct bound conforma-tion of a ligand-protein complex, its ability to

    assign better energy scores to known high-

    affinity ligands than to a large set of decoy

    compounds (enrichment), and the ability togenerate scores that correlate with measured

    binding affinities of known ligands. Over-all, existing methods frequently yield accurate

    bound complexes and can usually provide sig-nificant enrichment. However, the results are

    system dependent, and different methods per-form better on different systems. In addition,

    scores are at best weakly predictive of affini-

    ties across a series of known ligands. Carryingout a fair evaluation poses its own challenges

    (29); for example, enrichment studies pro-vide different results when the known ligands

    have nanomolar versus micromolar affinitiesDocking results also depend on procedural

    details outside the actual docking protocolsuch as the choice of protonation sites on the

    ligands (75, 120). Results are sometimes im-proved when multiple scoring functions are

    combined in a single consensus score, withthe idea that a ligand that ranks high on mul-

    tiple scoring functions is more likely to bind

    tightly (7, 24, 90). Nevertheless, the gains ofthis approach are limited by the simplicity of

    typical scoring functions, which are designedfor computational throughput.

    As a consequence, there is great interest inmore sophisticated and hence more accurate

    scoring functions. If these prove to be slow, itwill still be possible to prescreen compounds

    with a fast scoring function and apply themore accurate methods only to compounds

    that pass the prescreen. The fast method alsocan provide initial docked conformations to

    initiate calculations with the slower one. The

    following subsections review efforts to in-crease the sophistication of docking and scor-

    ing methods.

    Improved potential energy models. Oneimportant direction involves using potential

    energy models that are more advanced than atypical empirical force field. A particularly in-

    teresting study provides evidence that dock-ing accuracy is substantially better when the

    energy model accounts for the redistributionof ligand charges due to the electric field of

    28 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    9/24

    the binding site, i.e., for electronic polariza-

    tion (28). The procedure involves iterativelyadjusting partial atomic charges via quantum

    mechanicalcalculations during a docking pro-cedure. Rather large shifts in charge were ob-

    served, up to an average of 0.4 e across allatoms for some ligands. A number of other

    studies also explore force fields with more de-tailed treatments of electrostatics (31, 37, 87),and a different style of force field, designed for

    use with metalloenzymes, has now been usedto model protein-ligand systems (48).

    Another approach has been to use a quan-tum mechanical treatment of the ligand, and

    eventheprotein,ratherthanrelyonanempir-ical force field (47, 71, 98, 122). Because such

    methods capture dispersion forces poorly, avan der Waals dispersion term has been com-

    bined with the quantum approach, as done inan intriguing trial application to 165 protein-

    ligand systems (99). A recent review provides

    a useful perspective on the use of quantumcalculations in structure-based drug discovery

    (95).

    Improved solvent models. Many scoringfunctions have relied on simple models of the

    solvent, such as the distance-dependent di-electric. These fail to capture the sensitivity

    of the interactions between atomic charges tothe details of their locations relative to the

    solvent, as well as the fact that polar groupshave strong, favorable interactions with the

    solvent that may be lost upon binding. This

    issue has been addressed by incorporatingPBSA and GBSA solvation models into dock-

    ing and scoring, and a number of papers re-port improvements in accuracy (56, 84, 86,87,

    111, 135, 136). However, increased accuracyin some systems requires careful assignment

    of protonation and tautomer states (38, 56).It seems that, in these cases, removing one

    source of error exposes another one.Several studies also address the formation

    ofwaterbindingsitesduringdocking(41,100,123, and references therein), but the methods

    tried so far do not seem to markedly improve

    accuracy. In addition, it is not yet clear how to

    adjust the score of a ligand if it is predicted to

    bind with a water molecule.

    Accounting for conformational changes of

    the receptor. The first docking algorithms

    typically treated the protein as rigid, althoughbinding sites in reality adjust to accommodate

    different ligands. This simplification can beproblematic, because even a small steric over-

    lap can make a good ligand look like a misfit

    if it is docked into a single conformation ofthe receptor that was solved with a different

    ligand bound. However, because the recep-tor possesses many internal degrees of free-

    dom, treating it as flexible can vastly expandthe problem of discovering the most stable

    conformation of the protein-ligand complex.Moreover, as the number of candidate energy

    minima rises, the energy model must becomeincreasingly accurate in order to identify the

    global minimum correctly. A wide variety ofapproaches have been devised to address this

    problem, as previously reviewed (18). Meth-

    ods range from using a single conformation ofthe receptor, but softening it to approximate

    the consequences of flexibility (60); to dock-ing into a single representation of the binding

    site obtained by averaging across multiple re-ceptor structures (16, 74, 93); to docking into

    multiple discrete receptor conformations (1,20, 39, 130, 131); to reoptimizing the confor-

    mation of the binding site during the dockingcalculation itself (65, 70, 108, 133, 134). The

    relative merits of the various approaches arenot yet clear, though one study provides ev-

    idence that soft docking yields worse enrich-

    ment of known ligands than does docking intomultiple structures of the receptor (39).

    Configurational entropy. Even if a docking

    algorithm always identifies the global energyminimum of a protein-ligand complex with

    a perfectly accurate potential and solvationenergy function, it will not yield an accurate

    binding free energy. One reason is that thebinding free energy includes a loss in con-

    figurational entropy due to the decrease infreedom of the ligand and the protein upon

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 29

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    10/24

    binding. As noted above, the loss of configu-

    rational entropy is nearly as large as the gainin favorable binding energy and correlates

    roughly with the energy change. Therefore,it is expected that a model that accounts for

    changes in potential and solvation energy, butneglects the loss of configurational entropy,

    will tend to overestimate binding affinities. Infact, the energy terms in a physics-based scor-ing function had to be scaled down roughly

    10-fold to bring the results into range withexperiment (89), as previously noted (26); a

    similar result is observed in calculations withthe MM-PBSA method when configurational

    entropy is neglected (17).A number of scoring methods use the

    rigid-rotor/harmonic-oscillator formalism toseparate the ligands loss of entropy into a

    rotational and translational part that resultsfrom trapping in the binding site and another

    part associated with the ligands loss of flex-

    ibility. It is generally assumed that any en-tropy change associated with the hard degrees

    of freedom (bond stretches and angle bends)is negligible. This is reasonable because bond

    lengths andbond angles areunlikely to be per-turbed much by the noncovalent forces that

    drive protein-ligand binding. It is also widelyassumed that changes in rotational and trans-

    lational entropy are either negligible or con-stant across all binding reactions, but this view

    is not well founded. In fact, the translationaland rotational motions of a bound ligand de-

    pend on exactly how tightly it is held (23, 26),

    and this can vary not only from one complexto another, but also among alternative bound

    conformations of a single complex (26). Thistopic poses nontrivial theoretical issues and is

    now the subject of increasing computationalscrutiny (2, 33, 78, 85, 117, 138). For example,

    a recent study reports improved docking ac-curacy with a novel method of estimating the

    entropy of a bound ligand from the rangesof rotational, translational, and torsional co-

    ordinates associated with multiple dockedconformations (107).

    Changes in entropy due to the decreased

    flexibility of the ligand upon binding are com-

    monly regarded as resulting from a reduction

    in the number of energetically accessible lig-and rotamers upon binding and are often ac-

    counted for with empirical termsproportionalto the number of rotatable bonds of the ligand

    affected by binding (9, 36, 44, 101, 124). Thecorresponding free energy penalties are typi-

    cally on the order of 0.4 kcal mol1

    torsion1

    but run at least to 1 kcal mol1 torsion1 (99)However, a simple argument suggests that

    such estimates markedly overestimate the lossof entropy due to the drop in accessible ro-

    tamers. For a ligand with 10 rotatable bondsthe torsional penalty would be 4 kcal mol1

    corresponding to a drop of e4/RT = 800-fold in the number of equally stable ligand

    rotamers upon binding. This implies a mini-mum of 800 equally stable, rotamerically dis-

    tinct ligand conformations in the free stateThis number seems improbably large. It isnot clear whether the literature has directly

    addressed this issue, but programs designedto generate a large number of diverse ligand

    conformations typically yield far fewer con-formations (13, 72). Moreover, of these con-

    formations, only a few are likely to be within1 kcal mol1 of the global energy minimum

    Therefore, it is unlikely that rotamer count-ing accounts for much of a ligands entropy

    loss upon binding. Recent studies of modelsystems do reveal substantial losses in con-

    figurational entropy upon binding, but they

    attribute these losses to the fact that the en-ergy wells of a bound complex are narrower

    than those of the free species, rather than toa reduction in the number of stable rotamers

    (22, 26). Overall, it appears that a deeper un-derstanding of changes in configurational en-

    tropy upon binding could lead to improvedaccuracy in docking and scoring.

    Free Energy Methods

    Free energy methods use conformationa

    sampling to generate thermodynamic av-

    erages, in contrast with docking methodswhich focus on a single bound conforma-

    tion. The use of conformational averaging is

    30 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    11/24

    advantageous because it removes sensitivity to

    the details of the single representative confor-mation relied on by docking methods. On the

    other hand, free energy methods need morecomputer time in order to generate unam-

    biguous, converged results. Also, the confor-mational searches tend not to be wide ranging

    and thus may need to be seeded with severalplausible docked conformations. Two generalapproaches are considered here: end-point

    methods, which generate conformations ofonly the free and bound species and compute

    the binding free energy by taking a differ-ence, and pathway methods, which compute

    the sum of small changes along a multisteppathway connecting the initialand final states.

    End-Point methods. The linear interaction

    energy (LIE) method (3) involves runningtwo MD simulations with an empirical forcefield: one for the ligand in solution, and the

    other for the ligand in the protein bind-ing site. Snapshots saved from the simula-

    tions represent Boltzmann ensembles of con-

    formations and are used to compute theBoltzmann-averaged electrostatic (Uelec) and

    van der Waals (Uvdw) interaction energies ofthe ligand with its environment in the bound

    and free states. The binding free energy is

    then estimated as

    G o (Uelecbound Uelecfree)

    +(Uvdwbound Uvdwfree), 8.

    where the angle brackets indicate Boltzmannaverages. Changes in the internal energy of

    the solvent and the protein are considered tobe accounted for by the factors and , on

    the basis of their reorganization energy in re-sponse to the ligand (3). As recently summa-

    rized (19), good correlations between calcu-lation and experiment have been reported in

    a number of publications, both with the stan-dard coefficients = 0.18 and = 0.33 (19)

    and with system-specific values of these co-

    efficients (126). An additional surface-area-dependent term has also been incorporated

    (61), and more recent variants speed the MD

    LIE: linearinteraction energy

    calculations by using implicit treatments of

    the solvent (19, 135). In general, it does notseem that a single set of coefficients can be

    applied to all systems; therefore some experi-mental binding data for the system of interest

    may be needed before this method can be usedeffectively.

    Nonetheless, the quality of the results issomewhat surprising, because LIE does notaccount explicitly for standard concentration

    or for changes in the configurational entropyor the internal energy of the ligand. The

    method may be successful in part because itis generally used to compare ligands within a

    single chemical series. Entropy-energy com-pensation (23, 26), and cancellation of the

    neglected drop in solute entropy by the ne-glected rise in the solvation entropy, may also

    help.A second end-point approach, the

    MM-PBSA method, along with its GB vari-

    ant (MM-GBSA), uses MD simulations of thefree ligand, free protein, and their complex

    as a basis for calculating the average potentialand solvation energies in Equation 7 (46,

    114). The MD runs typically use an empiricalforce field and an explicit solvent model.

    The resulting snapshots i are postprocessedby stripping them of their explicit solvent

    molecules and computing their potentialenergies Ui with the empirical force field

    and their solvation energies Wi with either

    the PBSA or GBSA implicit solvent model.Averaging over each trajectory allows the

    changes in mean potential and solvationenergy to be calculated, as per Equation 7.

    The change in configurational entropy isthen estimated by energy-minimizing a few

    snapshots of the free and bound molecules,computing their entropy with the rigid-

    rotor/harmonic-oscillator approximation,and obtaining the average entropy change

    over these snapshots. This approach neglectsany contribution to the entropy change that

    may result from a change in the number

    of thermodynamically accessible energyminima. The quasiharmonic approximation

    (67) also has been used to estimate the change

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 31

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    12/24

    in configurational entropy from the MD sim-

    ulations (57, 79), despite concerns regardingits accuracy (21). It is not always clear from

    the literature whether or how the standardconcentration has been incorporated into

    MM-PBSA or MM-GBSA calculations, buta correct approach is available (equation 3

    in Reference 96); therefore the MM-PBSAmethod allows a clear connection with thetheory of binding.

    Although early applications of MM-PBSAappeared promising, it apparently is diffi-

    cult to converge the energy averages reli-ably. One reason for this difficulty may be

    that they encompass energy fluctuations notonly of the ligand and the binding site, but

    also of parts of the protein remote from thebinding site, which are less relevant to bind-

    ing. Convergence has been accelerated by asingle-trajectory approach (17, 79), in which

    only one MD simulation of the protein-ligand

    complex is carried out. Conformations of thenominally free ligand are then derived sim-

    ply by deleting the protein from the result-ing snapshots, and likewise for the free pro-

    tein. In this single-trajectory approach, onlyligand-protein distances contribute to the

    computed change in potential energy, becauseall other interatomic distances stay exactly the

    same. The single-trajectory approach resem-bles LIE because it similarly neglects changes

    in internal energy and configurational en-tropy of both ligand and protein. However,

    it differs from LIE in that it assumes the same

    conformational distribution for the free andbound ligand, whereas these change in LIE.

    Free energy pathway methods. The

    change in free energy between two states ofa system, such as before and after binding,

    can be written formally as RTln e U/RT,where U is the change in the energy

    function between the initial and final states,and the angle brackets indicate a Boltzmann

    average taken in the initial state (137). For

    binding, U would reflect the introductionof interactions between the protein and

    ligand, and the average could, in principle,

    be obtained by MD or MC simulations. In

    reality, such a simulation is extraordinarilydifficult to converge unless the initial and

    final states are similar. The free energyperturbation method solves this problem

    by breaking the change into N small stepsU (perturbations) and running a separate

    simulation for each resulting energy functionUi to obtain the stepwise free energy changesRTln

    e U/RT

    i

    associated with each step

    where i indicates a Boltzmann average withthe energy function Ui. Thus, the initial and

    final states are linked by a pathway of smallsteps for which the small free energy differ-

    ences can be computed. Another pathwayapproach, thermodynamic integration (TI)

    (73, 116), uses MD or MC to compute thefirst derivative of the free energy with respect

    to the distance along the path and thenestimates the total change in free energy via anumerical integral of the derivative along the

    path. Because the first derivative of the freeenergy is effectively a force, TI effectively

    involves a work integral. Either free energyperturbation or TI can be applied to a given

    free energy calculation, and both methodsare widely used.

    The first applications of such methods tobinding did not seek to compute the stan-

    dard free energy of binding, but rather tocompute the difference between the binding

    free energies of two similar ligands, Go

    by a method termed computational alchemy(119). This involves using pathways to com-

    pute the change in free energy when ligandA is changed to ligand B within the bind-

    ing site, as well as in solution. A free en-ergy cycle shows that the difference between

    these two free energy changes equals the dif-ference between the binding free energies

    of the two ligands. The calculations typi-cally employ lengthy MD or MC simulations

    (hours to many days, depending on the cal-culation) with an empirical force field and

    an explicit treatment of solvent. Relative to

    computing full-fledged standard free energiesof binding, in which a ligand must be re-

    moved from the binding site, computationa

    32 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    13/24

    alchemy has the merit of focusing on a rela-

    tively small perturbation, in which one ligandis changed to another similar one, and thus of

    converging relatively rapidly. Computationalalchemy also avoids the problem of account-

    ing for standard concentration, because it canbe shown from Equation 4 that Co cancels

    in Go

    . The method is theoretically soundand is supported by a literature reporting rela-tively accurate results, often within 1 to 2 kcal

    mol1 of experimental data. Still, computa-tional alchemy has not yet been subjected to

    extensive validation studies like those used toevaluate docking and scoring methods.

    The use of computational alchemy hasbeen limited in part by the amount of com-

    puter time it requires, especially when ligandA and ligand B are markedly different. How-

    ever, advances in computer technology andsimulation algorithms allow practical applica-tion of this method to drug discovery (4, 27,

    49, 104). Also, a clever extension of computa-tional alchemy by TI yields the relative affini-

    ties of multiple similar ligands from a singlesimulation run, at least for rigid ligands (92).

    At least two studies suggest that the methodis more accurate than MM-PBSA (46, 94), as

    noted above. Its convergence properties alsomay be better than that of the original imple-

    mentation of MM-PBSA because unlike MM-PBSA the work terms it computes are not di-

    rectly influenced by fluctuations far from the

    binding site.Pathway methods can also be used to com-

    putethestandardbindingfreeenergyofapro-tein and ligand. The Double Decoupling ap-

    proach (42), which was in effect first used tocompute the affinity of xenon for myoglobin

    (54), draws on related approaches (53, 106),notably the DoubleAnnihilationmethod (63).

    It involves using a pathway technique to com-pute the work of gradually decoupling the lig-

    and from the binding site and then effectivelycoupling it with an energy well or trap of de-

    fined size in bulk solution. The work of allow-ing the ligand to escape from the trap into a

    volume equal to 1/Co can be computed ana-

    lytically, so that the simulation results can be

    connectedwith the appropriatestandard state.

    This approach has been further developed andapplied to several protein-ligand systems (11,

    35). A computationally massive application ofthe similar but nonrigorous (42) Double An-

    nihilation method to the association of eightligands with the protein FKBP is also of con-

    siderable interest (40).It has been argued that Double De-coupling is difficult to converge for highly

    charged ligands because the final result mustaccurately balance strong interactions be-

    tween the ligand and the solvent against sim-ilarly strong interactions between the ligand

    and the protein (57). Alternative approachescompute the free energy of binding via a

    pathway in which the ligand is gradually ex-tracted from the binding site. Preliminary

    results from techniques of this type appearpromising (57, 79), though their applicability

    to ligands that bind in a closed cavity within

    the protein is not yet defined.Because the pathway methods are so

    promising and yet so difficult to run to con-vergence, there is great interest in accelerat-

    ing them or developing useful shortcuts, asrecently reviewed (105). Approaches include

    improved methods of extracting free energiesfrom the simulations (109, 132), novel sam-

    plingmethodstospeedconvergence(82,103),and hybrid implicit/explicit solvent models

    that reduce the computational load (5, 81).

    PROSPECTS AND CHALLENGES

    Accurate calculation of protein-ligand bind-

    ing affinities is an important and still unsolvedproblem. Nonetheless, significant progress is

    being made through a deepening understand-ing of the physical chemistry and the im-

    plementation of tractable algorithms basedon this understanding. Advances in com-

    puter hardware also continue to extend thelevel of detail that can be modeled. There

    is room for improvement in many aspects

    of existing methods, including models of thepotential energy and of solvent, treatment

    of protonation and tautomer equilibria, and

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 33

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    14/24

    Solvent entropy:entropy associatedwith motions of thesolvent molecules

    algorithms and heuristics for conformational

    sampling. Indeed, one should not expectmarked improvement in affinity calculations

    from progress against any single part of theproblem because theremainingparts still gen-

    erate large errors. Rather, success likely re-quires a concerted effort on all fronts. Un-

    derstanding the nature of the approximationsused is essential in this process, so that suc-cesses or failures will be interpretable and a

    path of improvement can be discerned. Oth-erwise, model development risks devolving to

    trial and error.Comparison with experimental data re-

    mains indispensable. Validation studies areeasier for computationally faster approaches,

    such as docking and scoring, but are alsoneeded for the more time-consuming meth-

    ods. Simple model systems can play animportant role in validation. For example, ar-

    tificial protein cavities that bind small, rigid

    ligands allow a focus on intermolecular inter-actions rather than on issues of ligand flexibil-

    ity (128, 129). Similarly, studies of host-guestmodel systems (23, 26) have provided insights

    into configurational entropy that bear on the

    development of improved scoring functionsExperimental data also are necessary as a basis

    for improving specific components of bind-ing models, and expansion of certain data sets

    would be welcome and valuable. For exam-ple, solvation models cannot be applied con-

    fidently to many candidate ligands becausethe existing database of solvation free ener-gies does not include many of the chemical

    moieties that frequently appear in drug-likecompounds.

    Finally, todays understanding of the phys-ical chemistry of molecular recognition may

    prove to be incomplete. Specific issues likelyto be of continuing interest include changes

    in ligand, protein, and solvent entropy uponbinding; the puzzling and apparently erro-

    neous tendency of many binding models toyield affinities that correlate strongly with the

    molecular weight of the ligand; electronic po-

    larization; the modeling of metal-containingbinding sites; and the thermodynamic impli-cations of water binding sites at the ligand-protein interface.

    SUMMARY POINTS

    1. Recent evaluations show that docking methods have considerable success in correctly

    identifying bound conformations and in enriching known ligands against a back-ground of decoy compounds, but yield less impressive results for ranking known

    ligand according to affinity.

    2. Important advances are being made in calculating the potential energy as a function

    of conformation, through more sophisticated force fields and the incorporation ofquantum mechanical treatments in ligand-protein modeling.

    3. Recent calculations suggest that changes in configurational entropy on binding arelarge, cancel much of the energy change that drives binding, and must be accounted

    for to obtain good correlations with measured affinities.

    4. Free energy pathway calculations with explicit solvent appear to yield good agreement

    with experiment, and it may be time to undertake more comprehensive validationstudies.

    5. Achieving accurate affinity calculations will likely require advances on multiple fronts:potential energy functions, solvation models, and the treatment of conformational

    flexibility.

    34 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    15/24

    DISCLOSURE STATEMENT

    MKG is a member of the company VeraChem LLC.

    ACKNOWLEDGMENTS

    The authors thank Dr. Hillary Gilson for valuable comments on the manuscript. This publica-tion was made possible by grant no. GM061300 to MKG and grant no. GM058187 to HXZ,

    from the National Institute of General Medical Sciences of the National Institutes of Health.Its contents are solely the responsibility of the authors and do not necessarily represent the

    official views of the National Institute of General Medical Sciences.

    LITERATURE CITED

    1. Alberts IL, Todorov NP, Dean PM. 2005. Receptor flexibility in de novo ligand design

    and docking. J. Med. Chem. 48:6585962. Alsallaq R, Zhou HX. 2007. Energy landscape and transition state of protein-protein

    association. Biophys. J. 92:14865023. Aqvist J, Medina C, Samuelsson JE. 1994. A new method for predicting binding affinity

    in computer-aided drug design. Protein Eng. 7:385914. Archontis G, Watson KA, Xie Q, Andreou G, Chrysina ED, et al. 2005. Glycogen phos-

    phorylase inhibitors: a free energy perturbation analysis of glucopyranose spirohydantoinanalogues. Protein Struct. Funct. Bioinform. 61:98498

    5. Banavali NK, Im W, Roux B. 2002. Electrostatic free energy calculations using the gen-

    eralized solvent boundary potential method. J. Chem. Phys. 117:7381886. Berendsen HJC, Grigera JR, Straatsma TP. 1987. The missing term in effective pair

    potentials. J. Phys. Chem. 91:6269717. Betzi S, Suhre K, ChetritB, Guerlesquin F, Morelli X. 2006. GFscore: a general nonlinear

    consensus scoring function for high-throughput docking.J. Chem. Inf. Model. 46:1704128. Bixon M, Lifson S. 1967. Potential functions and conformations in cycloalkanes. Tetra-

    hedron 23:76984

    9. Bohm HJ. 1994. The development of a simple empirical scoring function to estimate thebinding constant for a protein-ligand complex of known three-dimensional structure. J.

    Comput. Aided Mol. Des. 8:2435610. Bordner AJ, Cavasotto CN, Abagyan RA. 2002. Accurate transferable model for water,

    n-octanol, and n-hexadecane solvation free energies. J. Phys. Chem. B 106:110091511. Boresch S, Tettinger F, Leitgeb M, Karplus M. 2003. Absolute binding free energies: a

    quantitative approach for their calculation. J. Phys. Chem. B 107:95355112. Born M. 1920. Volumes and heats of hydration of ions. Z. Phys. 1:454813. Bostrom J. 2001. Reproducing the conformations of protein-bound ligands: a critical

    evaluation of several popular conformational searching tools. J. Comput. Aided Mol. Des.

    15:113752

    14. Brooijmans N, Kuntz ID. 2003. Molecular recognition and docking algorithms. Annu.Rev. Biophys. Biomol. Struct. 32:3357315. Brooks BR, Bruccoleri RE, Olafson BD, States DJ, Swaminathan S, Karplus M. 1983.

    CHARMM: a program for macromolecular energy, minimization and dynamics calcula-

    tions. J. Comput. Chem. 4:18721716. Broughton HB. 2000. A method for including protein flexibility in protein-ligand dock-

    ing: improving tools for database mining and virtual screening. J. Mol. Graph. Model.

    18:24757

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 35

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    16/24

    17. Brown SP, Muchmore SW. 2006. High-throughput calculation of protein-ligand bind-

    ing affinities: modification and adaptation of the MM-PBSA protocol to enterprise gridcomputing. J. Chem. Inf. Model. 46:9991005

    18. Carlson HA, McCammon JA. 2000. Accommodating protein flexibility in computationadrug design. Mol. Pharmacol. 57:21318

    19. Carlsson J, Ander M, Nervall M, Aqvist J. 2006. Continuum solvation models in the

    linear interaction energy method. J. Phys. Chem. B 110:1203441

    20. Cavasotto CN, Kovacs JA, Abagyan RA. 2005. Representing receptor flexibility in liganddocking through relevant normal modes. J. Am. Chem. Soc. 127:963240

    21. Chang C, Chang W, Gilson MK. 2005. Evaluating the accuracy of the quasiharmonic

    approximation. J. Chem. Theory Comput. 1:101728

    22. Chang C, Chen W, Gilson MK. 2007. Ligand configurational entropy and protein bind-ing. Proc. Natl. Acad. Sci. USA. In press

    23. Chang CE, Gilson MK. 2004. Free energy, entropy, and induced fit in host-guest recog-nition: calculations with the second-generation mining minima algorithm. J. Am. ChemSoc. 126:1315664

    24. Charifson PS, Corkery JJ, Murcko MA, Walters WP. 1999. Consensus scoring: a method

    for obtaining improved hit rates from docking databases of three-dimensional structures

    into proteins. J. Med. Chem. 42:5100925. Chen H, Lyne PD, Giordanetto F, Lovell T, Li J. 2005. On evaluating molecular-docking

    methods for pose prediction and enrichment factors. J. Chem. Inf. Model. 46:40115

    26. Computationalstudy of host-guestsystems, withinsights regardingconfigurationalentropy that areapplicable toprotein-ligandbinding.

    26. Chen W, Chang C, Gilson MK. 2004. Calculation of cyclodextrin binding affinities

    energy, entropy, and implications for drug design. Biophys. J. 87:303549

    27. Chipot C, Rozanska X, Dixit SB. 2005. Can free energy calculations be fast and accurateat the same time? Binding of low-affinity, nonpeptide inhibitors to the SH2 domain of

    the Src protein. J. Comput. Aided Mol. Des. 19:76570

    28. Evidence thatredistribution ofligand charge is animportantdeterminant ofconformation.

    28. Cho AE, Guallar V, Berne BJ, Friesner R. 2005. Importance of accurate charges

    in molecular docking: quantum mechanical/molecular mechanical (QM/MM) ap-

    proach. J. Comput. Chem. 26:91531

    29. Cole JC, Murray CW, Nissink JWM, Taylor RD, Taylor R. 2005. Comparing protein-ligand docking programs is difficult. Protein Struct. Funct. Bioinform. 60:32532

    30. Cornell WD, Cieplak P, Bayly CI, Gould IR, Merz KM, et al. 1995. A second generation

    force-field for the simulation of proteins, nucleic-acids, and organic-molecules. J. Am

    Chem. Soc. 117:517997

    31. Curioni A, Mordasini T, Andreoni W. 2004. Enhancing the accuracy of virtual screeningmolecular dynamics with quantum-refined force fields.J. Comput. Aided Mol. Des. 18:773

    84

    32. DesJarlais RL, Sheridan RP, Dixon JS, Kuntz ID, Venkataraghavan R. 1986. Dockingflexible ligands to macromolecular receptors by molecular shape. J. Med. Chem. 29:2149

    53

    33. Dixit SB, Chipot C. 2001. Can absolute free energies of association be estimated from

    molecular mechanical simulations? The biotin-streptavidin system revisited. J. Phys

    Chem. A 105:979599

    34. Dominy BN, Brooks CL III. 1999. Development of a generalized Born modeparametrization for proteins and nucleic acids. J. Phys. Chem. B 103:376573

    35. Donnini S, Juffer AH. 2004. Calculation of affinities of peptides for proteins. J. Comput

    Chem. 25:393411

    36 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    17/24

    36. Eldridge MD, Murray CW, Auton TR, Paolini GV, Mee RP. 1997. Empirical scoring

    functions: I. The development of a fast empirical scoring function to estimate the bindingaffinity of ligands in receptor complexes. J. Comput. Aided Mol. Des. 11:42545

    37. Ferrara P, Curioni A, Vangrevelinghe E, Meyer T, Mordasini T, et al. 2006. New scoringfunctions for virtual screening from molecular dynamics simulations with a quantum-

    refined force-field (QRFF-MD). Application to cyclin-dependent kinase 2. J. Chem. Inf.

    Model. 46:25463

    38. Ferrara P, Gohlke H, Price DJ, Klebe G, Brooks CL III.2004. Assessing scoring functionsfor protein-ligand interactions. J. Med. Chem. 47:30324739. Ferrari AM, Wei BQ, Costantino L, Shoichet BK. 2004. Soft docking and multiple

    receptor conformations in virtual screening. J. Med. Chem. 47:50768440. Fujitani H, Tanida Y, Ito M, Jayachandran G, Snow CD, et al. 2005. Direct calculation

    of the binding free energies of FKBP ligands. J. Chem. Phys. 123:08410841. Garcia-Sosa AT, Mancera RL, Dean PM. 2003. WaterScore: a novel method for distin-

    guishing between bound and displaceable water molecules in the crystal. J. Mol. Model.9:17282

    42. Gilson MK, Given JA, Bush BL, McCammon JA. 1997. The statistical-thermodynamicbasis for computation of binding affinities: a critical review. Biophys. J. 72:104769

    43. Gilson MK,Honig B. 1991. Theinclusion of electrostatic hydration energies in molecularmechanics calculations. J. Comput. Aided Mol. Des. 5:520

    44. Giordanetto F, Cotesta S, Catana C, Trosset JY, Vulpetti A, et al. 2004. Novel scoring

    functions comprising QXP, SASA, and protein side-chain entropy terms. J. Chem. Inf.Comput. Sci. 44:88293

    45. Goodsell DS, Olson AJ. 1990. Automated docking of substrates to proteins by simulatedannealing. Protein Struct. Funct. Genet. 8:195202

    46. Gouda H, Kuntz ID, Case DA, Kollman PA. 2003. Free energy calculations for theo-phylline binding to an RNA aptamer: comparison of MM-PBSA and thermodynamic

    integration methods. Biopolymers68:163447. Grater F, Schwarzl SM, Dejaegere A, Fischer S, Smith JC. 2005. Protein/ligand binding

    free energies calculated with quantum mechanics/molecular mechanics. J. Phys. Chem. B109:1047483

    48. Gresh N. 2006. Development, validation, and applications of anisotropic polarizablemolecular mechanics to study ligand and drug-receptor interactions. Curr. Pharmacol.Des. 12:212158

    49. Guimaraes CRW, Boger DL, Jorgensen WL. 2005. Elucidation of fatty acid amide hy-drolase inhibition by potent alpha-ketoheterocycle derivatives from Monte Carlo simu-

    lations. J. Am. Chem. Soc. 127:173778450. Hamelberg D, McCammon JA. 2004. Standard free energy of releasing a localized water

    molecule from the binding pockets of proteins: double-decoupling method.J. Am. Chem.

    Soc. 126:768389

    51. Hawkins GD, Cramer CJ, Truhlar DG. 1995. Pairwise solute descreening of solutecharges from a dielectric medium. Chem. Phys. Lett. 246:12229

    52. Head MS, Given JA, Gilson MK. 1997. Mining minima: direct computation of con-formational free energy. J. Phys. Chem. 101:160918

    53. Helms V, Wade R. 1995. Thermodynamics of water mediating protein-ligand interac-tions in cytochrome P450cam: a molecular dynamics study. Biophys. J. 69:81024

    54. Hermans J, Shankar S. 1986. The free energy of xenon binding to myoglobin from

    molecular dynamics simulation. Isr. J. Chem. 27:22527

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 37

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    18/24

    55. Honig B, Sharp K, Yang AS. 1993. Macroscopic models of aqueous solutions: biologica

    and chemical applications. J. Phys. Chem. 97:11019

    56. Use of aphysics-basedscoring function inlarge-scale virtualscreening.

    56. Huang N, Kalyanaraman C, Irwin JJ, Jacobson MP. 2006. Physics-based scoring

    of protein-ligand complexes: enrichment in large-scale virtual screening. J. Chem

    Inf. Model. 46:24353

    57. Calculation ofbinding affinity

    with a pathwaymethod thatinvolves pulling theligand from thebinding site.

    57. Woo HJ, Roux B. 2005. Calculation of absolute protein-ligand binding free energy

    from computer simulations. Proc. Natl. Acad. Sci. USA 102:682530

    58. Im W, Lee MS, Brooks CL III. 2003. Generalized Born model with a simple smoothingfunction. J. Comput. Chem. 24:1691702

    59. Jayaram B, Liu Y, Beveridge DL. 1998. A modification of the generalized Born theory

    for improved estimates of solvation energies and pK shifts. J. Chem. Phys. 109:146571

    60. Jiang F, Kim SH. 1991. Soft docking: matching of molecular surface cubes.J. Mol. Biol

    219:79102

    61. Jones-Hertzog DK, Jorgensen WL. 1997. Binding affinities for sulfonamide inhibitors

    with human thrombin using Monte Carlo simulations with a linear response method. JMed. Chem. 40:153949

    62. Jorgensen WL. 2004. Themany roles of computation in drug discovery. Science 303:1813

    18

    63. Jorgensen WL, Buckner JK, Boudon S, Tirado-Rives J. 1988. Efficient computation ofabsolute free energies of binding by computer simulations. Application to methane dimerin water. J. Chem. Phys. 89:374246

    64. Jorgensen WL, Jenson C. 1998. Temperature dependence of TIP3P, SPC, and TIP4Pwater from NPT Monte Carlo simulations: seeking temperatures of maximum density

    J. Comput. Chem. 19:117986

    65. Kairys V, Gilson MK. 2002. Enhanced docking with the mining minima optimizer: ac-

    celeration and side-chain flexibility. J. Comput. Chem. 23:165670

    66. Kaminski GA, Stern HA, Berne BJ, Friesner RA. 2004. Development of an accurate and

    robust polarizable molecular mechanics force field from ab initio quantum chemistry. J

    Phys. Chem. A 108:62127

    67. Karplus M, Kushick J. 1981. Methodfor estimating the configurational entropy of macro-molecules. Macromolecules14:32532

    68. Kauzmann W. 1959. Some factors in the interpretation of protein denaturation. AdvProtein Chem. 14:163

    69. Kellenberger E, Rodrigo J, Muller P, Rognan D. 2004. Comparative evaluation of eight

    docking tools for docking and virtual screening accuracy.Proteins Struct. Funct. Bioinform

    57:22542

    70. Keseru GM, Kolossvary I. 2001. Fully flexible low-mode docking: application to inducedfit in HIV integrase. J. Am. Chem. Soc. 123:127089

    71. Khandelwal A, Lukacova V, Comez D, Kroll DM, Raha S, Balaz S. 2005. A combinationof docking, QM/MM methods, and MD simulation for binding affinity estimation of

    metalloprotein ligands. J. Med. Chem. 48:543747

    72. Kirchmair J, Wolber G, Laggner C, Langer T. 2006. Comparative performance assess-

    ment of the conformational model generators Omega and Catalyst: a large-scale surveyon the retrieval of protein-bound ligand conformations. J. Chem. Inf. Model. 46:184861

    73. Kirkwood JG. 1935. Statistical mechanics of fluid mixtures. J. Chem. Phys. 3:30013

    74. Knegtel RMA, Kuntz ID, Oshiro CM. 1997. Molecular docking to ensembles of protein

    structures. J. Mol. Biol. 266:42440

    38 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    19/24

    75. Knox AJS, Meegan MJ, Carta G, Lloyd DG. 2005. Considerations in compound database

    preparationhidden impact on virtual screening results. J. Chem. Inf. Model. 45:190819

    76. Kuntz ID, Blaney JM, Oatley SJ, Langridge R, Ferrin TE. 1982. A geometric approach

    to macromolecule-ligand interactions. J. Mol. Biol. 161:2698877. Lamoureux G, Harder E, Vorobyov IV, Roux B, MacKerell AD. 2006. A polarizable

    model of water for molecular dynamics simulations of biomolecules. Chem. Phys. Lett.

    418:2454978. Lazaridis T, Masunov A, Gandolfo F. 2002. Contributions to the binding free energy of

    ligands to avidin and streptavidin. Protein Struct. Funct. Genet. 47:19420879. Lee MS, Olson MA. 2006. Calculation of absolute protein-ligand binding affinity using

    path and endpoint approaches. Biophys. J. 90:8647780. Lee MS, Salsbury FR Jr, Brooks CL III. 2002. Novel generalized Born methods.J. Chem.

    Phys. 116:1060614

    81. Theoretical andpractical analysis of

    MM-PBSA andfree energypathway methods.

    81. Lee MS, Salsbury FR, Olson MA. 2004. An efficient hybrid explicit/implicit solvent

    method for biomolecular simulations. J. Comput. Chem. 25:19677882. LeitgebM, Schroder C, Boresch S. 2005. Alchemical free energycalculationsand multiple

    conformational substates. J. Chem. Phys. 122:08410983. Li Z, Lazaridis T. 2005. The effect of water displacement on binding thermodynamics:

    concanavalin A. J. Phys. Chem. B 109:6627084. Liu HY, Kuntz ID, Zou X. 2004. Pairwise GB/SA scoring function for structure-based

    drug design. J. Phys. Chem. B 108:54536285. Luo H, Sharp K. 2002. On the calculation of absolute macromolecular binding free

    energies. Proc. Natl. Acad. Sci. USA 99:1039940486. Majeux N, Scarsi M, Apostolakis J, Ehrhardt C, Caflisch A. 1999. Exhaustive docking of

    molecular fragments with electrostatic solvation. Protein Struct. Funct. Genet. 37:8810587. Maple JR, Cao Y, Damm W, Halgren TA, Kaminski GA, et al. 2005. A polarizable force

    field and continuum solvation methodology for modeling of protein-ligand interactions.

    J. Chem. Theor. Comput. 1:69471588. Mihailescu M, Gilson MK. 2004. On the theory of noncovalent binding. Biophys. J.

    87:232689. Morris GM, Goodsell DS, Halliday RS, Huey R, Hart WE, et al. 1998. Automated

    docking using a Lamarckian genetic algorithm and an empirical binding free energy

    function. J. Comput. Chem. 19:16396290. Oda A, Tsuchida K, Takakura T, Yamaotsu N, Hirono S. 2006. Comparison of consensus

    scoring strategies for evaluating computational models of protein-ligand complexes. J.Chem. Inf. Model. 46:38091

    91. Olano LR, Rick SW. 2004. Hydration free energies and entropies for water in protein

    interiors. J. Am. Chem. Soc. 126:79918000

    92. A method ofpredicting theaffinities ofmultiple similarligands based upon

    a single simulation.

    92. Oostenbrink C, van Gunsteren WF. 2005. Free energies of ligand binding for

    structurally diverse compounds. Proc. Natl. Acad. Sci. USA 102:675054

    93. Osterberg F, Morris GM, Sanner MF, Olson AJ, Goodsell DS. 2002. Automated dock-ing to multiple target structures: incorporation of protein mobility and structural waterheterogeneity into AutoDock. Protein Struct. Funct. Genet. 46:3440

    94. Criticalcomparison of

    MM-PBSA andfree energypathway methods.

    94. Pearlman DA. 2005. Evaluating the molecular mechanics Poisson-Boltzmann sur-

    face area free energy method using a congeneric series of ligands to p38 MAP

    kinase. J. Med. Chem. 48:779680795. Peters MB, Raha K, Merz KM Jr. 2006. Quantum mechanics in structure-based drug

    design. Curr. Opin. Drug. Discov. 9:37079

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 39

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    20/24

    96. Potter MJ, Gilson MK. 2002. Coordinate systems and the calculation of molecular prop-

    erties. J. Phys. Chem. A 126:5636697. Qiu D, Shenkin PS, Hollinger FP, Still WC. 1997. The GB/SA continuum model for

    solvation: a fast analytical method for the calculation of approximate Born radii. J. Phys

    Chem. 101:30051498. Raha K, Merz KM Jr. 2004. A quantum mechanics-based scoring function: study of zinc

    ion-mediated ligand binding. J. Am. Chem. Soc. 126:10202199. Raha K, Merz KM Jr. 2005. Large-scale validation of a quantum mechanics based scoring

    function: predicting the binding affinity and the binding mode of a diverse set of protein-

    ligand complexes. J. Med. Chem. 48:455875100. Rarey M, Kramer B, Lengauer T. 1999. The particle concept: placing discrete water

    molecules during protein-ligand docking predictions. Protein Struct. Funct. Genet. 34:1728

    101. Rarey M, Kramer B, Lengauer T, Klebe G. 1996. A fast flexible docking method usingan incremental construction algorithm. J. Mol. Biol. 261:47089

    102. Ren PY, Ponder JW. 2004. Temperature andpressure dependence of the AMOEBAwatermodel. J. Phys. Chem. B 108:1342737

    103. Rick SW. 2006. Increasing the efficiency of free energy calculations using parallel tem-pering and histogram reweighting. J. Chem. Theor. Comput. 2:93946

    104. Rizzo R, Wang DP, Tirado-Rives J, Jorgensen WL. 2000. Validation of a model for thecomplex of HIV-1 reverse transcriptase with sustiva through computation of resistance

    profiles. J. Am. Chem. Soc. 122:12898900105. Rodinger T, Pomes R. 2005. Enhancing the accuracy, the efficiency and the scope of free

    energy simulations. Curr. Opin. Struct. Biol. 15:16470106. Roux B, Nina M, Pomes R, Smith JC. 1996. Thermodynamic stability of water molecules

    in the bacteriorhodopsin proton channel: a molecular dynamics free energy perturbation

    study. Biophys. J. 71:67081107. Ruvinsky AM, Kozintsev AV. 2005. New and fast statistical-thermodynamic method for

    computation of protein-ligand binding entropy substantially improves docking accuracyJ. Comput. Chem. 26:108995

    108. Sherman W, Day T, Jacobson MP, Friesner RA, Farid R. 2006. Novel procedure formodeling ligand/receptor induced fit effects. J. Med. Chem. 49:53453

    109. Shirts MR, Pande VS. 2005. Comparison of efficiency and bias of free energies computedby exponential averaging, the Bennett acceptance ratio, and thermodynamic integration.

    J. Chem. Phys. 122:144107110. Shoichet BK. 2004. Virtual screening of chemical libraries. Nature 432:86265111. Shoichet BK, Leach AR, Kuntz ID. 1999. Ligand solvation in molecular docking. Protein

    Struct. Funct. Genet. 34:416112. Sitkoff D, Sharp KA, Honig B. 1994. Accurate calculation of hydration free energies

    using macroscopic solvation models. J. Phys. Chem. 98:197888113. Sousa SF, Fernandes PA, Ramos MJ. 2006. Protein-ligand docking: current status and

    future challenges. Protein Struct. Funct. Bioinform. 65:1526114. Srinivasan J, Cheatham TE III, Cieplak P, Kollman PA, Case DA. 1998. Continuumsolvent studies of the stability of DNA, RNA, and phosphoramidate-DNA helices. J

    Am. Chem. Soc. 120:94019115. Stone AJ, Price SL. 1988. Some new ideas in the theory of intermolecular forces -

    anisotropic atom potentials. J. Phys. Chem. 92:332535116. Straatsma TP, McCammon JA. 1991. Multiconfiguration thermodynamic integration. J

    Chem. Phys. 95:117588

    40 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    21/24

    117. Theoreticalanalysis of

    MM-PBSA, andcalculation ofcontributions fromrotational andtranslationaldegrees of

    freedom.

    117. Swanson JMJ, Henchman RH, McCammon JA. 2004. Revisiting free energy cal-

    culations: a theoretical connection to MM/PBSA and direct calculation of the as-

    sociation free energy. Biophys. J. 86:6774

    118. Tashiro M, Stuchebrukhov AA. 2005. Thermodynamic properties of internal water

    molecules in the hydrophobic cavity around the catalytic center of cytochrome c oxi-dase. J. Phys. Chem. B 109:101522

    119. Tembe BL, McCammon JA. 1984. Ligand-receptor interactions. Comput. Chem. 8:28183

    120. Todorov NP, Monthoux PH, Alberts IL. 2006. The influence of variations of ligand pro-tonation and tautomerism on protein-ligand recognition and binding energy landscape.

    J. Chem. Inf. Model. 46:113442

    121. van Gunsteren WF, Berendsen HJC. 1987. Groningen Molecular Simulation (GROMOS)Library Manual. Groningen, The Nether.: BIOMOS

    122. Vasilyev V, Bliznyuk A. 2004. Application of semiempirical quantum chemical methods

    as a scoring function in docking. Theor. Chem. Acc. 112:31317

    123. Verdonk ML, Chessari G, Cole JC, Hartshorn MJ, Murray CW, et al. 2005. Modelingwater molecules in protein-ligand docking using GOLD. J. Med. Chem. 48:650415

    124. Wang R, Lai L, Wang S. 2002. Further development and validation of empirical scoringfunctions for structure-based binding affinity prediction.J. Comput. Aided Mol. Des. 16:11

    26125. Wang R, Lu Y, Fang X, Wang S. 2004. An extensive test of 14 scoring functions using

    the PDBbind refined set of 800 protein-ligand complexes. J. Chem. Inf. Comput. Sci.

    44:211424

    126. Wang W, Wang J, Kollman PA. 1999. What determines the van der Waals coefficientbeta in the LIE (linear interaction energy) method to estimate binding free energies using

    molecular dynamics simulations? Protein Struct. Funct. Genet. 34:395402

    127. Broad anddetailed evaluationof docking andscoring methods.

    127. Warren GL, Andrews CW, Capelli A, Clarke B, LaLonde J, et al. 2005. A critical

    assessment of docking programs and scoring functions. J. Med. Chem. 49:591231

    128. Wei BQ, Baase WA, Weaver LH, Matthews BW, Shoichet BK. 2002. A model binding

    site for testing scoring functions in molecular docking. J. Mol. Biol. 322:33955

    129. Wei BQ, Weaver LH, Ferrari AM, Matthews BW, Shoichet BK. 2004. Testing a flexible-receptor docking algorithm in a model binding site. J. Mol. Biol. 337:116182

    130. Wong CF, Kua J, Zhang Y, Straatsma TP, McCammon JA. 2005. Molecular dockingof balanol to dynamics snapshots of protein kinase A. Protein Struct. Funct. Bioinform.

    61:85058

    131. Yang AY, Kallblad P, Mancera RL. 2004. Molecular modelling prediction of ligand bind-ing site flexibility. J. Comput. Aided Mol. Des. 18:23550

    132. Yang W, Bitetti-Putzer R, Karplus M. 2004. Free energy simulations: use of reversecumulative averaging to determine the equilibrated region and the time required for

    convergence. J. Chem. Phys. 120:261828

    133. Zacharias M. 2004. Rapid protein-ligand docking using soft modes from molecular dy-namics simulations to account for protein deformability: binding of FK506 to FKBP.

    Protein Struct. Funct. Bioinform. 54:75967

    134. Zavodszky M, Kuhn L. 2005. Side-chain flexibility in protein-ligand binding: the minimalrotation hypothesis. Protein Sci. 14:110414

    135. Zhou R, Friesner RA, Ghosh A, Rizzo RC, Jorgensen WL, Levy RM. 2001. New linearinteraction method for binding affinity calculations using a continuum solvent model. J.

    Phys. Chem. B 105:1038897

    www.annualreviews.org Calculation of Protein-Ligand Binding Affinities 41

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    22/24

    136. Zou X, Sun YX, Kuntz ID. 1999. Inclusion of solvation in ligand binding free energy

    calculations using the generalized-Born model. J. Am. Chem. Soc. 121:803343137. Zwanzig RW. 1954. High-temperature equation of state by a perturbation method. I

    Nonpolar gases. J. Chem. Phys. 22:142026138. Lu B, Wong CF. 2005. Direct estimation of entropy loss due to reduced translational and

    rotational motions upon molecular binding. Biopolymers79:27785

    RELATED REVIEWS

    Brooijmans N, Kuntz ID. 2003. Molecular recognition and docking algorithms. Annu. RevBiophys. Biomol. Struct. 32:33573

    Peters MB, Raha K, Merz KM Jr. 2006. Quantum mechanics in structure-based drug design.Curr. Opin. Drug. Discov. 9:37079

    Rodinger T, Pomes R. 2005. Enhancing the accuracy, the efficiency andthe scope of free energysimulations. Curr. Opin. Struct. Biol. 15:16470

    42 Gilson Zhou

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    23/24

    Annual Review

    Biophysics and

    Biomolecular

    Structure

    Volume 36, 2007Contents

    Synthetic Gene Circuits: Design with Directed Evolution

    Eric L. Haseltine and Frances H. Arnold p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 1

    Calculation of Protein-Ligand Binding Affinities

    Michael K. Gilson and Huan-Xiang Zhou p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 21

    Visualizing Flexibility at Molecular Resolution: Analysis of Heterogeneity

    in Single-Particle Electron Microscopy ReconstructionsAndres E. Leschziner and Eva Nogales p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 43

    Phase Boundaries and Biological Membranes

    Gerald W. Feigenson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 63

    From Simple DNA-Protein Interactions to the Macromolecular

    Machines of Gene Expression

    Peter H. von Hippel p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 79

    Bilayer Thickness and Membrane Protein Function: An Energetic Perspective

    Olaf S. Andersen and Roger E. Koeppe, II p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 107

    Structural Mechanisms Underlying Posttranslational Modification by

    Ubiquitin-Like Proteins

    Billy T. Dye and Brenda A. Schulman p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 131

    Fluorescence Correlation Spectroscopy: Novel Variations of an

    Established Technique

    Elke Haustein and Petra Schwille p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 151

    High-Resolution, Single-Molecule Measurements of Biomolecular Motion

    William J. Greenleaf, Michael T. Woodside, and Steven M. Block p p p p p p p p p p p p p p p p p p p p 171

    Gene Regulation: Gene Control Network in DevelopmentSmadar Ben-Tabou de-Leon and Eric H. Davidson p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 191

    Microfluidic Large-Scale Integration: The Evolution of Design Rules

    for Biological Automation

    Jessica Melin and Stephen R. Quake p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p 213

    Deciphering Molecular Interactions of Native Membrane Proteins by

    Single-Molecule Force Spectroscopy

    Alexej Kedrov, Harald Janovjak, K. Tanuj Sapra, and Daniel J. M uller p p p p p p p p p p p p 233

    v

  • 7/30/2019 Calculations of Protein-Ligand Binding Affinities

    24/24

    Physics of Proteins

    Jayanth R. Banavar and Amos Maritan p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p

    Insights from Crystallographic Studies into the Structural and Pairing

    Properties of Nucleic Acid Analogs and Chemically Modified DNA

    and RNA Oligonucleotides

    Martin Egli and Pradeep S. Pallanp p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p

    Small-Angle X-Ray Scattering from RNA, Proteins, and Protein

    Complexes

    Jan Lipfert and Sebastian Doniach p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p

    Predictive Modeling of Genome-Wide mRNA Expression:

    From Modules to Molecules

    Harmen J. Bussemaker, Barrett C. Foat, and Lucas D. Ward p p p p p p p p p p p p p p p p p p p p p p p p

    New Fluorescent Tools for Watching Nanometer-Scale

    Conformational Changes of Single Molecules

    Erdal Toprak and Paul R. Selvinp p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p

    Single-Molecule Fluorescence Analysis of Cellular Nanomachinery

    Components

    Reiner Peters p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p

    Conformational Dynamics and Ensembles in Protein Folding

    Victor Munoz p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p

    Living with Noisy Genes: How Cells Function Reliably with Inherent

    Variability in Gene Expression

    Narendra Maheshri and Erin K. OShea p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p p