Вы находитесь на странице: 1из 17

proteins

STRUCTURE O FUNCTION O BIOINFORMATICS

On the diversity of physicochemical


environments experienced by identical ligands
in binding pockets of unrelated proteins
Abdullah Kahraman,1,2* Richard J. Morris,3 Roman A. Laskowski,1 Angelo D. Favia,4
and Janet M. Thornton1
1 European Bioinformatics Institute, Wellcome Trust Genome Campus, Hinxton, Cambridgeshire CB10 1SD, United Kingdom
2 Swiss Federal Institute of Technology Zurich, Institute of Molecular Systems Biology, Wolfgang-Pauli Strasse 16,

8106 Zurich, Switzerland


3 John Innes Centre, Computational and System Biology, Norwich Research Park, Norwich NR7 7UH, United Kingdom
4 Istituto Italiano di Tecnologia (IIT), Drug Discovery and Development, Via Morego 30, 16163 Genoa, Italy

ABSTRACT INTRODUCTION

Most function prediction methods that identify cognate It is generally assumed that the molecular recognition
ligands from binding site analyses work on the assumption of between protein receptors and ligand molecules requires
molecular complementarity. These approaches build on the in addition to geometrical complementarity, a high
conjectured complementarity of geometrical and physico- degree of physicochemical complementarity between both
chemical properties between ligands and binding sites so that binding partners.1 Shape complementarity in particular
similar binding sites will bind similar ligands. We found that plays a crucial role in the recognition process as it per-
this assumption does not generally hold for protein–ligand mits both binding partners to approach each other and
interactions and observed that it is not the chemical composi- form short-range noncovalent bonds.2,3 However, several
tion of ligand molecules that dictates the complementarity
studies have reported that the geometrical complementar-
between protein and ligand molecules, but that the ligand’s
share within the functional mechanism of a protein deter-
ity is often far from perfect.4–6 These studies revealed
mines the degree of complementarity. Here, we present for a that ligands sometimes are only partially enclosed by
set of cognate ligands a descriptive analysis and comparison of their binding sites with the rest of the ligand exposed to
the physicochemical properties that each ligand experiences in the solvent. Furthermore, even within the enclosed
various nonhomologous binding pockets. The comparisons in region, contacts between protein and ligand often occur
each ligand set reveal large variations in their experienced at relatively few points, involving strong hydrogen bonds
physicochemical properties, suggesting that the same ligand or van der Waals interactions7,8 leaving unoccupied
can bind to distinct physicochemical environments. In some space like a ‘‘buffer zone’’ around the remainder of the
protein ligand complexes, the variation was found to correlate ligand. The buffer zone is partially filled by water mole-
with the electrochemical characteristic of ligand molecules, cules enhancing the complementarity of the ‘‘naked’’
whereas in others it was disclosed as a prerequisite for the bio-
binding pocket. It has also been suggested that the buffer
chemical function of the protein. To achieve binding, proteins
were observed to engage in subtle balancing acts between elec- zones provide space for motions of protein, ligand, or
trostatic and hydrophobic interactions to generate stabilizing water molecules, thus avoiding the complete loss of the
free energies of binding. For the presented analysis, a new ligand’s and binding site residue’s entropy.9
method for scoring hydrophobicity from molecular environ- Physicochemical complementarity also contributes to
ments was developed showing high correlations with experi- molecular recognition, in particular, the complementarity
mental determined desolvation energies. The presented results
highlight the complexities of molecular recognition and Additional Supporting Information may be found in the online version of this article.
underline the challenges of computational structural biology This work was performed at the European Bioinformatics Institute, Hinxton, UK.
in developing methods to detect these important subtleties. Grant sponsor: BioSapiens Network of Excellence, through the European Commis-
sion within its FP6 Programme, under the thematic area ‘‘Life Sciences, Genomics
Proteins 2010; 00:000–000. and Biotechnology for Health’’; Grant number: LHSG-CT-2003-503265; Grant
V
C 2009 Wiley-Liss, Inc. sponsor: EMBL (Predoctoral Fellowship).
*Correspondence to: Abdullah Kahraman, ETHZ, IMSB, HPT C75, Wolfgang-Pauli
Strasse 16, 8093 Zurich, Switzerland. E-mail: abdullah@imsb.biol.ethz.ch
Key words: molecular recognition; protein–ligand interac- Received 29 September 2008; Revised 21 September 2009; Accepted 22 September
tion; electrostatics; hydrophobicity; noncomplementarity; 2009
Published online in Wiley InterScience (www.interscience.wiley.com).
cognate ligand; protein function; redox potential. DOI: 10.1002/prot.22633

V
C 2009 WILEY-LISS, INC. PROTEINS 1
A. Kahraman et al.

derived from long-range electrostatic interactions is AMP, ATP, FAD, FMN, glucose, heme, NAD, phosphate,
believed to be the driving force for molecular interac- or steroids (estradiol and dehydroepiandrosterone).6
tions. For example, proteins binding an adenine or gua- Countless applications have been developed in the past
nine were found to discriminate between both molecular for analysing active and binding sites, mainly in the field
moieties on the basis of electrostatics.10 Copper zinc of rational drug design. Most of the programs such as
superoxide dismutases, a protein family known to have LigBuilder,25 MCSS,26 PocketFinder,5 and Q-Site-
the highest reaction rate among enzymes, attract posi- Finder27 are derivatives of the program GRID.28 GRID,
tively charged metal ions into their binding sites through developed by Goodford, analyses the physicochemical
a highly negative electrostatic field,11 and DNA-binding properties of protein-binding sites by computing the
proteins attract negatively charged DNA molecules interaction energy between various spherical probes and
through positively charged patches on their molecular the atoms of the protein using a standard potential
surface.12 Different studies have been performed to energy function. Other programs provide different infor-
quantify the electrostatic complementarity between bind- mation about a given binding site. For example,
ing partners. In particular, early works by Nakamura GRASP,29 although primarily a powerful protein struc-
et al.,13–15 Chau and Dean,16–18 and Naray-Szabo and ture visualization software, has a built-in Poisson-Boltz-
Nagy19,20 analyzed the electrostatic lock-and-key model mann Solver, which allows the calculation and visualiza-
of single ligand- and protein-binding sites by mapping tion of the ESP in and around the protein-binding site.
the molecular electrostatic potential (MEP) of both bind- The program HINT30 calculates the hydrophobicity of a
ing partners on a set of reference points on their surfa- molecule using experimental octanol/water partition
ces. The MEPs on the reference points were eventually coefficients and constructs a hydropathy field or comple-
compared following the host-on-guest and guest-on- mentarity map for a protein binding site. The CASTp31
guest model. These studies came to the conclusion that database is a repository of protein clefts and voids and
long-range effects of partial charges are essential for cre- provides area and volume measurements for each cavity.
ating complementary electrostatic environments between Our own laboratory has developed CleftXplorer, which
binding site and ligand, but that partial charges on pro- uses spherical harmonic expansion coefficients to analyti-
tein and ligand lack face-to-face complementarity. cally describe the shape and size of binding sites and
The same conclusions were drawn for protein–protein ligands molecules.6,32–34 To obtain a more complete
interfaces.21 picture also of the physicochemical properties within
It has been suggested that physicochemical interactions binding pockets, the shape descriptor in CleftXplorer was
in natural complexes lack exact complementarity,22 most extended by various physicochemical descriptors that
likely to avoid irreversible ligand binding. Ledvina et al. characterize electrostatic charged–charged interactions,
showed that some phosphate receptors, sulfate-binding hydrophobic interactions, hydrogen bonds, and van der
proteins, flavodoxin structures, and DNase proteins exert Waals interactions (see Methods).
an intense negative electrostatic potential (ESP) at their Despite differences in their algorithms, all methodolo-
binding sites despite binding a highly negative ligand.23 gies that aim to identify a ligand that binds to a given
The proteins stabilize the anion charges by van der Waals protein binding site or, indeed, identify what the cognate
interactions and an extensive local hydrogen-bonding net- ligand might be assume that ligands and binding sites ex-
work comprising main-chain NH groups and hydroxyl hibit geometric and physicochemical complementarity. If
side chains. The question remains open whether in their this assumption is correct, binding sites that bind the
evolutionary past these enzymes were binding ligands with same cognate ligand should share very similar properties.
complementary ESPs. Similarly, Herschlag and coworkers’ Most studies so far have investigated the physicochemical
study of the electrostatic and geometric complementarity properties of individual protein–ligand complexes and
in the oxyanion hole of ketosteroid isomerase concluded have tried to highlight the complementarity between
that electrostatic complementarily makes only a modest those (see references earlier and Refs. 35–37). Our study
contribution to the enzyme’s catalytic mechanism, which sets itself apart by a detailed investigation and compari-
is primarily driven by geometrical complementarity.24 son of the physicochemical properties in sets of nonho-
Nakamura et al. listed four reasons for the absence of elec- mologous binding sites that all bind the same ligand,
trostatic complementary in some protein–ligand com- thereby revealing a surprising diversity in the protein
plexes: (1) the ligand interacts mainly with the solvent, (2) environments that a single ligand is able to bind.
hydrophobic interactions are strong, (3) dissociation of
the ionizable protein residues was incorrectly assigned,
and (4) additional ionic ligands were affecting the experi- METHODS
enced electrostatic energy of the ligand.13
Data set
To gain a wider perspective on physicochemical com-
plementarily, we have analyzed the physicochemical char- The analysis in this manuscript was performed on the
acteristics of 100 protein-binding sites that bind either same data set as published in Kahraman et al.6 The pro-

2 PROTEINS
Physicochemical Variation in Binding Pockets

tein structures in this data set were selected from the ESP and hydrophobicity score implicitly account for
protein quaternary structures (PQS)38 database, which surrounding water molecules.
contains the supposed biological unit of each crystallo- 2. For each ligand, identify and remove those polypep-
graphically determined entry in the Protein Data Bank39 tide chains and neighboring chemical compounds
(PDB). Only structures containing a cognate ligand were (NCCs) that lie within 9 Å of any ligand atom. NCCs
considered, that is, where the bound ligand is known to correspond to HET groups such as metals, cofactors,
be involved in the function of the protein as given by EC and coenzymes. The cutoff distance of 9 Å is in ac-
number40 or by the protein’s corresponding UniProt cordance with interaction ranges of nonbonded inter-
entry.41 These structures were filtered such that, for any actions.43 This speeds up the calculations while hav-
specific ligand type, the proteins binding the ligand were ing negligible effects10,44 on the calculated properties
nonhomologous, as defined by the homology (H) level (data not shown).
in the CATH protein structure classification database.42 3. At the center of each ligand atom, calculate the value
The final selection stage required that for each ligand of each physicochemical property, using the atoms of
type at least five nonhomologous protein–ligand struc- the retained protein chains and NCCs.
tures remained. The net result of this selection procedure
was nine ligand types—AMP, ATP, FAD, FMN, glucose,
heme, NAD, phosphate, and steroid molecules (estradiol
and dehydroepiandrosterone) and 100 protein-binding
Electrostatic potential
pockets containing one of these ligands.
Nine protein structures in the data set had loop The ESPs were calculated by solving the nonlinear
regions in the proximity of their binding sites for which, Poisson-Boltzmann equation45 using APBS46 (version
because of insufficient experimental data, no atom coor- 1.1). A grid size of 1 Å was chosen and the finite differ-
dinates were available in the PDB file. These loops ence discretization technique47 was applied. Calculations
regions probably exhibited high flexibility. Flexible loop were run at room temperature with a counterion concen-
regions can play a major role in binding different ligand tration of 0.1M. The continuum solvent model was allo-
molecules into a binding site, often enclosing the ligand cated a dielectric constant of es 5 78. Protein and ligand
in a cage-like structure. The flexible loop regions in our atoms as well as NCCs were assigned a dielectric constant
data set however remained or became flexible after the of e 5 4. Hydrogen atoms were added to the protein
ligand had bound to the protein structure, preventing structure using the program REDUCE48 (version 3.13),
them from being a stable integral part of the protein- which optimizes the protein’s hydrogen bond network,
binding site. As the contribution of these loop regions to while flipping ASN or GLN side-chain amides where
the ligand’s binding free energy varies with their flexibil- appropriate. Protonation states of histidine residues were
ity, we assumed that their contribution can be neglected approximated using PROPKA49 at a pH given for the
and therefore have not attempted to include them in the mother liquor in the PDB or in the primary literature of
physicochemical property calculations. the crystal structure. Partial charges and atom radii from
the PARSE50 parameter set were assigned to all protein
atoms using the automated procedure of PDB2PQR51
Mapping protein physicochemical properties (version 1.4.0). The reported potentials from APBS were
on ligand molecules
converted from kT/e to kcal/mol/e by applying the con-
The computer program CleftXplorer6 was modified so version factor of 0.592.
that in addition to describing a 3D shape of a binding The ligands in our data set as well as all NCCs were
pocket or a ligand molecule using spherical harmonics, it assigned Pauling’s van der Waals radii (which are also
could describe physicochemical properties mapped onto a used in the PARSE parameter set). All ligands in our
dot surface of that shape. The physicochemical properties data set were left uncharged in the binding sites, to
included ESPs, hydrophobicity scores, hydrogen bond do- account for the reduction of the screening effect by the
nor, acceptor, and van der Waals potential energies. These ligand molecule. Partial charges for the NCCs were
were computed using standard software packages, supple- obtained from the Merck molecular force field
mented by own code as described later. For the purpose of (MMFF94)52 as implemented in an in-house modified
this manuscript, the physicochemical properties were not version of the Chemistry Development Kit (CDK, version
mapped on the surface of a shape, but rather onto the atom 2.0.2)53 for Java. The correctness of the MMFF94 partial
center coordinates of each ligand in the data set, thereby charges was checked against the MMFF94 validation suite
following a similar approach to the GRID.28 CleftXplorer’s (http://www.ccl.net/cca/data/MMFF94/). For simplicity,
algorithm is simple and consists of three steps: metal ions were assigned partial charges equal to their
formal charge following the approach of MMFF94. Iron
1. Remove all crystallographically observed water mole- ions in heme molecules were treated as an integral part
cules from the protein structure file. The computed of the molecule and assigned a partial charge of 12e

PROTEINS 3
A. Kahraman et al.

irrespective of their oxidation state. Hydrogens were sigmoid potential function as suggested by Levitt and
added automatically to the ligands and NCCs at pH 7 Brylinski et al.62,63:
using OpenBabel54 (version 2.1.1) and in the few cases
where the automatic assignment failed were added man-  
ually using PyMOL (version 1.0).55 Phosphate molecules f distrel ¼ 1  0:5
 2 4 6 8

were always assumed to be unprotonated. The spatial 3 7 3 distrel  9 3 distrel þ 5 3 distrel  distrel ;
position of the hydrogen atoms was optimized within the
binding site using REDUCE.
where the value of distrel ranges from 0 to 1 and corre-
Scoring the hydrophobic environment sponds to the distance between a ligand’s atom center
and the center of the protein or NCC atom divided by
The calculation of CleftXplorer’s hydrophobicity scores
the cutoff distance of 9 Å.
follows the approach of HINT30 and the hydrophobicity
To calculate the total hydrophobic environment score
potential of Fauchère et al.56 in relating the hydropho-
(HES) at a ligand’s atom center, it was further necessary
bicity score to octanol–water partition coefficients
to sum the product between the distance function
(log P) values. CleftXplorer differs in that it uses atomic-
f(distrel) and the energy of transfer DGlog P over all n
based rather than fragment-based log P values. This
neighboring atoms. The following equation expresses this
makes CleftXplorer applicable to a wider spectrum of
simple sum and gives the final equation for HES:
molecules especially those for which HINT has no frag-
ment parameters. The atomic log P values were calcu- X
n  
lated using a modified version of the XlogP57 algorithm HES ¼
log P
f distrel i 3 DGi :
in the CDK. Modifications were required to add log P i¼1
values for charged molecules, as by default, log P values
are defined only for neutral molecules. Some of the Other physicochemical properties
ligands in our data set, as well as some of the NCCs,
contain charged moieties (e.g., PO4 groups), which CleftXplorer calculates van der Waals potential energies
increase the solubility of the molecule in water. If the using the ‘‘buffered 14-7’’ potential equation from the
molecule is treated as neutral its hydrophobicity will be MMFF94 force field.64 Hydrogen bond donor and
overestimated. To reduce the log P value for charged acceptor potential energies are calculated according to
atoms, a correction factor of 21.083 was introduced that the knowledge-based potential energies from Kortemme
corresponds to half of the zwitterion correction factor et al.65 These orientation-dependent hydrogen-bonding
used in the XlogP algorithm for the hydrophobicity cal- potentials were derived from geometric features found in
culation of amino acids. Metal ions were assigned an ar- a large set of high-resolution protein structures. For
bitrary log P 5 23, assuming that metal ions in solution more details, see the cited publications.
minimally interfere with the water network. For protein
atoms log P values were calculated considering each Average properties and their variation
amino acid within the tripeptide Gly-X-Gly, where X is Once the physicochemical properties of all proteins
the amino acid of interest. The values were calculated rel- in a ligand set were mapped onto their ligand mole-
ative to glycine, following the method of calculating the cules, the physicochemical property scores from differ-
Hansch p-constant,58 by subtracting the mean atomic ent binding sites were averaged and their standard
log P value for glycine atoms from each of the atomic deviation computed. For this purpose, an atom-map-
log P values giving an atomic log Prel value. Based on log ping table was created mapping corresponding atoms
Prel, the energy of transfer from water to organic solvent, between ligand A and ligand B, where both were from
DGlog P, was inferred following the approach of Eisenberg the same ligand set. In cases in which the molecular
and McLachlan59 by applying the Boltzmann equation moiety had rotational symmetry, atoms at similar spa-
from statistical thermodynamics: tial positions were associated to each other, after the
molecules were manually aligned on the neighboring
DG log P
¼ log P rel 3 2:30 3 RT ¼ log P rel 3 1:36 kcal=mol;
moieties. Atoms in the phosphate ligand set were
mapped randomly as a unique mapping due to sym-
where R is the Boltzmann factor and T is the room tem- metry was not possible. However, the small size of
perature in Kelvin. The XlogP-based ‘‘hydrophobicity phosphate molecules should generally result in similar
energy’’ correlates with the experimental free energy of physicochemical property scores over the whole mole-
transfer59 with R2 5 0.92. A similar approach to the cule.
molecular lipophilic potential (MLP)60,61 was used to In addition to the standard deviation, the relative
map the ‘‘hydrophobicity energy’’ of the protein to the number of sign changes in the property scores for each
atom center coordinates of the ligand molecules with a atom was calculated as a further means to assess the vari-

4 PROTEINS
Physicochemical Variation in Binding Pockets

ation of the physicochemical properties in binding sites. NCCs in the calculation of physicochemical properties
The sign change ratio (SCR) is given by was crucial as without them most proteins in our data
  set lacked the physicochemical properties required to
max Niþ ; Ni bind their cognate ligand.66 Of the 100 ligands in our
SCRi ¼ 1 
Nitotal dataset, 67 had 144 NCCs, of which 63 were metals,
mostly coordinated to AMP, ATP, or phosphate mole-
cules. For the remaining 33 ligands (mostly HEM, NAD,
for an atom i in a ligand. Ni1 is the number of positive and steroid), no NCC was found in the PQS entry, but
score values, Ni2 the number of negative score values, their presence in vivo can, of course, not be excluded.
and Nitotal is the size of the ligand set. The functional
range of SCR is between 0 and 0.5, with 0 denoting no
Physicochemical properties of proteins in
variation (all corresponding atoms have equal sign) and ligand-binding sites
0.5 indicating highest variation (one half of correspond-
ing atoms are positive, the other half negative). Electrostatic potential on ligands

Figure 1 shows examples of the ESP experienced by all


RESULTS ATP, NAD, and heme ligands in the data set. For the ESP
of the remaining ligand sets, see Figure X1 in Supporting
All calculations shown in this section were performed Information.
with all five physicochemical properties; that is, ESPs,
hydrophobicity scores, hydrogen bond donor, acceptor,
Adenosine-50 -triphosphate
and van der Waals potential energies. Hydrogen bond
patterns were found to vary most within all ligand sets, An adenosine-50 -triphosphate (ATP) molecule consists
whereas van der Waals potential energies were found to of an aromatic adenine ring, a ribose sugar, and a nega-
vary hardly. Nevertheless, all results and discussions in tively charged triphosphate group. For recognition and
the following were focused on the ESP and the HES, binding, the protein is expected to provide a positive
because of their importance and discriminative power in ESP for complementarity to the aromatic ring and in
molecular recognition events. particular the triphosphate tail. Figure 1(a) confirms this
For each ligand, ESP and HES were calculated in the expectation showing positive ESPs for all ATP-binding
presence of any NCC (see Methods). The inclusion of sites. An inspection of the ATP-binding sites revealed

Figure 1
Electrostatic potential that (a) ATP, (b) NAD, and (c) heme molecules ‘‘experience’’ within their binding sites. The ligands are horizontally ordered
from top left to bottom right with increasing positive potential. The potential is colored from red to white to blue for values below 25 kcal/mol/e
to 0 to values above 5 kcal/mol/e. The PQS-Id of the protein structure in which the ligand was found is given below each ligand. Note that to
make the visual comparison easier, each type of molecule is represented by a single conformer and may not be the conformer in the given PQS file.
The representative conformers are 1e2q, 1mi3, and 1po5 and have yellow labels. NAD molecules bound to oxidoreductases are labeled green; those
bound to transferases are labeled orange.

PROTEINS 5
A. Kahraman et al.

that the high positive potentials on the ATP molecules

21.94  3.93

2.87  1.19
Steroid
(on average 11.33 kcal/mol/e, see Table I) were caused
primarily by metal ions that are coordinated to one or
two phosphates at ATP’s triphosphate tail. Metal ions
were found coordinated in all ATP-binding sites except

17.14  14.64
in the DNA ligase (1a0i) and the biotin carboxylase

23.65  4.24
(1dv2). The function of those metals varies but most

PO4
often they act as catalytic stabilisers for transition states
of the enzyme–substrate complex67,68 or as charge
neutralizers between charged protein and phosphate
groups.69,70 In particular, the latter function is impor-
0.03  6.41

0.39  1.22
NAD

tant with respect to the molecular recognition of ATP as


metal ions can shield negative charges between the pro-
tein and ATP and lock the otherwise loose triphosphate
tail of ATP to the protein.71,72 The charge neutralizing
23.11  8.30

2.30  1.46

effect can increase ATP’s-binding affinity by several


HEM

orders of magnitude, as in the case of the bovine cAMP-


Average and Standard Deviation of the Experienced Electrostatic Potential and Hydrophobicity Environmental Scores for Different Ligand Sets

dependent protein kinase.73 The mouse homologue of


the same protein, 1rdq, in our data set shows, after the
exclusion of the divalent magnesium ion from the ESP
27.17  4.65

20.45  1.83

calculation, negative repulsive potentials toward the ATP


GLC

molecule. Only when the metal ion is included in the


calculation, the repulsive forces are neutralized and
become attractive (see Fig. 2). Similar neutralizing effects
of metal ions were observed in the PQS structures 1a49,
4.81  7.44

0.00  1.39

1b8a, 1e8x, 1o9t, and 1tid.


FMN

Not all ATP molecules require a metal ion to bind to


their receptor protein. As mentioned earlier, ATPs in the
binding sites of the DNA ligase (1a0i) and the biotin car-
5.27  5.93

0.94  1.21

boxylase (1dv2) were found without a metal ion or any


FAD

other NCCs. Although they are important for the cata-


The lowest and highest average score values for each property are colored purple and green, respectively.

lytic reaction of these enzymes, magnesium ions are not


essential for ATP binding. A detailed inspection of their
binding sites revealed that the proteins created attractive
11.33  9.13

21.02  2.00

positive ESPs [see Fig. 1(a)] toward the ATP phosphate


ATP

tails through phosphate-binding motifs. These motifs are


characterized by positive electrostatic fields originating
from positively polarized N-termini of a-helices and/or
positively charged lysine and arginine side chain and/or
1.53  5.87

20.65  2.53

ionic hydrogen bonds (O2HN) between backbone NH


AMP

groups and phosphate oxygen ions.74

Nicotinamide adenine dinucleotide


1.96  6.75

0.70  1.53

Nicotinamide adenine dinucleotide (NAD) molecules


All

consist of three functional groups, namely the mainly ar-


omatic adenosine moiety, the negatively charged pyro-
phosphate group, and the aromatic nicotinamide ring
Hydrophobicity Environmental

that can carry a formal positive charge when oxidized to


Physicochemical property

NAD1.75 The ESP that a NAD molecule experiences in


Electrostatic Potential

binding sites depends on the enzymatic function of the


NAD molecule. In our data set, the two most prominent
functions were the redox function in oxidoreductases
Score (HES)
(kcal/mol/e)

(1hex, 1ib0, 1jq5, 1mew, 1mi3_1, 1o04_1, 1qax, 1t2d,


Table I

2npx) and the group transfer function in transferases


(1ej2, 1og3, 1s7g, 1tox_1, 2a5f). As a redox partner in

6 PROTEINS
Physicochemical Variation in Binding Pockets

Figure 2
Electrostatic potential of the mouse cAMP-dependent protein kinase (PQS-Id 1rdq) as experienced by the cognate ligand ATP (a) without and (b)
with two coordinated magnesium ions (green-colored spheres). The potential is colored from red to white to blue for values below 25 kcal/mol/e
to 0 to values above 5 kcal/mol/e.

NAD-dependent dehydrogenases, a NAD1 molecule sup- card the nicotinamide moiety after cleavage for which re-
ports the oxidation of a substrate by accepting a hydride pulsive forces are advantageous, adenylyltransferases must
group (H2). As a substrate in ADP-ribosyltransferase reac- tightly bind the substrate nicotinamide ribonucleotide to
tions, NAD1 molecules are cleaved into an ADP-ribosyl fuse it to an ATP substrate molecule leading to the reac-
and a nicotinamide group, where the former is transferred tion product NAD1.
to an arginine group within the active site as part of a For the bacterial NADH peroxidase (2npx) repulsive
posttranslational modification and the latter is released potentials were found for the adenine moiety of NADH
into the solvent. For dehydrogenases, the nicotinamide (2npx is the only protein in our data set having bound
moiety should experience negative ESP due to the H2 NADH. All other NAD proteins have bound the oxidized
group transfer, whereas in ribosyl-transfer reactions the NAD1). The negative electrostatic field is created by six
positively charged nicotinamide moiety should experience negatively charged glutamate and aspartate amino acids
repulsive positive ESP. Figure 1(b) shows the ESP of all in close proximity. An examination of the residues
NAD molecules in our data set ordered by the total ESP around the nicotinamide moiety revealed that the repul-
experienced by the molecules from top left to bottom sive potentials were overridden by aromatic interactions
right. NAD molecules that are functioning as oxidation to a hydrophobic valine and an isoleucine side chain [see
partners were labeled green, whereas molecules acting as Fig. 4(a)]. Similar aromatic interactions were found for
substrates in transferase reactions were labeled orange.
All NAD molecules bound to oxidoreductases con-
firmed the expectation and showed an attractive negative
ESP at the majority of their nicotinamide moiety with
the exception of 1ib0, a rat cytochrome b5 reductase,
which experienced repulsive positive ESP. As the reduc-
tase name indicates 1ib0 distinguishes itself by using the
substrate NADH as a reducing agent. To increase the af-
finity toward NADH over NAD1, the enzyme was found
to use the observed positive ESP over the nicotinamide
moiety as a repulsive force against NAD1. The positive
ESP was created mainly at a conserved binding motif
‘‘CGPPPM’’76 with three prolines located at a positively
polarized N-terminus of an adjoining a-helix (see Fig. 3)
The transferase structure of the ADP-ribosyltransferase
1og3 and partially the structure of the NAD-dependent
deacetylase 1s7g exerted as expected repulsive electroposi-
tive potentials toward the nicotinamide moiety of NAD1. Figure 3
The structure of the nicotinamide mononucleotide NAD1-binding site of the rat cytochrome b5 reductase. The molecular
adenylyltransferase 1ej2 however exerted attractive nega- surface around the NAD molecule is shown transparent and colored
according to the electrostatic potential from red to white to blue for
tive ESP. Closer inspections showed that the attractive values below 25 kcal/mol/e to 0 to values above 5 kcal/mol/e. The
ESP was necessary for the function of 1ej2, namely to adjacent FAD molecule is shown in green, whereas the three proline
synthesize NAD1 molecules. Although transferases dis- residues of the conserved ‘‘CGPPPM’’ motif are shown in yellow.

PROTEINS 7
A. Kahraman et al.

Figure 4
(a) The repulsive electrostatic potential of NADH peroxidase toward the adenine moiety of NADH is compensated via p–CH interactions between
the adenine moiety of NADH (varicolored) and Ile 180 (green colored) and Val 242 (green colored). (b) ADP-ribosyltransferase stabilizes the
binding of NAD1 via a p–p interaction with Phe160 (green colored), p–NH interaction with the backbone amide of Ser148 (green colored), and an
internal hydrogen bond (black dashed line) between the amide group and the phosphate group.

the nicotinamide group of 1og3 [see Fig. 4(b)], where well known to contribute strongly to the binding energy
p–p stacking interaction to a phenylalanine from one side of heme molecules.79–81
and an aromatic-backbone-amide interaction from the Among all 16 heme-binding pockets analyzed, the bo-
other side provide the necessary binding free energy for vine endothelial nitric oxide synthase (eNOS) (PQS-Id:
the nicotinamide moiety of the NAD molecule. The im- 1d0c) was the only enzyme that exposes the Fe of its
portance of aromatic interactions in our data set will be heme ligand to a repulsive positive ESP. The positive ESP
described later in the text. Similar observations were made was mainly created by a loop region on the proximal site
on adenine groups that were found to bind protein-bind- of the heme molecule (see Fig. 5). The loop region
ing sites primarily with intermolecular hydrogen bonds to- includes the Cys 186, which coordinates Fe, and three ar-
gether with p–p stacking and cation–p interactions.77,78

Heme type B

The heme molecule consists of an aromatic porphyrin


ring system with a central coordinated iron ion and two
negatively charged propionate groups sitting ‘‘on top’’ of the
porphyrin ring system. Arginine residues form ionic bonds
to the propionate groups and anchor the heme molecule to
the protein structure, whereas histidine and less frequently
cysteine residues coordinate the heme’s iron ion and affect
the electrochemical character of the iron metal ion (Fe).
Based on these characteristics, one would expect posi-
tive ESP around the propionate groups and negative
potential in the porphyrin ring system around the iron
ion. Visually, this prediction seemed true for the PQS
structures 1d7c, 1dk0, 1iqc_1, 1naz, and 1np4. However,
the remaining heme-binding sites, in particular, 1d0c,
1po5, 1qla, and 2cpo, show repulsive positive electrostatic
fields around the heme’s porphyrin group, causing the
heme molecules to have a diverse range of ESP values
with an average of 23.11 kcal/mol/e and a standard devi- Figure 5
Binding site of the bovine endothelial nitric oxide synthase (eNOS)
ation of 8.30 kcal/mol/e (see Table I). Heme molecules
with the blue-colored electrostatic potential isosurfaces at 10 kcal/mol/e.
are bound by more than 20 different protein folds79 The heme molecule is shown in stick representation and on its
leading to a large diversity of environments around the proximal site are depicted the loop region with Cys 186 (yellow color)
molecules. In-depth inspection of the protein environ- coordinated to the heme’s iron ion (dashed line) and Arg 189 (red
color). The loop main-chain amine groups are also shown in stick
ments around the heme molecules shows that, like representation. NCCs, here the inhibitor 7-NIBr (PDB-ligand name
NAD’s nicotinamide moiety, repulsive ESPs are overrid- INE) and the crystallization agent acetate (PDB-ligand name INE), are
den by aromatic interactions. Aromatic interactions are shown in green color.

8 PROTEINS
Physicochemical Variation in Binding Pockets

higher ESP on the Fe ion generally induces a higher redox


potential for the heme molecule (see Fig. 6 and Table II).
The quinol:fumarate reductase (QFR) of Wolinella succi-
nogenes (PQS Id: 2bs2) binds two heme molecules, namely
a heme at the proximal site of the protein structure (heme
bP) and a heme at a distal site (heme bD). Both heme mol-
ecules have distinct redox potential. For many years it was
unclear, which of the heme molecules corresponded to the
low-potential (heme bL) and high-potential (heme bH)
heme molecule. With the computational simulation of the
electrochemical redox titration behavior of both heme mol-
ecules using electrostatic calculations, Haas and Lancaster97
were able to identify the low-potential heme as heme bD
and the high-potential heme as heme bP. Using the above
relationship between ESP and redox potential, we were
able to confirm Haas and Lancaster’s finding with an ESP
of 218.84 kcal/mol/e for heme bD and a much higher ESP
Figure 6
of 28.02 kcal/mol/e for heme bP.
Plot showing the correlation between heme redox potential and the
electrostatic potential that was measured on the heme’s iron ion (see
also Table II).
Remaining ligand sets

AMP-binding sites often generate positive ESPs toward


their ligand’s phosphate group (see Fig. X1 in Supporting
ginine residues of which Arg 189 is located right on top Information). Five of the proteins, namely 1amu, 1ct9_1,
of Fe. Main-chain NH groups in the loop region that are 1jp4, 1qb8, and 1tb7, had at least one metal ion bound
pointing inward into the loop were further strengthening within 9 Å proximity inducing a positive potential on
the positive ESP. Although the observed repulsive ESP the phosphate tail. The AMP molecule in asparagine syn-
seemed counterintuitive, it demonstrated once more that thetase (12as) compensated the repulsive global negative
the complementarity between proteins and ligands was ESPs through a local network of eight hydrogen bonds.
driven by the ligand’s role in the enzyme function and The flavin moiety in FAD as well as in FMN experienced
not by its chemical composition. eNOS must bind a pos- both negative and positive potentials in our data set. The
itively charged L-Arg substrate molecule to perform its phosphate group in both ligands was however always
function, namely creating nitric oxide. A heme Fe iron attracted by positive potentials. In the riboflavin kinase
ion in a reduced oxidation state Fe(II) is thereby advan-
tageous as it increases the binding affinity toward L-Arg
by 80-fold compared with the oxidized state Fe(III).82 Table II
Therefore, there is an interest from the enzyme to stabi- Comparison Between the Electrostatic Potential (ESP) on the Heme’s
Iron Ion and Its Redox Potential Against the Standard Hydrogen
lize the Fe(II). The positive ESP on the Fe of eNOS is Electrode (Em)
thereby advantageous, as it stabilizes the less positively
charged Fe(II) state. Furthermore, the positive electro- PQS-Id ESP (kcal/mol/e) Em (mV) Reference
static field increases the redox potential of Fe making the 1naz 242.81 ?
electron transfer from a distant FMN molecule thermo- 1dk0 233.13 2550 88
dynamically favorable.83 Finally, the field stabilizes the 1np4 224.67 2259 89
1qpa 223.65 2130 90
negatively charged Cys 186, which again increases the re- 1sox 221.45 168 91
dox potential of the heme molecule. 1pp9 220.03 210 92
In fact, the redox potential of the heme’s iron ion is 1qhu 19.20 ?
one of the most important electrochemical characteristics 1iqc_1 217.81 1450 85
1d7c 216.49 1169 93
of heme molecules. Heme redox potentials can vary 1eqg 210.69 250 94
between 2550 mV (in the hemophore HasA84) and 2bs2 28.02 29 95
1450 mV (in peroxidases85) versus the standard hydro- 1po5 27.30 ?
1ew0 25.82 ?
gen electrode. Correlating the redox potential to protein
1gwe 25.36 ?
electrostatics might be obvious; however, such a correla- 2cpo 20.61 2140 96
tion is not straightforward as the redox potential is influ- 1d0c 3.55 ?
enced by many different factors such as the type of the
See also Figure 6.
heme’s axial ligands, heme burial, and local ?, No redox potential was found in the literature for the protein at near crystalli-
charges.80,86,87 Nevertheless, our calculations show that zation condition.

PROTEINS 9
A. Kahraman et al.

(1p4m), the pyrophosphate of the product ADP located aliphatic hydrocarbons, such as valine, leucine, or isoleu-
next to FMN caused the repulsive negative potential at cine side chains, forming p–CH interactions.100,101 The
the phosphate group. The negative potential on the aro- second most often observed aromatic interactions were
matic ring of the same ligand was compensated by aro- p–p interactions between aromatic ring groups mostly
matic interactions. from phenylalanine and tyrosine residues.102 Other aro-
Glucose-binding sites were electrostatically most neg- matic interaction types were p–cation interactions
ative in our data set with an average ESP of 27.17 between aromatic rings and positively charged side chains
kcal/mol/e (see Table I). In contrast, phosphate-binding of arginine and lysine103 and p–HN interactions with
sites were most positive with an average potential of backbone amide groups that act as a hydrogen bond do-
17.14 kcal/mol/e (see Table I). Metal ions were found nor toward an aromatic ring.104 Also observed were p-
within 9 Å proximity in nine cases (1a6q, 1e9g, 1ew2, proline interactions.105 Aromatic interactions, in partic-
1h6l, 1ho5_1, 1l7m_1, 1lby, 1qf5, and 1tco) generating ular p–CH and p–p, are in general dominated by van der
positive potentials around the PO4 molecules. An inter- Waals dispersion and hydrophobic interactions.106,107
esting exception was the binding site of the bacterial Electrostatic interactions play only a secondary role and
dethiobiotin synthetase (1dak). This particular binding only contribute where aromatic interactions involve pro-
site apparently completely lacked the expected physico- tein hydroxyl or amine groups that have a high electro-
chemical properties to bind a phosphate molecule. Not static dipole moment.108 Accurate dispersion energy cal-
only it did exert repulsive negative potentials, but also culations require sophisticated and computationally
it lacked any compensating polar interactions and/or demanding quantum chemistry calculations106 that are
hydrogen bonds toward its ligand. Also, van der Waals beyond the scope of this work. Instead, we analyzed the
interactions could be ruled out as the phosphate mole- hydrophobicity of the surrounding protein residues and
cule was solvent accessible on up to 70% of its molec- NCCs and assessed the contribution of hydrophobic
ular surface. We could not find evidence that the mol- interactions in overriding electrostatic repulsion between
ecule has bound at a crystal contact interface. Unfortu- protein and ligand.
nately, the structure factors for 1dak were available at Figure 7 depicts the HES (see Methods) for the bind-
neither the PDB nor the Electron Density Server,98 ing sites in Figure 1, colored from magenta to white to
preventing further investigations of the electron density green for 21 to 0 to 1 HES for polar, neutral, and
of the phosphate molecule at its current location. The hydrophobic environments, respectively. From Figure 7,
50% occupancy value given for the entire phosphate it is clear that aromatic ring systems are generally located
molecule in the current crystal structure of 1dak might in a hydrophobic environment. In particular, the varia-
indicate an alternative location of the phosphate mole- tion observed for ESP in heme-binding sites does not
cule in the protein structure. And finally, as for heme exist for HES. All heme molecules are located entirely in
molecules, the molecular recognition of steroid mole- hydrophobic-binding pockets whether facing attractive or
cules is driven by hydrophobic and aromatic interac- repulsive ESP. Similar dominant HES were observed for
tions99 that override the observed repulsive ESP in ste- steroid molecules and flavin moieties (see Fig. X2 in Sup-
roid-binding sites in our data set. porting Information). The adenine moieties of ATP and
NAD were usually found in a hydrophobic environment,
whereas their charged phosphate groups tend to face a
Nonelectrostatic interactions between protein and ligand
polar environment. The NAD molecules (PQS Ids 1ibo,
Our results show that binding events are not solely 1hex, 1mew, 1og3, 1qax, and 2npx) that were found to
directed by classical electrostatic interactions between experience repulsive electrostatic forces at their adenine
partially charged atoms, as there are many examples of and nicotinamide moieties simultaneously form attractive
noncomplementary ESPs between protein and bound aromatic-hydrophobic interactions.
ligand. Thus, other nonelectrostatic forces acting between
ligand and protein must contribute and result in negative
Average and variation of physicochemical properties
binding free energies and lead to the binding of the
ligand to its binding site. With the exception of glucose Despite the various conformations of a ligand in dif-
and phosphate, all our ligands possess at least one aro- ferent protein-binding sites, one can calculate the average
matic ring complex, and most of the variation in terms ESP and HES and its standard deviation for each ligand
of ESP is observed at these ring systems. Ligand moieties atom in the protein-binding pocket. The average atomic
that are highly charged, like the phosphate groups, expe- ESP and HES were calculated for each ligand set and
rience rather constant complementary electrostatic fields. plotted in Figures 8 and 9. Next to the figures are
Close inspections of the binding site ligand complexes in depicted ESP and HES generated by the ligands in isola-
our data set showed that almost all aromatic ring groups tion. In contrast to many single cases (see in particular
undergo various forms of aromatic interactions. Most of NAD and heme in Fig. 1), the average values of the phys-
these interactions were of a hydrophobic nature toward icochemical properties show complementary characteris-

10 PROTEINS
Physicochemical Variation in Binding Pockets

Figure 7
Hydrophobicity experienced by (a) ATP, (b) NAD, and (c) heme molecules within their binding pockets. The level of experienced hydrophobicity is
given by the hydrophobic environment scores (HES) (see Methods) and is colored from magenta to white to green for polar environments with less
than 21 HES to 0 to hydrophobic environments with values above 1 HES. The ligands are ordered as in Figure 1. The PQS-Id of the protein
structure in which the ligand was found is given below each ligand. The representative conformer for each set has a yellow label.

tics between protein and ligand for the majority of the strating that the average physicochemical properties show
cases. From Figure 8, it becomes evident that usually a higher complementarity than individual protein–ligand
electrostatically positive protein potentials are neutralized pairs.
by electrostatically negative ligand potentials and vice The standard deviation of the absolute ESP and the
versa. HES is given in Table I. ESP shows large variations with
A closer look at the ESP in Figure 8(a) reveals average standard deviations several times larger than the average
positive ESPs for all adenine groups and all phosphate value. The highest variation is seen for phosphate-bind-
groups, whether as separate entities or part of larger mol- ing pockets having a standard deviation of 14.64 kcal/
ecules. Note that despite the similarity between AMP and mol/e, whereas the lowest variation is seen in steroid-
ATP, the latter is usually bound in pockets with higher binding pockets with 3.93 kcal/mol/e. The variation of
positive potential because of the larger number of metal HES is generally less and is around four times smaller
ions coordinated to the phosphate tail of ATP. The highly than the average standard deviation of ESP.
hydrophobic heme and steroid molecules, as well as glu- Although the absolute values of ESP and HES have a
cose and the nicotinamide moiety in NAD, are mainly high standard deviation and thus are very variable, their
surrounded by negative potential. Figure 9(a) shows that values are consistently of the same sign. Figure 10 shows
steroids and the porphyrin core of heme molecules are for each ligand atom the SCR (see Methods) for ESP and
bound in very hydrophobic environments together with HES. Cyan color indicates no variation (0%) for a ligand
the flavin moieties in FAD and FMN. Because of the atom with score values all with the same sign; the highest
large number of p–CH and p–p interactions involving variation (50%) is colored orange reflecting half positive
adenine moieties in our data set, these moieties experi- and half negative score values. According to Figure 10,
ence large HES despite being of moderate polar nature. half of all ligand atoms in the data set experience ESP
The comparison of the averaged ESPs from the pro- and hydrophobicity scores of low SCR, whereas half ex-
teins [Fig. 8(a)] and the ligands [Fig. 8(b)] on each perience high SCR. Hydrophobic moieties in ligand mol-
ligand atom reveals a correlation of only R2 5 0.25. The ecules often occupy equivalent hydrophobic environ-
correlation for HES is higher with R2 5 0.66. Both num- ments, like the flavin groups in FAD, FMN or the por-
bers are relatively small and indicate rather little interde- phyrin core in heme molecules or steroid molecules, but
pendence between the protein’s or ligand’s physicochemi- have their electrostatic environments change. Similar
cal properties. However, they still exceed the correlation observations can be made for central phosphate groups
between all 100 individual binding site/ligand complexes, in FAD and NAD. Thus, the interactions are on average
with R2 5 0.14 for ESP and R2 5 0.35 for HES, demon- complementary but vary widely in their magnitude.

PROTEINS 11
A. Kahraman et al.

Figure 8
(a) Average electrostatic potential that each ligand set in the data set ‘‘experiences’’ within protein-binding sites. (b) Electrostatic potential
calculated for each ligand molecule in isolation. The electrostatic potential is colored from red to white to blue for values below 25 kcal/mol/e
to 0 to above 5 kcal/mol/e.

Figure 9
(a) Average hydrophobicity that each ligand set in the data set ‘‘experiences’’ within protein-binding sites. (b) Hydrophobicity of the ligand
molecule in isolation. The level of experienced hydrophobicity is given by the hydrophobic environment scores (HES) (see Methods) and is colored
here from magenta to white to green for polar environments with less than 21 HES to 0 to hydrophobic environments with values above 1 HES.

12 PROTEINS
Physicochemical Variation in Binding Pockets

Figure 10
Variation of physicochemical property scores measured by the sign change ratio of the scores at each atom. Variation is colored from cyan to white
to orange for relative sign changes of 0–25–50%. 0% represents no score variation for the property. 50% illustrates highest variation with half of
the ligand set having positive and half of the ligand set having negative score values.

DISCUSSION ognition, including the contribution of hydrophobic


interactions to molecular recognition.109,110 To investi-
Our starting point for this work was to explore the va- gate hydrophobic interactions, we have developed a new
lidity of the assumption that binding sites, which bind hydrophobicity environment score that is based on a sta-
the same ligand, achieve their selectivity by providing a tistical thermodynamic description of atomic log P values.
similar physicochemical environment. To analyze this In this work, the hydrophobicity was observed to vary rel-
ligand binding pocket relationship, we have performed atively little among each set of ligands. In contrast, the
semiempirical calculations to compute and then compare ESP varies widely and is highly influenced by NCCs. The
the electrostatic and hydrophobic environments in 100 importance of entropic energy contributions is also sup-
protein binding sites. The binding sites were analyzed in ported by the ‘‘buffer zone,’’ that is, the free space
groups with each group member binding the same between a binding pocket and its bound ligand,6 which
ligand. The results reveal large variations in protein envi- can stabilize entropically the protein–ligand complex by
ronments experienced by each ligand. These large quanti- allowing the ligand and the binding site residues to retain
tative differences indicate the absence of perfect physico- some of their vibrational motion and flexibility.9 It
chemical complementarity between binding sites and should be pointed out that the correlation of our com-
ligands. A biochemical explanation for this variation puted properties with experimental observations (e.g., re-
could often be deduced from the differing requirements dox potentials) is modest at best. Herein, we have not
of proteins to carry out their function. Nevertheless, this considered any kinetics that accompany protein–ligand
work gives evidence that—similar to the concept of con- interaction or the binding specificity and discrimination
vergent evolution of gene and protein function—nature power of proteins toward their cognate ligands, which all
has evolved multiple binding solutions for the same add a further level of complexity to the analysis of the
ligand. For some proteins, it may not matter how the protein–ligand interaction in silico. It is well known that
ligand binds to its receptor, but merely that it binds and conformational changes in particular in allosteric enzymes
so a diverse range of strategies of binding the same have large effects on the molecular recognition of sub-
ligand have evolved. strate molecules and thus on the protein function. In fact,
Regardless of the simplicity of our approach, we were recent results on in vivo proteins have shown that pro-
able to observe some general principles of molecular rec- teins, in particular cell signaling proteins and transcrip-

PROTEINS 13
A. Kahraman et al.

tion factors, are commonly partially or fully intrinsically rigid.115 Our own docking calculations with a MM-
disordered.111 They usually undergo transitions from GBSA model116 on our data set showed unfavorable pos-
their disordered conformation to an ordered structure itive energies of up to 1170 kcal/mol/e for 20 protein–
upon ligand binding, thereby adopting different isomeric ligand complexes (data not shown). The results in this ar-
states. Experiments on antibody–antigen complexes ticle show that computational shortcuts based on comple-
showed that each of the isomeric states could bind differ- mentarity can be dangerous and that more sophisticated
ent antigens with high specificity though with low affin- models would be required for accurate predictions of
ity.112 Furthermore, most allosteric and induced-fit binding energies and binding poses that take into account
enzymes lack a correlation between binding affinity and all factors and forces currently known to contribute to
specificity because of the smaller magnitude of DDG molecular binding. In particular, desolvation energies117
when compared with the free binding energy DG.113 at the protein-binding site and entropy gain/loss of ligand
Consequently, proteins often lack physicochemical com- and binding site molecules109 will need addressing. This
plementarity toward their ligands, which can reduce the complexity will provide a challenge for computational
binding affinity but retain a high binding specificity. biology, be it for the accurate calculation of binding free
Our observations confirm the results found by some energies or the derivation of more empirical approaches.
earlier studies that were made on limited datasets (see
references in the Introduction), yet the myth of exact CONCLUSIONS
complementarity remains. The results of our analysis
have consequences for a number of applications, particu- This manuscript expands on previous work on the ge-
larly those related to function prediction and virtual ometrical variation in protein-binding pockets and
screening. Function prediction methods from structure ligands6 by analyzing the same binding-pocket/ligand-
are frequently based on detecting similarities between complexes with respect to their physicochemical proper-
annotated and functionally unknown binding sites. The ties. Our analysis of the binding pocket geometries
very basis of these approaches is, however, challenged if showed that binding pockets that bind the same ligand
complementarity is not a given. Such methods must cope had a greater variation in their shapes than can be
with this variation and include a probabilistic term in accounted for by the conformational variability of their
their similarity search that accounts for our observation ligand. Here, we have continued our analysis on the
that the same ligand can bind in one binding site via same binding sites and found that it is not the chemical
electrostatic interactions, in another via entropic contri- composition of a ligand molecule that determines the
butions or in a third via purely van der Waals interac- degree of complementarity between protein and ligand,
tions. A workaround to the probabilistic description but rather the ligand’s role within the function of the
could be the utilization of ‘‘3D consensus binding pro- protein. Our analyses show that the physicochemical
files.’’ Such profiles would represent the physicochemical properties ligands experience when bound to different
environment that a ligand on average experiences in vari- binding pockets can vary significantly (Figs. 1 and 7 and
ous proteins. A similarity search would then involve the Table I). This high variation reflects large energy fluctua-
comparison of the physicochemical properties of a poten- tions that are sometimes functionally necessary (Figs. 3
tial ATP-binding site against a set of 3D consensus bind- and 5). Some of the variation included even changes in
ing profiles. Not only would this decrease the screening the sign of the potentials for corresponding atoms in a
time for a whole database, but it also might increase the ligand set (see Fig. 10). To overcome repulsive electro-
prediction accuracy, as the average properties tend to static interactions, proteins were observed to use NCCs
have higher complementarity to the ligand properties (see Fig. 2) and attractive aromatic interactions (see Fig.
than single binding-site/ligand-complexes. 4) to their ligands. Complementarity was observed to
Despite well-developed theory and extensive progress some extent only for averaged properties in the ligand
in molecular simulations, computational approaches to set (Figs. 8 and 9). With our in-depth analysis, we were
calculate both enthalpic and entropic energies are still also able to identify false binding sites and determine the
limited in accuracy.109 The accurate computation of identity of electrochemically active molecules in proteins
these terms is, however, necessary to understand the fine (Fig. 6 and Table II).
balance of binding forces underlying protein–ligand inter-
actions and to make reliable predictions of binding ener-
gies. Scoring functions implemented in docking applica- ACKNOWLEDGMENTS
tions typically use a rather simplistic model of molecular The authors thank Rafael J. Najmanovich for valuable
recognition to allow the screening of a myriad binding comments and discussion throughout the work and
poses.114 As a result, the average accuracy of docking James D. Watson for his valuable comments on early
applications lies within 1.5–2 Å root-mean-square devia- drafts of this manuscript. All figures containing mole-
tion in about 70–80% of cases. Most of these successes cules were rendered using PyMOL (W.L. DeLano, http://
are cases in which ligand and protein are relatively pymol.sourceforge.net/).

14 PROTEINS
Physicochemical Variation in Binding Pockets

REFERENCES 22. Kangas E, Tidor B. Electrostatic complementarity at ligand binding


sites: application to chorismate mutase. J Phys Chem B 2001;105:
1. Tsai C-J, Norel R, Wolfson HJ, Maizel JV, Nussinov R. Protein- 880–888.
Ligand Interactions: Energetic Contributions and Shape Comple- 23. Ledvina PS, Yao N, Choudhary A, Quiocho FA. Negative electro-
mentarity. In: Encyclopedia of Life Sciences. John Wiley & Sons, static surface potential of protein sites specific for anionic ligands.
Ltd: Chichester http://www.els.net/ [doi: 10.1038/npg.els.0001343]. Proc Natl Acad Sci USA 1996;93:6786–6791.
2. Sotriffer C, Klebe G. Identification and mapping of small-molecule 24. Kraut DA, Sigala PA, Pybus B, Liu CW, Ringe D, Petsko GA, Hers-
binding sites in proteins: computational tools for structure-based chlag D. Testing electrostatic complementarity in enzyme catalysis:
drug design. Farmaco 2002;57:243–251. hydrogen bonding in the ketosteroid isomerase oxyanion hole. Plos
3. Fersht AR. Catalysis, binding and enzyme-substrate complementar- Biol 2006;4:501–519.
ity. Proc R Soc Lond B Biol Sci 1974;187:397–407. 25. Wang RX, Gao Y, Lai LH. LigBuilder: a multi-purpose program for
4. Liang J, Edelsbrunner H, Woodward C. Anatomy of protein pockets structure-based drug design. J Mol Model 2000;6:498–516.
and cavities: measurement of binding site geometry and implica- 26. Miranker A, Karplus M. Functionality maps of binding sites: a mul-
tions for ligand design. Protein Sci 1998;7:1884–1897. tiple copy simultaneous search method. Proteins 1991;11:29–34.
5. An J, Totrov M, Abagyan R. Pocketome via comprehensive identifi- 27. Laurie AT, Jackson RM. Q-SiteFinder: an energy-based method for
cation and classification of ligand binding envelopes. Mol Cell Pro- the prediction of protein-ligand binding sites. Bioinformatics 2005;
teomics 2005;4:752–761. 21:1908–1916.
6. Kahraman A, Morris RJ, Laskowski RA, Thornton JM. Shape varia- 28. Goodford PJ. A computational procedure for determining energeti-
tion in protein binding pockets and their ligands. J Mol Biol 2007; cally favorable binding sites on biologically important macromole-
368:283–301. cules. J Med Chem 1985;28:849–857.
7. Smith RD, Hu L, Falkner JA, Benson ML, Nerothin JP, Carlson HA. 29. Nicholls A, Sharp KA, Honig B. Protein folding and association:
Exploring protein-ligand recognition with binding MOAD. J Mol insights from the interfacial and thermodynamic properties of
Graph Model 2006;24:414–425. hydrocarbons. Proteins 1991;11:281–296.
8. Babine RE, Bender SL. Molecular recognition of protein-ligand 30. Kellogg GE, Semus SF, Abraham DJ. HINT: a new method of em-
complexes: applications to drug design. Chem Rev 1997;97:1359– pirical hydrophobic field calculation for CoMFA. J Comput-Aided
1472. Mol Des 1991;5:545–552.
9. Boehm HJ, Klebe G. What can we learn from molecular recognition 31. Binkowski TA, Naghibzadeh S, Liang J. CASTp: computed atlas of
in protein-ligand complexes for the design of new drugs? Angew surface topography of proteins. Nucleic Acids Res 2003;31:3352–
Chem Int Ed Engl 1996;35:2588–2614. 3355.
10. Basu G, Sivanesan D, Kawabata T, Go N. Electrostatic potential of 32. Kahraman A, Morris RJ, Laskowski RA, Thornton JM. Variation of
nucleotide-free protein is sufficient for discrimination between ade- geometrical and physicochemical properties in protein binding
nine and guanine-specific binding sites. J Mol Biol 2004;342:1053– pockets and their ligands. BMC Bioinformatics 2007;8:S1.
1066. 33. Morris RJ. An evaluation of spherical designs for molecular-like
11. Livesay DR, Jambeck P, Rojnuckarin A, Subramaniam S. Conserva- surfaces. J Mol Graph Model 2006;24:356–361.
tion of electrostatic properties within enzyme families and superfa- 34. Morris RJ, Najmanovich RJ, Kahraman A, Thornton JM. Real
milies. Biochemistry 2003;42:3464–3473. spherical harmonic expansion coefficients as 3D shape descriptors
12. Tsuchiya Y, Kinoshita K, Nakamura H. Structure-based prediction for protein binding pocket and ligand comparisons. Bioinformatics
of DNA-binding sites on proteins using the empirical preference of 2005;21:2347–2355.
electrostatic potential and the shape of molecular surfaces. Proteins 35. Gruber J, Zawaira A, Saunders R, Barrett CP, Noble ME. Computa-
2004;55:885–894. tional analyses of the surface properties of protein-protein interfa-
13. Nakamura H, Komatsu K, Nakagawa S, Umeyama H. Visualization ces. Acta Crystallogr D Biol Crystallogr 2007;63:50–57.
of electrostatic recognition by enzymes for their ligands and cofac- 36. Gabb HA, Jackson RM, Sternberg MJ. Modelling protein docking
tors. J Mol Graph 1985;3:2–11. using shape complementarity, electrostatics and biochemical infor-
14. Nakamura H, Komatsu K, Umeyama H. Electrostatic complemen- mation. J Mol Biol 1997;272:106–120.
tarities between guest ligands and host enzymes. J Phys Soc Jpn 37. Gerczei T, Asboth B, Naray-Szabo G. Conservative electrostatic
1985;54:3257–3260. potential patterns at enzyme active sites: the anion-cation-anion
15. Tsuchiya Y, Kinoshita K, Nakamura H. Analyses of homo-oligomer triad. J Chem Inf Comput Sci 1999;39:310–315.
interfaces of proteins from the complementarity of molecular sur- 38. Henrick K, Thornton JM. PQS: a protein quaternary structure file
face, electrostatic potential and hydrophobicity. Protein Eng Des Sel server. Trends Biochem Sci 1998;23:358–361.
2006;19:421–429. 39. Berman HM, Westbrook J, Feng Z, Gilliland G, Bhat TN, Weissig
16. Chau PL, Dean PM. Electrostatic complementarity between proteins H, Shindyalov IN, Bourne PE. The Protein Data Bank. Nucleic
and ligands. I. Charge disposition, dielectric and interface effects. Acids Res 2000;28:235–242.
J Comput-Aided Mol Des 1994;8:513–525. 40. Bairoch A. The ENZYME database in 2000. Nucleic Acids Res 2000;
17. Chau PL, Dean PM. Electrostatic complementarity between proteins 28:304–305.
and ligands. II. Ligand moieties. J Comput-Aided Mol Des 1994;8: 41. Apweiler R, Bairoch A, Wu CH, Barker WC, Boeckmann B, Ferro
527–544. S, Gasteiger E, Huang H, Lopez R, Magrane M, Martin MJ, Natale
18. Chau PL, Dean PM. Electrostatic complementarity between proteins DA, O’Donovan C, Redaschi N, Yeh LS. UniProt: the Universal Pro-
and ligands. III. Structural basis. J Comput-Aided Mol Des 1994;8: tein Knowledgebase. Nucleic Acids Res 2004;32:115.
545–564. 42. Pearl F, Todd A, Sillitoe I, Dibley M, Redfern O, Lewis T, Bennett
19. Naray-Szabo G, Nagy P. Electrostatic complementarity in molecular C, Marsden R, Grant A, Lee D, Akpor A, Maibaum M, Harrison A,
aggregates. IX. Protein-ligand complexes. Int J Quantum Chem Dallman T, Reeves G, Diboun I, Addou S, Lise S, Johnston C, Sil-
1989;35:215–221. lero A, Thornton JM, Orengo C. The CATH domain structure data-
20. Naray-Szabo G. Electrostatic complementarity in molecular associa- base and related resources Gene3D and DHS provide comprehen-
tions. J Mol Graph 1998;7:76–81. sive domain family information for genome analysis. Nucleic Acids
21. McCoy AJ, Chandana Epa V, Colman PM. Electrostatic complemen- Res 2005;33:247.
tarity at protein/protein interfaces. J Mol Biol 1997;268:570– 43. Mackerell AD, Jr. Empirical force fields for biological macromole-
584. cules: overview and issues. J Comput Chem 2004;25:1584–1604.

PROTEINS 15
A. Kahraman et al.

44. Nielsen JE, McCammon JA. Calculating pKa values in enzyme ase complexed with substrates and products reveal the methionine-
active sites. Protein Sci 2003;12:1894–1901. ATP recognition and give insights into the catalytic mechanism.
45. Honig B, Nicholls A. Classical electrostatics in biology and chemis- J Mol Biol 2003;331:407–416.
try. Science 1995;268:1144. 68. Larsen TM, Benning MM, Rayment I, Reed GH. Structure of the
46. Baker NA, Sept D, Joseph S, Holst MJ, McCammon JA. Electro- bis(Mg21)-ATP-oxalate complex of the rabbit muscle pyruvate ki-
statics of nanosystems: application to microtubules and the ribo- nase at 2.1 A resolution: ATP binding over a barrel. Biochemistry
some. Proc Natl Acad Sci USA 2001;98:10037–10041. 1998;37:6247–6255.
47. Klapper I, Hagstrom R, Fine R, Sharp K, Honig B. Focusing of elec- 69. Zheng J, Knighton DR, ten Eyck LF, Karlsson R, Xuong N, Taylor
tric fields in the active site of Cu-Zn superoxide dismutase: effects SS, Sowadski JM. Crystal structure of the catalytic subunit of
of ionic strength and amino-acid modification. Proteins 1986;1:47– cAMP-dependent protein kinase complexed with MgATP and pep-
59. tide inhibitor. Biochemistry 1993;32:2154–2161.
48. Word JM, Lovell SC, Richardson JS, Richardson DC. Asparagine 70. Schmitt E, Moulinier L, Fujiwara S, Imanaka T, Thierry JC, Moras
and glutamine: using hydrogen atom contacts in the choice of side- D. Crystal structure of aspartyl-tRNA synthetase from Pyrococcus
chain amide orientation. J Mol Biol 1999;285:1735–1747. kodakaraensis KOD: archaeon specificity and catalytic mechanism of
49. Li H, Robertson AD, Jensen JH. Very fast empirical prediction adenylate formation. EMBO J 1998;17:5227–5237.
and rationalization of protein pKa values. Proteins 2005;61:704– 71. Masuda S, Murakami KS, Wang S, Anders Olson C, Donigian J,
721. Leon F, Darst SA, Campbell EA. Crystal structures of the ADP
50. Sitkoff D, Sharp KA, Honig B. Accurate calculation of hydration and ATP bound forms of the bacillus anti-sigma factor SpoIIAB in
free-energies using macroscopic solvent models. J Phys Chem 1994; complex with the anti-anti-sigma SpoIIAA. J Mol Biol 2004;340:941–
98:1978–1988. 956.
51. Dolinsky TJ, Nielsen JE, McCammon JA, Baker NA. PDB2PQR: an 72. Bilwes AM, Quezada CM, Croal LR, Crane BR, Simon MI. Nucleo-
automated pipeline for the setup of Poisson-Boltzmann electro- tide binding by the histidine kinase CheA. Nat Struct Biol 2001;8:
statics calculations. Nucleic Acids Res 2004;32:W665–W667. 353–360.
52. Halgren TA. Merck molecular force field. II. MMFF94 van der 73. Armstrong RN, Kondo H, Granot J, Kaiser ET, Mildvan AS. Mag-
Waals and electrostatic parameters for intermolecular interactions. netic resonance and kinetic studies of the manganese(II) ion and
J Comput Chem 1996;17:520–552. substrate complexes of the catalytic subunit of adenosine 30 ,50 -
53. Steinbeck C, Han Y, Kuhn S, Horlacher O, Luttmann E, Willighagen monophosphate dependent protein kinase from bovine heart. Bio-
E. The chemistry development kit (CDK): an open-source Java chemistry 1979;18:1230–1238.
library for chemo-and bioinformatics. J Chem Inf Comput Sci 74. Hirsch AK, Fischer FR, Diederich F. Phosphate recognition in struc-
2003;43:493–500. tural biology. Angew Chem Int Ed Engl 2007;46:338–352.
54. Guha R, Howard MT, Hutchison GR, Murray-Rust P, Rzepa H, 75. Smith PE, Tanner JJ. Conformations of nicotinamide adenine dinu-
Steinbeck C, Wegner J, Willighagen E. The Blue Obelisk-interoper- cleotide (NAD(1)) in various environments. J Mol Recognit 2000;
ability in chemical informatics. J Chem Inf Model 2006;46:991–998. 13:27–34.
55. DeLano WL. The PyMOL molecular graphics system. San Carlos, 76. Percy MJ, Crowley LJ, Boudreaux J, Barber MJ. Expression of
CA: DeLano Scientific; 2002. a novel P275L variant of NADH: cytochrome b(5) reductase
56. Fauchère JL, Quarendon P, Kaetterer L. Estimating and representing gives functional insight into the conserved motif important for
hydrophobicity potential. J Mol Graph 1988;6:203. pyridine nucleotide binding. Arch Biochem Biophys 2006;447:59–
57. Wang RX, Gao Y, Lai LH. Calculating partition coefficient by atom- 67.
additive method. Perspect Drug Discov Des 2000;19:47–66. 77. Mao L, Wang Y, Liu Y, Hu X. Molecular determinants for ATP-
58. Fauchère JL, Pliska V. Hydrophobic parameters-pi of amino-acid binding in proteins: a data mining and quantum chemical analysis.
side-chains from the partitioning of N-acetyl-amino-acid amides. J Mol Biol 2004;336:787–807.
Eur J Med Chem 1983;18:369–375. 78. Denessiouk KA, Rantanen VV, Johnson MS. Adenine recognition: a
59. Eisenberg D, McLachlan AD. Solvation energy in protein folding motif present in ATP-, CoA-, NAD-, NADP-, and FAD-dependent
and binding. Nature 1986;319:199–203. proteins. Proteins 2001;44:282–291.
60. Heiden W, Moeckel G, Brickmann J. A new approach to analysis 79. Schneider S, Marles-Wright J, Sharp KH, Paoli M. Diversity and
and display of local lipophilicity/hydrophilicity mapped on molecu- conservation of interactions for binding heme in b-type heme pro-
lar surfaces. J Comput-Aided Mol Des 1993;7:503–514. teins. Nat Prod Rep 2007;24:621–630.
61. Kelly MD, Mancera RL. A new method for estimating the impor- 80. Reedy CJ, Gibney BR. Heme protein assemblies. Chem Rev 2004;
tance of hydrophobic groups in the binding site of a protein. J Med 104:617–649.
Chem 2005;48:1069–1078. 81. Roberts SA, Montfort WR. Haem proteins. In: Encyclopedia of Life
62. Levitt M. A simplified representation of protein conformations for Sciences. John Wiley & Sons, Ltd: Chichester http://www.els.net/
rapid simulation of protein folding. J Mol Biol 1976;104:59–107. [doi: 10.1002/9780470015902.a0003054].
63. Brylinski M, Konieczny L, Roterman I. Ligation site in proteins rec- 82. Presta A, Weber-Main AM, Stankovich MT, Stuehr DJ. Comparative
ognized in silico. Bioinformation 2006;1:127–129. effects of substrates and pterin cofactor on the heme midpoint
64. Halgren TA. Representation of van der Waals (vdW) interactions in potential in inducible and neuronal nitric oxide synthases. J Am
molecular mechanics force-fields—potential form, combination Chem Soc 1998;120:9460–9465.
rules, and vdW parameters. J Am Chem Soc 1992;114:7827–7843. 83. Stuehr DJ. Enzymes of the L-arginine to nitric oxide pathway.
65. Kortemme T, Morozov AV, Baker D. An orientation-dependent J Nutr 2004;134:2748S–2751S.
hydrogen bonding potential improves prediction of specificity and 84. Caillet-Saguy C, Turano P, Piccioli M, Lukat-Rodgers GS, Czjzek M,
structure for proteins and protein-protein complexes. J Mol Biol Guigliarelli B, Izadi-Pruneyre N, Rodgers KR, Delepierre M,
2003;326:1239–1259. Lecroisey A. Deciphering the structural role of histidine 83 for
66. Thompson D, Simonson T. Molecular dynamics simulations show heme binding in hemophore HasA. J Biol Chem 2008;283:5960–
that bound Mg21 contributes to amino acid and aminoacyl adenyl- 5970.
ate binding specificity in aspartyl-tRNA synthetase through long 85. Bradley AL, Chobot SE, Arciero DM, Hooper AB, Elliott SJ. A
range electrostatic interactions. J Biol Chem 2006;281:23792–23803. distinctive electrocatalytic response from the cytochrome c
67. Gonzalez B, Pajares MA, Hermoso JA, Guillerm D, Guillerm G, peroxidase of Nitrosomonas europaea. J Biol Chem 2004;279:13297–
Sanz-Aparicio J. Crystal structures of methionine adenosyltransfer- 13300.

16 PROTEINS
Physicochemical Variation in Binding Pockets

86. Mao J, Hauser K, Gunner MR. How cytochromes with different 100. Tsuzuki S, Fujii A. Nature and physical origin of CH/pi interaction:
folds control heme redox potentials. Biochemistry 2003;42:9829– significant difference from conventional hydrogen bonds. Phys
9840. Chem Chem Phys 2008;10:2584–2594.
87. Gunner MR, Honig B. Electrostatic control of midpoint potentials 101. Tsuzuki S, Honda K, Uchimaru T, Mikami M, Tanabe K. The mag-
in the cytochrome subunit of the Rhodopseudomonas viridis reaction nitude of the CH/pi interaction between benzene and some model
center. Proc Natl Acad Sci USA 1991;88:9151–9155. hydrocarbons. J Am Chem Soc 2000;122:3746–3753.
88. Reedy CJ, Elvekrog MM, Gibney BR. Development of a heme pro- 102. Hunter CA, Lawson KR, Perkins J, Urch CJ. Aromatic interactions.
tein structure electrochemical function database. Nucleic Acids Res J Chem Soc Perkin Trans 2001;2:651–669.
2008;36:D307–D313. 103. Cauet E, Rooman M, Wintjens R, Lievin J, Biot C. Histidine-aro-
89. Andersen JF, Ding XD, Balfour C, Shokhireva TK, Champagne DE, matic interactions in proteins and protein-ligand complexes: quan-
Walker FA, Montfort WR. Kinetics and equilibria in ligand binding tum chemical study of X-ray and model structures. J Chem Theory
by nitrophorins 1–4: evidence for stabilization of a nitric oxide-fer- Comput 2005;1:472–483.
riheme complex through a ligand-induced conformational trap. 104. Levitt M, Perutz MF. Aromatic rings act as hydrogen bond accept-
Biochemistry 2000;39:10118–10131. ors. J Mol Biol 1988;201:751–754.
90. Choinowski T, Blodig W, Winterhalter KH, Piontek K. The crystal 105. Toth G, Murphy RF, Lovas S. Stabilization of local structures by pi-
structure of lignin peroxidase at 1.70 A resolution reveals a hydroxy CH and aromatic-backbone amide interactions involving prolyl and
group on the cbeta of tryptophan 171: a novel radical site formed aromatic residues. Protein Eng 2001;14:543–547.
during the redox cycle. J Mol Biol 1999;286:809–827. 106. Luo R, Gilson HS, Potter MJ, Gilson MK. The physical basis
91. Spence JT, Kipke CA, Enemark JH, Sunde RA. Stoichiometry of of nucleic acid base stacking in water. Biophys J 2001;80:140–148.
electron uptake and the effect of anions and pH on the molybde- 107. Marsili S, Chelli R, Schettino V, Procacci P. Thermodynamics of
num and heme reduction potentials of sulfite oxidase. Inorg Chem stacking interactions in proteins. Phys Chem Chem Phys 2008;10:
1991;30:3011–3015. 2673–2685.
92. Rich PR, Jeal AE, Madgwick SA, Moody AJ. Inhibitor effects on re- 108. Tsuzuki S, Honda K, Uchimaru T, Mikami M, Tanabe K. Origin of
dox-linked protonations of the B-hemes of the mitochondrial-Bc1 the attraction and directionality of the NH/pi interaction: compari-
complex. Biochim Biophys Acta 1990;1018:29–40. son with OH/pi and CH/pi interactions. J Am Chem Soc 2000;122:
93. Hallberg BM, Bergfors T, Backbro K, Pettersson G, Henriksson G, 11450–11458.
Divne C. A new scaffold for binding haem in the cytochrome do- 109. Gilson MK, Zhou HX. Calculation of protein-ligand binding affin-
main of the extracellular flavocytochrome cellobiose dehydrogenase. ities. Annu Rev Biophys Biomol Struct 2007;36:21–42.
Structure 2000;8:79–88. 110. Davis AM, Teague SJ. Hydrogen bonding, hydrophobic interactions,
94. Tsai AL, Kulmacz RJ, Wang JS, Wang Y, Vanwart HE, Palmer G. and failure of the rigid receptor hypothesis. Angew Chem Int Ed
Heme coordination of prostaglandin-H synthase. J Biol Chem 1993; 1999;38:736–749.
268:8554–8563. 111. James LC, Tawfik DS. Conformational diversity and protein evolu-
95. Lancaster CRD, Gross R, Haas A, Ritter M, Mantele W, Simon J, tion—a 60-year-old hypothesis revisited. Trends Biochem Sci 2003;
Kroger A. Essential role of Glu-C66 for menaquinol oxidation indi- 28:361–368.
cates transmembrane electrochemical potential generation by Woli- 112. James LC, Roversi P, Tawfik DS. Antibody multispecificity mediated
nella succinogenes fumarate reductase. Proc Natl Acad Sci USA by conformational diversity. Science 2003;299:1362–1367.
2000;97:13051–13056. 113. Schneider HJ. Ligand binding to nucleic acids and proteins: does se-
96. Makino R, Chiang R, Hager LP. Oxidation-reduction potential lectivity increase with strength? Eur J Med Chem 2008;43:2307–2315.
measurements on chloroperoxidase and its complexes. Biochemistry 114. Jain AN. Scoring functions for protein-ligand docking. Curr Protein
1976;15:4748–4754. Pept Sci 2006;7:407–420.
97. Haas AH, Lancaster CRD. Calculated coupling of transmembrane 115. Sousa SF, Fernandes PA, Ramos MJ. Protein-ligand docking: current
electron and proton transfer in dihemic quinol: fumarate reductase. status and future challenges. Proteins 2006;65:15–26.
Biophys J 2004;87:4298–4315. 116. Lyne PD, Lamb ML, Saeh JC. Accurate prediction of the relative
98. Kleywegt GJ, Harris MR, Zou JY, Taylor TC, Wahlby A, Jones TA. potencies of members of a series of kinase inhibitors using molecu-
The Uppsala electron-density server. Acta Crystallogr D Biol Crys- lar docking and MM-GBSA scoring. J Med Chem 2006;49:4805–
tallogr 2004;60:2240–2249. 4808.
99. Wallimann P, Marti T, Furer A, Diederich F. Steroids in molecular 117. Koehl P. Electrostatics calculations: latest methodological advances.
recognition. Chem Rev 1997;97:1567–1608. Curr Opin Struct Biol 2006;16:142–151.

PROTEINS 17

Вам также может понравиться