You are on page 1of 19

192 Current Topics in Medicinal Chemistry, 2011, 11, 192-210

Protein Flexibility and Ligand Recognition: Challenges for Molecular


Modeling

Francesca Spyrakis1,2,*, Axel BidonChanal3, Xavier Barril3,4 and F. Javier Luque3,*

1
Department of General and Inorganic Chemistry, University of Parma, Parma, Italy, 2INBB, Biostructures and Biosys-
tems National Institute, 3Departament de Fisicoquímica and Institut de Biomedicina (IBUB), Facultat de Farmàcia,
Universitat de Barcelona, 08028, Barcelona, Spain, 4ICREA

Abstract: The intrinsic dynamics of macromolecules is an essential property to relate the structure of biomolecular sys-
tems with their function in the cell. In the field of ligand-receptor recognition, numerous evidences have revealed the limi-
tations of the lock-and-key theory, and the need to elaborate models that take into account the inherent plasticity of bio-
molecules, such as the induced-fit model or the existence of an ensemble of pre-equilibrated conformations. Depending on
the nature of the target system, ligand binding can be associated with small local adjustments in side chains or even the
backbone to large-scale motions of structural fragments, domains or even subunits. Reproducing the inherent flexibility of
biomolecules has thus become one of the most challenging issues in molecular modeling and simulation studies, as it has
direct implications in our understanding of the structure-function relationships, but even in areas such as virtual screening
and structure-based drug discovery. Given the intrinsic limitation of conventional simulation tools, only events occurring
in short time scales can be reproduced at a high accuracy level through all-atom techniques such as Molecular Dynamics
simulations. However, larger structural rearrangements demand the use of enhanced sampling methods relying on modi-
fied descriptions of the biomolecular system or the potential surface. This review illustrates the crucial role that structural
plasticity plays in mediating ligand recognition through representative examples. In addition, it discusses some of the
most powerful computational tools developed to characterize the conformational flexibility in ligand-receptor complexes.
Keywords: Molecular recognition, conformational flexibility, ligand-receptor complex, induced-fit, molecular dynamics, drug
design.

INTRODUCTION Tel: +34934024557; Fax: +34934035987; E-mail; fjluque@ub.edu


ligands to a macromolecular target [7]. Though there is con-
Biological processes are essentially based on the ability sensus that the problem of generating reasonable ligand ori-
of biomolecules to recognize each other within a noisy envi- entations is quite satisfactory, recognizing near-native geo-
ronment, where many look-alike molecules are present. Ex- metries and predicting their affinities is still achieved with
amples are given, for instance, by antibodies recognizing limited success [8-11]. Besides the specific nature of the
antigenic peptides [1, 2], regulatory proteins binding specific interactions that mediate ligand binding, several factors ac-
DNA sequences [3, 4], or enzymes catalyzing and transform- count for the challenging question of developing a physi-
ing their substrate [5]. A binding event, like the formation of cally-based accurate scoring function. First, one cannot ne-
a protein-ligand complex, normally occurs in aqueous solu- glect the fundamental role played by water molecules, which
tion and, under thermodynamic equilibrium conditions, the rearrange around the ligand and the protein active site upon
standard free energy of binding determines the reaction ten- the complex formation. The inclusion of conserved water
dency. The Gibbs’s free energy is decomposable into enthal- molecules in the active site of proteins and the correct esti-
pic and entropic contributions. The former is phenome- mation of their energetic contribution is crucial for docking
nologically attributed to the formation of non-bonded inter- analysis. Thus, one should be able to simulate the insertion
actions such as salt bridges, hydrogen bonds, hydrophobic of a ligand into a solvated pocket, displace only irrelevant
contacts, cation-pi interactions and metal complexation. The waters and estimate the total binding free energy including
latter is given by the loss of conformational mobility of both the contribution of the retained ones. Numerous efforts have
the protein and the ligand, and by the release of water mole- been made to predict structural waters [12-16], improving
cules from hydrophobic surfaces to the solvent [6]. both binding poses and energetic predictions [16-18]. Never-
The ability of the scoring functions implemented in dock- theless, a general strategy for the solvation treatment in
ing programs to retain the delicate balance between enthalpy docking and virtual screening is still elusive. Another aspect
and entropy is critical for the prediction of the binding of to be considered is the proper assignment of the ionization
and tautomerization states of the ligands and of residues in
the active sites, which can be affected by the microenviron-
*Address correspondence to these authors at the Department of General and ment at the binding site [19]. At this point, it is worth noting
Inorganic Chemistry, University of Parma, Parma, Italy, INBB, Biostruc-
tures and Biosystems National Institute; Tel: +390521905669; that waters can buffer the active site changing the number of
Fax: +390521905556; E-mail; francesca.spyrakis@unipr.it hydrogen-bond donors and acceptors [20, 21]. Finally, one
Departament de Fisicoquímica and Institut de Biomedicina (IBUB), Facultat cannot forget that proteins are inherently flexible systems
de Farmàcia, Universitat de Barcelona, 08028, Barcelona, Spain; able to undergo functionally relevant conformational transi-

1568-0266/11 $58.00+.00 © 2011 Bentham Science Publishers Ltd.


Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 193

tions even in native state conditions [22-26], and that flexi- or even global rearrangements. Depending on the degree of
bility is often essential for the protein function [24]. flexibility, proteins could be basically classified into three
main categories, i) rigid proteins, showing only small side-
It is generally assumed that the recognition process is
chain adjustments; ii) flexible proteins, able to undergo
regulated by the interplay between the interaction energy
gained from the proper alignment of the ligand in the binding larger motions around hinge points or structural elements of
the binding site, and iii) intrinsically unstable proteins, which
site, and the elastic energy required to deform the interacting
are only stabilized by the entrance of a ligand into the bind-
molecules. Savir and Tlusty suggested that the conforma-
ing pocket.
tional changes experienced by both ligand and target could
arise from an evolutionary optimization of the binding proc- The strong belief that the protein conformational diver-
ess, since the target-ligand mismatch improves the recogni- sity represents an evolutionary and functional advantage
tion specificity in a noisy environment as a conformational pushed the lock-and-key model further away, and now an
proofreading mechanism [27]. Moreover, as reported by Mit- appropriate handling of the binding pocket flexibility, or
tag et al. [28], the conformational disorder may provide an more generally of the protein dynamics, is fundamental to
electrostatic and steric advantage due to the generation of shed light into the protein-ligand binding. Unfortunately,
fluctuating electrostatic fields, allowing disordered proteins predicting which conformation a flexible target will adopt
to bind several surfaces of the partners by altering the buried upon the binding of a specific molecule is still a challenging
surface area. In this context, the conformational variability question and, whether the occurring conformational changes
increases plasticity and malleability, since the same molecule have to be attributed to the ligand entry (i.e, induced-fit
can interact with different partners, and provides an evolu- model), or depend on the intrinsic structural and dynamical
tionary advantage, facilitating alternative splicing, domain properties of the protein is a subject of current interest.
shuffling, protein modularity and mutations.
This dilemma was first addressed in the late fifties by
Knowledge of the intrinsic dynamics of macromolecules Koshland, who proposed the induced-fit theory, where the
has direct implications in our understanding of the relation- ligand plays a key role in inducing the target conformational
ship between structure and function, and particularly in areas change [33], and later by Monod, who suggested the exis-
such as virtual screening and structure-based drug discovery. tence of a preexisting ensemble of conformations, where the
Numerous evidences have revealed the limitations of the ligand only selects the most suitable one and shifts the con-
lock-and-key theory and the need to elaborate more sophisti- formational equilibrium [34]. According to the induced-fit
cated models that account for the plasticity of biomolecules. model, the ligand binding appreciably reduces the conforma-
In this context, this work pursues to illustrate the implica- tional entropy of the target bringing an energetic penalty to
tions between ligand binding and target flexibility. A series the overall binding process (Fig. (1)). On the contrary, in the
of representative examples, involving either small local ad- model of equilibrium conformational states, the ligand se-
justments in specific residues or even large-scale motions of lects and stabilizes one of the available conformations, with-
structural fragments, will be discussed. Moreover, the broad out significant reduction in the conformational entropy but
range of structural rearrangements that might mediate ligand favorably contributing to the binding free energy (Fig. (1))
binding raises challenging issues for the capability of model- [35]. The induced-fit model could easily account for local
ing techniques to capture the conformational flexibility of motions as side-chain reorientations or transitions between
biomolecules. Given the intrinsic limitation of conventional isomeric states. Nevertheless, cooperative changes or rear-
modeling tools, only events occurring in short time scales rangements of entire domains can be hardly attributed to the
can be reproduced by all-atom simulation methods, and entrance of small ligands into a slightly flexible binding
larger structural rearrangements demand the use of enhanced pocket.
sampling techniques. This review also discusses some of the
Experimental and computational studies have strongly
most powerful computational tools developed to characterize
supported the view of an existing equilibrium of conforma-
the role of structural flexibility in the formation of ligand-
tional substates in native conditions [36, 37], as originally
receptor complexes.
reported by Monod, Wyman and Changeux [22, 30, 38-42],
relating the different conformational states to an intrinsic
PROTEIN FLEXIBILITY AND LIGAND BINDING predisposition and dynamics of proteins. This approach
The energy landscape of a protein is usually displayed as would also agree with the view of a dynamic energy land-
a folding funnel, lined by unfavorable states collapsing scape in which the protein’s “open” form corresponds to the
through various routes into few favorable folded states, made lowest energy minimum, while the “closed” form is trapped
by sub-ensembles of structurally and energetically equivalent in a less stable minimum [43-45]. As an example (see be-
conformations of the protein [14, 23, 29-32]. The use of a low), Xu et al. have suggested that binding of ligands to ace-
single or even few structures (typically provided by X-ray tylcholinesterase (AChE) likely affects the pre-existing equi-
crystallography or NMR) neglects or poorly represents the librium dynamics selecting the most suitable conformation
existence of these conformational substates, which are dy- rather than promoting the formation of non-native structures
namically interchanging depending on “external” conditions [46].
such as ionic strength, pH, temperature and the presence of More recently, Okazaki and Takada [47], and Sullivan
ligands. Generally the highest flexibility is related to con- and Holyoak [48] sustained that protein flexibility could be
formational rearrangements in loops, but there is increasing better explained by joining both intrinsic dynamics and in-
evidence about the occurrence of larger structural modifica- duced-fit theories, since the selection of a pre-existing
tions in the structural elements that delineate the binding site
194 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

lecular dynamics (MD) simulations to post-process rigid


docking solutions [68]. The availability of a large number of
structures a priori represents the best strategy to dock
ligands into the most suitable and reliable conformation,
while predicting protein-ligand interactions by modeling a
single receptor structure would only be useful for ligands
targeting a single particular narrow state of the conforma-
tional ensemble. Nevertheless one should be aware that -due
to current methodological limitations- ensemble docking
does not necessarily translate into better predictions and can,
in fact, degrade the results [69, 70]. Therefore, a judicious
choice of the target conformations seems necessary for a
successful identification of new hits in virtual screening stu-
dies.

LOCAL STRUCTURAL CHANGES INDUCED UPON


BINDING
Most protein side chains undergo small structural ad-
justments during ligand binding. Thus, the inspection of the
side chains motions promoted by the ligand in 63 complexes
involving 20 proteins supports the assumption that protein
side chains move as little as necessary in order to achieve a
collision-free complex [59]. In fact, Zavodszky and Kuhn
reported that ligand binding did not typically involve
Fig. (1). Modification of the protein energy landscape upon ligand changes between rotamers, but mostly small (<15 degrees)
binding in the models of (a) induced fit and (b) conformational changes in the torsional angles. In turn, these findings sug-
states. In the induced-fit model the ligand promotes a structural gest that docking into the apo forms of their binding targets
rearrangement in the binding site of the receptor, which reduces the should a priori lead to satisfactory predictions about the best
conformational entropy. In the model of conformational states the poses of ligands when combined with simple energy minimi-
ligand selects and stabilizes one of the pre-existing conformations, zation or steric optimization procedures.
shifting the conformational equilibrium towards the bound state. In spite of the preceding comments, local adjustment of
the side chains of residues have been noticed in certain cases,
conformation could be reasonably followed by additional such as the low-molecular-weight protein tyrosine phospha-
rearrangements induced by a specific molecule [49]. In tase (LMW-PTP), where Trp49 is responsible for the sub-
particular, to shed more light on the dualism between strate specificity but also for the conformational exchange
induced-fit and population-shift mechanisms, Okazaki and quenching associated with the monomer-dimer transition
Takada developed and applied an implicit ligand binding [71], or in the copper-binding proteins superoxide dismutase
model combined with a double basin Hamiltonian, finding [72] and azurin [73], where a single histidine regulates the
that both models have their own range of applicability. Thus, accessibility of the copper-binding site. In these cases, the
strong and long-range interactions favor essentially an inclusion of local flexibility would be mandatory for the suc-
induced-fit mechanism, as in the case of protein-protein or cessful prediction of the ligand pose. Thus, binding sites can
protein-DNA interactions [50-52], whereas weak and short- be seen as a combination of preorganized regions, whose
range contacts better support the population-shift model, as structure is well preserved due to a network of contacts with
observed for antigen-antibody binding and substrate binding neighboring residues, and other regions exhibiting a larger
to enzymes [53-55]. The real situation likely resembles both structural plasticity likely due to a limited involvement in
mechanisms, since the ligands might first select a partially intramolecular contacts or to a large exposure to the solvent
closed conformation, and then induce conformational [21].
changes up to the closed state [47].
Acetylcholinesterase, an enzyme that catalyzes the hy-
While these issues deserve further clarifications, drug drolysis of the neurotransmitter acetylcholine, is well suited
design approaches have to somehow predict and simulate the to illustrate the subtle balance between structurally preserved
intrinsic flexibility of proteins, trying to find the best com- and flexible regions in the binding site. The 3D structure of
promise between accuracy, reliability and computational Torpedo californica AChE (initially solved at 2.8 Å resolu-
resources. This explains the intense effort paid to the devel- tion [74], and later refined to 2.5 Å resolution [75]) shows a
opment of docking algorithms dealing with flexible binding deep and narrow gorge (around 20 Å long) that penetrates
sites [24]. Simple and less time-consuming methods include halfway into the enzyme and widens out close to its base.
only binding pocket side-chain flexibility [56-59], while Fourteen highly conserved aromatic residues line a substan-
more complex approaches use constrained geometric simula- tial portion of the surface of the gorge. The catalytic active
tions [60], elasticity network theory [61, 62], docking into site is located at the bottom of the gorge, and the pi-electron
ensembles of conformations derived from experimental or distribution of Trp84 assists binding of the quaternary moi-
computational techniques [17, 63-67], or even use of mo- ety of choline through cation-pi interactions. In addition,
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 195

another triptophane residue (Trp279) at the entrance of the


gorge defines a peripheral site, which might act as an initial
binding site for substrate entry.
The structural complexity of the active site gorge ac-
counts for the large diversity of AChE inhibitors, which ex-
hibit notable differences in the mechanism of inhibition and
selectivity [76, 77]. In spite of such chemical diversity, the
main structural features of the residues at the catalytic and
peripheral binding sites are generally preserved in the differ-
ent AChE complexes with reversible inhibitors, except for
Phe330 at the catalytic site and Trp279 at the peripheral one.
Thus, these residues adopt drastic conformational changes
for different inhibitors (Fig. (2)). In the catalytic pocket, both
tacrine [78] and huprine X [79] bind the enzyme through
stacking of the aminoacridine ring with the aromatic rings of
Trp84 and Phe330. In particular, the side chain of this latter
residue is characterized by a N-C-C-C (1) torsional angle
of ~160°. The interaction of ()-huperzine A [75] involves a
cation-pi interaction between the amino group of the inhibi-
tor with the aromatic rings of Trp84 and Phe330. In this
case, the Phe330 side chain adopts a conformation defined
by 1 ~ -170°, which thus mimics the conformation found for
the AChE complex with edrophonium [78]. Finally, donepe-
zil extends from the catalytic site to the peripheral one along
the gorge, and its benzyl ring interacts through stacking in-
teraction with Trp84, while the piperidine ring lies onto the
benzene moiety of Phe330, whose side chain shows a tor-
sional angle of ~ -130° [80]. A similar orientation is found
for the binding of decamethonium [78].
The structural features of the binding of AChE inhibitors
at the peripheral binding site are more delicate due to the
conformational flexibility of Trp279, as noted in the X-ray
crystallographic structures available for several peripheral
site ligands [81] and dual binding site inhibitors [82-85]. In
fact, three main arrangements of the indole ring of Trp279
can be identified upon inspection of the X-ray crystallo-
graphic structures (Fig. (2)). The first orientation is charac-
terized by dihedral angles N-C-C-C3 (1) and C-C-C3-C3a
(2) close to -60° and -80°, respectively, as found in the apo
form of the enzyme [74]. Similar values are observed in
complexes with catalytic binding site inhibitors such as
huprine X [79] or (-)-huperzine A [75], and with peripheral
binding site inhibitors such as propidium, decidium and gal-
lamine [81]. Moreover, this arrangement is also found in
complexes with dual binding site inhibitors, such as de-
camethonium [78], donepezil [80], tacrine(10)-hupyridone Fig. (2). Representation of selected residues in both catalytic and
[84], and anti-TZ2PA6 [82], as well as in the complex with peripheral sites of acetylcholinesterase (AChE). Top: residues
fasciculin [86]. Dihedral angles close to -120° (1) and +50° Phe330 and Trp84 at the catalytic binding site for AChE complexes
(2) are found in the complexes with tacrine(8)-4-amino- with tacrine, ()-huperzine A and donepezil. Bottom: residue
quinoline [85] and NF595 [83]. Finally, an alternative orien- Trp279 in the peripheral site for the AChE complexes with donepe-
tation defined by dihedral angles close to -160° (1) and - zil, bis(7)-tacrine and syn-TZ2PA6.
120° (2) is found in the complex with syn-TZ2PA6 [82].
As a final remark, it is worth noting how an apparently -121.4°, 2: +43.7°). As noted by Xu et al. [46], the confor-
minor chemical change, such as the different length of the mational plasticity observed for Trp279 should not be con-
methylenic tether in bis(5)-tacrine and bis(7)-tacrine, leads ceived as the result of an induced-fit mechanism, but rather
to a different arrangement of the indole ring of Trp279 [85], as the preferential binding of the inhibitor to pre-existing
as the former binds to the peripheral site by imposing an conformational states, as those conformations are accessed in
apo-like conformation for Trp279 (1: -76.3°, 2: -87.0°), MD simulations of the apo form of the enzyme [46]. In this
whereas this residue is completely reoriented in the latter (1: context, rational drug design could largely benefit from the
identification of regions with a high intrinsic conformational
196 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

plasticity in the binding site, facilitating the right choice of


conformational states for successful docking studies [87].

SMALL-AMPLITUDE BACKBONE MOVEMENTS


Small-molecule binding sites in proteins usually tend to
be more rigid than the rest of the protein. For instance, in the
case of enzymes, changes in the backbone skeleton are usu-
ally not significant and the functional atoms move generally
less than 1 Å upon ligand binding [88]. Nevertheless, some
enzymes and receptors, as cytochromes or the liver X recep-
tors [89], have adaptable binding sites, which enable them to
recognize a wider range of ligands. Though large conforma-
tional changes in the backbone are relatively rare, the struc-
tural features of the binding site may be completely trans-
formed when they occur. Moreover, these changes mainly
affect regions in the binding site where flexibility plays a
functional role, as will be illustrated here for different pro-
teins.

Heat-Shock Protein 90
The oncology target Hsp90, and in particular its ATP
binding site, has attracted much interest in recent years.
Early structures of the N-terminus domain already identified Fig. (3). Superimposition of open, closed and helical conformations
a loop that could adopt two different conformations (open of Hsp90. In the open conformation, the binding site becomes con-
and closed) depending on the nature of the bound ligand siderably larger due to the opening of a hydrophobic pocket (black
(Fig. (3)) [90]. Later, the inhibitor PU3 -initially thought to surface). Upon dimerization, this loop (shown in the center, above
bind the open conformation [91]- revealed the existence of a the black surface) undergoes a major conformational rearrange-
third conformation, known as helical. While the open-closed ment, closing onto the binding site.
transition affects the periphery of the binding site, the helical
conformation results in the creation of a large lipophilic approximately 2.5% in humans [94]). More recently, a small
pocket at the bottom of the cavity, which changes substan- molecule of unrelated structure (identified by means of high-
tially the volume and the physicochemical properties of the throughput screening) was shown to bind in a newly created
binding site [92]. Interestingly, the crystallographic structure cavity. The ligand inserts itself between residues Trp40 and
of the full-length dimer has shown that the flexible region is Phe114, causing a partial melting of two -strands that form
part of a larger loop that undergoes a major conformational a lid over the binding site [95]. Molecules exploiting this
transition (about 40 Å) as part of the chaperoning cycle [93]. cryptic site can achieve better ligand efficiency and pharma-
This suggests that the flexibility of this loop, which is a func- cokinetic properties than most peptidomimetics [96].
tional requirement, may be exploited by non-natural ligands
to reshape the binding site and obtain better complementar- Kinase DFG-Out Binders
ity. Unfortunately the knowledge of where conformational
changes may take place cannot be used (at present) to predict A common mechanism of control of kinases is the phos-
which molecule will induce what change, and it is down to phorylation of the so-called activation loop (AL). Recogni-
serendipity that ligands are discovered (e.g. random screen- tion of the AL by an upstream kinase is thought to involve a
ing, naïf design) and that the conformational change is eluci- partial unfolding, and its regulatory function is also likely to
dated (as it may be incompatible with crystal formation). involve some conformational diversity. The flexibility of this
This system also illustrates how experimentally observed protein segment can also be attested by the lack of electron
conformations may not be treated as equal. Thus, due to its density and, therefore, missing residues in some X-ray struc-
larger volume and lipophilicity, the helical conformation tures. This flexibility has been exploited by several kinase
offers many more binding opportunities but is energetically inhibitors that bind to a hydrophobic pocket created by the
penalized and, unless the conformational strain is taken into displacement of a phenylalanine in the conserved DFG se-
account, docking results may degrade [69]. quence in the AL. DFG-out ligands were initially reported
for p38 [97], and they are often referred to as allosteric in-
Renin hibitors or indirect competitors because, although the inhibi-
tor binds in the same pocket as ATP, the substrate is com-
This protease, which converts angiotensinogen into an- peted off by the conformational change induced by the
giotensin I, has been studied as an antihypertensive target for ligand rather than by the ligand itself. As there is more se-
many years. The structure of the protein and its complex quence diversity in the pocket created by DFG-out binders,
with the peptidic substrate was used to design peptidomimet- these inhibitors can confer increased selectivity, a property
ics, often of high molecular weight and suboptimal pharma- of critical importance in the kinase field. As the displace-
cokinetics (e.g. the oral bioavailability of Aliskiren is ment of the AL happens at relatively long timescales, an-
other property of DFG-out ligands is slower association and
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 197

dissociation rates, which may affect the pharmacokinetics of


the compounds [98].

Protein Tyrosine Phosphatase 1B


PTP1B is a drug target for the treatment of type II diabe-
tes and obesity that has proven to be extremely challenging.
Many inhibitors acting on the phosphotyrosine binding site
have been described, but potency is heavily dependent on the
presence of a negative charge, which greatly damages the
pharmacokinetics properties [99]. In this context, the discov-
ery of allosteric inhibitors opened a new window of opportu-
nity [100]. Previously, two distinct conformations had been
described for the so-called WPD loop (residues 179-184),
which lines the binding site. In the apo form, it adopts an
open conformation, whereas substrate binding induces a
closing of the loop, thus reducing the size of the cavity, now
tightly fitting around the phosphotyrosine [101]. Interest-
ingly, this conformational change is coupled to a larger am-
plitude transition in the 7-helix (residues 287-295), which
is located about 20 Å away (Fig. (4)). In the WPD-closed
conformation this helix is ordered and in contact with the
3-helix, but in the WPD-closed conformation it is disor-
dered and separated from the rest of the protein. The allos-
teric inhibitors bind to the latter form, occupying a hydro-
phobic pocket that appears upon displacement of Trp291
(part of the 7-helix). It would seem then that these ligands
stabilize a conformation able to preclude the closing of the
WPD-loop. In addition to providing a completely different
chemotype, in this case the allosteric inhibitors bind in an
Fig. (4). Superposition of active (light grey) and inactive (dark
area that is poorly conserved amongst phosphatases, making grey) conformations of PTP-1B. In the active form, the active site
these compounds highly selective for PTP1B [100]. Very
loop is closed and 7-helix (in the top of the figure) is packed
recently, the WPD-open conformation (apo form) has been
against the rest of the protein. In the inactive form, the loop is open
exploited to identify non-competitive inhibitors. Although
(shown in the center), leaving a large and shallow binding site.
they bind to the phosphotyrosine binding site rather than
Allosteric inhibitors compete with the 7-helix for binding, thereby
competing with the substrate, they simply stabilize the inac-
stabilizing the inactive conformation.
tive conformation, reducing the concentration of the catalyti-
cally competent enzyme. Unlike most direct inhibitors, these
molecules do not bear a negative charge and can thus cross movements can be identified: i) motions involving fragments
membranes, achieving cellular activity [102]. of the protein chains, ii) hinge-bending motions involving
protein domains, and iii) hinge-bending motions between
Overall, even though the rigidity of small-molecule bind- covalently unconnected subunits. Generally these move-
ing sites has certainly been one of the factors contributing to ments are not attributed to an induced-fit mechanism, but as
the success of structure-based drug design, molecules ex- resulting from the conformational variability experienced by
ploiting other mechanisms of action -such as protein-protein proteins. Accordingly, the transition from an “open” to a
inhibitors or allosteric modulators- tend to bind in areas that “closed” conformation is not forced by the incoming ligand,
are much more flexible, thus adding another layer of uncer- but by an equilibrium shift towards an already available con-
tainty and complexity to rational design [103, 104]. Consid- formation. Interestingly, the physical principles that regulate
ering our increased knowledge about biological systems and the shifting of the conformational equilibrium at the bottom
the benefits provided by molecules acting on non-typical of a binding funnel should be comparable to the principles
binding sites (e.g. expansion in the number of druggable driving the identification of the native structures in folding
proteins, increased selectivity, potential to activate as well as funnels [107].
inhibit) these type of effectors –as well as protein flexibility–
are expected to play an ever-increasing role in drug design Relevant conformational adjustments have been reported
[105, 106]. in a variety of cases, such as the acyl carrier protein [108] or
the glutamate receptor [109], where the ligand binding trans-
LARGE-SCALE PROTEIN MOTIONS forms a local adjustment into a global allosteric effect. Other
systems where large-scale conformational transitions take
Many proteins experience significant conformational place are the NS3 RNA helicase from hepatitis C virus
transitions through hinge bending motions in order to adopt [110], the protein kinase Zap70 involved in T-cell activation
the active state and perform their function. As reported by [111], the DNA-binding domain of PU1 [112], calsensin
Kumar et al. in an outstanding review [107], three type of [113], the ATP-binding cassette transporters [114], the het-
erodimeric transcription factor CBF [115], the insulin-like
198 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

growth factor binding protein [116], the C-terminal zinc fin-


gers of human MTF-1 [117], the HIV-1 nucleocapsid NCp7
[118], or the CTP:glycerol-3-phosphate cytidylyl-transferase
dimeric enzyme [119], as reported in detail by Valente et al.
[35]. Few selected cases will be examined in more detail in
the following.

Kinase Proteins
Proteins belonging to the kinase family represent a nota-
ble example of systems undergoing large-scale movements.
These proteins share a well preserved catalytic core for the
transfer of the -phosphate of ATP to serine, threonine or
tyrosine in different target enzymes [120]. This activity is
regulated at different levels, such as phosphorylation by
other kinases, membrane and organelle localization via scaf-
folding proteins, protein-protein or domain-domain interac-
tions through regulator modules [121]. Protein kinases could
be viewed as molecular switches that can adopt two extreme
states: an “on” (active) conformation and an “off” (inactive or
minimally active) one. While all kinases show structurally
very similar active conformations, many differences have
been observed in the off states, where the formation of a cata-
lytically active pocket is not required [120]. From a struc-
tural viewpoint, these proteins contain two distinct lobes (the
smaller N lobe and the larger C lobe), forming a deep cleft at
the interface where ATP is bound (Fig. (5)). A highly con-
served phosphate loop, also known as the P loop, contains a
flexible Gly-rich sequence, which can thus closely contact
the phosphates and easily bind small inhibitors in the ab-
sence of ATP. The AL occupies a more central position and
needs to be phosphorylated to allow the protein adopting the
active conformation. Since a proper regulation of kinases is
crucial for cellular growth and development, misregulation
can lead to cell transformation or even cancer. Therefore,
understanding the conformational flexibility of these struc-
tural elements is fundamental for both a regulatory and a
drug therapy perspective.
Hyeon and co-workers computationally investigated the
dynamics of nucleotide binding in the catalytic domain of
protein kinase A (PKA) [121], as previously done by differ-
ent groups for other kinases as adenylate kinase (AK) [53,
122, 123], c-Src [124, 125], Abl [126], and cyclin-dependent Fig. (5). Superposition of the open (inactive; dark grey) and closed
kinase 5 [127]. PKA is activated by the phosphorylation of a (catalytically active; light grey) forms of (top) Protein Kinase A and
threonine in the AL and repressed by the binding of a regula- (bottom) Adenylate Kinase. ATP is shown in space fill cartoon. The
tory subunit. Its catalytic domain has been crystallized in key regions are labeled and indicated by black arrows.
diverse conformations that reveal differences in both ATP
and substrate-binding lobes, and significant movements in its binding site, the native contacts are temporarily removed
many secondary structure elements [128-130]. Activation of to accommodate the incoming substrate. Some of the secon-
PKA involves an open-close conformational transition in- dary motifs partially unfold to subsequently refold in the so-
duced by ATP binding. Interestingly, the major changes are called cracking mechanism (occurring in a 1-5 ms time
found in regions distant from the binding pocket. The com- range), which implies a non-monotonic change in flexibility
putational analyses showed that PKA maintains an open con- along the transition route. In particular, local unfoldings are
formation until ATP enters the catalytic domain. This event exploited to relieve specific areas of high stress during con-
leads to a clockwise rotation of the C-terminal tail that al- formational deformations and to overcome high energy bar-
lows Phe327 to contact ATP and enhances the compaction riers [131].
between small and large lobes of the catalytic domain. These
local dynamical effects occur at faster rate than the global An unfolding (cracking) model was also proposed for
dynamics, suggesting that the allosteric signal propagates AK [123], though an innovative approach has been sug-
from the binding pocket to the exterior sites through a net- gested by Hanson and co-workers [54]. This enzyme cata-
work of native contact pairs. When the nucleotide occupies lyzes the conversion of ATP and AMP in two ADP mole-
cules. It can adopt distinct (open and closed) limit conforma-
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 199

tions. The large hinge movement assisting the transition be- NMR and MD simulations studies demonstrated that the
tween open and closed forms is critical for function [132], cCaM has greater Ca2+ affinity and is more open than nCaM
since the binding of ATP and AMP induces the closing mo- when bound to Ca2+ [136-138]. It was also suggested that
tion of both the ATP lid domain and the nucleoside mono- cCaM could be partially unfolded in native conditions [139,
phosphate binding (NMP bind) domain (Fig. (5)), which thus 140], and that the conformational exchange may imply a
keeps water molecules far from the catalytic pocket during more complicated mechanism than a simple two-state (open-
the phosphoryl transfer reaction. The authors observed that closed) process, as suggested by the analysis performed by
the enzyme experiences a dynamical equilibrium between Tripathi and Portman using a coarse-grained variational
open and closed forms, with a preference towards the closed model [133]. While the transition from the closed to the open
state, even in the absence of substrates, which appears to be form of the more flexible nCaM domain does not require a
unnecessary to induce the formation of the active-site cavity. cracking event, the same shifting in the cCaM domain only
In turn, this large-scale conformational motion seems to be occurs via a transient unfolding in the helix linker between
an intrinsic feature of the protein. The natural flexibility and the binding loops. Similar processes also occur upon binding
the major stability of the closed state undoubtedly help in of CaM with its biological peptidic substrates (Fig. (6);
minimizing the energy required to shift the equilibrium to- [141]). The large-scale transition leading from the holo CaM
wards the active configuration characterized by a higher form to the complexed structure occurs via a hinge-based
catalytic rate. Accordingly, ligand binding does not block the motion, involving the breaking of the long helix-shaped
enzyme in a single conformation, but shifts the ATP lid do- linker, which partly unfolds around the hinge and generates
main distribution doubling its closing rate, and restricts the two different helices. This high flexibility represents an in-
range of conformational changes. Once the reaction has teresting strategy for enabling the binding of a number of
taken place, the products can be released following the lid peptides with different sequences, and is the expression of a
opening, which indeed represents the limiting step. This en- complex energy funnel characterized by a large number of
zyme thus appears to be designed to exploit large-amplitude isoenergetic minima around the funnel bottom [107].
conformational dynamics to achieve the fast catalytic turn-
over. Glutamate Dehydrogenase
A significant hinge motion has also been observed for
Calmodulin
glutamate dehydrogenase (GDH), which catalyzes the re-
CaM is a calcium binding protein able to bind and regu- versible oxidative deamination of L-glutamate into 2-
late a number of different targets as kinases, phosphatases, oxoglutamate and ammonia using NAD + as a cofactor. GDH
NO synthase, Ca2+ pumps, proteins active in motility and is formed by two trimers stacked on top of each other, whose
many others. Thus, it can participate in multiple biological subunits are, at least, composed of three domains (Fig. (7)).
processes, such as inflammation, metabolism, apoptosis, The NAD + binding domain is located in the upper part of the
muscle contraction, etc. Calmodulin consists of two structur- subunits, overhung by a long protrusion known as antenna.
ally similar N- (nCaM) and C-terminal (cCaM) globular do- This last part likely plays a relevant role in regulating the
mains bonded by a flexible interdomain linker. The two do- enzyme, allosterically controlled by numerous ligands and
mains display quite different properties, in particular in cofactors as GTP, ATP, ADP, or by hydrophobic compounds
terms of flexibility and Ca2+ affinity [133]. The Ca2+ binding as palmitoyl-CoA and steroid hormones ([142] and refer-
to the apo/closed form induces significant conformational ences therein). GDH undergoes significant conformational
changes leading to the holo/open form (Fig. (6); [134, 135]). rearrangements during each catalytic cycle (Fig. (7)). When
the NADH enters and binds into the coenzyme binding

Fig. (6). Different conformations of calmodulin: (a) apo structure, (b) holo structure (Ca2+ ions are represented by spheres), and (c) structure
of a calmodulin-peptide complex.
200 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

Fig. (7). Top: Superposition of the open (dark grey) and closed (light grey) forms of Glutamate Dehydrogenase. Bottom, left: Inset of the
open apo conformation. ATP is represented in space fill cartoons. Bottom, right: Inset of the closed catalytically active form. The GTP in-
hibitor, the NADH coenzyme and the substrate (glutamate; shown in dark grey) are represented in space fill cartoons.

pocket, the coenzyme domain experiences a 18° rotation


with respect to the other domain, via a hinge bending move- COMPUTATIONAL STUDIES AND MODELING OF
ment, to close down both the coenzyme and the natural sub- CONFORMATIONAL FLEXIBILITY
strate (i.e. glutamate). At the same time, the ascending helix Much of our knowledge on the structural features of
of the antenna moves towards the pivot helix of the adjacent biomolecular targets comes from high-resolution experimen-
subunit, which performs a counterclockwise rotation along tal techniques. Structural models obtained from X-ray crys-
the helical axis. These movements induce a compression of tallography are usually static, time- and space-average struc-
the entire hexamer, since the upper three pairs of subunits tures solved at extremely low temperatures. Therefore, they
move as a unique rigid entity towards the lower subunits, provide information on one or at most few structural con-
compressing the cavity at the interface. formations of the proteins, but they cannot account for a
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 201

comprehensive understanding of the inherent protein flexi- tions can be used not only for the refinement of docked
bility. Nevertheless, in the last few years technical advances complexes, but also during the preparation of the protein
such as time-resolved measurement on single crystals have receptor before docking in order to optimize its structure and
demonstrated to be extremely valuable to shed light into dy- account for protein flexibility [151]. As examples, we quote
namical aspects of fast processes, such as the breathing mo- two independent studies on the docking of small ligands,
tions of cavities coupled to ligand migration [143]. NMR which report a receptor-based pharmacophore method that
techniques provide ensembles of low-energy conformations, relies on a collection of multiple protein structures to ac-
and are particularly useful for proteins or polypeptides diffi- count for protein flexibility in HIV-1 protease [152], and to
cult to crystallyze or that provide too disordered crystals the molecular mechanism of resistance of the mutant
[144]. Moreover, it has the large advantage of being per- T66I/M154I HIV-1 integrase to the inhibitor L-731,988
formed under conditions and in solutions that mimic the bio- [153]. In particular, this strategy can be relevant for studies
logical environment. Unfortunately, the intrinsic experimen- dealing with membrane-bound proteins (i.e., G-protein cou-
tal limitations of NMR make this technique applicable to a pled receptors), which have proven very difficult to crystal-
limited set of targets. Even if the equilibrium between ener- lize for examination by X-ray crystallography, and are often
getically close conformational states may be investigated, too insoluble for NMR analysis [154, 155]. Nevertheless, the
larger movements can be hardly predicted. success of MD simulations coupled to docking will depend
MD simulations represent an alternative strategy to ex- on the structural diversity of the sampled conformations,
which in turn will be an issue of particular relevance in those
plore the inherent flexibility of proteins and the impact of
proteins with a high degree of conformational flexibility and
structural fluctuations on the binding of ligands or in protein-
when docking involves compounds with diverse chemical
protein interactions [145]. In essence, classical MD simula-
scaffolds.
tions rely on the iterative numerical calculation of instanta-
neous forces acting on the atoms of a system, which in turn The suitability of atomistic MD techniques to character-
consists of particles that move in response to their interac- ize slower motions such as bending of domains through
tions according to the equations of motion defined in classi- hinge regions or allosterism is more questionable, as the time
cal mechanics. MD can a priori be routinely applied to in- scales of those conformational changes are not accessible via
vestigate a wide range of dynamic processes, and coupled to conventional MD techniques. To address this issue one has
statistical mechanics permits to derive thermodynamic prop- to resort to enhanced sampling techniques [156-161]. One
erties of the system. Accordingly, one can predict, for in- possibility is the use of simplified representations of the
stance, changes in the binding free energy of a ligand or the biomolecular system that enhance sampling of certain de-
mechanisms and energetic consequences of conformational grees of freedom at the expense of others (typically those
alterations in proteins. Nevertheless, applications are largely related to the fastest motions). The goal is to enhance sam-
limited in practice by the time scale of the dynamical proc- pling of the “slow” motions via simplified descriptions
esses, which in biomolecular systems can range from femto- through integration of the “fast” degrees of freedom into a
seconds to hours. few (i.e., coarse graining) ones [162]. To this end, the system
is modeled as a collection of beads. As the number of beads
Conventional atomistic MD simulations are well suited to
decreases, the simulation is less expensive and the system
study local elastic vibrations of atoms/groups of atoms and
that can be simulated is larger. However, developing an ac-
rotations of side chains, and can therefore assist in identify-
curate and transferable parametrized force field capable of
ing small structural rearrangements triggered upon ligand
describing the general dynamics of systems becomes in-
binding. For instance, MD simulations of AChE complexed
with syn-TZ2PA6 [146] revealed a change in the side chain creasingly difficult as the graining is ‘coarser’. Coarse-
grained MD approaches, where some of the fine atomistic
of Trp286 (numbering in mouse AChE; this residue is
details of the system are smoothed over or averaged out, are
equivalent to Trp279 in Torpedo californica AChE) from the
promising to analyze large-scale structural changes, such as
orientation found in the apo form of the enzyme to that seen
reshaping of binding pockets. Nevertheless, while these
in the X-ray crystallographic complex with the inhibitor
techniques can be informative about the general trends of
[82]. This change, which took place in a 2 ns window, makes
the Trp286 side chain shooting off the hydrophobic core and protein dynamics, the absence of atomistic details limits se-
riously the impact of the sampled conformations for docking
adopting another conformation, which permits a stable stack-
experiments. Nevertheless, interesting hybrid methods com-
ing with the inhibitor. Moreover, this change was accompa-
bining coarse-grained models of proteins with an atomistic
nied by other major changes in His287, Trp86, Tyr133,
description of the active site have been recently developed,
Tyr337 and Phe338. A similar structural rearrangement of
to enhance the potential of these approaches in examining
the tryptophan residue at the peripheral site has also been
reported in an independent study of the AChE complex with molecular recognition processes [163].
indole-tacrine hybrids, which were designed as dual binding- Enhanced sampling can also be achieved through tech-
site inhibitors [147]. niques that rely on the modification of the conventional MD
sampling. In this case, the system is forced to explore the
MD simulations provide an efficient way to generate
conformational space by facilitating the escape from local
unbiased structures to be used in docking or virtual screening
analyses [148-150]. While inexpensive and fast docking al- energy minimum wells by using non-Boltzmann sampling.
This is achieved, for instance, in the Locally Enhanced Sam-
gorithms can be used to scan large compound libraries and
pling method by considering a given number of non-
reduce their size, more accurate MD simulations can be ap-
interacting copies of the fragment to be explored (i.e., a resi-
plied when few ligand candidates remain. Thus, MD simula-
due side chain or a ligand), whereas the interaction of each
202 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

fragment with the rest of the system is reduced by a suitable in local minima, and the number of replicas should be large
factor from their original magnitudes [164]. Another exam- enough to ensure the swapping of adjacent replicas.
ple is the use of multiple copies of the system that are simul-
RE methods have been generally applied to investigate
taneously explored by MD, such as Replica Exchange (RE)
folding mechanisms and intermediate state structures (see for
MD [165]. Alternatively, enhanced sampling can also be instance refs. [171-175]). The number of studies devoted to
attempted via the explicit modification of the potential sur-
rational drug design has increased in the last years. For in-
face. The aim is to reduce the time spent by the simulated
stance, Verkhivker et al. [176] used parallel tempering simu-
system in a local energy minimum well, which would thus
lations to rank the different complexes according to the aver-
favor the transition from the conformational region corre-
age ligand-protein interaction energies in the binding of
sponding to such well to another. Accordingly, the alteration
SB203386 inhibitor to HIV-1 protease. This approach
of the potential surface reduces the propensity of energy yielded better sampling of the inhibitor conformational space
wells to act as conformational traps and promotes the system
during the binding process compared to MC simulations,
to sample the rest of the available conformational space. This
where the protein was trapped in local minima even at high
approach includes techniques such as umbrella sampling
temperature. Nagashima and co-workers developed a new
[166] and accelerated MD [167].
software framework enhancing the efficiency of conforma-
In the following we examine in more detail a selection of tional sampling by RE [177]. The toolkit employed object-
computational techniques that have been applied to explore oriented design to overcome the complexity associated with
the impact of conformational flexibility on the interaction biomolecules and parallelization, and to better interact with
between ligands and their biomolecular targets. other software components, in the perspective of being ap-
plied to receptor-drug docking analyses.
Replica Exchange Methods Different RE variants have been reported in the literature.
The original RE method, first reported by Swendens and Berne and co-workers developed a RE method with solute
Wang in 1986 [168], was developed to enhance conforma- tempering variant able to explicitly include the solvent con-
tional transitions by crossing energy barriers in the rugged tribution, but also to reduce the number of replicas by avoid-
energy landscape of biomolecules [169, 170]. This tech- ing the evaluation of solvent-solvent interactions [178]. The
nique, also known as parallel tempering, involves running a strategy allows the potential energy to scale with the tem-
number of simulations at different temperature, and exchang- perature in a way that the molecule becomes hotter while the
ing temperature and coordinates value every fixed number of waters stay cold. The system is thus divided into two parts,
steps based on Metropolis criteria (Fig. (8)). Thus, improved the protein and the waters or, eventually, the protein and its
samplings are obtained by exchanging complete configura- solvation shell as the central group, and the remaining water
tions and allowing low temperature systems to access a rep- as the bath. Then, the temperature of the central group is
resentative set of low-temperature regions of phase space. changed, while the temperature of the bath is kept at the
Since the simulation of multiple replicas requires a propor- same target value. In this way the number of required repli-
tional higher computational effort, the set of temperatures cas dramatically reduces, since the acceptance probability is
should be chosen to ensure that no replica becomes trapped independent from the solvent-solvent interaction energy.

Fig. (8). Schematic representation of replica exchange formalism. Several replicas of the system are simulated at different temperatures. En-
hanced sampling is achieved by allowing the systems at different temperatures to exchange complete configurations. The inclusion of higher
temperature systems ensures that the lower temperature ones can access a representative set of low-temperature regions of phase space.
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 203

Simmerling et al. applied a hybrid implicit/explicit solvation Accelerated Molecular Dynamics


model, a REMD variant where simulations are performed
McCammon and coworkers developed the Accelerated
with a fully explicit solvent, but the exchange probability is
Molecular Dynamics method [167, 184], based on previous
determined by combining the first solvation shell and a con-
works by Voter [185], to improve and enhance the explora-
tinuum solvation model [179]. The consequent reduction of
the system size significantly reduces the computational cost, tion of the free energy landscape of large biological systems.
In this method a bias potential is added to the true one reduc-
while maintaining a very good agreement with results ob-
ing the energy barriers between states by raising the potential
tained from the standard explicit solvent REMD.
energy in the minima. A threshold energy is defined so that
Instead of investigating a set of different temperatures, a below this energy, called the boost energy, the simulation is
set of modified potentials is used in Hamiltonian REMD, as performed on the modified potential, and when the system
proposed by Kwak et al. [180]. Analogously the technique has an energy above the threshold, the simulation is per-
designed by Hritz and Oostenbrik is based on a Hamiltonian formed on the true potential.
REMD scheme, but adopts soft-core interactions between
When the bias potential is applied, the probability to es-
parts of the systems mostly contributing to define high en-
cape from a potential basin increases depending on the boost
ergy barriers [181]. Differently from other Hamiltonian
energy value and the form of the bias potential. In the formu-
REMD algorithms, a lower number of replicas is required in
this case, with locally larger differences between the indi- lation proposed by Hamelberg et al., the boost bias potential
is of the form:
vidual Hamiltonians.
An alternative RE method, known as distance replica V (r )=
(E  V (r ))2 ,
exchange method, has been used by Lou and Cukier to in-  + (E  V (r ))
vestigate the conformational fluctuations of the apo form of
AK [122], which undergoes a large fluctuation in the transi- where E is the boost energy, and  is a tunable parameter
tion between open and closed conformations linked to the that controls the deepness of the modified potential energy
adjustment of the AMP-binding and lid domains. The ap- basins.
proach is used to sample the system along a reaction coordi-
nate monitoring the distance between the AMP-binding and High values of the parameter  will produce small boost
core domains. As the reaction coordinate decreases, the lid bias potentials, and accordingly the modified potencial en-
domain reduces its mobility moving toward more con- ergy basins will resemble the ones of the normal potential
strained fluctuations, until reaching a stable closed-form con- and no effective sampling gain will be achieved. After con-
formation, while the more stable core structure is well main- sidering different combinations of the boost energy and 
tained at all stages along the reaction coordinate. These re- parameters, Hamelberg et al. concluded that the best choice
sults suggest that a closed form state can be sampled even in combines an  value close to E–Vmin, where Vmin is the value
the absence of ligand. When compared to other techniques, of the potential energy minimum nearest to the starting struc-
this method demonstrated to greatly accelerate the rate and ture, and a value of E greater than Vmin but sufficiently small
the extent of configurational sampling. to prevent the random walk of the system over the potential
energy surface.
More recently, Rodinger et al. applied the distributed
replica method as an efficient Boltzmann sampling of con- This technique has been used by Markwick and cowork-
formational space that enables large-scale computing [182]. ers to explore the intrinsic conformational flexibility of pro-
Thus, one of the fundamental drawbacks of REMD is the tein GB3 [186], and by Grant et al. to investigate the con-
large CPU time required to simulate a complex biological formational switching in Ras protein [187]. Moreover, it has
system, since the replicas must run synchronously and a been used to examine the folding mechanism of a beta-
dedicated and homogeneous cluster is necessary for an effi- hairpin, trpzip2, thus enabling the generation of multiple
cient implementation of the algorithm. In the distributed rep- protein folding and unfolding trajectories in relatively short
lica method multiple replica of the system covering a range simulations [188], and the identification of folding pathways.
of temperature or of reaction coordinate are independently Metadynamics
simulated, but the replicas, instead of experiencing a pair-
wise exchange, are stochastically moved one at a time. This technique [189] has been successfully applied to a
wide variety of systems in different fields ranging from bio-
Finally, RE techniques have also been combined with physics [190-194] to materials science [195-199], crystal
free energy perturbation (FEP) and finite difference thermo- structure prediction [200-203] or chemistry [204-208], as a
dynamic integration (FDTI) in order to design new methods, useful tool to reconstruct free energy surfaces and to simu-
respectively identified as REFEP and RETI, able to estimate late rare events. In this method, the evolution of the system
the relative free energies of systems involving large reor- is biased by adding a history-dependent potential energy
ganizations of the environment [183]. These techniques have function of a determined set of collective variables to the
been applied to the calculation of the hydration free energy Hamiltonian of the system. The potential is constructed
of methane and water. Compared to other standard methods, through successive addition of Gaussian functions that act as
REFEP and RETI showed lower random sampling and stan- repulsive potentials to prevent the system from revisiting
dard errors and very little hysteresis, plus an excellent points of the free energy surface (Fig. (9)). Consequently, the
agreement with experimental data, with a little or no extra system can escape the wells in the rugged landscape and
computational cost. efficiently explore it. The sum of Gaussian functions de-
204 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

Fig. (9). Schematic representation of time-dependent reconstruction of the multidimensional free energy. The potential is constructed as a
sum of gaussians centered along the trajectory of the collective variables. The dynamic evolution is labeled by the number of gaussians.

posed along the trajectory of the system is then used to re- pled conformations. Thus, it is possible for example to inves-
construct the free energy. tigate large conformational changes, like the folding energy
landscape of proteins [212], or the transition between open
A delicate aspect of the method is to identify the vari-
and closed states of a kinase protein (CDK5; [127]) with
ables that are of particular relevance for the motion of the
biomolecular system and that are difficult to sample by con- reasonable computational resources.
ventional simulation techniques [189]. Ideally the collective Another scenario where metadynamics have proved to be
variables should i) clearly distinguish between the initial valuable is the inclusion of the receptor flexibility in docking
state, the final states and the intermediates, ii) should de- studies [213] and the evaluation of small ligand-induced con-
scribe all the slow events that are relevant to the process of formational changes. As metadynamics is coupled to an MD
interest, and iii) their number should not be too large in order simulation, it is possible, in principle, to reproduce the bind-
to avoid a too long simulation time to fill the free energy ing pathway of a ligand with a receptor from the solvent to
surface. Unfortunately, there is no a priori recipe for finding the binding site, thus allowing the necessary rearrangements
the correct set of collective variables, and in many cases a in the side chains of residues to occur. For example, the
trial-and-error process is required to detect the most adequate method has been applied to study the folding inhibition of
choice. These variables are functions of the coordinates of the HIV-1 protease with a small peptide [214], where the
the system. The simplest definition involves geometry- large number of rotatable bonds and the necessity to include
related variables, such as interatomic distances, angles or the flexibility of the receptor limits the use of simple MD or
torsions. However, more elaborate definitions are also valid, common docking methods.
such as the potential energy, which is particularly useful in
Although metadynamics simulations imply a substantial
phase transitions [209] definitions, normal modes and essen-
computational cost, its advantage relies in the fact that it
tial coordinates derived from essential dynamics, which have
makes possible the identification of bottlenecks or unfavor-
been used to explore the conformational space of peptides able interactions that occur along the ligand’s binding path.
[210], or specific variables in proteins, such as the ‘helicity
This is clearly illustrated by studies of the translocation
of the backbone’ or the ‘dihedral correlation’ [211]. A his-
process of ampicillin through OmpF in Escherichia coli
tory-dependent dynamics is constructed in the space of these
[215], or of the trimethylamonium migration through the
variables in order to compensate the underlying free energy.
gorge of acetylcholinesterase [216]. In both cases, a good
As stated before, the use of methods with atomic detail, description of the mechanism and energetics of ligand migra-
like all-atom explicit MD simulations, provide sufficient tion is obtained, which in turn could be used for the design
accuracy to investigate large fluctuations and conformational of more potent and novel drugs.
changes, but the computational time needed to get enough
sampling is often too high. With metadynamics, the time CONCLUSIONS
required to explore large portions of conformational space is
drastically reduced as the biasing potential forces the system Protein flexibility undoubtedly plays a critical role in
to evolve to a fixed final state through non-previously sam- determining molecular recognition. The interaction between
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 205

a drug-like compound and its macromolecular target can FDTI = Finite Difference Thermodynamic Inte-
involve either local changes in the conformational distribu- gration
tion of the side chain of specific residues in the binding site,
FEP = Free Energy Perturbation
small amplitude readjustements in the backbone of structural
elements, or even large-scale conformational alterations af- GDH = Glutamate Dehydrogenase
fecting the spatial location of domains in the protein. Model- HSP = Heat-Shock Protein
ing efforts in drug design generally disregard these effects,
as they represent a formidable challenge to computational LMW-PTP = Low-Molecular-Weight Protein Tyro-
and simulation studies. Nevertheless, it is also clear that a sine Phosphatase
precise knowledge of the structural plasticity of the target MD = Molecular Dynamics
and its implication in mediating the binding of a drug is ex-
tremely valuable to assist the development of more potent, NMP = Nucleoside Monophosphate
selective compounds in structure-based drug design. Impor- PKA = Protein Kinase A
tantly, such knowledge can also lead to the discovery of
novel effective strategies to modulate the activity of a bio- PTP = Protein Tyrosine Phosphatase
molecular target, either by identifying alternative binding RE = Replica Exchange
pockets present in specific conformations of the protein or
by exploiting distinct pharmacophoric features of the binding REFERENCES
site associated with conformational transitions. Overall, this
[1] Jimenez, R.; Salazar, G.; Baldridge, K. K.; Romesberg, F. E. Flexi-
information will be translated into new opportunities for bility and molecular recognition in the immune system. Proc. Natl.
drug discovery, such as enhanced rates of hit detection in Acad. Sci. USA, 2003, 100, 92-97.
pharmacophore-guided virtual screening or the identification [2] Sundberg, E. J.; Mariuzza, R. A. Molecular recognition in anti-
of lead compounds incorporating novel scaffolds. body-antigen complexes. Adv. Protein Chem., 2002, 61, 119-160.
[3] Pabo, C. O.; Sauer, R. T. Protein-DNA recognition. Annu. Rev.
MD techniques are extremely powerful tools that allow Biochem., 1984, 53, 293-321.
us to gain insight into the link between target flexibility and [4] Pabo, C. O.; Sauer, R. T. Transcription factors: structural families
ligand binding. They provide a firm basis on which structural and principles of DNA recognition. Annu. Rev. Biochem., 1992, 61,
1053-1095.
information and biological data may be reconciled. All-atom [5] Hammes, G. G. Multiple conformational changes in enzyme ca-
molecular dynamics simulations have become a standard talysis. Biochemistry, 2002, 41, 8221-8228.
tool to examine the fine structural details of ligand-receptor [6] Hunter, C. A. Quantifying intermolecular interactions: guidelines
interactions, as they can reveal subtle structural rearrange- for the molecular recognition toolbox. Angew. Chem. Int. Ed. Engl.,
2004, 43, 5310-5324.
ments due to the mutual accommodation between the ligand [7] Gohlke, H.; Klebe, G. Approaches to the description and prediction
and the residues that define the walls of the pocket upon of the binding affinity of small-molecule ligands to macromolecu-
ligand binding in few nanoseconds. The challenge, neverthe- lar receptors. Angew. Chem. Int. Ed. Engl., 2002, 41, 2644-2676.
less, consists of extending the capability of MD techniques [8] Bissantz, C.; Folkers, G.; Rognan, D. Protein-based virtual screen-
to describe conformational transitions that require longer ing of chemical databases. 1. Evaluation of different dock-
ing/scoring combinations. J. Med. Chem., 2000, 43, 4759-4767.
time scales. The rapid progress in algorithms and computa- [9] Charifson, P. S.; Corkery, J. J.; Murcko, M. A.; Walters, W. P.
tional resources make it feasible to shed light into the mo- Consensus scoring: A method for obtaining improved hit rates from
lecular mechanism underlying those large-scale conforma- docking databases of three-dimensional structures into proteins. J.
tion transitions and their functional implications, opening Med. Chem., 1999, 42, 5100-5109.
[10] Spyrakis, F.; Cozzini, P.; Kellogg, G. E. Docking and scoring in
new avenues for drug discovery. Enhanced sampling tech- drug discovery. In: Burger's Medicinal Chemistry; Eds.; Wiley-
niques contribute decisively to the progress in this field, VCH: New York, 2010.
however, the synergy between experiments and simulations [11] Spyrakis, F.; Kellogg, G. E.; Amadasi, A.; Cozzini, P. Scoring
shall undoubdtedly represent the fundamental step towards functions for virtual screening. In Frontiers in Drug Design and
the understanding of the role played by dynamics properties Discovery; Eds.; Bentham Science Publisher, 2007, pp 317-379.
[12] Moitessier, N.; Westhof, E.; Hanessian, S. Docking of aminoglyco-
in a variety of biochemical systems. sides to hydrated and flexible RNA. J. Med. Chem., 2006, 49,
1023-1033.
ACKNOWLEDGEMENTS [13] Schnecke, V.; Kuhn, L. A. Virtual screening with solvation and
ligand-induced complementarity. Perspect. Drug Discov. Des.,
We acknowledge the financial support from the Ministe- 2000, 20, 171-190.
rio de Innovación y Ciencia (MICINN; SAF2008-05595 and [14] Shoemaker, B. A.; Wang, J.; Wolynes, P. G. Structural correlations
in protein folding funnels. Proc. Natl. Acad. Sci. USA, 1997, 94,
SAF2009-08811) and Generalitat de Catalunya (2009- 777-782.
SGR00298). [15] van Dijk, A. D.; Bonvin, A. M. Solvated docking: introducing
water into the modelling of biomolecular complexes. Bioinformat-
ABBREVIATIONS ics, 2006, 22, 2340-2347.
[16] Verdonk, M. L.; Chessari, G.; Cole, J. C.; Hartshorn, M. J.;
AK = Adenylate Kinase Murray, C. W.; Nissink, J. W.; Taylor, R. D.; Taylor, R. Modeling
water molecules in protein-ligand docking using GOLD. J. Med.
AL = Activation Loop Chem., 2005, 48, 6504-6515.
[17] Osterberg, F.; Morris, G. M.; Sanner, M. F.; Olson, A. J.; Goodsell,
AChE = Acetylcholinesterase D. S. Automated docking to multiple target structures: incorpora-
tion of protein mobility and structural water heterogeneity in
CaM = Calmodulin AutoDock. Proteins, 2002, 46, 34-40.
206 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

[18] Rarey, M.; Kramer, B.; Lengauer, T. The particle concept: placing [45] Tsai, C. J.; Ma, B.; Nussinov, R. Folding and binding cascades:
discrete water molecules during protein-ligand docking predictions. shifts in energy landscapes. Proc. Natl. Acad. Sci. USA, 1999, 96,
Proteins, 1999, 34, 17-28. 9970-9972.
[19] Czodrowski, P.; Dramburg, I.; Sotriffer, C. A.; Klebe, G. Devel- [46] Xu, Y.; Colletier, J. P.; Jiang, H.; Silman, I.; Sussman, J. L.; Weik,
opment, validation, and application of adapted PEOE charges to es- M. Induced-fit or preexisting equilibrium dynamics? Lessons from
timate pKa values of functional groups in protein-ligand com- protein crystallography and MD simulations on acetylcho-
plexes. Proteins, 2006, 65, 424-437. linesterase and implications for structure-based drug design. Pro-
[20] Brenk, R.; Vetter, S. W.; Boyce, S. E.; Goodin, D. B.; Shoichet, B. tein Sci., 2008, 17, 601-605.
K. Probing molecular docking in a charged model binding site. J. [47] Okazaki, K.; Takada, S. Dynamic energy landscape view of cou-
Mol. Biol., 2006, 357, 1449-1470. pled binding and protein conformational change: induced-fit versus
[21] Deng, W.; Verlinde, C. L. Evaluation of different virtual screening population-shift mechanisms. Proc. Natl. Acad. Sci. USA, 2008,
programs for docking in a charged binding pocket. J. Chem. Inf. 105, 11182-11187.
Model., 2008, 48, 2010-2020. [48] Sullivan, S. M.; Holyoak, T. Enzymes with lid-gated active sites
[22] Bahar, I.; Chennubhotla, C.; Tobi, D. Intrinsic dynamics of en- must operate by an induced fit mechanism instead of conforma-
zymes in the unbound state and relation to allosteric regulation. tional selection. Proc. Natl. Acad. Sci. USA, 2008, 105, 13829-
Curr. Opin. Struct. Biol., 2007, 17, 633-640. 13834.
[23] Carlson, H. A.; McCammon, J. A. Accommodating protein flexibil- [49] Bakan, A.; Bahar, I. The intrinsic dynamics of enzymes plays a
ity in computational drug design. Mol. Pharmacol., 2000, 57, 213- dominant role in determining the structural changes induced upon
218. inhibitor binding. Proc. Natl. Acad. Sci. USA, 2009, 106, 14349-
[24] Cozzini, P.; Kellogg, G. E.; Spyrakis, F.; Abraham, D. J.; Costan- 14354.
tino, G.; Emerson, A.; Fanelli, F.; Gohlke, H.; Kuhn, L. A.; Morris, [50] Bui, J. M.; McCammon, J. A. Protein complex formation by ace-
G. M.; Orozco, M.; Pertinhez, T. A.; Rizzi, M.; Sotriffer, C. A. tylcholinesterase and the neurotoxin fasciculin-2 appears to involve
Target flexibility: an emerging consideration in drug discovery and an induced-fit mechanism. Proc. Natl. Aca. Sci. USA, 2006, 103,
design. J. Med. Chem., 2008, 51, 6237-6255. 15451-15456.
[25] Henzler-Wildman, K.; Kern, D. Dynamic personalities of proteins. [51] Levy, Y.; Onuchic, J. N.; Wolynes, P. G. Fly-casting in protein-
Nature 2007, 450, 964-972. DNA binding: frustration between protein folding and electrostatics
[26] Tsai, C. J.; Kumar, S.; Ma, B.; Nussinov, R. Folding funnels, bind- facilitates target recognition. J. Am. Chem. Soc., 2007, 129, 738-
ing funnels, and protein function. Protein Sci., 1999, 8, 1181-1190. 739.
[27] Savir, Y.; Tlusty, T. Conformational proofreading: the impact of [52] Sugase, K.; Dyson, H. J.; Wright, P. E. Mechanism of coupled
conformational changes on the specificity of molecular recognition. folding and binding of an intrinsically disordered protein. Nature,
PLoS One, 2007, 2, e468. 2007, 447, 1021-1025.
[28] Mittag, T.; Kay, L. E.; Forman-Kay, J. D. Protein dynamics and [53] Arora, K.; Brooks, C. L., 3rd. Large-scale allosteric conformational
conformational disorder in molecular recognition. J. Mol. Recog- transitions of adenylate kinase appear to involve a population-shift
nit., 2010, 23, 105-116. mechanism. Proc. Natl. Acad. Sci. USA, 2007, 104, 18496-18501.
[29] Freire, E. The propagation of binding interactions to remote sites in [54] Hanson, J. A.; Duderstadt, K.; Watkins, L. P.; Bhattacharyya, S.;
proteins: analysis of the binding of the monoclonal antibody D1.3 Brokaw, J.; Chu, J. W.; Yang, H. Illuminating the mechanistic roles
to lysozyme. Proc. Natl. Acad. Sci. USA, 1999, 96, 10118-10122. of enzyme conformational dynamics. Proc. Natl. Acad. Sci. USA,
[30] Ma, B.; Kumar, S.; Tsai, C. J.; Nussinov, R. Folding funnels and 2007, 104, 18055-18060.
binding mechanisms. Protein Eng., 1999, 12, 713-720. [55] James, L. C.; Roversi, P.; Tawfik, D. S. Antibody multispecificity
[31] Onuchic, J. N. Contacting the protein folding funnel with NMR. mediated by conformational diversity. Science, 2003, 299, 1362-
Proc. Natl. Acad. Sci. USA, 1997, 94, 7129-7131. 1367.
[32] Rejto, P. A.; Freer, S. T. Protein conformational substates from X- [56] Claussen, H.; Buning, C.; Rarey, M.; Lengauer, T. FlexE: efficient
ray crystallography. Prog. Biophys. Mol. Biol., 1996, 66, 167-196. molecular docking considering protein structure variations. J. Mol.
[33] Koshland, D. E. Application of a Theory of Enzyme Specificity to Biol., 2001, 308, 377-395.
Protein Synthesis. Proc. Natl. Acad. Sci. USA, 1958, 44, 98-104. [57] Huey, R.; Morris, G.; Olson, A.; Goodsell, D. A semiempirical free
[34] Monod, J.; Wyman, J.; Changeux, J. P. On the nature of allosteric energy force field with charge-based desolvation. J. Comput.
transitions: A plausible model. J. Mol. Biol., 1965, 12, 88-118. Chem., 2007, 28, 1145-1152.
[35] Valente, A. P.; Miyamoto, C. A.; Almeida, F. C. Implications of [58] Verdonk, M. L.; Cole, J. C.; Hartshorn, M. J.; Murray, C. W.; Tay-
protein conformational diversity for binding and development of lor, R. D. Improved protein-ligand docking using GOLD. Proteins,
new biological active compounds. Curr. Med. Chem., 2006, 13, 2003, 52, 609-623.
3697-3703. [59] Zavodszky, M. I.; Kuhn, L. A. Side-chain flexibility in protein-
[36] Bosshard, H. R. Molecular recognition by induced fit: how fit is the ligand binding: the minimal rotation hypothesis. Protein Sci., 2005,
concept? News Physiol. Sci., 2001, 16, 171-173. 14, 1104-1114.
[37] James, L. C.; Tawfik, D. S. Conformational diversity and protein [60] Rosenfeld, R. J.; Goodsell, D. S.; Musah, R. A.; Morris, G. M.;
evolution-a 60-year-old hypothesis revisited. Trends Biochem. Sci., Goodin, D. B.; Olson, A. J. Automated docking of ligands to an ar-
2003, 28, 361-368. tificial active site: augmenting crystallographic analysis with com-
[38] Kern, D.; Zuiderweg, E. R. The role of dynamics in allosteric regu- puter modeling. J. Comput. Aided Mol. Des., 2003, 17, 525-536.
lation. Curr. Opin. Struct. Biol., 2003, 13, 748-757. [61] Zacharias, M.; Sklenar, H. Harmonic modes as variables to ap-
[39] Mittermaier, A.; Kay, L. E. New tools provide new insights in proximately account for receptor flexibility in ligand-receptor
NMR studies of protein dynamics. Science, 2006, 312, 224-228. docking simulations: applications to DNA minor groove ligand
[40] Showalter, S. A.; Bruschweiler, R. Quantitative molecular ensem- complex. J. Comput. Chem., 1999, 20, 287-300.
ble interpretation of NMR dipolar couplings without restraints. J. [62] Zavodszky, M. I.; Lei, M.; Thorpe, M. F.; Day, A. R.; Kuhn, L. A.
Am. Chem. Soc., 2007, 129, 4158-4159. Modeling correlated main-chain motions in proteins for flexible
[41] Tang, C.; Schwieters, C. D.; Clore, G. M. Open-to-closed transition molecular recognition. Proteins, 2004, 57, 243-261.
in apo maltose-binding protein observed by paramagnetic NMR. [63] Cavasotto, C. N.; Abagyan, R. A. Protein flexibility in ligand dock-
Nature, 2007, 449, 1078-1082. ing and virtual screening to protein kinases. J. Mol. Biol., 2004,
[42] Tobi, D.; Bahar, I. Structural changes involved in protein binding 337, 209-225.
correlate with intrinsic motions of proteins in the unbound state. [64] Cavasotto, C. N.; Kovacs, J. A.; Abagyan, R. A. Representing
Proc. Natl. Acad. Sci. USA, 2005, 102, 18908-18913. receptor flexibility in ligand docking through relevant normal
[43] Gulukota, K.; Wolynes, P. G. Statistical mechanics of kinetic modes. J. Am. Chem. Soc., 2005, 127, 9632-9640.
proofreading in protein folding in vivo. Proc. Natl. Acad. Sci. USA, [65] Kallblad, P.; Dean, P. M. Efficient conformational sampling of
1994, 91, 9292-9296. local side-chain flexibility. J. Mol. Biol., 2003, 326, 1651-1665.
[44] Kumar, S.; Ma, B.; Tsai, C. J.; Sinha, N.; Nussinov, R. Folding and [66] Nabuurs, S. B.; Wagener, M.; de Vlieg, J. A flexible approach to
binding cascades: dynamic landscapes and population shifts. Pro- induced fit docking. J. Med. Chem., 2007, 50, 6507-6518.
tein Sci., 2000, 9, 10-19.
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 207

[67] Totrov, M.; Abagyan, R. Flexible ligand docking to multiple recep- fasciculin-II. Acta Crystallogr. D Biol. Crystallogr., 2000, 56,
tor conformations: a practical alternative. Curr. Opin. Struct. Biol., 1385-1394.
2008, 18, 178-184. [87] Camps, P.; Formosa, X.; Galdeano, C.; Munoz-Torrero, D.; Rami-
[68] Krol, M.; Tournier, A. L.; Bates, P. A. Flexible relaxation of rigid- rez, L.; Gomez, E.; Isambert, N.; Lavilla, R.; Badia, A.; Clos, M.
body docking solutions. Proteins, 2007, 68, 159-169. V.; Bartolini, M.; Mancini, F.; Andrisano, V.; Arce, M. P.; Rodri-
[69] Barril, X.; Morley, S. D. Unveiling the full potential of flexible guez-Franco, M. I.; Huertas, O.; Dafni, T.; Luque, F. J. Pyrano[3,2-
receptor docking using multiple crystallographic structures. J. Med. c]quinoline-6-chlorotacrine hybrids as a novel family of acetylcho-
Chem., 2005, 48, 4432-4443. linesterase- and beta-amyloid-directed anti-Alzheimer compounds.
[70] Craig, I.; Essex, J.; Spiegel, K. Ensemble docking into multiple J. Med. Chem., 2009, 52, 5365-5379.
crystallographically-derived protein structures: an evaluation based [88] Gutteridge, A.; Thornton, J. Conformational changes observed in
on the statistical analysis of enrichments. J. Chem. Inf. Model., enzyme crystal structures upon substrate binding. J. Mol. Biol.,
2010, 50, 511-524. 2005, 346, 21-28.
[71] Akerud, T.; Thulin, E.; Van Etten, R. L.; Akke, M. Intramolecular [89] Barril, X.; Fradera, X. Incorporating protein flexibility into docking
dynamics of low molecular weight protein tyrosine phosphatase in and structure-based drug design. Exp. Opin. Drug Discov., 2006, 1,
monomer-dimer equilibrium studied by NMR: a model for changes 335-349.
in dynamics upon target binding. J. Mol. Biol., 2002, 322, 137-152. [90] Stebbins, C. E.; Russo, A. A.; Schneider, C.; Rosen, N.; Hartl, F.
[72] Banci, L.; Bertini, I.; Cantini, F.; D'Onofrio, M.; Viezzoli, M. S. U.; Pavletich, N. P. Crystal structure of an Hsp90-geldanamycin
Structure and dynamics of copper-free SOD: The protein before complex: targeting of a protein chaperone by an antitumor agent.
binding copper. Protein Sci., 2002, 11, 2479-2492. Cell, 1997, 89, 239-250.
[73] Korzhnev, D. M.; Karlsson, B. G.; Orekhov, V. Y.; Billeter, M. [91] Chiosis, G.; Timaul, M. N.; Lucas, B.; Munster, P. N.; Zheng, F.
NMR detection of multiple transitions to low-populated states in F.; Sepp-Lorenzino, L.; Rosen, N. A small molecule designed to
azurin. Protein Sci., 2003, 12, 56-65. bind to the adenine nucleotide pocket of Hsp90 causes Her2 degra-
[74] Sussman, J. L.; Harel, M.; Frolow, F.; Oefner, C.; Goldman, A.; dation and the growth arrest and differentiation of breast cancer
Toker, L.; Silman, I. Atomic structure of acetylcholinesterase from cells. Chem. Biol., 2001, 8, 289-299.
Torpedo californica: a prototypic acetylcholine-binding protein. [92] Wright, L.; Barril, X.; Dymock, B.; Sheridan, L.; Surgenor, A.;
Science, 1991, 253, 872-879. Beswick, M.; Drysdale, M.; Collier, A.; Massey, A.; Davies, N.;
[75] Raves, M. L.; Harel, M.; Pang, Y. P.; Silman, I.; Kozikowski, A. Fink, A.; Fromont, C.; Aherne, W.; Boxall, K.; Sharp, S.; Work-
P.; Sussman, J. L. Structure of acetylcholinesterase complexed with man, P.; Hubbard, R. E. Structure-activity relationships in purine-
the nootropic alkaloid, (-)-huperzine A. Nat. Struct. Biol., 1997, 4, based inhibitor binding to HSP90 isoforms. Chem. Biol., 2004, 11,
57-63. 775-785.
[76] Barril, X.; Kalko, S. G.; Orozco, M.; Luque, F. J. Rational design [93] Ali, M. M.; Roe, S. M.; Vaughan, C. K.; Meyer, P.; Panaretou, B.;
of reversible acetylcholinesterase inhibitors. Mini Rev. Med. Chem., Piper, P. W.; Prodromou, C.; Pearl, L. H. Crystal structure of an
2002, 2, 27-36. Hsp90-nucleotide-p23/Sba1 closed chaperone complex. Nature,
[77] Muñoz-Torrero, D. Acetylcholinesterase inhibitors as disease- 2006, 440, 1013-1017.
modifying therapies for Alzheimer’s disease. Curr. Med. Chem., [94] Staessen, J. A.; Li, Y.; Richart, T. Oral renin inhibitors. Lancet,
2008, 15, 2433-2455. 2006, 368, 1449-1456.
[78] Harel, M.; Schalk, I.; Ehret-Sabatier, L.; Bouet, F.; Goeldner, M.; [95] Vieira, E.; Binggeli, A.; Breu, V.; Bur, D.; Fischli, W.; Guller, R.;
Hirth, C.; Axelsen, P. H.; Silman, I.; Sussman, J. L. Quaternary Hirth, G.; Marki, H. P.; Muller, M.; Oefner, C.; Scalone, M.;
ligand binding to aromatic residues in the active-site gorge of ace- Stadler, H.; Wilhelm, M.; Wostl, W. Substituted piperidines--
tylcholinesterase. Proc. Natl. Acad. Sci. USA, 1993, 90, 9031-9035. highly potent renin inhibitors due to induced fit adaptation of the
[79] Dvir, H.; Wong, D. M.; Harel, M.; Barril, X.; Orozco, M.; Luque, active site. Bioorg. Med. Chem. Lett., 1999, 9, 1397-1402.
F. J.; Munoz-Torrero, D.; Camps, P.; Rosenberry, T. L.; Silman, I.; [96] Bezencon, O.; Bur, D.; Weller, T.; Richard-Bildstein, S.; Remen,
Sussman, J. L. 3D structure of Torpedo californica acetylcho- L.; Sifferlen, T.; Corminboeuf, O.; Grisostomi, C.; Boss, C.; Prade,
linesterase complexed with huprine X at 2.1 A resolution: kinetic L.; Delahaye, S.; Treiber, A.; Strickner, P.; Binkert, C.; Hess, P.;
and molecular dynamic correlates. Biochemistry, 2002, 41, 2970- Steiner, B.; Fischli, W. Design and preparation of potent, nonpep-
2981. tidic, bioavailable renin inhibitors. J. Med. Chem., 2009, 52, 3689-
[80] Kryger, G.; Silman, I.; Sussman, J. L. Structure of acetylcho- 3702.
linesterase complexed with E2020 (Aricept): implications for the [97] Pargellis, C.; Tong, L.; Churchill, L.; Cirillo, P. F.; Gilmore, T.;
design of new anti-Alzheimer drugs. Structure, 1999, 7, 297-307. Graham, A. G.; Grob, P. M.; Hickey, E. R.; Moss, N.; Pav, S.;
[81] Bourne, Y.; Taylor, P.; Radic, Z.; Marchot, P. Structural insights Regan, J. Inhibition of p38 MAP kinase by utilizing a novel allos-
into ligand interactions at the acetylcholinesterase peripheral ani- teric binding site. Nat. Struct. Biol., 2002, 9, 268-272.
onic site. EMBO J., 2003, 22, 1-12. [98] Regan, J.; Pargellis, C.; Cirillo, P. F.; Gilmore, T.; Hickey, E. R.;
[82] Bourne, Y.; Kolb, H. C.; Radic, Z.; Sharpless, K. B.; Taylor, P.; Peet, G.; Proto, A.; Swinamer, A.; Moss, N. The kinetics of binding
Marchot, P. Freeze-frame inhibitor captures acetylcholinesterase in to p38MAP kinase by analogues of BIRB 796. Bioorg. Med. Chem.
a unique conformation. Proc. Natl. Acad. Sci. USA, 2004, 101, Lett., 2003, 13, 3101-3104.
1449-1454. [99] Abad-Zapatero, C. Ligand efficiency indices for effective drug
[83] Colletier, J. P.; Sanson, B.; Nachon, F.; Gabellieri, E.; Fattorusso, discovery. Exp. Opin. Drug Discov., 2007, 2, 469-488.
C.; Campiani, G.; Weik, M. Conformational flexibility in the pe- [100] Wiesmann, C.; Barr, K. J.; Kung, J.; Zhu, J.; Erlanson, D. A.; Shen,
ripheral site of Torpedo californica acetylcholinesterase revealed W.; Fahr, B. J.; Zhong, M.; Taylor, L.; Randal, M.; McDowell, R.
by the complex structure with a bifunctional inhibitor. J. Am. S.; Hansen, S. K. Allosteric inhibition of protein tyrosine phospha-
Chem. Soc., 2006, 128, 4526-4527. tase 1B. Nat. Struct. Mol. Biol., 2004, 11, 730-737.
[84] Haviv, H.; Wong, D. M.; Greenblatt, H. M.; Carlier, P. R.; Pang, Y. [101] Jia, Z.; Barford, D.; Flint, A. J.; Tonks, N. K. Structural basis for
P.; Silman, I.; Sussman, J. L. Crystal packing mediates enantiose- phosphotyrosine peptide recognition by protein tyrosine phospha-
lective ligand recognition at the peripheral site of acetylcho- tase 1B. Science, 1995, 268, 1754-1758.
linesterase. J. Am. Chem. Soc., 2005, 127, 11029-11036. [102] Liu, S.; Zeng, L. F.; Wu, L.; Yu, X.; Xue, T.; Gunawan, A. M.;
[85] Rydberg, E. H.; Brumshtein, B.; Greenblatt, H. M.; Wong, D. M.; Long, Y. Q.; Zhang, Z. Y. Targeting inactive enzyme conforma-
Shaya, D.; Williams, L. D.; Carlier, P. R.; Pang, Y. P.; Silman, I.; tion: aryl diketoacid derivatives as a new class of PTP1B inhibitors.
Sussman, J. L. Complexes of alkylene-linked tacrine dimers with J. Am. Chem. Soc., 2008, 130, 17075-17084.
Torpedo californica acetylcholinesterase: Binding of Bis5-tacrine [103] Betts, M. J.; Sternberg, M. J. An analysis of conformational
produces a dramatic rearrangement in the active-site gorge. J. Med. changes on protein-protein association: implications for predictive
Chem., 2006, 49, 5491-5500. docking. Protein Eng., 1999, 12, 271-283.
[86] Kryger, G.; Harel, M.; Giles, K.; Toker, L.; Velan, B.; Lazar, A.; [104] Hardy, J. A.; Wells, J. A. Searching for new allosteric sites in en-
Kronman, C.; Barak, D.; Ariel, N.; Shafferman, A.; Silman, I.; zymes. Curr. Opin. Struct. Biol., 2004, 14, 706-715.
Sussman, J. L. Structures of recombinant native and E202Q mutant [105] Lindsley, J. E.; Rutter, J. Whence cometh the allosterome? Proc.
human acetylcholinesterase complexed with the snake-venom toxin Natl. Acad. Sci. USA, 2006, 103, 10533-10535.
208 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

[106] Whitty, A.; Kumaravel, G. Between a rock and a hard place? Nat. [128] Johnson, D. A.; Akamine, P.; Radzio-Andzelm, E.; Madhusudan,
Chem. Biol., 2006, 2, 112-118. M.; Taylor, S. S. Dynamics of cAMP-dependent protein kinase.
[107] Kumar, S.; Ma, B.; Tsai, C. J.; Wolfson, H.; Nussinov, R. Folding Chem. Rev., 2001, 101, 2243-2270.
funnels and conformational transitions via hinge-bending motions. [129] Knighton, D. R.; Bell, S. M.; Zheng, J.; Ten Eyck, L. F.; Xuong, N.
Cell Biochem. Biophys., 1999, 31, 141-164. H.; Taylor, S. S.; Sowadski, J. M. 2.0 Å refined crystal structure of
[108] Li, Q.; Khosla, C.; Puglisi, J. D.; Liu, C. W. Solution structure and the catalytic subunit of cAMP-dependent protein kinase complexed
backbone dynamics of the holo form of the frenolicin acyl carrier with a peptide inhibitor and detergent. Acta Cryst. D Biol. Cryst.,
protein. Biochemistry, 2003, 42, 4648-4657. 1993, 49, 357-361.
[109] Valentine, E. R.; Palmer, A. G., 3rd. Microsecond-to-millisecond [130] Zheng, J.; Knighton, D. R.; Xuong, N. H.; Taylor, S. S.; Sowadski,
conformational dynamics demarcate the GluR2 glutamate receptor J. M.; Ten Eyck, L. F. Crystal structures of the myristylated cata-
bound to agonists glutamate, quisqualate, and AMPA. Biochemis- lytic subunit of cAMP-dependent protein kinase reveal open and
try, 2005, 44, 3410-3417. closed conformations. Protein Sci., 1993, 2, 1559-1573.
[110] Seeliger, M. A.; Spichty, M.; Kelly, S. E.; Bycroft, M.; Freund, S. [131] Miyashita, O.; Onuchic, J. N.; Wolynes, P. G. Nonlinear elasticity,
M.; Karplus, M.; Itzhaki, L. S. Role of conformational heterogene- proteinquakes, and the energy landscapes of functional transitions
ity in domain swapping and adapter function of the Cks proteins. J. in proteins. Proc. Natl. Acad. Sci. USA, 2003, 100, 12570-12575.
Biol. Chem., 2005, 280, 30448-30459. [132] Yan, H.; Tsai, M. D. Nucleoside monophosphate kinases: structure,
[111] Folmer, R. H.; Geschwindner, S.; Xue, Y. Crystal structure and mechanism, and substrate specificity. Adv. Enzymol. Relat. Areas
NMR studies of the apo SH2 domains of ZAP-70: two bikes rather Mol. Biol., 1999, 73, 103-134.
than a tandem. Biochemistry, 2002, 41, 14176-14184. [133] Tripathi, S.; Portman, J. J. Inherent flexibility determines the transi-
[112] McKercher, S. R.; Lombardo, C. R.; Bobkov, A.; Jia, X.; Assa- tion mechanisms of the EF-hands of calmodulin. Proc. Natl. Acad.
Munt, N. Identification of a PU.1-IRF4 protein interaction surface Sci. USA, 2009, 106, 2104-2109.
predicted by chemical exchange line broadening. Proc. Natl. Acad. [134] Chattopadhyaya, R.; Meador, W. E.; Means, A. R.; Quiocho, F. A.
Sci. USA, 2003, 100, 511-516. Calmodulin structure refined at 1.7 A resolution. J. Mol. Biol.,
[113] Venkitaramani, D. V.; Fulton, D. B.; Andreotti, A. H.; Johansen, K. 1992, 228, 1177-1192.
M.; Johansen, J. Solution structure and backbone dynamics of [135] Kuboniwa, H.; Tjandra, N.; Grzesiek, S.; Ren, H.; Klee, C. B.; Bax,
Calsensin, an invertebrate neuronal calcium-binding protein. Pro- A. Solution structure of calcium-free calmodulin. Nat. Struct. Biol.,
tein Sci., 2005, 14, 1894-1901. 1995, 2, 768-776.
[114] Wang, C.; Karpowich, N.; Hunt, J. F.; Rance, M.; Palmer, A. G. [136] Chou, J. J.; Li, S.; Klee, C. B.; Bax, A. Solution structure of
Dynamics of ATP-binding cassette contribute to allosteric control, Ca(2+)-calmodulin reveals flexible hand-like properties of its do-
nucleotide binding and energy transduction in ABC transporters. J. mains. Nat. Struct. Biol., 2001, 8, 990-997.
Mol. Biol., 2004, 342, 525-537. [137] Evenas, J.; Forsen, S.; Malmendal, A.; Akke, M. Backbone dynam-
[115] Yan, J.; Liu, Y.; Lukasik, S. M.; Speck, N. A.; Bushweller, J. H. ics and energetics of a calmodulin domain mutant exchanging be-
CBFbeta allosterically regulates the Runx1 Runt domain via a dy- tween closed and open conformations. J. Mol. Biol., 1999, 289,
namic conformational equilibrium. Nat. Struct. Mol. Biol., 2004, 603-617.
11, 901-906. [138] Vigil, D.; Gallagher, S. C.; Trewhella, J.; Garcia, A. E. Functional
[116] Yao, S.; Headey, S. J.; Keizer, D. W.; Bach, L. A.; Norton, R. S. C- dynamics of the hydrophobic cleft in the N-domain of calmodulin.
terminal domain of insulin-like growth factor (IGF) binding protein Biophys. J., 2001, 80, 2082-2092.
6: conformational exchange and its correlation with IGF-II binding. [139] Chen, Y. G.; Hummer, G. Slow conformational dynamics and
Biochemistry, 2004, 43, 11187-11195. unfolding of the calmodulin C-terminal domain. J. Am. Chem. Soc.,
[117] Giedroc, D. P.; Chen, X.; Pennella, M. A.; LiWang, A. C. Confor- 2007, 129, 2414-2415.
mational heterogeneity in the C-terminal zinc fingers of human [140] Lundstrom, P.; Mulder, F. A.; Akke, M. Correlated dynamics of
MTF-1: an NMR and zinc-binding study. J. Biol. Chem., 2001, consecutive residues reveal transient and cooperative unfolding of
276, 42322-42332. secondary structure in proteins. Proc. Natl. Acad. Sci. USA, 2005,
[118] Ramboarina, S.; Srividya, N.; Atkinson, R. A.; Morellet, N.; 102, 16984-16989.
Roques, B. P.; Lefevre, J. F.; Mely, Y.; Kieffer, B. Effects of tem- [141] Meador, W. E.; Means, A. R.; Quiocho, F. A. Target enzyme rec-
perature on the dynamic behaviour of the HIV-1 nucleocapsid ognition by calmodulin: 2.4 A structure of a calmodulin-peptide
NCp7 and its DNA complex. J. Mol. Biol., 2002, 316, 611-627. complex. Science, 1992, 257, 1251-1255.
[119] Stevens, S. Y.; Sanker, S.; Kent, C.; Zuiderweg, E. R. Delineation [142] Li, M.; Smith, C. J.; Walker, M. T.; Smith, T. J. Novel inhibitors
of the allosteric mechanism of a cytidylyltransferase exhibiting complexed with glutamate dehydrogenase: allosteric regulation by
negative cooperativity. Nat. Struct. Biol., 2001, 8, 947-952. control of protein dynamics. J. Biol. Chem., 2009, 284, 22988-
[120] Huse, M.; Kuriyan, J. The conformational plasticity of protein 23000.
kinases. Cell, 2002, 109, 275-282. [143] Tomita, A.; Sato, T.; Ichiyanagi, K.; Nozawa, S.; Ichikawa, H.;
[121] Hyeon, C.; Jennings, P. A.; Adams, J. A.; Onuchic, J. N. Ligand- Chollet, M.; Kawai, F.; Park, S. Y.; Tsuduki, T.; Yamato, T.; Ko-
induced global transitions in the catalytic domain of protein kinase shihara, S. Y.; Adachi, S. Visualizing breathing motion of internal
A. Proc. Natl. Acad. Sci. USA, 2009, 106, 3023-3028. cavities in concert with ligand migration in myoglobin. Proc. Natl.
[122] Lou, H.; Cukier, R. I. Molecular dynamics of apo-adenylate kinase: Acad. Sci. USA, 2009, 106, 2612-2616.
a principal component analysis. J. Phys. Chem. B, 2006, 110, [144] Billeter, M.; Wagner, G.; Wuthrich, K. Solution NMR structure
12796-12808. determination of proteins revisited. J. Biomol. NMR, 2008, 42, 155-
[123] Whitford, P. C.; Miyashita, O.; Levy, Y.; Onuchic, J. N. Conforma- 158.
tional transitions of adenylate kinase: switching by cracking. J. [145] Adcock, S. A.; McCammon, J. A. Molecular dynamics: survey of
Mol. Biol., 2007, 366, 1661-1671. methods for simulating the activity of proteins. Chem. Rev., 2006,
[124] Faraldo-Gomez, J. D.; Roux, B. On the importance of a funneled 106, 1589-1615.
energy landscape for the assembly and regulation of multidomain [146] Senapati, S.; Bui, J. M.; McCammon, J. A. Induced fit in mouse
Src tyrosine kinases. Proc. Natl. Acad. Sci. USA, 2007, 104, 13643- acetylcholinesterase upon binding a femtomolar inhibitor: a mo-
13648. lecular dynamics study. J. Med. Chem., 2005, 48, 8155-8162.
[125] Yang, S.; Roux, B. Src kinase conformational activation: thermo- [147] Munoz-Ruiz, P.; Rubio, L.; Garcia-Palomero, E.; Dorronsoro, I.;
dynamics, pathways, and mechanisms. PLoS Comput. Biol., 2008, del Monte-Millan, M.; Valenzuela, R.; Usan, P.; de Austria, C.;
4, e1000047. Bartolini, M.; Andrisano, V.; Bidon-Chanal, A.; Orozco, M.; Lu-
[126] Levinson, N. M.; Kuchment, O.; Shen, K.; Young, M. A.; Koldob- que, F. J.; Medina, M.; Martinez, A. Design, synthesis, and bio-
skiy, M.; Karplus, M.; Cole, P. A.; Kuriyan, J. A Src-like inactive logical evaluation of dual binding site acetylcholinesterase inhibi-
conformation in the abl tyrosine kinase domain. PLoS Biol., 2006, tors: new disease-modifying agents for Alzheimer's disease. J.
4, e144. Med. Chem., 2005, 48, 7223-7233.
[127] Berteotti, A.; Cavalli, A.; Branduardi, D.; Gervasio, F. L.; Reca- [148] Kua, J.; Zhang, Y.; McCammon, J. A. Studying enzyme binding
natini, M.; Parrinello, M. Protein conformational transitions: the specificity in acetylcholinesterase using a combined molecular dy-
closure mechanism of a kinase explored by atomistic simulations. namics and multiple docking approach. J. Am. Chem. Soc., 2002,
J. Am. Chem. Soc., 2009, 131, 244-250. 124, 8260-8267.
Protein Flexibility and Ligand Recognition Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 209

[149] Lin, J. H.; Perryman, A. L.; Schames, J. R.; McCammon, J. A. [176] Verkhivker, G.; Rejto, P.; Bouzida, D.; Arthurs, S.; Colson, A.;
Computational drug design accommodating receptor flexibility: the Freer, S. T.; Gehlhaar, D.; Larson, V.; Luty, B.; Marrone, T.; Rose,
relaxed complex scheme. J. Am. Chem. Soc., 2002, 124, 5632- P. Parallel simulated tempering dynamics of ligand-protein binding
5633. with ensembles of protein conformations. Chem. Phys. Lett., 2001,
[150] Lin, J. H.; Perryman, A. L.; Schames, J. R.; McCammon, J. A. The 337, 181-189.
relaxed complex method: Accommodating receptor flexibility for [177] Goto, H.; Obata, S.; Kamakura, T.; Nakayama, N.; Sato, M.; Naka-
drug design with an improved scoring scheme. Biopolymers, 2003, jima, Y.; Nagashima, U.; Watanabe, T.; Inadomi, Y.; Ito, M.;
68, 47-62. Nishikawa, T.; Nakano, T.; Nilsson, L.; Tanaka, S.; Fukuzawa, K.;
[151] Alonso, H.; Bliznyuk, A. A.; Gready, J. E. Combining docking and Inagaki, Y.; Hamada, M.; Chuman, H. Drug discovery using grid
molecular dynamic simulations in drug design. Med. Res. Rev., technology. In Modern Methods for Theoretical Physical Chemis-
2006, 26, 531-568. try of Biopolymers; Starikov, E., Lewis, J. and Tanaka, S.; Eds.; El-
[152] Meagher, K. L.; Carlson, H. A. Incorporating protein flexibility in sevier: Amsterdam, 2006, pp 227-248.
structure-based drug discovery: using HIV-1 protease as a test case. [178] Liu, P.; Kim, B.; Friesner, R. A.; Berne, B. J. Replica exchange
J. Am. Chem. Soc., 2004, 126, 13276-13281. with solute tempering: a method for sampling biological systems in
[153] Brigo, A.; Lee, K. W.; Fogolari, F.; Mustata, G. I.; Briggs, J. M. explicit water. Proc. Natl. Acad. Sci. USA , 2005, 102, 13749-
Comparative molecular dynamics simulations of HIV-1 integrase 13754.
and the T66I/M154I mutant: binding modes and drug resistance to [179] Cheng, X.; Cui, G.; Hornak, V.; Simmerling, C. Modified replica
a diketo acid inhibitor. Proteins, 2005, 59, 723-741. exchange simulation methods for local structure refinement. J.
[154] Deupi, X.; Dolker, N.; Lopez-Rodriguez, M. L.; Campillo, M.; Phys. Chem. B, 2005, 109, 8220-8230.
Ballesteros, J. A.; Pardo, L. Structural models of class a G protein- [180] Kwak, W.; Hansmann, U. H. Efficient sampling of protein struc-
coupled receptors as a tool for drug design: insights on transmem- tures by model hopping. Phys. Rev. Lett., 2005, 95, 138102.
brane bundle plasticity. Curr. Top. Med. Chem., 2007, 7, 991-998. [181] Hritz, J.; Oostenbrink, C. Hamiltonian replica exchange molecular
[155] Fanelli, F.; De Benedetti, P. G. Computational modeling ap- dynamics using soft-core interactions. J. Chem. Phys., 2008, 128,
proaches to structure-function analysis of G protein-coupled recep- 144121.
tors. Chem. Rev., 2005, 105, 3297-3351. [182] Rodinger, T.; Howell, P. L.; Pomes, R. Calculation of absolute
[156] Berne, B. J.; Straub, J. E. Novel methods of sampling phase space protein-ligand binding free energy using distributed replica sam-
in the simulation of biological systems. Curr. Opin. Struct. Biol., pling. J. Chem. Phys., 2008, 129, 155102.
1997, 7, 181-189. [183] Woods, C.; Essex, J.; King, M. The development of replica-
[157] Hansmann, U. H.; Okamoto, Y. New Monte Carlo algorithms for exchange-based free-energy methods. J. Phys. Chem, B, 2003, 107,
protein folding. Curr. Opin. Struct. Biol., 1999, 9, 177-183. 13703-13710.
[158] Lei, H.; Duan, Y. Improved sampling methods for molecular simu- [184] Hamelberg, D.; McCammon, J. A. Accelerating conformational
lation. Curr. Opin. Struct. Biol., 2007, 17, 187-191. transitions in biomolecular systems. Ann. Rep. Comp. Chem., 2006,
[159] Mitsutake, A.; Sugita, Y.; Okamoto, Y. Generalized-ensemble 2, 221-232.
algorithms for molecular simulations of biopolymers. Biopolymers, [185] Voter, A. A method for accelerating the molecular dynamics simu-
2001, 60, 96-123. lation of infrequent events. J. Chem. Phys., 1997, 106, 4665-4677.
[160] Okamoto, Y. Generalized-ensemble algorithms: enhanced sampling [186] Markwick, P. R.; Bouvignies, G.; Blackledge, M. Exploring multi-
techniques for Monte Carlo and molecular dynamics simulations. J. ple timescale motions in protein GB3 using accelerated molecular
Mol. Graph. Model., 2004, 22, 425-439. dynamics and NMR spectroscopy. J. Am. Chem. Soc., 2007, 129,
[161] Trebst, S.; Troyer, M.; Hansmann, U. H. Optimized parallel tem- 4724-4730.
pering simulations of proteins. J. Chem. Phys., 2006, 124, 174903. [187] Grant, B. J.; Gorfe, A. A.; McCammon, J. A. Ras conformational
[162] Tozzini, V. Coarse-grained models for proteins. Curr. Opin. Struct. switching: simulating nucleotide-dependent conformational transi-
Biol., 2005, 15, 144-150. tions with accelerated molecular dynamics. PLoS Comput. Biol.,
[163] Neri, M.; Anselmi, C.; Cascella, M.; Maritan, A.; Carloni, P. 2009, 5, e1000325.
Coarse-grained model of proteins incorporating atomistic detail of [188] Yang, L.; Shao, Q.; Gao, Y. Q. Thermodynamics and folding
the active site. Phys. Rev. Lett., 2005, 95, 218102. pathways of trpzip2: an accelerated molecular dynamics simulation
[164] Czerminski, R.; Elber, R. Computational studies of ligand diffusion study. J. Phys. Chem. B, 2009, 113, 803-808.
in globins: I. Leghemoglobin. Proteins, 1991, 10, 70-80. [189] Laio, A.; Parrinello, M. Escaping free-energy minima. Proc. Natl.
[165] Sugita, Y.; Okamoto, Y. Replica-exchange molecular dynamics Acad. Sci. USA, 2002, 99, 12562-12566.
method for protein folding. Chem. Phys. Lett., 1999, 314, 141-151. [190] Biarnes, X.; Ardevol, A.; Planas, A.; Rovira, C.; Laio, A.; Par-
[166] Chipot, C.; Pohorille, A., Eds. Free Energy Calculations: Theory rinello, M. The conformational free energy landscape of beta-D-
and Applications in Chemistry and Biology. Springer, Berlin, 2007. glucopyranose. Implications for substrate preactivation in beta-
[167] Hamelberg, D.; Mongan, J.; McCammon, J. A. Accelerated mo- glucoside hydrolases. J. Am. Chem. Soc., 2007, 129, 10686-10693.
lecular dynamics: a promising and efficient simulation method for [191] Domene, C.; Klein, M. L.; Branduardi, D.; Gervasio, F. L.; Par-
biomolecules. J. Chem. Phys., 2004, 120, 11919-11929. rinello, M. Conformational changes and gating at the selectivity fil-
[168] Swendsen, R. H.; Wang, J. S. Replica Monte Carlo simulation of ter of potassium channels. J. Am. Chem. Soc., 2008, 130, 9474-
spin glasses. Phys. Rev. Lett., 1986, 57, 2607-2609. 9480.
[169] Earl, D. J.; Deem, M. W. Parallel tempering: theory, applications, [192] Fiorin, G.; Pastore, A.; Carloni, P.; Parrinello, M. Using metady-
and new perspectives. Phys. Chem. Chem. Phys., 2005, 7, 3910- namics to understand the mechanism of calmodulin/target recogni-
3916. tion at atomic detail. Biophys. J., 2006, 91, 2768-2777.
[170] Liwo, A.; Czaplewski, C.; Oldziej, S.; Scheraga, H. A. Computa- [193] Petraglio, G.; Bartolini, M.; Branduardi, D.; Andrisano, V.; Reca-
tional techniques for efficient conformational sampling of proteins. natini, M.; Gervasio, F. L.; Cavalli, A.; Parrinello, M. The role of
Curr. Opin. Struct. Biol., 2008, 18, 134-139. Li+, Na+, and K+ in the ligand binding inside the human acetyl-
[171] Garcia, A. E.; Onuchic, J. N. Folding a protein in a computer: an cholinesterase gorge. Proteins, 2008, 70, 779-785.
atomic description of the folding/unfolding of protein A. Proc. [194] Rohrig, U. F.; Laio, A.; Tantalo, N.; Parrinello, M.; Petronzio, R.
Natl. Acad. Sci. USA, 2003, 100, 13898-13903. Stability and structure of oligomers of the Alzheimer peptide
[172] Sanbonmatsu, K. Y.; Garcia, A. E. Structure of Met-enkephalin in Abeta16-22: from the dimer to the 32-mer. Biophys. J., 2006, 91,
explicit aqueous solution using replica exchange molecular dynam- 3217-3229.
ics. Proteins, 2002, 46, 225-234. [195] Behler, J.; Martonak, R.; Donadio, D.; Parrinello, M. Metadynam-
[173] Zhou, R. Trp-cage: folding free energy landscape in explicit water. ics simulations of the high-pressure phases of silicon employing a
Proc. Natl. Acad. Sci. USA, 2003, 100, 13280-13285. high-dimensional neural network potential. Phys. Rev. Lett., 2008,
[174] Zhou, R. Exploring the protein folding free energy landscape: 100, 185501.
coupling replica exchange method with P3ME/RESPA algorithm. [196] Di Pietro, E.; Pagliai, M.; Cardini, G.; Schettino, V. Solid-state
J. Mol. Graph. Model, 2004, 22, 451-463. phase transition induced by pressure in LiOH x H2O. J. Phys.
[175] Zhou, R.; Berne, B. J. Can a continuum solvent model reproduce Chem. B, 2006, 110, 13539-13546.
the free energy landscape of a beta-hairpin folding in water? Proc.
Natl. Acad. Sci. USA, 2002, 99, 12777-12782.
210 Current Topics in Medicinal Chemistry, 2011, Vol. 11, No. 2 Spyrakis et al.

[197] Donadio, D.; Bernasconi, M.; Boero, M. Ab initio simulations of [207] Michel, C.; Laio, A.; Mohamed, F.; Krack, M.; Parrinello, M.;
photoinduced interconversions of oxygen deficient centers in Milet, A. Free energy ab initio metadynamics: A new tool for the
amorphous silica. Phys. Rev. Lett., 2001, 87, 195504. theoretical study of organometallic reactivity? Example of the CC
[198] Prestipino, S.; Giaquinta, P. V. Liquid-solid coexistence via the and CH reductive eliminations from platinum(IV) complexes. Or-
metadynamics approach. J. Chem. Phys., 2008, 128, 114707. ganometallics, 2007, 26, 1241-1249.
[199] Zipoli, F.; Bernasconi, M.; Martonak, R. Constant pressure reactive [208] Schreiner, E.; Nair, N. N.; Marx, D. Influence of extreme thermo-
molecular dynamics simulations of phase transitions under pres- dynamic conditions and pyrite surfaces on peptide synthesis in
sure: The graphite to diamond conversion revisited. Eur. Phys. J. B, aqueous media. J. Am. Chem. Soc., 2008, 130, 2768-2770.
2004, 39, 41-47. [209] Trudu, F.; Donadio, D.; Parrinello, M. Freezing of a Lennard-Jones
[200] Karamertzanis, P. G.; Raiteri, P.; Parrinello, M.; Leslie, M.; Price, fluid: from nucleation to spinodal regime. Phys. Rev. Lett., 2006,
S. L. The thermal stability of lattice-energy minima of 5- 97, 105701.
fluorouracil: metadynamics as an aid to polymorph prediction. J. [210] Spiwok, V.; Lipovova, P.; Kralova, B. Metadynamics in essential
Phys. Chem. B, 2008, 112, 4298-4308. coordinates: free energy simulation of conformational changes. J.
[201] Martonak, R.; Donadio, D.; Oganov, A. R.; Parrinello, M. Crystal Phys. Chem. B, 2007, 111, 3073-3076.
structure transformations in SiO2 from classical and ab initio [211] Piana, S.; Laio, A. A bias-exchange approach to protein folding. J.
metadynamics. Nat. Mater., 2006, 5, 623-626. Phys. Chem. B, 2007, 111, 4553-4559.
[202] Martonak, R.; Laio, A.; Bernasconi, M.; Ceriani, C.; Raiteri, P.; [212] Bussi, G.; Gervasio, F. L.; Laio, A.; Parrinello, M. Free-energy
ZIpoli, F.; Parrinello, M. Simulation of structural phase transitions landscape for beta hairpin folding from combined parallel temper-
by metadynamics. Zeitschrift für Kristallographie 2005, 220, 489- ing and metadynamics. J. Am. Chem. Soc., 2006, 128, 13435-
498. 13441.
[203] Oganov, A. R.; Martonak, R.; Laio, A.; Raiteri, P.; Parrinello, M. [213] Gervasio, F. L.; Laio, A.; Parrinello, M. Flexible docking in solu-
Anisotropy of Earth's D'' layer and stacking faults in the MgSiO3 tion using metadynamics. J. Am. Chem. Soc., 2005, 127, 2600-
post-perovskite phase. Nature, 2005, 438, 1142-1144. 2607.
[204] Blumberger, J.; Ensing, B.; Klein, M. L. Formamide hydrolysis in [214] Bonomi, M.; Gervasio, F. L.; Tiana, G.; Provasi, D.; Broglia, R. A.;
alkaline aqueous solution: insight from Ab initio metadynamics Parrinello, M. Insight into the folding inhibition of the HIV-1 pro-
calculations. Angew. Chem. Int. Ed. Engl., 2006, 45, 2893-2897. tease by a small peptide. Biophys. J., 2007, 93, 2813-2821.
[205] Ensing, B.; De Vivo, M.; Liu, Z.; Moore, P.; Klein, M. L. Metady- [215] Ceccarelli, M.; Danelon, C.; Laio, A.; Parrinello, M. Microscopic
namics as a tool for exploring free energy landscapes of chemical mechanism of antibiotics translocation through a porin. Biophys. J.,
reactions. Acc. Chem. Res., 2006, 39, 73-81. 2004, 87, 58-64.
[206] Ensing, B.; Laio, A.; Gervasio, F. L.; Parrinello, M.; Klein, M. L. A [216] Branduardi, D.; Gervasio, F. L.; Cavalli, A.; Recanatini, M.; Par-
minimum free energy reaction path for the E2 reaction between rinello, M. The role of the peripheral anionic site and cation-pi in-
fluoro ethane and a fluoride ion. J. Am. Chem. Soc., 2004, 126, teractions in the ligand penetration of the human AChE gorge. J.
9492-9493. Am. Chem. Soc., 2005, 127, 9147-9155.

Received: December 1, 2009 Accepted: April 15, 2010

You might also like