You are on page 1of 51

Review

pubs.acs.org/CR

Cooperativity in Noncovalent Interactions


A. Subha Mahadevi and G. Narahari Sastry*
Centre for Molecular Modelling, CSIR-Indian Institute of Chemical Technology, Tarnaka, Hyderabad, India 500607
11. 3. Catalysis 2802
11. 4. Material Science 2803
12. Prospects and Outlook 2805
Author Information 2806
Corresponding Author 2806
Notes 2806
Downloaded via BENEMERITA UNIV AUTONOMA DE PUEBLA on May 13, 2019 at 14:25:16 (UTC).

Biographies 2806
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Acknowledgments 2806
References 2806

1. INTRODUCTION
CONTENTS After conquering the atomic structure about a century ago,
1. Introduction 2775 chemists have been largely interested in understanding the
2. Origin of Cooperativity 2777 concept of chemical bond and the formation of a molecule from
3. Hydrogen Bonding and Cooperativity 2778 its constituent atoms.1 Electrons, nucleus, as well as the nuclear
3. 1. Hydrogen-Bonded Clusters 2778 particles have their individual characteristics and give rise to
3. 2. Hydrogen Bonding in Peptides and Carbo- different types of elements, which make up our periodic table.
hydrates 2782 Molecules, which are formed from these atoms, possess their
3. 3. Hydrogen Bonding-Interplay with Other characteristically exquisite properties depending not only on the
Noncovalent Interactions 2783 type of atoms but also on the way in which they are connected.
4. Quantitative Definitions for Cooperativity 2784 These connections, known as chemical bonds, are at the heart of
4. 1. Many Body Analysis 2784 chemistry and provide the rational power to engage in a beautiful
4. 2. Double and Triple Mutant Analysis 2784 excursion toward the design of molecules through making and
4. 3. Measures for Estimation of Cooperativity in breaking bonds. Therefore, in the contemporary age whenever a
Noncovalent Interactions 2785 chemist, or for that matter any natural scientist, refers to a pure
4. 3. 1. Interaction Energy and Energy Decom- compound, such as benzene, water, or even more complicated
position Analysis 2785 structures such as taxol, fullerenes, etc., they refer to their
4. 3. 2. Spectoscopic Parameters 2786 molecular structures.
4. 3. 3. Structural Analysis 2786 Clearly, molecules are individual building blocks, and the way
5. How Does a Pair of Noncovalent Interactions in which they organize themselves have a great bearing on their
Mutually Influence Each Other? 2786 properties. We all know that the structure of most matter is
5.1. Cation-Induced Cooperativity 2786 determined by the way in which molecules associate with
5.2. Anion-Induced Cooperativity 2790 themselves, giving rise to different states of matter, and it is
6. Strong Manifestation of Weak Interactions 2791 particularly interesting that in the condensed phases, the
7. Interplay of π−π Interactions 2792 properties of molecules behave very differently depending on
7.1. With Other Noncovalent Interactions 2792 the temperature, pressure, volume, the presence of other
7.2. In Molecular Clusters 2793 molecules, interfaces with surfaces, environment, and under
8. Modulation of Cooperativity Due to Solvation 2794 varying electric or magnetic fields. In all cases, while the molecule
9. Cooperativity in Biochemistry 2795 as an entity is intact, the way in which noncovalent forces operate
9.1. Allostery and Binding Cooperativity 2795 with molecules of the same type or with other molecules or
9.2. Multimeric Proteins, Ion Channels, and surfaces is the key in determining these variations. Under-
Enzymes 2797 standing the nature of noncovalent interactions is thus extremely
9.3. Drug Receptor Interactions 2798 important to see what causes these variations in the proper-
9.4. Protein Folding 2799 ties.2−9 The most dramatic changes in the properties of the
9.5. Transcription 2800 molecules occur in condensed phases, in liquid and solid, as the
9.6. DNA/RNA 2800 way in which noncovalent interactions operate. Further, the way
10. Anticooperativity 2801 in which two molecules interact with each other are largely
11. Manifestation of Cooperativity 2801 governed by a combination of dispersion2,8 and electrostatic
11. 1. Self-Assembly 2801
11. 2. Chelate Cooperativity 2802 Received: June 30, 2014
Published: February 3, 2016

© 2016 American Chemical Society 2775 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

interactions. So much so that hydrogen bonding is one of the Noncovalent interactions were first recognized by J. D. van der
most popular noncovalent interactions and in a colloquial sense Waals in the later part of 19th century helping in reformulation of
is almost synonymous with noncovalent interactions.10−13 On the equation of state for real gases.2 Unlike covalent interaction
the basis of the structure of interacting molecules, cation−π which leads to the formation of a classical molecule, noncovalent
interaction,14−17 π−π stacking,18 CH···π interaction,18 anion−π interaction leads to the formation of molecular clusters. Some
interaction,19−23 halogen bond,24 lone pair interaction25 and so noncovalent interactions are also known to act at long distances
on and so forth have emerged (Figure 1). Indeed, the origin of of several angstroms unlike covalent bonds.3−7 For example, the
electrostatic, induction, and dispersion interactions are long-
range interactions, while the exchange-repulsion and charge-
transfer interactions are classified as short-range interactions.
The short-range interactions have their origin in the overlap of
molecular orbitals and do not act at long distances. Although
these interactions are traditionally considered to be weak, their
strength covers a substantial range, starting from a few kJ/mol to
several hundreds of kJ/mol depending on the type of interaction.
Earlier reports by Kollman have indicated how using the study of
noncovalent complexes indeed involves a fruitful interplay
between theory and experiment.3
Conventionally, from a chemical point of view, the nature of a
given chemical bond, for example a C−C bond, remains largely
unaffected by its surroundings.1 This is the basis for developing
molecular mechanics force fields, hybridization, and as such
giving an estimation of bond energies of a typical bond pair as
shown below.
ki ki
E (r N ) = ∑ (li − li ,0)2 + ∑ (θi − θi ,0)2
bonds 2 angles 2
Figure 1. Representation of the different kinds of noncovalent
interactions. Vn
+ ∑ [1 + cos(nω − γ )]
torsions 2 (1)
attraction in these interactions is electrostatic, induction, and
dispersion interactions. While most of these interactions are in Signature of a typical covalent bond is that its bond length,
general weaker than covalent interactions, the strength of some bond strength, as well as the spectroscopic signatures are usually
of the bonds such as hydrogen bond, cation−π, and anion−π can characteristic, and any modulations in their strength by
have a varied range and in some particular instances might be neighboring covalent bonds tend to be a small percentage of
even slightly stronger than a weak covalent bond. its absolute value. Thus, in general, variations in the covalent

Figure 2. Manifestation of cooperativity in highly varied fields.

2776 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

bonds are not very dramatic. However, noncovalent interactions assemblies, catalysis, and biological function is taken up. An
can vary remarkably. Let us consider the C−C single bond in extensive review of the available literature on cooperativity also
ethane and propane which are comparable in their length and reveals how cooperative behavior has aided substantially toward
strength which remains largely so in the congeneric series of understanding the conformational transitions and allosteric
alkanes. However, let us take a dimer of water connected through interactions in proteins. Cooperativity is thus a concept of vital
a hydrogen bond and compare it with a trimer, tetramer, and a importance in chemistry, biology, material science, and allied
larger polymer. It has been unambiguously established that the areas.26−36 The synergistic interplay between different kinds of
average hydrogen bond strength dramatically increases, with the noncovalent interactions is crucial to maintain the structure of
concomitant shortening of hydrogen bond length, becoming important biomacromolecules like DNA and proteins and in
almost double when it reaches a decamer.13 Thus, the way in retaining the fidelity of information processing needed for
which the noncovalent interactions interact with themselves is normal life. In the field of chemistry too, cooperativity of
certainly not additive. Hydrogen bond strength ranges from a noncovalent interactions has an essential role in template-
very weak interaction to somewhat comparable with a weak directed synthesis, the transmission of stereochemical informa-
covalent bond. Similarly, the strengths of cation−π and π−π tion, and in determining the structure and properties of
interactions can be altered by two to three times, higher or lower materials.39−43
depending on the environment. One more important aspect in
any given process of aggregation is the large number of 2. ORIGIN OF COOPERATIVITY
noncovalent interactions that operate simultaneously. Thus, An explanation of the origin of cooperativity is helpful for
noncovalent interactions are a topic of outstanding importance, understanding various cooperative effects. The cooperativity
in all areas of chemistry, material science, biology, and medicine. reported in this review can be classified into the following types
The additive scheme used in eq 1 for covalent bonds based on origin: (a) many body interaction, (b) secondary
encounters utter inadequacies when applied to systems governed interaction, (c) chelate effects, and (d) cooperativity and
by noncovalent interactions. The underlying reason for this anticooperativity induced by conformation change.
failure is well-known nonadditivity. This nonadditivity arises due The many body interaction is a major source of the
to the cooperativity or anticooperativity of noncovalent cooperativity in the clusters of small molecules and is the origin
interactions. In the current review, we try to describe the of hydrogen bond cooperativity. It not only considers the sum of
concept of cooperativity among noncovalent interactions.26−32 pairwise interactions but also includes three-body terms, four-
When a pair of noncovalent interactions strengthen each other, it body terms, and so on. The contribution of different
is called cooperative while when they weaken each other they are intermolecular forces to many body interaction is very well
defined as operating in an anticooperative manner. Cooperativity summarized in the book, The Theory of Intermolecular Forces by A.
implies that the sum of at least two interactions is larger than the J. Stone.44 The main contributions to forces between molecules
simple addition of the individual interactions. The terms may be classified into long-range and short-range. Three kinds of
cooperative and anticooperative have different origins and long-range effects are mentioned below. The electrostatic effect
meanings in different fields28,33−36 and are extensively used in arises from the classical interaction between the static charge
biology (Figure 2). Although the qualitative concept of distributions of the two molecules. They are strictly pairwise
cooperativity is quite clear, quantitative experimental measures additive and may be either attractive or repulsive in nature.
barring some spectroscopic approaches are rather scarce. Dispersion refers to the attractive term of the vdW equation. This
Computational quantum chemical methods provide the most attractive part of vdW interactions is an electron correlation
relevant means of probing and quantifying cooperativity. It is (quantum mechanical many-body) effect known as London
interesting to examine the strength and limitations of quantum dispersion.9 The contribution of dispersion to the many body
chemical methods and the dependence of the cooperativity interaction is small. Induction effects arise from the distortion of
measure on method, basis set employed, basis set superposition a particular molecule in the electric field of all its neighbors and
error, etc. Knowledge of how noncovalent interactions manifest are always attractive. As the fields of several neighboring
themselves in small molecular clusters and the quantification of molecules may reinforce each other or cancel out, induction is
their cooperativity and anticooperativity thus appears to be strongly nonadditive.44 Because the origin of the induction
essential for comprehension of supramolecular assembly.37,38 interaction is the induced polarization by electric field, the many
In the current review, we start with a discussion on the origin of body interaction is particularly significant in cases where the
cooperativity. Historically, cooperativity has been studied interacting molecule or ion has a strong electric field such as the
extensively in hydrogen-bonded systems and is thus discussed Mg2+ ion. Another vital aspect of the many body effects is the
as a prelude to the section on its quantification and measurement. modification of intermolecular forces between molecules in the
From a fundamental point of view, it is interesting to examine presence of a solvent.
how a pair of noncovalent interactions mutually influences each The secondary interaction is another important source of
other. In addition to hydrogen bonding, noncovalent inter- cooperativity.45−51 The secondary electrostatic interactions
actions such as cation−π, anion−π, and other weak interactions rationalize the relative stabilities of multiple hydrogen-bonded
have a strong manifestation in controlling supramolecular complexes. In multiple hydrogen-bonded complexes, the
structure and function. Among the noncovalent interactions, arrangement of the hydrogen bond donor (D) and acceptor
cation−π may be argued to be stronger, while π−π and CH−π (A) decides the total stability of the complex. Using molecular
interactions are weaker compared to hydrogen bonding. We mechanics calculations on nucleic acid base pairs, Jorgensen et al.
divide the review into sections on cooperativity in hydrogen- elucidated that secondary interactions between neighboring
bonded clusters, cation−π induced cooperativity, anion−π heteroatoms can contribute additional stabilization, provided
induced cooperativity, and cooperativity among weak inter- that hydrogen bond donors with their positive partial charges are
actions. Finally, the impact of cooperativity in understanding present in one molecule and acceptors in the other molecule
various phenomena such as formation of supramolecular (DD, AA).45−47 Repulsion between diagonally placed acceptor
2777 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

and donor functions will destabilize the primary hydrogen bond quite a remarkable statement, and spectroscopic advances and
when both molecules contain alternating sequences (e.g., AD, computational methods have enabled one to quantitatively
DA, etc.). Jorgensen et al. also reported the importance of estimate the extent of cooperativity after almost 3−4 decades
secondary electrostatic interactions due to polar functional after the observation. Dannenberg and co-workers have
groups located closely to hydrogen bonding sites.46 Studies by undertaken a series of studies on amide clusters and ascertained
Uchimaru et al. revealed that although the main source of the the cooperativity in these hydrogen-bonded clusters.65−67
stabilization for hydrogen-bonding association was the electro- Besides, a plethora of experimental and computational studies
static contribution, the secondary electrostatic and polarization which evaluate cooperativity in hydrogen-bonded clusters have
interactions due to polar functional groups located closely to the also been reported over the last two decades.68−151 The areas
hydrogen bonding sites also significantly alter the magnitude of which have witnessed the liberal use of the word cooperativity in
hydrogen-bonding stabilization.48 Uchimaru and co-workers also the context of hydrogen bonding include delineation of peptide
reported that the hydrogen bond donor−acceptor orientation and carbohydrate structure152−176 and also the concept of
had an important role in cooperativity.49 Their case study in resonance-assisted hydrogen bonding.177−179
three iso-complexes of C8H9N5O2 showed that when a hydrogen 3. 1. Hydrogen-Bonded Clusters
bond is formed at the neighboring hydrogen bond site, the
intrinsic hydrogen bond forming abilities of hydrogen bonding Water clusters have remained as the most favored model systems
sites in the DA type molecules are increased, while those in the while investigating the intermolecular forces that act among
AA and DD type molecules are decreased. them, especially with regard to nonadditive interaction
Another important source of cooperativity is the chelate energies.11,68−95 Beginning in the late 1960s Morokuma and
effect.28,52−54 The loss of the entropy of translation which is Pedersen68 and Kollman and Allen69 were among the first to
associated with the cluster formation increases Gibbs free energy. carry out ab initio Hartree−Fock calculations for a pair of
If molecules can bind by two interactions, the formation of the interacting water molecules. In 1970, Moskowitz and co-workers
second binding does not accompany the loss of the entropy of employed SCF calculations to investigate the potential of
translation. This is the origin of chelate effect. Williams and co- interaction for dimers and trimers of water molecules.70 The
workers52,53 have extensively discussed the general relevance of study clearly revealed how three-molecule nonadditivities are
enthalpy/entropy compensations to binding interactions of large in magnitude and vary in sign depending on the hydrogen
biological importance in particular with respect to the binding of bond pattern involved. Huyskens reported a study on factors
agonists versus antagonists to a common receptor site. The governing the influence of a first hydrogen bond on the
chelate effect is one of the important sources of the stability of β- formation of a second one by the same molecule or ion.11
sheet structure, in which many hydrogen bonds exist. Several experimental studies have addressed the structure and
The major source of cooperativity in biological systems is one spectra of small water clusters, ice, as well as larger three-
which is induced by conformational change. Typical examples of dimensional water clusters.43,71−74,93 A few experimental studies
cooperativity in biology are the binding of oxygen to hemoglobin, which were employed to study hydrogen bond cooperativity are
wherein binding at each of its four binding sites increases the mentioned below. A study of self-association and hydrogen
oxygen affinity of the other sites, and in the folding of proteins bonding of 3,4,5-trichlorophenol with water using matrix-
and nucleic acids that are characterized by sharp melting isolation FT-IR spectroscopy provided experimental proof for
transitions.26 Indeed the origin of this kind of allosteric the dominating role of the induction energy term in the three
cooperativity is quite different from the origin of cooperativity body effect.77 Saykally and co-workers reported the quantifica-
found in clusters of small molecules bound by intermolecular tion of hydrogen bond cooperativity in the water by measuring
forces. the far-infrared vibration−rotation tunnelling spectrum of the
The hydrophobic interaction55 has its origin in the change of perdeuterated water tetramer.43 Pate and co-workers have
the structure of liquid water associated with the hydrophobic employed broadband rotational spectroscopy on water clusters
molecules. A study by Chandler56 suggests that water molecules produced in a pulsed molecular jet expansion to determine the
can still form a hydrogen bond network around a small solute oxygen atom geometry in three isomers of the nonamer and two
molecule, although it is constrained by the presence of the solute. isomers of the decamer.75 The cooperativity effects revealed by
The solvent entropy is reduced while the enthalpy remains the hydrogen bond O−O distance variations in this study were
unaffected. The hydrophobic force has long been considered as shown to be consistent with a simple model for hydrogen
the major driving force of protein folding.57 When a protein bonding in water that takes into account the cooperative and
folds, the hydrophobic side chains are embedded within the anticooperative bonding effects of nearby water molecules.
protein, leaving the hydrophilic side chains at the surface. Schmidt and co-workers reported an alternative interpretation of
the structure of the IR vibrational mode [υ(OH) band] of pure
3. HYDROGEN BONDING AND COOPERATIVITY water.76 The reinterpretation was based on the influence of the
Hydrogen bonding has been a subject of extensive study, and cooperative hydrogen bonding arising from a network of
several key reviews are available in literature revealing its wide- hydrogen bonds in the liquid. The insights obtained by foregoing
ranging importance.58−64 In most cases, hydrogen-bonded experimental studies have generated a lot of interest among
clusters have been employed to model cooperativity. Historically, theoretical and computational chemists.
the first mention (to our knowledge) of cooperativity can be seen Ab initio studies for the ground states of the linear water dimer
in the paper of H. S. Frank and W. Wen in Discussions of Faraday with Cs symmetry and cyclic water tetramer with S4 symmetry
Society.10 The authors made a dramatic observation that “the were undertaken by Lesyng and co-workers.79 Employing a
formation of hydrogen bonds in water is predominantly a cooperativity parameter based on the two-body, non-neighbor
cooperative phenomenon, so that, in most cases, when one bond interaction energy, plus three- and four-body contributions,
forms, several (perhaps “many”) will form, and when one bond including one-body deformation terms in relation to the total
breaks, then, typically, a whole cluster will “dissolve” ”. This is interaction energy of the water tetramer, they demonstrated an
2778 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 3. Pictorial representation of four different kinds of arrangements of water clusters (W1D, W2D, W2DH, and W3D clusters) employed to
evaluate cooperativity as a function of arrangement of individual molecules in a cluster of particular size. Edited and reprinted from ref 84. Copyright
2010 American Chemical Society.

energy gain of 29% based on cooperativity in the S4 water energy ratios ([H2O10:2], [H2O20:10]) revealed that the strength
tetramer with the MP3/6-31G** approximation. Suhai convinc- of complexation energy increases by 33, 38, 35, and 91%,
ingly elucidated that hydrogen bonding in ice is a highly respectively, as cluster size increases from 2−10 in four different
cooperative phenomenon where the optimization of the kinds of arrangements analyzed. As the cluster size increases from
structure of the infinite water polymer at the MP2 level yielded 10−20; however, the increase in strength of hydrogen bonding is
a relative enhancement of 47% over the corresponding dimer in the range of 4−6% only. Hence the nonadditivity of hydrogen
value.80 Further, the study revealed that the cohesive energy of bond strength or the associated cooperativity was much more
ice results from a delicate balance between different repulsive and evident as cluster size increased from 3−10 rather than from 10−
attractive terms in third and fourth order, which exhibit different 20, where the augmentation of hydrogen bond strength is only
long-range behaviors. Xantheas studied the significance of all marginal for all the four arrangements. Thus, the presence of a
higher-order components for interaction energy, in particular, definite amount of cooperativity in hydrogen-bonded water
the three-body term among the nonadditive terms for water clusters where the addition of a subsequent monomer to a
clusters ranging in size from clusters trimer through pentamer.81 hydrogen-bonded cluster augments the strength of existing
Hydrogen-bonding networks wherein donor−acceptor arrange- interactions was inferred. The mode of arrangement, in
ments existed between all water molecules were associated with particular, how these interactions were arranged was shown to
the largest nonadditivities among other networks present in low- play a crucial role on the strength of interaction.
lying minima of small water clusters. Yáñez and co-workers Application of combined infrared spectroscopy and DFT
performed an ab initio study on water trimers and identified a approaches by Ohno et al. demonstrated how the formation of
global minimum which corresponded to an asymmetric cyclic one hydrogen bond in a hydrogen-bonded water chain
structure.82 This structure presented significant cooperative cooperatively enhances or diminishes the strength of another
effects compared to a Cs dimer as reflected in several parameters hydrogen bond.85 An observation of how cooperativity of the
such as a stiffer intermolecular potential, shorter O−O distances, hydrogen bonding of water molecules affects the corresponding
longer donor O−H bond lengths, larger energies per hydrogen OH stretching bands was made. Quantum calculations
bond, and greater shifts of the donor O−H bond stretching performed on a series of water clusters in order to mimic water
frequencies than the Cs dimer. Studies on hydrogen bonding in molecules found in restricted environment unlike bulk water
phenol, water, and phenol-water clusters by Subramanian and co- show that cooperative effects must be taken into account in the
workers have demonstrated the enhanced stability on account of treatment of hydrogen bonds and water clusters in such bounded
hydrogen bonding, although for clusters with a similar hydrogen- systems.86 Ruckenstein and co-workers focused on the role of
bonding pattern, intermolecular interaction in phenol clusters many-body interactions on the structure of ordinary ice and
was slightly stronger than in water clusters.83 liquid water revealing that 62−63% of hydrogen bonds must be
The impact of varying arrangements of water molecules in broken to disintegrate a large hexagonal ice piece used to
different kinds of clusters as a function both of size of cluster (n = simulate water molecules in the form of a cube into small
2−20) and method of computation used have been studied by clusters.87 Stokely et al. performed a study on the low-
our group using extensive density functional theory (DFT) and temperature phase behavior of liquid water by combining
ab initio calculations (Figure 3).84 These model systems served mean field calculations and Monte Carlo simulations.88 While
to explain associated cooperativity in hydrogen bonding. The emphasizing key physical quantities that determine scenarios
ratio of complexation energy per hydrogen bond from decamer which describe water, they report that it is the amount of
to dimer [H2O10:2] and eicosamer to decamer [H2O20:10] was cooperativity in relation to the strength of the directional
calculated to quantify the strength of hydrogen bond with component of the hydrogen bond that establishes which scenario
increase in cluster size and was used as an indicator of prevails. A recent study explores the impact of hydrogen bond
cooperativity that is seen in the water clusters. The complexation cooperativity in eight low lying water hexamers.89 Stabilizing
2779 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

cooperativity observed in linear hydrogen-bonded water systems subjected to free optimization without any emphasis on the kind
diminished as clusters move from nearly planar to three- of order in the structure that is attained on optimization. The
dimensional structures. Water molecules which donated both interaction energy per monomer increased from dimer to 15mer
hydrogens to form double-donating interactions had increased by 90% in the case of circular arrangement, by 76% in the case of
stabilization, whereas waters which accepted two hydrogen linear arrangement, and by 34% in the case of a third standard
bonds experienced a decrease in stabilization due to cooperative arrangement, respectively. The cooperativity in hydrogen
and anticooperative effects. bonding for acetamide clusters was evaluated by considering
Studies on hydrogen-bonded clusters are not limited to water the variation in geometrical parameters including bond distances
clusters alone. 116−148 Dannenberg and co-workers have and electron density values at the bond critical point in the
contributed substantially to quantifying hydrogen bond different model systems. Similarly, DFT calculations were
cooperativity by performing quantum chemical calculations on performed on four arrangements of formamide clusters
several molecular clusters.65−67,101−107 Starting with ab initio and [HCONH2]n, (n = 1−10) linear, circular, helical, and stacked
semiempirical studies on 1,3 diones,65 dimers, and trimers of forms.109 These studies reveal maximum cooperativity, based on
acetic acid66 and acetylacetone67 to evaluate the strength of interaction energy per monomer, in the stacked arrangement
hydrogen bonding, they further investigated how cooperative followed by the circular, helical, and linear arrangements,
interactions dictate the hydrogen-bonding structure. Employing respectively. In all arrangements, an increasing trend in
DFT calculations on chain and ribbonlike arrangement of cooperativity of hydrogen bonding was observed as a function
molecules, they demonstrated a high degree of cooperativity in of increasing cluster size. Parra and co-workers reported DFT
hydrogen-bonded chains of urea and formamide mole- studies on intermolecular bifurcated hydrogen-bonding inter-
cules.101−103 Linear chains of hydrogen-bonded formamide actions in diformamide chains110 and carbonic acid clusters,111
molecules containing from 3 to 15 monomeric units revealed a showing the existence of significant cooperative effects in a linear
cooperative effect, wherein the strongest hydrogen bonds, network of three-center bifurcated hydrogen bonds as well as
typically those nearest the center of the formamide chain, have two-dimensional ringlike networks. There are two types of
a bond strength which approaches 200% that of the dimer.104 bifurcated three-center hydrogen bond interactions they have
Further, vibrational frequencies of the coupled N−H, CO explored: (a) one that involves a hydrogen atom and two
stretches, and C−N stretch/CNH bend revealed clear shifts that acceptor atoms denoted as A1HA2 and (b) one that involves one
reflect the cooperative stabilization of the hydrogen bonds.105 acceptor atom and two hydrogen atoms denoted as H1AH2.
We have performed a structural and energetic comparison of Positive cooperativity observed in these studies helps to
linear, circular and standard arrangements of (acetamide)n rationalize the common occurrence of three-center H bonds in
clusters (n = 1−15) at the B3LYP/D95** level of theory to the crystals structures of many molecular systems. Cooperativity
reveal significant cooperativity of hydrogen-bonding and size- in intramolecular bifurcated hydrogen-bonding interactions has
dependent structural preferences (Figure 4).108 The standard also been reported by them.112 Ludwig et al. analyzed the
arrangement in this study is defined as one where the maximum temperature dependence of NMR chemical shifts and quadru-
number of hydrogen-bonding interactions are found while being pole coupling constants in neat N-methylacetamide using
experimental and ab initio quantum cluster equilibrium (QCE)
theory.113 Strong cooperative effects were found in the linear N-
methylacetamide molecular clusters (n = 3−5) as reflected in the
chemical shifts and quadrupole coupling values for each species.
Theoretical calculations on smaller (HCN)n clusters where n =
1−4 at the RHF and MP2 level and for larger clusters where n =
5−7 at the RHF level of theory have shown that large cooperative
effects of hydrogen bonding are reflected in increasing hydrogen
bond energies, decreasing intermolecular separations, increasing
average dipole moments and in C−H stretching frequencies
shifts and the corresponding intensities as a consequence of
increasing cluster size.114 Theoretical computations on systems
consisting of up to four hydrazine molecules illustrate a
significant contribution of cooperative phenomena to the
interaction energy, amounting to as much as 12% of the overall
interaction energy.115 DFT calculations to examine the effect of
hydrogen bond cooperativity on the magnitude of the NMR
chemical shifts and spin−spin coupling constants in a C4h-
symmetric guanine-quartet and in structures consisting of six
cyanamide monomers have been reported.116 The study showed
that the magnitude of the NMR properties along the hydrogen
bond network, for example the |1JNH| coupling and 1H and 15N
chemical shifts of the hydrogen-bonding amino N−H group and
the |h2JNN| trans-hydrogen bond coupling, increased for
Figure 4. Two factors considered while studying cooperativity in
acetamide clusters (ref 108), size of clusters and difference in types of structures containing a larger number of monomers. Wang and
arrangements for clusters of the same size. BSSE corrected interaction co-workers have proposed a rapid method to predict the
energy per hydrogen bond (in kcal/mol) calculated employing the hydrogen bond cooperativity in long N-methylacetamide chains
B3LYP/D95** level of theory with dispersion correction is indicated in containing up to 200 monomers.117,118 It is based on parameters
the figure for selected clusters. obtained from fittings to the hydrogen-bonding energies in N-
2780 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 5. Representation of (a) different water, formamide, and acetamide clusters employed to understand modulation of hydrogen bond cooperativity
in the presence of H+, Na+, Mg2+, OH−, and Cl− ions using DFT calculations (ref 131). (b) Plot showing BSSE uncorrected sequential binding energy
values obtained for one particular arrangement of water molecules named as W1D with five different ions and increasing cluster size.

methylacetamide chains containing 2 to 7 monomeric units H···π) in a cyclic complex involving one acetylene and two water
generated using ab initio calculations. A DFT study has revealed molecules has been studied.128 The total interaction energy of
significant contribution of cooperativity in the interactions of the three hydrogen bonds was shown to increase as the number
hydrazoic acid clusters consisting of up to four monomers.119 of methyl group substitutions in the complex increases. MP2 and
Scheiner’s group probed into the underlying nature of the DFT studies on linear (urea)n=3−10 clusters revealed how
CH···O interaction using the ab initio methods.120−124 They cooperativity effects significantly enhance the N−H···O hydro-
thoroughly explored the cooperativity aspect of hydrogen bonds, gen bond from its strength of −7.90 kcal/mol in a dimer to
wherein a chain of n hydrogen-bonded molecules is held together −11.39 kcal/mol in a decamer.129 Gadre and co-workers recently
more strongly than would be expected based on the energetics of reported the cooperative contribution toward hydrogen bonding
the single hydrogen bond within a dimer.122 As mentioned earlier in para-substituted calix[n]arenes (CX[n]) (n = 4, 5) and their
in the section on origin of cooperativity, hydrogen bond thio analogues.130 Cooperativity was found to be nearly 5 times
cooperativity is typically attributed in large measure to the larger in the parent structure than that to its thio analogue.
polarization induced in each subunit by the presence of its More recently, work done in our group evaluated the impact
hydrogen-bonding partner. Scheiner and co-workers focused on that the presence of ions, such as Mg2+, Na+, H+, Cl−, and OH−,
evaluating the ability of one hydrogen bond in a chain to affect has on hydrogen-bonded clusters of increasing size (water,
others by comparing the CH···O bonds in (H2CO)n and formamide, and acetamide [n = 1−10]) in the context of their
(HFCO)n to the OH···O bonds in (H2O)n.120,121 Although the associated cooperativity using DFT calculations (Figure 5).131
degree of cooperativity is generally proportional to the strength The presence of an ion provides contrasting insight into the
of the hydrogen bond, the CH···O bonds in (HFCO)n were evaluated sequential binding energies of hydrogen-bonded
shown to display a disproportionately high degree of clusters of different sizes. Dramatically higher sequential binding
cooperativity. It was estimated that the mean hydrogen bond energies were seen in the presence of ions initially for smaller-
energy in an infinite chain of H2CO molecules was 25% greater sized molecular clusters compared to parent clusters. With
than the same quantity in a dimer, while the long water chain increasing cluster size, the difference in the sequential binding
exhibits a 66% enhancement over (H2O)2. Inspite of containing energies between ionic clusters and parent clusters becomes
substantially weaker individual hydrogen bonds than those in reduced. However, Mg2+-bearing clusters continued to exhibit
water chains, the study revealed that (HFCO)n manifests an substantially higher sequential binding energies and coopera-
energetic cooperativity that is nearly as large as that of the OH··· tivity, even with increasing cluster size. Monovalent ions had a
O congeners. Cooperativity between the O−H···O and C−H··· reduced impact on hydrogen bond cooperativity from hexamer
O hydrogen bonds have been investigated by quantum chemical onward, and this may be indicative of the short-range over which
calculations.125 The interaction energies of the O−H···O and C− their influence prevails. The study helped reinforce the extent of
H···O hydrogen bonds were increased by 53% and 58% impact that divalent cations have on hydrogen-bonded chains as
respectively, demonstrating the presence of large cooperativity. against monovalent ions.
Grabowski and co-workers reported a study on H2CO···(HF)n Zabardasti et al. reported presence of anticooperativity in
(n = 1, ..., 9) complexes using the MP2 method and showed how hydrogen-bonded clusters in two separate studies.132,133 Ab
cooperativity effect significantly enhances F−H···O hydrogen initio calculations on dihydrogen-bonded clusters of BeH42− with
bond.126 Chen and co-workers employed energy decomposition 1−4 molecules of NH3 revealed how cooperative effect
approaches to quadruple and double hydrogen-bonded model decreased with the increasing size of the clusters.132 In another
systems to emphasize how cooperativity is the most important study employing ab initio and DFT calculations on hydrogen-
factor determining stability of the complex particularly in bonded clusters of water cyanuric acid, the presence of both
quadruple hydrogen bond dimers.127 The cooperativity between cooperative and anticooperative effects were seen depending on
three types of hydrogen bonds (O−H···O, C−H···O, and O− the geometry of the structures.133 Albrecht et al. considered the
2781 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 6. Selected crystal structures showing coexistence of different noncovalent interactions along with corresponding references where these
structures are reported. (a) Hydrogen bonding, π−π, and halogen interactions as seen in ref 182, (b) hydrogen bonding and π−π interactions noted in
refs 375, 372, and 373, (c) anion−π and π−π interactions seen in ref 302, (d) cation−π and anion−π interactions from ref 19, and (e) π−π interactions
and halogen bonds observed in refs 381 and 382.

changes in atomic energy in the clusters versus the isolated series. Using theoretical studies, Wu et al. clearly established the
monomer for small clusters of methanol, water, and form- existence of significant cooperativity in the formation of 310 and
aldehyde. A variety of stabilities were observed within these α-helices, whereas no cooperativity was found in the formation of
hydrogen-bonded clusters, including indications of cooperative β-strands and 27-ribbons for a series of polyglycine models
and anticooperative interactions.134 containing up to 14 amino acid residues.163 The absence of
3. 2. Hydrogen Bonding in Peptides and Carbohydrates significant cooperativity in terms of enthalpy contribution in β-
sheets was further shown by employing a model system
Hydrogen bonds are the most important interactions between consisting of a dimer of a tripeptide to obtain repeating units
amino acid residues in peptides and proteins as they provide the for the parallel and antiparallel β-sheets.164 Baker and co-workers
stability for secondary structures like helices or sheets. Several dwelt on the origin and relative importance of the contributions
theoretical studies using α helix and β sheet model systems have to helical cooperativity.165 They also reported the role of
been performed to understand their inherent cooperativ- cooperative hydrogen bonding in amyloid formation.166
ity.153−173 Kemp et al. used the experimental demonstration of Dannenberg and co-workers performed ONIOM (DFT/
an enthalpic component to the cooperativity of α-helical peptides AM1) calculations on capped parallel-β-sheets of acetyl-
as evidence for hydrogen bond cooperation in these VQIVYK-NHCH 3 , acetyl-Q-NHCH 3, and acetyl-alanine-
structures.162 This was done considering the helicity measured NHCH3 as model systems to study the importance of hydrogen
at different temperatures in water for a solubilized polyalanine bonding between glutamine side chains to the formation of
2782 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

amyloid VQIVYK parallel β-sheets.167,168 The Q sheets in these reported.183 On the basis of a reductionist approach for the
studies exhibit cooperativity that is reminiscent of chains of construction of a micelle, Fernández and co-workers have shown
formamides104,105 and 4-pyridones,107 where the interactions how highly directional hydrogen bonds build a frame on top of
become stronger as the chain or sheet grows. Using both classical which the dispersive forces give the aggregate its final shape.184
electrostatic methods and DFT methods, the energetics of fibril Lesarri and co-workers have determined the structures for
formation for the first three layers was shown to be cooperative. phenol dimer and trimer through the use of chirped pulse Fourier
Ireta et al. emphasized the importance of cooperativity in the transform microwave spectroscopy in the 2−8 GHz band.185 The
stability of α-helices.169 Two limiting cases were considered, one dimeric structure was found to represent a case, where an
of an isolated hydrogen bond and the other of an infinite α helical interplay between dispersion and hydrogen bonding played an
chain for DFT studies. Cooperativity within an infinite network essential role in fixing the complexation geometry.
of hydrogen bonds was clearly shown to strengthen each A few studies also explore cooperativity of the dihydrogen
individual bond by more than a factor of 2. bond with other noncovalent interactions.150,151,186,187 Ren and
Jensen and co-workers reported a computational methodology co-workers describe a cooperativity effect between the
for backbone amide proton chemical shift (δH).170 Interestingly, dihydrogen bonding and H−M···π interactions (M = Li, Na,
cooperative hydrogen-bonding effects were found to have a and K) in ternary complexes of FH···HM···C2H2/C2H4/C6H6
significant impact on δH values by affecting the primary hydrogen using the DFT and MP2 level calculations.186 Analyses of the
bond geometry and polarizing the electron density around the charge of the hydrogen atoms in H···H moiety, atoms in
amide proton. Magistrato and co-workers employed DFT and molecule (AIM) analysis, and electron density shifts methods
hybrid DFT/MM simulations both in vacuum and in aqueous were adopted, and the cooperativity effect of the dihydrogen
solution on model polyglutamine β-sheet structures character- bond on the H−M···π interaction was shown to be more
istically observed in Huntington’s disease.153 Cooperativity of pronounced than that of the M···π bond on the H···H
glutamine side chains was shown to affect both directions, interaction. The coexistence of both dihydrogen bonding and
perpendicular and parallel to the backbone. The unusual metal−σ interaction was explored by Grabowski and co-workers
behavior observed in these β-sheets provided significant extra- in the case of three aggregates H2···LiH···H2, H2···NaH···H2, and
stabilization of polyglutamine aggregation. Schreiber and co- H2···HBeH···H2 from among a series of complexes formed by
workers evaluated direct and cooperative contributions while hydrogen with metal hydrides.187 They however report that the
studying the strength of buried hydrogen bonds and salt modulus of the cooperativity energy is not greater than 0.05 kcal/
bridges.171 Employing a modifed multiple mutant cycle protocol mol, suggesting that no significant cooperative effect is observed
to selected interactions between TEM-1-β-lactamase and its in these instances.
protein inhibitor BLIP, they demonstrated how formation of In recent years, several studies on the interplay between
network of interactions helps establish cooperative effects which hydrogen bonding and the halogen bond have been
can convert even unfavorable interactions to favorable ones, reported.188−195 Li and co-workers have performed ab initio
more so for salt bridges and to a lesser extent for hydrogen bonds. calculations on H3N···XY···HF triads (X, Y = F, Cl, and Br), each
Dashnau and co-workers used molecular dynamics (MD) having a halogen bond and a hydrogen bond.191 An analysis of
simulations combined with water−water hydrogen bond angle molecular geometries, binding energies, and infrared spectra of
analysis, calculation of solvent accessible surface area, and monomers, dyads, and triads to examine cooperative effects
approximate free energy of solvation to show how intramolecular indicate significant cooperativity between the halogen and
hydrogen bond cooperativity was closely associated with changes hydrogen bonds in these complexes. Further effect of a halogen
in water structure surrounding the aldohexopyranose stereo- bond on a hydrogen bond was more pronounced than that of a
isomers.174 Gadre and co-workers report a direct estimation of hydrogen bond on a halogen bond. In another study, by means of
individual intramolecular O−H···O interaction energies in sugar cooperativity of halogen bond with hydrogen bond and
molecules using the molecular tailoring approach (MTA).175 substitution effect, the change of halogen bond from the
These studies reveal a contribution to the hydrogen bond energy chlorine-shared one to an ion-pair one was successfully
from the cooperativity to be typically between 0.1 and 0.6 kcal/ realized.192 The substitution of alkali metal greatly strengthened
mol when hydrogen bonds were a part of a relatively weak the halogen bond, while the cooperativity not only made the
equatorial−equatorial hydrogen bond network and higher halogen bond have a large change but also led to a large change in
between 0.5 and 1.1 kcal/mol when hydrogen bonds participated the strength of the hydrogen bond. Meng and co-workers noted
in an axial−axial hydrogen bond network. the positive cooperativity between the HOX···OH/SH halogen
bond and the Y−H···(H)OX hydrogen bond in OH/SH···
3. 3. Hydrogen Bonding-Interplay with Other Noncovalent HOX···HY (X = Cl and Br; Y = F, Cl, and Br) complexes by
Interactions
means of MP2 level calculations.193 Further, the interplay
Hydrogen bond operates in combination with numerous other between hydrogen bonding and lithium bonding in the HLi-
noncovalent interactions (Figure 6). A brief discussion on some NCH-NCH complex was studied with ab initio calculations. An
of the studies exploring their coexistence and their mutual impact increase in binding energies by about 19% and 61% for the
is given in the following section.180−214 In this context, Geerlings lithium and hydrogen bonds, respectively, in the trimer were
and co-workers have investigated the interplay between aromatic noted.196
stacking and hydrogen bonding in nucleobases by employing Yáñ ez and co-workers recently addressed cooperativity
high-level quantum chemical calculations.180 Hydrogen-bonding between hydrogen bonds and beryllium bonds in (H2O)n BeX2
capacity of the N3 and O2 atoms of cytosine was shown to (n = 1−3, X = H, F) complexes.197 The changes in the atomic
increase linearly with the electrostatic repulsion between the energy components were correlated with the changes in the
stacked rings. Superstructures of diketopyrrolopyrrole donors strength of the interactions, thereby accounting for cooperative
and perylenediimide acceptors formed by a combination of or anticooperative effects. Russo and co-workers studied the
hydrogen bonding and π···π stacking have also been recently interplay between hydrogen bonding and both anion−π and
2783 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

lone-pair−π interactions for the global stability of halide- approach to study condensed phases as it offers a method to
dichlorotetraoxacalix[2]-arene[2]triazine using first principle arrive at the energy of an N-molecule system that can be
computations, highlighting that their combination leads to a expressed as a sum of lower-order interactions which may be
cooperativity effect.198 Emphasizing how hydrogen bonds have obtained with reasonably high accuracy. Xantheas reported the
indeed become a crucial functional and structural element in importance of the nonadditive terms in particular the three-body
modern inorganic chemistry, a combination of transition metal terms as being responsible for determining the relative stabilities
ions and hydrogen-bonding interactions have been studied.199 of various trimer through pentamer isomers of water clusters.81
Cooperativity between weak hydrogen bonds was revealed in Almeida and co-workers demonstrate through many body
molecular clusters by examination of the structure, internal decomposition analysis, the importance of a proper representa-
dynamics, and origin of the weak intermolecular forces between tion of the cooperative effects in hydrogen fluoride by way of
sevoflurane (1,1,1,3,3,3-hexafluoro-2-(fluoromethoxy)propane) inclusion of at least the three-body terms.145 Recent advances in
and a benzene molecule, using multi-isotopic broadband this context include the development of a 3-body:many-body
rotational spectra.200 Thus, a large number of systematic integrated fragmentation method for weakly bound clusters,
investigations mentioned in the section above have established specifically finding application in the case of water clusters
the modulation of hydrogen bond strength through coopera- (H2O)n, n = 3−10, 16, and 17.217 This technique captures all 1, 2,
tivity. and 3-body interactions with a high-level electronic structure
method, whereas a less demanding method is used to obtain 4-
4. QUANTITATIVE DEFINITIONS FOR COOPERATIVITY body and higher-order interactions. Another variant in this
The quantification of cooperativity differs based on the approach approach particularly with respect to evaluation of the interaction
and the nature of the system employed.215−253 energies of large clusters (H2O)n=6,16,24 from many-body
4. 1. Many Body Analysis expansion involves reconstruction of the cluster from small
subclusters with a much lower computational cost.218 This is
Several studies attempt to get a quantitative account of the
done by applying progressively lower-level methods for
cooperative effects by decomposing the interaction energy of a
subsequent terms in the many-body expansion. Rapid con-
system of n bodies and by inclusion of the many body terms in
vergence of the many-body expansion is the key to such a
the analysis.145,215−227 Axilrod and Teller215 and Muto216 were
stratified approximation many-body approach. While studying
apparently the first to provide an explicit treatment of the
complex microscopic nonadditivity effects in hydrophobic
nonadditive interaction energies in 1943. Hankins et al.
interactions and their impact on protein folding, Chan and co-
investigated the potential of interaction for pairs and triplets of
workers evaluated potentials of mean force of three-body
water molecules and made a representation for the individual
hydrophobic association of three methanes in water.219 Recently,
terms in the many body expansion for these small water
Matsumoto has reported a study on the four-body cooperativity
clusters.70 A quantitative measure of the cooperative effect is
made by decomposing the interaction energy of a system of n observed while surveying the effective attraction force between
bodies. Typically, the energy of the cluster in consideration can hydrated methane molecules.220
be broken down into 1, 2,..., n body contributions via the many 4. 2. Double and Triple Mutant Analysis
body decomposition, where the 1-body term provides the Thermodynamic double mutant cycles and triple mutant boxes
monomer distortion or relaxation energy. The many body are widely employed for the experimental quantification of
expansion of interaction energy for a system with N monomers in noncovalent interactions and cooperative effects in pro-
it is given below teins.26,228−234 Double mutant cycles were originally devised to
N N investigate the interactions in proteins.228,229 Chemical double
E(1, .., N ) = ∑V 1B
(i ) + ∑ V 2B(i , j) mutant cycles help to measure the magnitude of a particular
i i<j functional group interaction in both weakly and strongly bound
N complexes.231 Any difference between the functional group
+ ∑ V 3B(i , j , k) + ... + V nB(1, ..., N ) interaction energies in two systems was shown to provide a
i<j<k measure of the magnitude of the enthalpic chelate effect in the
(2) complexes.233 Two double mutant cycles can be formally
combined to produce a triple mutant box, providing a general
where method for quantification of cooperative effects. Hunter and co-
2 workers have performed extensive studies using double and triple
V 2B = E(1, 2) − ∑ E (i ) mutant cycle experiments to evaluate cooperativity in non-
i=1 (3) covalent interactions.26,231−233 Quantification of intermolecular
3 3
functional group interactions in hydrogen-bonded zipper
V 3B = E(1, 2, 3) − complexes in chloroform231 and in two different doubly
∑ E ( i , j) + ∑ E ( i ) hydrogen-bonded motifs in carbon tetrachloride, chloroform,
i<j i=1 (4)
1,1,2,2-tetrachloroethane, and cyclohexane form part of their
1B
V (i) is the one-body (1B) potential, which gives the energy studies.233
that is required to deform an individual molecule from its Schreiber and co-workers have provided an experimental
equilibrium geometry. V2B and V3B are the two-body interaction approach to evaluate the net binding free energy of buried
and three-body interaction terms where E(1, 2) is the dimer hydrogen bonds and salt bridges in selected interactions between
energy and E(1, 2, 3) represents the trimer energy. The three- TEM-1-β-lactamase and its protein inhibitor, BLIP.171 Results
body and higher interaction terms reflect the nonadditive or from this study demonstrate the importance of forming networks
cooperative effects typically seen in hydrogen-bonded clusters. of buried salt bridges and suggest that the cooperative
The many-body expansion indeed represents a powerful networking effect results from the favorable contribution of the
2784 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

protein to the interaction. Ullmann et al. have recently reported acceptor (1), phenol (2), and metal cation (3) proceeds with
free-energy calculations within a continuum electrostatics model reaction energy ΔE123.
to analyze the coupling of protonation, reduction, and
Ecoop = ΔE123 − ΔE12 − ΔE32 − ΔE13 (5)
conformational change in azurin from Pseudomonas aerugino-
sa.230 Free energy measures of cooperativity are used for a
ΔE123 = E123 − E1 − E2 − E3 (6)
detailed analysis of the above-mentioned changes revealing how
cooperative free energies are useful in detecting and quantifying Two different routes adopted for the stepwise formation of the
thermodynamic coupling between events in biomolecular ternary complex were explored. The first one involved initial
systems. formation of the hydrogen-bonded complex followed by
Measures of cooperativity based on a combination of subsequent addition of the metal cation to form the ternary
mathematical modeling and experimental data are available in complex. A second stepwise pathway was constructed in a
few studies.55,29,137,235−245 Mariuzza and co-workers used surface completely analogous mode by initial addition of the metal cation
plasmon resonance techniques along with modeling on a to the phenol π-system and subsequent formation of the
trimolecular protein complex as a model system for quantifying hydrogen bond. The computed positive cooperativity between
cooperativity.237 The complex considered in this instance was cation−π and hydrogen-bonding interactions was rationalized
obtained by the simultaneous interaction of a superantigen with through RVS (reduced variational space) analysis. A repulsive
major histocompatibility complex and T cell receptor. This study contribution of electrostatic interaction working in an anti-
reports both cooperativity and anticooperativity, besides cooperative fashion during the formation of the ternary complex
revealing an augmentation of the temperature dependence of was noted. This anticooperativity however was smaller in
binding kinetics upon the cooperative interaction of different magnitude compared to the polarization component of phenol
protein components in the complex. Jauch and co-workers have which was essentially responsible for the net positive
developed a method to measure heterodimer cooperativity cooperativity.
through a factor which reveals the mode of Sox-Oct transcription In another study, we gauged cooperativity in ternary
factor heterodimerization on composite DNA elements.238 complexes formed by two aromatic benzene rings (BB) and a
Roubeau and co-workers have reported a direct, experimental, benzene ring with a cation (viz. Li+, Na+, K+, NH4+, PH4+, OH3+,
quantitative measure of the cooperative character of the spin- and SH3+) (IB) in representative complexes.248 Similar to the
crossover phenomenon in triazole-bridged Fe(II) spin-crossover study mentioned earlier, the magnitude of cooperativity in a
one-dimensional materials.240 The cooperative character of the ternary complex was calculated by deducting all pairwise
spin-crossover in these materials was quantified using the so- interactions from the total complexation energy of the trimer.
called Slichter and Drickamer and domain models based on data The complexation energy of the trimer corresponded to the
obtained from differential scanning calorimetry. Mas and co- energy involved in the direct assembly of the ternary complex
workers have emphasized the hierarchical approach to from its constituent monomers which were two benzene units
cooperativity in macromolecular and self-assembling binding and a cation. While two of the three two-body terms represent
systems.241 A global association quotient which was defined as Kc the cation−π and π−π interaction energies, the third term
= [occupied sites]/([free sites] L), with L being the free ligand represents the through space cation−π interaction energy
concentration was employed. This coefficient Kc easily relates to between the cation and the second benzene ring with which it
other measures of cooperativity such as the Hill number or the does not have any contact. Besides several other studies have also
Scatchard plot and also to the free energies involved in the employed energy decomposition schemes to evaluate contribu-
binding processes at each ligand concentration. Douglas and co- tion of various factors to cooperativities.129,189,246,247,249−251
workers investigated specific models of self-assembly where the Grimme and co-workers subjected the intermolecular interaction
self-assembly arises through the presence of nonlocal constraints energy of nucleic acid trimers from a benchmark data set to an
such as thermal activation, chemical initiation, and formally energy decomposition analysis (EDA) to reveal that nonadditive
constrained reaction order on the self-assembly process.242 A effects are mainly due to induction, while exchange repulsion,
comparison of basic thermodynamic characteristics of these electrostatic, and dispersion contributions are essentially
models to those obtained for a cooperative model and an additive.251 In a study employing quantum chemical calculations
idealized noncooperative reference model were reported. Piguet to investigate N−H···O hydrogen-bonding properties in linear
and co-workers suggest a combination of two simple indexes (urea)n=1−10 clusters, Morokuma analysis for all the clusters and
measuring intermetallic (IcMM) and interligand (IcLL) interactions DFT-SAPT analysis for the urea dimer were performed.129 The
as part of a thermodynamic model developed for quantitatively DFT-SAPT analysis of the interaction energy components
estimating cooperativity in supramolecular polymetallic [MmLn] indicates that the electrostatic and dispersive interactions are
self-assemblies.243 indeed the most important attractive terms in the urea dimer.
The interaction energies of several structures of guanidinium−
4. 3. Measures for Estimation of Cooperativity in
benzene complexes and anion approaching in different directions
Noncovalent Interactions
were further dissected into their electrostatic, exchange,
4. 3. 1. Interaction Energy and Energy Decomposition repulsion, polarization, and dispersion contributions by means
Analysis. We have defined the cooperative energy Ecoop in a of local molecular orbital energy decomposition analysis.250
ternary complex of phenol, metal cation, and hydrogen bond Besides the electrostatic cation−anion attraction, the effect of the
acceptors in our group.181 The extent to which the hydrogen anion over the cation−π interaction was shown to be mainly due
bonding and cation−π interaction act in concert in these ternary to polarization and could be rationalized following the changes in
systems can be deduced quantitatively by comparing the overall the anion−π and the nonadditive terms of the interaction. Datta
reaction energy ΔE123 with the three individual interaction and co-workers have applied an energy decomposition analysis
energy terms i.e., ΔE12, ΔE32, and ΔE13. The direct assembly of scheme to mixed cation (Li+, Na+, and K+) and anion (F−, Cl−,
the ternary complex from its constituents hydrogen bond and Br−) complexes on π-surfaces.252 Electrostatic interactions
2785 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

were found to be repulsive between the π-surface of benzene and its intensity increases. The extent of decrease in N−H stretching
the anion F−•. Dispersion interactions were also significant in frequency depends on the strength of hydrogen-bonding
providing stability to these complexes. While studying the interaction. With an intention of demonstrating variation in
cooperativity of hydrogen- and halogen-bonding interactions in cooperativity among different arrangements of acetamide and
Cl−···HCCH···HF, Cl−···ClCCH···HF, and F−···ClCCH···HF formamide clusters of increasing size, we compared the
complexes, Grabowski showed that all hydrogen bonds and magnitude of shift in N−H stretching frequency in each
halogen bonds which existed in the corresponding dyads were arrangement.108,109 This was done using two parameters, average
enhanced in the triads.189 The polarization term of the energy of N−H stretching frequency value for each cluster, and deviation
interaction was found to be most important for the halogen- of average N−H stretching value from that computed for a
bonded systems. Symmetry-adapted perturbation theory monomer. The decrease in value of average N−H stretching
(SAPT) was applied to systems where pnicogen bonds and frequencies leading to a red shift in larger clusters correlates to
cation−π interactions coexist in the same complex, in order to enhanced hydrogen bonding and was used to explain
understand the nature of the interactions and the mechanisms of cooperativity.
cooperative and diminutive effects operating in such com- 4. 3. 3. Structural Analysis. Analysis of variation of the
plexes.249 Electrostatic interaction played a major role in the structural parameters of different groups involved in the
change of cation−π interaction strength in the ternary systems hydrogen bonding as a consequence of their cooperative nature
considered in this study, whereas the change of pnicogen is performed in several studies.111,112,118,159,254 Studying
bonding strength was shown to be caused jointly by the cooperativity in amide hydrogen bonding lead Kobko and co-
electrostatic, induction, and dispersion interactions. workers to examine the relationship between the CO and C−
4. 3. 2. Spectoscopic Parameters. Maes et al. have N bond distances as a function of the hydrogen-bonding chain
employed matrix-isolation FT-IR spectroscopy as a suitable length and the position of the formamide in linear chains clearly
experimental method for evaluating hydrogen bond coopera- reflecting a structural response by each formamide unit to the
tivities in a case study for self-association and hydrogen bonding hydrogen bonding within its chain.105 Futher as the chains
of 3,4,5-trichlorophenol with water.77 Using two cooperativity become longer and the formamide monomer becomes more
factors, one representing the fortification of a hydrogen bond A− central, the CO distances were shown to increase, while the
H···A−H by a further A−H···B interaction in a 1:2 complex A− C−N distance decreased. Dannenberg and co-workers also
H···A−H···B and another representing the inverse fortification, compared the structures and energies of β-strands, α-helices, and
differentiation between open, linear structures, and closed, cyclic 310-helices for capped polyalanines, acetyl(ala)nNH2, for values
trimers was shown to be possible. Ohno et al. have also employed of n from 2 to 18, using mixed DFT/AM1 calculations.158 Potent
matrix-isolation infrared spectroscopy and DFT to understand cooperativity of α-helices were manifested structurally by
the effect of cooperative hydrogen bonding on the OH decreasing O···H distances as n increases, energetically by
stretching-band shift for water clusters.85 While establishing increased stability per hydrogen bond, and from dipole
cooperativity in amide hydrogen-bonding chains, Dannenberg moments. The effect of the methyl group on the cooperativity
and co-workers have compared between vibrational coupling between three types of hydrogen bond (O−H···O, C−H···O,
through hydrogen bonds and covalent bonds for polyglycines in and O−H···π) in a cyclic complex was explored in a study by Li et
β-strands and chains of hydrogen-bonding formamides.104,156,167 al.128 Addition of methyl group lead to enhancement of O−H···π
Through a study on G-quartet, cyanamide rings, and the linear and O−H···O hydrogen bonds, and weakening of C−H···O
cyanamide chain with identical monomers, van Mourik et al. hydrogen bond. Performing DFT and hybrid DFT/MM
showed that in the absence of geometric variations, the simulations of polyglutamine β-sheet structures in vacuo and in
magnitude of the NMR parameters changes when hydrogen- aqueous solution, it was observed that the cooperativity of
bonding monomers are progressively added to extending ring or glutamine side chains affects both the directions perpendicular
chain structures.116 The influence of cooperativity on the and parallel to the backbone.153 Energetic as well as structural
frequency shift of the Ar−H stretch vibration in HArF complexes aspects such as shortening of hydrogen bond lengths and
has also been studied.253 Varfolomeev et al. have used a increasing the number of hydrogen bonds involved were
combination of FTIR spectroscopy and quantum chemical undertaken. The cooperativity of hydrogen bonding in different
calculations on complexes of catechol and phenol along with arrangements of water, acetamide, and formamide clusters of
different proton acceptors.239 A measure of cooperativity based increasing size was also manifested by a regular decrease in
on stretching vibration frequency shifts was developed, and this average O···H and C−N bond distances, while average CO
frequency shift in a complex with cooperative hydrogen bonds and N−H bond lengths increase with increasing cluster
was determined from the IR spectra of catechol with equimolar size.84,108,109
quantity of bases in the gas phase. The cooperativity measure
(Ab) employed for the same is shown below. 5. HOW DOES A PAIR OF NONCOVALENT
O − H ···· O − H ··· B
INTERACTIONS MUTUALLY INFLUENCE EACH
ΔνHB OTHER?
Ab = O − H ··· B
ΔνHB (7)
5.1. Cation-Induced Cooperativity
where ΔνO−H···O−H···B
HB represents X−H (X−O or N) stretching Cation−π interactions are arguably the strongest among
vibration frequency shift in the complexes with the adjacent noncovalent interactions and have been shown to be modulated
hydrogen bonds and ΔνHB O−H···B
indicates X−H stretching by several factors.14−17,41,64,248,255−278 A number of experimental
vibration frequency shift in a complex with a single hydrogen and theoretical studies which elucidate cation-induced cooper-
bond. ativity have also emerged.249,279−299 Our group has investigated
In the case of clusters where hydrogen-bonding interaction is the fundamental question of how a pair of noncovalent
manifested, N−H stretching frequency shifts to a lower value and interactions mutually impact each other, especially when one
2786 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 7. Schematic representation of approach taken to demonstrate how M−π and π−π interactions work in concert, and the subtle π−π interaction
becomes substantial in the presence of a metal ion. (a) Interaction energies (in kcal/mol) of the metal ion with the benzene monomer and dimers and
(b) π−π interaction energies (in kcal/mol) of the various benzene dimers in the presence of metal ions at the MP2/6-311++G**// MP2/6-31G* level
of theory are shown. Edited and reprinted from refs 270 and 271. Copyright 2006 American Chemical Society.

of them is a cation−π interaction. In this regard, two scenarios beside the T) orientations. Besides the database analysis, a
were studied systematically. In the first case, cooperativity computational study was also undertaken. Interestingly, the
between cation−π interaction and hydrogen bonding181 was enhancement of π−π interaction in the presence of metal ion was
dealt with and in the second instance between cation−π and π−π found to be quite substantial, especially when the metal had a
interactions.248 In an earlier effort, the common coexistence of higher charge. Mg2+ ion enhanced the strength of the π−π
M−π and π−π interactions in biology and chemistry has been interactions by almost more than 5-fold in the three
explored to garner a better understanding of how one kind of configurations of benzene, S−M, PD−M, and TB−M. There-
noncovalent interaction affects the strength of another (Figure fore, the strength of subtle π−π interactions transformed to
7).270,271 This mutual impact is typically described in terms of substantial, in the presence of a metal ion. Significantly, the
cooperativity and anticooperativity in bonding. The occurrence interaction between the two benzene rings in most orientations is
of M−π−π (M = Li+, K+, Na+, Mg2+, and Ca2+) interactions in the enhanced to an almost comparable extent. The influence that the
CSD (Cambridge Structural Database) and PDB (Protein Data metal ion has on the first ring could be the major source of the
Bank) databases have been investigated and compared with the enhanced interaction. Thus, metal ion-assisted π−π interaction
number of exclusive M−π motifs. Besides, different forms of strengths may become comparable in magnitude to that of the
benzene dimers (PD-parallel displaced, S-stacked, and T- hydrogen-bonding interaction.
shaped) were also subjected to ab initio calculations as part of As mentioned earlier, we have performed quantum chemical
this study, to establish the relative preference of differently calculations to gauge the effect of cation−π and hydrogen-
oriented aromatic moieties to bind to each other. The PDB was bonding interactions on each other.181 A model ternary system,
then searched for metal ion containing structures having less than M-phenol-acceptor (M = Li+ and Mg2+; acceptor = H2O,
3 Å resolution and a R value < 0.3. A total of 1941 protein HCOOH, HCN,CH3OH, HCONH2, and NH3) was taken as it
structures containing the 5 metal ions were then subjected to exhibits both cation−π and hydrogen-bonding interactions
analysis following removal of redundancy. The database analysis (Figure 8). Cooperativity was clearly quantified and defined,
revealed several salient points: (i) there was a high occurrence of and the computed positive cooperativity between cation−π and
M−π and M−π−π interactions in chemistry and biology. (ii) hydrogen-bonding interactions further rationalized through an
The prevalence of M−π−π was observed to be higher in most energy decomposition scheme, reduced variational space (RVS)
cases or comparable to that of exclusively M−π configurations. and charge analyses. Although the direct interaction between the
(iii) The number of M−π−π motifs present was only marginally metal ion and acceptor contributed to the cooperative
lower than the number of metal ions available. (iv) In proteins, interaction, the enhancement of both the cation-π and the
the aromatic amino acid side chains seemed to prefer the PD−M hydrogen bonding energy, even in the absence of such
and TB−M (T-shaped alignment of the benzene rings with M stabilization, was shown to be significant.
2787 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

forms have been recently reported by our group (Figure


9).272−274 In all these studies, the conformational space of
aromatic amino acid dimers has been explored by considering the
relative orientation of two aromatic moieties such as parallel
displaced, T-shaped, and stacked orientations, and the possible
hydrogen-bonding interactions between COOH and NH2
groups and also other noncovalent interactions between
aromatic rings and COOH and NH2 groups such as OH-π,
NH−π, and π−π interactions. The comparison of stability of
different dimeric forms studied reveals that the most stable
dimers predominantly possess hydrogen-bonding interactions,
while the ones with aromatic side chain interactions are less
stable. Thus, a delicate balance of noncovalent interactions and
their interplay governed the stability of different forms of the
studied aromatic amino acids and their respective dimers.
Rooman and co-workers have focused on the relative
importance of simultaneous cation−π and hydrogen-bonding
interactions in DNA staircase motifs.279 Nonadditive effects
induced by the different energy and entropy contributions both
in vacuum and in different solvents were explored, revealing how
in the absence of an estimation of the dispersion nonadditivity,
the cooperativity for guanine-argnine-guanine, cation−π, and
hydrogen-bonded stair motifs arises from the environment. 1H
NMR experiments and quantum chemical calculations by
Dutasta and co-workers emphasize the competing cation−π
and anion−π interactions involved simultaneously in the
molecular recognition process of selected zwitterionic neuro-
transmitters and a particular hemicryptophane.280 Li and co-
Figure 8. Cooperativity between cation−π and hydrogen-bonding workers showed the prominent enhancing effect of the cation−π
interactions. Edited and reprinted from ref 181. Copyright 2008
American Chemical Society.
interaction on the halogen−hydridehalogen bond in the
model systems, M1···C6H5X···HM2 (M1 = Li+, Na+; X = Cl, Br;
M2 = Li, Na, BeH, and MgH).281 The study showed that
Extensive DFT studies and conformational analysis of halogen−hydride halogen bonding is indeed strengthened
tryptophan, phenyalanine, and tyrosine in their ionic and dimeric greatly by a cation−π interaction. The interaction energy in

Figure 9. Pictorial representation of various noncovalent interactions observed in dimers of tryptophan (ref 272), tyrosine (ref 273), and phenylalanine
(ref 274), characterized using AIM analysis.

2788 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 10. (a) Schematic representation of ternary model systems containing 4-amino-2-iodophenol, Li+, NH2CH3, C6H6, and NH3CH3+ exhibiting
different kinds of noncovalent interactions, including cation-π (CP), hydrogen bond (HB), halogen bond (XB), π−π (PP), metal ion-lone pair (ML)
and charge-assisted hydrogen bond (CHB). (b) Complexation energy of the ternary systems, sum of complexation energy of binary systems (in free
state) present in the ternary systems and cooperativity energy of the ternary systems as obtained at the M06-2X/6-31G(d) (I: DGDZVP) level of theory.
Edited and reprinted from ref 275. Copyright 2015 American Chemical Society.

the triads was two to six times as much as that in the dyads. corresponding binary systems were explored by them.282 A
Moreover, the nature of the cation, halogen donor, and the metal methylated pyrrole was shown to be favorable for the cation−π
hydride were shown to influence the nature of the halogen bond. interaction, whereas it was unfavorable for the NH---O hydrogen
In another study, the effect of the methyl group on the bond. Further, an ab initio study of synergetic effects of two
cooperativity between cation−π interaction and NH---O hydro- strong interactions namely cation−π interaction and lithium
gen bonding using ternary systems constituted with Na+, bond in the M+···phenyl lithium···N model system where (M =
pyrrole/methylated pyrrole, and water/dimethyl ether and the Li, Na, K; N = H2O and NH3) revealed that cation−π interaction
2789 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

and lithium bonding in the trimers become stronger relative to By careful consideration of all possible combinations of these
the dimers.283 An increase in interaction energy by 4.4−6.3% was noncovalent interactions, based on energy, geometry, charge,
reported for cation−π interaction, while that of lithium bonding and AIM analysis, we have systematically evaluated the
increased by 5.2−15.9%. Lu and co-workers have investigated the cooperativity among 40 ternary systems and 105 quaternary
mutual influence between halogen bonds and cation−π systems. A study of the quaternary systems revealed how
interactions using MP2 level calculations.284 hydrogen bond, halogen bond, and π−π interaction work
Song and co-workers studied cation−π or cation−π−π together by enhancing each other’s strength. The study
interaction between one cation and one or two structures highlights that the positively charged species enhance hydrogen
bearing electron-rich π systems.285 A new type of interaction, bond−hydrogen bond and hydrogen bond−π−π interactions
cation⊗3π, in which one cation simultaneously binds with three forming cooperative systems. Surprisingly, some of the
separate π-electron-rich structures was explored. An anomalous quaternary systems were also found to be cooperative in nature
increase in the order of the one-benzene binding strength of the despite the electrostatic repulsion between two positive charge
cation⊗3π interaction, with K+ > Na+ > Li+ at odds with the species involved.
conventional ranking of the binding strength, was noted. This Through a study of substituted anilines and benzylamines,
unexpected observation was owing to the cooperative interaction cooperative cation−π, π−π, and van der Waals interactions were
of the cation with the three benzenes and also between the three shown to be responsible for an increase in aromatic cationic
benzenes, wherein steric-exclusion effect between the three amine sorption to Na/Ca-montmorillonite, well beyond the
benzenes played an important role. Frontera and co-workers extent expected by cation exchange alone.295 Besides compounds
have demonstrated how the cation−π and hydrogen-bonding with greater amine charge/area and electron-donating sub-
interactions influence each other in complexes where both stituents that allowed for greater electron density at the center of
interactions are present and also if such an interplay monotoni- the aromatic ring showed a greater potential for cation−π
cally increases by adding further hydrogen bonds to the aromatic interactions on montmorillonite surfaces. Hence, numerous
systems.286 Cooperativity effects were demonstrated in another studies of cation-induced cooperativity have been in evidence in
study, when cation−π and lone pair−π interactions coexist in the the past few years.
same system.287 Employing a combination of DFT and MP2 5.2. Anion-Induced Cooperativity
calculations on a binary complex formed by an O-cresol moiety
and a Na+ cation or two O-cresol monomers, as well as the In this section, several studies which employ theoretical
ternary system that contains a Na+ cation and two O-cresol units, calculations and experimental examples to illustrate how
Ren and co-workers investigated the cooperativity effects interplay exists between anion−π interactions and other
between the cation−molecule and hydrogen bonding.288 noncovalent interactions have been dealt with. Quite in contrast
Cooperativity effect was observed in the conformations formed to cation−π interactions, the generality of anion−π interactions
by the moderate Na+···π, H···π, and H···O interactions, while the has come to light only in recent years. A few recent reviews also
anticooperativity effect was present in the complex with the highlight the increasing understanding of the role played by
strong Na+···O cation−molecule interaction. The cooperativity anion−π interactions in supramolecular assembly.21−23 Mount-
effect between dihydrogen bonding and H−M···π interactions ing experimental evidence for the presence of anion−π
was also investigated where H···H and H−M···π interactions interactions through elucidation of crystal structure of inorganic
were shown to be strengthened in the ternary complexes in complexes and biological macromolecules along with establish-
comparison with the corresponding binary system.289 Studying ing their functional significance has been reported.19,20
the influence of the Li···π interaction of C6H6···LiOH on the The coexistence of anion−π interactions with other non-
H···π interaction of C6H6···HOX (X = F, Cl, Br, and I) and the covalent interactions has also been revealed.250,252,300−322
X···π interaction of C6H6···XOH (X = Cl, Br, and I), Meng and Atwood and co-workers provided preliminary studies on anion
co-workers revealed that the addition of the Li···π interaction to binding within the cavity of π-metalated calixarenes.19 They
benzene weakens the H···π and X···π interactions.290 Ebrahimi et reported the synthesis of tetrametallic hosts as anion receptors by
al. were able to reveal the influence of cation−π and anion−π X-ray crystal structure investigations, clearly demonstrating a
interactions on the strength and nature of the N···H hydrogen cooperative effect arising from the arrangement of four metal
bond based on MP2 calculations in model systems.291 centers about a common, rigid binding pocket, resulting in
Cooperative and diminutive effects were found when halogen anion−host contacts. Hay et al. have summarized evidence for
and lithium−π bonds coexisted in the same complex considering existence of four distinct binding modes for complexes of anions
the case of CF3X-NCLi-C2H4, CF3X-CNLi-C2H4, and CF3X- with charge-neutral arenes, including C−H hydrogen bonding
C2H4−LiCN (X = Cl, Br, I) complexes.292 Cheng et al. also and anion−π interaction.306 Manzano and co-workers studied
reported interesting cooperative and diminutive effects when anion encapsulation in metallic grids showing the presence of
pnicogen bonds and cation−π interactions coexist in the same two counteranions (BF4− or PF6−) hosted in the cavities,
complex.249 exhibiting C−H···F and anion···π interactions in the solid
In a recent study from our group, we examined the key factors state.307 In the case of complexes with a methylated ligand,
that control the structures and energetics of the coexistence of aromatic π−π stacking interactions were also noted. Using a
multiple noncovalent interactions.275 4-Amino-2-iodophenol combined crystallographic and theoretical approach, Frontera
was taken as a model system which exhibits nine different and co-workers studied anion−π interactions in a series of
kinds of noncovalent interactions, including cation−π, hydrogen bisadenine derivatives dwelling in particular on the existence of
bond through oxygen and nitrogen, halogen bond, π−π synergistic effects between π−π and anion−π interactions.308
interaction, metal ion−lone pair interaction through oxygen Berryman et al. developed the experimental evidence for
and nitrogen and charge-assisted hydrogen bond through oxygen interactions between anions and electron-deficient aromatic
and nitrogen (Figure 10). Indeed, very few studies have been rings.309 They were able to demonstrate how tandem hydrogen
reported on the cooperativity in quarternary structures.141,293,294 bonds and anion−arene interactions augment halide binding in
2790 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

solution.310 Jentzsch et al. have recently demonstrated the 6. STRONG MANIFESTATION OF WEAK
general functional relevance of anion−π interactions elucidating INTERACTIONS
how anion transport is achieved with halogen bonds.311 Notash
Interestingly, a large number of studies on cooperativity among
et al. established how a coordinated anion could control the
several other weaker noncovalent interactions have come to light
structural motif of one-dimensional coordination polymers
in the past few years.323−361 Atwood et al. have shown the impact
through multiple cooperative weak noncovalent interactions
of extensive cooperativity in maintaining the structural integrity
such as hydrogen bonds, halogen---π and π−π interactions.304
of a nonporous organic solid operating as a host compound for
Giese et al. investigated an anion−π complex between bromide
guest transport.8 The host molecules were shown to be mobilized
and a neutral receptor both in solid state and solution.312 A
definite role of a cooperative effect of the N−H--anion and in response to relatively weak van der Waals interactions with
anion−π interaction was reported therein. Zaccheddu et al. small guest species such as vinyl bromide or toluene. Li et al.
employed a combination of dispersion-corrected DFT and studied the interplay between halogen bond and lithium bond in
quantum Monte Carlo calculations to rationalize the unusual the complexes of the form MCN-LiCN-XCCH where (M = H,
structural features observed in a nitrate−triazine−triazine Li, and Na and X = Cl, Br, and I).24 The study reports distinct
complex.313 This study emphasized that π−π stacking was not cooperativity between the lithium bond and halogen bond where
only enforced by the coordination of the triazines within the the lithium bond has a larger enhancing effect on the halogen
particular crystal structure but was regulated by cooperative bond, in a range of 11.7−29.4%. Using ab initio calculations,
anion−π and π−π interactions, wherein the cooperative effect Meng and co-workers have demonstrated the presence of
amounted to 6% of the total binding energy. cooperativity in binary and ternary complexes of OCS···C6H6,
The specific aspect of anion-induced cooperativity has also C6H6···Rg and OCS···C6H6···Rg (Rg = He, Ne, Ar, and Kr)
been addressed through numerous theoretical studies.315−322 governed by van der Waals forces.324 Li and co-workers have
Extensive theoretical studies on the role of anion-induced emphasized the importance of both electrostatic and van der
cooperativity have come from the group of Frontera and co- Waals interactions as being essential for describing the
workers.315−319,321,322 Studying the interplay between cation−π, cooperative interaction in long helices.325 In a specific study on
anion−π, and π−π interactions in trifluorobenzene π systems long α- and 310-helical polyalanines, the greater cooperativity of
along with selected metal cations and anions by applying α-helices over 310-helices in long helices was shown to originate
resolution of the identity MP2 level calculations revealed that mainly from the much stronger van der Waals interaction in α-
both the cation−π and the anion−π interactions have a strong helices. Recently, Grimme and co-workers have calculated the
influence upon the π−π interaction and vice versa.315 They also energy profiles of the activation reaction of small molecules (H2,
reported cooperativity effects in complexes, wherein anion−π Br2, and CO2) with boron/phosphorus frustrated Lewis pairs.326
and hydrogen-bonding interactions coexist even at very large While investigating the cooperative nature of the reactions using
distances (as long as 11 Å).316 Cooperativity effects have also interaction energies in the ternary system and for reactant pairs,
been reported by them in complexes where interplay between they proposed an isosurface representation of the many-body
lone pair−π or anion−π interactions and halogen bonding exists. deformation density (Δρmb) to serve as a qualitative tool to
Cooperativity was observed even when the distance between the visualize cooperative, nonadditive effects in complex chemical
anion and halogen bond donor molecule was longer than 9 Å.317 systems. Ren and co-workers have employed DFT and ab initio
Unexpected nonadditivity effects were also reported in the case calculations on ternary systems of FH···HM···C2H2/C2H4/
of complexes of fluorine-substituted ethyne, ethene, butadiene, C6H6, where M = Li, Na, and K and C2H2/C2H4/C6H6 represent
benzene, and [n] radialenes (n = 3−5) with two anions at the RI- the π electron donor.186 The study reveals the cooperative nature
MP2/aug-cc-pVTZ level of theory.318 of interaction between dihydrogen bonding and H−M···π
Ab initio MP2 level claculations on 1,4-diiodo-perfluoroben- interaction. Esrafili and co-workers have revealed a size-
zene working both as electron-deficient π aromatic ring and as dependent variation in atomic multipole moments as a useful
halogen bond donor with anions were performed by Lu et al.25 signature of lithium-bonding cooperativity in case of linear
Distinct additive and diminutive effects were observed when clusters of (LiCN)n and (LiNC)n (n = 3−7), which are
halogen bonds and anion−π/lone pair−π interactions were connected via lithium bonds.327 Cooperative effects in the
present in the same complex. Cabaleiro-Lago and co-workers geometry, energy, and electron density in the linear and cyclic
clearly showed how the interaction energy was decomposed into arrangements of clusters of interhalogen derivatives have also
different two- and three-body contributions in complexes been reported.328 Cooperative interplay between covalent and
consisting of guanidinium, benzene, and one anion.250 The noncovalent interactions have been explored both from the
three-body interaction was demonstrated to be anticooperative perspective of crystal structure analysis341 as well as employing
in the case of the presence of the cation and the anion on the theoretical calculations.330
same side of the π system, whereas when the anion and the cation Another important class of noncovalent interactions are the
were on opposite sides of the π system, the three-body σ−σ interactions which are decisive for alkyl group inter-
interaction was cooperative. Datta and co-workers have actions.362−369 Indeed, in the absence of polar functional groups,
employed dispersion-corrected DFT on mixed cation anion π dispersion dominates intermolecular interactions between
systems to reveal cooperative stabilization of the otherwise hydrocarbon moieties and is characteristic for CH−σ and
weakly stable anion···π complexes.252 Decomposition of the total CC−π as well as for various other σ−π-mixed clusters. Schreiner
interaction energies in these systems has shown that the et al. reported the preparation and characterization of alkanes
cooperativity among the interactions is indeed substantial and with the longest C−C bonds observed to date (up to 1.704 Å),
as large as 100 kcal/mol. Thus, the anion-induced cooperativity whose surprisingly high thermal stabilities are due to a favorable
has emerged as an important mode of tuning several noncovalent balance of repulsive and attractive van der Waals interactions
interactions such as hydrogen bonding, halogen bonding, and between the CH surfaces.363 σ Stacking can reach the energy of
cation−π interactions. chemical bonds, and a study by Schreiner et al. illustrates the
2791 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

point that σ−σ and π−π interactions are equally important.362 reported the synthesis of a flavoenzyme model bearing a
With regard to the properties of single-sheet [n]graphanes, their diamidopyridine recognition element and several aromatic
corresponding double-layered forms along with multilayered units to mimic π−π and donor atom−π interactions commonly
[n]graphanes (n = 10−97) were studied employing the found in nature.385 The study demonstrated that a subtle
dispersion-corrected DFT in the above-mentioned study. interplay occurs between the hydrogen-bonding network and the
Interestingly, there is currently no evidence for cooperativity aromatic units which allows for maximum stabilization of the
for these interactions, and the association energies of multi- cofactor in the reduced state. In an attempt to put the discussion
layered graphanes per carbon atom are rather similar to their of the interplay of the various interaction energy contributions on
double-layered forms and independent of the number of quantitative grounds, Hesselmann et al. have employed DFT-
layers.362 A benchmark comparison of σ−σ and π−π dispersion SAPT methodology (symmetry adapted perturbation theory
in dimers of naphthalene and decalin and coronene and based on DFT) on the Watson−Crick and stacked structures of
perhydrocoronene showed that strong dispersion interaction the adenine-thymine and guanine-cytosine base pairs.386 A good
requires rigid subsystems and good fits of their repulsive agreement of total interaction energies for the cbs-extrapolated
potential walls.367 DFT-SAPT deviating by an upward shift of 1.2−1.6 kcal/mol
from estimated cbs CCSD(T) results for stacked and Watson−
7. INTERPLAY OF π−π INTERACTIONS Crick structures of AT and GC was reported.
The cooperativity of π−π and cation−π interactions has been
7.1. With Other Noncovalent Interactions clearly demonstrated by our group taking benzene as the
Several examples of coexistence of π−π interactions along with aromatic model system and a series of monovalent cationic
other noncovalent interactions abound in litera- species, such as Li+, Na+, K+, NH4+, PH4+, OH3+, and SH3+
ture.251,294,308,370−387 The presence of halogen bonding and (Figure 11).248 Systematic quantum chemical studies were
π−π interactions in the solid state structure of a butadiynylene
linked bidentate halogen bond donor was recently shown by
Taylor and co-workers.378 A cocrystal with tetra-n-butylammo-
nium iodide was also studied. An interplay of coordination and
π−π stacking interactions has been studied in cadmium(II)
iodide and thiocyanate complexes adopted by polycyclic 1,4-
bis(pyridazin-4-yl)benzene.377 Mattay and co-workers have
reported the synthesis of arene-functionalized 2-aminopyrimi-
dines, where hydrogen bonding and π−π interaction have a
significant contribution in the process of self-assembly on
account of their coexistence.372 Fu et al. revealed the crystal
structure of methyl 4-hydroxy-3-nitrobenzoate where 12 hydro-
gen bonding and two π-stacking interactions are shown linking
the molecules into infinite stacked sheets.373 Shishkin et al. have
conducted an analysis of the crystal structure of a set of
substituted 1,3,5-trihalobenzenes from the viewpoint of their
energy of intermolecular interactions.379 The study showed that
all crystals studied possess a columnar structure where the
presence of shearing mechanical properties is not caused by the
layered structure of the crystal but the character of interactions
between neighboring columns, including stacking interactions
and halogen bonds which are located near planes of molecules
belonging to different columns. Figure 11. Model systems employed to understand cooperativity
Theoretical and experimental studies on model systems which between cation−π and π−π interactions of ref 248.
mimic π−π interactions along with other noncovalent
interactions have also been reported.388−396 Rooman and co- performed to estimate the effects of cation−π and π−π
workers focused on a stair motif involving a cation−π and interactions on each other in cation-π−π systems. This study
hydrogen bond occurring in the DNA-binding domain of Tc3 has provided a simple model for cooperativity. All possible
transposase from Caenorhabditis elegans (1TC3) between Arg orientations of onium ions to form cation−π complexes were
C236 and the two successive Gua A7 and A8.279 This complex explored, and the most stable conformation was used to further
was considered as a model system for studying the relative evaluate its effect on π−π interaction. A notable increase of 2−5
importance of the three simultaneous interactions that is, kcal/mol in the π−π interaction energy in the presence of the
cation−π, hydrogen bonding, and stacking, with the resulting cations was seen. The cation−π interaction energy was also
nonadditive effects induced by the different energy and entropy enhanced in the presence of π−π interaction albeit to a smaller
contributions, both in a vacuum and in different solvents. extent. The wide varieties of cations including both metal and
Geerlings and co-workers subjected pyridine-substituted ben- inorganic ions, employed in the study underline the generality of
zene model complexes to MP2 level calculations in order to the results obtained.
ascertain influence of stacking on hydrogen bonding.384 This Martiń and co-workers have shown cooperativity between
study showed that the hydrogen-bonding capacity of the π−π and hydrogen-bonding interactions through the formation
nitrogen atom was directly related to the electrostatic interaction of a highly stable supramolecular complex formed from C60 and
between the cycles, more so to the electron-donating/with- exTTF (2-[9-(1,3-dithiol-2-ylidene)anthracen-10(9H)-yli-
drawing character of the substituents. Rotello and co-workers dene]-1,3-dithiole).387 Grimme et al. studied the three-body
2792 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 12. Pictorial representation of absence of significant cooperativity in benzene clusters (ref 406) employed as model systems to understand
behavior of multiple π−π and CH−π coexisting with each other. Uniformly similar values of complexation energy per interaction in the two varied
arrangements in spite of increasing cluster size suggest additive nature of stacking and CH−π interactions in benzene clusters.

contribution to the interaction energy for a benchmark MP2 ring by analyzing data in CSD.393 Zhu et al. have also undertaken
database of nucleic acid base trimers.251 They employed an a survey of the CSD for structures in which halogen bonds and
energy decomposition analysis scheme having similar definition aromatic stacking interactions are present.394 They reported
of energy terms to the Morokuma decomposition scheme to cooperative effects in complexes in which the aromatic rings
reveal how nonadditive effects were mainly attributed to the behave both as electron donors in halogen bonds and as electron
induction term, while exchange repulsion, electrostatic, and acceptors in π−π stacking interactions. Strong cooperativity was
dispersion contributions were essentially additive. Frontera and also shown to occur in the complexes involving pyrazine rings as
co-workers have reported the presence of strong synergistic Lewis bases in some of the complexes.
effects in systems, wherein π−π interactions modeled using 7.2. In Molecular Clusters
benzene and hexafluorobenzene, coexisted with cation−π and
Besides the above-mentioned studies, investigations into
anion−π interactions in the same complex.294 Very high
molecular clusters which emphasize interplay between stacking
nonadditivity energies ranging up to −20.3 kcal/mol were
and CH−π interactions are also discussed in order to evaluate the
observed, indicating how cation−π and/or anion−π interactions
presence of cooperativity and anticooperativity in these
have a strong influence on the π−π interaction and vice versa.
interactions.397−414 The benzene dimer has served as a prototype
Presence of synergistic effects in complexes wherein edge-to-face for π−π and CH−π interactions.397,398 Tsuzuki and co-workers
aromatic interactions and hydrogen-bonding interactions coexist have reported the CCSD(T) interaction energy of the benzene
have also been studied.388 Besides, a minor impact of X−H/π (X dimer at the basis set limit for parallel, T-shaped, and slipped-
= C, N, and O) interaction on the π−π interaction and vice versa parallel orientations revealing that dispersion interaction is
has also been reported.389 Ab initio calculations employing the indeed the major source of attraction in the benzene dimer,
RI-MP2 method on a choice model of interacting molecules, although electrostatic interactions appear to be very important
namely 1,4-diaminobenzene and terephthaldehyde demonstra- for determining the directionality of the constituent benzene
ted that synergetic effects were present in complexes where molecules in cluster formation.397 Computations have been
hydrogen-bonding interactions, cation−π, and π−π interactions limited to dimers, trimers, and few tetramers of benzene
coexist.390 Ebrahimi et al. performed ab initio calculations on X- molecules due to the huge requirement of computational
ben∥pyr···H−F (X = NO2, CF3, CN, F, Cl, CH3, and OH) resources for the higher clusters.399−404 Heuristic search
complexes to gauge the effect of π-stacking and hydrogen- techniques employing genetic algorithm for benzene cluster
bonding interactions on each other.391 Cooperativity of optimization have also been pursued.405
interactions in these complexes was noted where face-to-face Through an exhaustive study on clusters of benzene (Bz)n, n =
aromatic interactions and hydrogen-bonding interactions 2 to 8, at the MP2/6-31++G** level of theory, we demonstrated
coexist. Balasubramanian and co-workers have used several how the cooperativity in benzene clusters is quite insignificant,
structural parameters such as hydrogen bond length to measure which is in sharp contrast to the way in which hydrogen bonding
cooperativity in oligomers of benzene-1,3,5-tricarboxamide and and some other noncovalent interactions operate.406 In order to
to study stacking based on quantum chemical calculations.392 quantify the extent of cooperativity in a set of benzene clusters
Zarić and co-workers have shown remarkably stronger stacking (Figure 12), the complexation energy (CE) per interaction
interactions of pyridines with hydrogen bonds due to local between a pair of benzene molecules (CE/(n − 1), n = number of
parallel alignment interactions of OH bonds with the aromatic benzene monomers in the cluster) was compared from trimer to
2793 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

hexamer. A near uniform CE/(n − 1) value of ∼3.3 kcal/mol proteins was undertaken. The study was also directed to reveal
from trimer to hexamer in the case of parallel displaced stacked any specific network and connectivity of aromatic rings in the
clusters at the MP2/6-31G* level of theory was observed. The proteins. Although the distribution of π−π networks was found
corresponding value of CE/(n − 1) is approximately −6.5 kcal/ to be ubiquitous across all family of proteins, a higher propensity
mol at the MP2/6-31++G** level of theory for the same to be perpendicular to each other (CH···π) rather than parallel
structures. The addition of another benzene molecule did not was observed, considering orientation of π−π networks.
seem to enhance the strength of each individual interaction. This Using R2PI (resonant two-photon ionization) spectroscopy in
observation was quite in contrast to earlier reported studies concert with electronic structure calculations, Reid and co-
dealing with noncovalent interactions such as cation−π, π−π, workers have recently shown studies on the homoclusters of
and hydrogen-bonding interactions. In the case of clusters where chlorobenzene and also on bromobenzene and mixed homo-
CH−π interactions were more prominent, a similar absence of benzene clusters, as prototypical systems where π−π stacking,
cooperativity was observed. The CE per interaction values there CH/π interactions, and halogen-bonding interactions may all be
were in the range from −2.7 to −4.6 kcal/mol at the MP2/6- present.395,408 Sherrill and co-workers demonstrated that the
31G* level and consistently about a couple of kcal/mol higher at interaction energies in larger benzene clusters were obtained as
the MP2/6-31++G** level. An inspection of all the clusters and the sum of interaction energies for isolated benzene dimers, with
their CEs revealed that the interaction energies were additive an error of less than 10% for the systems considered.398,399
irrespective of the number or type (CH···π or π−π) of Waller and co-workers revealed parallel π-stacking was not a
interactions. Hence stacking interactions in benzene clusters good structural motif for promoting cooperative bindings in gas-
seem to be essentially additive in nature. The cooperativity phase homotrimers.409 Further, anthracene trimers which
observed in clusters of polar molecules is due to induction adopted predominantly stacked geometries showed additive
interactions which play an important role in determining their three body interaction energy reflective of the absence of
structure. As electrostatic interactions contribute little in the case cooperativity, similar to pyrazine oxide and uracil trimers.
of benzene clusters, the resultant small induction interactions Additivity of π−π or π+−π stacking and T-shaped interactions
may be the cause for negligible cooperativity that is observed. between natural or damaged DNA nucleobases and histidine was
In order to investigate how large π-networks can enhance the investigated by Wetmore and co-workers.410 Although both
stability of a protein conformation and influence protein−ligand stacked and T-shaped interactions, as well as both π−π and π+−π
interactions, our group developed an Aromatic−Aromatic interactions, were shown to exhibit geometric additivity, they
Interaction Database, A2ID (Figure 13).407 In continuation of obtained a small magnitude of additivity energy < 0.48 kcal/mol,
earlier efforts on understanding how cooperativity can suggesting that these interactions are additive. The interaction
strengthen the multiple interactions considerably, a qualitative energy analysis however does not reveal whether the
and quantitative study estimate for the 2π, 3π, and higher nucleobase−amino acid interactions are equivalent in dimers
aromatic repeat units within a threshold distance cutoff in and trimers, or whether they differ in different-sized complexes
but sum to yield an overall additive interaction. An analysis of
how the biological backbone affects the π−π stacking
interactions between conjugated nucleobases and amino acids
has also been reported.411

8. MODULATION OF COOPERATIVITY DUE TO


SOLVATION
The presence of solvent and its impact on cooperativity have
been considered using both experimental and theoretical
studies.415−434 By determination of dynamic vapor pressure
measurements, Wolfenden et al. explored cooperativity and
anticooperativity in solvation by water for imidazoles, quinones,
nitrophenols, nitrophenolate, and nitrothiophenolate ions.415
Collum and co-workers showed that mixtures of ether and amine
solvents offer relative dimer binding energies in some cases and
provided evidence of cooperative solvation due to an absence of
mixed solvates.416
Addressing the question on whether cooperativity in hydro-
phobic association could be reproduced correctly by implicit
solvation models, Scheraga and co-workers evaluated the
potential of mean force (PMF) of the hydrophobic association
of pairs and triplets of nonpolar solute molecules with various
sizes and strengths of van der Waals interactions.417 Umbrella
sampling MD simulations combined with the weighted histo-
gram analysis method were employed for this purpose. The study
Figure 13. Schematic representation of the selection of π−π motifs
using a virtual sphere in the Aromatic−Aromatic Interaction Database
revealed that distance dependence of the PMF, including the
(A2ID), ref 407. The coordinate system used to define the centroid cooperative contribution, was described qualitatively correctly by
position of the observed molecule (CB) relative to the reference the variation of either the molecular surface area or the excluded
molecule (CA). The interplanar angle ϕ and normals to the planes of π- solvation-shell molecular volume with the distance between the
systems, Nref and Nn+1. RNM is the distance between the two π-systems, solute molecules. Jordan and co-workers have studied the
and R is the cutoff distance. infrared spectra of I−·(CH3OH)n·Arm, n = 1,2 clusters, obtained
2794 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

via argon and methanol predissociation, with the aid of ab initio complexes.430 The solvent effect on the halogen bond in the
calculations of their OH stretching fundamentals.418 A dramatic H2CS−BrCl complex was also studied, and it was found that the
OH red-shift relative to both the symmetric isomer and the n = 1 solvent has a prominent enhancing effect on it. Evaluating
complex was traced to a cooperative hydrogen bonding which is solvation effect, particularly in terms of its impact on
only operative in the asymmetric form. cooperativity, is vital as there are a number of different molecules
Using correlated ab initio calculations, Scheiner and co- of real biological relevance which are capable of forming a
workers have shown that conventional and OH···ϕ hydrogen multiplicity of hydrogen bonds and other noncovalent
bonds obey the normal principles of cooperativity, wherein these interactions all at the same time. Although the last few decades
bonds are strengthened when a central molecule serves have seen a remarkable development of methods, techniques,
simultaneously as both proton donor and acceptor, whereas and algorithms to study solvation problems computationally,
CH···O bonds do not appear to be amenable to such positive there is still a need to dwell further into understanding the role of
cooperativity.419 As such, when placed in a polarizable medium, solvation on cooperativity.
all hydrogen bonds weaken as the dielectric constant of the
solvent increases. Spectroscopic studies of solvated hydrogen 9. COOPERATIVITY IN BIOCHEMISTRY
and hydroxide ions at aqueous surfaces have been reported.321 The term cooperativity in biochemistry is used in a range of
The role of hydrogen-bonding cooperativity is investigated in processes such as preorganization, avidity, allostery, and some
these studies.420 types of macro molecular assembly. The underlying phenomen-
Microsolvation of the cation−π interaction has been explored on in cooperative processes as is typically encountered is one in
in detail by our group.421,422 The sequential attachment of water which the binding of a ligand to one site alters the affinity of a
molecules to cation−π (Li+-benzene, K+-benzene, and Mg2+- separate ligand molecule for binding at another site. Such a
benzene) systems revealed how solvation of the metal ion cooperativity requires allosteric communication between binding
decreases its interaction energy with the π system, while the sites. Whitty describes the role of configurational preorganization
solvation of the π system increased its interaction energy with the to be a second important phenomena underlying different
metal ion. The cation binding to benzene enhanced the ability of cooperative processes, for example those involving polyvalent
aromatic protons to participate in hydrogen bonding with water ligands or even as observed in protein folding.435 Besides,
molecules. An interesting modulation of the strength of the consideration of protein folding illustrates that cooperativity is
cation−π interaction as a result of other noncovalent interactions not solely an equilibrium phenomenon but also a strategy used
was observed through these studies. Subsequently, in separate by nature to increase reaction rates.435 Cooperativity may also be
studies on metal ion solvation423 including one modeling Na+ more broadly defined as a process for which intermediates are
solvation424 and a second on Mg2+ solvation425 with increasing disfavored (resulting, for example, in a two-state conformational
number of water molecules (n = 1−20) employing DFT and change).436 Indeed, Williamson suggests that understanding
MP2 level calculations, the variation in the stability of the cooperativity provides an insight into the mechanisms that
complexes with the same number of water molecules could be govern assembly and disassembly of multicomponent com-
attributed to the hydrogen-bonding patterns observed in them. plexes.437
The strength of hydrogen bonding between the surrounding 9.1. Allostery and Binding Cooperativity
water molecules in the different coordination shells along with The concept of cooperativity in binding is typically introduced
the metal−water interaction was shown to play a key role in through a discussion of oxygen binding to hemoglobin (Figure
determining the stability of the complex. 14).438−447 Eaton and co-workers have provided an account of
Kuo and co-workers adopted a combined theoretical,
spectroscopic approach to study how when a solvent molecule
is replaced by another molecule with larger proton affinity, the
strength of all other hydrogen bonds decreases.426 The concept
of anticooperativity by successive substitution was considered in
a mixed solvation system. H+(CH3OH)m(H2O)n (m + n = 5 and
6) mixed clusters were employed to demonstrate hydrogen bond
weakening. Recently Bakker and co-workers have shown an
important study wherein a combined terahertz and femtosecond
infrared spectroscopic study of water dynamics around different
ions including magnesium, lithium, sodium, and cesium cations,
as well as sulfate, chloride, iodide, and perchlorate anions, reveals
that the effect of ions and counterions on water can be strongly
interdependent and nonadditive and in certain cases extends well Figure 14. Depiction of cooperative binding in hemoglobin.
beyond the first solvation shell of water molecules directly
surrounding the ion.427 Peruchena and co-workers have
performed ab initio calculations to demonstrate that the the novel physical experiments and rigorous quantitative analysis
hydrogen bonds are stronger in the dihydrated complexes than conducted by several scientists over the decades to elucidate two-
in the corresponding monohydrated complexes, wherein the σ- state allosteric mechanism of hemoglobin.438 Starting with the
and π-electron delocalization were found.428 The importance of original work of Hill in 1910 on the possible effects of the
ion cooperativity in understanding the osmotic coefficients of aggregation of the molecules of hemoglobin on its dissociation
salts and their effects on water surface tension has been explored curves,439 a large number of studies have been evolved that
by Gao through simple theoretical models.429 The cooperative provide models to rationalize the O2 binding curve of
effect of halogen bonding was considered by Cheng and co- hemoglobin and its associated cooperativity.440−443 Important
workers in H2CS−XY (XY = FF, ClF, ClCl, BrF, BrCl, and BrBr) among the models formulated to explain this cooperativity was
2795 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 15. Structures from PDB showing coexistence of different noncovalent interactions.

the two-state concerted MWC model put forth by Monod, ligand binding along with consideration for the earlier mentioned
Wyman, and Changeux in which with only three parameters MWC model and structural data.446 Ackers et al. later reported
allosteric cooperativity was adequately described.444 Employing a how the allosteric mechanism was controlled by quaternary
comparison between X-ray structures of liganded hemoglobin switching from form T to form R occurs whenever heme-site
and its deoxy form, Perutz filled in the lack of a structural
mechanism in the MWC model.445 Karplus and co-workers binding creates a tetramer with at least one ligated subunit on
subsequently presented an allosteric model of the statistical each dimeric half-molecule.447 Cooperativity was demonstrated
mechanics of cooperativity to account for the dependence on to arise from both concerted quaternary switching and sequential
proton concentration, as indicated by the pH-dependence of modulation of binding within each quaternary form, T and R.
2796 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Early discussions on cooperativity of biological membranes in the importance of understanding how cooperativity provides
terms of their lattice structure448 and suggestions of cooperative insight into the mechanisms that govern assembly and
transitions triggered by either a change in electric field449 disassembly of multicomponent complexes.437 The quantifica-
(membrane potential) or in a ligand concentration or both have tion of cooperativity in various multicomponent assemblies
been reported. Rebek et al. have shown that processes involving through the construction of thermodynamic cycles and under-
small molecules in solution can be subjected to control involving standing the underlying dynamic equilibrium of the various
conformational changes induced by binding, and they also binding components, their rates of the association and
reported an example of a model system which shows binding disassociation and associated free energies were discussed.437
cooperativity between two remote sites.450 Studies exploring the Mariuzza and co-workers237 used surface plasmon resonance
consequences of thermodynamic stability on generalized binding techniques and mathematical modeling to describe the
phenomena in biopolymers have also been undertaken.451,452 energetics of cooperativity in a trimolecular protein complex.
Allosteric cooperativity continues to capture current inter- They studied the ternary complex formed by the simultaneous
est.453−473 Bashford and co-workers suggested a mechanism interaction of a superantigen with major histocompatibility
which requires no conformational change to produce a complex and T cell receptor to show how the system exhibits
cooperative effect considering cooperative binding of charged both positive and negative cooperativity. Sleat et al. provided
ligands to proteins.461 The ligand binding sites were electro- genetic evidence for nonredundant functional cooperativity
statically coupled to protein side chains that can undergo between two different proteins involved in lipid transport.479
protonation and deprotonation. Laughton and co-workers also Zandany and co-workers report a strategy for direct analysis of
provided an example of the concept of allostery without cooperativity in multisubunit allosteric proteins.480 They
conformational change. This was done through a series of MD measured effects of all possible combinations of mutated
simulations on the free DNA, the 1:1 complex, and the 2:1 subunits on protein function, along with analysis of the hierarchy
complex with ligand Hoechst 33258 designed to enable of inter subunit interactions, assessed by using high-order double
calculation of thermodynamic parameters associated with the mutant cycle coupling analysis. Blacklow and co-workers
molecular recognition events.462 Fully solvated MD simulations reported the X-ray structure of a dimeric human Notch1
and quasiharmonic normal-mode analysis were performed on transcription complex loaded on the paired site from the human
chloroeremomycin, vancomycin, and dechlorovancomy- HES1 promoter.481 Analysis of this structure helped reveal how
cin.463,464 The study revealed that biochemical cooperativity promoter organizational features control cooperativity and the
could be mediated through changes in vibrational activity, responsiveness of different promoters to Notch signaling.
irrespective of the presence or absence of concomitant structural Selected examples of crystal structures from PDB showing
change. An exoditopic artificial receptor based on a hydrindacene coexistence of different noncovalent interactions are shown in
platform provided evidence that cooperativity in amide hydrogen Figure 15.
bonding by polarization plays a part in the positive homotropic A large number of studies have shown that the activation of ion
allosteric binding property with benzenediols.465 Through a channels are mediated through cooperativity.482−487 The highly
study on a tris-cation receptor wherein a clustering of cooperative opening of calcium channels by nanomolar
electrostatic interactions allows the efficient complexation of concentrations of inositol 1,4,5-trisphosphate (IP3) enables
amino acid carboxylates in water, Schmuck et al. have cells to detect and amplify very small changes in the
emphasized how cooperativity functions even in small and concentration of IP3 in response to hormonal, sensory, and
flexible artificial receptors.466 Hilser and co-workers have growth control stimuli was shown by Holowka and co-
investigated whether, in addition to serving as structural scaffolds workers.488 They investigated the kinetics of IP3-induced Ca2+
for recognition and catalysis, active-site residues may also play a
release from intracellular storage pools in permeabilized rat
role in modulating the cooperative network.467 They employed
basophilic leukemia cells to gain insight into the molecular
an experimentally validated ensemble-based description of
mechanism of action of IP3. Qian and co-workers examined the
proteins to elucidate the extent to which perturbations at
mechanism of cooperative activation of big potassium (BK)
different sites can influence the cooperative network in the
channels by Ca2+.489 Smith-Maxwell et al. have studied the role of
protein. A study on a saccharide receptor achieving concen-
the S4 transmembrane segment in channel gating by constructed
tration-dependent mannoside selectivity through two distinct
chimeras in which S4 segments from several divergent potassium
cooperative binding pathways has been reported.468
channels, Shab, Shal, Shaw, and Kv3.2, were inserted into a
Goodey et al. have reviewed how conformational mobility
serves as the common route between allosteric regulation and Shaker potassium channel background.490 Tombola et al. have
catalysis, more so in the context of the role played by allosteric clearly shown how the opening of the two pores of the Hv1
effects in drug design and protein engineering.469 Veglia and co- voltage-gated proton channel is tuned by cooperativity.491 Zhou
workers mapped the allosteric network in the catalytic subunit of investigating cooperativity in twin gatings reported patch-clamp
protein kinase A using NMR spectroscopy.470,471 Positive fluorometry measurements of cAMP binding and channel
allosteric cooperativity was shown to be generated by nucleotide opening combined with global kinetic fitting to reveal that
and substrate binding during the transitions through the major ligand binding to the tetrameric hyperpolarization-activated
conformational states including apo, intermediate, and closed cyclic nucleotide-gated channel is favored at the second and
conformers. Kremer et al. have recently emphasized the fourth binding events.492 Fujiwara et al. focused on the molecular
implementation of allosteric control in artificial receptors with basis for dimer formation and intersubunit coupling in Hv1/VSO
different combinations of effectors.472 P, a dimeric voltage-gated H+ channel in which the gating of one
subunit is coupled to that of the other subunit within the
9.2. Multimeric Proteins, Ion Channels, and Enzymes dimer.493 A cytoplasmic coiled-coil mediates the cooperative
Cooperativity serves as an important factor in the functioning of gating temperature sensitivity in this voltage-gated H+ channel
multimeric complexes.474−478 Williamson has recently addressed Hv1.
2797 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

The activity of many enzymes are regulated by allosteric ative effects of extended hydrogen networks were also shown to
effectors.34,457,494−498 Salahub and co-workers reported a study make significant contributions to δH. Houk and co-workers
employing DFT calculations to reveal a large cooperative effect studied the strong reversible binding of biotin by avidin and
induced by the interaction of an anion with a peptide hydrogen streptavidin using DFT and ab initio methods.510 Five hydrogen
bond network, directly or through an imidazole molecule.499 An bonds were shown to act cooperatively leading to stabilization
increase of the cooperative effect on going from a neutral to a that is larger than the sum of individual hydrogen-bonding
negatively charged ligand was shown to be sufficiently large that energies. Also a charged aspartate was shown to be the key
it could make an important contribution to the catalysis by residue that provides the driving force for cooperativity in the
stabilizing the charge formation along the reaction pathway in the hydrogen-bonding network for both avidin and streptavidin by
enzyme-catalyzed reactions. Wulff and co-workers have explored greatly polarizing the urea of biotin. Recently Gibson and co-
cooperativity of an amidinium and a copper ion in a transition- workers have captured cooperative interaction data in a
state imprinted cavity giving rise to high catalytic activity based structured format that is backward compatible with PSI-MI
on a model of carboxypeptidase A.500 Single molecule (Proteomics Standards Initiative of the Human Proteome
enzymology have demonstrated dynamic effects that were Organization)-based data and applications.511 Oxyanion holes
concealed entirely by ensemble-averaging techniques.501 Studies commonly found in many enzyme structures are crucial for the
with mutant human Monoamine Oxidase B revealed an allosteric stabilization of high-energy oxyanion intermediates or transition
mechanism where the increased binding affinity of ligand results states through hydrogen bonding.512−517 Site-directed muta-
from a cooperative increase in hydrogen bond strength through genesis studies on the catalytic triad, a group of three amino acids
formation of a more hydrophobic milieu.502 Hammes-Schiffer that are found in the active sites of some proteases, reveal the
and co-workers have proposed a general mechanism for enzyme importance of the oxyanion hole for catalysis.518 Thus, the
catalysis that includes multiple intermediates and a complex, importance of cooperativity has been established in the structure
multidimensional standard free energy surface.503 Flexibility, and function of a large number of proteins.438−520
diversity, and cooperativity have found emphasis as pillars of 9.3. Drug Receptor Interactions
enzyme catalysis.
Positive cooperativity in some reactions involving cytochrome Williams and co-workers have contributed substantially to
P450 3A4 which oxidizes a diversity of substrates, including understanding associations between two or more molecules
various drugs, steroids, carcinide natural products, have been besides considering the effect of interactions with solvent
reported.504 Guengerich and co-workers observed the presence molecules and the changes in the internal structure of the
of positive cooperativity in oxidations of several substrates, associating molecules on binding.28−30,521−524 Of particular
including testosterone, 17β-estradiol, amitriptyline, and aflatoxin significance is the study on the cooperativity between binding of
(AF) B1 in systems containing purified recombinant bacterial cell wall precursor analogs/ligands to and antibiotic dimerization
P450 3A4.505 Studies on construction and characterization of a of the clinically important vancomycin group antibiotics.523 An
site-directed mutant that displays hyperbolic steroid hydrox- analysis of the origins of a cooperative binding energy of
ylation kinetics to analyze human cytochrome P450 3A4 dimerization illustrates how, when a protein has a loose structure,
cooperativity were also reported.506 Oostenbrink and co-workers the binding energy of another molecule to the protein can derive
adopted a combination of MD simulations and free-energy in part from changes occurring within the protein. The above-
calculations to elucidate the physicochemical origin of the mentioned result was indeed important for drug design as it is
observed positive homotropic cooperativity in ketoconazole indeed common practice to seek the origins of binding affinity at
binding to cytochrome P450 3A4.507 The binding of the first the interface formed between associating entities. An illustration
ketoconazole molecule was established to increase the affinity for of how partitioning of binding free energy for each of the
the binding of the second ketoconazole molecule by 1.2 kcal/ interactions which are made or broken on binding, in particular
mol, thereby explaining the experimentally observed cooperative for the hydrophobic effect, ignores the associated cooperativity
behavior of cytochrome P450 3A4. Shape complementarity made by them.29 An exploration of the relationship between
through van der Waals interactions was identified as the main noncovalent structure and free energy of binding was undertaken
driving force of such a binding. Qian and co-workers have to include the roles of enthalpy and entropy of association.
suggested a cooperativity in the phosphorylation-dephosphor- Considering dimerization of vancomycin-type antibiotics pos-
ylation cyclic reactions which is temporal, with energy “stored” in itive cooperativity was shown to manifest itself in a more
time rather than in space as seen in allosteric cooperativity.508,509 favorable enthalpy of association and a partially compensating
This was done through stochastic models for different molecular less favorable entropy of association.30 The study suggests that
interactions involving signal transduction modules. sets of noncovalent interactions that are made with positive
Phenolic ligand binding characteristics of the insulin hexamer cooperativity must be mutually enhancing in their free-energy
were employed to develop a quantitative model for a symmetry- benefits. While conversely, sets of noncovalent interactions that
asymmetry-based cooperativity mechanism.236 A systematic are made with negative cooperativity must be mutually
analysis of crystal data using Secbase suggests that cooperative weakening in their free-energy benefits.524 Besides, changes in
effects which mutually polarize spatially well-aligned hydrogen the relative populations of the monomer and asymmetric dimer
bonds are present either in α helices and parallel β sheets, forms of ristocetin A, upon binding of two molecules of ligand,
whereas such influences seem to be lacking in 310 helices and showed that ligand binding was negatively cooperative with
antiparallel β sheets. Jensen and co-workers developed a respect to dimerization.
computational methodology for backbone amide proton Several studies have established the role of different non-
chemical shift (δH) predictions based on ab initio quantum covalent interactions in drug receptor interactions observed in
chemical treatment of part of protein G and ubiquitin.170 therapeutically important targets.525−546 By use of crystal
Hydrogen bond geometry was indicated to be the most structure analysis and isothermal titration calorimetry for a
important δH-determinant, besides which longer-range cooper- congeneric series of thrombin inhibitors, Klebe and co-workers
2798 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

have shown how extensive cooperative effects between hydro- contributions to the folding barrier were found to be
phobic contacts and hydrogen bond formation are intimately considerable. They subsequently reported the noncooperative
coupled via dynamic properties of the formed complexes.525 A folding behavior of the de novo designed protein Top7, thereby
complex picture of mutually competing and partially compensat- raising the possibility that evolutionary selection for biological
ing enthalpic and entropic effect determined the nonadditivity of functions might have led to folding cooperativity.562−564 Plaxco
free-energy contributions to ligand binding at the molecular and co-workers demonstrated that traditional Go̅-potential
level. Two separate series of thrombin inhibitors containing polymers in which only inteactions present in the native state
hydrophobic side chains of increasing size that bind in the S3 are favorable lack the extreme cooperativity that characterizes the
pocket with or without an adjacent amine that engages in a folding of naturally occurring, two-state proteins.565,566 The
hydrogen bond with Gly 216 were studied. The presence of an folding rates of a diverse set of Go̅ 27-mers were shown to be
adjacent hydrogen bond was shown to enhance binding affinity poorly dispersed and effectively uncorrelated with native state
per Å2 of hydrophobic contact surface in the S3 pocket by 75% topology. Zhou et al. employed an all-atom off-lattice model of
and 59%, respectively, over the inhibitors lacking this hydrogen the fragment B of staphylococcal protein A to suggest that a large
bond.526 This demonstrated cooperativity between the hydro- well-packed solid core is the origin of the folding cooperativity of
phobic interaction and the hydrogen bond. Hangauer and co- proteins.567 Thirumalai and co-workers have also probed into the
workers have shown water-mediated ligand functional group effect of finite size on cooperativity and the rates of protein
cooperativity through a study on four thermolysin phosphona- folding using lattice models with side chains, off-lattice Go̅
midate inhibitors.527 The differential binding free energy and models, and the available experimental data.568,569
enthalpy was caused by replacement of a hydrogen with a Me A few recent experimental studies570−573 considering
group, which binds in the well-hydrated S2′ pocket and is more cooperativity of protein folding are discussed below. Dobson
favorable in the presence of a ligand carboxylate. Klebe and co- and co-workers employed hydrogen-exchange experiments
workers have recently shown a surprising nonadditivity of monitored by NMR and mass spectroscopy to reveal the impact
functional group contributions for the carboxylate and/or methyl of a specific mutation in human lysozyme, in terms of its ability to
groups in the binding of four congeneric peptide-like reduce the stability of the beta domain and an adjacent C-helix in
thermolysin inhibitors where the ligands differ only by a terminal the native structure.571 Further, mass spectroscopy data revealed
carboxylate and/or methyl group.528 the high degree of cooperativity involved in the transient folding
9.4. Protein Folding
of these regions. Clarke and co-workers have addressed the
apparent cooperativity in the folding of multidomain proteins, in
A large number of experimental and theoretical studies have particular that of pairs of three-helix bundle spectrin domains and
aided in providing valuable insights into understanding the found that it depends on the relative rates of folding of the
mechanism of protein folding.55,162,547−597 These include constituent domains.572 Equilibrium studies done in the study
seminal studies on different limiting mechanisms of protein indicate that equilibrium data alone are insufficient to describe
folding highlighted by the two main models at the time, the the folding of multidomain proteins and to quantify the effects
nucleation-propagation mechanism and the diffusion-collision that one domain can have on its neighbor. Shank and co-workers
model and identification and characterization of protein-folding have used a single molecule optical tweezers approach to induce
pathways and their intermediate states.548 Several subtopics the selective unfolding of particular regions of T4 lysozyme to
addressed in this context included proline isomerization, protein investigate how the topological organization of a protein affects
fragments, and disulfide bond intermediates.549 Simulation and the coupling and folding cooperativity between its domains.573
theoretical models have played an important role in shaping the The study showed that the topological organization of the
understanding of protein folding.550 Kinetically distinct pathways polypeptide chain critically determines the folding cooperativity
for folding were identified showing that the folding energy between domains. Proteins may have evolved to select certain
landscape was more complex than had been originally topologies that increase coupling between regions to avoid areas
thought.551 The process of protein folding is characterized by a of the landscape that lead to misfolding. Gellman and co-workers
high degree of cooperativity manifested by the almost negligible have described a 20-residue peptide that adopts a triple-stranded
population levels of partially stable intermediate states observed antiparallel β-sheet conformation in aqueous solution.574 This
during transition between folded and unfolded states.552,553 Dill model system was used to show that antiparallel β -sheet
et al. suggested a hydrophobic zipper hypothesis to determine formation is cooperative, perpendicular to the strand direction
the physical basis of cooperativity by which proteins avoid based on site-specific conformational data from NMR spectros-
exhaustive searching of conformational space.557 Hydrophobic copy. Indeed peptide β-sheet systems have emerged as context
zipper cooperativity for heteropolymer collapse was shown to be independent models for probing secondary structure propen-
driven by nonlocal interactions mediated by the solvent leading sities and aspects of cooperativity both parallel and perpendicular
to helical, sheet, and irregular structures, depending on the amino to the strand direction.575,576 Simmerling and co-workers
acid sequence. More recently, a theory for protein-folding performed microseconds of standard- and replica-exchange
stability and cooperativity for helix bundle proteins has also been MD simulations on designed three-stranded β-sheet structures
suggested by them.558 to understand cooperativity in their folding.577 Using MD
Chan and co-workers studied protein folding cooperativity by simulations, Hills et al. demonstrated that the hydrophobic
comparing different theoretical models.55,559−561 On the basis of cooperativity seen in the self-association of a model amyloido-
simple lattice models, they suggest that proteinlike thermody- genic peptide STVIYE was sufficient to allow for nucleation-
namic cooperativity may require a cooperative interplay between dependent polymerization with a pentamer critical nucleus.578
local and nonlocal interactions.560 Baker and co-workers Findings of possible noncooperative, folding have also been
reported an important statistical study which proposed that reported.582−585 Two types of downhill folding, one where
native state topology, and not sequence itself, was the main certain mutants of λ-repressor behave in a two-state manner near
determinant of folding pathways.555 Besides the entropic their transition midpoint and under mildly folding conditions,
2799 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

but they can fold downhill under strongly folding conditions, and anucleosidic units on both strands. Deyà and co-workers
second where some proteins can undergo global downhill computed the cooperativity energies in two models of Z-
folding, which means that the protein always behaves in a one- DNA.611 The first model demonstrated how the lone pair−π
state manner for the entire range of experimental conditions, interaction and hydrogen bonding reinforce between guanine
have been reported.561 The results of a more recent study of a and ether, whereas a second model studied the influence of
model oligomer of human islet amyloid polypeptide (hIAPP) stacking interaction between the cytosine bases upon the
segments explored by means of all-atom MD simulations under hydrogen bonds and vice versa. Marek and co-workers
different force fields supports a picture of downhill-like investigated structural and energetic features of model guanine
cooperative assembly of β strands during the fibrillation and xanthine containing DNA quadruplexes.612 Emphasizing the
process.586 Diehl et al. have developed a biosynthetic approach
physicochemical nature of the noncovalent interactions involved,
to control and probe cooperativity in multiunit biomotor
assemblies by linking molecular motors to artificial protein they established that hydrogen bonding makes the greatest
scaffolds.587 Cooperative interactions between monomeric contribution to the internal stability of the DNA quadruplexes,
kinesin-1 motors attached to protein scaffolds were shown to whereas the aromatic base stacking and ion coordination terms
enhance hydrolysis activity and microtubule gliding velocity. which are similar account for the rest. EDA analysis also revealed
Thus, cooperativity in protein folding has been addressed in how guanine containing structures seemed to benefit from a high
varying model systems of interest.55,162,547−597 degree of hydrogen bond cooperativity, whereas xanthine-
9.5. Transcription containing models appeared to be characterized by a more
favorable and cooperative π−π stacking. In a summary on the
Transcription regulation in prokaryotes and eukaryotes is known design and development of a class of molecular duplexes with
to occur through the coordinated action of transcription factors programmable hydrogen-bonding sequences and adjustable
and DNA.524,598−608 The role of cooperativity in this process has
stabilities, Gong has emphasized cooperativity of multiple
been investigated by several studies, some of which are
mentioned below. Hammer and co-workers have investigated hydrogen bonding as one of the most important features
the long-range cooperativity between gene regulatory sequences associated with the self-assembly of DNA.613 In a recent study on
in prokaryote by reporting that transcription initiation can be DNA-based nanostructures built on a long single-stranded DNA
regulated from an operator site placed one to five kilobases scaffold known as DNA origamis, an attempt was made to model
downstream of the promoter and the transcribed gene.601 Zhang the annealing and melting properties of DNA constructions.614
and co-workers developed a method which integrates genome- Cooperativity between staples was shown to be critical to
wide gene expression data and chromatin immunoprecipitation quantitatively explain the folding process of DNA origamis.
(ChIP-chip) data to discover biologically relevant synergistic Oddershede and co-workers have recently demonstrated the
interactions between different transcription factors.602 Given any importance of DNA supercoiling as a factor to enhanced
pair of transcription factors, a novel measure of cooperativity cooperativity and efficiency of an epigenetic switch operating in
between two transcription factors based on the expression bacteriophage λ.615
patterns of sets of target genes was developed. Herschlag and co-workers report a direct quantitative measure
9.6. DNA/RNA of tertiary contact cooperativity in a folded RNA.616 The stability
Although there has been a substantial emphasis on under- of an independently folding P4−P6 domain from the
standing cooperativity in proteins with regard to structure, Tetrahymena thermophila group I intron was measured by single
stability, and also their binding with other molecules, the extent molecule fluorescence resonance energy transfer. An enhance-
of cooperativity both in DNA and RNA molecules has been ment of P4−P6 stability by 3.2 ± 0.2 kcal/mol as a consequence
analyzed to a much lesser extent.609−620 The binding of of cooperativity between the two tertiary contacts was reported
surfactant ions to DNA macromolecules is highly cooperative by them. Strobel and cowokers employed nucleotide analog
due to a hydrophobic effect.609 Mel’nikov et al. have shown on interference mapping and mutagenesis to determine the possible
the level of individual DNAs, that salt-induced unfolding sites for glycine binding and tertiary cooperativity in case of the
transition of the globules is largely discrete.609 For the ensemble glycine riboswitch.617
average of the DNAs, the transition is discrete with Chance and co-workers have addressed questions related to
acetyltrimethylammonium bromide but is continuous with RNA structure by combining chemical denaturation and
another cationic surfactant distearyldimethylammonium bro- hydroxyl radical footprinting methods.579 They determined
mide. The discreteness for the coil−globule transition in the unfolding isotherms for each of 26 discrete sites of protection
ensemble of DNAs complexed with the first surfactant is located throughout the Tetrahymena thermophila group I
attributed to the existence of the phase transition in whole over ribozyme and showed how intradomain contact and three
the bulk solution. The cooperativity of assembly of two nucleic interdomain contacts show high cooperativity, indicating that
acid strands is determined by the stacking interactions of
these contacts exhibit global cooperativity in their folding
consecutively formed base pairs and by the nature and
conformational properties of the sugar−phosphate backbone. behavior. Bevilaqua and co-workers found that the small DNA
Leumann and co-workers investigated the dependence of duplex hairpins fold in a highly cooperative manner with indirect
melting cooperativity between two double helical domains on the coupling, while their RNA counterparts fold in a much less
length and the conformational properties of the intervening cooperative fashion and display direct coupling.580 The study
sugar−phosphate backbone in a model system.610 They found also indicated that folding cooperativity in RNA and DNA is
that structural communication between two double helical dependent on position in the helix. Brooks and co-workers have
domains is transferred along the DNA backbone over the also shown cooperative and directional folding in the case of
equivalent of about 12−20 backbone units, based on whether preQ1 Riboswitch aptamer domain using an all atom Go̅ model
there is a symmetric or asymmetric distribution of the simulation.581
2800 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

10. ANTICOOPERATIVITY no cooperativity. Employing kinetic analyses and surface


plasmon resonance detection, studies on regulation of tyrosine
There are several examples of anticooperative binding in proteins
hydroxylase by the concentration of tetrahydrobiopterin cofactor
and in materials.621−630 A combined thermodynamic, kinetic,
have shown a presence of negative cooperativity in their
and X-ray crystallographic study on ligand-binding properties of
binding.634 Using NMR, Stevens et al. report a case of strong
recombinant ferric vitreoscilla homodimeric hemoglobin helped
negative cooperativity by delineation of the allosteric mechanism
to establish anticooperative ligand-binding behavior in sol-
of a dimeric cytidylyltransferase enzyme between the first and
ution.621 Anticooperativity was reflected by a 300-fold decrease
second binding of its substrate, CTP.635 A role for prion protein
in the second-order rate constant for cyanide and imidazole in the function of cell protection by scavenging excess copper has
binding to the monoligated ferric vitreoscilla hemoglobin with been suggested by Millhauser and co-workers based on electron
respect to that for ligand association to the ligand free paramagnetic resonance and fluorescence quenching experi-
homodimer in solution. Doig and co-workers studied Glu-Lys ments to determine the copper affinity to an N-terminal
and Lys-Glu salt bridges formed when the residues are spaced i, i octarepeat domain of the protein.636 Decomposition of the
+ 4 in an isolated α-helix in aqueous solution.622 When the side EPR spectra revealed the proportions of all coordination species
chains were spaced i, i + 4, i + 8, forming a Glu-Lys-Glu triplet, throughout the copper concentration range and identified
the second salt bridge provided no additional stabilization to the significant populations of intermediates, consistent with negative
helix. They suggested that the triplet may be less stabilizing than cooperativity. Thus, several salient aspects of cooperativity in
the sum of the two individual interactions, giving rise to biochemistry have been addressed.
anticooperativity.
In a study modeling the antigen−antibody encounter, a
11. MANIFESTATION OF COOPERATIVITY
formula that enables one to calculate the anticooperativity as a
function of the size of the binding site for any values of the 11. 1. Self-Assembly
separation between the two active lobes and of the antigen size
Consequences of cooperativity in supramolecular systems have
was proposed.623 Liang and co-workers introduced a geometric
been exploited for template-directed molecular synthesis,
model for quantifying 3-body interactions in native proteins.624
development of methods to probe weak supramolecular
The study defined a nonadditive coefficient that characterized
interactions besides assessing cooperativity in self-assembly
cooperativity and anticooperativity of residue interactions in
processes.638−689 Meijer and co-workers have studied the role of
native proteins by measuring the deviation of 3-body interactions
cooperativity in supramolecular polymers.640−647 Spectroscopi-
from 3 independent pairwise interactions. While comparing the
cally monitoring the nucleation process in chemical self-assembly
3-body propensity value from what would be expected if only of π-conjugated molecules leading to formation of helical
pairwise interactions were considered, the study revealed how supramolecular fibrillar structures revealed the presence of a
hydrophobic interactions and hydrogen-bonding interactions remarkably high degree of cooperativity.641 The study also
make nonadditive contributions to protein stability. The elucidated a significant role for organized shell of solvent
nonadditive nature was shown to depend on whether such molecules in rigidifying the aggregates and guiding them toward
interactions were located in the protein interior or on the protein further assembly into bundles and gels. In another study, they
surface. In order to explain multibody effects in hydrophobic demonstrated a balancing out of attractive forces based on dipole
association on a molecular level, Czaplewski et al. showed that interactions, π−π stacking, solvophobic effects, with electrostatic
the three-methane potential of mean force (PMF) in water at 298 repulsive interactions, in the process of self-assembly of discotic
K for their system (i.e., the trimer case with isosceles-triangle amphiphiles into helical columnar aggregates of controlled
geometry) is anticooperative at the contact minimum.625 length.642 The presence or absence of cooperativity in the self-
Nonadditivity properties of hydrophobic interactions were assembly mechanism was determined to be a key feature in
shown to be temperature-dependent by Chan and co-work- manipulating the physical properties of synthetic supramolecular
ers.55,626 Kudrev developed a method for the mathematical polymers. Stoddart and co-workers have assessed the impact of
processing of experimental titration curves of a homogeneous multivalency and cooperativity in the context of self-assembling
one-dimensional polynucleotide.627 This method helps in both as models for biological processes and to get a handle on
calculation of the cooperativity and anticooperativity parameters how to order molecules into complex nanoarchitectures.648,649 In
of the interaction of a ligand with an infinite homogeneous a recent study on extended π···π stacking between contiguous
polymer. arene units in the rings of a series of oligorotaxanes, progressively
Negative cooperativity has been reported in studies on enhanced molecular recognition and increased positive cooper-
regulatory enzymes several decades ago.631−637 While elucidat- ativity were shown to be important.649 Studies on m-phenyl-
ing negative cooperativity in the case of CTP synthetase, an eneethynylene oligomers demonstrate how critical chain length
allosteric enzyme, Levitzki et al. gave strong support for the phenomena mark the onset of cooperativity in a macromolecule
sequential model of subunit interactions which postulates that and indeed how the supramolecular polymerization mechanism
ligand-induced conformational changes are responsible for is sensitive to length of the starting oligomer and geometry of the
regulatory and cooperative phenomena in enzymes.631 DeMeyts monomer units.650,651
et al. reported the characterization of negative cooperativity in DFT studies by de Greef and co-workers rationalized
site−site interactions of insulin receptor.632 Crystal structures of experimentally observed cooperative growth of C3-symmetrical
the negatively cooperative aspartate receptor caught at trialkylbenzene-1,3,5-tricarboxamide supramolecular polymers
intermediate stages in the binding process were studied by that self-assemble into ordered 1D structures through hydrogen
Koshland.633 On the basis of frequency of occurrence of bonding.652 Cooperative growth of these structures was shown
negatively cooperative proteins studied, sequential changes in to be due to electrostatic interactions and nonadditive effects
binding patterns were shown to be extensive not only in positive emanating from redistribution of the electron density along the
as well as in negatively cooperative proteins but also in those with aggregate length. Hanke and co-workers report an investigation
2801 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

of the cooperative effects that control the synthesis of a graphene cause substantial changes in the magnitude of cooperative
nanoribbon on the Au(111) surface starting from an anthracene binding as expressed in the large differences in effective molarity.
polymer while considering the associated van der Waals 11. 3. Catalysis
interactions using DFT calculations.653 Cooperativity in self-
assembled complexes of twenty-nine homo and hetero metallic Enzyme catalysis and associated cooperativity has served as an
polynuclear triple-stranded helicates is explained by a combina- inspiration for the design of new chemical reactions that proceed
tion of experimental synthesis and a simple point-charge model with efficiency, high selectivity, and minimal waste.499,503,695
which fine-tunes the balancing between electrostatic repulsion Recent studies reveal cooperative effects with organocatalysts
and solvation energies. 654 A combined theoretical and used in a wide variety of chemical transformations.696−727 The
experimental approach to probe weak supramolecular inter- emphasis in these studies lies in the demonstration of the precise
actions in bisurea solution focuses on evolving a supramolecular control of a cooperatively catalyzed reaction with noncovalent
platform that presents a cooperative transition, so that a interactions alone.
perturbation of such a transition can be monitored by a Schreiner and co-workers presented a mild and efficient
temperature scanning experiment.655 method for the completely regioselective alcoholysis of styrene
Waters and co-workers pursued mutagenesis studies in [2]- oxides.722 This study suggested a hydrogen-bonding mediated
catenanes which mimic numerous key properties of large protein cooperative Brønsted-acid catalysis mechanism for a system
systems.656 Expounding on the role of multivalency and consisting of a thiourea derivative and mandelic acid. In
cooperativity on the structure and stability of these systems subsequent studies, Schreiner and co-workers reported a
where multiple noncovalent interactions create a web of thiourea−Brønsted acid cooperative catalytic system for the
interdependent interactions, they demonstrate how changes to enantioselective cyanosilylation of aldehydes.720 Combined
a component of the web leads to compensating effects in some of experimental and computational investigations showed that the
the linked interactions. In others, the perturbations result in a key catalytic species resulted from the cooperative interaction of
cascade of destabilizing interactions that lead to disproportionate bifunctional thioureas and an achiral acid that formed well-
losses in stability. Ercolani proposed a method to assess defined chiral hydrogen-bonding environments. Jacobsen and
cooperativity in self-assembly processes in the cases of self- co-workers have conducted extensive studies on the cooperative
assembly of helicates and of porphyrin ladders and suggested in mechanisms responsible for stereoinduction in highly enantio-
contrast that positive cooperativity in artificial self-assembling selective reactions such as the Claisen rearrangement, cationic
systems is probably much more rare than it was previously polycyclizations, and Strecker reaction promoted by noncovalent
thought to be.657 Cooperativity and multivalency have also been catalysts.696 They also showed how creating cooperative catalysts
implicated in the case of different kinds of materials like lies in controlling the competition of the two catalysts for the
multipolymers,658 polyethylenimine gels,650 etc. Prohens et al. reactants as well as their mutual interaction through the Povarov
have worked on cooperativity in solid state squaramides where reaction using urea and thiourea derivatives.698 Krautwald et al.
hydrogen-bonded catemers play a crucial role in defining the have reported studies on α-allylation of branched aldehydes
solid-state synthon of disecondary squaramides, overriding their where two independent catalysts working synergistically allow
preferred association mode in solution.660 Zhao and co-workers the controlled synthesis of all four stereoisomers in a reaction
designed aromatically functionalized oligocholate pore-forming that forms carbon carbon bonds.699
materials in which the aromatic interactions and hydrophobic Martiń and co-workers developed a simple synthetic receptor
interactions could work cooperatively.661 that displays an enantioselective positive cooperativity between
intrareceptor interactions and guest binding leading to a
11. 2. Chelate Cooperativity reinforcement of chiral recognition.700 The role of cooperativity
Several studies have emerged on the concept of chelate has also been addressed by several other studies on asymmetric
cooperativity,690−694 which is essentially an increase in binding catalysis. Kanai et al. reported the development of Lewis acid
affinity due to intramolecular interactions. Hunter et al. have Lewis base bifunctional asymmetric catalysts with an emphasis
demonstrated the influence of hydrogen bond strength on on cooperativity specifically for a wide range of enantioselective
chelate cooperativity using the metalloporphyrin-ligand sys- reactions including cyanosilylation of aldehydes, Strecker
tem.690 A systematic survey of the relationship between chemical reaction of aldimines, Reissert reactions of quinolines, isoquino-
structure and cooperativity to measure the free-energy lines, and pyridines, cyanosilylation of ketones, and Strecker
contributions due to intramolecular ether-phenol hydrogen reaction of ketoimines.701 Scheidt and co-workers have also
bonding in different supramolecular architectures using chemical reported how precise orchestration of interactions to activate
double mutant cycles in toluene has been reported. Emphasizing reactants simultaneously or sequentially leads to cooperative
the role of preorganization in porphyrin oligomer complexes, catalysis.702 This concept has been clearly shown through a study
Anderson and co-workers demonstrate how a highly cooperative on cooperative N-heterocyclic carbene/Lewis acid catalytic
two-state model describes stabilities of the studied systems.691 system that promotes the addition of homoenolate equivalents
Denaturation titrations of complexes of cyclic and linear zinc- to hydrazones, generating highly substituted γ-lactams with a
porphyrin hexamer with multidentate ligands reveal that the high degree of diastereo- and enantioselectivity. Further studies
stepwise effective molarities for the third through sixth based on the approach of cooperative catalysis include a report by
intramolecular coordination events with the cyclic hexamer are Patil where merging metal and N-heterocyclic carbene catalysis
extremely high compared to linear porphyrin oligomers. Schalley has been employed as a mode to explore enantioselective organic
and co-workers have recently reported the effects of chelate transformations.703 Hapiot and co-workers have shown the
cooperativity and spacer length on the assembly, thermody- implication of cooperativity in explaining the catalytic perform-
namics, and kinetics of divalent pseudorotaxanes.692 These ances of cyclodextrin-substituted polymers in aqueous rhodium-
studies indicate how very subtle changes in the spacer length catalyzed hydroformylation of 1-hexadecene.704
2802 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Figure 16. Examples of interplay of noncovalent interactions in the field of material science: (a) in self-assembled nanotubes from flexible macrocycles,
ref 751, (b) in solid-state squaramides, ref 660, and in (c) oligophenyleneethynylene (OPE)-based supramolecular polymer, ref 749.

A recent review by Raynal et al. extensively focuses on molecular symmetry, rigidity, and steric hindrance and on the
supramolecular catalysts for which the observed catalytic activity other by the type and concentration of all dielectrically active
and/or selectivity have been imputed to noncovalent interaction species. Both molecular and dielectric architecture were observed
between the reaction partners.719 Sunoj and co-workers have to vary in the course of the chemical reaction, with the overall
explored the mechanistic insights on cooperative catalysis direction in which the cooperativity shifts being governed by the
through computational quantum chemical methods.718 Cooper- interplay between these two phenomena. The role of interplay
ative catalysis addressed in this review cover multicatalytic between CH−π and CF−π interactions in determining the
approaches under one pot reaction conditions, wherein the structure of N-heterocyclic carbene (NHC) palladium com-
complementary attributes of two or more catalysts are made to plexes proven to be effective catalysts in many important
work together. Metal−ligand cooperativity may lower kinetic reactions, especially cross-coupling reactions, was demonstrated
barriers and enable unique reaction chemistry specifically in the using 1H NMR spectroscopy, X-ray crystallography, and DFT
case of cleavage of N−N bonds by soluble transition metal calculations.712 Thus, several studies performed in recent years
complexes. On the basis of this concept, Chirik and co-workers underline the contemporary relevance of cooperativity in the
have prepared early transition metal analogs to exploit metal− design of catalysts whose mode of action is governed by
ligand cooperativity in reactions relevant to nitrogen fixation.705 cooperativity.696−727
Messerle and co-workers have reported studies on homobime- 11. 4. Material Science
tallic dirhodium and diiridium complexes as catalysts for the
intramolecular dihydroalkoxylation of a series of alkyne diol The impact of different noncovalent interactions on materials has
substrates.706 The enhanced activity and selectivity of bimetallic been highlighted through several recent studies.728−736 Some
catalysts as against a single metal catalyst was attributed to examples of interplay of noncovalent interactions in the field of
cooperativity between the metal pairs, and a bimetallic material science are shown in Figure 16. Design of several
intermediate for the reaction pathway is proposed. Motta et al. bioinspired novel materials in the recent years has emphasized
have also reported the distinctive center-to-center cooperative the role played by cooperativity of multiple noncovalent
catalytic properties exhibited by bimetallic constrained geometry interactions especially that of the hydrogen bond.737,738 Welland
catalysts (CGCs) and analyzed the impact of metal−metal and co-workers have shown a scalable self-assembly approach to
proximity effects on ethylene polymerization processes using making free-standing films from amyloid protein fibrils, wherein
DFT calculations.707 Mankad and co-workers showed that the films were well-ordered and highly rigid, with a Young’s
catalytic C−H borylation, a transformation that previously modulus which is comparable to the highest values for
required noble metal catalysts, could be achieved with base metal proteinaceous materials found in nature.739 Buehler has
catalysts by using a mechanistic paradigm that depends on demonstrated how understanding self-assembly of amyloid
intimate cooperativity between two closely associated yet distinct proteins can be exploited to make well-ordered and strong
base metal sites.708,709 functional macroscopic materials.740 The presence of hydrogen
Besides, studies on a supramolecular system that is based on bonds in β sheet structures under nanoconfinement shows how a
cooperative self-assembled helicity, exhibiting strong chiral weakness can be turned into a strength. On the basis of a study of
amplification and having racemization and deracemization the shearing deformation of hydrogen bonds within β-strands
processes in thermodynamic equilibrium have also been and β-sheet nanocrystals applying an elastic structural model, the
reported.710 Fitz et al. have focused on the intermolecular number of hydrogen bonds that deform cooperatively was
cooperativity of reactive polymers considering epoxy amine identified. The studies of Buehler and co-workers740−745 further
model systems for their case study.711 The molecular level emphasize how the intrinsic strength limitation of hydrogen
characteristics that govern the intermolecular cooperativity of bonds can be overcome by the formation of a nanocomposite
these reactive systems were determined on one hand by structure of hydrogen bond clusters, thereby enabling the
2803 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

formation of larger and much stronger beta-sheet structures.741 H···X hydrogen-bonding interactions was demonstrated. A novel
Besides the nanoscale organization of hydrogen bonds was method for fabricating tough hydrogels that are totally physically
shown to be critical for the mechanical behavior of protein cross-linked by cooperative hydrogen bonding between a pre-
materials such as silk.742 More recently, they have also analyzed existing polymer and an in situ polymerized polymer was recently
the size of region in which hydrogen bond deformation in reported by Fang and co-workers.750 Comparative synthesis
functional groups attached to carbon nanotubes (CNT) is experiments, DSC characterization, and molecular modeling
cooperative, focusing specifically on CNTs functionalized with indicated that the formation of strong cooperative hydrogen
carboxyl groups as a model system.743 These studies reveal that bonding between the pre-existing poly(N-vinylpyrrolidone) and
the effect of structural organization of functional groups is the in situ formed P-aqueous acrylamide chains contributes to
significant not only in very small diameter CNTs but also in the gel formation and the toughening of the hydrogels.
larger diameter CNTs which are most commonly used for Martiń and co-workers have reported crystal structures of self-
engineering applications. Larger diameter CNTs show a larger assembled nanotubes through stacking of macrocycles.751 A
cooperative deformation range, providing a good basis for further combination of assembly driven by the propensity of the
improvement by adding structured functional groups. Tuning macrocycles to create nearly flat structures displaying a void
the water content in graphene oxide paper nanocomposites was space within them and the cooperativity of weak directional
shown to allow for the translation of the nanoscale hydrogen interactions such as dipole−dipole interactions and CH···O
bond network into macroscale mechanical properties.744 For hydrogen bonds and nondirectional interactions such as van der
graphene oxide paper at an optimal concentration of water, the Waals contacts resulted in yielding unexpected self-assembled
degree of cooperative hydrogen bonding within the network nanotubes. Tokmachev et al. conducted an analysis of all
comprising adjacent nanosheets and water molecules was found hydrogen bond networks for finite elements of ice nanotubes
to enhance the modulus of the paper. Hu et al. have employed a formed by up to 32 water molecules.752 The hydrogen bonds in
scalable strategy to produce biomimic continuous fibers with the growing ice nanotubes were classified according to their
brick and mortar structures of alternating graphene sheets and structural positions and significant dependencies on the
hyperbranched polyglycerol (HPG) binders via wet-spinning cooperativity energy, and bond lengths on the system’s
assembly technology.746 The resulting macroscopic supra- morphology were revealed. Ju and co-workers recently reported
molecular fibers exhibit excellent mechanical properties the binding affinity and thermodynamic parameters between a
comparable to bone or nacre. The study also provides evidence surfactant and carbon nanotube using flavin mononucleotide-
of the role played by strong cooperative hydrogen bonding wrapped carbon nanotubes as model systems.753 The binding
among adjacent HPGs upon different thickness of HPG- affinity of metallic tubes was weaker than that of semiconducting
enveloped graphene sheets. A strategy based on the idea that tubes, and equilibrium constant patterns from semiconducting
incorporation of a perfluorophenyl group as the end-capped π- tubes showed a preference to certain SWNT chiralities and
conjugated part, and a linking amino acid of phenylalanine with a surfactant-specific cooperativity according to nanotube chirality.
phenyl moiety in the side chain, forms a new intramolecular Jin et al. have reported the self-assembly of a gemini-shaped,
building block in a peptide hydrogelator considering the efficient chiral amphiphilic hexa-perihexabenzocoronene having two
formation of supramolecular nanofibers and hydrogels was chiral oxyalkylene side chains, along with two lipophilic side
recently established.747 chains, yielding graphitic nanotubes with one-handed helical
Yam and co-workers have employed a series of luminescent chirality.754 A high level of chirality amplification observed
metallogels of PtII terpyridyl complexes showing gelation therein indicated a long-range cooperativity in the self-
properties driven by an interplay between Pt···Pt and π−π assembling process.
interactions in addition to hydrophobic−hydrophobic inter- The role of dipolar interactions and hydrogen-bonding
actions.396 Schneider and co-workers studied the impact of interactions in fine-tuning the linear and first nonlinear optical
cooperativity and selectivity in chemomechanical polyethyleni- (NLO) responses in molecular aggregates were investigated by
mine gels.659 The gel particles obtained from polyethylenimine Pati and co-workers.755 The nonadditivity of optical properties in
(PEI) cross-linked with 10% ethylene glycol diglycidyl ether their systems were shown to arise specifically due to cooperativity
demonstrated a considerable selectivity in size changes on of weak interactions. Calculations performed on model systems
exposure to solutions of different effector compounds. A striking starting from dimers to multiple aggregates provide insights into
cooperativity was noted if aromatic compounds such as the interaction mechanisms and strategies to enhance the first
naphthoic acid were used simultaneously with amino acids as hyperpolarizabilities of π-conjugated molecular assemblies.
effectors. Miravet and co-workers highlighted the key role played Overney and co-workers have studied a second-order nonlinear
by solubility in influencing the gelation process.748 A model to optical material of a class of self-assembling molecular glasses
relate thermal stability and minimum gelation concentration involving quadrupolar phenyl-perfluorophenyl (Ph-PhF) inter-
values of small-molecule gelation in terms of the solubility and actions by analyzing its molecular relaxation phenomena and
cooperative self-assembly of gelator building blocks was phase behavior using a nanoscale methodology based on
developed. scanning probe microscopy.756 The mesoscale dynamics and
Fernández and co-workers have recently reported an cooperativity involved in relaxation processes were captured and
oligophenyleneethynylene (OPE)-based amphiphilic PtII com- quantified in terms of dynamic entropy and enthalpy. The
plex that forms supramolecular polymeric structures in aqueous importance of entropic cooperativity in the design of novel
and polarmedia driven by π−π and different weak C−H···X (X = materials that entail a multitude of interactive constraints form
Cl, O) interactions involving chlorine atoms attached to the PtII dipole field interactions to local dendritic bonding have also been
centers as well as oxygen atoms and polarized methylene groups studied by analyzing a wide-range of dendritic organic NLO self-
belonging to the peripheral glycol chains.749 By employing a assembly molecular glasses.757 Suponitsky et al. have explored
collection of experimental techniques, self-assembly and hydro- the effect of π···π stacking aggregation on hyperpolarizability in
gelation triggered by cooperative π−π and unconventional C− the context of supramolecular design of NLO materials.758 They
2804 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

showed cases of positive cooperativity and more frequently cases 12. PROSPECTS AND OUTLOOK
of negative cooperativity. A decomposition approach was
The current review on cooperativity in noncovalent interactions
employed to analyze the reasons of deviation of aggregate
elucidates the broad range of fields as diverse as structural
molecular hyperpolarizability from additivity as well as the
biology, supramolecular chemistry, and material science, etc. to
cooperative effect of intermolecular interactions on hyper-
be influenced by the subtle modulation and amplification of
polarizability for stacks of growing size. Mafé and co-workers
different noncovalent interactions. Indeed, the search for
proposed a nanoscale switch, giving a nonlinear function with
quantitative aspects of cooperative effects have added a new
two conductive states separated by a sharp transition region on
level of complexity and thus reactivated the well-developed field
the basis of an array of molecular dipoles.759 Employing
of noncovalent interactions. We have attempted to show the
theoretical calculations, they revealed that the local interactions
topological and structure-oriented factors considered in the
between dipoles result in cooperative phenomena that can
study of noncovalent interactions and also in analysis of
significantly improve the switching characteristics.
Pettersson and co-workers examined the balance of surface cooperativity as a function of interaction. Cooperativity and
bonding and hydrogen bonding in the mixed OH + H2O anticooperativity in life sciences in general and in biochemistry in
overlayer on Pt(111), Cu(111), and Cu(110) using DFT particular is an old concept. However, there are several new
calculations.144 A cooperativity effect between surface bonding aspects of biology unfolding such as DNA origami, temporal
and hydrogen bonding that underlies the stability of the mixed cooperativity, etc., and all of them have a role in which
phase at metal surfaces was noted. Water molecules become noncovalent interactions influence each other. A number of
more strongly bonded to the surface upon hydrogen bonding to instances in catalysis and also materials where structure,
OH, while the OH surface bonding is instead weakened through property, and function are critically dependent on the
hydrogen bonding to water. Jones and co-workers have explored cooperativity of noncovalent interactions have emerged.
mechanisms to tune cooperativity between aminoalkylsilanes The contribution of theoretical chemistry and electronic
and silanols on a silica surface by varying the length of the alkyl structure theory are phenomenal in understanding the molecular
linker attaching the amine to the silica surface from methyl to structure and function. Linking the molecules with the clusters of
pentyl.760 The linker length was shown to strongly affect the various length scales all the way to bulk properties warrants a
catalytic cooperativity of amines and silanols in aldol detailed qualitative and quantitative understanding on the way in
condensations as well as the adsorptive cooperativity for CO2 which the noncovalent interactions manifest themselves. Thus,
capture. Bouteiller et al. studied the role played by high in this direction, “How a pair of noncovalent interactions
cooperativity in the thickness transition of a rigid supramolecular mutually influence themselves?” remains a question of prime
polymer made from low molecular weight bisurea.761 Self- significance. Such an understanding is necessary to develop
assembly by hydrogen bonding into two distinct high molecular accurate methods to model macromolecular structures and
weight structures revealed the tunability of this transition. Jain supramolecular assemblies, which are prevalent in both material
and co-workers have very recently found evidence of atomic level science and biology. Recent years have seen rapid experimental
cooperativity in an inorganic material.762 A thousand-atom advances in the design of supramolecular assemblies, smart
nanocrystal of the inorganic solid cadmium selenide exhibits materials, catalysts, and biomolecules where noncovalent
strong positive cooperativity in its reaction with copper ions. A interactions play a prominent role. Experiment and theory
nanocrystal doped with a few copper impurities becomes highly have been the two important paradigms since the inception of
prone to be doped even further, as manifested by a strongly science. Upon the spectacular emergence of computer hardware
sigmoidal response in optical spectroscopy and electron and software, equations based on theory, which are prohibitive to
diffraction measurements. solve manually, have given rise to computational sciences or
Wu and co-workers investigated the adsorption and surface modeling, and this has become the third paradigm in science. In
forces due to multivalent polymers based on coarse-grained the last couple of decades, data intensive research based on
polymer models.658 Polymers with multiple binding segments informatics has gained outstanding significance and emerged as
exhibit strong cooperativity by providing higher levels of surface the fourth paradigm of science.773−775 Thus, in contemporary
protection than those with only one binding site. Negative science, it has become indispensable to integrate the conven-
cooperativity was observed for an architecture with weakly tional paradigms of “experiment” and “theory” with the modern
attractive anchoring segments. Gröhn and co-workers have very paradigms of “modeling” and “informatics” to solve problems in
recently reported the formation of a new type of organic− chemistry along with its interdisciplinary subjects. Rigorous
inorganic hybrid particle with tunability of particle size, through theoretical understanding is needed also to develop more robust
self-assembly driven by the interplay of three noncovalent multiscale modeling approaches. It appears that theoretical
interactions namely ionic, π−π, and Hamaker interactions.763 understanding of how various noncovalent interactions manifest
Hamaker interactions are van-der-Waals interactions on the themselves in supramolecular assemblies is rather vague. While
colloidal scale. This is a short-range force that in a diluted system methods based on molecular dynamics give an excellent
can only come into play once particles are brought close to each dynamics picture it needs to be backed by highly reliable and
other through other forces. Gibbs-Davis and co-workers have accurate first-principles computations. Cooperativity as a
studied the influence of gap length on cooperativity and rate of concept is extensively employed across various disciplines and
association in DNA-modified gold nanoparticle aggregates.764 in slightly varying contexts as has been explained in the review.
The flexibility in aggregates decreases the rate of aggregation as There is an underlying need to develop vigorous procedures to
well as the extent of cooperativity, suggesting implications in quantitatively determine the extent of cooperativity. However,
genomic DNA detection. Thus, a substantial number of studies quantitative estimation of coopeativity effect is challenging both
including the above-mentioned specific examples clearly computationally and experimentally as the size of the system
illustrate the relevance and importance of cooperativity among increases. First priniciples calcualtions have limitations when the
noncovalent interactions in wide-ranging novel materials.765−772 size of the systems become more than a few hundred molecules,
2805 DOI: 10.1021/cr500344e
Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

and the empirical and other approximate methods are rather REFERENCES
unreliable. The qualitative models developed so far have (1) Pauling, L. The Nature of the Chemical Bond and the Structure of
limitations to estimate the effect of cooperativity in multimeric Molecules and Crystals: An Introduction to Modern Structural Chemistry;
systems. Similarly, the terms anticooperativity and negative Cornell University Press: New York, 1960.
cooperativity need to be used cautiously. We believe more (2) van der Waals, J. D. Over de Continuı̈teit van den Gas- en
rigorous models need to be evolved and recommend the use of Vloeistoftoestand (On the continuity of the gaseous and liquid state).
anticooperativity in place of negative cooperativity in literature. It Ph.D. Dissertation, University of Leiden, 1873.
is important to develop both qualitative and quantitative models (3) Kollman, P. A. Noncovalent Interactions. Acc. Chem. Res. 1977, 10,
365−371.
of cooperativity and anticoopertivity at the molecular level. As it (4) Müller-Dethlefs, K.; Hobza, P. Noncovalent Interactions: A
is apparent now, the number of empirical evidence emerging in Challenge for Experiment and Theory. Chem. Rev. 2000, 100, 143−167.
the areas of chemistry, material science, and biology based on (5) Riley, K. E.; Pitoňaḱ , M.; Jurečka, P.; Hobza, P. Stabilization and
cooperativity is increasing at a rapid pace. This vast and varied Structure Calculations for Noncovalent Interactions in Extended
information coming out of a number of observations needs to be Molecular Systems Based on Wave Function and Density Functional
put together and informatics based approaches need to be Theories. Chem. Rev. 2010, 110, 5023−5063.
employed in understanding and prediction of new materials. (6) Mati, I. K.; Cockroft, S. L. Molecular Balances for Quantifying
Noncovalent Interactions. Chem. Soc. Rev. 2010, 39, 4195−4205.
(7) Schalley, C. Analytical Methods in Supramolecular Chemistry; Wiley-
AUTHOR INFORMATION VCH: Weinheim, 2007; p 1.
Corresponding Author (8) Atwood, J. L.; Barbour, L. J.; Jerga, A.; Schottel, B. L. Guest
Transport in a Nonporous Organic Solid via Dynamic van der Waals
*E-mail: gnsastry@gmail.com. Cooperativity. Science 2002, 298, 1000−1002.
Notes (9) Wagner, J. P.; Schreiner, P. R. London Dispersion in Molecular
Chemistry-Reconsidering Steric Effects. Angew. Chem., Int. Ed. 2015, 54,
The authors declare no competing financial interest. 12274−12296.
(10) Frank, H. S.; Wen, W. Y. Structural Aspects of Ion-Solvent
Biographies Interaction in Aqueous Solutions: A Suggested Picture of Water
A. Subha Mahadevi has obtained her Bachelor’s degree from Nizam Structure. Discuss. Faraday Soc. 1957, 24, 133−140.
(11) Huyskens, P. L. Factors Governing the Influence of a First
College, Osmania University, and her Masters degree in Biotechnology Hydrogen Bond on the Formation of a Second One by the Same
from University of Hyderabad in 1998. After a gap of 9 years, she Molecule or Ion. J. Am. Chem. Soc. 1977, 99, 2578−2582.
pursued the advanced course in bioinformatics at CSIR-IICT, (12) Elrod, M. J.; Saykally, R. J. Many-Body Effects in Intermolecular
Hyderabad, which triggered her interest in the field of computational Forces. Chem. Rev. 1994, 94, 1975−1999.
biology. She joined the Molecular Modeling Group at IICT in 2008 and (13) Prins, L. J.; Reinhoudt, D. N.; Timmerman, P. Noncovalent
has been awarded the DST Woman Scientist Fellowship in 2010. She Synthesis Using Hydrogen Bonding. Angew. Chem., Int. Ed. 2001, 40,
has obtained her Ph.D in Bioinformatics under the guidance of Dr G. N. 2382−2426.
(14) Mahadevi, A. S.; Sastry, G. N. Cation-π Interaction: Its Role and
Sastry. Her research interests include understanding the role of
Relevance in Chemistry, Biology, and Material Science. Chem. Rev.
noncovalent interactions such as cation-π, stacking and hydrogen 2013, 113, 2100−2138.
bonding in molecular clusters, supramolecular assemblies, and their (15) Ma, J. C.; Dougherty, D. A. The Cation-π Interaction. Chem. Rev.
cooperativity. 1997, 97, 1303−1324.
G. Narahari Sastry has obtained his early education in Andhra Pradesh, (16) Reddy, A. S.; Sastry, G. N. Cation [M = H+, Li+, Na+, K+, Ca2+,
Mg2+, NH4+, and NMe4+] Interactions with the Aromatic Motifs of
Bachelor’s and Master’s degrees from Osmania University before
Naturally Occurring Amino Acids: A Theoretical Study. J. Phys. Chem. A
obtaining his Ph.D. from the University of Hyderabad under the 2005, 109, 8893−8903.
supervision of Professor E. D. Jemmis. After a couple of post doctoral (17) Reddy, A. S.; Sastry, G. M.; Sastry, G. N. Cation-Aromatic
stints with Professor Sason Shaik and Professor Thomas Bally, he started Database. Proteins: Struct., Funct., Genet. 2007, 67, 1179−1184.
his independent career in 1997 from Pondicherry University. In 2002, he (18) Kim, K. S.; Tarakeshwar, P.; Lee, J. Y. Molecular Clusters of π-
moved to the CSIR-IICT, Hyderabad, to head the Molecular Modeling Systems: Theoretical Studies of Structures, Spectra, and Origin of
Group. He has wide ranging research interests in computational Interaction Energies. Chem. Rev. 2000, 100, 4145−4185.
chemistry and computational biology. The focus of the group currently (19) Staffilani, M.; Hancock, K. S. B.; Steed, J. W.; Holman, K. T.;
Atwood, J. L.; Juneja, R. K.; Burkhalter, R. S. Anion Binding within the
is on understanding the way in which nonbonding interactions operate
Cavity of π-Metalated Calixarenes. J. Am. Chem. Soc. 1997, 119, 6324−
in chemical and biological systems with special attention to cation−π 6335.
and π−π interactions. He was awarded the Shanti Swarup Bhatnagar (20) Fairchild, R. M.; Holman, K. T. Selective Anion Encapsulation by
Award in Chemical Sciences for the year 2011. He has published over a Metalated Cryptophane with a π-Acidic Interior. J. Am. Chem. Soc.
255 scientific papers and is a fellow of the Indian Academy of Sciences, 2005, 127, 16364−16365.
Andhra Pradesh Academy of Sciences, and also the National Academy of (21) Schottel, B. L.; Chifotides, H. T.; Dunbar, K. R. Anion-π
Sciences, India. He is currently a J.C. Bose National Fellow. interactions. Chem. Soc. Rev. 2008, 37, 68−83.
(22) Frontera, A.; Quiñonero, D.; Deyà, P. M. Cation−π and anion−π
interactions. WIREs Comput. Mol. Sci. 2011, 1, 440−459.
ACKNOWLEDGMENTS (23) Frontera, A.; Gamez, P.; Mascal, M.; Mooibroek, T. J.; Reedijk, J.
Putting Anion−π Interactions into Perspective. Angew. Chem., Int. Ed.
We acknowledge the liberal financial support from CSIR 12th
2011, 50, 9564−9583.
Five Year Plan, GENESIS, and INTELCOAT as well as from (24) Li, Q.; Li, R.; Liu, Z.; Li, W.; Cheng, J. Interplay between Halogen
DST, DBT, and DAE-BRNS. We also thank many of our co- Bond and Lithium Bond in MCN-LiCN-XCCH (M = H, Li, and Na; X =
workers who participated in the studies related to noncovalent Cl, Br, and I) Complex: The Enhancement of Halogen Bond by a
interactions over many years. Lithium Bond. J. Comput. Chem. 2011, 32, 3296−3303.

2806 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(25) Lu, Y.; Liu, Y.; Li, H.; Zhu, X.; Liu, H.; Zhu, W. Energetic Effects (48) Uchimaru, T.; Korchowiec, J.; Tsuzuki, S.; Matsumura, K.;
between Halogen Bonds and Anion-π or Lone Pair-π Interactions: A Kawahara, S. Importance of Secondary Electrostatic Interactions in
Theoretical Study. J. Phys. Chem. A 2012, 116, 2591−2597. Hydrogen-Bonding Complexes: An Investigation Using the Self-
(26) Hunter, C. A.; Anderson, H. L. What is Cooperativity? Angew. Consistent Charge and Configuration Method for Subsystems. Chem.
Chem., Int. Ed. 2009, 48, 7488−7499. Phys. Lett. 2000, 318, 203−209.
(27) Hughes, A. D.; Anslyn, E. V. A Cationic Host Displaying Positive (49) Kawahara, S.; Taira, K.; Uchimaru, T. Hydrogen Bond
Cooperativity in Water. Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 6538− Cooperativity Derived from Neighboring Hydrogen Bond Formation:
6543. Case Study in Three Iso-Complexes of C8H9N5O2. Chem. Phys. 2003,
(28) Williams, D. H.; Westwell, M. S. Aspects of Weak Interactions. 290, 79−83.
Chem. Soc. Rev. 1998, 27, 57−63. (50) Jeong, K.-S.; Tjivikua, T.; Rebek, J., Jr. Relative Hydrogen-
(29) Williams, D. H.; Bardsley, B. Estimating Binding Constants − The Bonding Affinities of Imides and Lactams. J. Am. Chem. Soc. 1990, 112,
Hydrophobic Effect and Cooperativity. Perspect. Drug Discovery Des. 3215−3217.
1999, 17, 43−59. (51) Clark, S. L.; Hammond, P. T. The Role of Secondary Interactions
(30) Calderone, C. T.; Williams, D. H. An Enthalpic Component in in Selective Electrostatic Multilayer Deposition. Langmuir 2000, 16,
Cooperativity: The Relationship between Enthalpy, Entropy, and 10206−10214.
Noncovalent Structure in Weak Associations. J. Am. Chem. Soc. 2001, (52) Searle, M. S.; Williams, D. H. The Cost of Conformational Order:
123, 6262−6267. Entropy Changes in Molecular Associations. J. Am. Chem. Soc. 1992,
(31) Bissantz, C.; Kuhn, B.; Stahl, M. A Medicinal Chemist’s Guide to 114, 10690−10697.
Molecular Interactions. J. Med. Chem. 2010, 53, 5061−5084. (53) Searle, M. S.; Sharman, G. J.; Groves, P.; Benhamu, B.;
(32) Rodriguez-Docampo, Z.; Pascu, S. I.; Kubik, S.; Otto, S. Beauregard, D. A.; Westwell, M. S.; Dancer, R. J.; Maguire, A. J.; Try,
Noncovalent Interactions within a Synthetic Receptor Can Reinforce A. C.; Williams, D. H. Enthalpic (Electrostatic) Contribution to the
Guest Binding. J. Am. Chem. Soc. 2006, 128, 11206−11210. Chelate Effect: A Correlation between Ligand Binding Constant and a
(33) Goyon, J.; Colin, A.; Ovarlez, G.; Ajdari, A.; Bocquet, L. Spatial Specific Hydrogen Bond Strength in Complexes of Glycopeptide
Cooperativity in Soft Glassy Flows. Nature 2008, 454, 84−87. Antibiotics with Cell Wall Analogues. J. Chem. Soc., Perkin Trans. 1 1996,
(34) Boyer, P. D. Energy, Life, and ATP. Angew. Chem., Int. Ed. 1998, 1, 2781−2786.
37, 2296−2307. (54) Murphy, K. P.; Xie, D.; Thompson, K. S.; Amzel, L. M.; Freire, E.
(35) Zhong, Z.; Li, X.; Zhao, Y. Enhancing Binding Affinity by the Entropy in Biological Binding Processes: Estimation of Translational
Cooperativity between Host Conformation and Host-Guest Inter- Entropy Loss. Proteins: Struct., Funct., Genet. 1994, 18, 63−67.
actions. J. Am. Chem. Soc. 2011, 133, 8862−8865. (55) Moghaddam, M. S.; Shimizu, S.; Chan, H. S. Temperature
(36) Carrillo, R.; Morales, E. Q.; Martín, V. S.; Martín, T. A Novel Dependence of Three-Body Hydrophobic Interactions: Potential of
Approach for the Evaluation of Positive Cooperative Guest Binding: Mean Force, Enthalpy, Entropy, Heat Capacity, and Nonadditivity. J.
Kinetic Consequences of Structural Tightening. Chem. - Eur. J. 2013, 19, Am. Chem. Soc. 2005, 127, 303−316.
7042−7048. (56) Chandler, D. Interfaces and the Driving Force of Hydrophobic
(37) Hasenknopf, B.; Lehn, J. M.; Kneisel, B. O.; Baum, G.; Fenske, D. Assembly. Nature 2005, 437, 640−647.
(57) Dill, K. A. Dominant Forces in Protein Folding. Biochemistry
Self-Assembly of a Circular Double Helicate. Angew. Chem., Int. Ed. Engl.
1990, 29, 7133−7155.
1996, 35, 1838−1840.
(58) Ludwig, R. Water: From Clusters to the Bulk. Angew. Chem., Int.
(38) Lehn, J. M. Toward Complex Matter: Supramolecular Chemistry
Ed. 2001, 40, 1808−1827.
and Self-organization. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 4763−
(59) Steiner, T. The Hydrogen Bond in the Solid State. Angew. Chem.,
4768.
Int. Ed. 2002, 41, 48−76.
(39) Hussain, M. A.; Soujanya, Y.; Sastry, G. N. Evaluating the Efficacy
(60) Grabowski, S. J. What Is the Covalency of Hydrogen Bonding?
of Amino Acids as CO2 Capturing Agents: A First Principles Chem. Rev. 2011, 111, 2597−2625.
Investigation. Environ. Sci. Technol. 2011, 45, 8582−8588. (61) Ball, P. Water as an Active Constituent in Cell Biology. Chem. Rev.
(40) Higashibayashi, S.; Onogi, S.; Srivastava, H. K.; Sastry, G. N.; Wu, 2008, 108, 74−108.
Y.-T.; Sakurai, H. Stereoelectronic Effect of Curved Aromatic (62) Grabowski, S. J. Hydrogen Bonding. New Insights; Springer:
Structures: Favoring the Unexpected Endo Conformation of Benzylic- Heidelberg, 2006.
Substituted Sumanene. Angew. Chem., Int. Ed. 2013, 52, 7314−7316. (63) Hobza, P.; Havlas, Z. Blue-Shifting Hydrogen Bonds. Chem. Rev.
(41) Umadevi, D.; Sastry, G. N. Molecular and Ionic Interaction with 2000, 100, 4253−4264.
Graphene Nanoflakes: A Computational Investigation of CO2, H2O, Li, (64) Mahadevi, A. S.; Sastry, G. N. Computational Approaches
Mg, Li+, and Mg2+ Interaction with Polycyclic Aromatic Hydrocarbons. Towards Modeling Finite Molecular Assemblies: Role of cation-π, π-π
J. Phys. Chem. C 2011, 115, 9656−9667. and hydrogen bonding interactions. In Practical Aspects of Computational
(42) Schuster, P.; Wolschann, P. Hydrogen Bonding: From Small Chemistry I: An Overview of the Last Two Decades and Current Trends;
Clusters to Biopolymers. Monatsh. Chem. 1999, 130, 947−960. Leszczynski, J., Shukla, M. K., de Rode, H., Eds.; Springer: The
(43) Cruzan, J. D.; Braly, L. B.; Liu, K.; Brown, M. G.; Loeser, J. G.; Netherlands, 2012; p 517.
Saykally, R. J. Quantifying Hydrogen Bond Cooperativity in Water: (65) Turi, L.; Dannenberg, J. J. Molecular Orbital Studies of Crystal
VRT Spectroscopy of the Water Tetramer. Science 1996, 271, 59−62. Formation: The Aggregation and Nucleation of 1, 3-Diones. J. Phys.
(44) Stone, A. J. The Theory of Intermolecular Forces, 2 ed.; Oxford Chem. 1992, 96, 5819−5824.
University Press, 2013. (66) Turi, L.; Dannenberg, J. J. Molecular Orbital Study of Acetic Acid
(45) Jorgensen, W. L.; Pranata, J. Importance of Secondary Aggregation. 1. Monomers and Dimers. J. Phys. Chem. 1993, 97, 12197−
Interactions in Triply Hydrogen Bonded Complexes: Guanine-Cytosine 12204.
vs Uracil-2,6-Diaminopyridine. J. Am. Chem. Soc. 1990, 112, 2008− (67) Dannenberg, J. J.; Rios, R. Theoretical Study of the Enolic Forms
2010. of Acetylacetone. How Strong Is the H-Bond? J. Phys. Chem. 1994, 98,
(46) Jorgensen, W. L.; Severance, D. L. Chemical Chameleons: 6714−6718.
Hydrogen Bonding with Imides and Lactams in Chloroform. J. Am. (68) Morokuma, K.; Pedersen, L. Molecular-Orbital Studies of
Chem. Soc. 1991, 113, 209−216. Hydrogen Bonds. An Ab Initio Calculation for Dimeric H20. J. Chem.
(47) Pranata, J.; Wierschke, S. G.; Jorgensen, W. L. OPLS Potential Phys. 1968, 48, 3275−3282.
Functions for Nucleotide Bases. Relative Association Constants of (69) Kollman, P. A.; Allen, L. C. Theory of the Hydrogen Bond:
Hydrogen Bonded Base Pairs in Chloroform. J. Am. Chem. Soc. 1991, Electronic Structure and Properties of the Water Dimer. J. Chem. Phys.
113, 2810−2819. 1969, 51, 3286−3293.

2807 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(70) Hankins, D.; Moskowitz, J. W.; Stillinger, F. H. Water Molecule (92) Luck, W. A. P. The Importance of Cooperativity for the Properties
Interactions. J. Chem. Phys. 1970, 53, 4544−4554. of Liquid Water. J. Mol. Struct. 1998, 448, 131−142.
(71) Keutsch, F. N.; Cruzan, J. D.; Saykally, R. J. The Water Trimer. (93) Keutsch, F. N.; Saykally, R. J. Water Clusters: Untangling the
Chem. Rev. 2003, 103, 2533−2577. Mysteries of the Liquid, One Molecule at a Time. Proc. Natl. Acad. Sci. U.
(72) Buck, U.; Ettischer, I.; Melzer, M.; Buch, V.; Sadlej, J. Structure S. A. 2001, 98, 10533−10540.
and Spectra of Three-Dimensional (H2O)n Clusters, n = 8, 9, 10. Phys. (94) Agmon, N. Liquid Water: From Symmetry Distortions to
Rev. Lett. 1998, 80, 2578−2581. Diffusive Motion. Acc. Chem. Res. 2012, 45, 63−73.
(73) Wernet, Ph.; Nordlund, D.; Bergmann, U.; Cavalleri, M.; Odelius, (95) Saha, S.; Sastry, G. N. Quantifying Cooperativity in Water
M.; Ogasawara, H.; Näslund, L. Å.; Hirsch, T. K.; Ojamäe, L.; Glatzel, P.; Clusters: An Attempt towards Obtaining a Generalized Equation. Mol.
Pettersson, L. G. M.; Nilsson, A. The Structure of the First Coordination Phys. 2015, 113, 3031−3041.
Shell in Liquid Water. Science 2004, 304, 995−999. (96) Hannachi, Y.; Silvi, B.; Bouteiller, Y. Ab initio Study of the
(74) Head-Gordon, T.; Johnson, M. E. Tetrahedral Structure or Structure, Cooperativity, and Vibrational Properties of the H2O: (HF)2
Chains for Liquid Water. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 7973− Hydrogen Bonded Complex. J. Chem. Phys. 1992, 97, 1911−1918.
7977. (97) González, L.; Mó, O.; Yáñez, M. High Level Ab Initio and Density
(75) Pérez, C.; Zaleski, D. P.; Seifert, N. A.; Temelso, B.; Shields, G. C.; Functional Theory Studies on Methanol-Water Dimers and Cyclic
Kisiel, Z.; Pate, B. H. Hydrogen Bond Cooperativity and the Three- Methanol(water)2 Trimer. J. Chem. Phys. 1998, 109, 139−150.
Dimensional Structures of Water Nonamers and Decamers. Angew. (98) González, L.; Mó, O.; Yáñez, M. Density Functional Theory
Chem., Int. Ed. 2014, 53, 14368−14372. Study on Ethanol Dimers and Cyclic Ethanol Trimers. J. Chem. Phys.
(76) Schmidt, D. A.; Miki, K. Structural Correlations in Liquid Water: 1999, 111, 3855−3861.
A New Interpretation of IR Spectroscopy. J. Phys. Chem. A 2007, 111, (99) Nagaraju, M.; Sastry, G. N. Comparative Study on Formamide-
10119−10122. Water Complex. Int. J. Quantum Chem. 2010, 110, 1994−2003.
(77) Maes, G.; Smets, J. Hydrogen Bond Cooperativity. A Quantitative (100) Nagaraju, M.; Sastry, G. N. Effect of Alkyl Substitution on H-
Study Using Matrix-Isolation m-IR Spectroscopy. J. Phys. Chem. 1993, bond Strength of Substituted Amide-Alcohol Complexes. J. Mol. Model.
97, 1818−1825. 2011, 17, 1801−1816.
(78) Tse, Y.-C.; Newton, M. D. Theoretical Observations on the (101) Masunov, A.; Dannenberg, J. J. Theoretical Study of Urea and
Structural Consequences of Cooperativity in H–O Hydrogen Bonding. Thiourea. 2. Chains and Ribbons. J. Phys. Chem. B 2000, 104, 806−810.
J. Am. Chem. Soc. 1977, 99, 611−613. (102) Kobko, N.; Paraskevas, L.; del Rio, E.; Dannenberg, J. J.
(79) Koehler, J. E. H.; Saenger, W.; Lesyng, B. Cooperative Effects in Cooperativity in Amide Hydrogen Bonding Chains: Implications for
Extended Hydrogen Bonded Systems Involving O-H Groups. Ab Initio Protein-Folding Models. J. Am. Chem. Soc. 2001, 123, 4348−4349.
Studies of the Cyclic S4 Water Tetramer. J. Comput. Chem. 1987, 8, (103) Dannenberg, J. J. Cooperativity in hydrogen bonded aggregates.
1090−1098. Models for crystals and peptides. J. Mol. Struct. 2002, 615, 219−226.
(80) Suhai, S. Cooperative effects in hydrogen bonding: Fourth-order (104) Kobko, N.; Dannenberg, J. J. Cooperativity in Amide Hydrogen
many-body perturbation theory studies of water oligomers and of an Bonding Chains. A Comparison between Vibrational Coupling through
infinite water chain as a model for ice. J. Chem. Phys. 1994, 101, 9766− Hydrogen Bonds and Covalent Bonds. Implications for Peptide
9782. Vibrational Spectra. J. Phys. Chem. A 2003, 107, 6688−6697.
(81) Xantheas, S. S. Cooperativity and hydrogen bonding network in (105) Kobko, N.; Dannenberg, J. J. Cooperativity in Amide Hydrogen
water clusters. Chem. Phys. 2000, 258, 225−231.
Bonding Chains. Relation between Energy, Position, and H-Bond Chain
(82) Mó, O.; Yáñez, M.; Elguero, J. Cooperative (nonpairwise) effects
Length in Peptide and Protein Folding Models. J. Phys. Chem. A 2003,
in water trimers: An ab initio molecular orbital study. J. Chem. Phys.
107, 10389−10395.
1992, 97, 6628−6638.
(106) Asensio, A.; Kobko, N.; Dannenberg, J. J. Cooperative
(83) Parthasarathi, R.; Subramanian, V.; Sathyamurthy, N. Hydrogen
Hydrogen-Bonding in Adenine-Thymine and Guanine-Cytosine Base
Bonding in Phenol, Water, and Phenol-Water Clusters. J. Phys. Chem. A
Pairs. Density Functional Theory and Møller-Plesset Molecular Orbital
2005, 109, 843−850.
(84) Neela, Y. I.; Mahadevi, A. S.; Sastry, G. N. Hydrogen Bonding in Study. J. Phys. Chem. A 2003, 107, 6441−6443.
Water Clusters and Their Ionized Counterparts. J. Phys. Chem. B 2010, (107) Chen, Y.; Dannenberg, J. J. Cooperative 4-Pyridone H-Bonds
114, 17162−17171. with Extraordinary Stability. A DFT Molecular Orbital Study. J. Am.
(85) Ohno, K.; Okimura, M.; Akai, N.; Katsumoto, Y. The effect of Chem. Soc. 2006, 128, 8100−8101.
cooperative hydrogen bonding on the OH stretching-band shift for (108) Mahadevi, A. S.; Neela, Y. I.; Sastry, G. N. A Theoretical Study
water clusters studied by matrix-isolation infrared spectroscopy and on Structural, Spectroscopic and Energetic Properties of Acetamide
density functional theory. Phys. Chem. Chem. Phys. 2005, 7, 3005−3014. Clusters [CH3CONH2] (n = 1−15). Phys. Chem. Chem. Phys. 2011, 13,
(86) Znamenskiy, V. S.; Green, M. E. Quantum Calculations on 15211−15220.
Hydrogen Bonds in Certain Water Clusters Show Cooperative Effects. J. (109) Mahadevi, A. S.; Neela, Y. I.; Sastry, G. N. Hydrogen Bonded
Chem. Theory Comput. 2007, 3, 103−114. Networks in Formamide [HCONH2]n (n = 1 − 10) clusters: A
(87) Ruckenstein, E.; Shulgin, I. L.; Shulgin, L. I. Cooperativity in Computational Exploration of Preferred Aggregation Patterns. J. Chem.
Ordinary Ice and Breaking of Hydrogen Bonds. J. Phys. Chem. B 2007, Sci. 2012, 124, 35−42.
111, 7114−7121. (110) Parra, R. D.; Bulusu, S.; Zeng, X. C. Cooperative Effects in One-
(88) Stokely, K.; Mazza, M. G.; Stanley, H. E.; Franzese, G. Effect of Dimensional Chains of Three-Center Hydrogen Bonding Interactions.
Hydrogen Bond Cooperativity on the Behavior of Water. Proc. Natl. J. Chem. Phys. 2003, 118, 3499−3509.
Acad. Sci. U. S. A. 2010, 107, 1301−1306. (111) Parra, R. D.; Bulusu, S.; Zeng, X. C. Cooperative Effects in Two-
(89) Albrecht, L.; Chowdhury, S.; Boyd, R. J. Hydrogen Bond Dimensional Ring-Like Networks of Three-Center Hydrogen Bonding
Cooperativity in Water Hexamers: Atomic Energy Perspective of Local Interactions. J. Chem. Phys. 2005, 122, 184325−8.
Stabilities. J. Phys. Chem. A 2013, 117, 10790−10799. (112) Parra, R. D.; Ohlssen, J. Cooperativity in Intramolecular
(90) Sciortino, F.; Fornili, S. L. Hydrogen Bond Cooperativity in Bifurcated Hydrogen Bonds: An Ab Initio Study. J. Phys. Chem. A 2008,
Simulated Water: Time Dependence Analysis of Pair Interactions. J. 112, 3492−3498.
Chem. Phys. 1989, 90, 2786−2792. (113) Ludwig, R.; Weinhold, F.; Farrar, T. C. Structure of Liquid N-
(91) Ojamaee, L.; Hermansson, K. Ab Initio Study of Cooperativity in Methylacetamide: Temperature Dependence of NMR Chemical Shifts
Water Chains: Binding Energies and Anharmonic Frequencies. J. Phys. and Quadrupole Coupling Constants. J. Phys. Chem. A 1997, 101, 8861−
Chem. 1994, 98, 4271−4282. 8870.

2808 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(114) King, B. F.; Weinhold, F. Structure and Spectroscopy of (HCN)n (136) Guo, H.; Karplus, M. Ab Initio Studies of Hydrogen Bonding of
Clusters: Cooperative and Electronic Delocalization Effects in C−HN N-Methylacetamide: Structure, Cooperativity, and Internal Rotational
Hydrogen Bonding. J. Chem. Phys. 1995, 103, 333−347. Barriers. J. Phys. Chem. 1992, 96, 7273−7287.
(115) Cabaleiro-Lago, E. M.; Ríos, M. A. Ab Initio Study of Interactions (137) Missopolinou, D.; Panayiotou, C. Hydrogen Bonding
in Hydrazine Clusters of One To Four Molecules: Cooperativity in the Cooperativity and Competing Inter- and Intramolecular Associations:
Interaction. J. Phys. Chem. A 1999, 103, 6468−6474. A Unified Approach. J. Phys. Chem. A 1998, 102, 3574−3581.
(116) van Mourik, T.; Dingley, A. J. Characterizing the Cooperativity (138) Dkhissi, A.; Ramaekers, R.; Houben, L.; Adamowicz, L.; Maes,
in H-Bonded Amino Structures. J. Phys. Chem. A 2007, 111, 11350− G. DFT/B3-LYP study of the hydrogen-bonding cooperativity:
11358. application to (2-pyridone)2, 2-pyridone-H2O, 2-pyridone-CH3OH
(117) Jiang, X.; Wang, C. Rapid Prediction of the Hydrogen Bond and 2-pyridone-CH3OCH3. Chem. Phys. Lett. 2000, 331, 553−560.
Cooperativity in N-methylacetamide Chains. ChemPhysChem 2009, 10, (139) Sum, A. K.; Sandler, S. I. Ab Initio Calculations of Cooperativity
3330−3336. Effects on Clusters of Methanol, Ethanol, 1-Propanol, and Meth-
(118) Sun, C.; Wang, C. Cooperative Influence of Water Binding to anethiol. J. Phys. Chem. A 2000, 104, 1121−1129.
Peptides by N-H···OH2 and CO···HOH Hydrogen Bonds: Study by (140) Zabardasti, A.; Amani, S.; Solimannejad, M.; Salehnassaj, M.
Ab Initio Calculations. Int. J. Quantum Chem. 2012, 112, 2336−2341. Theoretical Study and Atoms in Molecule Analysis of Hydrogen
(119) Xia, Q. Y.; Xiao, H. M.; Ju, X. H.; Gong, X. D. DFT Study on Bonded Clusters of Ammonia and Isocyanic Acid. Struct. Chem. 2009,
Cooperativity in the Interactions of Hydrazoic Acid Clusters. Int. J. 20, 1087−1092.
Quantum Chem. 2003, 94, 279−286. (141) Mandal, A.; Prakash, M.; Kumar, R. M.; Parthasarathi, R.;
(120) Kar, T.; Scheiner, S. Comparison of Cooperativity in CH···O Subramanian, V. Ab Initio and DFT Studies on Methanol-Water
and OH···O Hydrogen Bonds. J. Phys. Chem. A 2004, 108, 9161−9168. Clusters. J. Phys. Chem. A 2010, 114, 2250−2258.
(121) Scheiner, S. Solvation of Hydrogen Bonded Systems: CH···O, (142) Li, Q.; Jiang, L.; Wang, X.; Li, W.; Cheng, J.; Sun, J. Ab Initio
OH···O, and Cooperativity. In Solvation Effects On Molecules And Study of the Structure, Cooperativity, and Vibrational Properties in the
Biomolecules; Canuto, S., Ed; Springer: The Netherlands, 2008; p 407. Mixed Hydrogen-Bonded Trimers of Hydrogen Isocyanide and Water.
(122) Scheiner, S. Hydrogen Bonding: A Theoretical Perspective; Oxford Int. J. Quantum Chem. 2011, 111, 1072−1080.
University Press: New York, 1997. (143) Albrecht, L.; Boyd, R. J. Visualizing Internal Stabilization in
(123) Azofra, L. M.; Scheiner, S. Complexation of n SO2 Molecules (n Weakly Bound Systems Using Atomic Energies: Hydrogen Bonding in
= 1, 2, 3) with Formaldehyde and Thioformaldehyde. J. Chem. Phys. Small Water Clusters. J. Phys. Chem. A 2012, 116, 3946−3951.
2014, 140, 034302−10. (144) Schiros, T.; Ogasawara, H.; Näslund, L.-Å.; Andersson, K. J.;
(124) Azofra, L. M.; Scheiner, S. Tetrel, chalcogen, and CH··O Ren, J.; Meng, S.; Karlberg, G. S.; Odelius, M.; Nilsson, A.; Pettersson, L.
Hydrogen Bonds in Complexes Pairing Carbonyl-Containing Mole-
G. M. Cooperativity in Surface Bonding and Hydrogen Bonding of
cules with 1, 2, and 3 molecules of CO2. J. Chem. Phys. 2015, 142,
Water and Hydroxyl at Metal Surfaces. J. Phys. Chem. C 2010, 114,
034307.
10240−10248.
(125) Li, Q.; An, X.; Gong, B.; Cheng, J. Cooperativity between OH···
(145) Rincón, L.; Almeida, R.; Aldea, D. G. Many-Body Energy
O and CH···O Hydrogen Bonds Involving Dimethyl Sulfoxide-H2O-
Decomposition Analysis of Cooperativity in Hydrogen Fluoride
H2O Complex. J. Phys. Chem. A 2007, 111, 10166−10169.
Clusters. Int. J. Quantum Chem. 2005, 102, 443−453.
(126) Ziółkowski, M.; Grabowski, S. J.; Leszczynski, J. Cooperativity in
(146) Li, Q.; Liu, Z.; Cheng, J.; Li, W.; Gong, B.; Sun, J. Theoretical
Hydrogen-Bonded Interactions: Ab Initio and “Atoms in Molecules”
Study on the Cooperativity of Hydrogen Bonds in (HNC)2···HF
Analyses. J. Phys. Chem. A 2006, 110, 6514−6521.
(127) Xu, W.; Li, X.; Tan, H.; Chen, G. Theoretical Study on Stabilities Complexes. J. Mol. Struct.: THEOCHEM 2009, 896, 112−115.
of Multiple Hydrogen Bonded Dimers. Phys. Chem. Chem. Phys. 2006, 8, (147) Qi, H. W.; Leverentz, H. R.; Truhlar, D. G. Water 16-mers and
4427−4433. Hexamers: Assessment of the Three-Body and Electrostatically
(128) Li, Q.; An, X.; Luan, F.; Li, W.; Gong, B.; Cheng, J.; Sun, J. The Embedded Many-Body Approximations of the Correlation Energy or
Effect of Methyl Group on the Cooperativity between Three Types of the Nonlocal Energy As Ways to Include Cooperative Effects. J. Phys.
Hydrogen Bond: O-H···O, C-H···O, and O-H···π. Int. J. Quantum Chem. Chem. A 2013, 117, 4486−4499.
2008, 108, 558−566. (148) Shibl, M. F.; Pietrzak, M.; Limbach, H.; Kühn, O. Geometric H/
(129) Esrafili, M. D.; Beheshtian, J.; Hadipour, N. L. Computational D Isotope Effects and Cooperativity of the Hydrogen Bonds in
Study on the Characteristics of the Interaction in Linear Urea Clusters. Porphycene. ChemPhysChem 2007, 8, 315−321.
Int. J. Quantum Chem. 2011, 111, 3184−3195. (149) Li, Q.; An, X.; Luan, F.; Li, W.; Gong, B.; Cheng, J.; Sun, J.
(130) Khedkar, J. K.; Deshmukh, M. M.; Gadre, S. R.; Gejji, S. P. Cooperativity between Two Types of Hydrogen Bond in H3C−HCN−
Hydrogen Bond Energies and Cooperativity in Substituted Calix[n]- HCN and H3C−HNC−HNC Complexes. J. Chem. Phys. 2008, 128,
arenes (n = 4, 5). J. Phys. Chem. A 2012, 116, 3739−3744. 154102−6.
(131) Mahadevi, A. S.; Sastry, G. N. Modulation of Hydrogen Bonding (150) Li, Q.; Hu, T.; An, X.; Gong, B.; Cheng, J. Cooperativity between
upon Ion Binding: Insights into Cooperativity. Int. J. Quantum Chem. the Dihydrogen Bond and the N···HC Hydrogen Bond in LiH−
2014, 114, 145−153. (HCN)n Complexes. ChemPhysChem 2008, 9, 1942−1946.
(132) Zabardasti, A.; Kakanejadi, A.; Moosavi, S.; Bigleri, Z.; (151) Alkorta, I.; Blanco, F.; Elguero, J. Dihydrogen Bond
Solimannejad, M. Anticooperativity in Dihydrogen Bonded Clusters Cooperativity in Aza-borane Derivatives. J. Phys. Chem. A 2010, 114,
of Ammonia and BeH42‑. J. Mol. Struct.: THEOCHEM 2010, 945, 97− 8457−8462.
100. (152) Latajka, Z.; Scheiner, S. Structure, Energetics and Vibrational
(133) Zabardasti, A.; Kakanejadi, A.; Ghenaatian, F.; Bigleri, Z. A Spectra of Dimers, Trimers, and Tetramers of HX(X = C1, Br, I). Chem.
Theoretical Study of Cooperative and Anticooperative Effects on Phys. 1997, 216, 37−52.
Hydrogen-Bonded Clusters of Water and the Cyanuric Acid. Mol. Simul. (153) Rossetti, G.; Magistrato, A.; Pastore, A.; Carloni, P. Hydrogen
2010, 36, 960−968. Bonding Cooperativity in polyQ β-Sheets from First Principle
(134) Albrecht, L.; Boyd, R. J. Atomic Energy Analysis of Calculations. J. Chem. Theory Comput. 2010, 6, 1777−1782.
Cooperativity, Anticooperativity, and Noncooperativity in Small (154) Gung, B. W.; Zhu, Z.; Zou, D.; Everingham, B.; Oyeamalu, A.;
Clusters of Methanol, Water, and Formaldehyde. Comput. Theor. Crist, R. M.; Baudlier, J. Requirement for Hydrogen-Bonding
Chem. 2015, 1053, 328−336. Cooperativity in Small Polyamides: A Combined VT-NMR and VT-
(135) Kleeberg, H.; Klein, D.; Luck, W. A. P. Quantitative Infrared IR Investigation. J. Org. Chem. 1998, 63, 5750−5761.
Spectroscopic Investigations of Hydrogen-Bond Cooperativity. J. Phys. (155) Knott, M.; Chan, H. S. Exploring the Effects of Hydrogen
Chem. 1987, 91, 3200−3203. Bonding and Hydrophobic Interactions on the Foldability and

2809 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Cooperativity of Helical Proteins Using a Simplified Atomic Model. (176) Vicente, V.; Martin, J.; Jiménez-Barbero, J.; Chiara, J. L.; Vicent,
Chem. Phys. 2004, 307, 187−199. C. Hydrogen-Bonding Cooperativity: Using an Intramolecular Hydro-
(156) Wieczorek, R.; Dannenberg, J. J. Hydrogen Bond Cooperativity, gen Bond To Design a Carbohydrate Derivative with a Cooperative
Vibrational Coupling, and Dependence of Helix Stability on Changes in Hydrogen-Bond Donor Centre. Chem.−Eur. J. 2004, 10, 4240−4251.
Amino Acid Sequence in Small 310-Helical Peptides. A Density (177) Gilli, G.; Bellucci, F.; Ferretti, V.; Bertolasi, V. Evidence for
Functional Theory Study. J. Am. Chem. Soc. 2003, 125, 14065−14071. Resonance-Assisted Hydrogen Bonding from Crystal-Structure Corre-
(157) Wieczorek, R.; Dannenberg, J. J. H-Bonding Cooperativity and lations on the Enol Form of the β-Diketone Fragment. J. Am. Chem. Soc.
Energetics of α-Helix Formation of Five 17-Amino Acid Peptides. J. Am. 1989, 111, 1023−1028.
Chem. Soc. 2003, 125, 8124−8129. (178) Bertolasi, V.; Cilli, P.; Ferretti, V.; Gilli, G. Evidence for
(158) Wieczorek, R.; Dannenberg, J. J. Comparison of Fully Optimized Resonance-Assisted Hydrogen Bonding. 2.1 Intercorrelation between
α- and 310-Helices with Extended β-Strands. An ONIOM Density Crystal Structure and Spectroscopic Parameters in Eight Intra-
Functional Theory Study. J. Am. Chem. Soc. 2004, 126, 14198−14205. molecularly Hydrogen Bonded 1,3-Diaryl-1,3-propanedione Enols. J.
(159) Salvador, P.; Kobko, N.; Wieczorek, R.; Dannenberg, J. J. Am. Chem. Soc. 1991, 113, 4917−4925.
Calculation of trans-Hydrogen-Bond 13C-15N Three-Bond and Other (179) Bertolasi, V.; Pretto, L.; Gilli, G.; Gilli, P. π-Bond Cooperativity
Scalar J-Couplings in Cooperative Peptide Models. A Density and Anticooperativity Effects in Resonance-Assisted Hydrogen Bonds
Functional Theory Study. J. Am. Chem. Soc. 2004, 126, 14190−14197. (RAHBs). Acta Crystallogr., Sect. B: Struct. Sci. 2006, 62, 850−863.
(160) Sheridan, R. P.; Lee, R. H.; Peters, N.; Allen, L. C. Hydrogen- (180) Mignon, P.; Loverix, S.; Steyaert, J.; Geerlings, P. Influence of the
Bond Cooperativity in Protein Secondary Structure. Biopolymers 1979, π−π Interaction on the Hydrogen Bonding Capacity of Stacked DNA/
18, 2451−2458. RNA Bases. Nucleic Acids Res. 2005, 33, 1779−1789.
(161) Van Duijnen, P. T.; Thole, B. T. Cooperative Effects in α- (181) Vijay, D.; Zipse, H.; Sastry, G. N. On the Cooperativity of
Helices: An Ab Initio Molecular-Orbital Study. Biopolymers 1982, 21, Cation-π and Hydrogen Bonding Interactions. J. Phys. Chem. B 2008,
1749−1761. 112, 8863−8867.
(162) Kennedy, R. J.; Tsang, K.; Kemp, D. S. Consistent Helicities (182) Kruszynski, R.; Trzesowska-Kruszynska, A. Halogen, Hydrogen
from CD and Template t/c Data for N-Templated Polyalanines: and Electrostatic Interactions in 2-amino-5-chloro-1,3-benzoxazol-3-
Progress toward Resolution of the Alanine Helicity Problem. J. Am. ium nitrate and 2-amino-5-chloro-1,3-benzoxazol-3-ium perchlorate.
Chem. Soc. 2002, 124, 934−944. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2010, 66, o449−o454.
(163) Wu, Y.; Zhao, Y. A Theoretical Study on the Origin of (183) Rieth, S.; Li, Z.; Hinkle, C. E.; Guzman, C. X.; Lee, J. J.; Nehme,
Cooperativity in the Formation of 310- and α-Helices. J. Am. Chem. Soc. S. I.; Braunschweig, A. B. Superstructures of Diketopyrrolopyrrole
2001, 123, 5313−5319. Donors and Perylenediimide Acceptors Formed by Hydrogen-Bonding
(164) Zhao, Y.; Wu, Y. A Theoretical Study of β-Sheet Models: Is the and π···π Stacking. J. Phys. Chem. C 2013, 117, 11347−11356.
Formation of Hydrogen-Bond Networks Cooperative? J. Am. Chem. Soc. (184) León, I.; Millán, J.; Cocinero, E. J.; Lesarri, A.; Fernández, J. A.
2002, 124, 1570−1571. Shaping Micelles: The Interplay between Hydrogen Bonds and
(165) Morozov, A. V.; Tsemekhman, K.; Baker, D. Electron Density Dispersive Interactions. Angew. Chem., Int. Ed. 2013, 52, 7772−7775.
Redistribution Accounts for Half the Cooperativity of α Helix (185) Seifert, N. A.; Steber, A. L.; Neill, J. L.; Pérez, C.; Zaleski, D. P.;
Formation. J. Phys. Chem. B 2006, 110, 4503−4505. Pate, B. H.; Lesarri, A. The Interplay of Hydrogen Bonding and
(166) Tsemekhman, K.; Goldschmidt, L.; Eisenberg, D.; Baker, D. Dispersion in Phenol Dimer and Trimer: Structures from Broadband
Cooperative Hydrogen Bonding in Amyloid Formation. Protein Sci. Rotational Spectroscopy. Phys. Chem. Chem. Phys. 2013, 15, 11468−
2007, 16, 761−764. 11477.
(167) Wieczorek, R.; Dannenberg, J. J. Amide I Vibrational (186) Guo, J.; Shi, W.; Ren, F.; Cao, D.; Zhang, Y. A B3LYP and
Frequencies of α-Helical Peptides Based upon ONIOM and Density MP2(full) Theoretical Investigation into the Cooperativity Effect
Functional Theory (DFT) Studies. J. Phys. Chem. B 2008, 112, 1320− between Dihydrogen-Bonding and H−M•••π (M = Li, Na, K)
1328. Interactions among HF, MH with the π-Electron Donor C2H2, C2H4 or
(168) Plumley, J. A.; Dannenberg, J. J. The Importance of Hydrogen C6H6. J. Mol. Model. 2013, 19, 3153−3163.
Bonding between the Glutamine Side Chains to the Formation of (187) Alkorta, I.; Elguero, J.; Solimannejad, M.; Grabowski, S. J.
Amyloid VQIVYK Parallel β-Sheets: An ONIOM DFT/AM1 Study. J. Dihydrogen Bonding vs Metal-σ Interaction in Complexes between H2
Am. Chem. Soc. 2010, 132, 1758−1759. and Metal Hydride. J. Phys. Chem. A 2011, 115, 201−210.
(169) Ireta, J.; Neugebauer, J.; Scheffler, M.; Rojo, A.; Galván, M. (188) Kovács, A.; Varga, Z. Halogen Acceptors in Hydrogen Bonding.
Density Functional Theory Study of the Cooperativity of Hydrogen Coord. Chem. Rev. 2006, 250, 710−727.
Bonds in Finite and Infinite α-Helices. J. Phys. Chem. B 2003, 107, (189) Grabowski, S. J. Cooperativity of Hydrogen and Halogen Bond
1432−1437. Interactions. Theor. Chem. Acc. 2013, 132, 1347.
(170) Parker, L. L.; Houk, A. R.; Jensen, J. H. Cooperative Hydrogen (190) Jing, B.; Li, Q.; Li, R.; Gong, B.; Liu, Z.; Li, W.; Cheng, J.; Sun, J.
Bonding Effects Are Key Determinants of Backbone Amide Proton Competition and Cooperativity between Hydrogen Bond and Halogen
Chemical Shifts in Proteins. J. Am. Chem. Soc. 2006, 128, 9863−9872. Bond in HNC···(HOBr)n and (HNC)n···HOBr (n = 1 and 2) Systems.
(171) Albeck, S.; Unger, R.; Schreiber, G. Evaluation of Direct and Comput. Theor. Chem. 2011, 963, 417−421.
Cooperative Contributions Towards the Strength of Buried Hydrogen (191) Li, Q.; Lin, Q.; Li, W.; Cheng, J.; Gong, B.; Sun, J. Cooperativity
Bonds and Salt Bridges. J. Mol. Biol. 2000, 298, 503−520. between the Halogen Bond and the Hydrogen Bond in H3N···XY···HF
(172) Kung, V. M.; Cornilescu, G.; Gellman, S. H. Impact of Strand Complexes (X, Y = F, Cl, Br). ChemPhysChem 2008, 9, 2265−2269.
Number on Parallel β-Sheet Stability. Angew. Chem., Int. Ed. 2015, 54, (192) Li, Q.; Ma, S.; Liu, X.; Li, W.; Cheng, J. Cooperative and
14336. Substitution Effects in Enhancing Strengths of Halogen Bonds in
(173) Nochebuena, J.; Ireta, J. On Cooperative Effects and FClCNX Complexes. J. Chem. Phys. 2012, 137, 084314−8.
Aggregation of GNNQQNY and NNQQNY Peptides. J. Chem. Phys. (193) Wu, W.; Zeng, Y.; Li, X.; Zhang, X.; Zheng, S.; Meng, L. Interplay
2015, 143, 135103. between Halogen Bonds and Hydrogen Bonds in OH/SH···HOX···HY
(174) Dashnau, J. L.; Sharp, K. A.; Vanderkooi, J. M. Carbohydrate (X = Cl, Br; Y = F, Cl, Br) Complexes. J. Mol. Model. 2013, 19, 1069−
Intramolecular Hydrogen Bonding Cooperativity and Its Effect on 1077.
Water Structure. J. Phys. Chem. B 2005, 109, 24152−24159. (194) McDowell, S. A. C.; Yarde, H. K. Cooperative effects of
(175) Deshmukh, M. M.; Bartolotti, L. J.; Gadre, S. R. Intramolecular Hydrogen, Lithium and Halogen Bonding on F−H/Li···OH 2
Hydrogen Bonding and Cooperative Interactions in Carbohydrates via Complexes. Phys. Chem. Chem. Phys. 2012, 14, 6883−6888.
the Molecular Tailoring Approach. J. Phys. Chem. A 2008, 112, 312− (195) Domagała, M.; Palusiak, M. The Influence of Substituent Effect
321. on Noncovalent Interactions in Ternary Complexes Stabilized by

2810 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Hydrogen-Bonding and Halogen-Bonding. Comput. Theor. Chem. 2014, (214) Anand, M.; Fernandez, I.; Schaefer, H. F., III; Wu, J. I. Hydrogen
1027, 173−178. Bond−Aromaticity Cooperativity in Self-Assembling 4-Pyridone
(196) Li, Q.; Hu, T.; An, X.; Li, W.; Cheng, J.; Gong, B.; Sun, J. Chains. J. Comput. Chem. 2016, 37, 59.
Theoretical Study of the Interplay between Lithium Bond and (215) Axilrod, B. M.; Teller, E. Interaction of the van der Waals Type
Hydrogen Bond in Complexes Involved with HLi and HCN. between Three Atoms. J. Chem. Phys. 1943, 11, 299−300.
ChemPhysChem 2009, 10, 3310−3315. (216) Muto, Y. Force between Nonpolar Molecules. Proc. Phys.-Math
(197) Albrecht, L.; Boyd, R. J.; Mó, O.; Yáñez, M. Cooperativity Soc. Jpn. 1943, 17, 629−631.
between Hydrogen Bonds and Beryllium Bonds in (H2O)nBeX2 (n = 1− (217) Bates, D. M.; Smith, J. R.; Janowski, T.; Tschumper, G. S.
3, X = H, F) Complexes. A New Perspective. Phys. Chem. Chem. Phys. Development of a 3-Body: Many-Body Integrated Fragmentation
2012, 14, 14540−14547. Method for Weakly Bound Clusters and Application to Water Clusters
(198) Alberto, M. E.; Mazzone, G.; Russo, N.; Sicilia, E. The Mutual (H2O)n=3−10, 16, 17. J. Chem. Phys. 2011, 135, 044123−8.
Influence of Noncovalent Interactions in π-Electron Deficient Cavities: (218) Góra, U.; Podeszwa, R.; Cencek, W.; Szalewicz, K. Interaction
The Case of Anion Recognition by Tetraoxacalix[2]arene[2]triazine. Energies of Large Clusters from Many-Body Expansion. J. Chem. Phys.
Chem. Commun. 2010, 46, 5894−5896. 2011, 135, 224102.
(199) Natale, D.; Mareque-Rivas, J. C. The Combination of Transition (219) Shimizu, S.; Chan, H. S. AntiCooperativity and Cooperativity in
Metal Ions and Hydrogen Bonding Interactions. Chem. Commun. 2008, Hydrophobic Interactions: Three-Body Free Energy Landscapes and
425−437. Comparison with Implicit-Solvent Potential Functions for Proteins.
(200) Seifert, N. A.; Zaleski, D. P.; Pérez, C.; Neill, J. L.; Pate, B. H.; Proteins: Struct., Funct., Genet. 2002, 48, 15−30.
Vallejo-López, M.; Lesarri, A.; Cocinero, E. J.; Castaño, F.; Kleiner, I. (220) Matsumoto, M. Four-Body Cooperativity in Hydrophonic
Probing the C-H···π Weak Hydrogen Bond in Anesthetic Binding: The Association of Methane. J. Phys. Chem. Lett. 2010, 1, 1552−1556.
Sevoflurane−Benzene Cluster. Angew. Chem., Int. Ed. 2014, 53, 3210− (221) von Lilienfeld, O. A.; Tkatchenko, A. Two- and Three-Body
3213. Interatomic Dispersion Energy Contributions to Binding in Molecules
(201) Kobayashi, Y.; Saigo, K. Periodic Ab Initio Approach for the and Solids. J. Chem. Phys. 2010, 132, 234109−11.
Cooperative Effect of CH/π Interaction in Crystals: Relative Energy of (222) Wang, L.; Friesner, R. A.; Berne, B. J. Hydrophobic Interactions
CH/π and Hydrogen-Bonding Interactions. J. Am. Chem. Soc. 2005, 127, in Model Enclosures from Small to Large Length Scales: Nonadditivity
15054−15060. in Explicit and Implicit Solvent Models. Faraday Discuss. 2010, 146,
(202) Li, Q.; Li, H.; Li, R.; Jing, B.; Liu, Z.; Li, W.; Luan, F.; Cheng, J.; 247−262.
Gong, B.; Sun, J. Influence of Hybridization and Cooperativity on the (223) Farina, C.; Santos, F. C.; Tort, A. C. A Simple Way of
Properties of Au-Bonding Interaction: Comparison with Hydrogen Understanding the Nonadditivity of van der Waals Dispersion Forces.
Bonds. J. Phys. Chem. A 2011, 115, 2853−2858. Am. J. Phys. 1999, 67, 344−349.
(203) Grabowski, S. J.; Leszczynski, J. The Enhancement of X−H···π (224) Sponer, J.; Gabb, H. A.; Leszczynski, J.; Hobza, P. Base-Base and
Hydrogen Bond by Cooperativity Effects − Ab initio and QTAIM
Deoxyribose-Base Stacking Interactions in B-DNA and Z-DNA: A
calculations. Chem. Phys. 2009, 355, 169−176.
Quantum-Chemical Study. Biophys. J. 1997, 73, 76−87.
(204) Mo, Y. Can QTAIM Topological Parameters Be a Measure of
(225) Valiron, P.; Mayer, I. Hierarchy of Counterpoise Corrections for
Hydrogen Bonding Strength? J. Phys. Chem. A 2012, 116, 5240−5246.
N-body Clusters: Generalization of the Boys-Bernardi Scheme. Chem.
(205) Gholipour, A. R.; Saydi, H.; Neiband, M. S.; Neyband, R. S.
Phys. Lett. 1997, 275, 46−55.
Simultaneous Interactions of Pyridine with Substituted Benzene Ring
(226) Tkatchenko, A.; von Lilienfeld, O. A. Popular Kohn-Sham
and H−F in X-ben⊥pyr···H−F Complexes: A Cooperative Study. Struct.
Density Functionals Strongly Overestimate Many-Body Interactions in
Chem. 2012, 23, 367−373.
(206) Solimannejad, M.; Rezaei, Z.; Esrafili, M. D. Competition and van der Waals Systems. Phys. Rev. B: Condens. Matter Mater. Phys. 2008,
Interplay between the Lithium Bonding and Hydrogen Bonding: R3C··· 78, 045116−6.
HY···LiY and R3C···LiY···HY Triads as a Working Model (R = H, CH3; (227) Chmela, J.; Harding, M. E.; Matioszek, D.; Anson, C. E.; Breher,
Y = CN, NC). J. Mol. Model. 2013, 19, 5031−5035. F.; Klopper, W. Differential Many-Body Cooperativity in Electronic
(207) Solimannejad, M.; Malekani, M.; Alkorta, I. Cooperativity Spectra of Oligonuclear Transition-Metal Complexes. ChemPhysChem
between the Hydrogen Bonding and Halogen Bonding in F3CX··· 2015, DOI: 10.1002/cphc.201500626.
NCH(CNH)···NCH(CNH) complexes (X = Cl, Br). Mol. Phys. 2011, (228) Schreiber, G.; Fersht, A. R. Energetics of Protein-Protein
109, 1641−1648. Interactions: Analysis of the Barnase-Barstar Interface by Single
(208) Tian, Q.; Wang, Y.; Shi, W.; Song, S.; Tang, H. A Theoretical Mutations and Double Mutant Cycles. J. Mol. Biol. 1995, 248, 478−486.
Investigation into the Cooperativity Effect between the H···O and H··· (229) Searle, M. S.; Griffiths-Jones, S. R.; Skinner-Smith, H. Energetics
F−Interactions and Electrostatic Potential upon 1:2 (F − :N- of Weak Interactions in a β-hairpin Peptide: Electrostatic and
(Hydroxymethyl)acetamide) Ternary-System Formation. J. Mol. Hydrophobic Contributions to Stability from Lysine Salt Bridges. J.
Model. 2013, 19, 5171−5185. Am. Chem. Soc. 1999, 121, 11615−11620.
(209) Kříž, J.; Dautzenberg, H. Cooperative Interactions of Unlike (230) Ullmann, R. T.; Ullmann, G. M. Coupling of Protonation,
Macromolecules 2: NMR and Theoretical Study of Electrostatic Binding Reduction, and Conformational Change in Azurin from Pseudomonas
of Sodium Poly(styrenesulfonate)s to Copolymers with Variably aeruginosa Investigated with Free Energy Measures of Cooperativity. J.
Distributed Cationic Groups. J. Phys. Chem. A 2001, 105, 3846−3854. Phys. Chem. B 2011, 115, 10346−10359.
(210) Kříž, J.; Dybal, J. Cooperative Hydrogen Bonds of Macro- (231) Hunter, C. A.; Jones, P. S.; Tiger, P.; Tomas, S. Chemical Triple-
molecules. 3. A Model Study of the Proximity Effect. J. Phys. Chem. B Mutant Boxes for Quantifying Cooperativity in Intermolecular
2007, 111, 6118−6126. Interactions. Chem. - Eur. J. 2002, 8, 5435−5446.
(211) Remya, K.; Suresh, C. H. Cooperativity and Cluster Growth (232) Hunter, C. A.; Tomas, S. Cooperativity, Partially Bound States,
Patterns in Acetonitrile: A DFT Study. J. Comput. Chem. 2014, 35, 910− and Enthalpy-Entropy Compensation. Chem. Biol. 2003, 10, 1023−
922. 1032.
(212) Mohan, N.; Suresh, C. H. Accurate Binding Energies of (233) Camara-Campos, A.; Musumeci, D.; Hunter, C. A.; Turega, S.
Hydrogen, Halogen, and Dihydrogen Bonded Complexes and Cation Chemical Double Mutant Cycles for the Quantification of Cooperativity
Enhanced Binding Strengths. Int. J. Quantum Chem. 2014, 114, 885− in H-Bonded Complexes. J. Am. Chem. Soc. 2009, 131, 18518−18524.
894. (234) Hunter, C. A.; Misuraca, M. C.; Turega, S. M. Solvent Effects on
(213) Zhuo, H.; Li, Q.; Li, W.; Cheng, J. Non-additivity between Chelate Cooperativity. Chem. Sci. 2012, 3, 589−601.
Substitution and Cooperative Effects in Enhancing Hydrogen Bonds. J. (235) Acerenza, L.; Mizraji, E. Cooperativity: A Unified View. Biochim.
Chem. Phys. 2014, 141, 244305. Biophys. Acta, Protein Struct. Mol. Enzymol. 1997, 1339, 155−166.

2811 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(236) Bloom, C. R.; Kaarsholm, N. C.; Ha, J.; Dunn, M. F. Half-Site (257) Mahadevi, A. S.; Sastry, G. N. A Theoretical Study on Interaction
Reactivity, Negative Cooperativity, and Positive Cooperativity: of Cyclopentadienyl Ligand with Alkali and Alkaline Earth Metals. J.
Quantitative Considerations of a Plausible Model. Biochemistry 1997, Phys. Chem. B 2011, 115, 703−710.
36, 12759−12765. (258) Vijay, D.; Sastry, G. N. Exploring the Size Dependence of Cyclic
(237) Andersen, P. S.; Schuck, P.; Sundberg, E. J.; Geisler, C.; and Acyclic π-Systems on Cation−π Binding. Phys. Chem. Chem. Phys.
Karjalainen, K.; Mariuzza, R. A. Quantifying the Energetics of 2008, 10, 582−590.
Cooperativity in a Ternary Protein Complex. Biochemistry 2002, 41, (259) Vijay, D.; Sastry, G. N. A Computational Study on π and σ
5177−5184. Modes of Metal Ion Binding to Heteroaromatics (CH)5‑mXm and
(238) Ng, C. K. L.; Li, N. X.; Chee, S.; Prabhakar, S.; Kolatkar, P. R.; (CH)6‑mXm (X = N and P): Contrasting Preferences between Nitrogen
Jauch, R. Deciphering the Sox-Oct Partner Code by Quantitative and Phosphorous-Substituted Rings. J. Phys. Chem. A 2006, 110,
Cooperativity Measurements. Nucleic Acids Res. 2012, 40, 4933−4941. 10148−10154.
(239) Varfolomeev, M. A.; Klimovitskii, A. E.; Abaidullina, D. I.; (260) Sharma, B.; Umadevi, D.; Sastry, G. N. Contrasting Preferences
Madzhidov, T. I.; Solomonov, B. N. Additive” Cooperativity of of N and P Substituted Heteroaromatics Towards Metal Binding:
Hydrogen Bonds in Complexes of Catechol with Proton Acceptors in Probing the Regioselectivity of Li+ and Mg2+ binding to (CH)6‑m‑nNmPn.
the Gas Phase: FTIR Spectroscopy and Quantum Chemical Phys. Chem. Chem. Phys. 2012, 14, 13922−13932.
Calculations. Spectrochim. Acta, Part A 2012, 91, 75−82. (261) Priyakumar, U. D.; Punnagai, M.; Krishna Mohan, G. P.; Sastry,
(240) Roubeau, O.; Castro, M.; Burriel, R.; Haasnoot, J. G.; Reedijk, J. G. N. A Computational Study of Cation−π Interactions in Polycyclic
Calorimetric Investigation of Triazole-Bridged Fe(II) Spin-Crossover Systems: Exploring the Dependence on the Curvature and Electronic
One-Dimensional Materials: Measuring the Cooperativity. J. Phys. Factors. Tetrahedron 2004, 60, 3037−3043.
Chem. B 2011, 115, 3003−3012. (262) Priyakumar, U. D.; Sastry, G. N. Cation-π Interactions of Curved
(241) Garcés, J. L.; Acerenza, L.; Mizraji, E.; Mas, F. A Hierarchical Polycyclic Systems: M+ (M = Li and Na) Ion Complexation with
Approach to Cooperativity in Macromolecular and Self-Assembling Buckybowls. Tetrahedron Lett. 2003, 44, 6043−6046.
Binding Systems. J. Biol. Phys. 2008, 34, 213−235. (263) Priyakumar, U. D.; Sastry, G. N. Structures, Energetics, Relative
(242) Douglas, J. F.; Dudowicz, J.; Freed, K. F. Lattice Model of Stabilities, and Out-of-Plane Distortivities of Skeletally Disubstituted
Equilibrium Polymerization. VII. Understanding the Role of “Cooper- Benzenes, (CH)4X2 (X = N, P, C−, Si−, O+, and S+): An Ab Initio and
ativity” in Self-Assembly. J. Chem. Phys. 2008, 128, 224901−17. DFT Study. J. Am. Chem. Soc. 2000, 122, 11173−11181.
(243) Hamacek, J.; Borkovec, M.; Piguet, C. A Simple Thermodynamic (264) Priyakumar, U. D.; Sastry, G. N. Heterobuckybowls: A
Model for Quantitatively Addressing Cooperativity in Multicomponent Theoretical Study on the Structure, Bowl-to-Bowl Inversion Barrier,
Self-Assembly ProcessesPart 1: Theoretical Concepts and Applica- Bond Length Alternation, Structure-Inversion Barrier Relationship,
tion to Monometallic Coordination Complexes and Bimetallic Helicates Stability, and Synthetic Feasibility. J. Org. Chem. 2001, 66, 6523−6530.
Possessing Identical Binding Sites. Chem. - Eur. J. 2005, 11, 5217−5226. (265) Priyakumar, U. D.; Sastry, G. N. First Ab Initio and Density
(244) Douglass, E. F., Jr.; Miller, C. J.; Sparer, G.; Shapiro, H.; Spiegel, Functional Study on the Structure, Bowl-to-Bowl Inversion Barrier, and
D. A. A Comprehensive Mathematical Model for Three-Body Binding Vibrational Spectra of the Elusive C3v-Symmetric Buckybowl:
Equilibria. J. Am. Chem. Soc. 2013, 135, 6092−6099. Sumanene, C21H12. J. Phys. Chem. A 2001, 105, 4488−4494.
(245) Wang, R.; Nganso, D. S.; Kaabouchi, E. A.; Wang, Q. A. (266) Rao, J. S.; Sastry, G. N. Structural and Energetic Preferences of π,
Investigation of an Energy Nonadditivity for Nonextensive Systems. σ, and Bidentate Cation Binding (Li+, Na+, and Mg2+) to Aromatic
Chin. Sci. Bull. 2011, 56, 3661−3665. Amines (Ph-(CH2)n-NH2, n = 2−5): A Theoretical Study. J. Phys. Chem.
(246) Kumar, R. M.; Vijay, D.; Sastry, G. N.; Subramanian, V. A 2009, 113, 5446−5454.
Intermolecular Interactions through Energy Decomposition A Chem- (267) Srivastava, H. K.; Sastry, G. N. Viability of Clathrate Hydrates as
ists’ Perspective. In Concepts and Methods in Modern Theoretical CO2 Capturing Agents: A Theoretical Study. J. Phys. Chem. A 2011, 115,
Chemistry; Ghosh, S. K., Chattaraj, P. K., Eds.; CRC Press, 2013; p 313. 7633−7637.
(247) White, J. C.; Davidson, E. R. An Analysis of the Hydrogen Bond (268) Dinadayalane, T. C.; Sastry, G. N.; Leszczynski, J.
in Ice. J. Chem. Phys. 1990, 93, 8029−8035. Comprehensive Theoretical Study Towards the Accurate Proton
(248) Vijay, D.; Sastry, G. N. The Cooperativity of Cation−π and π−π Affinity Values of Naturally Occurring Amino Acids. Int. J. Quantum
Interactions. Chem. Phys. Lett. 2010, 485, 235−242. Chem. 2006, 106, 2920−2933.
(249) Li, Q.; Zhuo, H.; Yang, X.; Cheng, J.; Li, W.; Loffredo, R. E. (269) Kumar, M. K.; Rao, J. S.; Prabhakar, S.; Vairamani, M.; Sastry, G.
Cooperative and Diminutive Effects of Pnicogen Bonds and Cation−π N. The Effect of Spacer Chain Length on Ion Binding to Bidentate α, ω-
Interactions. ChemPhysChem 2014, 15, 500−506. diamines: Contrasting Ordering for H+ and Li+ ion affinities. Chem.
(250) Carrazana-García, J. A.; Rodríguez-Otero, J.; Cabaleiro-Lago, E. Commun. 2005, 1420−1422.
M. A Computational Study of Anion-Modulated Cation−π Interactions. (270) Reddy, A. S.; Vijay, D.; Sastry, G. M.; Sastry, G. N. From Subtle
J. Phys. Chem. B 2012, 116, 5860−5871. to Substantial: Role of Metal Ions on π-π Interactions. J. Phys. Chem. B
(251) Antony, J.; Brüske, B.; Grimme, S. Cooperativity in Noncovalent 2006, 110, 2479−2481.
Interactions of Biologically Relevant Molecules. Phys. Chem. Chem. Phys. (271) Reddy, A. S.; Vijay, D.; Sastry, G. M.; Sastry, G. N. Reply to
2009, 11, 8440−8447. “Comment on ‘From Subtle to Substantial: Role of Metal Ions on π-π
(252) Mandal, T. K.; Samanta, S.; Chakraborty, S.; Datta, A. An Interactions’. J. Phys. Chem. B 2006, 110, 10206−10207.
Interplay of Cooperativity between Cation···π, Anion···π and CH··· (272) Purushotham, U.; Sastry, G. N. A Comprehensive Conforma-
Anion Interactions. ChemPhysChem 2013, 14, 1149−1154. tional Analysis of Tryptophan, its Ionic and Dimeric Forms. J. Comput.
(253) Liu, X.; Li, Q.; Cheng, J.; Li, W. Influence of Cooperativity on the Chem. 2014, 35, 595−610.
Frequency Shift of the Ar−H Stretch Vibration in HArF Complexes. (273) Purushotham, U.; Sastry, G. N. Exploration of Conformations
Mol. Phys. 2013, 111, 497−504. and Quantum Chemical Investigation of L-Tyrosine Dimers, Anions,
(254) Li, Q.; Zhu, H.; An, X.; Gong, B.; Cheng, J. Nonadditivity of Cations and Zwitterions: a DFT study. Theor. Chem. Acc. 2012, 131,
Methyl Group in Single-Electron Hydrogen Bond of Methyl Radical- 1093−14.
Water Complex. Int. J. Quantum Chem. 2009, 109, 605−611. (274) Purushotham, U.; Vijay, D.; Sastry, G. N. A Computational
(255) Vijay, D.; Sakurai, H.; Subramanian, V.; Sastry, G. N. Where to Investigation and the Conformational Analysis of Dimers, Anions,
Bind in Buckybowls? The Dilemma of a Metal Ion. Phys. Chem. Chem. Cations, and Zwitterions of L-Phenylalanine. J. Comput. Chem. 2012, 33,
Phys. 2012, 14, 3057−3065. 44−59.
(256) Premkumar, J. R.; Vijay, D.; Sastry, G. N. The Significance of the (275) Saha, S.; Sastry, G. N. Cooperative or Anticooperative: How
Alkene Size and the Nature of the Metal Ion in Metal−Alkene Non-covalent Interactions Influence Each Other. J. Phys. Chem. B 2015,
Complexes: A Theoretical Study. Dalton Trans. 2012, 41, 4965−4975. 119, 11121−11135.

2812 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(276) Sharma, B.; Srivastava, H. K.; Gayatri, G.; Sastry, G. N. Energy Aromatic Amine Sorption to Aluminosilicates and Soils: Role of
Decomposition Analysis of Cation-π, Metal Ion-Lone Pair, Hydrogen Intermolecular Cation−π Interactions. Environ. Sci. Technol. 2013, 47,
Bonded, Charge Assisted Hydrogen Bonded and π-π Interactions. J. 14119−14127.
Comput. Chem. 2015, 36, 529−538. (296) Song, J.; Ng, S. C.; Tompa, P.; Lee, K. A. W.; Chan, H. S.
(277) Hussain, M. A.; Mahadevi, A. S.; Sastry, G. N. Estimating the Polycation-π Interactions Are a Driving Force for Molecular
Binding Ability of Onium Ions with CO2 and π Systems: A Recognition by an Intrinsically Disordered Oncoprotein Family. PLoS
Computational Investigation. Phys. Chem. Chem. Phys. 2015, 17, Comput. Biol. 2013, 9, e1003239−12.
1763−1775. (297) Dimitrijević, B. P.; Borozan, S. Z.; Stojanović, S. Đ. π−π and
(278) Neela, Y. I.; Sastry, G. N. Theoretical Investigation of Anion (F−, Cation−π Interactions in Protein−Porphyrin Complex Crystal
Cl−) and Cation (Na+) Interactions with Substituted Benzene Structures. RSC Adv. 2012, 2, 12963−12972.
[C6H6‑nYn (Y = -F, -CN, -NO2; n = 1−6)]. Mol. Phys. 2015, 113, (298) Francisco, V.; Basílio, N.; García-Río, L. Ionic Exchange in p-
137−148. Sulfonatocalix[4]arene-Mediated Formation of Metal−Ligand Com-
(279) Biot, C.; Wintjens, R.; Rooman, M. Stair Motifs at Protein-DNA plexes. J. Phys. Chem. B 2014, 118, 4710−4716.
Interfaces: Nonadditivity of H-Bond, Stacking, and Cation-π (299) Feng, G.; Qi, T.; Shi, W.; Guo, Y.; Zhang, Y.; Guo, J.; Kang, L. A
Interactions. J. Am. Chem. Soc. 2004, 126, 6220−6221. B3LYP and MP2(full) Theoretical Investigation on the Cooperativity
(280) Perraud, O.; Robert, V.; Gornitzka, H.; Martinez, A.; Dutasta, J. Effect between Hydrogen-Bonding and Cation-Molecule Interactions
Combined Cation−π and Anion−π Interactions for Zwitterion
and Thermodynamic Property in the 1:2 (Na+:N-(Hydroxymethyl)-
Recognition. Angew. Chem., Int. Ed. 2012, 51, 504−508.
acetamide) Ternary Complex. J. Mol. Model. 2014, 20, 2154.
(281) Li, R.; Li, Q.; Cheng, J.; Liu, Z.; Li, W. The Prominent Enhancing
(300) Zhang, D.; Chatelet, B.; Serrano, E.; Perraud, O.; Dutasta, J.;
Effect of the Cation−π Interaction on the Halogen−Hydride Halogen
Bond in M1···C6H5X···HM2. ChemPhysChem 2011, 12, 2289−2295. Robert, V.; Martinez, A. Insights into the Complexity of Weak
(282) Li, Q.; Li, W.; Cheng, J.; Gong, B.; Sun, J. Effect of Methyl Group Intermolecular Interactions Interfering in Host−Guest Systems.
on the Cooperativity between Cation−π Interaction and NH···O ChemPhysChem 2015, 16, 2931−2935.
Hydrogen Bonding. J. Mol. Struct.: THEOCHEM 2008, 867, 107−110. (301) Li, Y.; Pink, M.; Karty, J. A.; Flood, A. H. Dipole-Promoted and
(283) Li, Q.; Li, R.; Liu, X.; Cheng, J.; Li, W. Ab initio Study of Size-Dependent Cooperativity between Pyridyl-Containing Triazolo-
Synergetic Effects of Two Strong Interactions of Cation−π Interaction phanes and Halides Leads to Persistent Sandwich Complexes with
and Lithium Bond in M+ ··· Phenyl Lithium ··· N (M = Li, Na, K; N = Iodide. J. Am. Chem. Soc. 2008, 130, 17293−17295.
H2O and NH3) Complex. Mol. Phys. 2012, 110, 457−465. (302) Hall, B. R.; Manck, L. E.; Tidmarsh, I. S.; Stephenson, A.; Taylor,
(284) Lu, Y.; Liu, Y.; Li, H.; Zhu, X.; Liu, H.; Zhu, W. Mutual Influence B. F.; Blaikie, E. J.; Griend, D. A. V.; Ward, M. D. Structures, Host−
between Halogen Bonds and Cation−π Interactions: A Theoretical Guest Chemistry and Mechanism of Stepwise Self-Assembly of M4L6
Study. ChemPhysChem 2012, 13, 2154−2161. Tetrahedral Cage Complexes. Dalton Trans. 2011, 40, 12132−12145.
(285) Duan, M.; Song, B.; Shi, G.; Li, H.; Ji, G.; Hu, J.; Chen, X.; Fang, (303) Quiñonero, D.; Deyà, P. M.; Carranza, M. P.; Rodríguez, A. M.;
H. Cation⊗3π: Cooperative Interaction of a Cation and Three Jalón, F. A.; Manzano, B. R. Experimental and Computational Study of
Benzenes with an Anomalous Order in Binding Energy. J. Am. Chem. the Interplay between C−H/π and Anion−π Interactions. Dalton Trans.
Soc. 2012, 134, 12104−12109. 2010, 39, 794−806.
(286) Estarellas, C.; Frontera, A.; Quiñonero, D.; Deyà, P. M. Interplay (304) Notash, B.; Safari, N.; Khavasi, H. R. Anion-Controlled
between Cation−π and Hydrogen Bonding Interactions: Are Non- Structural Motif in One-Dimensional Coordination Networks via
Additivity Effects Additive? Chem. Phys. Lett. 2009, 479, 316−320. Cooperative Weak Noncovalent Interactions. CrystEngComm 2012, 14,
(287) Estarellas, C.; Frontera, A.; Quiñonero, D.; Deyà, P. M. Can 6788−6796.
Lone Pair-π and Cation-π Interactions Coexist? A Theoretical Study. (305) Arcus, V.; Bernstein, D.; Crombie, C.; Saunders, G. Infinite
Cent. Eur. J. Chem. 2011, 9, 25−34. Stacking of Alternating Polyfluoroaryl Rings and Bromide Anions.
(288) Yan, Y.; Shi, W.; Feng, G.; Ren, F.; Wang, Y. A B3LYP and CrystEngComm 2013, 15, 9841−9843.
MP2(full) Theoretical Investigation on the Cooperativity Effect (306) Hay, B. P.; Bryantsev, V. S. Anion−Arene Adducts: C−H
between Cation−Molecule and Hydrogen-Bonding Interactions in the Hydrogen Bonding, Anion−π Interaction, and Carbon Bonding Motifs.
O-Cresol Complex with Na+. Comput. Theor. Chem. 2012, 996, 91−102. Chem. Commun. 2008, 2417−2428.
(289) Li, B.; Shi, W.; Ren, F. A B3LYP and MP2 Theoretical (307) Manzano, B. R.; Jalón, F. A.; Ortiz, I. M.; Soriano, M. L.; de la
Investigation on the Cooperativity Effect between the XH···HM Torre, F. G.; Elguero, J.; Maestro, M. A.; Mereiter, K.; Claridge, T. D. W.
(X = F, Cl, Br; M = Li, Na, K) Dihydrogen-Bonding and HM···π Self-Assembly of Ligands Designed for the Building of a New Type of [2
Interactions Involving C6H6. Comput. Theor. Chem. 2013, 1020, 81−90. × 2] Metallic Grid. Anion Encapsulation and Diffusion NMR
(290) Zeng, Y.; Wu, W.; Li, X.; Zheng, S.; Meng, L. Influence of the
Spectroscopy. Inorg. Chem. 2008, 47, 413−428.
Li···π Interaction on the H/X···π Interactions in HOLi···C6H6···HOX/
(308) Garcia-Raso, A.; Albertí, F. M.; Fiol, J. J.; Tasada, A.; Barceló-
XOH (X = F, Cl, Br, I) Complexes. ChemPhysChem 2013, 14, 1591−
Oliver, M.; Molins, E.; Escudero, D.; Frontera, A.; Quiñonero, D.; Deyà,
1600.
(291) Ebrahimi, A.; Masoodi, H. R.; Khorassani, M. H.; Ghaleno, M. H. P. M. Anion-π Interactions in Bisadenine Derivatives: A Combined
The Influence of Cation-π and Anion-π Interactions on the Strength and Crystallographic and Theoretical Study. Inorg. Chem. 2007, 46, 10724−
Nature of N···H Hydrogen Bond. Comput. Theor. Chem. 2012, 988, 48− 10735.
55. (309) Berryman, O. B.; Johnson, D. W. Experimental Evidence for
(292) Esrafili, M. D.; Esmailpour, P.; Mohammadian-Sabet, F.; Interactions between Anions and Electron-Deficient Aromatic Rings.
Solimannejad, M. Theoretical Study of the Interplay between Halogen Chem. Commun. 2009, 3143−3153.
Bond and Lithium−π Interactions: Cooperative and Diminutive Effects. (310) Berryman, O. B.; Hof, F.; Hynes, M. J.; Johnson, D. W. Anion−π
Chem. Phys. Lett. 2013, 588, 47−50. Interaction Augments Halide Binding in Solution. Chem. Commun.
(293) Frontera, A.; Quiñonero, D.; Garau, C.; Costa, A.; Ballester, P.; 2006, 506−508.
Deyà, P. M. MP2 Study of Cation−(π)n−π Interactions (n = 1−4). J. (311) Jentzsch, A. V.; Emery, D.; Mareda, J.; Metrangolo, P.; Resnati,
Phys. Chem. A 2006, 110, 9307−9309. G.; Matile, S. Ditopic Ion Transport Systems: Anion−π Interactions and
(294) Frontera, A.; Quiñonero, D.; Costa, A.; Ballester, P.; Deyá, P. M. Halogen Bonds at Work. Angew. Chem., Int. Ed. 2011, 50, 11675−11678.
MP2 Study of Cooperative Effects between Cation−π, Anion−π and (312) Giese, M.; Albrecht, M.; Krappitz, T.; Peters, M.; Gossen, V.;
π−π interactions. New J. Chem. 2007, 31, 556−560. Raabe, G.; Valkonen, A.; Rissanen, K. Cooperativity of H-bonding and
(295) Vasudevan, D.; Arey, T. A.; Dickstein, D. R.; Newman, M. H.; Anion−π Interaction in the Binding of Anions with Neutral π-
Zhang, T. Y.; Kinnear, H. M.; Bader, M. M. Nonlinearity of Cationic Acceptors. Chem. Commun. 2012, 48, 9983−9985.

2813 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(313) Zaccheddu, M.; Filippi, C.; Buda, F. Anion-π and π-π (334) Solimannejad, M.; Malekani, M.; Alkorta, I. Cooperative and
Cooperative Interactions Regulating the Self-Assembly of Nitrate- Diminutive Unusual Weak Bonding In F3CX···HMgH···Y and F3CX···
Triazine-Triazine Complexes. J. Phys. Chem. A 2008, 112, 1627−1632. Y···HMgH Trimers (X = Cl, Br; Y = HCN, and HNC). J. Phys. Chem. A
(314) Gural’skiy, I. A.; Escudero, D.; Frontera, A.; Solntsev, P. V.; 2010, 114, 12106−12111.
Rusanov, E. B.; Chernega, A. N.; Krautscheid, H.; Domasevitch, K. V. (335) Zhao, Q.; Feng, D.; Hao, J. The Cooperativity between
1,2,4,5-Tetrazine: An Unprecedented μ4-Coordination that Enhances Hydrogen and Halogen Bond in the XY···HNC···XY (X, Y = F, Cl, Br)
Ability for Anion···π Interactions. Dalton Trans. 2009, 2856−2864. complexes. J. Mol. Model. 2011, 17, 2817−2823.
(315) Quiñonero, D.; Frontera, A.; Garau, C.; Ballester, P.; Costa, A.; (336) McDowell, S. A. C.; Joseph, J. A. Cooperative Effects of
Deyá, P. M. Interplay Between Cation−π, Anion− π and π−π Noncovalent Bonds to the Br Atom of Halogen-Bonded H3N···BrZ and
Interactions. ChemPhysChem 2006, 7, 2487−2491. HCN···BrZ (Z = F, Br) Complexes. J. Chem. Phys. 2012, 137, 074310−8.
(316) Lucas, X.; Estarellas, C.; Escudero, D.; Frontera, A.; Quiñonero, (337) McDowell, S. A. C.; Joseph, J. A. Cooperative Effects in Novel
D.; Deyà, P. M. Very Long-Range Effects: Cooperativity between LiF/HFLiFXF (X = F, Cl, Br) Clusters. J. Chem. Phys. 2013, 138,
Anion−π and Hydrogen-Bonding Interactions. ChemPhysChem 2009, 164313−5.
10, 2256−2264. (338) Wu, W.; Lu, Y.; Liu, Y.; Li, H.; Peng, C.; Liu, H.; Zhu, W. Weak
(317) Estarellas, C.; Frontera, A.; Quiñonero, D.; Deyà, P. M. Energetic Effects between X−π and X−N Halogen Bonds: CSD Search
Theoretical Study on Cooperativity Effects between Anion−π and and Theoretical Study. Chem. Phys. Lett. 2013, 582, 49−55.
Halogen-Bonding Interactions. ChemPhysChem 2011, 12, 2742−2750. (339) Mitra, M.; Manna, P.; Seth, S. K.; Das, A.; Meredith, J.; Helliwell,
(318) Estarellas, C.; Frontera, A.; Quiñonero, D.; Deyà, P. M. M.; Bauzá, A.; Choudhury, S. R.; Frontera, A.; Mukhopadhyay, S. Salt-
Unexpected Nonadditivity Effects in Anionπ Complexes. J. Phys. bridge−π (sb−π) Interactions at Work: Associative Interactions of
Chem. A 2011, 115, 7849−7857. sb−π, π−π and Anion−π in Cu(II)-malonate−2-aminopyridine−
(319) Estarellas, C.; Frontera, A.; Quiñonero, D.; Alkorta, I.; Deyà, P. hexafluoridophosphate Ternary System. CrystEngComm 2013, 15,
M.; Elguero, J. Energetic vs Synergetic Stability: A Theoretical Study. J. 686−696.
Phys. Chem. A 2009, 113, 3266−3273. (340) Albrecht, L.; Boyd, R. J.; Mó, O.; Yáñez, M. Changing Weak
(320) Alkorta, I.; Blanco, F.; Elguero, J. Simultaneous Interaction of Halogen Bonds into Strong Ones through Cooperativity with Beryllium
Tetrafluoroethene with Anions and Hydrogen-Bond Donors: A Bonds. J. Phys. Chem. A 2014, 118, 4205−4213.
Cooperativity Study. J. Chem. Theory Comput. 2009, 5, 1186−1194. (341) An, X.; Li, R.; Li, Q.; Liu, X.; Li, W.; Cheng, J. Substitution,
(321) Alkorta, I.; Blanco, F.; Deyá, P. M.; Elguero, J.; Estarellas, C.; Cooperative, and Solvent Effects on π Pnicogen Bonds in the FH2P and
Frontera, A.; Quiñonero, D. Cooperativity in Multiple Unusual Weak FH2As Complexes. J. Mol. Model. 2012, 18, 4325−4332.
Bonds. Theor. Chem. Acc. 2010, 126, 1−14. (342) Li, Q.; Sun, L.; Liu, X.; Li, W.; Cheng, J.; Zeng, Y. Enhancement
(322) Escudero, D.; Frontera, A.; Quiñonero, D.; Deyà, P. M. Interplay of Iodine−Hydride Interaction by Substitution and Cooperative Effects
between Anion-π and Hydrogen Bonding Interactions. J. Comput. Chem.
in NCX−NCI−HMY Complexes. ChemPhysChem 2012, 13, 3997−
2009, 30, 75−82.
4002.
(323) Kumbhar, S.; Jana, S.; Anoop, A.; Waller, M. P. Cooperativity in
(343) Suresh, C. H.; Mohan, N.; Vijayalakshmi, K. P.; George, R.;
Bimetallic Glutathione Complexes. J. Mol. Graphics Modell. 2015, 62, 1−
Mathew, J. M. Typical Aromatic Noncovalent Interactions in Proteins:
10.
A Theoretical Study Using Phenylalanine. J. Comput. Chem. 2009, 30,
(324) Zeng, Y.; Hao, J.; Zheng, S.; Meng, L. Cooperativity between
S···π and Rg···π in the OCS···C6H6···Rg (Rg = He, Ne, Ar, and Kr) van 1392−1404.
(344) Podgornik, R.; French, R. H.; Parsegian, V. A. Nonadditivity in
der Waals Complexes. J. Phys. Chem. A 2011, 115, 11057−11066.
(325) Hua, S.; Xu, L.; Li, W.; Li, S. Cooperativity in Long α- and 310- van der Waals Interactions within Multilayers. J. Chem. Phys. 2006, 124,
Helical Polyalanines: Both Electrostatic and van der Waals Interactions 044709.
Are Essential. J. Phys. Chem. B 2011, 115, 11462−11469. (345) Alkorta, I.; Elguero, J.; Yáñez, M.; Mó, O. Cooperativity in
(326) Mück-Lichtenfeld, C.; Grimme, S. Theoretical Analysis of Beryllium Bonds. Phys. Chem. Chem. Phys. 2014, 16, 4305−4312.
Cooperative Effects of Small Molecule Activation by Frustrated Lewis (346) Xu, T.; Chen, E. Y. -X. Probing Site Cooperativity of Frustrated
Pairs. Dalton Trans. 2012, 41, 9111−9118. Phosphine/Borane Lewis Pairs by a Polymerization Study. J. Am. Chem.
(327) Solimannejad, M.; Ghafari, S.; Esrafili, M. D. Theoretical Insight Soc. 2014, 136, 1774−1777.
into Cooperativity in Lithium-Bonded Complexes: Linear Clusters of (347) Malenov, D. P.; Janjić, G. V.; Veljković, D. Ž .; Zarić, S. D. Mutual
LiCN and LiNC. Chem. Phys. Lett. 2013, 577, 6−10. Influence of Parallel, CH/O, OH/π and Lone Pair/π Interactions in
(328) Alkorta, I.; Blanco, F.; Elguero, J. A Computational Study of the Water/Benzene/Water System. Comput. Theor. Chem. 2013, 1018, 59−
Cooperativity in Clusters of Interhalogen Derivatives. Struct. Chem. 65.
2009, 20, 63−71. (348) Wang, Y.; Wu, W.; Liu, Y.; Lu, Y. Influence of Transition Metal
(329) Ghoshal, D.; Maji, T. K. Fabrication of Supramolecular Coordination on Halogen Bonding: CSD Survey and Theoretical Study.
Frameworks by Tuning the Binding Site of a Tripodal Ligand with d10 Chem. Phys. Lett. 2013, 578, 38−42.
Metal Ions: Interplay of Covalent and Noncovalent Interactions in (349) Campo-Cacharrón, A.; Cabaleiro-Lago, E. M.; Carrazana-
Solid-State Structure. J. Chem. Sci. 2010, 122, 801−806. García, J. A.; Rodríguez-Otero, J. Interaction of Aromatic Units of
(330) Lv, H.; Zhuo, H.; Li, Q.; Yang, X.; Li, W.; Cheng, J. Mutual Amino Acids with Guanidinium Cation: The Interplay of π···π, X
Influence between Covalent and Noncovalent Interactions in H3N− H···π, and M+···π Contacts. J. Comput. Chem. 2014, 35, 1290−1301.
MCN−XF (X = H, Li, Cl, Br; M = Ag, Cu, Au). Mol. Phys. 2014, 112, (350) Yu, G.; Hua, B.; Han, C. Proton Transfer in Host−Guest
1081−1088. Complexation between a Difunctional Pillar[5]arene and Alkyldi-
(331) Esrafili, M. D.; Esmailpour, P.; Mohammadian-Sabet, F.; amines. Org. Lett. 2014, 16, 2486−2489.
Solimannejad, M. Substituent Effects on Cooperativity between Lithium (351) George, J.; Deringer, V. L.; Dronskowski, R. Cooperativity of
Bonds. Int. J. Quantum Chem. 2014, 114, 295−301. Halogen, Chalcogen, and Pnictogen Bonds in Infinite Molecular Chains
(332) Esrafili, M. D.; Fatehi, P.; Solimannejad, M. Mutual Influence by Electronic Structure Theory. J. Phys. Chem. A 2014, 118, 3193−3200.
between Conventional and Unconventional Lithium Bonds. J. Mol. (352) Gunning, P. T. Positive Ion Pair Cooperativity Exhibited for the
Graphics Modell. 2014, 49, 129−137. Binding of Phosphate under Physiological Conditions. Org. Biomol.
(333) Esrafili, M. D.; Vakili, M. Halogen Bonds Enhanced by σ-hole Chem. 2005, 3, 3877−3879.
and π-hole Interactions: A Comparative Study on Cooperativity and (353) Burress, C. N.; Bodine, M. I.; Elbjeirami, O.; Reibenspies, J. H.;
Competition Effects between X···N and S···N Interactions in H3N··· Omary, M. A.; Gabbaï, F. P. Enhancement of External Spin-Orbit
XCN···SF2 and H3N···XCN···SO2 Complexes (X = F, Cl, Br and I). J. Coupling Effects Caused by Metal-Metal Cooperativity. Inorg. Chem.
Mol. Model. 2014, 20, 1−9. 2007, 46, 1388−1395.

2814 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(354) Kuriakose, N.; Kadam, S.; Vanka, K. A Theoretical Study of (374) Dey, S. K.; Das, G. Selective inclusion of PO43− within Persistent
Metal−Metal Cooperativity in the Homogeneous Water Gas Shift Dimeric Capsules of a Tris(thiourea) Receptor and Evidence of Cation/
Reaction. Inorg. Chem. 2012, 51, 377−385. Solvent Sealed Unimolecular Capsules. Dalton Trans. 2012, 41, 8960−
(355) Medel, R.; Heger, M.; Suhm, M. A. Molecular Docking via 8972.
Olefinic OH···π Interactions: A Bulky Alkene Model System and Its (375) Zaman, I. G.; Delibas, N. C.; Necefoğlu, H.; Hökelek, T. trans-
Cooperativity. J. Phys. Chem. A 2015, 119, 1723−1730. Tetraaquabis(isonicotinamide-κN1)zinc bis(3-hydroxybenzoate) tetra-
(356) Marín-Luna, M.; Alkorta, I.; Elguero, J.; Mó, O.; Yáñez, M. hydrate. Acta Crystallogr., Sect. E: Struct. Rep. Online 2013, 69, m198−
Interplay between Beryllium Bonds and Anion-π Interactions in m199.
BeR2:C6X6:Y− Complexes (R = H, F and Cl, X = H and F, and Y = (376) Al-Mawsawi, L. Q.; Hombrouck, A.; Dayam, R.; Debyser, Z.;
Cl and Br). Molecules 2015, 20, 9961−9976. Neamati, N. Four-Tiered π Interaction at the Dimeric Interface of HIV-
(357) Guo, X.; An, X.; Li, Q. Se···N Chalcogen Bond and Se···X 1 Integrase Critical for DNA Integration and Viral Infectivity. Virology
Halogen Bond Involving F2CSe: Influence of Hybridization, 2008, 377, 355−363.
Substitution, and Cooperativity. J. Phys. Chem. A 2015, 119, 3518−3527. (377) Degtyarenko, A. S.; Domasevitch, K. V. Cadmium(II) Iodide
(358) Gao, M.; Cheng, J.; Yang, X.; Li, W.; Xiao, B.; Li, Q. Influence of and Thiocyanate Complexes Adopted by Polycyclic 1,4-bis(pyridazin-4-
Substituents on the Nature of Metal···π Interaction and its yl)benzene: Interplay of Coordination and π−π Stacking Interactions.
Cooperativity with Halogen Bond. J. Chem. Phys. 2015, 143, 054308−7. Acta Crystallogr., Sect. C: Cryst. Struct. Commun. 2013, 69, 219−224.
(359) Ghafari, S.; Gholipour, A. Simultaneous Interactions of (378) Walter, S. M.; Sarwar, M. G.; Chudzinski, M. G.; Herdtweck, E.;
Pyrimidine Ring with BeF2 and BF3 in BeF2···X−Pyr···BF3 Complexes: Lough, A. J.; Huber, S. M.; Taylor, M. S. Halogen Bonding and π−π
Non-cooperativity. J. Mol. Model. 2015, 21, 253. Interactions in the Solid-State Structure of a Butadiynylene-Linked
(360) Marín-Luna, M.; Alkorta, I.; Elguero, J.; Mó, O.; Yáñez, M. Bis(iodoperfluoroarene). CrystEngComm 2013, 15, 3097−3101.
Simultaneous Aromatic−Beryllium Bonds and Aromatic−Anion (379) Shishkin, O. V.; Medvediev, V. V.; Zubatyuk, R. I. Supra-
Interactions: Naphthalene and Pyrene as Models of Fullerenes, Carbon molecular Architecture of Molecular Crystals Possessing Shearing
Single-Walled Nanotubes, and Graphene. ChemPhysChem 2015, 16, Mechanical Properties: Columns versus Layers. CrystEngComm 2013,
2680−2686. 15, 160−167.
(361) Varadwaj, P. R.; Varadwaj, A.; Jin, B. Unusual Bonding Modes of (380) Chen, J.; Lai, Y.; Wan, D.; Jin, M.; Pu, H. Cooperative
Perfluorobenzene in its Polymeric (Dimeric, Trimeric and Tetrameric) Entrapment of Xanthene Dyes by a Core-Engineered Unimolecular
Forms: Entirely Negative Fluorine Interacting Cooperatively with Micelle. Macromol. Chem. Phys. 2013, 214, 1817−1828.
Entirely Negative Fluorine. Phys. Chem. Chem. Phys. 2015, 17, 31624. (381) Eryazici, I.; Moorefield, C. N.; Durmus, S.; Newkome, G. R.
(362) Fokin, A. A.; Gerbig, D.; Schreiner, P. R. σ/σ- and π/π- Synthesis and Single-Crystal X-ray Characterization of 4,4″-Function-
Interactions Are Equally Important: Multilayered Graphanes. J. Am.
alized 4′-(4-Bromophenyl)-2,2′:6′,2″-terpyridines. J. Org. Chem. 2006,
Chem. Soc. 2011, 133, 20036−20039.
71, 1009−1014.
(363) Schreiner, P. R.; Chernish, L. V.; Gunchenko, P. A.; Tikhonchuk,
(382) Britton, D. o-Chloro- and o-bromobenzo-nitrile: Pseudosym-
E. Y.; Hausmann, H.; Serafin, M.; Schlecht, S.; Dahl, J. E. P.; Carlson, R.
metry and Pseudo-Isostructural Packing. Acta Crystallogr., Sect. C: Cryst.
M. K.; Fokin, A. A. Overcoming Lability of Extremely Long Alkane
Struct. Commun. 2007, 63, o14−o16.
Carbon−Carbon Bonds through Dispersion Forces. Nature 2011, 477,
(383) Vijay, D.; Sakurai, H.; Sastry, G. N. The Impact of Basis Set
308−312.
Superposition Error on the Structure of π-π Dimers. Int. J. Quantum
(364) Alonso, M.; Woller, T.; Martin-Martinez, F. J.; Contreras-García,
Chem. 2011, 111, 1893−1901.
J.; Geerlings, P.; De Proft, F. Understanding the Fundamental Role of π/
(384) Mignon, P.; Loverix, S.; De Proft, F.; Geerlings, P. Influence of
π, σ/σ, and σ/π Dispersion Interactions in Shaping Carbon-Based
Stacking on Hydrogen Bonding: Quantum Chemical Study on Pyridine-
Materials. Chem. - Eur. J. 2014, 20, 4931−4941.
(365) Kemnitz, C. R.; Mackey, J. L.; Loewen, M. J.; Hargrove, J. L.; Benzene Model Complexes. J. Phys. Chem. A 2004, 108, 6038−6044.
(385) Gray, M.; Goodman, A. J.; Carroll, J. B.; Bardon, K.; Markey, M.;
Lewis, J. L.; Hawkins, W. E.; Nielsen, A. F. Origin of Stability in
Branched Alkanes. Chem. - Eur. J. 2010, 16, 6942−6949. Cooke, G.; Rotello, V. M. Model Systems for Flavoenzyme Activity:
(366) Wodrich, M. D.; Wannere, C. S.; Mo, Y.; Jarowski, P. D.; Houk, Interplay of Hydrogen Bonding and Aromatic Stacking in Cofactor
K. N.; von Rague Schleyer, P. The Concept of Protobranching and Its Redox Modulation. Org. Lett. 2004, 6, 385−388.
Many Paradigm Shifting Implications for Energy Evaluations. Chem. - (386) Hesselmann, A.; Jansen, G.; Schütz, M. Interaction Energy
Eur. J. 2007, 13, 7731−7744. Contributions of H-Bonded and Stacked Structures of the AT and GC
(367) Janowski, T.; Pulay, P. A Benchmark Comparison of σ/σ and π/ DNA Base Pairs from the Combined Density Functional Theory and
π Dispersion: The Dimers of Naphthalene and Decalin, and Coronene Intermolecular Perturbation Theory Approach. J. Am. Chem. Soc. 2006,
and Perhydrocoronene. J. Am. Chem. Soc. 2012, 134, 17520−17525. 128, 11730−11731.
(368) Gonthier, J. F.; Wodrich, M. D.; Steinmann, S. N.; Corminboeuf, (387) Santos, J.; Grimm, B.; Illescas, B. M.; Guldi, D. M.; Martín, N.
C. Branched Alkanes Have Contrasting Stabilities. Org. Lett. 2010, 12, Cooperativity between π-π and H-Bonding Interactions−A Supra-
3070−3073. molecular Complex Formed by C60 and exTTF. Chem. Commun. 2008,
(369) Jiemchooroj, A.; Sernelius, B. E.; Norman, P. C6 Dipole-Dipole 5993−5995.
Dispersion Coefficients for the n-Alkanes: Test of an Additivity (388) Escudero, D.; Frontera, A.; Quiñonero, D.; Deya, P. M. Interplay
Procedure. Phys. Rev. A: At., Mol., Opt. Phys. 2004, 69, 044701. between Edge-to-Face Aromatic and Hydrogen-Bonding Interactions. J.
(370) Hatano, B.; Aikawa, A.; Tagaya, H.; Takahashi, H. New Phys. Chem. A 2008, 112, 6017−6022.
Approach for the Recovery of Bisphenol A from Water Using Inclusion (389) Quiñonero, D.; Frontera, A.; Escudero, D.; Ballester, P.; Costa,
Complex with Quinoline Derivatives. Chem. Lett. 2004, 33, 1276−1277. A.; Deyà, P. M. MP2 Study of Synergistic Effects between X−H/π (X =
(371) Bourne, S. A.; Corin, K. C.; Nassimbeni, L. R.; Toda, F. Selective C,N,O) and π−π Interactions. Theor. Chem. Acc. 2008, 120, 385−393.
Enclathration of Picolines. Cryst. Growth Des. 2005, 5, 379−382. (390) Estarellas, C.; Escudero, D.; Frontera, A.; Quiñonero, D.; Deya,
(372) Stoll, I.; Brodbeck, R.; Neumann, B.; Stammler, H.; Mattay, J. P. M. Theoretical Ab Initio Study of the Interplay between Hydrogen
Controlling the Self Assembly of Arene Functionalised 2-amino- Bonding, Cation−π and π−π Interactions. Theor. Chem. Acc. 2009, 122,
pyrimidines by arene-perfluoroarene Interaction and by Silver(I) 325−332.
Complex Formation. CrystEngComm 2009, 11, 306−317. (391) Ebrahimi, A.; Habibi, M.; Neyband, R. S.; Gholipour, A. R.
(373) Fu, X.; Li, J.; Simpson, J. Non-Covalent Interactions in the Cooperativity of π-Stacking and Hydrogen Bonding Interactions and
Crystal Structure of Methyl 4-Hydroxy-3-Nitrobenzoate. Crystals 2012, Substituent Effects on X-ben∥pyr···H−F Complexes. Phys. Chem. Chem.
2, 669−674. Phys. 2009, 11, 11424−11431.

2815 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(392) Kulkarni, C.; Reddy, S. K.; George, S. J.; Balasubramanian, S. (412) Meyer, M.; Albrecht-Gary, A.; Dietrich-Buchecker, C. O.;
Cooperativity in the Stacking of Benzene-1, 3, 5-Tricarboxamide: The Sauvage, J. π-π Stacking-Induced Cooperativity in Copper(I)
Role of Dispersion. Chem. Phys. Lett. 2011, 515, 226−230. Complexes with Phenanthroline Ligands. Inorg. Chem. 1999, 38,
(393) Ninković, D. B.; Janjić, G. V.; Zarić, S. D. Crystallographic and 2279−2287.
Ab Initio Study of Pyridine Stacking Interactions. Local Nature of (413) Engkvist, O.; Hobza, P.; Selzle, H. L.; Schlag, E. W. Benzene
Hydrogen Bond Effect in Stacking Interactions. Cryst. Growth Des. 2012, Trimer and Benzene Tetramer: Structures and Properties Determined
12, 1060−1063. by the Nonempirical Model (NEMO) Potential Calibrated from the
(394) Li, H.; Lu, Y.; Liu, Y.; Zhu, X.; Liu, H.; Zhu, W. Interplay CCSD(T) Benzene Dimer Energies. J. Chem. Phys. 1999, 110, 5758−
between Halogen Bonds and π−π Stacking Interactions: CSD Search 5762.
and Theoretical Study. Phys. Chem. Chem. Phys. 2012, 14, 9948−9955. (414) Iimori, T.; Ohshima, Y. S1−S0 Vibronic Spectra of Benzene
(395) Muzangwa, L.; Nyambo, S.; Uhler, B.; Reid, S. A. On π-Stacking, Clusters Revisited. I. The Tetramer. J. Chem. Phys. 2002, 117, 3656−
C-H/π, and Halogen Bonding Interactions in Halobenzene Clusters: 3674.
Resonant Two-Photon Ionization Studies of Chlorobenzene. J. Chem. (415) Wolfenden, R.; Liang, Y.; Matthews, M.; Williams, R.
Phys. 2012, 137, 184307−9. Cooperativity and Anticooperativity in Solvation by Water: Imidazoles,
(396) Tam, A. Y.; Wong, K. M.; Wang, G.; Yam, V. W. Luminescent Quinones, Nitrophenols, Nitrophenolate, and Nitrothiophenolate Ions.
Metallogels of Platinum(II) Terpyridyl Complexes: Interplay of J. Am. Chem. Soc. 1987, 109, 463−467.
Metal···Metal, π−π and Hydrophobic−Hydrophobic Interactions on (416) Lucht, B. L.; Collum, D. B. Lithium Ion Solvation: Amine and
Gel Formation. Chem. Commun. 2007, 20, 2028−2030. Unsaturated Hydrocarbon Solvates of Lithium Hexamethyldisilazide
(397) Tsuzuki, S.; Honda, K.; Uchimaru, T.; Mikami, M.; Tanabe, K. (LiHMDS). J. Am. Chem. Soc. 1996, 118, 2217−2225.
Origin of Attraction and Directionality of the π/π Interaction: Model (417) Czaplewski, C.; Ripoll, D. R.; Liwo, A.; Rodziewicz-Motowidło,
Chemistry Calculations of Benzene Dimer Interaction. J. Am. Chem. Soc. S.; Wawak, R. J.; Scheraga, H. A. Can Cooperativity in Hydrophobic
2002, 124, 104−112. Association Be Reproduced Correctly by Implicit Solvation Models? Int.
(398) Sinnokrot, M. O.; Sherrill, C. D. High-Accuracy Quantum J. Quantum Chem. 2002, 88, 41−55.
Mechanical Studies of π-π Interactions in Benzene Dimers. J. Phys. (418) Robertson, W. H.; Karapetian, K.; Ayotte, P.; Jordan, K. D.;
Chem. A 2006, 110, 10656−10668. Johnson, M. A. Infrared Predissociation Spectroscopy of I−.(CH3OH)n,
(399) Tauer, T. P.; Sherrill, C. D. Beyond the Benzene Dimer: An n = 1,2: Cooperativity in Asymmetric Solvation. J. Chem. Phys. 2002,
Investigation of the Additivity of π-π Interactions. J. Phys. Chem. A 2005, 116, 4853−4857.
109, 10475−10478. (419) Kar, T.; Scheiner, S. Cooperativity of Conventional and
(400) Gonzalez, C.; Lim, E. C. Ab Initio Study of the Intermolecular Unconventional Hydrogen Bonds Involving Imidazole. Int. J. Quantum
Interactions in Small Benzene Clusters: The Equilibrium Structures of Chem. 2006, 106, 843−851.
(420) Tarbuck, T. L.; Ota, S. T.; Richmond, G. L. Spectroscopic
Trimer, Tetramer, and Pentamer. J. Phys. Chem. A 2001, 105, 1904−
Studies of Solvated Hydrogen and Hydroxide Ions at Aqueous Surfaces.
1908.
J. Am. Chem. Soc. 2006, 128, 14519−14527.
(401) Gonzalez, C.; Allison, T. C.; Lim, E. C. Hartree-Fock Dispersion
(421) Reddy, A. S.; Zipse, H.; Sastry, G. N. Cation-π Interactions of
Probe of the Equilibrium Structures of Small Microclusters of Benzene
Bare and Coordinatively Saturated Metal Ions: Contrasting Structural
and Naphthalene: Comparison with Second-Order MØeller-Plesset
and Energetic Characteristics. J. Phys. Chem. B 2007, 111, 11546−
Geometries. J. Phys. Chem. A 2001, 105, 10583−10587.
11553.
(402) Ye, X.; Li, Z.; Wang, W.; Fan, K.; Xu, W.; Hua, Z. The Parallel
(422) Rao, J. S.; Zipse, H.; Sastry, G. N. Explicit Solvent Effect on
π−π Stacking: A Model Study with MP2 and DFT Methods. Chem. Phys. Cation-π Interactions: A First Principle Investigation. J. Phys. Chem. B
Lett. 2004, 397, 56−61. 2009, 113, 7225−7236.
(403) Swart, M.; van der Wijst, T.; Fonseca Guerra, C.; Bickelhaupt, F. (423) Rao, J. S.; Dinadayalane, T. C.; Sastry, G. N.; Leszczynski, J.
M. π-π Stacking Tackled with Density Functional Theory. J. Mol. Model. Comprehensive Study on the Solvation of Mono- and Divalent Metal
2007, 13, 1245−1257. Cations: Li+, Na+, K+, Be2+, Mg2+ and Ca2+. J. Phys. Chem. A 2008, 112,
(404) Pullan, W. J. Structure Prediction of Benzene Clusters Using a 12944−12953.
Genetic Algorithm. J. Chem. Inf. Model. 1997, 37, 1189−1193. (424) Neela, Y. I.; Mahadevi, A. S.; Sastry, G. N. First Principles Study
(405) Takeuchi, H. Novel Method for Geometry Optimization of and Database Analyses of Structural Preferences for Sodium ion (Na+)
Molecular Clusters: Application to Benzene Clusters. J. Chem. Inf. Solvation and Coordination. Struct. Chem. 2013, 24, 67−79.
Model. 2007, 47, 104−109. (425) Neela, Y. I.; Mahadevi, A. S.; Sastry, G. N. Analyzing
(406) Mahadevi, A. S.; Rahalkar, A.; Gadre, S. R.; Sastry, G. N. Ab Initio Coordination Preferences of Mg2+ Complexes: Insights from Computa-
Investigation of Benzene Clusters: Molecular Tailoring Approach. J. tional and Database Study. Struct. Chem. 2013, 24, 637−650.
Chem. Phys. 2010, 133, 164308−12. (426) Bing, D.; Hamashima, T.; Fujii, A.; Kuo, J. Anticooperative Effect
(407) Chourasia, M.; Sastry, G. M.; Sastry, G. N. Aromatic−Aromatic Induced by Mixed Solvation in H+(CH3OH)m(H2O)n (m + n = 5 and
Interactions Database, A2ID: An Analysis of Aromatic π-Networks in 6): A Theoretical and Infrared Spectroscopic Study. J. Phys. Chem. A
Proteins. Int. J. Biol. Macromol. 2011, 48, 540−552. 2010, 114, 8170−8177.
(408) Reid, S. A.; Nyambo, S.; Muzangwa, L.; Uhler, B. π-Stacking, C− (427) Tielrooij, K. J.; Garcia-Araez, N.; Bonn, M.; Bakker, H. J.
H/π, and Halogen Bonding Interactions in Bromobenzene and Mixed Cooperativity in Ion Hydration. Science 2010, 328, 1006−1009.
Bromobenzene−Benzene Clusters. J. Phys. Chem. A 2013, 117, 13556− (428) Angelina, E. L.; Peruchena, N. M. Strength and Nature of
13563. Hydrogen Bonding Interactions in Mono- and Di-Hydrated Formamide
(409) Yang, J.; Waller, M. P. A Systematic Approach to Identify Complexes. J. Phys. Chem. A 2011, 115, 4701−4710.
Cooperatively Bound Homotrimers. J. Phys. Chem. A 2013, 117, 174− (429) Gao, Y. Q. Simple Theoretical Model for Ion Cooperativity in
182. Aqueous Solutions of Simple Inorganic Salts and Its Effect on Water
(410) Rutledge, L. R.; Churchill, C. D. M.; Wetmore, S. D. A Surface Tension. J. Phys. Chem. B 2011, 115, 12466−12472.
Preliminary Investigation of the Additivity of π-π or π+-π Stacking and T- (430) Li, Q.; Li, R.; Zhou, Z.; Li, W.; Cheng, J. S···X Halogen Bonds and
Shaped Interactions between Natural or Damaged DNA Nucleobases H···X Hydrogen Bonds in H2CS−XY (XY = FF, ClF, ClCl, BrF, BrCl,
and Histidine. J. Phys. Chem. B 2010, 114, 3355−3367. and BrBr) Complexes: Cooperativity and Solvent Effect. J. Chem. Phys.
(411) Churchill, C. D. M.; Rutledge, L. R.; Wetmore, S. D. Effects of 2012, 136, 014302−8.
the Biological Backbone on Stacking Interactions at DNA−Protein (431) Sharma, B.; Rao, J. S.; Sastry, G. N. Effect of Solvation on Ion
Interfaces: The Interplay between the Backbone··· π and π··· π Binding to Imidazole and Methylimidazole. J. Phys. Chem. A 2011, 115,
Components. Phys. Chem. Chem. Phys. 2010, 12, 14515−14526. 1971−1984.

2816 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(432) Miller, D. J.; Lisy, J. M. Entropic Effects on Hydrated Alkali- (460) Huma, Z. E.; Ludeman, J. P.; Wilkinson, B. L.; Payne, R. J.;
Metal Cations: Infrared Spectroscopy and ab Initio Calculations of Stone, M. J. NMR Characterization of Cooperativity: Fast Ligand
M+(H2O)x=2−5 Cluster Ions for M = Li, Na, K, and Cs. J. Am. Chem. Soc. Binding Coupled to Slow Protein Dimerization. Chem. Sci. 2014, 5,
2008, 130, 15393−15404. 2783−2788.
(433) Prell, J. S.; O’Brien, J. T.; Williams, E. R. Structural and Electric (461) Spassov, V.; Bashford, D. Electrostatic Coupling to pH-Titrating
Field Effects of Ions in Aqueous Nanodrops. J. Am. Chem. Soc. 2011, Sites as a Source of Cooperativity in Protein-Ligand Binding. Protein Sci.
133, 4810−4818. 1998, 7, 2012−2025.
(434) Hishida, M.; Tanaka, K.; Yamamura, Y.; Saito, K. Cooperativity (462) Harris, S. A.; Gavathiotis, E.; Searle, M. S.; Orozco, M.;
between Water and Lipids in Lamellar to Inverted-Hexagonal Phase Laughton, C. A. Cooperativity in Drug-DNA Recognition: A Molecular
Transition. J. Phys. Soc. Jpn. 2014, 83, 044801−8. Dynamics Study. J. Am. Chem. Soc. 2001, 123, 12658−12663.
(435) Whitty, A. Cooperativity and Biological Complexity. Nat. Chem. (463) Jusuf, S.; Loll, P. J.; Axelsen, P. H. Configurational Entropy and
Biol. 2008, 4, 435−439. Cooperativity between Ligand Binding and Dimerization in Glycopep-
(436) Editorial. Capturing Cooperativity. Nat. Chem. Biol. 2008, 4, 433. tide Antibiotics. J. Am. Chem. Soc. 2003, 125, 3988−3994.
(437) Williamson, J. R. Cooperativity in Macromolecular Assembly. (464) Jusuf, S.; Loll, P. J.; Axelsen, P. H. The Role of Configurational
Nat. Chem. Biol. 2008, 4, 458−465. Entropy in Biochemical Cooperativity. J. Am. Chem. Soc. 2002, 124,
(438) Eaton, W. A.; Henry, E. R.; Hofrichter, J.; Mozzarelli, A. Is 3490−3491.
Cooperative Oxygen Binding by Hemoglobin Really Understood? Nat. (465) Kawai, H.; Katoono, R.; Nishimura, K.; Matsuda, S.; Fujiwara,
Struct. Biol. 1999, 6, 351−358. K.; Tsuji, T.; Suzuki, T. Positive Homotropic Allosteric Binding of
(439) Hill, A. V. The Possible Effects of the Aggregation of the Benzenediols in a Hydrindacene-Based Exoditopic Receptor: Cooper-
Molecules of Haemoglobin on its Dissociation Curves. Proc. Physiol. Soc. ativity in Amide Hydrogen Bonding. J. Am. Chem. Soc. 2004, 126, 5034−
1910, 40, iv−vii. 5035.
(440) Haber, J. E.; Koshland, D. E., Jr. Relation of Protein Subunit (466) Schmuck, C.; Geiger, L. Efficient Complexation of N-Acetyl
Interactions to the Molecular Species Observed During Cooperative Amino Acid Carboxylates in Water by an Artificial Receptor:
Binding Of Ligands. Proc. Natl. Acad. Sci. U. S. A. 1967, 58, 2087−2093. Unexpected Cooperativity in the Binding of Glutamate but Not
(441) Ogata, R. T.; McConnell, H. M. Mechanism of Cooperative Aspartate. J. Am. Chem. Soc. 2005, 127, 10486−10487.
Oxygen Binding to Hemoglobin. Proc. Natl. Acad. Sci. U. S. A. 1972, 69, (467) Liu, T.; Whitten, S. T.; Hilser, V. J. Functional Residues Serve a
335−339. Dominant Role in Mediating the Cooperativity of the Protein Ensemble.
(442) Lee, A. W.; Karplus, M. Structure-Specific Model of Hemoglobin Proc. Natl. Acad. Sci. U. S. A. 2007, 104, 4347−4352.
Cooperativity. Proc. Natl. Acad. Sci. U. S. A. 1983, 80, 7055−7059. (468) Rieth, S.; Miner, M. R.; Chang, C. M.; Hurlocker, B.;
(443) Perutz, M. F. Mechanisms of cooperativity and allosteric Braunschweig, A. B. Saccharide Receptor Achieves Concentration
regulation in proteins. Q. Rev. Biophys. 1989, 22, 139−236. Dependent Mannoside Selectivity through Two Distinct Cooperative
(444) Monod, J.; Wyman, J.; Changeux, J. P. On the Nature of Binding Pathways. Chem. Sci. 2013, 4, 357−367.
Allosteric Transitions: A Plausible Model. J. Mol. Biol. 1965, 12, 88−118. (469) Goodey, N. M.; Benkovic, S. J. Allosteric Regulation and
(445) Perutz, M. F. Stereochemistry of Cooperative Effects in Catalysis Emerge via a Common Route. Nat. Chem. Biol. 2008, 4, 474−
Haemoglobin. Nature 1970, 228, 726−734. 482.
(446) Szabo, A.; Karplus, M. A Mathematical Model for Structure- (470) Masterson, L. R.; Mascioni, A.; Traaseth, N. J.; Taylor, S. S.;
Function Relations in Hemoglobin. J. Mol. Biol. 1972, 72, 163−197. Veglia, G. Allosteric Cooperativity in Protein Kinase A. Proc. Natl. Acad.
(447) Ackers, G. K.; Doyle, M. L.; Myers, D.; Daugherty, M. A. Sci. U. S. A. 2008, 105, 506−511.
Molecular Code for Cooperativity in Hemoglobin. Science 1992, 255, (471) Li, G.; Srivastava, A. K.; Kim, J.; Taylor, S. S.; Veglia, G. Mapping
54−63. the Hydrogen Bond Networks in the Catalytic Subunit of Protein Kinase
(448) Changeux, J.; Thiéry, J.; Tung, Y.; Kittel, C. On the A using H/D Fractionation Factors. Biochemistry 2015, 54, 4042−4049.
Cooperativity of Biological Membranes. Proc. Natl. Acad. Sci. U. S. A. (472) Kremer, C.; Lützen, A. Artificial Allosteric Receptors. Chem. -
1967, 57, 335−341. Eur. J. 2013, 19, 6162−6196.
(449) Hill, T. L. Electric Fields and the Cooperativity of Biological (473) Dutta, S.; Weiner, L.; Sheves, M. Cation Binding to
Membranes. Proc. Natl. Acad. Sci. U. S. A. 1967, 58, 111−114. Halorhodopsin. Biochemistry 2015, 54, 3164−3172.
(450) Rebek, J., Jr.; Wattley, R. V.; Costello, T.; Gadwood, R.; (474) Yifrach, O.; Horovitz, A. Nested Cooperativity in the ATPase
Marshall, L. Allosteric Effects: Binding Cooperativity in a Subunit Activity of the Oligomeric Chaperonin GroEL. Biochemistry 1995, 34,
Model. Angew. Chem., Int. Ed. Engl. 1981, 20, 605−606. 5303−5308.
(451) Di Cera, E.; Gill, S. J.; Wyman, J. Binding Capacity: (475) Mammen, M.; Choi, S.; Whitesides, G. M. Polyvalent
Cooperativity and Buffering in Biopolymers. Proc. Natl. Acad. Sci. U. S. Interactions in Biological Systems: Implications for Design and Use of
A. 1988, 85, 449−452. Multivalent Ligands and Inhibitors. Angew. Chem., Int. Ed. 1998, 37,
(452) Di Cera, E. Site-Specific Thermodynamics: Understanding 2754−2794.
Cooperativity in Molecular Recognition. Chem. Rev. 1998, 98, 1563− (476) Falke, J. J. Cooperativity between Bacterial Chemotaxis
1592. Receptors. Proc. Natl. Acad. Sci. U. S. A. 2002, 99, 6530−6532.
(453) Cui, Q.; Karplus, M. Allostery and Cooperativity Revisited. (477) Bialek, W.; Setayeshgar, S. Cooperativity, Sensitivity, and Noise
Protein Sci. 2008, 17, 1295−1307. in Biochemical Signaling. Phys. Rev. Lett. 2008, 100, No. 258101.
(454) Cárdenas, M. L. Michaelis and Menten and the Long Road to the (478) Chabre, M.; Deterre, P.; Antonny, B. The Apparent
Discovery of Cooperativity. FEBS Lett. 2013, 587, 2767−2771. Cooperativity of Some GPCRs Does Not Necessarily Imply
(455) Huang, H. W. Molecular Mechanism of Antimicrobial Peptides: Dimerization. Trends Pharmacol. Sci. 2009, 30, 182−187.
The Origin of Cooperativity. Biochim. Biophys. Acta, Biomembr. 2006, (479) Sleat, D. E.; Wiseman, J. A.; El-Banna, M.; Price, S. M.; Verot, L.;
1758, 1292−1302. Shen, M. M.; Tint, G. S.; Vanier, M. T.; Walkley, S. U.; Lobel, P. Genetic
(456) Edelstein, S. A Novel Equation for Cooperativity of the Evidence for Nonredundant Functional Cooperativity between NPC1
Allosteric State Function. J. Mol. Biol. 2014, 426, 39−42. and NPC2 in Lipid Transport. Proc. Natl. Acad. Sci. U. S. A. 2004, 101,
(457) Jencks, W. P. On the Attribution and Additivity of Binding 5886−5891.
Energies. Proc. Natl. Acad. Sci. U. S. A. 1981, 78, 4046−4050. (480) Zandany, N.; Ovadia, M.; Orr, I.; Yifrach, O. Direct Analysis of
(458) Tsai, C.; Nussinov, R. A Unified View of ‘‘How Allostery Cooperativity in Multisubunit Allosteric Proteins. Proc. Natl. Acad. Sci.
Works’’. PLoS Comput. Biol. 2014, 10, e1003394−12. U. S. A. 2008, 105, 11697.
(459) Bouhaddou, M.; Birtwistle, M. R. Dimerization-Based Control of (481) Arnett, K. L.; Hass, M.; McArthur, D. G.; Ilagan, M. X. G.; Aster,
Cooperativity. Mol. BioSyst. 2014, 10, 1824−1832. J. C.; Kopan, R.; Blacklow, S. C. Structural and Mechanistic Insights into

2817 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Cooperative Assembly of Dimeric Notch Transcription Complexes. (505) Ueng, Y. F.; Kuwabara, T.; Chun, Y. J.; Guengerich, F. P.
Nat. Struct. Mol. Biol. 2010, 17, 1312−1318. Cooperativity in Oxidations Catalyzed by Cytochrome P450 3A4.
(482) Patlak, J. B. Cooperating to Unlock the Voltage-dependent K Biochemistry 1997, 36, 370−381.
Channel. J. Gen. Physiol. 1999, 113, 385−387. (506) Harlow, G. R.; Halpert, J. R. Analysis of Human Cytochrome
(483) Oana Popa, M.; Alekov, A. K.; Bail, S.; Lehmann-Horn, F.; P450 3A4 Cooperativity: Construction and Characterization of a Site-
Lerche, H. Cooperative Effect of S4-S5 Loops in Domains D3 and D4 Directed Mutant that Displays Hyperbolic Steroid Hydroxylation
on Fast Inactivation of the Na+ Channel. J. Physiol. 2004, 561, 39−51. Kinetics. Proc. Natl. Acad. Sci. U. S. A. 1998, 95, 6636−6641.
(484) Pathak, M.; Kurtz, L.; Tombola, F.; Isacoff, E. The Cooperative (507) Bren, U.; Oostenbrink, C. Cytochrome P450 3A4 Inhibition by
Voltage Sensor Motion that Gates a Potassium Channel. J. Gen. Physiol. Ketoconazole: Tackling the Problem of Ligand Cooperativity Using
2004, 125, 57−69. Molecular Dynamics Simulations and Free-Energy Calculations. J.
(485) Sack, J. T.; Aldrich, R. W. Binding of a Gating Modifier Toxin Chem. Inf. Model. 2012, 52, 1573−1582.
Induces Intersubunit Cooperativity Early in the Shaker K Channel’s (508) Qian, H. Thermodynamic and Kinetic Analysis of Sensitivity
Activation Pathway. J. Gen. Physiol. 2006, 128, 119−132. Amplification in Biological Signal Transduction. Biophys. Chem. 2003,
(486) Nekouzadeh, A.; Silva, J. R.; Rudy, Y. Modeling Subunit 105, 585−593.
Cooperativity in Opening of Tetrameric Ion Channels. Biophys. J. 2008, (509) Qian, H.; Cooper, J. A. Temporal Cooperativity and Sensitivity
95, 3510−3520. Amplification in Biological Signal Transduction. Biochemistry 2008, 47,
(487) Richards, M. J.; Gordon, S. E. Cooperativity and Cooperation in 2211−2220.
Cyclic Nucleotide-Gated Ion Channels. Biochemistry 2000, 39, 14003− (510) DeChancie, J.; Houk, K. N. The Origins of Femtomolar Protein-
14011. Ligand Binding: Hydrogen-Bond Cooperativity and Desolvation
(488) Meyer, T.; Holowka, D.; Stryer, L. Highly Cooperative Opening Energetics in the Biotin-(Strept)Avidin Binding Site. J. Am. Chem. Soc.
of Calcium Channels by Inositol 1, 4, 5-Trisphosphate. Science 1988, 2007, 129, 5419−5429.
240, 653−656. (511) Van Roey, K.; Orchard, S.; Kerrien, S.; Dumousseau, M.; Ricard-
(489) Qian, X.; Niu, X.; Magleby, K. L. Intra- and Intersubunit Blum, S.; Hermjakob, H.; Gibson, T. J. Capturing Cooperative
Cooperativity in Activation of BK Channels by Ca2+. J. Gen. Physiol. Interactions with the PSI-MI Format. Database 2013, 2013, bat066.
2006, 128, 389−404. (512) Clausen, T.; Kaiser, M.; Huber, R.; Ehrmann, M. HTRA
(490) Smith-Maxwell, C. J.; Ledwell, J. L.; Aldrich, R. W. Role of the S4 Proteases: Regulated Proteolysis in Protein Quality Control. Nat. Rev.
in Cooperativity of Voltage-Dependent Potassium Channel Activation. Mol. Cell Biol. 2011, 12, 152−162.
J. Gen. Physiol. 1998, 111, 399−420. (513) Di Cera, E. Serine Proteases. IUBMB Life 2009, 61, 510−515.
(491) Tombola, F.; Ulbrich, M. H.; Kohout, S. C.; Isacoff, E. Y. The (514) Gohara, D. W.; Di Cera, E. Allostery in Trypsin-like Proteases
Opening of the Two Pores of the Hv1 Voltage-Gated Proton Channel is Suggests New Therapeutic Strategies. Trends Biotechnol. 2011, 29, 577−
585.
Tuned by Cooperativity. Nat. Struct. Mol. Biol. 2010, 17, 44−50.
(515) Warshel, A.; Sharma, P. K.; Kato, M.; Xiang, Y.; Liu, H.; Olsson,
(492) Zhou, L. Ion Channels: Cooperativity in Twin Gatings. Nat.
M. H. M. Electrostatic Basis for Enzyme Catalysis. Chem. Rev. 2006, 106,
Chem. Biol. 2012, 8, 136−137.
3210−3235.
(493) Fujiwara, Y.; Kurokawa, T.; Takeshita, K.; Kobayashi, M.;
(516) Sohn, J.; Grant, R. A.; Sauer, R. T. Allostery is an Intrinsic
Okochi, Y.; Nakagawa, A.; Okamura, Y. The Cytoplasmic Coiled-Coil
Property of the Protease Domain of DegS. J. Biol. Chem. 2010, 285,
Mediates Cooperative Gating Temperature Sensitivity in the Voltage-
34039−34047.
Gated H+ Channel Hv1. Nat. Commun. 2012, 3, 816−11. (517) Polgár, L. The Catalytic Triad of Serine Peptidases. Cell. Mol. Life
(494) Saito, M. Subunit Cooperativity in the Action of Lactate Sci. 2005, 62, 2161−2172.
Dehydrogenase. Biochim. Biophys. Acta 1972, 258, 17−26. (518) Lee, G. M.; Shahian, T.; Baharuddin, A.; Gable, J. E.; Craik, C. S.
(495) Koshland, D. E., Jr.; Hamadani, K. Proteomics and Models for Enzyme Inhibition by Allosteric Capture of an Inactive Conformation. J.
Enzyme Cooperativity. J. Biol. Chem. 2002, 277, 46841−46844. Mol. Biol. 2011, 411, 999−1016.
(496) Qian, H. Cooperativity and Specificity in Enzyme Kinetics: A (519) Chen, H.; Zhou, X.; Ou-Yang, Z.-C. Classification of Amino
Single-Molecule Time-Based Perspective. Biophys. J. 2008, 95, 10−17. Acids Based On Statistical Results of Known Structures and
(497) Levin, M. K.; Wang, Y.; Patel, S. S. The Functional Interaction of Cooperativity of Protein Folding. Phys. Rev. E Stat. Nonlin. Soft Matter
the Hepatitis C Virus Helicase Molecules is Responsible for Unwinding Phys. 2002, 65, 061907.
Processivity. J. Biol. Chem. 2004, 279, 26005−26012. (520) Koch, O.; Bocola, M.; Klebe, G. Cooperative Effects in
(498) Sikora, B.; Eoff, R. L.; Matson, S. W.; Raney, K. D. DNA Hydrogen-Bonding of Protein Secondary Structure Elements: A
Unwinding by Escherichia coli DNA Helicase I (TraI) Provides Systematic Analysis of Crystal Data Using Secbase. Proteins: Struct.,
Evidence for a Processive Monomeric Molecular Motor. J. Biol. Chem. Funct., Genet. 2005, 61, 310−317.
2006, 281, 36110−36116. (521) Cho, Y. R.; Maguire, A. J.; Try, A. C.; Westwell, M. S.; Groves, P.;
(499) Guo, H.; Salahub, D. R. Cooperative Hydrogen Bonding and Williams, D. H. Cooperativity and AntiCooperativity between Ligand
Enzyme Catalysis. Angew. Chem., Int. Ed. 1998, 37, 2985−2990. Binding and the Dimerization of Ristocetin A: Asymmetry of a
(500) Liu, J. Q.; Wulff, G. Functional Mimicry of the Active Site of Homodimer Complex and Implications for Signal Transduction. Chem.
Carboxypeptidase A by a Molecular Imprinting Strategy: Cooperativity Biol. 1996, 3, 207−215.
of an Amidinium and a Copper Ion in a Transition-State Imprinted (522) Williams, D. H.; Stephens, E.; O’Brien, D. P.; Zhou, M.
Cavity Giving Rise to High Catalytic Activity. J. Am. Chem. Soc. 2004, Understanding Noncovalent Interactions: Ligand Binding Energy and
126, 7452−7453. Catalytic Efficiency from Ligand-Induced Reductions in Motion within
(501) van Oijen, A. M. Cutting the Forest to See a Single Tree? Nat. Receptors and Enzymes. Angew. Chem., Int. Ed. 2004, 43, 6596−6616.
Chem. Biol. 2008, 4, 440−443. (523) Williams, D. H.; Maguire, A. J.; Tsuzuki, W.; Westwell, M. S. An
(502) Bonivento, D.; Milczek, E. M.; McDonald, G. R.; Binda, C.; Holt, Analysis of the Origins of a Cooperative Binding Energy of
A.; Edmondson, D. E.; Mattevi, A. Potentiation of Ligand Binding Dimerization. Science 1998, 280, 711−714.
through Cooperative Effects in Monoamine Oxidase B. J. Biol. Chem. (524) Williams, D. H.; Davies, N. L.; Zerella, R.; Bardsley, B.
2010, 285, 36849−36856. Noncovalent Interactions: Defining Cooperativity. Ligand Binding
(503) Hammes, G. G.; Benkovic, S. J.; Hammes-Schiffer, S. Flexibility, Aided by Reduced Dynamic Behavior of Receptors. Binding of Bacterial
Diversity, and Cooperativity: Pillars of Enzyme Catalysis. Biochemistry Cell Wall Analogues to Ristocetin. A. J. Am. Chem. Soc. 2004, 126,
2011, 50, 10422−10430. 2042−2049.
(504) Denisov, I. G.; Frank, D. J.; Sligar, S. G. Cooperative Properties (525) Baum, B.; Muley, L.; Smolinski, M.; Heine, A.; Hangauer, D.;
of Cytochromes P450. Pharmacol. Ther. 2009, 124, 151−167. Klebe, G. Non-Additivity of Functional Group Contributions in

2818 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

Protein−Ligand Binding: A Comprehensive Study by Crystallography Performance of MM-PBSA and MM-GBSA Approaches. J. Chem. Inf.
and Isothermal Titration Calorimetry. J. Mol. Biol. 2010, 397, 1042− Model. 2012, 52, 3088−3098.
1054. (544) Srivastava, H. K.; Chourasia, M.; Kumar, D.; Sastry, G. N.
(526) Muley, L.; Baum, B.; Smolinski, M.; Freindorf, M.; Heine, A.; Comparison of Computational Methods to Model DNA Minor Groove
Klebe, G.; Hangauer, D. G. Enhancement of Hydrophobic Interactions Binders. J. Chem. Inf. Model. 2011, 51, 558−571.
and Hydrogen Bond Strength by Cooperativity: Synthesis, Modeling, (545) Bindu, P. H.; Sastry, G. M.; Sastry, G. N. Characterization of
and Molecular Dynamics Simulations of a Congeneric Series of Calcium and Magnesium Binding Domains of Human5-Lipoxygenase.
Thrombin Inhibitors. J. Med. Chem. 2010, 53, 2126−2135. Biochem. Biophys. Res. Commun. 2004, 320, 461−467.
(527) Nasief, N. N.; Tan, H.; Kong, J.; Hangauer, D. Water Mediated (546) Chourasia, M.; Sastry, G. M.; Sastry, G. N. Proton Binding Sites
Ligand Functional Group Cooperativity: The Contribution of a Methyl and Conformational Analysis of H+K+-ATPase. Biochem. Biophys. Res.
Group to Binding Affinity is Enhanced by a COO− Group Through Commun. 2005, 336, 961−966.
Changes in the Structure and Thermodynamics of the Hydration Waters (547) Levinthal, C. Are there Pathways for Protein Folding? J. Chim.
of Ligand−Thermolysin Complexes. J. Med. Chem. 2012, 55, 8283− Phys. 1968, 65, 44−45.
8302. (548) Karplus, M.; Weaver, D. L. Protein-Folding Dynamics. Nature
(528) Biela, A.; Betz, M.; Heine, A.; Klebe, G. Water Makes the 1976, 260, 404−406.
Difference: Rearrangement of Water Solvation Layer Triggers Non- (549) Kim, P. S.; Baldwin, R. L. Specific Intermediates in the Folding
additivity of Functional Group Contributions in Protein−Ligand Reactions of Small Proteins and the Mechanism of Protein Folding.
Binding. ChemMedChem 2012, 7, 1423−1434. Annu. Rev. Biochem. 1982, 51, 459−489.
(529) Kuhn, B.; Fuchs, J. E.; Reutlinger, M.; Stahl, M.; Taylor, N. R. (550) Bryngelson, J. D.; Onuchic, J. N.; Socci, N. D.; Wolynes, P. G.
Rationalizing Tight Ligand Binding through Cooperative Interaction Funnels, Pathways, and the Energy Landscape of Protein Folding: A
Networks. J. Chem. Inf. Model. 2011, 51, 3180−3198. Synthesis. Proteins: Struct., Funct., Genet. 1995, 21, 167−195.
(530) Houk, K. N.; Leach, A. G.; Kim, S. P.; Zhang, X. Binding (551) Radford, S. E.; Dobson, C. M.; Evans, P. A. The Folding of Hen
Affinities of Host−Guest, Protein−Ligand, and Protein−Transition- Lysozyme Involves Partially Structured Intermediates and Multiple
State Complexes. Angew. Chem., Int. Ed. 2003, 42, 4872−4897. Pathways. Nature 1992, 358, 302−307.
(531) Said, A. M.; Hangauer, D. G. Binding Cooperativity Between a (552) Freire, E.; Murphy, K. P.; Sanchez-Ruiz, J. M.; Galisteo, M. L.;
Ligand Carbonyl Group and a Hydrophobic Side Chain Can be Privalov, P. L. The Molecular Basis of Cooperativity in Protein Folding.
Enhanced by Additional H-Bonds in a Distance Dependent Manner: A Thermodynamic Dissection of Interdomain Interactions in Phospho-
Case Study with Thrombin Inhibitors. Eur. J. Med. Chem. 2015, 96, glycerate Kinase. Biochemistry 1992, 31, 250−256.
405−424. (553) Staniforth, R. A.; Burston, S. G.; Smith, C. J.; Jackson, G. S.;
(532) Nasief, N. N.; Hangauer, D. Additivity or Cooperativity: Which Badcoe, I. G.; Atkinson, T.; Holbrook, J. J.; Clarke, A. R. The Energetics
Model Can Predict the Influence of Simultaneous Incorporation of Two and Cooperativity of Protein Folding: A Simple Experimental Analysis
or More Functionalities in a Ligand Molecule? Eur. J. Med. Chem. 2015, based upon the Solvation of Internal Residues. Biochemistry 1993, 32,
90, 897−915.
3842−3851.
(533) Janardhan, S.; Srivani, P.; Sastry, G. N. Choline Kinase: An
(554) Udgaonkar, J. B.; Baldwin, R. L. NMR Evidence for an Early
Important Target for Cancer. Curr. Med. Chem. 2006, 13, 1169−1189.
Framework Intermediate on the Folding Pathway of Ribonuclease A.
(534) Janardhan, S.; Srivani, P.; Sastry, G. N. 2D and 3D Quantitative
Nature 1988, 335, 694−699.
Structure-Activity Relationship Studies on a Series of bis-Pyridinium
(555) Baker, D. A Surprising Simplicity to Protein Folding. Nature
Compounds as Choline Kinase Inhibitors. QSAR Comb. Sci. 2006, 25,
2000, 405, 39−42.
860−872.
(556) Kuwajima, K. The Molten Globule State as a Clue for
(535) Srivani, P.; Sastry, G. N. Potential Choline Kinase Inhibitors: A
Understanding the Folding and Cooperativity of Globular-Protein
Molecular Modeling Study of bis-Quinolinium Compounds. J. Mol.
Structure. Proteins: Struct., Funct., Genet. 1989, 6, 87−103.
Graphics Modell. 2009, 27, 676−688.
(557) Dill, K. A.; Fiebig, K. M.; Chan, H. S. Cooperativity in Protein-
(536) Kulkarni, R.; Achaiah, G.; Sastry, G. N. Novel Targets for Anti-
Inflammatory and Anti-Arthritic Agents. Curr. Pharm. Des. 2006, 12, Folding Kinetics. Proc. Natl. Acad. Sci. U. S. A. 1993, 90, 1942−1946.
(558) Ghosh, K.; Dill, K. A. Theory for Protein Folding Cooperativity:
2437−2454.
(537) Badrinarayan, P.; Sastry, G. N. Specificity Rendering ‘Hot-Spots’ Helix Bundles. J. Am. Chem. Soc. 2009, 131, 2306−2312.
for Aurora Kinase Inhibitor Design: The Role of Non-Covalent (559) Chan, H. S. Modeling Protein Density of States: Additive
Interactions and Conformational Transitions. PLoS One 2014, 9, Hydrophobic Effects Are Insufficient for Calorimetric Two-State
e113773. Cooperativity. Proteins: Struct., Funct., Genet. 2000, 40, 543−571.
(538) Badrinarayan, P.; Sastry, G. N. Sequence, Analysis of p38 MAP (560) Kaya, H.; Chan, H. S. Towards a Consistent Modeling of Protein
Kinase: Exploiting DFG-Out Conformation as a Strategy to Design New Thermodynamic and Kinetic Cooperativity: How Applicable is the
Type II leads. J. Chem. Inf. Model. 2011, 51, 115−129. Transition State Picture to Folding and Unfolding? J. Mol. Biol. 2002,
(539) Badrinarayan, P.; Sastry, G. N. Virtual Screening Filters for the 315, 899−909.
Design of Type II p38 MAP Kinase Inhibitors: A Fragment Based (561) Badasyan, A.; Liu, Z.; Chan, H. S. Interplaying Roles of Native
Library Generation Approach. J. Mol. Graphics Modell. 2012, 34, 89− Topology and Chain Length in Marginally Cooperative and Non-
100. cooperative Folding of Small Protein Fragments. Int. J. Quantum Chem.
(540) Badrinarayan, P.; Sastry, G. N. Rational Approaches towards 2009, 109, 3482−3499.
Lead Optimization of Kinase Inhibitors: The Issue of Specificity. Curr. (562) Scalley-Kim, M.; Baker, D. Characterization of the Folding
Pharm. Des. 2013, 19, 4714−4738. Energy Landscapes of Computer Generated Proteins Suggests High
(541) Choudhury, C.; Priyakumar, U. D.; Sastry, G. N. Dynamics Folding Free Energy Barriers and Cooperativity may be Consequences
Based Pharmacophore Models for Screening Potential Inhibitors of of Natural Selection. J. Mol. Biol. 2004, 338, 573−583.
Mycobacterial Cyclopropane Synthase. J. Chem. Inf. Model. 2015, 55, (563) Watters, A. L.; Deka, P.; Corrent, C.; Callender, D.; Varani, G.;
848−860. Sosnick, T.; Baker, D. The Highly Cooperative Folding of Small
(542) Reddy, A. S.; Pati, S. P.; Kumar, P. P.; Pradeep, H. N.; Sastry, G. Naturally Occurring Proteins is likely the Result of Natural Selection.
N. Virtual Screening in Drug Discovery − A Computational Perspective. Cell 2007, 128, 613−624.
Curr. Protein Pept. Sci. 2007, 8, 329−351. (564) Plaxco, K. W.; Simons, K. T.; Baker, D. Contact Order,
(543) Srivastava, H. K.; Sastry, G. N. A Molecular Dynamics Transition State Placement and the Refolding Rates of Single Domain
Investigation on a Series of HIV Protease Inhibitors: Assessing the Proteins. J. Mol. Biol. 1998, 277, 985−994.

2819 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(565) Jewett, A. I.; Pande, V. S.; Plaxco, K. W. Cooperativity, Smooth (588) Orlova, A.; Prochniewicz, E.; Egelman, E. H. Structural
Energy Landscapes and the Origins of Topology-Dependent Protein Dynamics of F-Actin: II. Cooperativity in Structural Transitions. J.
Folding Rates. J. Mol. Biol. 2003, 326, 247−253. Mol. Biol. 1995, 245, 598−607.
(566) Faísca, P. F. N.; Plaxco, K. W. Cooperativity and the Origins of (589) Kolinski, A.; Galazka, W.; Skolnick, J. On the Origin of the
Rapid, Single-Exponential Kinetics in Protein Folding. Protein Sci. 2006, Cooperativity of Protein Folding: Implications from Model Simulations.
15, 1608−1618. Proteins: Struct., Funct., Genet. 1996, 26, 271−287.
(567) Zhou, Y.; Linhananta, A. Thermodynamics of an All-Atom Off- (590) Luo, Y.; Kay, M. S.; Baldwin, R. L. Cooperativity of Folding of
Lattice Model of the Fragment B of Staphylococcal Protein A: the Apomyoglobin pH4 Intermediate Studied by Glycine and Proline
Implication for the Origin of the Cooperativity of Protein Folding. J. Mutations. Nat. Struct. Biol. 1997, 4, 925−930.
Phys. Chem. B 2002, 106, 1481−1485. (591) Fernandez, A.; Colubri, A.; Berry, R. S. Three-body Correlations
(568) Klimov, D. K.; Thirumalai, D. Cooperativity in Protein Folding: in Protein Folding: The Origin of Cooperativity. Phys. A 2002, 307,
From Lattice Models with Sidechains to Real Proteins. Folding Des. 235−259.
1998, 3, 127−139. (592) Keskin, O.; Ma, B.; Rogale, K.; Gunasekaran, K.; Nussinov, R.
(569) Kouza, M.; Li, M. S.; O’Brien, E. P., Jr.; Hu, C.; Thirumalai, D. Protein−Protein Interactions: Organization, Cooperativity and Map-
Effect of Finite Size on Cooperativity and Rates of Protein Folding. J. ping in a Bottom-Up Systems Biology Approach. Phys. Biol. 2005, 2,
Phys. Chem. A 2006, 110, 671−676. S24−S35.
(570) Yang, X.; Kathuria, S. V.; Vadrevu, R.; Matthews, C. R. PLoS One (593) Kloss, E.; Courtemanche, N.; Barrick, D. Repeat-Protein
2009, 4, e7179−9. Folding: New Insights into Origins of Cooperativity, Stability, and
(571) Canet, D.; Last, A. M.; Tito, P.; Sunde, M.; Spencer, A.; Archer, Topology. Arch. Biochem. Biophys. 2008, 469, 83−99.
D. B.; Redfield, C.; Robinson, C. V.; Dobson, C. M. Local Cooperativity (594) Yang, L.; Song, G.; Jernigan, R. L. Protein Elastic Network
in the Unfolding of an Amyloidogenic Variant of Human Lysozyme. Models and the Ranges of Cooperativity. Proc. Natl. Acad. Sci. U. S. A.
Nat. Struct. Biol. 2002, 9, 308−315. 2009, 106, 12347−12352.
(572) Batey, S.; Clarke, J. Apparent Cooperativity in the Folding of (595) Sengupta, D.; Kundu, S. Role of Long- and Short-Range
Multidomain Proteins Depends on the Relative Rates of Folding of the Hydrophobic, Hydrophilic and Charged Residues Contact Network in
Constituent Domains. Proc. Natl. Acad. Sci. U. S. A. 2006, 103, 18113− Protein’s Structural Organization. BMC Bioinf. 2012, 13, 142−14.
(596) Gruszka, D. T.; Whelan, F.; Farrance, O. E.; Fung, H. K.; Paci, E.;
18118.
Jeffries, C. M.; Svergun, D. I.; Baldock, C.; Baumann, C. G.; Brockwell,
(573) Shank, E. A.; Cecconi, C.; Dill, J. W.; Marqusee, S.; Bustamante,
D. J.; Potts, J. R.; Clarke, J. Cooperative Folding of Intrinsically
C. The Folding Cooperativity of a Protein is Controlled by its Chain
Disordered Domains Drives Assembly of a Strong Elongated Protein.
Topology. Nature 2010, 465, 637−640.
Nat. Commun. 2015, 6, 7271−7279.
(574) Schenck, H. L.; Gellman, S. H. Use of a Designed Triple-
(597) Malhotra, P.; Udgaonkar, J. B. Tuning Cooperativity on the Free
Stranded Antiparallel β-Sheet to Probe β-Sheet Cooperativity in
Energy Landscape of Protein Folding. Biochemistry 2015, 54, 3431−
Aqueous Solution. J. Am. Chem. Soc. 1998, 120, 4869−4870.
3441.
(575) Searle, M. S.; Ciani, B. Design of β-Sheet Systems for
(598) Schüle, R.; Muller, M.; Otsuka-Murakami, H.; Renkawitz, R.
Understanding the Thermodynamics and Kinetics of Protein Folding. Cooperativity of the Glucocorticoid Receptor and the CACCC-Box
Curr. Opin. Struct. Biol. 2004, 14, 458−464. Binding Factor. Nature 1988, 332, 87−90.
(576) Xu, Y.; Bunagan, M. R.; Tang, J.; Gai, F. Probing the Kinetic (599) Ogata, K.; Sato, K.; Tahirov, T. H. Eukaryotic Transcriptional
Cooperativity of β-Sheet Folding Perpendicular to the Strand Direction. Regulatory Complexes: Cooperativity from Near and Afar. Curr. Opin.
Biochemistry 2008, 47, 2064. Struct. Biol. 2003, 13, 40−48.
(577) Roe, D. R.; Hornak, V.; Simmerling, C. Folding Cooperativity in (600) Dodd, I. B.; Shearwin, K. E.; Perkins, A. J.; Burr, T.; Hochschild,
a Three-Stranded β-Sheet Model. J. Mol. Biol. 2005, 352, 370−381. A.; Egan, J. B. Cooperativity in Long-Range Gene Regulation by the λ CI
(578) Hills, R. D., Jr.; Brooks, C. L., III Hydrophobic Cooperativity as a Repressor. Genes Dev. 2004, 18, 344−354.
Mechanism for Amyloid Nucleation. J. Mol. Biol. 2007, 368, 894−901. (601) Dandanell, G.; Valentin-Hansen, P.; Larsen, J. E.; Hammer, K.
(579) Ralston, C. Y.; He, Q.; Brenowitz, M.; Chance, M. R. Stability Long-Range Cooperativity between Gene Regulatory Sequences in a
and Cooperativity of Individual Tertiary Contacts in RNA Revealed Prokaryote. Nature 1987, 325, 823−826.
through Chemical Denaturation. Nat. Struct. Biol. 2000, 7, 371−374. (602) Banerjee, N.; Zhang, M. Q. Identifying Cooperativity among
(580) Siegfried, N. A.; Metzger, S. L.; Bevilacqua, P. C. Folding Transcription Factors Controlling the Cell Cycle in Yeast. Nucleic Acids
Cooperativity in RNA and DNA is Dependent on Position in the Helix. Res. 2003, 31, 7024−7031.
Biochemistry 2007, 46, 172−181. (603) Dehner, A.; Klein, C.; Hansen, S.; Müller, L.; Buchner, J.;
(581) Feng, J.; Walter, N. G.; Brooks, C. L. Cooperative and Schwaiger, M.; Kessler, H. Cooperative Binding of p53 to DNA:
Directional Folding of the preQ1 Riboswitch Aptamer Domain. J. Am. Regulation by Protein−Protein Interactions through a Double Salt
Chem. Soc. 2011, 133, 4196−4199. Bridge. Angew. Chem., Int. Ed. 2005, 44, 5247−5251.
(582) Zuo, G.; Wang, J.; Wang, W. Folding with Downhill Behavior (604) Krell, T.; Terán, W.; Mayorga, O. L.; Rivas, G.; Jiménez, M.;
and Low Cooperativity of Proteins. Proteins: Struct., Funct., Genet. 2006, Daniels, C.; Molina-Henares, A. J.; Martínez-Bueno, M.; Gallegos, M.
63, 165−173. T.; Ramos, J. L. Optimization of the Palindromic Order of the TtgR
(583) Zhou, Z.; Bai, Y. Structural Biology: Analysis of Protein-Folding Operator Enhances Binding Cooperativity. J. Mol. Biol. 2007, 369,
Cooperativity. Nature 2007, 445, E16−E17. 1188−1199.
(584) Sadqi, M.; Fushman, D.; Muñoz, V. Atom-by-Atom Analysis of (605) King, J. C.; Xu, J.; Wongvipat, J.; Hieronymus, H.; Carver, B. S.;
Global Downhill Protein Folding. Nature 2006, 442, 317−321. Leung, D. H.; Taylor, B. S.; Sander, C.; Cardiff, R. D.; Couto, S. S.;
(585) Ferguson, N.; Sharpe, T. D.; Johnson, C. M.; Schartau, P. J.; Gerald, W. L.; Sawyers, C. L. Cooperativity of TMPRSS2-ERG with PI3-
Fersht, A. R. Structural Biology: Analysis of ’Downhill’ Protein Folding. Kinase Pathway Activation in Prostate Oncogenesis. Nat. Genet. 2009,
Nature 2007, 445, E14−E15. 41, 524−526.
(586) Li, Y.; Ji, C.; Xu, W.; Zhang, J. Z. H. Dynamical Stability and (606) Mirny, L. A. Nucleosome-Mediated Cooperativity between
Assembly Cooperativity of β-Sheet Amyloid Oligomers − Effect of Transcription Factors. Proc. Natl. Acad. Sci. U. S. A. 2010, 107, 22534−
Polarization. J. Phys. Chem. B 2012, 116, 13368−13373. 22539.
(587) Diehl, M. R.; Zhang, K.; Lee, H. J.; Tirrell, D. A. Engineering (607) Mekler, V.; Severinov, K. Cooperativity and Interaction Energy
Cooperativity in Biomotor-Protein Assemblies. Science 2006, 311, Threshold Effects in Recognition of the − 10 Promoter Element by
1468−1471. Bacterial RNA Polymerase. Nucleic Acids Res. 2013, 41, 7276−7285.

2820 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(608) Kerppola, T. K.; Curran, T. Fos-Jun Heterodimers and Jun (630) Antony, J.; Grimme, S. Structures and Interaction Energies of
Homodimers Bend DNA in Opposite Orientations: Implications for Stacked Graphene−Nucleobase Complexes. Phys. Chem. Chem. Phys.
Transcription Factor Cooperativity. Cell 1991, 66, 317−326. 2008, 10, 2722−2729.
(609) Mel’nikov, S. M.; Sergeyev, V. G.; Yoshikawa, K.; Takahashi, H.; (631) Levitzki, A.; Koshland, D. E., Jr. Negative Cooperativity in
Hatta, I. Cooperativity or Phase Transition? Unfolding Transition of Regulatory Enzymes. Proc. Natl. Acad. Sci. U. S. A. 1969, 62, 1121−1128.
DNA Cationic Surfactant Complex. J. Chem. Phys. 1997, 107, 6917− (632) DeMeyts, P.; Bainco, A. R.; Roth, J. Site-Site Interactions among
6924. Insulin Receptors. J. Biol. Chem. 1976, 251, 1877−1888.
(610) Becaud, J.; Pompizi, I.; Leumann, C. J. Propagation of Melting (633) Koshland, D. E., Jr. The Structural Basis of Negative
Cooperativity along the Phosphodiester Backbone of DNA. J. Am. Cooperativity: Receptors and Enzymes. Curr. Opin. Struct. Biol. 1996,
Chem. Soc. 2003, 125, 15338−15342. 6, 757−761.
(611) Escudero, D.; Estarellas, C.; Frontera, A.; Quiñonero, D.; Deyà, (634) Flatmark, T.; Almås, B.; Knappskog, P. M.; Berge, S. V.; Svebak,
P. M. Cooperativity Effects between Noncovalent Interactions: Are they R. M.; Chehin, R.; Muga, A.; Martínez, A. Tyrosine Hydroxylase Binds
important for Z-DNA Stability? Chem. Phys. Lett. 2010, 485, 221−225. Tetrahydrobiopterin Cofactor with Negative Cooperativity, as Shown
(612) Yurenko, Y. P.; Novotný, J.; Sklenár,̌ V.; Marek, R. Exploring by Kinetic Analyses and Surface Plasmon Resonance Detection. Eur. J.
Noncovalent Interactions in Guanine- and Xanthine-Based Model DNA Biochem. 1999, 262, 840−849.
Quadruplex Structures: A Comprehensive Quantum Chemical (635) Stevens, S. Y.; Sanker, S.; Kent, C.; Zuiderweg, E. R. Delineation
Approach. Phys. Chem. Chem. Phys. 2014, 16, 2072−2084. of the Allosteric Mechanism of a Cytidylyltransferase Exhibiting
(613) Gong, B. Molecular Duplexes with Encoded Sequences and Negative Cooperativity. Nat. Struct. Biol. 2001, 8, 947−952.
Stabilities. Acc. Chem. Res. 2012, 45, 2077−2087. (636) Walter, E. D.; Chattopadhyay, M.; Millhauser, G. L. The Affinity
(614) Arbona, J.; Aimé, J.; Elezgaray, J. Cooperativity in the Annealing of Copper Binding to the Prion Protein Octarepeat Domain: Evidence
of DNA Origamis. J. Chem. Phys. 2013, 138, 015105−10. for Negative Cooperativity. Biochemistry 2006, 45, 13083−13092.
(615) Norregaard, K.; Andersson, M.; Sneppen, K.; Nielsen, P. E.; (637) Aramaki, H.; Kabata, H.; Takeda, S.; Itou, H.; Nakayama, H.;
Brown, S.; Oddershede, L. B. DNA Supercoiling Enhances Coopera- Shimamoto, N. Formation of Repressor-Inducer-Operator Ternary
tivity and Efficiency of an Epigenetic Switch. Proc. Natl. Acad. Sci. U. S. A. Complex: Negative Cooperativity of D-Camphor Binding to CamR.
2013, 110, 17386−17391. Genes Cells. 2011, 16, 1200−1207.
(616) Sattin, B. D.; Zhao, W.; Travers, K.; Chu, S.; Herschlag, D. Direct (638) Wang, C.; Wang, Z.; Zhang, X. Amphiphilic Building Blocks for
Measurement of Tertiary Contact Cooperativity in RNA Folding. J. Am. Self-Assembly: From Amphiphiles to Supra-Amphiphiles. Acc. Chem.
Chem. Soc. 2008, 130, 6085−6087. Res. 2012, 45, 608−618.
(617) Kwon, M.; Strobel, S. A. Chemical Basis of Glycine Riboswitch (639) Mali, K. S.; Lava, K.; Binnemans, K.; De Feyter, S. Hydrogen
Cooperativity. RNA 2007, 14, 25−34. Bonding Versus van der Waals Interactions: Competitive Influence of
(618) Overgaard, M.; Borch, J.; Jørgensen, M. G.; Gerdes, K. Noncovalent Interactions on 2D Self-Assembly at the Liquid−Solid
Messenger RNA Interferase RelE Controls relBE Transcription by
Interface. Chem. - Eur. J. 2010, 16, 14447−14458.
Conditional Cooperativity. Mol. Microbiol. 2008, 69, 841−857.
(640) Hoeben, F. J. M.; Jonkheijm, P.; Meijer, E. W.; Schenning, A. P.
(619) Said, H.; Schüller, V. J.; Eber, F. J.; Wege, C.; Liedl, T.; Richert,
H. J. About Supramolecular Assemblies of π-Conjugated Systems. Chem.
C. M1.3 − A Small Scaffold for DNA Origami. Nanoscale 2013, 5, 284−
Rev. 2005, 105, 1491−1546.
290.
(641) Jonkheijm, P.; van der Schoot, P.; Schenning, A. P. H. J.; Meijer,
(620) Ares, S.; Voulgarakis, N. K.; Rasmussen, K. Ø.; Bishop, A. R.
E. W. Probing the Solvent-Assisted Nucleation Pathway in Chemical
Bubble Nucleation and Cooperativity in DNA Melting. Phys. Rev. Lett.
Self-Assembly. Science 2006, 313, 80−83.
2005, 94, No. 035504.
(642) Besenius, P.; Portale, G.; Bomans, P. H. H.; Janssen, H. M.;
(621) Bolognesi, M.; Boffi, A.; Coletta, M.; Mozzarelli, A.; Pesce, A.;
Tarricone, C.; Ascenzi, P. Anticooperative Ligand Binding Properties of Palmans, A. R. A.; Meijer, E. W. Controlling the Growth and Shape of
Recombinant Ferric Vitreoscilla Homodimeric Hemoglobin: A Chiral Supramolecular Polymers in Water. Proc. Natl. Acad. Sci. U. S. A.
Thermodynamic, Kinetic and X-ray Crystallographic Study. J. Mol. 2010, 107, 17888−17893.
Biol. 1999, 291, 637−650. (643) Markvoort, A. J.; ten Eikelder, H. M. M.; Hilbers, P. A. J.; de
(622) Iqbalsyah, T. M.; Doig, A. J. Anticooperativity in a Glu−Lys− Greef, T. F.A.; Meijer, E. W. Theoretical Models of Nonlinear Effects in
Glu Salt Bridge Triplet in an Isolated α-Helical Peptide. Biochemistry Two-Component Cooperative Supramolecular Copolymerizations.
2005, 44, 10449−10456. Nat. Commun. 2011, 2, 509−9.
(623) Piazza, F.; De Los Rios, P.; Fanelli, D.; Bongini, L.; Skoglund, U. (644) García, F.; Korevaar, P. A.; Verlee, A.; Meijer, E. W.; Palmans, A.
Anticooperativity in Diffusion-Controlled Reactions with Pairs of R. A.; Sánchez, L. The Influence of π-Conjugated Moieties on the
Partially Absorbing Domains: A Model for the Antigen-Antibody Thermodynamics of Cooperatively Self-Assembling Tricarboxamides.
Encounter. Eur. Biophys. J. 2005, 34, 899−911. Chem. Commun. 2013, 49, 8674−8676.
(624) Li, X.; Liang, J. Geometric Cooperativity and Anticooperativity (645) Smulders, M. M. J.; Schenning, A. P. H. J.; Meijer, E. W. Insight
of Three-Body Interactions in Native Proteins. Proteins: Struct., Funct., into the Mechanisms of Cooperative Self-Assembly: The “Sergeants-
Genet. 2005, 60, 46−65. and-Soldiers” Principle of Chiral and Achiral C3-Symmetrical Discotic
(625) Czaplewski, C.; Liwo, A.; Ripoll, D. R.; Scheraga, H. A. Triamides. J. Am. Chem. Soc. 2008, 130, 606−611.
Molecular Origin of Anticooperativity in Hydrophobic Association. J. (646) van der Weegen, R.; Korevaar, P. A.; Voudouris, P.; Voets, I. K.;
Phys. Chem. B 2005, 109, 8108−8119. de Greef, T. F.; Vekemans, J. A.; Meijer, E. W. Small Sized Perylene-
(626) Shimizu, S.; Moghaddam, M. S.; Chan, H. S. Comment on Bisimide Assemblies Controlled by Both Cooperative and AntiCooper-
“Molecular Origin of Anticooperativity in Hydrophobic Association. J. ative Assembly Processes. Chem. Commun. 2013, 49, 5532−5534.
Phys. Chem. B 2005, 109, 21220−21221. (647) Tian, Y.; Meijer, E. W.; Wang, F. Cooperative Self-Assembly of
(627) Kudrev, A. G. Calculation of the Cooperativity and Platinum(II) Acetylide Complexes. Chem. Commun. 2013, 49, 9197−
Anticooperativity Parameters of the Interaction of a Ligand with an 9199.
Infinite Homogeneous Polymer. J. Anal. Chem. 2001, 56, 232−237. (648) Badjić, J. D.; Nelson, A.; Cantrill, S. J.; Turnbull, W. B.; Stoddart,
(628) Cysewski, P. Non-Additive Interactions of Nucleobases in J. F. Multivalency and Cooperativity in Supramolecular Chemistry. Acc.
Model Dinucleotide Steps Occurring in B-DNA Crystals. J. Mol. Model. Chem. Res. 2005, 38, 723−732.
2010, 16, 1721−1729. (649) Belowich, M. E.; Valente, C.; Smaldone, R. A.; Friedman, D. C.;
(629) Szczięśniak, M. M.; Ratajczak, H. Nonadditivity of the SCF Thiel, J.; Cronin, L.; Stoddart, J. F. Positive Cooperativity in the
Interaction Energy in the (LiH)3 Complex. Int. J. Quantum Chem. 1980, Template-Directed Synthesis of Monodisperse Macromolecules. J. Am.
17, 1069−1074. Chem. Soc. 2012, 134, 5243−5261.

2821 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(650) Stone, M. T.; Heemstra, J. M.; Moore, J. S. The Chain-Length Development of Proteomimetic Polyenes. Macromolecules 2003, 36,
Dependence Test. Acc. Chem. Res. 2006, 39, 11−20. 9752−9762.
(651) Wackerly, J. W.; Moore, J. S. Cooperative Self-Assembly of (669) Lanigan, N.; Wang, X. Supramolecular Chemistry of Metal
Oligo(m-phenyleneethynylenes) into Supramolecular Coordination Complexes in Solution. Chem. Commun. 2013, 49, 8133−8144.
Polymers. Macromolecules 2006, 39, 7269−7276. (670) Prince, R. B.; Saven, J. G.; Wolynes, P. G.; Moore, J. S.
(652) Filot, I. A.W.; Palmans, A. R. A.; Hilbers, P. A. J.; van Santen, R. Cooperative Conformational Transitions in Phenylene Ethynylene
A.; Pidko, E. A.; de Greef, T. F. A. Understanding Cooperativity in Oligomers: Chain-Length Dependence. J. Am. Chem. Soc. 1999, 121,
Hydrogen-Bond-Induced Supramolecular Polymerization: A Density 3114−3121.
Functional Theory Study. J. Phys. Chem. B 2010, 114, 13667−13674. (671) Stone, M. T.; Moore, J. S. Supramolecular Chelation Based on
(653) Björk, J.; Stafström, S.; Hanke, F. Zipping Up: Cooperativity Folding. J. Am. Chem. Soc. 2005, 127, 5928−5935.
Drives the Synthesis of Graphene Nanoribbons. J. Am. Chem. Soc. 2011, (672) Camara-Campos, A.; Hunter, C. A.; Tomas, S. Cooperativity in
133, 14884−14887. the Self-Assembly of Porphyrin Ladders. Proc. Natl. Acad. Sci. U. S. A.
(654) Riis-Johannessen, T.; Favera, N. D.; Todorova, T. K.; Huber, S. 2006, 103, 3034−3038.
M.; Gagliardi, L.; Piguet, C. Understanding, Controlling and (673) Ercolani, G.; Schiaffino, L. Allosteric, Chelate, and Interannular
Programming Cooperativity in Self-Assembled Polynuclear Complexes Cooperativity: A Mise au Point. Angew. Chem., Int. Ed. 2011, 50, 1762−
in Solution. Chem. - Eur. J. 2009, 15, 12702−12718. 1768.
(655) Roman, M.; Cannizzo, C.; Pinault, T.; Isare, B.; Andrioletti, B.; (674) Mayoral, M. J.; Rest, C.; Stepanenko, V.; Schellheimer, J.;
van der Schoot, P.; Bouteiller, L. Supramolecular Balance: Using Albuquerque, R. Q.; Fernández, G. Cooperative Supramolecular
Cooperativity To Amplify Weak Interactions. J. Am. Chem. Soc. 2010, Polymerization Driven by Metallophilic Pd···Pd Interactions. J. Am.
132, 16818−16824. Chem. Soc. 2013, 135, 2148−2151.
(656) Chung, M. K.; Lee, S. J.; Waters, M. L.; Gagné, M. R. Self- (675) Fernández, G.; Stolte, M.; Stepanenko, V.; Würthner, F.
Assembled Multi-Component Catenanes: The Effect of Multivalency Cooperative Supramolecular Polymerization: Comparison of Different
and Cooperativity on Structure and Stability. J. Am. Chem. Soc. 2012, Models Applied on the Self-Assembly of Bis(merocyanine) Dyes. Chem.
134, 11430−11443. - Eur. J. 2013, 19, 206−217.
(657) Ercolani, G. Assessment of Cooperativity in Self-Assembly. J. (676) Ke, C.; Smaldone, R. A.; Kikuchi, T.; Li, H.; Davis, A. P.;
Am. Chem. Soc. 2003, 125, 16097−16103. Stoddart, J. F. Quantitative Emergence of Hetero[4]rotaxanes by
(658) Cao, D.; Wu, J. Theoretical Study of Cooperativity in Template-Directed Click Chemistry. Angew. Chem., Int. Ed. 2013, 52,
Multivalent Polymers for Colloidal Stabilization. Langmuir 2005, 21, 381.
9786−9791. (677) Jena, H. S. Effect of Cooperative Noncovalent Interactions on
(659) Kato, K.; Schneider, H. Cooperativity and Selectivity in
the Solid State Heterochiral Self-Assembly: The Concepts of Isotactic
Chemomechanical Polyethylenimine Gels. Langmuir 2007, 23,
and Syndiotactic Arrangements in Coordination Complex. Inorg. Chim.
10741−10745.
Acta 2014, 410, 156−170.
(660) Prohens, R.; Portell, A.; Puigjaner, C.; Tomàs, S.; Fujii, K.;
(678) Aparicio, F.; Sánchez, L. Thermodynamics of the Helical,
Harris, K. D. M.; Alcobé, X.; Font-Bardia, M.; Barbas, R. Cooperativity
Supramolecular Polymerization of Linear Self-Asembling Molecules:
in Solid-State Squaramides. Cryst. Growth Des. 2011, 11, 3725−3730.
Influence of Hydrogen Bonds and π Stacking. Chem. - Eur. J. 2013, 19,
(661) Widanapathirana, L.; Zhao, Y. Aromatically Functionalized
Cyclic Tricholate Macrocycles: Aggregation, Transmembrane Pore 10482−10486.
(679) Sun, H.; Navarro, C.; Hunter, C. A. Influence of Non-covalent
Formation, Flexibility, and Cooperativity. J. Org. Chem. 2012, 77, 4679−
4687. Preorganization on Supramolecular Effective Molarities. Org. Biomol.
(662) Rekharsky, M. V.; Yamamura, H.; Kawai, M.; Osaka, I.; Arakawa, Chem. 2015, 13, 4981−4992.
R.; Sato, A.; Ko, Y. H.; Selvapalam, N.; Kim, K.; Inoue, Y. Sequential (680) Chan, A. K.; Wong, K. M.; Yam, V. W. Supramolecular Assembly
Formation of a Ternary Complex among Dihexylammonium, of Isocyanorhodium(I) Complexes: An Interplay of Rhodium(I)···
Cucurbit[6]uril, and Cyclodextrin with Positive Cooperativity. Org. Rhodium(I) Interactions, Hydrophobic−Hydrophobic Interactions,
Lett. 2006, 8, 815−818. and Host−Guest Chemistry. J. Am. Chem. Soc. 2015, 137, 6920−6931.
(663) Leclercq, L.; Noujeim, N.; Sanon, S. H.; Schmitzer, A. R. Study of (681) Hu, W.; Hu, W.; Liu, Y. A.; Li, J.; Jiang, B.; Wen, K. Negative
the Supramolecular Cooperativity in the Multirecognition Mechanism Cooperativity in the Binding of Imidazolium and Viologen Ions to a
of Cyclodextrins/Cucurbituril/Disubstituted Diimidazolium Bromides. Pillar[5]arene-Crown Ether Fused Host. Org. Lett. 2015, 17, 2940−
J. Phys. Chem. B 2008, 112, 14176−14184. 2943.
(664) Sato, H.; Tashiro, K.; Shinmori, H.; Osuka, A.; Murata, Y.; (682) Beyeh, N. K.; Ala-Korpi, A.; Pan, F.; Jo, H. H.; Anslyn, E. V.;
Komatsu, K.; Aida, T. Positive Heterotropic Cooperativity for Selective Rissanen, K. Cooperative Binding of Divalent Diamides by N-Alkyl
Guest Binding via Electronic Communications through a Fused Zinc Ammonium Resorcinarene Chlorides. Chem. - Eur. J. 2015, 21, 9556−
Porphyrin Array. J. Am. Chem. Soc. 2005, 127, 13086−13087. 9562.
(665) Ménand, M.; de Beaumais, S. A.; Chamoreau, L.; Derat, E.; (683) Mäkelä, T.; Kalenius, E.; Rissanen, K. Cooperatively Enhanced
Blanchard, S.; Zhang, Y.; Bouteiller, L.; Sollogoub, M. Solid-State Ion Pair Binding with a Hybrid Receptor. Inorg. Chem. 2015, 54, 9154−
Hierarchical Cyclodextrin-Based Supramolecular Polymer Constructed 9165.
by Primary, Secondary, and Tertiary Azido Interactions. Angew. Chem., (684) Yokoyama, S.; Hirose, T.; Matsuda, K. Effects of Alkyl Chain
Int. Ed. 2014, 53, 7238−7242. Length and Hydrogen Bonds on the Cooperative Self-Assembly of 2-
(666) Howe, E. N. W.; Bhadbhade, M.; Thordarson, P. Cooperativity Thienyl-Type Diarylethenes at a Liquid/Highly Oriented Pyrolytic
and Complexity in the Binding of Anions and Cations to a Tetratopic Graphite (HOPG) Interface. Chem. - Eur. J. 2015, 21, 13569−13576.
Ion-Pair Host. J. Am. Chem. Soc. 2014, 136, 7505−7516. (685) Kulkarni, C.; Bejagam, K. K.; Senanayak, S. P.; Narayan, K. S.;
(667) Bejagam, K. K.; Fiorin, G.; Klein, M. L.; Balasubramanian, S. Balasubramanian, S.; George, S. J. Dipole-Moment-Driven Cooperative
Supramolecular Polymerization of Benzene-1, 3, 5-tricarboxamide: A Supramolecular Polymerization. J. Am. Chem. Soc. 2015, 137, 3924−
Molecular Dynamics Simulation Study. J. Phys. Chem. B 2014, 118, 3932.
5218−5228. (686) Wang, X.; Miller, D. S.; Bukusoglu, E.; de Pablo, J. J.; Abbott, N.
(668) Cheuk, K. K. L.; Lam, J. W. Y.; Lai, L. M.; Dong, Y.; Tang, B. Z. L. Topological Defects in Liquid Crystals as Templates for Molecular
Syntheses, Hydrogen-Bonding Interactions, Tunable Chain Helicities, Self-assembly. Nat. Mater. 2015, 15, 106.
and Cooperative Supramolecular Associations and Dissociations of (687) Xu, J.; Chen, L.; Zhang, X. How to Make Weak Noncovalent
Poly(Phenylacetylene)s Bearing l-Valine Pendants: Toward the Interactions Stronger. Chem. - Eur. J. 2015, 21, 11938−11946.

2822 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(688) Akhuli, B.; Ghosh, P. Selective Recognition and Extraction of (707) Motta, A.; Fragalà, I. L.; Marks, T. J. Proximity and Cooperativity
KBr via Cooperative Interactions with a Urea Functionalized Crown Effects in Binuclear d0 Olefin Polymerization Catalysis. Theoretical
Ether Dual-host. Chem. Commun. 2015, 51, 16514. Analysis of Structure and Reaction Mechanism. J. Am. Chem. Soc. 2009,
(689) Seth, S. K.; Sarkar, D.; Jana, A. D.; Kar, T. On the Possibility of 131, 3974−3984.
Tuning Molecular Edges To Direct Supramolecular Self-Assembly in (708) Mazzacano, T. J.; Mankad, N. P. Base Metal Catalysts for
Coumarin Derivatives through Cooperative Weak Forces: Crystallo- Photochemical C−H Borylation that Utilize Metal−Metal Coopera-
graphic and Hirshfeld Surface Analyses. Cryst. Growth Des. 2011, 11, tivity. J. Am. Chem. Soc. 2013, 135, 17258−17261.
4837−4849. (709) Mankad, N. P. Non-Precious Metal Catalysts for C−H
(690) Hunter, C. A.; Misuraca, M. C.; Turega, S. M. Influence of H- Borylation Enabled by Metal−Metal Cooperativity. Synlett 2014, 25,
Bond Strength on Chelate Cooperativity. J. Am. Chem. Soc. 2011, 133, 1197−1201.
20416−20425. (710) Cantekin, S.; ten Eikelder, H. M. M.; Markvoort, A. J.; Veld, M.
(691) Hogben, H. J.; Sprafke, J. K.; Hoffmann, M.; Pawlicki, M.; A. J.; Korevaar, P. A.; Green, M. M.; Palmans, A. R. A.; Meijer, E. W.
Anderson, H. L. Stepwise Effective Molarities in Porphyrin Oligomer Consequences of Cooperativity in Racemizing Supramolecular Systems.
Complexes: Preorganization Results in Exceptionally Strong Chelate Angew. Chem., Int. Ed. 2012, 51, 6426−6431.
Cooperativity. J. Am. Chem. Soc. 2011, 133, 20962−20969. (711) Fitz, B.; Andjelić, S.; Mijovic, J. Reorientational Dynamics and
(692) Jiang, W.; Nowosinski, K.; Löw, N. L.; Dzyuba, E. V.; Klautzsch, Intermolecular Cooperativity of Reactive Polymers. 1. Model Epoxy−
F.; Schäfer, A.; Huuskonen, J.; Rissanen, K.; Schalley, C. A. Chelate Amine Systems. Macromolecules 1997, 30, 5227−5238.
Cooperativity and Spacer Length Effects on the Assembly Thermody- (712) Xu, X.; Pooi, B.; Hirao, H.; Hong, S. H. CH−π and CF−π
namics and Kinetics of Divalent Pseudorotaxanes. J. Am. Chem. Soc. Interactions Lead to Structural Changes of N-Heterocyclic Carbene
2012, 134, 1860−1868. Palladium Complexes. Angew. Chem., Int. Ed. 2014, 53, 1283−1287.
(693) Nowosinski, K.; von Krbek, L. K. S.; Traulsen, N. L.; Schalley, C. (713) Broussard, M. E.; Juma, B.; Train, S. G.; Peng, W. J.; Laneman, S.
A. Thermodynamic Analysis of Allosteric and Chelate Cooperativity in A.; Stanley, G. G. A Bimetallic Hydroformylation Catalyst: High
Di- and Trivalent Ammonium/Crown-Ether Pseudorotaxanes. Org. Regioselectivity and Reactivity through Homobimetallic Cooperativity.
Lett. 2015, 17, 5076−5079. Science 1993, 260, 1784−1788.
(694) Traulsen, N. L.; Traulsen, C. H.-H.; Deutinger, P. M.; Müller, S.; (714) Suss-Fink, G. Cooperativity in Rh2 Complexes: High Catalytic
Schmidt, D.; Linder, I.; Schalley, C. A. Chelate Cooperativity Effects on Activity and High Regioselectivity in the Hydroformylation of Olefins.
the Formation of Di- and Trivalent Pseudo[2]rotaxanes with Angew. Chem., Int. Ed. Engl. 1994, 33, 67−69.
Diketopiperazine Threads and Tetralactam Wheels. Org. Biomol. (715) Chiang, K. P.; Bellows, S. M.; Brennessel, W. W.; Holland, P. L.
Multimetallic Cooperativity in Activation of Dinitrogen at Iron−
Chem. 2015, 13, 10881.
Potassium Sites. Chem. Sci. 2014, 5, 267−274.
(695) Knowles, R. R.; Jacobsen, E. N. Attractive Noncovalent
(716) Al-Amin, M.; Roth, K. E.; Blum, S. A. Mechanistic Studies of
Interactions in Asymmetric Catalysis: Links between Enzymes and
Gold and Palladium Cooperative Dual-Catalytic Cross-Coupling
Small Molecule Catalysts. Proc. Natl. Acad. Sci. U. S. A. 2010, 107,
Systems. ACS Catal. 2014, 4, 622−629.
20678−20685.
(717) Cooper, O.; Camp, C.; Pécaut, J.; Kefalidis, C. E.; Maron, L.;
(696) Xu, H.; Zuend, S. J.; Woll, M. G.; Tao, Y.; Jacobsen, E. N.
Gambarelli, S.; Mazzanti, M. Multimetallic Cooperativity in Uranium-
Asymmetric Cooperative Catalysis of Strong Brønsted Acid−Promoted
Mediated CO2 Activation. J. Am. Chem. Soc. 2014, 136, 6716−6723.
Reactions Using Chiral Ureas. Science 2010, 327, 986−990. (718) Jindal, G.; Kisan, H. K.; Sunoj, R. B. Mechanistic Insights on
(697) Schreiner, P. R. Cooperativity Tames Reactive Catalysts. Science Cooperative Catalysis through Computational Quantum Chemical
2010, 327, 965−966. Methods. ACS Catal. 2015, 5, 480−503.
(698) Schindler, C. S.; Jacobsen, E. N. A New Twist on Cooperative (719) Raynal, M.; Ballester, P.; Vidal-Ferran, A.; van Leeuwen, P. W. N.
Catalysis. Science 2013, 340, 1052−1053. M. Supramolecular Catalysis. Part 1: Noncovalent Interactions as a Tool
(699) Krautwald, S.; Sarlah, D.; Schafroth, M. A.; Carreira, E. M. for Building and Modifying Homogeneous Catalysts. Chem. Soc. Rev.
Enantio- and Diastereodivergent Dual Catalysis: α-Allylation of 2014, 43, 1660−1733.
Branched Aldehydes. Science 2013, 340, 1065−1068. (720) Zhang, Z.; Lippert, K. M.; Hausmann, H.; Kotke, M.; Schreiner,
(700) Carrillo, R.; Feher-Voelger, A.; Martín, T. Enantioselective P. R. Cooperative Thiourea−Brønsted Acid Organocatalysis: Enantio-
Cooperativity between Intra-Receptor Interactions and Guest Binding: selective Cyanosilylation of Aldehydes with TMSCN. J. Org. Chem.
Quantification of Reinforced Chiral Recognition. Angew. Chem., Int. Ed. 2011, 76, 9764−9776.
2011, 50, 10616−10620. (721) Kim, H. Y.; Oh, K. Highly Diastereo- and Enantioselective Aldol
(701) Kanai, M.; Kato, N.; Ichikawa, E.; Shibasaki, M. Power of Reaction of Methyl r-Isocyanoacetate: A Cooperative Catalysis
Cooperativity: Lewis Acid−Lewis Base Bifunctional Asymmetric Approach. Org. Lett. 2011, 13, 1306−1309.
Catalysis. Synlett 2005, 10, 1491−1508. (722) Weil, T.; Kotke, M.; Kleiner, C. M.; Schreiner, P. R. Cooperative
(702) Raup, D. E. A.; Cardinal-David, B.; Holte, D.; Scheidt, K. A. Brønsted Acid-Type Organocatalysis: Alcoholysis of Styrene Oxides.
Cooperative Catalysis by Carbenes and Lewis Acids in a Highly Org. Lett. 2008, 10, 1513−1516.
Stereoselective Route to γ-Lactams. Nat. Chem. 2010, 2, 766−771. (723) Perdriau, S.; Zijlstra, D. S.; Heeres, H. J.; de Vries, J. G.; Otten, E.
(703) Patil, N. T. Merging Metal and N-Heterocyclic Carbene A Metal−Ligand Cooperative Pathway for Intermolecular Oxa-Michael
Catalysis: On the Way to Discovering Enantioselective Organic Additions to Unsaturated Nitriles. Angew. Chem., Int. Ed. 2015, 54,
Transformations. Angew. Chem., Int. Ed. 2011, 50, 1759−1761. 4236−4240.
(704) Potier, J.; Menuel, S.; Fournier, D.; Fourmentin, S.; Woisel, P.; (724) Gan, Q.; Ronson, T. K.; Vosburg, D. A.; Thoburn, J. D.;
Monflier, E.; Hapiot, F. Cooperativity in Aqueous Organometallic Nitschke, J. R. Cooperative Loading and Release Behavior of a Metal-
Catalysis: Contribution of Cyclodextrin-Substituted Polymers. ACS Organic Receptor. J. Am. Chem. Soc. 2015, 137, 1770−1773.
Catal. 2012, 2, 1417−1420. (725) Schmidt, V. A.; Hoyt, J. M.; Margulieux, G. W.; Chirik, P. J.
(705) Milsmann, C.; Turner, Z. R.; Semproni, S. P.; Chirik, P. J. Azo Cobalt-Catalyzed [2π+2π] Cycloadditions of Alkenes: Scope, Mecha-
NN Bond Cleavage with a Redox-Active Vanadium Compound nism and Elucidation of Electronic Structure of Catalytic Intermediates.
Involving Metal−Ligand Cooperativity. Angew. Chem., Int. Ed. 2012, 51, J. Am. Chem. Soc. 2015, 137, 7903−7914.
5386−5390. (726) Li, H.; Marks, T. J. Nuclearity and Cooperativity Effects in
(706) Ho, J. H. H.; Choy, S. W. S.; Macgregor, S. A.; Messerle, B. A. Binuclear Catalysts and Cocatalysts for Olefin Polymerization. Proc.
Cooperativity in Bimetallic Dihydroalkoxylation Catalysts Built on Natl. Acad. Sci. U. S. A. 2006, 103, 15295−15302.
Aromatic Scaffolds: Significant Rate Enhancements with a Rigid (727) Rummelt, S. M.; Radkowski, K.; Roşca, D.; Fürstner, A.
Anthracene Scaffold. Organometallics 2011, 30, 5978−5984. Interligand Interactions Dictate the Regioselectivity of trans-Hydro-

2823 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

metalations and Related Reactions Catalyzed by [Cp*RuCl].Hydrogen Weight Gelators: Elucidating the Principles of Gelation Based on
Bonding to a Chloride Ligand as a Steering Principle in Catalysis. J. Am. Gelator Solubility and a Cooperative Self-Assembly Model. J. Am. Chem.
Chem. Soc. 2015, 137, 5506−5519. Soc. 2008, 130, 9113−9121.
(728) Umadevi, D.; Panigrahi, S.; Sastry, G. N. Noncovalent (749) Rest, C.; Mayoral, M. J.; Fucke, K.; Schellheimer, J.; Stepanenko,
Interaction of Carbon Nanostructures. Acc. Chem. Res. 2014, 47, V.; Fernández, G. Self-Assembly and (Hydro)gelation Triggered by
2574−2581. Cooperative π−π and Unconventional CH···X Hydrogen Bonding
(729) Remya, K.; Suresh, C. H. Non-covalent Intermolecular Carbon− Interactions. Angew. Chem., Int. Ed. 2014, 53, 700−705.
Carbon Interactions in Polyynes. Phys. Chem. Chem. Phys. 2015, 17, (750) Song, G.; Zhang, L.; He, C.; Fang, D.; Whitten, P. G.; Wang, H.
27035−27044. Facile Fabrication of Tough Hydrogels Physically Cross-Linked by
(730) Umadevi, D.; Sastry, G. N. Quantum Mechanical Study of Strong Cooperative Hydrogen Bonding. Macromolecules 2013, 46,
Physisorption of Nucleobases on Carbon Materials: Graphene versus 7423−7435.
Carbon Nanotubes. J. Phys. Chem. Lett. 2011, 2, 1572−1576. (751) Carrillo, R.; López-Rodríguez, M.; Martín, V. S.; Martín, T.
(731) Umadevi, D.; Sastry, G. N. Impact of the Chirality and Curvature Crystal Structures of Self-Assembled Nanotubes from Flexible Macro-
of Carbon Nano Structure on their Interaction with Aromatics and cycles by Weak Interactions. CrystEngComm 2010, 12, 3676−3683.
Aminoacids. ChemPhysChem 2013, 14, 2570−2578. (752) Tokmachev, A. M.; Dronskowski, R. Hydrogen-Bond Networks
(732) Umadevi, D.; Sastry, G. N. Feasibility of Carbon Nanomaterials in Finite Ice Nanotubes. J. Comput. Chem. 2011, 32, 99−105.
as Gas Sensors: A Computational Investigation. Curr. Sci. 2014, 106, (753) Oh, H.; Sim, J.; Ju, S. Binding Affinities and Thermodynamics of
1224−1234. Noncovalent Functionalization of Carbon Nanotubes with Surfactants.
(733) Panigrahi, S.; Sastry, G. N. Reducing Polyaromatic Hydro- Langmuir 2013, 29, 11154−11162.
carbons: The Capability and Capacity of Lithium. RSC Adv. 2014, 4, (754) Jin, W.; Fukushima, T.; Niki, M.; Kosaka, A.; Ishii, N.; Aida, T.
14557−14563. Self-Assembled Graphitic Nanotubes with One-Handed Helical Arrays
(734) Hussain, M. A.; Soujanya, Y.; Sastry, G. N. Computational of a Chiral Amphiphilic Molecular Graphene. Proc. Natl. Acad. Sci. U. S.
Design of Functionalized Imidazolate Linkers of Zeolitic Imidazolate A. 2005, 102, 10801−10806.
Frameworks for Enhanced CO2 Adsorption. J. Phys. Chem. C 2015, 119, (755) Datta, A.; Pati, S. K. Dipolar Interactions and Hydrogen Bonding
23607−23618. in Supramolecular Aggregates: Understanding Cooperative Phenomena
(735) Umadevi, D.; Sastry, G. N. Graphane Versus Graphene: A for 1st Hyperpolarizability. Chem. Soc. Rev. 2006, 35, 1305−1323.
Computational Investigation of the Interaction of Nucleobases, (756) Gray, T.; Kim, T.; Knorr, D. B., Jr.; Luo, J.; Jen, A. K.-Y.;
Aminoacids, Heterocycles, Small Molecules (CO2, H2O, NH3, CH4, Overney, R. M. Mesoscale Dynamics and Cooperativity of Networking
H2), Metal ions and Onium Ions. Phys. Chem. Chem. Phys. 2015, 17, Dendronized Nonlinear Optical Molecular Glasses. Nano Lett. 2008, 8,
30260−30269. 754−759.
(736) Hussain, M. A.; Vijay, D.; Sastry, G. N. Buckybowls as (757) Knorr, D. B., Jr.; Benight, S. J.; Krajina, B.; Zhang, C.; Dalton, L.
Adsorbents for CO2, CH4, and C2H2: Binding and Structural Insights
R.; Overney, R. M. Nanoscale Phase Analysis of Molecular
from Computational Study. J. Comput. Chem. 2016, 37, 366.
Cooperativity and Thermal Transitions in Dendritic Nonlinear Optical
(737) Chou, C.; Buehler, M. J. Bond Energy Effects on Strength,
Glasses. J. Phys. Chem. B 2012, 116, 13793−13805.
Cooperativity and Robustness of Molecular Structures. Interface Focus
(758) Suponitsky, K. Y.; Masunov, A. E. Supramolecular Step in
2011, 1, 734−743.
Design of Nonlinear Optical Materials: Effect of π···π Stacking
(738) Qin, Z.; Buehler, M. Bioinspired Design of Functionalised
Aggregation on Hyperpolarizability. J. Chem. Phys. 2013, 139,
Graphene. Mol. Simul. 2012, 38, 695−703.
(739) Knowles, T. P. J.; Oppenheim, T. W.; Buell, A. K.; Chirgadze, D. 094310−16.
Y.; Welland, M. E. Nanostructured Films from Hierarchical Self- (759) Mafé, S.; Manzanares, J. A.; Reiss, H. Nanoscale Switch Based on
Assembly of Amyloidogenic Proteins. Nat. Nanotechnol. 2010, 5, 204− Interacting Molecular Dipoles: Cooperativity can Improve the Device
207. Characteristics. J. Appl. Phys. 2011, 109, No. 044302.
(740) Buehler, M. J. Strength in Numbers. Nat. Nanotechnol. 2010, 5, (760) Brunelli, N. A.; Didas, S. A.; Venkatasubbaiah, K.; Jones, C. W.
172−174. Tuning Cooperativity by Controlling the Linker Length of Silica-
(741) Keten, S.; Buehler, M. J. Geometric Confinement Governs the Supported Amines in Catalysis and CO2 Capture. J. Am. Chem. Soc.
Rupture Strength of H-bond Assemblies at a Critical Length Scale. Nano 2012, 134, 13950−13953.
Lett. 2008, 8, 743−748. (761) Bouteiller, L.; Colombani, O.; Lortie, F.; Terech, P. Thickness
(742) Keten, S.; Xu, Z.; Ihle, B.; Buehler, M. J. Nanoconfinement Transition of a Rigid Supramolecular Polymer. J. Am. Chem. Soc. 2005,
Controls Stiffness, Strength and Mechanical Toughness of β-Sheet 127, 8893−8898.
Crystals in Silk. Nat. Mater. 2010, 9, 359−367. (762) White, S. L.; Smith, J. G.; Behl, M.; Jain, P. K. Cooperativity in a
(743) Nair, A. K.; Qin, Z.; Buehler, M. J. Cooperative Deformation of Nanocrystalline Solid-State Transition. Nat. Commun. 2013, 4, 2933.
Carboxyl Groups in Functionalized Carbon Nanotubes. Int. J. Solids (763) Düring, J.; Hölzer, A.; Kolb, U.; Branscheid, R.; Gröhn, F.
Struct. 2012, 49, 2418−2423. Supramolecular Organic−Inorganic Hybrid Assemblies with Tunable
(744) Compton, O. C.; Cranford, S. W.; Putz, K. W.; An, Z.; Brinson, Particle Size: Interplay of Three Noncovalent Interactions. Angew.
L. C.; Buehler, M. J.; Nguyen, S. T. Tuning the Mechanical Properties of Chem., Int. Ed. 2013, 52, 8742−8745.
Graphene Oxide Paper and Its Associated Polymer Nanocomposites by (764) Sikder, Md. D. H.; Gibbs-Davis, J. M. The Influence of Gap
Controlling Cooperative Intersheet Hydrogen Bonding. ACS Nano Length on Cooperativity and Rate of Association in DNA-Modified
2012, 6, 2008−2019. Gold Nanoparticle Aggregates. J. Phys. Chem. C 2012, 116, 11694−
(745) Qin, Z.; Buehler, M. J. Cooperative Deformation of Hydrogen 11701.
Bonds in Beta-Strands and Beta-Sheet Nanocrystals. Phys. Rev. E 2010, (765) Liu, J.; Lu, Y. Smart Nanomaterials Responsive to Multiple
82, No. 061906. Chemical Stimuli with Controllable Cooperativity. Adv. Mater. 2006, 18,
(746) Hu, X.; Xu, Z.; Gao, C. Multifunctional, Supramolecular, 1667−1671.
Continuous Artificial Nacre Fibres. Sci. Rep. 2012, 2, 767. (766) Busseron, E.; Ruff, Y.; Moulin, E.; Giuseppone, N. Supra-
(747) Hsu, S.; Lin, Y.; Chang, J.; Liu, Y.; Lin, H. Intramolecular molecular Self-Assemblies as Functional Nanomaterials. Nanoscale
Interactions of a Phenyl/Perfluorophenyl Pair in the Formation of 2013, 5, 7098−7140.
Supramolecular Nanofibers and Hydrogels. Angew. Chem., Int. Ed. 2014, (767) Liu, W.; Wang, Q.; Wang, Y.; Huang, Z.; Wang, D. Designed
53, 1921−1927. Self-Assemblies Based on Cooperative Noncovalent Interactions
(748) Hirst, A. R.; Coates, I. A.; Boucheteau, T. R.; Miravet, J. F.; Including Anion−π, Lone-Pair Electron−π and Hydrogen Bonding.
Escuder, B.; Castelletto, V.; Hamley, I. W.; Smith, D. K. Low-Molecular- RSC Adv. 2014, 4, 9339−9342.

2824 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825
Chemical Reviews Review

(768) Kumagai, T.; Hanke, F.; Gawinkowski, S.; Sharp, J.; Kotsis, K.;
Waluk, J.; Persson, M.; Grill, L. Controlling Intramolecular Hydrogen
Transfer in a Porphycene Molecule with Single Atoms or Molecules
Located Nearby. Nat. Chem. 2013, 6, 41−46.
(769) Suttipong, M.; Grady, B. P.; Striolo, A. Self-Assembled
Surfactants on Patterned Surfaces: Confinement and Cooperative
Effects on Aggregate Morphology. Phys. Chem. Chem. Phys. 2014, 16,
16388−16398.
(770) Xavier, P.; Bose, S. Non-Equilibrium Segmental Dynamics
Driven by Multiwall Carbon Nanotubes in PS/PVME Blends. Phys.
Chem. Chem. Phys. 2014, 16, 9309−9316.
(771) Bharati, A.; Xavier, P.; Kar, G. P.; Madras, G.; Bose, S.
Nanoparticle-Driven Intermolecular Cooperativity and Miscibility in
Polystyrene/Poly(vinyl methyl ether) Blends. J. Phys. Chem. B 2014,
118, 2214−2225.
(772) Wei, X.; Nangreave, J.; Liu, Y. Uncovering the Self-Assembly of
DNA Nanostructures by Thermodynamics and Kinetics. Acc. Chem. Res.
2014, 47, 1861−1870.
(773) The Fourth Paradigm: Data-Intensive Scientific Discovery; Hey, T.,
Tansley, S., Tolle, K., Eds.; Microsoft Research, 2009.
(774) Magoulas, R.; Lorica, B. Introduction to Big Data in Release 2.0;
O’Reilly Media: Sebastopol, CA, 2009.
(775) Snijders, C.; Matzat, U.; Reips, U.-D. ″Big Data″: Big Gaps of
Knowledge in the Field of Internet Science. Int. J. Internet Sci. 2012, 7,
1−5.

2825 DOI: 10.1021/cr500344e


Chem. Rev. 2016, 116, 2775−2825

You might also like