Vous êtes sur la page 1sur 11

FULL PAPER

DOI: 10.1002/chem.201200497

Supramolecular Binding Thermodynamics by Dispersion-Corrected Density


Functional Theory

Stefan Grimme*[a]

Abstract: The equilibrium association um solvation model (based on DFT A semilocal (TPSS) and a hybrid densi-
free enthalpies DGa for typical supra- data) provides solvation free enthalpies ty functional (PW6B95) have been
molecular complexes in solution are and the remaining ro-vibrational en- tested. Although the DGa values result
calculated by ab initio quantum chemi- thalpic/entropic contributions are ob- as a sum of individually large terms
cal methods. Ten neutral and three pos- tained from harmonic frequency calcu- with opposite sign (DE vs. solvation
itively charged complexes with experi- lations. Low-lying vibrational modes and entropy change), the approach
mental DGa values in the range 0 to are treated by a free-rotor approxima- provides unprecedented accuracy for
21 kcal mol1 (on average 6 kcal tion. The accurate account of London DGa values with errors of only 2 kcal
mol1) are investigated. The theoretical dispersion interactions is mandatory mol1 on average. Relative affinities
approach employs a (nondynamic) with contributions in the range 5 to for different guests inside the same
single-structure model, but computes 60 kcal mol1 (up to 200 % of DE). host are always obtained correctly. The
the various energy terms accurately Inclusion of three-body dispersion ef- procedure is suggested as a predictive
without any special empirical adjust- fects improves the results considerably. tool in supramolecular chemistry and
ments. Dispersion corrected density can be applied routinely to semirigid
functional theory (DFT-D3) with ex- systems with 300400 atoms. The vari-
Keywords: density functional calcu-
tended basis sets (triple-z and quadru- ous contributions to binding and en-
lations dispersion corrections
ple-z quality) is used to determine thalpyentropy compensations are dis-
noncovalent interactions theoreti-
structures and gas-phase interaction en- cussed.
cal chemistry thermodynamics
ergies (DE), the COSMO-RS continu-

Introduction on the binding mechanisms prevails.[6, 7] One of the reasons


for this is that already the smallest experimentally consid-
The noncovalent interactions between atoms and molecules ered systems often consist of a hundred or more atoms and
play an important role in structural biology and supramolec- that solvation as well as entropic effects play an important
ular chemistry. They control the structures of proteins and role in the binding. These three factors contribute to large
DNA, hostguest systems, enzymesubstrate binding, anti- and often unaffordable computational costs.
genantibody recognition, or the orientation of molecules It is shown in this work for the first time in a systematic
on surfaces or in molecular films.[1, 2] Because of their ubiqui- study, that the free enthalpy of association of a host and
tous role in diverse fields, the investigation and understand- a guest molecule to form a supramolecular complex in solu-
ing of these weak interactions has become one of the major tion (DGa, equilibrium association constant Ka) can be com-
goals of modern chemistry. Noncovalent interactions[35] puted with reasonable accuracy from well-established, stan-
form the basis of what is usually called a supramolecular dard quantum chemical methods without any special adjust-
interaction. However, quantitative description of its ther- ments. This has some perspective for an ab initio computa-
modynamics by means of ab initio electronic structure tion of absolute proteinligandACHTUNGRE(drug) binding constants. In
theory is not well developed and a phenomenological view a pilot study on proteinligand model complexes, including
also the use of accurate wave-function-based methods,[8] we
have already shown that quantum chemical methods based
on London dispersion[9, 10] corrected density functional
[a] Prof. Dr. S. Grimme
Mulliken Center for Theoretical Chemistry theory (DFT-D3[11]) provide very accurate interaction ener-
Institut fr Physikalische und Theoretische Chemie gies for large complexes (for recent benchmarks on smaller
Universitt Bonn, Beringstr. 4, 53115 Bonn (Germany) complexes see references [12, 13]). The typical DFT-D3
E-mail: grimme@thch.uni-bonn.de error for the basic gas-phase interaction energy DE has
Supporting information for this article is available on the WWW been estimated to be < 5 % in several studies,[8, 13, 14] which
under http://dx.doi.org/10.1002/chem.201200497. It contains Cartesian
coordinates of all molecules and complexes, all thermodynamic data
would enable an accurate prediction of total DGa values
and a detailed description of the SCC-DFTB-D3H and PM6-D3H within 12 kcal mol1 if the other contributions (solvation
methods used for semi-empirical frequency calculations. and entropy change[15]) could be obtained within similar

Chem. Eur. J. 2012, 00, 0 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim &1&
These are not the final page numbers!
error limits. That this is indeed possible will be shown These systems cover a broad range of typical supramolec-
herein. The major advantage of such an ab initio treatment ular interactions ranging from nonpolar-dispersion-dominat-
of supramolecular interactions as opposed to empirical ed ones for complexes 1 and 2 and p-stacked structures in 3
force-field methods (for two examples see referen- and 4 to hydrogen bonding in 5 and 7 b and cationdipolar
ces [16, 17]) is that systems of any composition including binding modes in 6 and 7 a. For each of the six hosts, two
transition metals can be studied and that an unbiased pic- complexes with different guest molecules are investigated.
ture and understanding of the fundamental physical interac- This should shed light on the ability of the approach to cal-
tion mechanisms can be obtained. For a critical evaluation culate accurate relative affinities, which is in some applica-
of force-field methods for noncovalent interactions see ref- tions even more important than the absolute magnitude of
erence [18]. DGa. The benzene dimer in water 1[24] is included as a small
We have investigated the following systems (see Figure 1): system for which accurate gas-phase interaction energies are
a benzene dimer (1), two tweezer complexes with tetra- known theoretically[2528] and to demonstrate that the ap-
cyanoquinone (TCNQ) and 1,4-dicyanobenzene (2 a and 2 b proach is completely general and works for small as well as
in CHCl3) investigated in the group of Klrner,[19] two extended complexes. It also represents a limiting case with
pincer complexes of organic p systems (3 a and 3 b with a very small (near zero) binding affinity.
guests denoted as 3 c and 3 d in CH2Cl2) that have been in- Experimentally observed DGa values for supramolecular
vestigated recently also theoretically,[20] fullerenes C60 and complexes consisting of a few hundred atoms are typically
C70 in a so-called buckycatcher (4 a and 4 b in toluene) in the range 1 to 10 kcal mol1.[2] Complexes 7 a and 7 b
from reference [21], an amide macrocycle (mcycle) that with values of 21 and 14 kcal mol1, respectively, repre-
binds benzoquinone (BQ) and glycine anhydride (GLH) sent extremes. In any case this is almost an order of magni-
from the work of Hunter et al.[22] (5 a and 5 b in CHCl3) and tude less than the computed DE values for complexes of this
the cucurbit[6]uril (CB6) cation complexes of butylammoni- size.[8, 21, 20] This indicates that solvent and entropic correc-
um (BuNH4) and propylammonium (PrNH4), respectively, tions are large and positive and are effectively quenching
from the work of Mock and Shih[23] (6 a and 6 b in an 1:1 the many, in total rather strong intermolecular, noncovalent
mixture of formic acid and water). The cucurbit[7]uril interactions as noted already many years ago.[15] This situa-
(CB7) host forms an extremely stable inclusion complex 7 a tion in which a small quantity results from addition of sever-
with the dicationic ferrocene derivative bis(trimethylammo- al large terms with opposite signs is the worst-case scenario
niomethyl)ferrocene (FECP) with an equilibrium associa- for any nonempirical approach. Systematic error cancella-
tion constant of 3  1015 m1, which represents some kind of tion cannot be expected and each term individually must be
world record for synthetic systems in water.[17] The corre- computed rather accurately to obtain reliable DGa values.
sponding neutral complex 7 b with 1-hydroxyadamantane With this in mind one could consider this project successful
(ADOH) as guest in water from the same work is investigat- already if the order of magnitude and affinity trends for dif-
ed here for comparison. ferent complexes are correct. How difficult the problem is

Figure 1. Gas-phase structure of investigated complexes.

&2& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 0

These are not the final page numbers!


Supramolecular Binding Thermodynamics
FULL PAPER
has been exemplified in a DFT study for the binding of tems it can be calculated now routinely by well-established dispersion-
a neutral guest inside a charged [2]pseudorotaxane with ex- corrected DFT (atom pair-wise DFT-D3[11, 41] or density-dependent DFT-
nonlocal[42, 43] (NL) variants). Only DFT-D3 is considered in this work
perimental values of DHa = 4.9 kcal mol1 and DGa = (for a recent overview of various other dispersion-corrected DFT meth-
1.6 kcal mol1.[29] Very strongly deviating theoretical DHa ods[4448] see reference [49]). We have tested two different density func-
values between 28 and + 23 kcal mol1 (in MeCN using tionals (the TPSS meta-GGA[50] and the PW6B95 meta-hybrid[51]) to
a continuum solvation model, with and without counterions) obtain an estimate for the error of DE. The DGRRHO contribution is com-
puted by lower-level quantum chemical methods (SCC-DFTB[52, 53] and
have been computed using various (dispersion-uncorrected)
PM6[54, 55]) in order to make the entire approach routinely applicable to
density functional methods.[29] Before we show that one can systems with about 300400 atoms (although this is not a prerequisite
significantly improve the accuracy, the general theoretical and these semi-empirical MO methods could be avoided entirely).
approach is outlined highlighting a few theoretical aspects No special adjustments for the systems considered here are made to the
which require special attention. published procedures that are employed to compute the individual parts
in Equation (1). It should be stressed that the theoretical approach fol-
lowed is ab initio in a sense that all physically relevant terms are consid-
ered and computed as accurately as possible, but are not adjusted empiri-
Computational Methods cally for the present purpose. However, the methods themselves necessar-
ily contain a few empirical parameters that are essential to treat large
General approach to free enthalpies of association in solution: The free systems (several hundreds of atoms) with reasonable accuracy. A similar
enthalpy of association of molecules A and B (e.g., host and guest) to approach as used here has been proposed very recently by Jacquemin
form a noncovalently bound complex C at temperature T in solvent X is and co-workers[20] and has been applied to 3 a and 3 b. These authors also
given by Equation (1) in which DE is the gas-phase association energy as employ DFT-D for the interaction energy, but use the PCM solvation
calculated in the super-molecule approach DE = E(C)E(A)E(B) (E model[35] and a relatively small AO basis set for structure optimization.
being the total electronic energies), GTRRHO is the sum of corrections from Three-body dispersion effects as well as the low-frequency mode problem
energy to free enthalpy in the rigid-rotor-harmonic-oscillator approxima- (see below) were not considered. However, good agreement with the ex-
tion (RRHO) also including zero-point-vibrational energy for each spe- perimental DGa values (errors of only 13 kcal mol1) has been obtained.
cies in the gas phase, and dGTsolv(X) is their corresponding free solvation The catcher complex 4 a was investigated previously by Zhao and Truhlar
enthalpy. using DFT without explicit dispersion corrections.[56] Beside DE and
DGRRHO values, only estimates for the solvation effects were given. For
DGa DE DGTRRHO DdGTsolv X 1 previous quantum chemical studies on molecular tweezer or clip
complexes see, for example, references [5761].
Equation (1) is based on the fundamental assumption that the compo-
In the following we use the short notation DGa (and similar for DHa and
nents in the association reaction are distinct chemical species with a well-
the other terms, i.e., drop the T and X indices), but note that these
defined molecular structure. It would easily be possible to apply the ap-
values refer to specific solvents and temperature for each of the systems.
proach for a few well-defined conformers (the corresponding quantities
The prefix D refers to differences regarding the reaction A + B!C, while
have to be thermodynamically averaged), but extremely flexible, highly
d indicates that the quantity implicitly represents a difference between
dynamic systems cannot be treated in this way (for the thermodynamic
thermodynamic states (i.e., gassolution).
treatment of very flexible, but smaller systems see reference [62]). Force-
The last term in Equation (1) is taken unmodified from the COSMO-RS field studies by Gilson and co-workers[17, 63] for similar complexes have
continuum solvation model,[3032] which is based on DFT (BP86,[33]/ shown that only the few lowest-lying conformers contribute and that the
TZVP[34]) single-point energy computations for the electrostatic solute herein pursued nondynamic, single-structure approach may introduce
solvent interaction energy (as expressed by the screening-charges through errors of merely about 1 kcal mol1. Note, in this context, that the present
the so-called s-profiles[30]) and including further molecular surface pro-
study represents a proof-of-principle and that thermodynamic averaging
portional corrections to directly yield free solvation energies. The corre-
over a few conformers is possible in principle also at an ab initio level.
sponding enthalpies dHsolv are computed from the temperature depend-
The focus here is clearly on the basic intermolecular contacts and the
ence of the free enthalpies, but are used here merely for analysis purpos-
electronic structure aspect, which is complementary to more dynamic
es and do not represent fundamental quantities. The dGsolv terms ob-
treatments.[17, 63, 64]
tained from COSMO-RS refer to the 1 mol gas/1 mol solvent frame and
can directly be used in Equation (1) to compute standard state data. The examples have been taken either from our previous work[21, 61] or are
Note that very strongly and specifically coordinating solvent molecules more or less randomly chosen from a standard textbooks or reviews on
can be included in a quantum chemical supermolecular system and supramolecular chemistry.[2, 7] No biased selection of complexes for which
should not be treated by COSMO-RS (which, however, is not considered the methods work well has been attempted. The only prerequisite was
here). The COSMO-RS model normally provides accurate dGsolv values a certain molecular rigidity in addition to the coverage of different bind-
with an estimated error of about 510 % (< 0.51 kcal mol1 for neutral ing types and a wide range of affinities.
systems[35, 36]). Because the error for the two other terms in Equation (1) Computation of (gas-phase) interaction energies: The species A, B, and
is of the same order of magnitude (or even somewhat larger) and be- C were fully optimized in the gas phase at the TPSS-D3/def2-TZVP level
cause other means to account for solvation effects quantitatively are still and these structures were also used for the COSMO-RS treatment. Scans
not routine, we take this continuum model as is and concentrate on an for low-lying conformers and pre-optimizations were performed at the
accurate evaluation of the DE and DGRRHO terms. A further advantage of semi-empirical level, but always only the lowest energy structure found
COSMO-RS is that solvent mixtures can be treated on equal footing. For was further considered for DFT-D3. The semilocal TPSS[50] functional
a thorough discussion of the current status and performance of this and represents a robust, well-tested DFT method which provides good molec-
related solvation models see also Cramer and Truhlar.[3739] The depend- ular structures[41] at low computational cost. For some years it has been
ency of the performance of COSMO-RS on the underlying DFT method the default functional for structure optimizations in our group (for the
(here: B86/TZVP//TPSS-D3/def2-TZVP) has recently been investigated most recent example see reference [65]).
in detail,[40] but is found to be very weak. The molecular electronic energies E were computed by dispersion-cor-
The gas-phase complexation energy DE is expected to be the dominant rected DFT given by Equation (2), in which EDFT is the (all-electron) KS-
contribution in Equation (1); for a further discussion of this point see DFT SCF energy for a particular density functional, E2 disp is the standard
below. It is very specific for the mode of interaction and its accurate com- atom pair-wise London dispersion energy from D3 theory[11] (using
putation is mandatory in order to obtain reliable results. For large sys- BeckeJohnson damping[41, 47, 48]), and E3
disp is a three-body dispersion term

Chem. Eur. J. 2012, 00, 0 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &3&
These are not the final page numbers!
S. Grimme

(of AxilrodTellerMutto type[66, 67]), which was calculated as described in proaches that require use of internal coordinates (and that have never
reference [11] using our freely available program dftd3.[68] been extended to large systems) see reference [82]. A related but compu-
tationally more involved alternative is the harmonic mode scanning
method, see reference [64].
E EDFT E2 3
disp Edisp 2
The vibrational entropy of an harmonic oscillator with frequency w is
given by Equation (3) in which k is Boltzmanns constant, h is Plancks
Note that this last term is independent of the chosen density functional
constant, T is the absolute temperature and R is the usual gas constant.
in our approach. These three-body corrections to the dispersion interac-
tion of small complexes (< 1020 atoms) are negligible for, for example,  
the S66 benchmark set[69] (average E3 1 hw
disp corrections are 0.06 kcal mol SV R  ln1  ehw=kT 3
1
for an average total interaction energy of 6 kcal mol ; similar results ke hw=kT  1
have been reported in references [70, 71]). However, as already noted in
reference [11], three-body dispersion effects become more important in Because the first term on the right-hand side of Equation (3) asymptoti-
larger and dense systems. As will be shown below for the first time, its cally approaches infinity for w!0 we replace the contribution of low-
account improves computed supramolecular interaction energies consid- lying modes to the entropy by a corresponding rotational entropy (quasi-
erably (the more commonly used three-body uncorrected energy E = RRHO approach). For any normal mode we compute the moment of in-
EDFT + E2 ertia m for a free-rotor with the same frequency [Eq. (4)].
disp is also discussed). Furthermore, for a consistent description of
small and concomitantly large complexes it is essential to describe the
asymptotic 1/R6 dependence of the interaction energy (in which R is h
m 4
the intermolecular distance) correctly also in the large distance regime, 8p2 w
which is not possible with for example, uncorrected, (semilocal) meta-
In order to define an effective moment of inertia and to restrict it to
hybrid functionals like M06-2X[72] (for more details see reference [13]).
a reasonable value, we introduce the average molecular moment of iner-
The relatively large systems considered here should be used in the future
tia Bav as a limiting value for very small w (large m) by taking the follow-
as benchmarks to investigate this point also for other dispersion correc-
ing reduced value from Equation (5).
tions in more detail.
The dependence of the results for DGa on the chosen density functional mBav
stems from the DE part. We investigate this by single-point calculations m0 5
m Bav
with the PW6B95[51] meta-hybrid functional developed by Zhao and
Truhlar, which has proven to be very accurate for a wide range of appli- The entropy for a low-frequency mode in this approximation is then
cations when dispersion corrections are properly taken into account.[13, 73] given by Equation (6).
In all geometry optimizations we apply the large def2-TZVP AO basis
set,[34, 74] which yields small basis set superposition errors (BSSE) on the " ( 1=2 )#
8p3 m0 kT
order of 5 % of DE for larger complexes as considered here. Computa- SR R 1=2 ln 6
h2
tionally demanding counterpoise corrections are hence not applied and
we prefer to investigate the basis set dependence with the huge def2-
QZVP basis set (def2-QZVP[74, 75] with discarded g- and f-functions on In order to continuously interpolate between the rotational and harmonic
non-hydrogen and hydrogen atoms, respectively; this is denoted in short vibrational approximations, SV and SR are combined by using a weighting
QZ in the following). In one case the full def2-QZVP and cc-pVQZ[76] function of w as given by Equation (7) in which w(w) is given by the
basis has been applied for convergence checking purposes and differen- Head-Gordon damping function [Eq. (8)],[83] with a = 4 (very similar re-
ces of only < 0.2 kcal mol1 have been observed. Remaining (tiny) BSSE sults have been obtained by a Fermi-weighting function that has also
effects have been absorbed into the short-range D3 parameters as deter- been tested).
mined in original work[11, 41] at a similar basis set level. Note that the QZ
basis already contains semidiffuse AO basis functions that are known to S wwSv 1  wwSR 7
be sufficient to describe molecular polarizeabilities at the GGA level.[11]
1
Diffuse functions from reference [77] (e.g., def2-TZVPD) have been ww 8
1 w0 =wa
tested, but often lead to serious SCF convergence problems due to near-
linear dependencies in the basis and are generally not recommended
This interpolates between the harmonic vibrational entropy for w @ w0
(and not necessary for simulations of large systems in solution).
and a pure (and always finite) rotational entropy for small w. We chose
The largest SCF calculation performed here consists of 8250 contracted a value of w0 = 100 cm1 as a default (about 1/2kT at room temperature).
Gaussian AO basis functions, but can be computed nowadays routinely Equation (7) then effectively replaces the vibrational entropy for all
on conventional PC clusters. Note that this technical level is much more modes with frequencies < 100 cm1 by a corresponding free-rotor entro-
sophisticated than in many previous applications for systems of similar py. A similar cut-off value has been used by Truhlar and co-workers re-
size (e.g., references [16, 56, 78]) that employ only double-z basis sets. For cently.[56, 84] In the Supporting Information it is shown that the computed
a careful basis set study on 3 see reference [20]. entropies are only weakly dependent on the value of w0 if chosen within
All DFT calculations were performed with TURBOMOLE.[79] The RI reasonable limits (50150 cm1). Figure 2 shows the entropy for one
approximation[80] for the Coulomb integrals is always applied, which mode as a function of w at T = 298 K and for a typical value of Bav =
speeds up the computations (in particular for TPSS) significantly without 1044 kg m2. As can be seen the average entropy given by Equation (7) re-
any significant loss of accuracy. Matching default auxiliary basis sets are mains reasonable even for very low frequencies of < 23 cm1.
used.[81] The numerical quadrature grid in TURBOMOLE denoted as m4 This allows us to use computed entropies (free enthalpies) in a black-
(m5 for the final single-point energies) has been employed for the inte- box fashion, even in cases when considerable numerical noise in the fre-
gration of the exchange-correlation contributions. The computations of quency calculations cannot be avoided for technical reasons. As will be
the harmonic vibrational frequencies at the DFT-D3 level for comparison discussed further below, the differences between the results from Equa-
were performed seminumerically in parallel using the TURBOMOLE tions (3) and (7) for bimolecular reaction (association) entropies for mol-
script NumForce. ecules with only about 100 atoms already is often 12 kcal mol1. Test cal-
Computation of internal (gas-phase) entropies: Low-lying vibrational fre- culations on model systems with 300400 atoms have shown that the
quencies are notoriously inaccurate in the harmonic approximation and errors from Equation (3) can easily reach 34 kcal mol1, because very
numerical noise in typical quantum chemical calculations causes further small w values on the order of 510 cm1 are quite common in such sys-
errors. This will be considered in the following and a completely black- tems. Note that for mid-sized molecules without low-frequency modes
box type procedure is proposed to remedy the problem. For previous ap- (for which experimental data would be available) the difference to the

&4& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 0

These are not the final page numbers!


Supramolecular Binding Thermodynamics
FULL PAPER
frequencies for DFTB-D3H and PM3-D3H methods has been considered
but optimum values (in comparison to DFT-D3/def2-TZVP computed
free enthalpies at 298 K for a set of 14 organic molecules) are close to
unity and hence frequencies are finally taken unscaled. Change of a fre-
quency scale factor by  5 % typically changes computed Ga values by
less than  0.2 kcal mol1.

Results

Gas-phase interaction energies and RRHO free enthalpy


contributions: The rotational part of the entropy depends
on the molecular symmetry of the host, guest, and the com-
plex. For highly symmetric molecules the effect of the sym-
metry number s is significant (12 kcal mol1) and requires
Figure 2. Computed entropies for one vibrational mode as a function of some consideration. For a thorough discussion of this issue
w for T = 298.15 K, Bav = 1044 kg m2, and w0 = 100 cm1. see reference [86]. We mention only cases in which symme-
try plays a role. The cucurbit[6]uril host has D6d symmetry,
standard treatment is negligible so that direct tests of the quality of the
while both hostguest complexes have only CS symmetry.
approximation cannot be made. However, as shown below, the better The cucurbit[7]uril is only close to D7d or D7h symmetry and
agreement of computed and experimental DGa values for some of the hence s = 1 is taken (the effect on DGRRHO is 1.3 kcal mol1).
here investigated systems strongly supports the reliability of the above The pincer host and glycineanhydride have C2 symmetry.
approach. The tweezer host is C2v-symmetric and benzoquinone be-
Computation of harmonic vibrational frequencies: The zero-point vibra- longs to the D2h group. The fullerenes C60 and C70 are treat-
tional energy, translational and rotational contributions to enthalpy and
free enthalpy were calculated in the usual rigid-rotor-harmonic-oscillator
ed in Ih and D5h groups, respectively, and the C60 complex
approximation (RRHO) using the equilibrium structure and seminumeri- 4 a is C2v-symmetric. The benzene dimer is close to C2h sym-
cally computed vibrational frequencies. Their computation with DFT-D3 metry, but the true minimum is found, in agreement with
is without any theoretical problems possible, but requires a huge amount previous studies, to have no symmetry.[45] Because the corre-
of computational resources for systems with several hundreds of atoms. sponding hypersurfaces are extremely flat the complex can
In fact the DFT-D3/large basis set frequency calculations would repre-
sent by far the computational bottleneck of the entire procedure. We
be considered as having effectively symmetry. Nevertheless,
hence propose to compute those by lower-level SCF methods that are, we consistently use s = 1, but expect a relatively large inher-
however, dispersion-corrected and hence can treat weakly bound and ent error of the DGRRHO term in this case.
supramolecular systems at very modest computational cost. In a few The data for DFT-D3 interaction energies, two- and
cases we compare the results from these more approximate methods to three-body dispersion contributions and RRHO free enthal-
full DFT-D3 frequency calculations. In all cases the rotational constants
computed with DFT-D3 structures have been employed for the rotational
py contributions are given in Table 1. We compare DGRRHO
part of the partition functions. values from PM6-D3H and DFTB-D3H methods in selected
We have tested D3-dispersion-correct-
ed semi-empirical methods PM6[54, 55]
and SCC-DFTB,[52] which are known Table 1. Computed gas-phase interaction energies DE (i.e., not including three-body terms), two-body DE2 disp
to be reasonably accurate for common (TPSS short-range damping parameters) and three-body DE3 disp dispersion contributions to the interaction
noncovalent interactions.[85] However, energy, and gas-phase RRHO free enthalpy contributions. For PM6 the values without the low-frequency-free-
these methods do not perform well for rotor approximation are given in parenthesis. All methods are D3-dispersion-corrected, but the D3-label has
hydrogen-bonded systems, which re- been omitted for clarity. All values in kcal mol1.
quire further, special corrections [see
Complex DE2
disp DE3
disp DE DGRRHO
the Supporting Information for a de-
TPSS/TZ TPSS/QZ PW6B95/QZ PM6 DFTB
tailed description of the corrections
that are used here and that are denot- 1 benzene dimer 5.1 0.1 3.0 2.9 3.0 7.4 (3.1) 8.8
ed D3H (zero-damping D3-disper- 2a TCNA@tweezer 38.2 1.8 33.5 33.0 32.1 16.7 (15.5) 16.5
sion correction plus hydrogen-bond 2b DCB@tweezer 27.7 1.2 22.2 21.8 21.7 14.7 (12.8) 15.0
correction)]. DFTB has sometimes 3a 3c@pincer 31.3 1.8 26.8 25.8 26.2 17.1 (16.6) 17.6[a]
problems with very polar structures 3b 3d@pincer 16.8 0.7 20.6 19.9 21.1 16.4 (15.9) [b]
(e.g. 6 or 7 b) and PM6 is known to 4a C60@catcher 55.1 3.2 36.9 36.5[c] 35.0 16.1 (13.7) 16.5[d]
yield too short alkylalkyl contacts.[85] 4b C70@catcher 58.6 3.5 38.6 38.1 36.1 16.9 (14.4) 17.7
The ferrocene part in 7 a is somewhat 5a GLH@mcyle 23.7 1.0 34.2 32.9 32.9 17.3 (17.7) 17.7[e]
distorted by PM6 and represents a bor- 5b BQ@mcyle 22.1 1.0 23.1 22.1 21.4 16.0 (16.3) 17.2
derline case for its applicability. After 6a BuNH4@CB6 25.1 2.2 84.2 83.2 84.4 16.0 (16.9) 17.6
some testing it was concluded that 6b PrNH4@CB6 21.0 1.8 80.7 79.8 80.7 15.1 (15.6) 18.1
semi-empirically computed frequencies 7a FECP@CB7 52.9 4.6 132.8 130.3 134.2 23.4 (26.5) [b]
could be used but other options (e.g., 7b ADOH@CB7 35.5 3.3 28.5 27.3 29.0 17.4 (16.5) 18.6
DFT-D3/very small basis set) should [a] The TPSS-D3 value is 16.8 kcal mol1. [b] Not computed due to missing SCC-DFTB parametrization for
be tested in the future. Applying scale some atoms. [c] 36.7 kcal mol1 using the cc-pVQZ basis. [d] The TPSS-D3 value is 16.8 kcal mol1. [e] The
factors for the harmonic vibrational TPSS-D3 value is 17.4 kcal mol1.

Chem. Eur. J. 2012, 00, 0 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &5&
These are not the final page numbers!
S. Grimme

cases with those from a full TPSS-D3 frequency calculation. (1). In all these cases substantially better agreement be-
Table 1 also contains data in the standard formalism tween the theoretical and experimental DGa values is ob-
[Eq. (3)]. The DE values are given for the semilocal TPSS tained when the low-frequency vibrations are treated in the
and PW6B95 hybrid meta-GGA functionals and for TPSS free-rotor approximation as proposed above.
with def2-TZVP (TZ) and def2-QZVPACHTUNGRE(-g,-f) (QZ) AO basis
sets. Total association free enthalpies and comparison with exper-
Interaction energies at the EDFT + E2 disp level range from imental data: In Table 2 the total gas-phase interaction ener-
3 kcal mol1 for the benzene dimer to large values of 20 gies including three-body terms with the best functional
to 30 kcal mol1 for the other neutral complexes.
The largest values of about 80 and 130 kcal
mol1 are computed for the cationic systems 6 and
Table 2. Best estimate total gas-phase interaction energies (PW6B95-D3/QZ//TPSS-
7 a, respectively. The two-body dispersion correc- D3/TZ including three-body dispersion), solvent free enthalpy contributions
tions DE2
disp to the raw DFT interaction energy are (COSMO-RS), total enthalpies, entropies and free enthalpies of association, and cor-
huge, often on the order of about 100 % of DE and responding experimental values. If not stated otherwise the data refer to standard
for the catcher complexes even almost 200 %. Al- state conditions in the solvents mentioned in the text.
though DE2 disp is a model (density functional)-depen-
Complex DE DdGsolv DHa,calcd TDSa,calcd DGa,calcd DGa,exptl
[a]
dent quantity and only the total DE represents the 1 benzene dimer 2.8 5.8 0.7 0.6 1.3 0.1
basic term, this very clearly shows that an accurate 2 a TCNA@tweezer 30.3 9.0 8.7 4.1 4.6 4.2
2 b DCB@tweezer 20.5 4.4 6.4 4.9 1.5 1.4
account of the dispersion energy is mandatory for
3 a 3c@pincer 24.4 5.4 7.8 6.0 1.8 2.3
obtaining overall reliable supramolecular interac- 3 b 3d@pincer 20.4 3.0 8.3 7.2 1.0 1.3
tions. 4 a C60@catcher[b] 31.8 6.5 14.1 4.9 9.2 5.3
The three-body dispersion term DE3 is always 4 b C70@catcher[b] 32.6 7.3 13.7 5.2 8.4 5.1
disp
31.9 5.4 5.5 8.3
repulsive and roughly an order of magnitude small- 5 a GLH@mcyle 9.2 0.1
5 b BQ@mcyle 20.4 3.9 5.4 4.8 0.6 3.3
er. However, the contributions are mostly 13 kcal 6 a BuNH @CB6 82.2 54.5 6.8 4.9 11.7 6.9
4
1
mol , which is very significant in relation to the 6 b PrNH4@CB6 78.9 56.2 0.2 7.3 7.5 5.7
magnitude of the DGa target quantity which is typi- 7 a FECP@CB7 129.6 86.9 8.5 10.7 19.2 21.1
cally in the range 0 to 8 kcal mol1. In the case of 7 b ADOH@CB7 25.7 8.9 21.3 4.1 17.2 14.1
3 a and 3 b even the affinity difference for the same [a] At 303 K. [b] At 293 K.
host is influenced by 1.1 kcal mol1 due to three-
body effects. The complexes with the biggest DE3 disp
terms are the large 4 and 7 and with an exceptionally high (PW6B95) and the largest basis set are given together with
contribution of 4.6 kcal mol1 for the iron-containing 7 a. solvation enthalpies, free enthalpies, and experimental data.
Basis set effects (i.e., difference between TZ and QZ A comparison of computed and experimental DGa values is
values) as estimated at the TPSS level are weak in most also displayed graphically in Figure 3. In this plot we include
cases (< 1 kcal mol1), which gives confidence that the inter- estimated theoretical error bars, which have been obtained
action energies with the largest QZ basis are sufficiently by adding the following three main estimates (absolute
converged for the present purpose. In addition, the differen- values): 1) the difference between PM6 and DFTB comput-
ces between the two density functionals (using the same ge- ed RRHO corrections (if available), 2) the difference be-
ometries) are small, mostly less than 1 kcal mol1 (at most
3.8 kcal mol1 corresponding to 3 % for the most strongly
bound 7 a), which is very encouraging. Calculations using
lower-level optimized structures reveal insignificant error
due to the single-point approximation (i.e., that TPSS-D3/
TZ structures are used for PW6B95/QZ).
Finally, we shall discuss methodological issues regarding
the calculation of the DGRRHO values. These contributions
are always positive (complex destabilizing) and show moder-
ate variation (with the exception of 1 due to its small size
and of 7 a) for the different complexes. In the three cases
for which we also computed TPSS-D3/TZ frequencies (un-
scaled), good mutual agreement between the semi-empirical
results within 0.5 kcal mol1 is observed. However, a notable
impact of the proposed free-rotor approximation for the
low-frequency vibrations is found. The difference to the
standard treatment (at the PM6-D3H level) can reach size-
able values of 22.5 kcal mol1 (for 2 a and 4 b), 3.1 kcal Figure 3. Comparison of experimental and computed total free associa-
mol1 for 7 a, and even 4.3 kcal mol1 for the benzene dimer tion enthalpies.

&6& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 0

These are not the final page numbers!


Supramolecular Binding Thermodynamics
FULL PAPER
tween TPSS/QZ and PW6B95/QZ results, and 3) 3 % of the between the gas-phase interaction energy and DdGsolv, but
DdGsolv term. also contains an electronically difficult transition-metal com-
Before discussing the values in some detail it should be plex.
noted that the overall agreement between theory and exper- The other problematic system seems to be the catcher
imental data can be considered as exceptionally good. For 2 complexes 4, which are overbound by 34 kcal mol1. Care-
and 3 the theoretical and experimental data almost exactly ful searches for conformers with different orientations of
match and for 5 b and 6 b the deviations are only about the guest than those reported in reference [21] were success-
1 kcal mol1. Neglecting the three-body-dispersion energy ful, but revealed only minor energy changes (about 0.5 kcal
would cause systematic overbinding by about 2 kcal mol1. mol1 for DE compared to the minimum structure reported
Overall we obtain a respectable mean absolute deviation previously). The computed gas-phase interaction energies of
(MAD) and mean (signed) deviation (MD) of 2.1 and about 32 kcal mol1 for 4 (including three-body dispersion)
0.8 kcal mol1, respectively (3.4 and 3.1 kcal mol1 with- are lower than the plain M06-2X value from reference [56]
out DE3
disp). However, these statistical data should be taken (26.4 kcal mol1 for 4 a). This result can be expected be-
with some care because of the small number of comparisons cause attractive long-range dispersion is not included in the
(13) and a few outliers that strongly influence these statis- M06 functionals. Accounting for these by using D3 (zero-
tics. damping parameters for M06-2X taken from reference [14]),
Beside this very good agreement of the absolute values a value of 35.2 kcal mol1 is obtained, which is in excellent
for complexes of very different structure we furthermore agreement with our DE(PW6B95-D3) result of 36 kcal
note a good relative accuracy. The DDGa values for the two mol1. Hence, because the gas-phase interaction seems to be
guests in the same host are always predicted qualitatively rather certain we can only speculate that the treatment of
right and often also quantitative agreement with the experi- a nonpolar solute in a nonpolar solvent (toluene) in
mental difference is obtained. As already mentioned, the COSMO-RS is responsible for the problem. Again as noted
benzene dimer has a very flat potential energy surface for above, the similar affinity of the host for C60 and C70 (with
which the present approach is limited. Nevertheless by using a slight preference for the former) is nicely reproduced by
the free-rotor approximation for the low-frequency modes, the calculations (computed DDGa = 0.8 vs. exptl. 0.2 kcal
the error of about 1.4 kcal mol1 for DGa in water (0.5 kcal mol1).
mol1 using DFTB-D3H frequencies) seems acceptable. The
here proposed approach also yields much better affinities Thermodynamic analysis and discussion of the binding
for 3 (1.8 and 1.0 kcal mol1 compared to 2.3 and mechanisms: Beside the free enthalpy also DHa and TDSa
1.3 kcal mol1 experimentally) than reported in refer- are observable and have been measured for a variety of
ence [20]. Somewhat surprisingly very reasonable results are supramolecular complexes.[2, 7] Because the experimental
also obtained for the cationic cucurbituril complexes for values sometimes vary considerably also for structurally sim-
which the cancellation between large gas-phase DE values ilar complexes and are not available for most of the here
and the solvation contribution is very pronounced. This in considered compounds, a detailed comparison of these ex-
particular holds for the dicationic complex 7 a, for which perimental and theoretical data is beyond the scope of this
DdGsolv = 87 kcal mol1, while DE = 130 kcal mol1. These work. Nevertheless it should be mentioned that the DHa and
results indicate the high accuracy and reliability of the TDSa values given in Table 2 tend to be individually larger
COSMO-RS treatment although this should be verified for than observed in most experiments meaning that enthalpy
other charged complexes, in particular if multiply charged. entropy compensation effects appear stronger theoretically.
Inspection of Figure 3 shows that there are basically only Similar observations have been made in a force-field study
three supramolecular systems which might be considered as on cyclodextrin complexes of small molecules[63] (for
outliers. The hydrogen-bonded complex 5 a with a mea- a recent discussion see also reference [87]). However,
sured Ka = 106 m1 is computed to be too weakly bound a basic discussion of this point requires further studies on
(about 2.8 kcal mol1 error). The deviation for the similar specifically selected complexes for which very solid experi-
complex 5 b is of about the same size, so that some systemat- mental data are available and therefore we will discuss here
ic problems for this system (maybe also in the experiment) only the implications from the present theoretical results.
seem to be present. For the overbinding of the complex 6 a Figure 4a shows a comparison of computed gas-phase in-
(by 4.8 kcal mol1) there is also no evident reason, as 6 b is teraction energies and solution free association enthalpies.
described well and the structural difference between the two This may shed light on the important question of how much
guests (propyl vs. butyl moiety) is small. Qualitatively the supramolecular interactions are influenced by secondary
computed difference between 6 a and 6 b also seems reason- effects like solvation and change of entropy.
able (better binding of the larger guest in 6 a and more re- We notice some correlation between DE and DGa values.
pulsive solvent contribution of the smaller cation guest in Of course it cannot be used to quantitatively estimate en-
6 b). The error for 6 a is also difficult to understand in light thalpies from energies because the structural dependency of
of the excellent results for the similar cucurbit[7]uril com- the other contributions is substantial. Nevertheless, the in-
plexes 7. This in particular holds for 7 a, which not only is teraction energy can be considered as the basic driving force
a dication with the already mentioned huge compensation for binding in most cases and should be employed as guiding

Chem. Eur. J. 2012, 00, 0 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &7&
These are not the final page numbers!
S. Grimme

ternal and solvation) factors. In contrast to most other sys-


tems, the cationic complexes have strong stabilizing entropic
contributions. This can be explained by the fact that the
host does not need to be ordered in the complexation (in
contrast to the solvent molecules) and the corresponding
solvent entropy is released. This is in qualitative agreement
with experimental observations for many cationic supra-
molecular complexes.[2, 7] The only other supramolecular
complex with a positive TDSa value is the strongly hydro-
gen-bonded 4 a.
The most striking result of the analysis of the thermody-
namic data is the neutral cucurbit[7]uril complex with 1-hy-
droxyadamantane 7 b that shows a very large affinity of
DGa = 14.1 kcal mol1 (calcd 17.2 kcal mol1), but only
a moderate DE value of 26 kcal mol1. It is clearly recog-
nized in Figure 4a as an outlier. According to the data in
Table 2 and in comparison to, for example, the energetically
similarly bound complex 5 a, favorable solvation contribu-
tions (enthalpic and entropic) seem to be the descisive
factor. Tentatively it can be assigned to the semipolar char-
acter (relatively large hydrocarbon with a small polar OH
side group) of the guest (similar to the case of the benzene
dimer in water). Upon complexation this releases the sol-
ventsolvent interactions in water, but nevertheless, the
guest matches perfectly with the host regarding pure ener-
getics. It can be forseen that valuable insight of this kind
will lead in the future to specifically tailored hostguest sys-
tems.

Figure 4. Comparison of computed a) gas-phase interaction energies with


free association enthalpies in solution (the solid line shows the result of
a linear regression with a correlation coefficient of 0.79) and b) associa- Conclusion
tion solution enthalpies with corresponding entropies (multiplied by the
measurement temperature, the regression correlation coefficient is 0.45). A nonempirical quantum chemical procedure to compute
free enthalpies of association for supramolecular complexes
in solution (common organic solvents or water at ambient
principle when new supramolecular systems are developed. temperature) is proposed. The approach can be applied in
Note that the DE discussed here does not represent some the present form routinely to complexes of about 300400
model-dependent quantity as in empirical methods, but rep- atoms and the only prerequisite is a certain conformational
resents (within its error limit) a physically well-defined and rigidity. In contrast to classical force-field based approaches,
fundamental quantity. This is an important advantage of the the method can handle almost any type of electronic struc-
present approach. Although strong energy binders can ture including charged systems without special adjustments.
suffer from large entropy penalty (see below), it is unlikely The unprecedented accuracy obtained here with errors for
that high binding affinities can be obtained without energet- absolute DGa values of typically only about 12 kcal mol1
ically strong interactions. Although this seems to be a some- (average deviation of 2 kcal mol1 with a maximum of 4 kcal
what trivial conclusion, it is based here for the first time on mol1) is rooted in the high accuracy of the DFT-D3 method
a set of fundamentally correct ab initio data for large sys- to compute gas-phase interaction energies DE. It has been
tems. previously estimated to provide errors for DE of about 5 %
Figure 4b shows a comparison of computed association or less, which is consistent with the findings reported here.
enthalpies and the corresponding entropy penalties TDSa. In passing it is noted that the accuracy of DFT-D3 (and this
Although the plot shows a lot of scatter in the data, a rough holds for the many other published dispersion corrections)
enthalpyentropy compensation can be deduced; that is, has never been tested for large molecular systems in com-
strong binding causes freezing of internal and solvent de- parison to experimental data. The very good results ob-
grees of freedom. Interestingly, for most neutral complexes tained here for the first time in such a real-life situation
(except the very nonpolar systems) the association enthal- are encouraging and support the basic ideas in the D3 devel-
pies DHa fall in a rather narrow range of 6 to 10 kcal opment. However, a new and important outcome of the
mol1. Hence, the substantial variations in DGa for the neu- present work is that three-body dispersion effects (which are
tral complexes are also strongly influenced by entropic (in- unimportant in small benchmark systems and that have not

&8& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 0

These are not the final page numbers!


Supramolecular Binding Thermodynamics
FULL PAPER
been considered in great detail so far) must be considered [6] E. A. Meyer, R. K. Castellano, F. Diederich, Angew. Chem. 2003,
for accurate results. Neglecting those would lead to over- 115, 1244; Angew. Chem. Int. Ed. 2003, 42, 1210.
[7] H.-J. Schneider, Angew. Chem. 2009, 121, 3982; Angew. Chem. Int.
binding for the majority of the investigated complexes by Ed. 2009, 48, 3924.
typically 13 kcal mol1 (with a maximum value of 4.6 kcal [8] J. Antony, S. Grimme, D. G. Liakos, F. Neese, J. Phys. Chem. A
mol1). The other hitherto overlooked and important effect 2011, 115, 11210.
in the calculations is the entropy due to the ubiquitous low- [9] S. C. Wang, Phys. Z. 1927, 28, 663.
frequency vibrational modes in large molecules. If treated in [10] F. London, Z. Phys. 1930, 63, 245.
[11] S. Grimme, J. Antony, S. Ehrlich, H. Krieg, J. Chem. Phys. 2010,
the standard harmonic approximation, they can cause unpre- 132, 154104.
dictable errors for entropy differences and often introduce [12] L. A. Burns, A. Vazquez-Mayagoitia, B. G. Sumpter, C. D. Sherrill,
significant noise into the total free enthalpy results. This J. Chem. Phys. 2011, 134, 084107.
problem has been practically solved by applying a simple [13] L. Goerigk, H. Kruse, S. Grimme, ChemPhysChem 2011, 12, 3421.
[14] L. Goerigk, S. Grimme, Phys. Chem. Chem. Phys. 2011, 13, 6670.
free-rotor approximation. The third main conclusion con-
[15] B. Tidor, M. Karplus, J. Mol. Biol. 1994, 238, 405.
cerns the COSMO-RS solvation model that provides excep- [16] K. N. Houk, S. Menzer, S. P. Newton, F. M. Raymo, J. F. Stoddart,
tionally accurate solvation free energies. From the present D. J. Williams, J. Am. Chem. Soc. 1999, 121, 1479.
results and analysis it seems as if all three main factors [17] S. Moghaddam, C. Yang, M. Rekharsky, Y. H. Ko, K. Kim, Y. Inoue,
(DFT-D3, vibrational entropies, solvation) contribute rough- M. K. Gilson, J. Am. Chem. Soc. 2011, 133, 3570.
[18] C. D. Sherrill, B. G. Sumpter, M. O. Sinnokrot, M. S. Marshall, E. G.
ly with the same uncertainty to the theoretical results, at
Hohenstein, R. C. Walker, I. R. Gould, J. Comput. Chem. 2009, 30,
least for neutral complexes. For charged systems the solva- 2187.
tion model likely becomes the accuracy determining factor. [19] M. Kamieth, U. Burkert, P. S. Corbin, S. J. Dell, S. C. Zimmerman,
Beside the astonishingly good agreement between theory F.-G. Klrner, Eur. J. Org. Chem. 1999, 2741.
and experiment for absolute values it should be mentioned [20] J. Graton, J.-Y. Le Questel, B. Legouin, P. Uriac, P. van de Weghe,
D. Jacquemin, Chem. Phys. Lett. 2012, 522, 11.
that the relative affinities for different guests are obtained [21] C. Mck-Lichtenfeld, S. Grimme, L. Kobryn, A. Sygula, Phys.
accurately as well. In this context we want to highlight here Chem. Chem. Phys. 2010, 12, 7091.
only the compensation effects in the example of cucur- [22] C. Allott, H. Adams, P. L. Bernad Jr., C. A. Hunter, C. Rotger, J. A.
bit[7]uril, for which the two complexes have similarly huge Thomas, Chem. Commun. 1998, 2449.
[23] W. L. Mock, N.-Y. Shih, J. Am. Chem. Soc. 1989, 111, 2697.
experimental affinities of 21 and 14 kcal mol1, respec-
[24] E. E. Tucker, S. D. Christian, J. Phys. Chem. 1979, 83, 426.
tively (computed to be 19.2 and 17.2 kcal mol1) that [25] C. D. Sherrill, Rev. Comput. Chem. 2009, 26, 1.
result from totally different contributions, for example, the [26] M. O. Sinnokrot, E. F. Valeev, C. D. Sherrill, J. Am. Chem. Soc.
DDGsolv terms of 87 and 9 kcal mol1 and the gas-phase in- 2002, 124, 10887.
teraction energies of 130 and 26 kcal mol1, respectively. [27] P. Jurecka, J. Sponer, J. Cerny, P. Hobza, Phys. Chem. Chem. Phys.
2006, 8, 1985.
Similar analysis of the different contributions to relative
[28] T. Takatani, E. G. Hohenstein, M. Malagoli, M. S. Marshall, C. D.
DGa values in the other cases shows, that DE is the decisive Sherrill, J. Chem. Phys. 2010, 132, 144104.
factor for favorable association although to a varying [29] D. Benitez, E. Tkatchouk, I. Yoon, J. F. Stoddart, W. A. God-
degree. In all of the investigated pairs of hostguest com- dard III, J. Am. Chem. Soc. 2008, 130, 14928.
plexes a larger (absolute) DE value also leads to a better af- [30] A. Klamt, J. Phys. Chem. 1995, 99, 2224.
[31] F. Eckert, A. Klamt, AIChE J. 2002, 48, 369.
finity. Enthalpyentropy compensation plays a role, but its [32] F. Eckert, A. Klamt, COSMOtherm, Version C2.1, Release 01.11;
structural and solvent dependence is not fully clear and COSMOlogic GmbH & Co. KG, Leverkusen (Germany), 2010.
should be investigated in more detail further. In any case, [33] J. P. Perdew, Phys. Rev. B 1986, 33, 8822.
the future for an ab initio based understanding of noncova- [34] A. Schfer, C. Huber, R. Ahlrichs, J. Chem. Phys. 1994, 100, 5829.
[35] J. Tomasi, B. Mennucci, R. Cammi, Chem. Rev. 2005, 105, 2999.
lent interactions in large supra- and biomolecular systems is
[36] A. Klamt, B. Mennucci, J. Tomasi, V. Barone, C. Curutchet, M.
bright. Orozco, F. Luque, Acc. Chem. Res. 2009, 42, 489.
[37] C. J. Cramer, Essentials of Compuational Chemistry, Wiley, New
York, 2002.
[38] C. J. Cramer, D. G. Truhlar, Acc. Chem. Res. 2008, 41, 760.
[39] C. J. Cramer, D. G. Truhlar, Acc. Chem. Res. 2009, 42, 493.
Acknowledgements [40] R. Franke, B. Hannebauer, Phys. Chem. Chem. Phys. 2011, 13,
21344.
The author thanks the team of principal investigators of the SUPRA- [41] S. Grimme, S. Ehrlich, L. Goerigk, J. Comput. Chem. 2011, 32, 1456.
TEC DFG excellence initiative for stimulating discussions that have [42] O. A. Vydrov, T. Van Voorhis, J. Chem. Phys. 2010, 133, 244103.
triggered the present investigations and D. Jacquemin for providing coor- [43] W. Hujo, S. Grimme, J. Chem. Theory Comput. 2011, 7, 3866.
dinates of structures 3. [44] M. Dion, H. Rydberg, E. Schrder, D. C. Langreth, B. I. Lundqvist,
Phys. Rev. Lett. 2004, 92, 246401.
[45] J. Grfenstein, D. Cremer, J. Chem. Phys. 2009, 130, 124105.
[1] J. M. Lehn, Supramolecular Chemistry, Concepts and Perspectives, [46] A. Krishtal, D. Geldorf, K. Vanommeslaeghe, C. VanAlsenoy, P.
VCH, Weinheim, 1995. Geerlings, J. Chem. Theory Comput. 2012, 8, 125.
[2] J. L. Atwood, J. W. Steed, Supramolecular Chemistry. Wiley, 2009. [47] A. D. Becke, E. R. Johnson, J. Chem. Phys. 2005, 122, 154104.
[3] K. S. Kim, P. Tarakeshwar, J. Y. Lee, Chem. Rev. 2000, 100, 4145. [48] E. R. Johnson, A. D. Becke, J. Chem. Phys. 2005, 123, 024101.
[4] P. Hobza, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 3. [49] S. Grimme, Wiley Interdiscip. Rev. Comput. Mol. Sci. 2011, 1, 211.
[5] E. G. Hohenstein, C. D. Sherrill, Wiley Interdiscip. Rev. Comput. [50] J. Tao, J. P. Perdew, V. N. Staroverov, G. E. Scuseria, Phys. Rev. Lett.
Mol. Sci. 2012, 2, 304. 2003, 91, 146401.

Chem. Eur. J. 2012, 00, 0 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &9&
These are not the final page numbers!
S. Grimme

[51] Y. Zhao, D. G. Truhlar, J. Phys. Chem. A 2005, 109, 5656. [70] A. Tkatchenko, O. A. von Lilienfeld, Phys. Rev. B 2008, 78, 045116.
[52] M. Elstner, D. Porezag, G. Jungnickel, J. Elsner, M. Haugk, T. [71] O. A. von Lilienfeld, A. Tkatchenko, J. Chem. Phys. 2010, 132,
Frauenheim, S. Suhai, G. Seifert, Phys. Rev. B 1998, 58, 7260 7268. 234109.
[53] T. Frauenheim, DFTB + (Density Functional based Tight Binding) [72] Y. Zhao, D. G. Truhlar, Theor. Chem. Acc. 2008, 120, 215.
2008, see http://www.dftb.org/. [73] L. Goerigk, S. Grimme, J. Chem. Theory Comput. 2011, 7, 291.
[54] J. J. P. Stewart, J. Mol. Model. 2007, 13, 1173. [74] F. Weigend, R. Ahlrichs, Phys. Chem. Chem. Phys. 2005, 7, 3297.
[55] J. J. P. Stewart, Stewart Computational Chemistry, Version 11.038 L, [75] F. Weigend, F. Furche, R. Ahlrichs, J. Chem. Phys. 2003, 119, 12753.
see http://OpenMOPAC.net. [76] T. H. Dunning Jr., J. Chem. Phys. 1989, 90, 1007.
[56] Y. Zhao, D. G. Truhlar, Phys. Chem. Chem. Phys. 2008, 10, 2813. [77] D. Rappoport, F. Furche, J. Chem. Phys. 2010, 133, 134105.
[57] F. G. Klrner, J. Panitzky, D. Preda, L. T. Scott, J. Mol. Model. 2000, [78] M. Murata, S. Maeda, Y. Morinaka, Y. Murata, K. Komatsu, J. Am.
6, 318. Chem. Soc. 2008, 130, 15800.
[58] F. G. Klrner, B. Kahlert, Acc. Chem. Res. 2003, 36, 919. [79] R. Ahlrichs, M. Br, M. Hser, H. Horn, C. Klmel, Chem. Phys.
[59] C. E. Chang, M. K. Gilson, J. Am. Chem. Soc. 2004, 126, 13156. Lett. 1989, 162, 165.
[60] S. P. Brown, T. Schaller, U. P. Seelbach, F. Koziol, C. Ochsenfeld, [80] K. Eichkorn, O. Treutler, H. hm, M. Hser, R. Ahlrichs, Chem.
F. G. Klrner, H. W. Spiess, Angew. Chem. 2001, 113, 740; Angew. Phys. Lett. 1995, 240, 283.
Chem. Int. Ed. 2001, 40, 717. [81] K. Eichkorn, F. Weigend, O. Treutler, R. Ahlrichs, Theor. Chem.
[61] M. Parac, M. Etinski, S. Grimme, J. Chem. Theory Comput. 2005, 1, Acc. 1997, 97, 119.
1110. [82] P. Y. Ayala, H. B. Schlegel, J. Chem. Phys. 1998, 108, 2314.
[62] T. Yu, J. Zheng, D. G. Truhlar, Phys. Chem. Chem. Phys. 2012, 14, [83] J.-D. Chai, M. Head-Gordon, Phys. Chem. Chem. Phys. 2008, 10,
482. 6615.
[63] W. Chen, C.-E. Chang, M. K. Gilson, Biophys. J. 2004, 87, 3035. [84] R. F. Ribeiro, A. V. Marenich, C. J. Cramer, D. G. Truhlar, J. Phys.
[64] C.-E. Chang, M. J. Potter, K. Gilson, J. Phys. Chem. B 2003, 107, Chem. B 2011, 115, 14556.
1048. [85] J. Rezc, P. Hobza, J. Chem. Theory Comput. 2012, 8, 141.
[65] T. Wiegand, H. Eckert, O. Ekkert, R. Frhlich, G. Kehr, G. Erker, S. [86] M. K. Gilson, K. K. Irikura, J. Phys. Chem. B 2010, 114, 16304.
Grimme, J. Am. Chem. Soc. 2012, 134, 4236. [87] A. Choutko, W. F. van Gunsteren, P. H. Hnenberger, ChemPhys-
[66] B. M. Axilrod, E. Teller, J. Chem. Phys. 1943, 11, 299. Chem 2011, 12, 3214.
[67] J. Mutto, Proc. Phys. Math. Soc. Japan 1943, 17, 629.
[68] See http://www.thch.uni-bonn.de/.
[69] J. Rezc, K. E. Riley, P. Hobza, J. Chem. Theory Comput. 2011, 7, Received: February 15, 2012
2427. Published online: && &&, 0000

&10& www.chemeurj.org  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim Chem. Eur. J. 0000, 00, 0 0

These are not the final page numbers!


Supramolecular Binding Thermodynamics
FULL PAPER
Theoretical Chemistry

S. Grimme* . . . . . . . . . . . . . . . . . . . &&&&&&&&
Supramolecular Binding Thermody-
namics by Dispersion-Corrected
Density Functional Theory
Unprecedented accuracy for computed high-level quantum chemical proce-
association free enthalpies of supra- dures. The approach is quite general,
molecular hostguest complexes (some includes all basic physical effects quan-
examples are shown here) in solution titatively and requires no special
has been achieved by a combination of empirical adjustments.

Chem. Eur. J. 2012, 00, 0 0  2012 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim www.chemeurj.org &11&
These are not the final page numbers!

Vous aimerez peut-être aussi