predicting the tautomeric equilibrium of acetylacetone in solution. i. the right answer for the...

6
Predicting the Tautomeric Equilibrium of Acetylacetone in Solution. I. The Right Answer for the Wrong Reason? SEBASTIAN SCHLUND, 1 ELINE M. BASI ´ LIO JANKE, 2 KLAUS WEISZ, 3 BERND ENGELS 1 1 Institute for Organic Chemistry, University of Wu¨rzburg, Am Hubland, 97074 Wu¨rzburg, Germany 2 Institute for Chemistry, Free University of Berlin, Takustr. 3, 14195 Berlin, Germany 3 Institute for Biochemistry, Ernst-Moritz-Arndt University of Greifswald, Felix-Hausdorff-Str. 4, 17489 Greifwald, Germany Received 14 February 2009; Revised 14 May 2009; Accepted 19 May 2009 DOI 10.1002/jcc.21354 Published online 25 June 2009 in Wiley InterScience (www.interscience.wiley.com). Abstract: This study investigates how the various components (method, basis set, and treatment of solvent effects) of a theoretical approach influence the relative energies between keto and enol forms of acetylacetone, which is an important model system to study the solvent effects on chemical equilibria from experiment and theory. The compu- tations show that the most popular density functional theory (DFT) approaches, such as B3LYP overestimate the sta- bility of the enol form with respect to the keto form by 10 kJ mol 21 , whereas the very promising SCS-MP2 approach is underestimating it. MP2 calculations indicate that in particular the basis set size is crucial. The Dunning Huzinaga double f basis (D95z(d,p)) used in previous studies overestimates the stability of the keto form consider- ably as does the popular split-valence plus polarization (SVP) basis. Bulk properties of the solvent included by con- tinuum approaches strongly stabilize the keto form, but they are not sufficient to reproduce the reversal in stabilities measured by low-temperature nuclear magnetic resonance experiments in freonic solvents. Enthalpic and entropic effects further stabilize the keto form, however, the reversal is only obtained if also molecular effects are taken into account. Such molecular effects seem to influence only the energy difference between the keto and the enol forms. Trends arising due to variation in the dielectric constant of the solvent result from bulk properties of the solvent, i.e., are already nicely described by continuum approaches. As such this study delivers a deep insight into the abil- ities of various approaches to describe solvent effects on chemical equilibria. q 2009 Wiley Periodicals, Inc. J Comput Chem 31: 665–670, 2010 Key words: keto-enol equilibrium; solvent effects; method dependency Introduction The influence of the surrounding medium on both the reaction pathway (thermodynamics) and reaction rate (kinetics) is one of the major challenges in chemistry as nearly all reactions take place in solvent. 1 Several studies show that not only the bulk properties but also the molecular structure of the solvent can be of great significance. 2 Hence, the development of theoretical methods which allow for solvent effects has become an impor- tant field of research. For electronic structure calculations the continuum models approximating a homogeneous bulk of solvent molecules have gained large popularity in the recent past. 3 Hereby, the polariza- tion effect of a continuum on the electronic structure of the sol- ute is taken into account. Often used solvent models are the polarizable continuum model (PCM) method developed by Tom- asi or the conductor-like screening model (COSMO) by Klamt and Schu ¨urmann. 4 The molecular nature of the solvent environ- ment can be treated explicitly by the combined quantum mechanics/molecular mechanics simulation method. 5 The solute is calculated within a QM framework, whereas the solvent mole- cules are described by a classical force field. The interaction potentials between solvent and solute are included either by an electrostatic or polarized embedding approach. 6 However, due to the large computational costs, this approach is only feasible with a semiempirical or density functional based treatment of the QM part. A well-studied example of solvent effect is the keto-enol tau- tomerism of acetylacetone: either the keto or the enol form is favored depending on the surrounding solvent. Experiments have Additional Supporting Information may be found in the online version of this article. Correspondence to: B. Engels; e-mail: [email protected] Contract/grant sponsor: Deutsche Forschungsgemeinschaft; contract/grant numbers: EN1971/10, SFB 630 q 2009 Wiley Periodicals, Inc.

Upload: sebastian-schlund

Post on 11-Jun-2016

218 views

Category:

Documents


1 download

TRANSCRIPT

Page 1: Predicting the tautomeric equilibrium of acetylacetone in solution. I. The right answer for the wrong reason?

Predicting the Tautomeric Equilibrium of Acetylacetone

in Solution. I. The Right Answer for the Wrong Reason?

SEBASTIAN SCHLUND,1 ELINE M. BASILIO JANKE,2 KLAUS WEISZ,3 BERND ENGELS1

1Institute for Organic Chemistry, University of Wurzburg, Am Hubland,97074 Wurzburg, Germany

2Institute for Chemistry, Free University of Berlin, Takustr. 3, 14195 Berlin, Germany3Institute for Biochemistry, Ernst-Moritz-Arndt University of Greifswald, Felix-Hausdorff-Str. 4,

17489 Greifwald, Germany

Received 14 February 2009; Revised 14 May 2009; Accepted 19 May 2009DOI 10.1002/jcc.21354

Published online 25 June 2009 in Wiley InterScience (www.interscience.wiley.com).

Abstract: This study investigates how the various components (method, basis set, and treatment of solvent effects)

of a theoretical approach influence the relative energies between keto and enol forms of acetylacetone, which is an

important model system to study the solvent effects on chemical equilibria from experiment and theory. The compu-

tations show that the most popular density functional theory (DFT) approaches, such as B3LYP overestimate the sta-

bility of the enol form with respect to the keto form by �10 kJ mol21, whereas the very promising SCS-MP2

approach is underestimating it. MP2 calculations indicate that in particular the basis set size is crucial. The Dunning

Huzinaga double f basis (D95z(d,p)) used in previous studies overestimates the stability of the keto form consider-

ably as does the popular split-valence plus polarization (SVP) basis. Bulk properties of the solvent included by con-

tinuum approaches strongly stabilize the keto form, but they are not sufficient to reproduce the reversal in stabilities

measured by low-temperature nuclear magnetic resonance experiments in freonic solvents. Enthalpic and entropic

effects further stabilize the keto form, however, the reversal is only obtained if also molecular effects are taken into

account. Such molecular effects seem to influence only the energy difference between the keto and the enol forms.

Trends arising due to variation in the dielectric constant of the solvent result from bulk properties of the solvent,

i.e., are already nicely described by continuum approaches. As such this study delivers a deep insight into the abil-

ities of various approaches to describe solvent effects on chemical equilibria.

q 2009 Wiley Periodicals, Inc. J Comput Chem 31: 665–670, 2010

Key words: keto-enol equilibrium; solvent effects; method dependency

Introduction

The influence of the surrounding medium on both the reaction

pathway (thermodynamics) and reaction rate (kinetics) is one of

the major challenges in chemistry as nearly all reactions take

place in solvent.1 Several studies show that not only the bulk

properties but also the molecular structure of the solvent can be

of great significance.2 Hence, the development of theoretical

methods which allow for solvent effects has become an impor-

tant field of research.

For electronic structure calculations the continuum models

approximating a homogeneous bulk of solvent molecules have

gained large popularity in the recent past.3 Hereby, the polariza-

tion effect of a continuum on the electronic structure of the sol-

ute is taken into account. Often used solvent models are the

polarizable continuum model (PCM) method developed by Tom-

asi or the conductor-like screening model (COSMO) by Klamt

and Schuurmann.4 The molecular nature of the solvent environ-

ment can be treated explicitly by the combined quantum

mechanics/molecular mechanics simulation method.5 The solute

is calculated within a QM framework, whereas the solvent mole-

cules are described by a classical force field. The interaction

potentials between solvent and solute are included either by an

electrostatic or polarized embedding approach.6 However, due to

the large computational costs, this approach is only feasible with

a semiempirical or density functional based treatment of the QM

part.

A well-studied example of solvent effect is the keto-enol tau-

tomerism of acetylacetone: either the keto or the enol form is

favored depending on the surrounding solvent. Experiments have

Additional Supporting Information may be found in the online version of

this article.

Correspondence to: B. Engels; e-mail: [email protected]

Contract/grant sponsor: Deutsche Forschungsgemeinschaft; contract/grant

numbers: EN1971/10, SFB 630

q 2009 Wiley Periodicals, Inc.

Page 2: Predicting the tautomeric equilibrium of acetylacetone in solution. I. The right answer for the wrong reason?

shown that in gas phase and media of low polarity the enol form

is more stable than the keto form, whereas the keto form domi-

nates in more polar environments.2,7 The large effects result

from the intramolecular stabilization of the enol form that is

strongly reduced by competing intermolecular interactions with

the solvent shell (Fig. 1).

With the availability of new nuclear magnetic resonance

(NMR) experiments using the liquefied freonic gases (CDF3/

CDF2Cl), we were able to investigate the keto-enol tautomeric

equilibrium of acetylacetone in solvent with different dielectric

constants. The NMR measurements were performed at tempera-

tures as low as 123 K.8 Because the dielectric constant of the

freonic mixture depends strongly on temperature, it effectively

mimics very different solvent polarities if measurements are per-

formed over a large temperature range (from e 5 14 at 190 K to

e 5 34 at 120 K).9 Thus, unusually strong temperature effects

on the acetylacetone keto-enol equilibrium were experimentally

observed on lowering the temperature (see Supp. Info.). An enol

tautomeric preference typical for the presence of a nonpolar sol-

vent is observed at higher temperatures. The corresponding sig-

nal intensities show a enol:keto molar ratio of about 4:1 at T �173 K. With decreasing temperature, however, the keto tautomer

becomes increasingly populated and at 123 K the enol:keto

molar ratio has completely reversed (1:4).

To model these new NMR experiments, we have performed

state-of-the-art electronic structure calculations taking bulk

effects of the solvent into account. However, by using a contin-

uum approach for modeling the solvent shell, we were not able

to reproduce the experimentally observed shift in the keto-enol

ratio of acetylacetone upon increasing the solvent polarity.

This was curious since previous theoretical works seemed to

describe this reversal with less sophisticated approaches. Previ-

ous computations showed10,11 that the keto tautomer undergoes

a geometric change as the solvent polarity increases making its

dipole moment larger compared to the enol form.12,13 Finally,

Ishida et al.13 correctly reproduced the experimentally observed

enol ? keto transition in polar solvents by means of electronic

structure calculations on MP2/D95z(d,p) level of theory. The

solvent effects were taken into account by the reference site

interaction model self-consistent field (RISM-SCF) method.

The RISM-SCF by Kato and coworkers14 represents an inte-

gral equation method based on statistical mechanics of molecu-

lar liquids coupled with Hartree-Fock equations. Hence, the mo-

lecular effects of the solvent molecules are only described by

potential functions and do not explicitly include hydrogen bond-

ing from the solvent molecules to the solute. Therefore, the suc-

cess of this approach suggests that the shift from the enol form

to the keto form derives primarily from the bulk properties of

the solvent, whereas the molecular nature of the solvent is of

smaller relevance (Sakaki et al.15 have recently extended the

RISM-SCF model using a more realistic Coulomb interaction

between solvent and solute by introducing a spatial electron den-

sity distribution instead of using a set of grid points).

Because our improved computations were not able to reproduce

the experimentally observed shift, we reinvestigated the problem

to reveal possible error cancellations. Our calculations clearly

show that the good agreement of former theoretical studies with

corresponding experimental results indeed result from an error

compensation due to two opposite trends. The present results also

give a deeper insight into the accuracy of various ab initio meth-

ods in the description of keto-enol equilibria and indicate that the

molecular nature of the solvent plays an important role.

Theoretical Details

As in related studies,16 we first applied DFT approaches, which

are known for their favorable cost-benefit ratio.17 However, in

some cases, reliable predictions deserve higher level

approaches,18 we also applied perturbational approaches and

coupled cluster methods. All structure optimizations have been

performed with the Turbomole program package (V5.8-V5.10)19

and were done at RI-BP86, B3LYP, BHLYP, HF, and RI-MP2

level of theory using a TZVPP (valence triple zeta plus two sets

of polarization functions) basis set. Furthermore, MP2 single-

point calculations with D95v(d,p), SV(P), SVP, and TZVPP ba-

sis sets have been performed on HF/TZVPP optimized struc-

tures. The electrostatic contributions of the solvent were esti-

mated using the COSMO approach with dielectric constants

from e 5 3 to e 5 78 to simulate solvents of varying polarity.4

The CCSD(T)/TZVPP and (SCS)-MP2/TZVPP single-point cal-

culations were performed with MOLPRO on RI-MP2/TZVPP

optimized geometries.20

The frequencies and thermodynamic corrections in gas phase

were calculated analytically on DFT (BHLYP) level with the Tur-

bomole program package. Hereby, the geometries have been reopti-

mized using a Valence triple zeta plus two sets of polarization

functions (TZVPP) basis set because the analytical gradient was

not available for g-functions. Frequencies of the continuum sol-

vated molecules have been calculated numerically on a RI-

BHLYP/TZVPP level of theory as calculations of numerical deriv-

atives with COSMO were only possible using the RI approxima-

tion. All wave numbers were scaled by a factor of 0.9 and the

enthalpic and entropic contributions were determined at 273 and

123 K. The nonelectrostatic contributions to the free energy of sol-

vation have been determined with the COSMO approach as imple-

mented in Gaussian-03 (C-PCM with standard parameters). We

performed BHLYP/TZVP single-point calculations on the BHLYP/

TZVPP optimized geometries in the respective dielectric.21

For the explicitly solvated structures of acetylacetone, force

field-based conformational searches have been performed

(MMFF94 force field, Mixed MCMM/LM search algorithm,

5000 steps) using the Macromodel V8.0 suite by Schrodinger

(Portland, OR). The generated conformers were subsequently

optimized on DFT level (RI-BP86/SVP), and the lowest con-

formers have been reoptimized on BHLYP/TZVPP level to

locate the global minima on the PES.

Figure 1. Tautomeric structures of acetylacetone.

666 Schlund et al. • Vol. 31, No. 4 • Journal of Computational Chemistry

Journal of Computational Chemistry DOI 10.1002/jcc

Page 3: Predicting the tautomeric equilibrium of acetylacetone in solution. I. The right answer for the wrong reason?

Results

The optimized structures of the keto and enol tautomers of ace-

tylacetone are depicted in Figure 2. The respective electronic

energies in gas phase and solvent for the various levels of theory

are given in Table 1. Regarding first the gas phase values one

can see a discrepancy in the relative energies between pure

GGA functionals (BP86, BLYP) and hybrid functionals

(B3LYP, BHLYP). In comparison with the most accurate

CCSD(T) result, the energy difference between the keto and

enol forms is strongly overestimated if pure GGA functionals

(BP86, BLYP) are used. If an increasing amount of exact

exchange is included, the computed energy difference

approaches the CCSD(T) value. The BHLYP value deviates by

about 4 kJ mol21, which is of similar accuracy than the RI-MP2

results. Moreover, density functionals with less exact exchange

were not able to locate the second local minimum K2 on the

energy hypersurface which becomes more stable in continuum

solvation with large dielectrics than K1.

The gas phase energies obtained by Møller-Plesset perturba-

tion theory second order calculated with a triple-f-basis set with

additional polarization functions are generally in good agreement

with coupled-cluster results. Only the spin-component scaled

MP2 method describes the relative energies of the keto forms

K1 and K2 systematically as too stable. It seems that the

decreased ‘‘triplet’’ contribution of the SCS-MP2 method is

underestimating the intramolecularly hydrogen bonded enol form

E. A much stronger underestimation of the energy difference is

obtained if a double-f-basis is employed as it was done in the

study of Dannenberg and Rios.11 The relative energy of the K1

tautomer lies only 4 kJ mol21 higher than for the enol tautomer

if the D95** basis set by Dunning and Huzinaga is used.

The influence of bulk effects of a polar solvent was estimated

on BHLYP/TZVPP level. Going from gas phase to solution, two

effects can be observed: (1) with increasing dielectric constant

of the solvent, the keto forms are stabilized with respect to the

enol form and (2) the keto form K2 becomes more stable than

K1 for e values larger than 3. This behavior is reflected in a

larger increase of the dipole moment (Table 2) of K2 indicating

a stronger stabilization of this tautomer in polar solvents. The

measured dihedral angles between both carbonyl groups in K2

are in excellent agreement with previously published studies that

predicted a geometrical change of the keto tautomer on solva-

tion.12,13 However, by using a simple continuum model the trend

reversal in the stability of keto/enol tautomers as observed

experimentally is not found. BHLYP/TZVPP predicts K2 to be

about 6 kJ mol21 less stable than E. For gas phase, BHLYP

deviates from the more accurate CCSD(T) by 4 kJ mol21.

Transferring this difference to solvent, K2 is still less stable

than E by about 2 kJ mol21, even in a continuum simulating an

aqueous surrounding (e 5 78).

To estimate the influence of enthalpic and entropic contribu-

tions on the tautomeric equilibrium of acetylacetone, the thermo-

dynamic corrections have been calculated both for gas phase

and solution on DFT level with varying dielectric constants and

temperatures (Table 3).

In all cases enthalpic (DHcorr) and entropic effects (2TDScorr)stabilize the keto form relative to the enol form. The entropic

effect results from the more disordered state of the keto form.

The enthalpic effect is mainly connected with the zero-point

Figure 2. Tautomeric structures of acetylacetone in gas phase (RI-MP2/TZVPP).

Table 1. Relative Energies (in kJ mol21) of Optimized Structures of

Acetylacetone Tautomers Calculated in Gas Phase or for Continuum

Solvation (COSMO) With a TZVPP Basis Set (Exception: MP2/D95**).

Level of theory

Enol Keto

E K1 K2

CCSD(T)a 0.0 118 122

BP86 0.0 138 2BLYP 0.0 128 2B3LYP 0.0 127 2BHLYP 0.0 122 125

MP2a 0.0 122 128

RI-MP2 0.0 120 125

SCS-MP2a 0.0 111 121

MP2/D95**b 0.0 14 2BHLYP (e 5 3) 0.0 115 115

BHLYP (e 5 10) 0.0 112 19

BHLYP (e 5 40) 0.0 110 16

BHLYP (e 5 78) 0.0 110 16

aSingle-point calculations on RI-MP2/TZVPP optimized geometries.bThese recalculated energies employing Gaussian-03 are in excellent

agreement with the energies published by Dannenberg et al.11

Table 2. Dipole Moments in Debye of Acetylacetone Tautomers

Calculated in Gas Phase or Solvent (COSMO Calculations).

E K1 K2

MP2 2.98 1.50 3.80

BHLYP 3.15 1.65 3.77

BHLYP (e 5 3) 3.73 1.96 4.98

BHLYP (e 5 40) 4.36 2.29 6.45

667Predicting the Tautomeric Equilibrium of Acetylacetone in Solution

Journal of Computational Chemistry DOI 10.1002/jcc

Page 4: Predicting the tautomeric equilibrium of acetylacetone in solution. I. The right answer for the wrong reason?

energy. Because the varying bonding situation the zero-point

energy of the enol form is somewhat greater than that of the

keto form. Hence, the difference between both forms decreases

if one goes from DE to DH. The nonelectrostatic part of the sol-

vation free energy DEn.e. is composed of cavitation energy, dis-

persion energy, and repulsion energy terms. The keto forms

show a slightly larger cavity that is created within the polariz-

able continuum (�127 A3 vs. 125 A3). The larger volume of the

keto forms results in an increased contribution to the cavitation

and dispersion energy, whereas the repulsion energy remains

nearly unchanged. In total, the nonelectrostatic contributions to

the solvation free energy favor the enol form.

The Gibbs free energies show an energy difference of 111

kJ mol21 between the keto form K2 and the enol form at 273 K

in a dielectric continuum with e 5 3. The NMR experiment

measured a keto:enol ratio of about 1:4 for temperatures higher

than 173 K resulting in a free energy difference of about 13 kJ

mol21. At 123 K and with e 5 40 the keto form K2 is still pre-

dicted to be less stable than E, but the value decreases to 6 kJ

mol21. The experiment, however, clearly shows a trend reversal

at 123 K making the K2 tautomer more stable than the enol

form E (22 kJ mol21). The presented approach, thus, underesti-

mates the keto form in both cases by 8 kJ mol21.

The gas phase values in Table 1 indicate that the BHLYP

relative electronic energies of E are in general overestimated by

3–4 kJ mol21. Taking this into account, we estimated the K2

tautomer to be 7 kJ mol21 (273 K, e 5 3) and 2 kJ mol21 (123

K, e 5 40) less stable than E. This means that the stability of

the enol form in solvent is overestimated by 4 kJ mol21 for

both temperatures resp. dielectrics. In contrast for the relative

stability of the keto tautomer with respect to the enol form

(about 5 kJ mol21) experiment and theory agree very well.

Ishida et al.13 reproduced the experimentally found shift from

the enol form to the keto form. However, on the basis of our

gas phase findings, we suspected that the agreement results by

chance from the deficiency of the small double f basis set

{(9s5p1d/4s1p)/[3s2p1d/2s1p]} used in this study. To investigate

this assumption, the correlation energies for both gas phase and

solvation of the keto form K2 and the enol form E have been

determined by MP2 single-point calculations on HF/TZVPP

optimized structures using different basis sets of double f and

triple f size with varying amount of polarization. The valence

double f basis D95v(d,p) basis set is very similar to that used by

Ishida et al.

Table 4 reveals that the energies of the MP2/TZVPP single-

point calculations are already in very good agreement to the val-

ues obtained by full structure optimizations on MP2 and BHLYP

level of theory (Table 1). This clearly shows that the structural

parameters are alike for all methods. However, the relative HF

energy DE(K2-E) is very sensitive with respect to the quality of

the basis set. Table 4 indicates that the inclusion of polarization

on hydrogen atoms seems to be very crucial as the enol form is

stabilized stronger in gas phase by increasing p-function admix-

ture from SV(P) (216 kJ mol21) to TZVPP (14 kJ mol21). The

SVP basis set seems to overestimate the HF energy of the enol

form, which also holds for all solvated structures. Moreover, for

gas phase conditions, the inclusion of correlation effects strongly

favors the enol form. This effect is more than twice as strong

for a triple f basis as for a double f basis. For TZVPP this

amounts to 19 kJ mol21, whereas both SVP and D95v(d,p) give

a correlation energy of only 9 kJ mol21.

In solution, the same trends as in gas phase can be observed,

however, the relative energies are all shifted toward a stabiliza-

tion of the keto form K2. As is clearly seen, the trend reversal

that has been observed experimentally can only be obtained for

the D95v(d,p) basis set. Only for this basis the keto form

Table 3. Relative Thermodynamic Corrections (in kJ mol21) of

Optimized Structures of Acetylacetone Tautomers Calculated in Gas

Phase (Analytically) or in Solvent (Numerically) on a BHLYP/TZVP

(Gas Phase) and a RI-BHLYP/TZVPP (Solvent) Level of Theory. The

Nonelectrostatic Terms (DEn.e.) Have Been Determined by BHLYP/

TZVP Calculations.

DEne DEeleca ZPVE T [K] DHcorr b 2TDScorr b DG

e 5 0 D(K1-E) – 118 13 273 22 24 112

123 22 21 114

D(K2-E) – 122 14 273 22 25 115

123 23 22 117

e 5 3 D(K1-E) 14 115 14 273 22 25 113

123 23 22 115

D(K2-E) 13 115 15 273 23 25 111

123 24 21 114

e 5 40 D(K1-E) 14 110 13 273 22 24 19

123 23 21 111

D(K2-E) 14 16 14 273 23 24 14

123 23 21 16

aGas phase: CCSD(T) value on BHLYP optimized structure.bMinus-sign indicates that the difference becomes smaller, i.e., the keto

form is more stabilizied than the enol form.

Table 4. Relative Energies DE(K2-E) (in kJ mol21) of Hartree-Fock (HF) Optimized Structures of

Acetylacetone Tautomers Calculated in Gas Phase and Continuum Solvation (COSMO) as a Function of the

Basis Set Size.

HF MP2

SV(P) D95v(d,p) SVP TZVPP1 SV(P) D95v(d,p) SVP TZVPPa

Gas phase 216 11 17 14 27 110 114 123

e 5 40 233 219 212 216 219 25 11 18

e 5 78 233 219 212 217 219 25 11 17

aRI approximation (frozen-core).

668 Schlund et al. • Vol. 31, No. 4 • Journal of Computational Chemistry

Journal of Computational Chemistry DOI 10.1002/jcc

Page 5: Predicting the tautomeric equilibrium of acetylacetone in solution. I. The right answer for the wrong reason?

becomes more stable in a polar continuum than the enol form,

which indeed is in agreement with the experimental findings, but

for the wrong reasons. The SVP basis that is often used as a

compromise between costs and accuracy gives a better descrip-

tion. Nevertheless, it still overestimates the stability of the keto

form by 6–9 kJ mol21. The SV(P) basis (standard basis in the

Turbomole package) is describing the enol form already in gas

phase as being less stable than the keto form due to the lack of

polarization functions on the hydrogen atoms.

Molecular Solvent Effects on Acetylacetone

Tautomerism

The conductor-like screening model (COSMO) used in the pres-

ent approach can only give information about the polarization

effect of the continuum on the electronic structure of the solute,

but the molecular nature of the solvent is not explicitly taken

into account. Because both acetylacetone tautomers E and K2

possess two hydrogen bonding acceptor sites, it is very likely

that directed hydrogen bonds to a protic surrounding are formed

(we cannot exclude that different continuum approaches may

lead to different trends. One possibility would be approaches

with terms which take directed interaction more accurately into

account. Other effects may also result from differences in the

treatment of nonelectrostatic effects as mentioned in ref. 22).

The remaining differences between experiment and theory indi-

cate that differences in the strength of these hydrogen bonds

lead to the trend reversal that is found experimentally. This may

happen since the intramolecular hydrogen bond has to compete

with the intermolecular interactions between solvent and solute.

To estimate the influence of such hydrogen bonds on the relative

energies, we computed clusters in which the tautomers of acetyl-

acetone are explicitly solvated with up to six water molecules.

The rest of the solvation shell is approximated by a regular con-

tinuum model (COSMO) of either low (e 5 3) or high (e 5 40)

polarity. The conformational distribution of the water molecules

has been investigated on force field level and the most promis-

ing structures were fully optimized on a DFT level (BHLYP/

TZVPP/COSMO). Exemplary structures for tautomers sur-

rounded by six extra water molecules are shown in Figure 3.

Such cluster approaches can only provide trends because a com-

putations of free energies need to take all low lying minima into

account. In addition, more explicit water molecules would be

needed. Hence, we only concentrated on the shifts in the relative

energies.

Figure 4 shows the computed relative energies DE (K2-E) as

a function of the number of water molecules. The relative ener-

gies of the keto form show a significant decrease from initially

16 kJ mol21 (no water molecule) to 11 kJ mol21 (six water

molecules) surrounded by a continuum environment (e 5 40),

but even more water molecules are necessary for shift in the tau-

tomeric equilibrium which is found experimentally. An extrapo-

lation of Figure 4 indicates that at least eight surrounding water

molecules are necessary, but we would like to point out that

entropic effects which are not included may also be important.

A similar behavior is expected for other hydrogen donating sol-

vents like CHClF2/CHF3 (Freon).

Conclusion

Using state-of-the-art ab initio methods, this work investigates

how the keto-enol equilibrium of acetylacetone is influenced by

a varying polarity of the solvent. Unlike previous calculations,

the experimentally observed trend reversal for acetylacetone is

not found by using a continuum solvation model (COSMO). In

our opinion, this is due to two important reasons:

A basis set of double f size as used in various studies before

leads to an underestimation of the hydrogen bond strength of the

enol form, which erroneously results in the expected change in

the keto:enol molar ratio on solvation in a polar solvent (which

in turn further stabilizes the keto form due to its large dipole

moment).

Interactions arising from the molecular nature of the solvent,

which are not included in most continuum models such as

COSMO, are responsible for an overestimation of the hydrogen

bond in solvent. We could show that the inclusion of several

explicit solvent molecules (‘‘supermolecular ansatz’’) leads to a

lowering of the energy gap between keto and enol forms. By

extrapolation of the calculated energies, we expect a trend rever-

Figure 3. Examples for optimized structures of acetylacetone tauto-

mers E and K2 in continuum solvation (e 5 40) with additional

explicit six water molecules (BHLYP/TZVPP/COSMO).

Figure 4. Relative energies DE (K2-E) in kilojoules per mole of

optimized structures of acetylacetone in continuum solvation with

additional explicit water molecules calculated on a BHLYP/TZVPP/

COSMO level of theory.

669Predicting the Tautomeric Equilibrium of Acetylacetone in Solution

Journal of Computational Chemistry DOI 10.1002/jcc

Page 6: Predicting the tautomeric equilibrium of acetylacetone in solution. I. The right answer for the wrong reason?

sal for a solvation shell that comprises at least eight protic sol-

vent molecules.

In summary, the trend reversal in the equilibrium of keto-

enol tautomers of acetylacetone is footed on shifts in relative

stabilities of the keto and enol forms. Standard quantum chemi-

cal approaches in combination with continuum solvation models

are not able to predict this reversal. However, if the dielectric

constant is increased from e 5 3 to e 5 40, the relative shift in

the energy difference between the keto and enol tautomer is

reproduced very well (5 kJ mol21). This indicates that trends are

already covered by bulk effects, but absolute values of the keto-

enol energy difference are systematically overestimated by about

4 kJ mol21. This indicates that the change in tautomeric popula-

tion for different solvation conditions result from molecular sol-

vent effects that further destabilize the enol form. In order to

describe the shift correctly, it is necessary to calculate the vari-

ous tautomers within an explicit solvent shell by means of quan-

tum chemical methods that are known for treating hydrogen

bonding interactions and correlation effects correctly.

References

1. Reichardt, C. Solvents and Solvent Effects in Organic Chemistry,

3rd ed.; Wiley-VCH: Weinheim, 2003.

2. (a) Cook, G.; Feltman, P. M. J Chem Educ 2007, 84, 1827; (b)

Mills, S. G.; Beak, P. J Org Chem 1985, 50, 1216; (c) Emsley, J.;

Freeman, N. J. J Mol Struct 1987, 161, 193.

3. (a) Tomasi, J.; Persico, M.; Chem Rev 1994, 94, 2027; (b) Tomasi,

J.; Mennucci, B.; Cammi, R. Chem Rev 2005, 105, 2999.

4. (a) Miertus, S.; Scrocco, E.; Tomasi, J. Chem Phys 1981, 55, 117; (b)

Klamt, A.; Schuurmann, G. J Chem Soc Perkin Trans 2, 1993, 5, 799.

5. (a) Warshel, A.; Levitt, M. J Mol Biol 1976, 103, 227; (b) Cramer, C.

J. Essentials of Computational Chemistry; Wiley: Chichester, 2002.

6. Bakowies, D.; Thiel, W. J Phys Chem 1996, 100, 10580.

7. (a) Spencer, J. N.; Holmboe, E. S.; Kirshenbaum, M. R.; Firth, D.

W.; Pinto, P. B. Can J Chem 1982, 60, 1178; (b) Hibbert, F.; Ems-

ley, J. Hydrogen Bonding and Chemical Reactivity, Vol. 26; Aca-

demic Press: Amsterdam, 1990; p 256; (c) S. L. Wallen, C. R.

Yonker, C. L. Phelps, C. M. Wai, J. Chem. Soc., Faraday Trans.

1997, 93, 2391; (d) Moriyasu, M. Kato, A. Hashimoto, Y. J Chem

Soc, Faraday Trans 2 1986, 82, 515.

8. (a) Siegel, J. S.; Anet, F. A. L. J Org Chem 1988, 53, 2629; (b) Golu-

bev, N. S.; Denisov, G. S. J Mol Struct 1992, 270, 263; (c) Weisz, K.;

Jahnchen, J. Limbach, H.-H. J Am Chem Soc 1997, 119, 6436; (d)

Dunger, A.; Limbach, H.-H.; Weisz, K. Chem Eur J 1998, 4, 621; (e)

Smirnov, S. N.; Benedict, H.; Golubev, N. S.; Denisov, G. S.; Kreevoy,

M. M.; Schowen, R. L.; Limbach, H.-H. Can J Chem 1999, 77, 943.

9. (a) Golubev, N. S.; Denisov, G. S.; Smirnov, S. N.; Shchepkin, D.

N.; Limbach, H.-H. Z Phys Chem 1996, 196, 73; (b) Shenderovich,

I. G.; Burtsev, A. P.; Denisov, G. S.; Golubev, N. S.; Limbach, H.-

H. Magn Reson Chem 2001, 39, S91–S99.

10. (a) Buemi, G.; Gandolfo, C. J Chem Soc Faraday Trans 2 1989, 85,

215; (b) Rios, M.; Rodriguez, J. J Mol Struct (Theochem) 1990, 204,

137; (c) Karelson, M. Maran, U. Tetrahedron 1996, 34, 11325; (d)

Bauer, S. H.; Wilcox, C. F. Chem Phys Lett 1997, 279, 122; (e) Buemi,

G.; Zuccarello, F.; Electron J Theor Chem 1997, 2, 302; (f) Sliznev, V.

V.; Lapshina, S. B.; Girichev, G. V. J Struct Chem 2006, 47, 220.

11. Dannenberg, J. J.; Rios, R. J Phys Chem 1994, 98, 6714.

12. Cramer, C. J.; Truhlar, D. G. Solvent Effects and Chemical Reactiv-

ity; Tapia, O.; Bertran, J., Eds.; Kluwer: Dordrecht, 1996.

13. Ishida, T.; Hirata, F.; Kato, S. J Chem Phys 1999, 110, 3938.

14. (a) Ten-no, S.; Hirata, F.; Kato, S. Chem Phys Lett 1993, 214, 391;

(b) Ten-no, S.; Hirata, F.; Kato, S. J Chem Phys 1994, 100, 7443;

(c) Sato, H.; Hirata, F.; Kato, S. J Chem Phys 1996, 105, 1546; (d)

Yokogawa, D.; Sato, H.; Sakaki, S. J Chem Phys 2007, 126,

244504; (e) Yamazaki, T.; Sato, H.; Hirata, F. J Chem Phys 2001,

115, 8949; (f) Sato, H.; Hirata, F. J Chem Phys 1999, 115, 8545; (g)

Sato, H.; Hirata, F. J Chem Phys 2001, 115, 8949; (g) Yamazaki,

T.; Sato, H.; Hirata, F. Chem Phys Lett 2000, 325, 668.

15. Yokagawa, D.; Sato, H.; Imai, T.; Sakaki, S. J Chem Phys 2009,

130, 064111.

16. (a) Schlund, S.; Mladenovic, M.; Basılio Janke, E. M.; Engels, B.;

Weisz, K. J Am Chem Soc 2005, 127, 16151; (b) Hupp, T.; Sturm,

Ch.; Basılio Janke, E. M.; Perez Cabre, M.; Weisz, K.; Engels, B. J

Phys Chem A, 2005, 109, 1703; (c) Baslio Janke, E. I.; Schlund, S.;

Paasche, A.; Engels, B.; Dede, R.; Hussain, I.; Langer, P.; Rettig,

M.; Weisz, K. J Org Chem DOI: 10.1021/jo9004475; (d) Schlund,

S.; Schmuck, C.; Engels, B. J Am Chem Soc, 2005, 127, 11115; (e)

Schlund, S.; Muller, C.; Graßmann, C.; Engels, B. J Comput Chem

2008, 29, 407; (f) Schlund, S.; Schmuck. C.; Engels, B. Chemistry

2007, 13, 6644.

17. (a) Musch, P. W.; Engels, B. J Am Chem Soc 2005, 123, 5557; (b) Hel-

ten, H.; Schirmeister, T.; Engels, B. J Org Chem 2005, 70, 233; (c) Hel-

ten, H.; Schirmeister, T.; Engels, B. J Phys ChemA 2004, 108, 9442.

18. (a) Engels, B.; Peyerimhoff, S. D. J Phys Chem 1989, 93, 4462; (b)

Engels, B. J Chem Phys 1994, 100, 1380; (c) Bundgen, P.; Engels,

B.; Peyerimhoff, S. D. Chem Phys Lett 1991, 176, 407.

19. (a) Ahlrichs, R.; Bar, M.; Haser, M.; Horn, H. Kolmel, C. Chem Phys

Lett 1998, 162, 165; (b) Haser, M.; Ahlrichs, R. J Comput Chem 1989,

10, 104; (c) Haase, F.; Ahlrichs, R. J Comput Chem 1993, 14, 907; (d)

Treutler, O.; Ahlrichs, R. J Chem Phys 1995, 102, 346; (e) Deglmann,

P.; May, K.; Furche, F.; Ahlrichs, R. Chem Phys Lett 2004, 384, 103;

(f) Weigend, F.; Haser, M. Theor Chem Acc 1997, 97, 331; (g) Degl-

mann, P.; Furche, F.; Ahlrichs. R. Chem Phys Lett 2002, 362, 511; (h)

Deglmann, P.; Furche, F. J Chem Phys 2002, 117, 9535; (i) Schafer,

A.; Huber, C; Ahlrichs, R. J Chem Phys 1994, 100, 5829; (j) Vahtras,

O.; Almlof, J.; Feyereisen, M. W. Chem PhysLett 1993, 213, 514; (k)

Eichkorn, K.; Treutler, O.; Ohm, H.; Haser, M.; Ahlrichs, R. Chem

Phys Lett 1995, 242, 652; (l) Schafer, A.; Klamt, A.; Sattel, D.; Loh-

renz, J. C. W.; Eckert, F. Phys Chem Chem Phys 2000, 2, 2187.

20. Werner, H.-J.; Knowles, P. J.; Lindh, R.; Manby, F. R.; Schutz, M.;

Celani, P.; Korona, T.; Rauhut, G.; Amos, R. D.; Bernhardsson, A.;

Berning, A.; Cooper, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert,

F.; Hampel C.; Hetzer, G.; Lloyd, A. W.; McNicholas, S. J.; Meyer,

W.; Mura, M. E.; Nicklass, A.; Palmieri, P.; Pitzer, R.; Schumann, U.;

Stoll, H.; Stone, A. J.; Tarroni, R.; Thorsteinsson, T.MOLPRO, Ver-

sion 2006.1, a Package of ab Initio Programs.

21. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M.

A.; Cheeseman, J. R.; Montgomery, J. A., Jr; Vreven, T.; Kudin, K. N.;

Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; Men-

nucci, B.; Cossi, M.; Scalmani, G.; Rega, N.; Petersson, G. A.; Nakatsuji,

H.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida,

M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Klene, M.; Li, X.;

Knox, J. E.; Hratchian, H. P.; Cross, J. B.; Bakken, V.; Adamo, C.; Jara-

millo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.;

Cammi, R.; Pomelli, C.; Ochterski, J. W.; Ayala, P. Y.; Morokuma, K.;

Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Zakrzewski, V. G.; Dapprich,

S.; Daniels, A. D.; Strain, M. C.; Farkas, O.; Malick, D. K.; Rabuck, A.

D.; Raghavachari, K.; Foresman, J. B.; Ortiz, J. V.; Cui, Q.; Baboul, A.

G.; Clifford, S.; Cioslowski, J.; Stefanov, B. B.; Liu, G.; Liashenko, A.;

Piskorz, P.; Komaromi, I.; Martin, R. L.; Fox, D. J.; Keith, T.; Al-Laham,

M. A.; Peng, C. Y.; Nanayakkara, A.; Challacombe, M.; Gill, P. M. W.;

Johnson, B.; Chen, W.; Wong, M.W.; Gonzalez, C.; Pople, J. A. Gaussian

03, Revision C. 02; Gaussian, Inc.: Wallingford, CT, 2004.

22. Cramer, C. J.;, Truhlar, D. G. Acc Chem Res 2008, 41, 760.

670 Schlund et al. • Vol. 31, No. 4 • Journal of Computational Chemistry

Journal of Computational Chemistry DOI 10.1002/jcc