+ All Categories
Transcript

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 5105

Cite this: Phys. Chem. Chem. Phys., 2011, 13, 5105–5113

Halogen bonded complexes between volatile anaesthetics (chloroform,

halothane, enflurane, isoflurane) and formaldehyde: a theoretical study

Wiktor Zierkiewicz,*aRobert Wieczorek,

bPavel Hobza

cdand Danuta Michalska

a

Received 8th October 2010, Accepted 12th January 2011

DOI: 10.1039/c0cp02085k

The structures and intermolecular interactions in the halogen bonded complexes of anaesthetics

(chloroform, halothane, enflurane and isoflurane) with formaldehyde were studied by ab initio

MP2 and CCSD(T) methods. The CCSD(T)/CBS calculated binding energies of these complexes

are between �2.83 and �4.21 kcal mol�1. The largest stabilization energy has been found

for the C–Br� � �O bonded halothane� � �OCH2 complex. In all complexes the C–X bond length

(where X = Cl, Br) is slightly shortened, in comparison to a free compound, and an increase

of the C–X stretching frequency is observed. The electrostatic interaction was excluded as being

responsible for the C–X bond contraction. It is suggested that contraction of the C–X bond

length can be explained in terms of the Pauli repulsion (the exchange overlap) between the

electron pairs of oxygen and halogen atoms in the investigated complexes. This is supported by

the DFT-SAPT results, which indicate that the repulsive exchange energy overcompensates the

electrostatic one. Moreover, the dispersion and electrostatic contributions cover about 95% of

the total attraction forces, in these complexes.

1. Introduction

Anaesthesia is a reversible phenomenon and general

anaesthetics act by perturbing weak intermolecular inter-

actions without breaking or forming covalent bonds. Despite

the fact that volatile anaesthetics, such as chloroform,

halothane, enflurane and isoflurane, have been used clinically,

the mechanism of their action is still not fully understood.

They may act by a direct bonding to neuroreceptors.1–6

Sandorfy1 reported that they can form weak van der Waals

complexes or intermolecular hydrogen bonds, and the strength

of these interactions is in the range between 1.0 and

2.2 kcal mol�1. Investigation of the halogen bonds and

hydrogen bonds may help to explain anaesthetic properties

of polyhalogenated alkanes and ethers.7–9 A number of

experimental and theoretical evidences confirm that the halogen

bond plays an important role in a wide variety of biological

phenomena such as protein–ligand complexation.10–15

Eckenhoff and coworkers16 reported the crystal structure of

the halothane/apoferritin and isoflurane/apoferritin complexes.

For the halothane complex it has been shown that in the binding

pocket the distance between the bromine atom of halothane and

the carbonyl oxygen atom of leucine (3.11 A) is smaller than the

sum of the corresponding van der Waals radii (3.37 A).

Studies of the electrostatic potentials of the halogen bonded

systems show that the lone electron pairs of the halogen atom

bonded to the carbon atom form a belt of negative

electrostatic potential around its central part leaving the outer-

most region positive, the so-called s-hole.17,18 The halogen

bonding was explained as a noncovalent interaction between a

covalently bound halogen on one molecule and a negative site

of another.7,19–21 Recently, it has been shown that the

interaction energy in these complexes is dominated by the

dispersion and electrostatic contributions.22–24 Halogen bonds

are similar to hydrogen bonds in many respects. Therefore,

they may collaborate or compete with each other.25–30

The aim of this work is to elucidate the nature of inter-

actions in the chloroform, halothane, enflurane and isoflurane

complexes with formaldehyde using ab initio MP2 and

CCSD(T) methods. The studies on the halogen bond inter-

actions between anaesthetics and amino acids are important

for understanding their mechanism of action, in biological

systems. Therefore, the complexes investigated in this work

can serve as the models for such studies.

The structures, binding energies and harmonic frequencies

of these complexes are discussed. The C–X (XQCl, Br) bond

in anaesthetics becomes shorter and stronger upon formation

of a complex with formaldehyde. The possible explanation of

this phenomenon is presented.

a Faculty of Chemistry, Wroc!aw University of Technology,Wybrzeze Wyspianskiego 27, 50-370 Wroc!aw, Poland.E-mail: [email protected]; Fax: +48 71-320-4360;Tel: +48 71-320-3455

b Faculty of Chemistry, University of Wroclaw, F. Joliot-Curie 14,50-383 Wroclaw, Poland

c Institute of Organic Chemistry and Biochemistry, Academy ofSciences of the Czech Republic and Center for Biomolecules andComplex Molecular Systems, 166 10 Praha 6, Czech Republic

dDepartment of Physical Chemistry, Palacky University,771 46 Olomouc, Czech Republic

PCCP Dynamic Article Links

www.rsc.org/pccp PAPER

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

5106 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 This journal is c the Owner Societies 2011

2. Theoretical methods

Full geometry optimizations of the isolated anaesthetic

molecules (chloroform, halothane, enflurane, isoflurane) and

their halogen bonded complexes with formaldehyde were

performed by the ab initio second order Møller–Plesset

perturbation (MP2) method31 using the 6-311++G(d,p) basis

set.32,33 The counterpoise CP-corrected gradient optimization

has been used.34 Subsequently, vibrational harmonic

frequencies and infrared intensities have been calculated for

all species considered in this work.

To elucidate the relativistic effects for the halothane� � �formaldehyde complex with the C–Br� � �O halogen bond, we

have performed additional calculations using the LanL2DZdp

pseudopotential basis set35 for the bromine atom and the

6-311++G(d,p) basis set for the other atoms. Moreover, the

geometry of the halothane� � �formaldehyde complex has been

optimized at the MP2/cc-pVTZ level. Further, the interaction

energy has been calculated using both the aug-cc-pVTZ basis

set36,37 on all atoms and mixed aug-cc-pVTZ-PP basis set38 on

the bromine atom and aug-cc-pVTZ on other atoms.

The CCSD(T) complete basis set (CBS) limit interaction

energies have been obtained as a sum of the MP2/CBS

interaction energies and the CCSD(T) correction term, which

is defined as a difference between the MP2 and CCSD(T)

interaction energies determined in a smaller basis set (6-31G*).39,40

The CBS limit energies were evaluated by an extrapolation

of the Hartree–Fock energy and the MP2 correlation energy

from the cc-pVTZ and cc-pVQZ basis sets.36,37 The two-point

extrapolation methods of Helgaker et al. were used.41,42

To get a detailed insight into the nature of bonding in

these complexes, the symmetry adapted perturbation theory

DFT-SAPT43 calculations were carried out at the PBE0/aug-cc-

pVTZ level.44 The SAPT method determines the total inter-

action energy as the sum of the first- and second-order

perturbation terms. The favorable performance of the method

is due to a combination of the DFT treatment for monomers

and SAPT treatment for interaction between monomers.

A natural bond orbital (NBO) analysis45 was performed at

the MP2/6-311++G(d,p) level of theory. To calculate the

positive sigma hole potentials, the electrostatic potential

surface maxima on halogen atoms of isolated s-hole donors

have been computed at the B3PW91/6-31G(d,p) level using the

WFA (Wavefunction Analysis Program).46

All calculations were carried out with the GAUSSIAN 0947

or MOLPRO 200648 programs.

3. Results and discussion

3.1 Structures

Fig. 1 shows the structures of the chloroform� � �OCH2 and

halothane� � �OCH2 complexes optimized at the MP2/

6-311++G(d,p) level. It should be mentioned that the most

stable conformer of halothane has been selected for investiga-

tions.49–52 As is seen from this figure, two structural isomers of

the halothane� � �formaldehyde complex have been found. The

(A) complex with the C–Br� � �O halogen bond is more

stable, by 0.44 kcal mol�1, than the (B) complex with the

C–Cl� � �O bond.

Fig. 2 illustrates the structures of the formaldehyde

complexes with the most stable conformers of enflurane53–56

and isoflurane. In the case of the isoflurane� � �OCH2 complex,

the MP2 calculations have revealed the presence of two

structural isomers, (A) and (B), where (A) is more stable by

0.16 kcal mol�1. The formation of these two isomers can be

explained as a consequence of the weak secondary interaction

between the hydrogen atom of formaldehyde and the fluorine

atom of isoflurane.

Tables 1 and 2 list the selected geometrical and vibrational

parameters for the investigated complexes. As follows from

Table 1, the C1–X2 bond (where X = Cl, Br) is contracted

upon complexation. In the chloroform and isoflurane

complexes the C1–Cl2 bond length is contracted by �0.003 A,

while the smallest change (�0.001 A) is observed in the

halothane� � �OCH2 (A) complex. As follows from Table 2, a

Fig. 1 Optimized structures of chloroform� � �formaldehyde and two

halothane� � �formaldehyde (A and B) complexes with selected atom

numbering. Calculations were performed at the MP2/6-311++G(d,p)

level.

Fig. 2 Optimized structures of enflurane� � �formaldehyde and two

isoflurane� � �formaldehyde (A and B) complexes with selected atom

numbering. Calculations were performed at the MP2/6-311++G(d,p)

level.

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 5107

contraction of the C1–X2 bond is concomitant with a small

(up to 4 cm�1) increase of the C–X stretching frequency

(blue-shift). For the enflurane� � �OCH2 complex, the infrared

intensity of the corresponding mode decreases by�15 kmmol�1.

In other complexes the decrease of the IR intensity is smaller

(from 0 to �5 km mol�1).

Recently, van der Veken and coworkers57 have confirmed

experimentally the small (+1.7 cm�1) blue-shift of the n(C–Cl)stretching frequency in the trifluorochloromethyl dimethyl

ether complex containing the C–Cl� � �O halogen bond. The

MP2/6-311++G(d,p) calculated contraction of the C–Cl bond

was �0.006 A.

As shown in Table 1, upon complexation, the C3–O4 bond is

elongated by 0.001 A, in all complexes. This is accompanied by

a small (up to 2 cm�1) red-shift of the n(C3–O4) stretching

frequency and the decrease of the infrared intensity of this

mode, by about �11 km mol�1 in all complexes, except for

the halothane� � �formaldehyde (A) complex, where the IR

intensity decreases only by �1 km mol�1 (Table 2).

The sum of the van der Waals radii of the chlorine–oxygen

and the bromine–oxygen atoms is 3.27 and 3.37 A,

respectively.58 As follows from Table 1, in the chloroform

and halothane (A) and (B) complexes, the intermolecular

X2� � �O4 distances are smaller than the corresponding sum of

the vdW radii.

In the enflurane� � �OCH2 and isoflurane� � �OCH2 complexes

the Cl2� � �O4 distance is equal to (or is slightly larger than) the

sum of the vdW radii. Despite this fact, the theoretical

methods have predicted that these complexes are stable.

Auffinger et al.10 on the basis of the Protein Data Basis

(PDB) survey have reported that the average Br� � �O distance

(3.15 A) is longer, by 0.09 A, than the average Cl� � �O distance.

However, in the halothane–formaldehyde (A) complex the

MP2 calculated Br� � �O distance (3.14 A) is shorter (by 0.06 A)

than the Cl� � �O distance in the (B) complex, as shown in Table 1.

A similar effect was reported for the complexes of hypohalous

acids with formaldehyde59 and for the F3CBr� � �OCH2 and

F3CCl� � �OCH2 complexes.60 On the other hand, calculations

performed at the MP2/6-311++G(d,p) level for the F3CBr and

F3CCl complexes with dimethyl ether predicted the same Br� � �Oand Cl� � �O atom distances (2.89 A),57 whereas MP2 calculations

of the halothane� � �dimethyl ether predicted a longer Br� � �Odistance than Cl� � �O, by about 0.05 A.61 This comparison shows

that the intermolecular Br� � �O and Cl� � �O distances are different,

in various systems.

The sum of the van der Waals radii of the chlorine–

hydrogen and bromine–hydrogen atoms is 2.95 and 3.05 A,

respectively.58 In the CHCl3� � �OCH2, CHClBrCF3� � �OCH2 (A),

CHClBrCF3� � �OCH2 (B), enflurane� � �OCH2, isoflurane� � �OCH2 (A) and isoflurane� � �OCH2 (B) complexes, the shorter

distance between the H atom of OCH2 and the halogen atom is

equal: 3.29, 3.53, 3.29, 2.99, 3.91 and 3.39 A, respectively. All

these distances are larger than the corresponding sum of the

vdW radii. Nevertheless, some weak secondary interactions

may also occur in these complexes. Especially, in the

enflurane� � �OCH2 complex, where the closest H� � �F distance

is only slightly larger (0.04 A) than the sum of the vdW

radii.

In biological molecules with the halogen bond, the average

C–X� � �O and X� � �O–C angles are 1601–1801 and about 901,

respectively.10 The results collected in Table 1 show that the

C1–X2� � �O4 angles in the C–Cl� � �O bonded complexes are

smaller than the C–Br� � �O angle in the halothane (A) complex.

Table 1 Selected structural parameters of the anaesthetic� � �formaldehyde complexes (distances r, d, in A, angle y in 1). Results from the MP/6-311++G(d,p) and the MP/6-311++G(d,p)/LanL2DZdp calculations

r(C1–X2) r(C3–O4) d(X2� � �O4) y(C1–X2� � �O4) y(X2� � �O4–C3)

Chloroform� � �OCH2 1.762 1.214 3.23 166.9 101.2(�0.003)a (+0.001)

Halothane� � �OCH2 (A) 1.930 1.214 3.14 171.1 110.8(�0.001) (+0.001)

Halothane� � �OCH2 (A)b 1.934 1.214 3.13 171.1 110.6(�0.001) (+0.001)

Halothane� � �OCH2 (B) 1.758 1.214 3.20 164.5 101.8(�0.002) (+0.001)

Enflurane� � �OCH2 1.750 1.214 3.30 154.1 97.3(�0.002) (+0.001)

Isoflurane� � �OCH2 (A) 1.764 1.214 3.27 160.5 98.3(�0.003) (+0.001)

Isoflurane� � �OCH2 (B) 1.764 1.214 3.28 158.6 97.8(�0.003) (+0.001)

a In parentheses are shown differences between the values in the complex and in the isolated molecule. b Results from the MP2 calculations with

the LanL2DZdp basis set for the bromine atom and the 6-311++G(d,p) basis set for the other atoms.

Table 2 Selected vibrational parameters of the anaesthetic� � �formaldehyde complexes (vibrational frequency n, in cm�1, infraredintensity I, in km mol�1). Results from the MP/6-311++G(d,p)calculations

n(C1–X2) I(C1–X2) n(C3–O4) I(C3–O4)

Chloroform� � �OCH2 388 (+4)a 1 (0) 1760 (�1) 64 (�10)Halothane� � �OCH2 (A) 739 (+1) 14 (�5) 1759 (�2) 73 (�1)Halothane� � �OCH2 (A)b 741 (+2) 13 (�5) 1758 (�3) 72 (�2)Halothane� � �OCH2 (B) 849 (0) 70 (�5) 1760 (�1) 64 (�10)Enflurane� � �OCH2 862 (+2) 45 (�15) 1759 (�2) 63 (�11)Isoflurane� � �OCH2 (A) 764 (0) 28 (�1) 1760 (�1) 63 (�11)Isoflurane� � �OCH2 (B) 764 (+1) 26 (�3) 1760 (�1) 63 (�11)a In parentheses are shown differences between the values in the

complex and in the isolated molecule. b Results from the MP2 calcula-

tions with the LanL2DZdp basis set for the bromine atom and the

6-311++G(d,p) basis set for the other atoms.

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

5108 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 This journal is c the Owner Societies 2011

The deviation of the C1–X2� � �O4 angles from 1801 can be

attributed to the secondary weak interaction between the

hydrogen atom of formaldehyde and halogen atom or atoms

of the halogen donor. The biggest deviation (25.91) of this

angle has been found in the enflurane� � �OCH2 complex. It

should be mentioned that the H� � �O distance (2.97 A) between

the H atom of formaldehyde and O atom of enflurane is larger,

by 0.25 A, than the sum of the corresponding vdW radii.

The Cl� � �O4–C3 angles are in the range 97.31–101.81, and all

are smaller than the Br� � �O4–C3 angle (110.81) in the

halothane (A) complex. A similar effect was reported earlier

for the halomethane� � �formaldehyde complexes, and it was

attributed to larger contribution of a dispersion energy in the

Cl� � �O interaction.23

To answer the question whether the relativistic effects are

important in the halothane� � �formaldehyde complex (A) with

the C–Br� � �O halogen bond, the additional MP2 calculations

have been performed using the LanL2DZdp pseudopotential

basis set for the bromine atom and the 6-311++G(d,p) basis

set for the other atoms. A similar method was used by

Michielsen et al.61 The results are presented in Tables 1 and 2.

As follows from Table 1 the use of the LanL2DZdp basis set

for Br leads to an elongation of the C1–Br bond by 0.004 A,

and a shortening of the Br� � �O4 distance by 0.01 A. The

stabilization energies of this complex calculated at the

MP2/6-311++G(d,p) and MP2/6-311++G(d,p)/LanL2DZdp

levels are �2.00 and �2.03 kcal mol�1, respectively. Thus,

the difference between the binding energies is only 1.5%,

which indicates that the relativistic effects are negligible for

the halothane� � �formaldehyde complex with the C–Br� � �Obond. However, it should be mentioned that, in the case of

the H3CBr� � �OCH2 complex, the relativistic effects evaluated

at different levels of theory were slightly higher, the MP2

interaction energy of this complex calculated with the aug-cc-

pVTZ-PP basis set on the bromine atom and aug-cc-pVTZ on

all other atoms was larger by 4.2% than that calculated using

the aug-cc-pVTZ on all atoms.23 We have performed calcula-

tions on the same level of theory for the halothane� � �formaldehyde complex (A) and the results are very similar,

the difference between the two interaction energies is 4.5%.

Thus, the relativistic effects in the investigated system are

rather small.

In order to make the comparison between formaldehyde

and formamide as the s-hole acceptors, the chloroform� � �formamide complex has also been studied (Fig. 3).

In the latter, the shortening of the C1–Cl2 bond caused by

complexation (�0.005 A) is larger than that in the chloro-

form� � �formaldehyde complex (�0.003 A). Contraction of the

C1–Cl2 bond is concomitant with an increase of the C–Cl

stretching frequency (+4 cm�1). The infrared intensity of the

corresponding mode increases by +7 km mol�1. The Cl2� � �O3

distance in the chloroform� � �OCHNH2 complex is shorter (by

0.08 A) than that in the chloroform� � �formaldehyde complex.

The C1–Cl2� � �O4 angle (170.21) is larger by 3.31 despite the fact

that the shorter distance between the H atom of OCHNH2 and

the halogen atom (3.29 A) is the same as that in the formalde-

hyde complex.

3.2 NBO analysis

The selected NBO data are listed in Table 3. As follows from

this table, the halogen atom shows the largest change of the

atomic charge, upon complexation. The charge on X2

increases in the range between 0.012 and 0.018 e

(halothane� � �OCH2 (A) complex).

Examination of occupancies of the bonding s(C1–X2) and

antibonding s*(C1–X2) orbitals indicates that the change

in the electron density on these orbitals is negligible (smaller

than �0.002 e), upon complexation.

Fig. 3 Optimized structure of the chloroform� � �formamide complex

with selected atom numbering. Distances in A, angles in 1. Results

from the MP2/6-311++G(d,p) calculations.

Table 3 Charges (q in e) on selected atoms, charge transfer (CT in e) in anaesthetic� � �formaldehyde complexes calculated at the MP2/6-311++G(d,p) level

C1 X2a O4 C3 CTb

Chloroform� � �OCH2 �0.310 0.052 �0.469 0.287 0.002(�0.002)c (+0.013) (�0.005) (+0.001)

Halothane� � �OCH2 (A) �0.402 0.131 �0.473 0.290 0.007(�0.008) (+0.018) (�0.009) (+0.004)

Halothane� � �OCH2 (B) �0.397 0.065 �0.470 0.287 0.003(�0.003) (+0.012) (�0.006) (+0.001)

Enflurane� � �OCH2 0.160 0.036 �0.472 0.286 0.002(�0.007) (+0.013) (�0.008) (0.000)

Isoflurane� � �OCH2 (A) 0.018 0.034 �0.469 0.286 0.002(�0.006) (+0.012) (�0.005) (0.000)

Isoflurane� � �OCH2 (B) 0.018 0.034 �0.468 0.286 0.002(�0.006) (+0.012) (�0.004) (0.000)

a X2 means the bromine atom in the halothane� � �OCH2 (A) complex and chlorine atoms in all other complexes. b Charge transfer from

formaldehyde to a s-hole donor. c In parentheses are shown the changes of the charges caused by complexation, calculated as differences between

the values in the complex and in the isolated molecule.

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 5109

As follows from the last column of Table 3, the charge

transfer (CT) from formaldehyde to halothane (A complex) is

equal to 0.007 e, while in the remaining complexes it is about

twice smaller. It is worth to mention that the values of CT

are similar to those found for weak hydrogen bonds.62

The NBO analysis has been performed also for the

chloroform� � �formamide complex. These results have revealed

that the change of the charge on the chlorine atom caused by

complexation equals +0.023 e, which is much larger than that

in the chloroform� � �formaldehyde complex (+0.013 e). This

fact can be explained as a consequence of different charges on

the O4 atoms in the s-hole acceptors. The charges on the

oxygen atoms in formaldehyde and formamide are �0.464 and�0.564 e, respectively. The larger the electron density (ED) on

the oxygen atom, the larger the decrease of ED on the

chlorine atom.

3.3 Contraction of the C–X2 bond

Recently, it has been suggested that in the halogen bonded

complexes with weak hyperconjugative interaction, the

electrostatic interaction is responsible for contraction of the

Y–X bond (where Y = C, N, Si and X = Cl, Br).63 Thus, we

have also investigated the role of electrostatic and polarization

effects in chloroform and its complex with formaldehyde. The

Merz–Kollman charges of formaldehyde atoms (obtained

from the complex) were placed at the positions of the

respective atoms. In this treatment, the geometry changes were

caused by both electrostatic and polarization interactions

between chloroform and the inhomogenous electric field.

According to these calculations, the C1–Cl2 bond of chloro-

form was only slightly elongated (by 0.0002 A) upon the

electric field of the proton acceptor. In addition, geometry

optimization of the isolated halogen donors was performed in

the uniform electric field parallel to the C1–Cl2 bond. These

results have shown that in the homogenous electric field (in the

range from 0.004 to 0.024 au) the C1–Cl2 bond length is

elongated. This excludes the electrostatic interaction as being

responsible for the contraction of the C1–Cl2 bond length.

To examine the Hermansson64 and the Qian/Krimm65

models for predicting contraction or elongation of the

C1–X2 bond (blue-shifting or red-shifting complexes,

respectively) we calculated the derivatives of the permanent

dipole moments (qm0/qRC1–X2) and their directions for each

anaesthetic, see Table 4. According to these models the change

in the frequency of the C1–X2 bond should be related to the

magnitudes and directions of the derivatives qm0/qRC1–X2and

qmind/qRC1–X2, where mind is the dipole moment induced by the

electric field of the s-hole acceptor. Since qmind/qRC1–X2is in

the direction of the electric field, it always supports an

elongation of the C1–X2 bond.66

As is seen from Table 4 only in the case of halothane

(X2 = Br) the qm0/qRC–Br has a negative value. This indicates

that the negative dipole gradient for the C1–X2 bond is not the

necessary condition for the blue-shift of the C1–X2 stretching

vibration.

In the last column of Table 4 the values of angle between the

dipole moment and the bond from C1 to X2 are collected. An

angle of 1801 correlates with a contraction of this bond

(blue-shifting complexes), while an angle of 01 indicates possi-

bility of an elongation of the bond (red-shifting complexes).66

As follows from this table only in the case of enflurane the

value of this angle (1501) is nearest to 1801. In the other

molecules the dipole derivatives are more or less perpendicular

to the uniform electric field. Thus, it is impossible to derive the

definite conclusions regarding the nature of contraction of the

C1–X2 bond on the basis of these results.

In the case of the blue-shifted hydrogen bonded complexes

Schlegel et al.67 suggested that the Pauli repulsion between two

fragments is responsible for the contraction of the X–H bond.

Now, the question arises: can the analogous Pauli repulsion

(the exchange overlap) effect explain the changes of the C–X

bond length in the investigated complexes?

The halogen atom has three electron pairs which form a belt

of negative electrostatic potential around the central part of

this atom, leaving the outermost region positive (s-hole).17,18

The oxygen atom of formaldehyde has two lone pair orbitals

LP(1)O4 and LP(2)O4. The LP(2)O4 orbital is involved in the

formation of the halogen bond, and it overlaps with the

s*(C1X2) orbital of the investigated anaesthetic. This

orbital interacts with the halogen atom through the s-hole.Simultaneously, electrons of the LP(1)O4 orbital can

repulse the electron pairs on the halogen atom. In our opinion,

this repulsion is responsible for pushing the halogen atom

towards the carbon atom, which makes the C1–X2 bond

shorter.

Fig. 4 illustrates the contours of the selected orbital ampli-

tudes in the plane passing through the C1, Cl2 and O4 atoms, in

the CHCl3� � �OCH2 complex. The overlap of LP(2)O4 with the

s*(C1Cl2) orbitals is shown in Fig. 4A, while the overlap of

LP(1)O4 with LP(1)Cl2 and overlap of LP(1)O4 with LP(2)Cl2are depicted in Fig. 4B and C, respectively. It should be

mentioned that the contour of the third electron pair orbital

of the Cl atom (LP(3)Cl2) is not seen in the selected plane.

If the suggested idea of repulsion is responsible for this

effect, then in the case of a halogen acceptor with only one

electron pair (for instance the N atom), the contraction of the

C–X bond should be smaller (in comparison to the O atom

acting as the halogen acceptor) or elongation of this bond

should be observed.

In the C–X� � �O halogen bonded complexes of F3C–X with

dimethyl ether, the C–X bonds (X=Cl, Br, I) were contracted

by �0.005, �0.006 and �0.004 A, respectively. On the other

hand, in the C–X� � �N halogen bonded complexes of F3C–X

with trimethylamine, the C–X bonds (X = Br, I) were

elongated by 0.005 and 0.015 A, respectively, or negligibly

contracted (by �0.001 A for X = Cl).60,68 Analogous results

Table 4 Derivatives of the permanent dipole moments of anaesthetics(qm0/qRC–X2

in D A�1) and the angle between the dipole moment andthe bond from C1 to X

2(angles in 1). Calculations performed at the

MP2/6-311++G(d,p) level

qm0/qRC–X2Angle

Chloroform (X2= Cl) 0.30 108Halothane (X2= Br) �0.35 84Halothane (X2= Cl) 0.02 93Enflurane (X2= Cl) 2.12 150Isoflurane (X2= Cl) 1.13 106

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

5110 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 This journal is c the Owner Societies 2011

were obtained for the F3Y–X (Y = C, N, Si; X = Cl, Br)

complexes with H2O and NH3.63

These results seem to support our explanation of the origin

of the contraction of the C–X bond in the studied anaesthetics,

upon complexation.

3.4 Binding energies

The binding energies calculated by the CCSD(T)/CBS,

MP2/CBS and HF/CBS methods are compared in Table 5.

As is seen in this table, the interaction energies obtained at the

CCSD(T)/CBS level (DECCSD(T)/CBS) are in the range between

�2.83 and �4.21 kcal mol�1. All the theoretical methods

consistently indicate that the C–Br� � �O halogen bonded

complex of halothane� � �OCH2 (A) has the largest binding

energy. It is known that the magnitude of the binding energy

increases with the increasing halogen size.23 Indeed, in the

halothane� � �OCH2 (A) complex (C–Br� � �O bond) binding

energy is larger, by about 31%, than that in the halothane� � �OCH2 (B) complex (C–Cl� � �O bond), as revealed by the

CCSD(T)/CBS calculations. A similar effect was reported for

the H3CBr� � �OCH2 and H3CCl� � �OCH2 complexes, calcu-

lated at the same level of theory (CCSD(T)/CBS), the former

complex had a stronger binding energy (by about 32%) than

the latter.23 Moreover, for two halothane� � �dimethyl ether

complexes, that with the C–Br� � �O bond had a larger inter-

action energy (�3.82 kcal mol�1) than the other, with the

C–Cl� � �O bond (�2.58 kcal mol�1).61

The CCSD(T) correction terms (calculated as differences

between DECCSD(T)/CBS and DEHF/CBS) are about 60–70% of

the total interaction energies. This indicates that dispersion is

very important in the formation of these complexes.

For the chloroform� � �formamide complex the calculated

CCSD(T)/CBS, MP2/CBS and HF/CBS binding energies are

�3.42, �3.48 and �1.13 kcal mol�1, respectively. Thus,

the chloroform� � �OCHNH2 complex is more stable than

chloroform� � �OCH2 by �0.53 kcal mol�1.

It was shown that the positive surface potential maxima

Vs,max correlate with the interaction energies, in the bromo-

benzene and bromopyridine complexes with acetone.24 To

check this idea, the positive sigma hole potentials for isolated

anaesthetics have been calculated. The most positive values of

Vs,max on the chlorine and bromine atoms in anaesthetics

investigated are listed in Table 6.

For the chlorine atoms, the values of Vs,max range from

8.8 kcal mol�1 (isoflurane) to 14.3 kcal mol�1 (halothane)

while for the bromine atom in halothane Vs,max is

20.8 kcal mol�1. Trogdon et al.9 calculated the Vs,max values

on the chlorine and bromine atoms in the same anaesthetics at

the HF/6-31* level of theory. They obtained slightly larger

values: 14.5, 22.3, 15.3, 13.0, 14.1 kcal mol�1 for chloroform,

halothane (Br), halothane (Cl), enflurane and isoflurane,

respectively.

In the series of anaesthetics investigated in this work we

have found the linear relationship between the interaction

energies DECCSD(T)/CBS and the Vs,max with a correlation

coefficient of 0.77. A rather poor correlation clearly indicates

that the electrostatic term is not the dominant one and the

dispersion and polarization energy terms contribute to a

stabilization of the halogen bond as well.

3.5 Decomposition of interaction energies

The DFT symmetry adapted perturbation theory (DFT-SAPT)

analysis has been performed at the PBE0/aug-cc-pVTZ level of

theory. The results are collected in Table 7. The DFT-SAPT

interaction energies vary between �2.58 and �3.92 kcal mol�1.

In all Cl� � �O halogen bonded complexes considered in this

Fig. 4 Contours of the selected orbitals amplitude in plane passing

through C1, Cl2 and O4 atoms, in the CHCl3� � �OCH2 complex.

Overlap of the LP(2)O4 and s*(C1Cl2) orbitals (A), LP(1)O4 and

LP(1)Cl2 orbitals (B), LP(1)O4 and LP(2)Cl2 orbitals (C).

Table 5 Binding energies calculated by the CCSD(T)/CBS(DECCSD(T)/CBS), MP2/CBS (DEMP2/CBS), HF/CBS (DEHF/CBS) andcorrelation energies. All values are in kcal mol�1

DEHF/CBS DEMP2/CBS DECCSD(T)/CBS DEcorr a

Chloroform� � �OCH2 �0.95 �2.99 �2.89 �1.94Halothane� � �OCH2 (A) �1.61 �4.36 �4.21 �2.60Halothane� � �OCH2 (B) �1.13 �3.29 �3.21 �2.08Enflurane� � �OCH2 �1.39 �3.72 �3.67 �2.28Isoflurane� � �OCH2 (A) �0.91 �2.99 �2.93 �2.03Isoflurane� � �OCH2 (B) �0.75 �2.88 �2.83 �2.08a Difference between DECCSD(T)/CBS and DEHF/CBS.

Table 6 Vs,max (in kcal mol�1) on X2 (X2 = Cl or Br) atoms forisolated anaesthetics calculated at the B3PW91/6-31G(d,p) level

Vs,max on X2

Chloroform (X2 = Cl) 13.8Halothane (X2 = Br) 20.8Halothane (X2 = Cl) 14.3Enflurane (X2 = Cl) 11.1Isoflurane (X2 = Cl) 8.8

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 5111

work, the dispersion ED(2) energy is the dominant attraction

component representing about 50% of the total attraction

forces, while the electrostatic Eel(1) forces account for about

40–45% of the total energy. In the Br� � �O halogen bonded

complex of halothane (A), Eel(1) contributes 51%, while ED

(2)

contributes 38% to the total energy. Similar results were

obtained for the halomethane� � �formaldehyde complexes.23

In the chloroform� � �formamide complex, electrostatic and

dispersion energies contribute about 45% each to the total

energy. For all complexes investigated, the repulsive exchange

energy Eex(1) overcompensates Eel

(1), thus the first-order

energies E(1) are repulsive by less than 0.70 kcal mol�1. The

magnitude of the repulsive exchange energy supports the

conclusion that the Pauli repulsion (the exchange overlap)

effect is responsible for the contraction of the C–X bond in the

anaesthetics upon complexation.

The contribution of the induction Ei(2) energy to the total

attraction energy is about 5%. These results show that the

dispersion and electrostatic contributions cover about 95% of

the total attraction forces, in these complexes.

3.6 Comparison of the halogen bond and hydrogen bond

Both types of the intermolecular interactions share several

characteristics such as strength or the hyperconjugation

between antibonding orbital of the C–X bond (X = Cl, Br)

and the electron pair of the oxygen atom. In this work we

would like to point out the different mechanisms responsible

for the contraction of the C–H bond (in the blue-shifted

hydrogen bonded complexes) and the C–X bond (in the

complexes considered in this work). In both cases the oxygen

atom is the acceptor for the H or X atoms. In the former

complexes the electrostatic interaction plays the crucial role,

while in the latter complexes the repulsion between the

electron pairs of the oxygen and halogen atoms is the driving

force. In both cases the contraction of the C–H or C–X bonds

is accompanied by an increase of the C–H or C–X stretching

frequencies (blue-shift).

4. Conclusions

In this work the structures and binding energies of the halogen

bonded complexes of anaesthetics (chloroform, halothane,

enflurane and isoflurane) with formaldehyde were studied

using ab initio MP2 and CCSD(T) methods. The understand-

ing of the halogen bond interaction between anaesthetics and

amino acids is important for elucidation of their mechanism of

action, in biological systems. Thus, the complexes investigated

in this work can serve as the models for such studies. The most

important conclusions are the following:

(1) All anaesthetics studied in this work make stable

complexes with formaldehyde through the Cl� � �O halogen

bonding. In addition, halothane also forms the Br� � �O halogen

bonded complex (B). According to the MP2 method, the

halothane� � �OCH2 complex with the C–Br� � �O halogen bond

is more stable, by 0.44 kcal mol�1, than that the other one. In

the case of the isoflurane� � �OCH2 complex, the MP2 calcula-

tions have revealed the presence of two structural isomers,

which differ in energy by 0.16 kcal mol�1.

(2) Binding energies calculated by the CCSD(T) complete

basis set (CBS) limit method for six halogen bonded

anaesthetic� � �formaldehyde complexes vary between �2.83and �4.21 kcal mol�1. The largest DECCSD(T)/CBS has been

found for the C–Br� � �O halogen bonded halothane� � �OCH2

(A) complex, while the smallest binding energy has been

obtained for the isoflurane� � �OCH2 (B) complex.

(3) The C–X bond lengths (where X = Cl, Br) are

contracted upon complexation. It is suggested that the main

reason for contraction of the C–X bond is a repulsion between

the lone electron pair orbitals (the exchange overlap) of the

oxygen and halogen atoms.

(4) In the chloroform and two halothane complexes with

formaldehyde, the intermolecular (X� � �O) distance is smaller

than the sum of the corresponding van der Waals radii. In

the enflurane� � �OCH2 and isoflurane� � �OCH2 complexes the

intermolecular distance is equal to or slightly longer than the

sum of the corresponding vdW radii. However, all the calcula-

tions performed in this work predict these complexes to be

stable.

(5) The DFT-SAPT results show that the dispersion and

electrostatic contributions cover about 95% of the total

attraction forces, in these complexes. In the Cl� � �O halogen

bonded complexes, the dispersion is the dominant

attraction component representing about 50% of the total

attraction forces, while the electrostatic forces account for

about 40–45% of the total energy. In the Br� � �O halogen

bonded complex of halothane, the electrostatic and dispersion

terms contribute 51% and 38%, respectively, to the total

attraction forces.

(6) According to the DFT-SAPT results for all complexes

investigated, the repulsive exchange energy overcompensates

the electrostatic one. This supports our conclusion that the

Pauli repulsion (the exchange overlap) effect is responsible for

Table 7 First- and second-order perturbative DFT-SAPT (PBE0/aug-cc-pVTZ) energies (kcal mol�1) for the investigated anaesthetic� � �OCH2

complexes

Eel(1) Eex

(1) E(1) Ei(2) ED

(2) E(2) d(HF) DEDFT�SAPT

Chloroform� � �OCH2 �2.47 2.88 0.41 �0.31 �2.68 �2.99 �0.15 �2.73Halothane� � �OCH2 (A) �4.70 5.24 0.54 �0.65 �3.49 �4.14 �0.32 �3.92Halothane� � �OCH2 (B) �2.85 3.21 0.36 �0.35 �2.90 �3.25 �0.17 �3.06Enflurane� � �OCH2 �2.84 3.10 0.26 �0.34 �3.37 �3.71 �0.15 �3.60Isoflurane� � �OCH2 (A) �2.33 2.81 0.48 �0.29 �2.81 �3.10 �0.14 �2.76Isoflurane� � �OCH2 (B) �2.05 2.74 0.69 �0.29 �2.84 �3.13 �0.14 �2.58Eel

(1) electrostatic energy; Eex(1) exchange–repulsion energy; E(1) sum of Eel

(1) and Eex(1); Ei

(2) induction energy; ED(2) dispersion energy; E(2) sum of

Ei(2) and ED

(2); d(HF) sum of higher order HF contributions; DEDFT–SAPT sum of E(1), E(2) and d(HF) energies.

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

5112 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 This journal is c the Owner Societies 2011

the contraction of the C–X bond in the anaesthetics upon

complexation.

(7) Calculations performed for the chloroform� � �formamide

and chloroform� � �formaldehyde complexes at the CCSD(T)/

CBS level of theory have revealed that the former is more

stable by �0.53 kcal mol�1.

Acknowledgements

The authors are indebted to Wroclaw University of Techno-

logy for financial support, Grant No. 344113. The generous

computer time from the Poznan Supercomputer and

Networking Center as well as Wroclaw Supercomputer and

Networking Center is acknowledged. This work was a part of

the research project No. Z40550506 of the Institute of Organic

Chemistry and Biochemistry, Academy of Sciences of the

Czech Republic and it was also supported by Grants No.

LC512 and MSM6198959216 from the Ministry of Education,

Youth and Sports of the Czech Republic. The support of

Praemium Academiae, Academy of Sciences of the Czech

Republic, awarded to PH in 2007 is also acknowledged.

References

1 C. Sandorfy, Collect. Czech. Chem. Commun., 2005, 70, 539.2 V. Campagna-Slater and D. F. Weaver,Neurosci. Lett., 2007, 418, 28.3 R. G. Eckenhoff and J. S. Johansson, Pharmacol. Rev., 1997, 49,343.

4 G. A. Manderson and J. S. Johansson, Biochemistry, 2002, 41,4080.

5 C. Sandorfy, J. Mol. Struct., 2004, 708, 3.6 T. Cui, V. Bondarenko, D. Ma, C. Canlas, N. Brandon,J. Johansson, Y. Xu and P. Tang, Biophys. J., 2008, 94, 4464.

7 P. Metrangolo, H. Neukirch, T. Pilati and G. Resnati, Acc. Chem.Res., 2005, 38, 386.

8 T. Di Paolo and C. Sandorfy, Nature, 1974, 252, 471.9 G. Trogdon, J. S. Murray, M. C. Concha and P. Politzer, J. Mol.Model., 2007, 13, 313.

10 P. Auffinger, F. A. Hays, E. Westhof and P. S. Ho, Proc. Natl.Acad. Sci. U. S. A., 2004, 101, 16789.

11 R. Battistutta, M. Mazzorana, S. Sarno, Z. Kazimierczuk,G. Zanotti and L. A. Pinna, Chem. Biol., 2005, 12, 1211.

12 M. Ghosh, I. A. Meerts, T. M. A. Cook, A. Bergman, A. Brouwerand L. N. Johnson, Acta Crystallogr., Sect. D: Biol. Crystallogr.,2000, 56, 1085.

13 D. M. Himmel, K. Das, A. D. Clark, S. H. Hughes, A. Benjahad,S. Oumouch, J. Guillemont, S. Coupa, A. Poncelet, I. Csoka,C. Meyer, K. Andries, C. H. Nguyen, D. S. Grierson andE. Arnold, J. Med. Chem., 2005, 48, 7582.

14 Y. Jiang, A. A. Alcaraz, J. M. Chen, H. Kobayashi, Y. J. Lu andJ. P. Snyder, J. Med. Chem., 2006, 49, 1891.

15 M. L. Lopez-Rodriguez, M. Murcia, B. Benhamu, A. Viso,M. Campillo and L. Pardo, J. Med. Chem., 2002, 45, 4806.

16 R. Liu, P. J. Loll and R. G. Eckenhoff, FASEB J., 2005, 19, 567.17 T. Clark, M. Hennemann, J. S. Murray and P. Politzer, J. Mol.

Model., 2007, 13, 291.18 P. Politzer, P. Lane, M. C. Concha, Y. Ma and J. S. Murray,

J. Mol. Model., 2007, 13, 305.19 I. Alkorta, F. Blanco, M. Solimannejad and J. Elguero, J. Phys.

Chem. A, 2008, 112, 10856.20 P. Politzer, J. S. Murray and M. Concha, J. Mol. Model., 2008, 14,

659.21 J. P. M. Lommerse, A. J. Stone, R. Taylor and F. H. Allen, J. Am.

Chem. Soc., 1996, 118, 3108.22 K. E. Riley and K. M. Merz, J. Phys. Chem. A, 2007, 111, 1688.23 K. A. Riley and P. Hobza, J. Chem. Theory Comput., 2008, 4, 232.24 K. A. Riley, J. S. Murray, P. Politzer, M. C. Concha and P. Hobza,

J. Chem. Theory Comput., 2009, 5, 155.25 A. C. Legon, Angew. Chem., Int. Ed., 1999, 38, 2686.

26 E. Corradi, S. V. Meille, M. T. Messina, P. Metrangolo andG. Resnati, Angew. Chem., Int. Ed., 2000, 39, 1782.

27 C. B. Aakeroy, J. Desper, B. A. Helfrich, P. Metrangolo, T. Pilati,G. Resnati and A. Stevenazzi, Chem. Commun., 2007, 4236.

28 P. Metrangolo and G. Resnatil, Science, 2008, 231, 918.29 C. B. Aakeroy, M. Fasulo, N. Schultheiss, J. Desper and

C. Moore, J. Am. Chem. Soc., 2007, 129, 13772.30 A. Sun, J. W. Lauher and N. S. Goroff, Science, 2006, 312, 1030.31 C. Møller and M. S. Plesset, Phys. Rev., 1934, 46, 618.32 R. Krishnan, J. S. Binkley, R. Seeger and J. A. Pople, J. Chem.

Phys., 1980, 72, 650.33 M. J. Frisch, A. J. Pople and J. S. Binkley, J. Chem. Phys., 1984,

80, 3265.34 S. F. Boys and F. Bernardi, Mol. Phys., 1970, 19, 553.35 C. E. Check, T. O. Faust, J. M. Bailey, B. J. Wright, T. M. Gilbert

and L. S. Sunderlin, J. Phys. Chem. A, 2001, 105, 8111.36 R. A. Kendall, T. H. Dunning Jr. and R. J. Harrison, J. Chem.

Phys., 1992, 96, 6796.37 T. H. Dunning Jr., J. Chem. Phys., 1989, 90, 1007.38 K. A. Peterson, D. Figgen, E. Goll, H. Stoll and M. Dolg, J. Chem.

Phys., 2003, 119, 11113.39 K. Raghavachari, G. W. Trucks, J. A. Pople and M. Head-Gordon,

Chem. Phys. Lett., 1989, 157, 479.40 P. Jurecka and P. Hobza, Chem. Phys. Lett., 2002, 365, 89.41 A. Halkier, T. Helgaker, P. Jorgensen, W. Klopper and J. Olsen,

Chem. Phys. Lett., 1999, 302, 437.42 A. Halkier, T. Helgaker, P. Jorgensen, W. Klopper, H. Koch,

J. Olsen and A. K. Wilson, Chem. Phys. Lett., 1998, 286, 243.43 (a) R. Podeszwa, R. Bukowski and K. Szalewicz, J. Chem. Theory

Comput., 2006, 2, 400; (b) A. Hesselmann and G. Jansen, Chem.Phys. Lett., 2002, 362, 319; (c) A. Hesselmann and G. Jansen,Chem. Phys. Lett., 2002, 357, 464; (d) A. Misquitta, R. Podeszwa,B. Jeziorski and K. Szalewicz, J. Chem. Phys., 2005, 123, 214103.

44 J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1996,77, 3865.

45 (a) A. E. Reed, L. A. Curtiss and F. Weinhold, Chem. Rev., 1988,88, 899; (b) E. D. Glendening, A. E. Reed, J. E. Carpenter andF. Weinhold, NBO 3.1 Theoretical Chemistry Institute, Universityof Wisconsin, Madison, WI, 1996.

46 (a) F.A. Bulat and A. Toro-Labbe, ‘‘WFA: A suite of programsto analyse wavefunctions’’, unpublished; (b) F. A. Bulat,A. Toro-Labbe, T. E. Brinck, J. S. Murray and P. Politzer,J. Mol. Model., 2010, 16, 1679.

47 M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone,B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li,H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng,J. L. Sonnenberg, M. Hada, M. Ehara, K. Toyota, R. Fukuda,J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao,H. Nakai, T. Vreven, J. A. Montgomery Jr., J. E. Peralta,F. Ogliaro, M. Bearpark, J. J. Heyd, E. Brothers, K. N. Kudin,V. N. Staroverov, R. Kobayashi, J. Normand, K. Raghavachari,A. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi,N. Rega, J. M. Millam, M. Klene, J. E. Knox, J. B. Cross,V. Bakken, C. Adamo, J. Jaramillo, R. Gomperts, R. E.Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W.Ochterski, R. L. Martin, K. Morokuma, V. G. Zakrzewski, G. A.Voth, P. Salvador, J. J. Dannenberg, S. Dapprich, A. D. Daniels,O. Farkas, J. B. Foresman, J. V. Ortiz, J. Cioslowski andD. J. Fox, Gaussian, Inc., Wallingford, CT, 2009.

48 MOLPRO, a package of ab initio programs designed byH. J. Werner and P. J. Knowles, version 2006, R. D. Amos et al.

49 B. Czarnik-Matusewicz, D. Michalska, C. Sandorfy andTh. Zeegers-Huyskens, Chem. Phys., 2006, 322, 331.

50 Z. Liu, Y. Xu, A. C. Saladino, T. Wymore and P. Tang, J. Phys.Chem. A, 2004, 108, 781.

51 D. Scharf and K. Laasonen, Chem. Phys. Lett., 1996, 258, 276.52 F. J. Melendez and M. A. Palafox, J. Mol. Struct.

(THEOCHEM), 1999, 493, 179.53 W. Zierkiewicz, D. Michalska and Th. Zeegers-Huyskens, J. Mol.

Struct. (THEOCHEM), 2009, 911, 58.54 A. Pfeiffer, H. J. Mack and H. Oberhammer, J. Am. Chem. Soc.,

1998, 120, 6384.55 C. Zhao, P. L. Polavarapu, H. Grosenik and V. Schurig, J. Mol.

Struct., 2000, 550, 105.

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online

This journal is c the Owner Societies 2011 Phys. Chem. Chem. Phys., 2011, 13, 5105–5113 5113

56 D. Michalska, D. Bienko, B. Czarnik-Matusewicz, M. Wierzejewska,C. Sandorfy and Th. Zeegers-Huyskens, J. Phys. Chem. B, 2007, 11,12228.

57 D. Hauchecorne, R. Szostak, W. A. Herrebout and B. J. van derVeken, ChemPhysChem, 2009, 10, 2105.

58 A. Bondi, J. Phys. Chem., 1964, 68, 441.59 Q. Li, X. Xu, T. Liu, B. Jing, W. Li, J. Cheng, B. Gong and J. Sun,

Phys. Chem. Chem. Phys., 2010, 12, 6837.60 M. Palusiak, J. Mol. Struct. (THEOCHEM), 2010, 945, 89.61 B. Michielsen, W. A. Herrebout and B. J. van der Veken,

ChemPhysChem, 2007, 8, 1188.

62 W. Zierkiewicz, P. Jurecka and P. Hobza, ChemPhysChem, 2005,6, 609.

63 W. Wang and P. Hobza, J. Phys. Chem. A, 2008, 112, 4114.64 K. Hermansson, J. Phys. Chem. A, 2002, 106, 4695.65 W. Qian and S. Krimm, J. Phys. Chem. A, 2002, 106, 6628.66 J. S. Murray, M. C. Concha, P. Lane, P. Hobza and P. Politzer,

J. Mol. Model., 2008, 14, 699.67 X. Li, L. Liu and H. B. Schlegel, J. Am. Chem. Soc., 2002, 124,

9639.68 D. Hauchecorne, B. J. van der Veken, A. Moiana and

W. A. Herrebout, Chem. Phys., 2010, 374, 30.

Dow

nloa

ded

by W

rocl

aw U

nive

rsity

of

Tec

hnol

ogy

on 0

7 M

arch

201

1Pu

blis

hed

on 0

3 Fe

brua

ry 2

011

on h

ttp://

pubs

.rsc

.org

| do

i:10.

1039

/C0C

P020

85K

View Online


Top Related