computational study of weakly interacting complexes · computational study of weakly interacting...

191
Computational Study of Weakly Interacting Complexes Dissertation for the Degree of Doktor der Naturwissenschaften (Dr. rer. nat.) Ruhr-Universität Bochum Elsa Sánchez García 2006

Upload: vannguyet

Post on 01-Nov-2018

219 views

Category:

Documents


0 download

TRANSCRIPT

Computational Study of Weakly Interacting Complexes

Dissertation

for the Degree of

Doktor der Naturwissenschaften (Dr. rer. nat.)

Ruhr-Universität Bochum

Elsa Sánchez García

2006

2

This work was carried out between September 2002 and March 2006 under the

supervision of Prof. Dr Luis Montero, Laboratorio de Química Computacional y Teórica,

Universidad de la Habana and Prof. Dr. Wolfram Sander, Lehrstuhl für Organische

Chemie II, Ruhr Universität Bochum.

First Referee: Prof. Dr. M. Havenith-Newen

Second Referee: Prof. Dr. L. A. Montero Cabrera

Subsidiary Subject: Prof. Dr. B. Benecke (Biochemistry)

Dissertation submitted:

Disputation: 19.07.2006

3

4

List of Publications

1. Sánchez-García, Elsa; Montero, Luis. A; Sander, Wolfram. Computational

Study of Non-Covalent Complexes between Formamide and Formic Acid. Journal of Physical Chemistry A (2006), in press.

2. Montero, Luis. A; Sánchez-García, Elsa. Similarity Analysis of Molecular Systems Formed by Amylose and Organoleptic Compounds. Revista Cubana de Física (2006), in press.

3. Sánchez-García, Elsa; Studentkowski, Marc; Montero, Luis A.; Sander,

Wolfram. Non-covalent Complexes between Dimethyl ether and Formic Acid-an

Ab initio and Matrix Isolation Study. ChemPhysChem (2005), 6(4), 618-624.

4. Sánchez-García, Elsa; George, Lisa; Montero, Luis A.; Sander, Wolfram. 1:2

Formic Acid/Acetylene Complexes: Ab initio and Matrix Isolation Studies of

Weakly Interacting Systems. Journal of Physical Chemistry A (2004), 108(52),

11846-11854.

5. George, Lisa; Sánchez-García, Elsa; Sander, Wolfram. Matrix Isolation Infrared

and ab initio Study of Formic Acid-Acetylene Interaction: Example of H…π and

C-H…O Interaction. Journal of Physical Chemistry A (2003), 107(35), 6850-

6858.

6. García, Elsa Sánchez; Montero, Luis A.; Hermida, Jose M.; Cruz, Roberto;

Gonzalez, Gerardo. Calculation of Association Energy in Acetone Clusters by the

Multiple Minima Hypersurface Approximation. Revista Cubana de Física

(2000), 17(1-2), 41-46.

7. Sánchez-García, Elsa; Mardyukov, Arthur; Studentkowski, Marc; Montero,

Luis. A; Sander, Wolfram. Furan - Formic Acid Dimers – an Ab initio and Matrix

Isolation Study, (2006) submitted.

8. Sánchez-García, Elsa; Montero, Luis. A; Sander, Wolfram. Computational Study

of Weakly Interacting Complexes between Acetylene and Oxygen Heterocycles,

(2006) in preparation.

5

Scientific Meetings (Poster = P, Oral Presentation=O)

2006 1st Workshop Forschergruppe 618, Universität Bochum, Germany. (O)

2005 International Chemical Congress of Pacific Basin Societies (Pacifichem).

Area11 - Physical and Theoretical Chemistry. Computational Quantum

Chemistry: Methodology and Application. Honolulu, Hawaii, United

States. (P)

2003 Gordon Research Conference of Physical Organic Chemistry. Plymouth,

New Hampshire, United States. (P)

6

Index

1. General Introduction ..................................................................................................... 9

1.1 The formic acid molecule......................................................................................... 10

1.2 Hydrogen bonds and weak interactions.................................................................... 11

- Some definitions of the hydrogen bond ................................................... 11

- Classification of hydrogen bonds............................................................. 12

- Some applications of the hydrogen bond................................................. 15

- Cooperativity............................................................................................ 17

- Proton transfer.......................................................................................... 18

- Methods of studying hydrogen bonds ...................................................... 19

- Spectroscopy methods.............................................................................. 20

- Diffraction methods ................................................................................. 22

- Matrix isolation ........................................................................................ 24

1.3 Quantum mechanical calculations............................................................................ 26

- The quantum-mechanical treatment of molecules. .................................. 26

- Electron correlation.................................................................................. 28

- Semiempirical methods............................................................................ 30

- Density functional theory......................................................................... 36

- Basis sets .................................................................................................. 38

- Basis set superposition error .................................................................... 41

2. The Multiple Minima Hypersurface (MMH) Approach............................................. 46

2.1 Introduction.............................................................................................................. 46

- The Multiple minima problem ................................................................. 46

- Chemical similarity searching.................................................................. 49

2.2 The Multiple Minima Hypersurface (MMH) approach ........................................... 51

3. Formic Acid Complexes with Formamide and Dimethyl ether.................................. 58

3.1 Introduction.............................................................................................................. 58

3.2 Computational methods............................................................................................ 59

3.3 Formic acid – formamide complexes. Results and discussion................................. 60

- Formic acid – formamide dimers ............................................................. 60

7

- Geometries and binding energies. Analysis of the intermolecular

interactions ............................................................................................... 60

- Comparison with other dimers ................................................................. 67

- Methods and basis set influence on the calculated geometries and binding

energies of the FMA – FA dimers............................................................ 69

- Effect of the BSSE on the calculated geometries and binding energies .. 72

- Intramolecular distances and vibrational frequencies. Calculated spectra73

- Larger systems ......................................................................................... 76

- 1:2 Formic acid – formamide complexes................................................ 76

- Analysis of the intermolecular interactions in the trimers ....................... 79

- 1:4 Formic acid – formamide complexes................................................. 86

- Comparison of FMA – FA complexes with the crystal structure ............ 88

3.4 Formic acid – dimethyl ether dimers. Results and discussion ................................. 91

- Geometries and binding energies ............................................................. 91

- Geometry optimization including BSSE.................................................. 95

- Intramolecular distances and vibrational frequencies .............................. 97

- Comparison with matrix isolation spectroscopy results........................... 99

3.5 Conclusion.............................................................................................................. 101

4. Formic Acid Complexes with π systems .................................................................. 104

4.1 Introduction .......................................................................................................... 104

4.2 Computational methods........................................................................................ 106

4.3 Formic acid – furan dimers. Results and discussion ............................................ 108

- Geometries and binding energies. .......................................................... 108

- Type (i) complexes................................................................................. 111

- Type (ii) complexes................................................................................ 117

- Other FA – furan geometries................................................................. 118

- Basis set influence on the calculated geometries of the FA – furan dimers

................................................................................................................ 122

- Effect of the BSSE on the calculated geometries and binding energies 129

- Comparison with other furan complexes ............................................... 131

- Comparison with matrix isolation spectroscopy results......................... 132

8

4.4 1:2 Formic acid – acetylene complexes. Results and discussion ......................... 136

- Geometries and binding energies ........................................................... 136

- Intramolecular distances and vibrational frequencies ............................ 141

- Comparison with matrix isolation spectroscopy results......................... 145

- Analysis of the intermolecular interactions in the trimers ..................... 148

4.5 Conclusion ............................................................................................................. 152

5. Acetylene Complexes with Oxygen Heterocycles. An Outlook............................... 155

5.1 Introduction............................................................................................................ 155

5.2 Computational methods ......................................................................................... 158

5.3 Acetylene – furan dimers. Results and discussion................................................. 159

5.4 Acetylene – THF dimers. Results and discussion.................................................. 164

5.5 Acetylene – 1,4-dioxane dimers. Results and discussion ...................................... 167

5.6 Conclusion ............................................................................................................. 168

6. General Conclusion................................................................................................... 169

7. Summary ................................................................................................................... 174

8. References................................................................................................................. 181

9

1. General Introduction

Hydrogen bonds and weak interactions play important roles in molecular recognition,

properties of condensed phases, solid state reactions, crystal engineering, and in

determining the shapes and stabilities of biomolecules.[1-3] In contrast to the conventional

strong and moderate hydrogen bonds, which have been extensively described, the nature

and characteristics of weak interactions is not an undisputed field.[3] To find out a

representative and large set of the possible molecular arrangements (minima) of hydrogen

bonded and weakly interacting complexes is quite often one of the most complicated

questions.

Therefore, the structural analysis application of the Multiple Minima Hypersurface

(MMH) approach[4-6] as a tool for localizing minima is introduced here. Randomly

arranged clusters are generated as starting points and subsequently optimized. The results

are processed with programs especially written for this purpose and the geometries are

afterward re-optimized at higher level of theory. The bases of the MMH procedure for

searching local minima are presented and discussed.

Dimers and larger aggregates of formic acid with acetylene, dimethyl ether, formamide

and furan show both strong hydrogen bonds and weak interactions and are studied using

MMH in combination with high level ab initio calculations. The comparison of the

various minima and systems allows for a detailed discussion of the individual

contributions of intermolecular interactions in formic acid complexes. An outlook to the

dimers of acetylene with furan, tetrahydrofuran and 1,4-dioxane complements the

analysis of the intermolecular interactions.

The theoretical results are compared to data from matrix isolation spectroscopy or

crystal structure analysis. The influence of the theoretical methods and the basis set

superposition errors (BSSE) on the calculated geometries and binding energies of the

complexes is also studied. The Multiple Minima Hypersurface (MMH) approach,

combined with ab initio quantum-chemical calculations is established as a very reliable

procedure for localizing weakly and moderate hydrogen bonded minima.

10

1.1 The formic acid molecule

Formic acid (HCOOH) is the smallest monocarboxylic acid and one of the simplest

molecules that forms two hydrogen bonds.[7] Therefore, its structure in the gas and

condensed phases has been much studied.[8-29] The formic acid molecule displays

rotational isomerism[8, 9] between the experimentally well characterized s-trans[10, 11] and

s-cis conformers[9, 12] (Figure 1.1). The s-trans form is 4 kcal/mol lower in energy than

the s-cis form.[7-9] Lundell et al.[13] generated the s-cis conformer by multiphoton IR

irradiation of the s-trans conformer in low-temperature matrices.[7, 13]

Figure 1.1: s-trans and s-cis conformers of the formic acid molecule

In the gas phase the monomer and the dimer of formic acid are forming an equilibrium

in which the dimer is more stable by 14 kcal/mol.[7, 14] Like acetic acid, but unlike many

others carboxylic acids which retain the dimeric structure in the crystalline state, the

crystal structure of formic acid shows an infinite polymeric chain in which each molecule

is linked to two neighbors by a hydrogen bond.[8, 15] At very low temperatures (4.5 K) the

chains of the s-trans form are found in the crystal structure, whereas chains of the s-cis

form are found at higher temperatures.[8, 15] Formic acid is a strongly hydrogen bonded

liquid[16] which probably consists of short chains similar to those observed in the solid.[8]

However, the structure of the liquid formic acid is still a subject of debate, since the

cyclic dimer, an acyclic open dimer[17], polymeric chains [7, 18] and a mixture of several of

these species[7] have been proposed as main constituents. Due to the properties of formic

acid, the hydrogen bonding with other molecules can be used as a model for many

chemical and biochemical systems which exhibit the organic acidic type of bonding, like

proteins and the base pairs in nucleic acids.[19-22]

11

1.2 Hydrogen bonds and weak interactions

Some definitions of the hydrogen bond

The evidences of hydrogen bond were observed long before it was identified and given

a name.[1] Since the beginning of the last century, scientists like Werner (1902), Hantzsch

(1910) and Pfeiffer (1914)[23, 24] used the terms “Nebenvalenz” (near valence) and “innere

Komplexalzbildung” to describe both intra- and intermolecular hydrogen bonds.[1] Moore

and Winmill[25] in 1912 used the term weak union in describing properties of amines in

aqueous solutions.

In “An Introduction to Hydrogen Bonding”, Jeffrey describes that, according to

Pauling, the concept of the hydrogen bond is attributed to M.L. Huggins and

independently to W.M Latimer and W.H Rodebush.[1] In 1922, Huggins affirmed that “a

positively charge kernel containing no electrons in its valence shell reacting with an

atom containing a lone valence pair can form a weak hydrogen bond”.[1] But two years

earlier in 1920, Latimer and Rodebush published that “The hydrogen nucleus held by two

octets constitutes a weak bond”. [1, 26] Both of them mentioned the example of the amines

in aqueous solutions described previously by Moore and Winmill.

It was Pauling who really introduced the concept of the hydrogen bond with the

statements: “Under certain conditions an atom of hydrogen is attracted by rather strong

forces to two atoms instead of only one, so it may be considered to be acting as a bond

between them. This is called a hydrogen bond” and “A hydrogen atom with only one

stable orbital cannot form more than one pure covalent bond and the attraction of the

two atoms observed in hydrogen bond formation must be due largely to ionic forces”.[27]

Therefore, Jeffrey states that hydrogen bonds are formed when the electronegativity,

according to Pauling, of A relative to H in an A-H covalent bond is such as to withdraw

electrons and leave the proton partially unshielded. To interact with this donor A-H bond,

the acceptor B must have lone-pair electrons or polarizable π electrons.[1]

The first text devoted entirely to hydrogen bonding “The Hydrogen Bond” was written

by Pimentel and McClellan in 1960.[28] They give a more general definition of hydrogen

bond: “A hydrogen bond exists between the functional group, A-H, and an atom or a

12

group of atoms B, in the same or different molecules when (a) there is evidence of bond

formation (association or chelation), (b) there is evidence that this new bond linking A-H

and B specifically involves a hydrogen atom already bonded to A”.[28] It is important to

realize that the Pimentel and McClellan definition makes no assumptions about the nature

of the A and B atoms, and that it enables an evaluation of the hydrogen bonding potential

of groups like C-H and π acceptors.[29]

As Jeffrey points out, a lot has been written about hydrogen bonds, and some concepts

are been continuously rediscovered.[1] Thus, the C-H hydrogen bonds are currently a

point of interest of the scientific community, but they were reviewed more than 50 years

ago by Hunter.[30] However, nowadays the definition of hydrogen bond by Pimentel and

McClellan is the most accepted due to its practical applications and suitability for both

experimental and theoretical investigators.[1]

Classification of hydrogen bonds

The hydrogen bonds are classified by Jeffrey in three categories according to their

energies and the nature of the A-H…B interactions.[1] The strong hydrogen bonds have

bond energies between 15 – 40 kcal/mol and the A-H…B interaction is mostly covalent

with H…B bond lengths from 1.2 to 1.5 Å and bond angles of 175 – 180°. They have an

electron density deficient donor group or an acceptor group with an excess of electron

density.[1]

Moderate hydrogen bonds are those which have bond energies between 4 – 15 kcal/mol

and the A-H…B interactions are mostly electrostatics with H-B bond lengths of 1.5 – 2.2

Å and bond angles of 130 – 180°.[1] They are mostly formed by neutral donor and

acceptor groups in which the donor A atoms are more electronegative than the hydrogen

and the acceptor B atoms have lone-pair unshared electrons. According to Jeffrey, these

are the most common hydrogen bonds in chemistry and nature and essential components

of the structure and function of biological molecules.[1]

Weak hydrogen bonds have bond energies between 1 – 4 kcal/mol and the A-H…B

interactions are basically electrostatic although probably also involving electron

correlation, with H-B bond lengths of 2.2 – 3.2 Å and bond angles of 90 – 150°. They are

13

formed when the hydrogen atom is covalently bonded to a slightly more electroneutral

atom relative to hydrogen, as in C-H or when the acceptor group has π electrons, such as

C≡C or an aromatic ring. These interactions have similar energies and geometries than

van der Waals complexes, however, differ from the latter by a directional involvement of

the A-H bond.[1]

As can be seen in Figure 1.2,[31] Desiraju classifies the hydrogen bonds in a very

similar way to Jeffrey, but he names the strong hydrogen bonds “very strong”, and the

moderate “strong”. This distinction comes from supramolecular considerations, since

Desiraju means by “strong” bonds those that are able to control crystal and

supramolecular structure effectively. By weak, Desiraju and Steiner mean hydrogen

bonds whose influence on crystal structure and packing is variable.[29] In this sense, a

“strong” hydrogen bond is one which is much stronger than a van der Waals interaction

while a weak hydrogen bond is one which is not. In order to be consistent, the

classification made by Jeffrey is used here and Desiraju’s classification is referred in

“italics” if necessary. Thus, according to Desiraju`s classification, the O-H…O=C and N-

H…O=C interactions are “strong” (moderate for Jeffrey) hydrogen bonds. The C-H…O,

C-H…N, N-H…π, O-H…π and C-H…π interactions are weak hydrogen bonds.

Desiraju and Steiner also classify the hydrogen bonds as “conventional” and “non-

conventional”,[29] based on the “conventionality” of the donor and acceptor groups,

where the categories “strong” and “conventional” have many points in common.

However, there are “strong” non-conventional hydrogen bonds as well as there are weak

conventional hydrogen bond types. The classification of hydrogen bonds still is a

controversial field. For instance, methyl donors are borderline cases between weak

hydrogen bonds and van der Waals interactions because of their large dispersion

contribution.

14

Figure 1.2: Desiraju’s classification of hydrogen bonds. This figure has been taken

from “Hydrogen Bridges in Crystal Engineering: Interactions without Borders” by G. R.

Desiraju[31]

The C-H…π interaction is another borderline case, since it is considered by some

authors like Nishio[3] as the weakest hydrogen bond occurring between a soft acid (CH)

and a soft base (π electrons), whereas others will not name it as a hydrogen bond at all.

Nevertheless, it has gradually become accepted that the C-H…π interaction plays a role

in a variety of chemical and biochemical phenomena like the stabilization of proteins

structures, the conformation of coordination compounds and the selectivity in organic

reactions.

Desiraju and Steiner point out some differences between “strong” and weak hydrogen

bonds:[29]

• The van der Waals cut-off criterion in the H…B distance for the assignment of

hydrogen bond character is inappropriate for weak hydrogen bonds. This

15

criterion does not stand on experimental or theoretical ground, but has only

been established for reasons of apparent convenience. This criterion works

reasonably well for “strong” hydrogen bonds which are almost always short

enough to fulfill it. But even for these, due to sterical reasons, bonds like N-

H…O can be elongated beyond the van der Waals separation. Weak hydrogen

bonds, especially A-H…π interactions are even longer.[29]

• The results of crystallographic and spectroscopic investigations do not

necessarily agree to each other as well as they do for “strong” hydrogen bonds.

Unlike the “strong” hydrogen bonds, large distortions are possible with very

little changes of the energies in weakly bonded systems, due to their shallow

potential energy surfaces. Consequently, the correlation between

crystallographic and spectroscopic properties is very variable.[29]

• The hydrogen bond in general is considered as the initial state of a proton

transfer process, but only for “strong” hydrogen bonds do such proton transfer

processes occur with significant rates.[29]

Some applications of the hydrogen bond

As Nishio explains, noncovalent forces play an important role in chemical reactions,

molecular recognition, and in many biochemical and chemical processes. While strong

covalent bonds bind the atoms together in a molecule, the noncovalent and weak

interactions determine the shape and the conformation of the molecule.[3]

Therefore, hydrogen bonding is very relevant to supramolecular chemistry. For

instance, the role of hydrogen bonding in determining the packing motifs of molecules in

crystals requires the recognition and understanding of the cooperative systems of

hydrogen bonding. As with molecules, any description of supramolecular structure

requires the knowledge of connectivity and geometry. For Jeffrey the connectivity is the

hydrogen-bonding pattern.[1] A knowledge of commonly occurring hydrogen-bond

patterns associated with particular donor and acceptor function groups can be used to

synthesize new supramolecular complexes.[1, 32]

16

Intermolecular hydrogen bonding also plays an important role in the way molecules

assemble in liquid crystals.[1] According to the name, liquid crystals are supramolecular

assemblies in one and two dimensions which constitute a state of matter between crystals

and liquids. They are also known as ordered liquids.[1] Ferroelectric and others liquid

crystals have been made through hydrogen bonding, and in some liquid crystals the

function of the hydrogen bonding is to increase the length of the rods.[1]

Molecular inclusion is another large and growing field of supramolecular chemistry. In

inclusion compounds like the hydrates and the cyclodextrines, hydrogen bonding is an

essential component of the host structure.[1]

The hydrogen-bonded helical and sheet structures proposed for proteins by Pauling,

Corey and Branson and the hydrogen bonded base-pair in the structure of DNA by

Watson and Crick are evidences of the importance of hydrogen bonding in the structure

and function of biological macromolecules.[1] Jeffrey explains that strong hydrogen bonds

are rare in biological structures since they are too rigid and not easily broken. For

example, the salt bridges in proteins and the P-OH…O=P bonds in nucleic acids are

hydrogen bonds. These bonds are generally interrupted by water molecules, which do not

form very strong hydrogen bonds, either as donors or acceptors. The weak interactions

like the C-H…O hydrogen bonds play also a role in biological structures.[1]

Hydrogen bonding is the major factor in determining the structure of the nucleic acids.

Inter- and intramolecular hydrogen bonding schemes have been also proposed for

polysaccharides. In protein structures, the peptide N-H and C=O groups form

intramolecular N-H…O=C hydrogen bonds which determine the conformation of the

peptide main chain, being responsible for the formation of helical or sheet structures. The

formation of these hydrogen bonds results in additional π character of the peptide C-N

bond, which results in a more rigid planar conformation. The side groups contain

hydrogen bond donor and acceptor groups which form the hydrogen bonds between the

polypeptide chains. [1]

17

Cooperativity

Hydrogen bonds show cooperative effects. Accordingly, the energy of an array of n

interlinked hydrogen bonds is larger than the sum of n-isolated hydrogen bonds, as

described by Desiraju.[29] Or, according to Jeffrey’s definition, cooperativity or non-

additivity represents the difference between calculating energies using atom-pair

potentials and many-atom potentials.[1] This non-additive property can be applied in

general to all intermolecular interactions. Cooperativity occurs because of the ability of

donor and acceptor groups to form hydrogen bonds is further increased by an increase in

polarity when the hydrogen bonds are part of a collective ensemble. Two different

mechanisms to producing this effect are described:[29]

Functional groups acting simultaneously as hydrogen bond donors and acceptors form

extended chains or rings in which the individual hydrogen bonds enhance each other’s

strength by mutual polarization.[29] This occurs mainly with hydroxyl groups and was

recognized in the crystal structure of small carbohydrates by Jeffrey.[1] From the crystal

structure of the cyclodextrins this effect was identified by Saenger.[33, 34] Since there are

no multiple bonds involved, it has been described by Jeffrey and Saenger as σ-bond

cooperativity.[1]

Charge flow in suitably polarizable π-bond systems increases donor and acceptor

strengths. This cooperativity involves hydrogen bonding between molecules with

conjugated multiple π systems and is also described as Resonance-Assisted Hydrogen

Bonding (RAHB).[1, 29, 35] This descriptor was first applied to hydrogen bonding in β-

diketone moieties and has since been extended and made more general by Gilli, Bertolasi

and Ferreti.[35] In some biological structures it is called π-cooperativity by Jeffrey and

Saenger.[1]

Cooperativity is particularly important in hydrogen bonding because of the diffuse

nature and high polarizability of the hydrogen and lone-pair electron densities; and

according to Jeffrey,[1] the most evident structural manifestation of σ-bond cooperativity

is the predominance of linear chains of …O-H…O-H…O-H… bonds in the crystal

structures of the monosaccharides and the cyclic hydrogen bond structures of

18

oligosaccharides and cyclodextrins. In carbohydrate hydrates, the water molecules use

their double-donor double-acceptor properties to link the chains of hydroxyl bonds into

three-dimensional nets. [1]

As Jeffrey explains, RAHB is important in many biological structures and has been

observed in the crystal structures of purines, pyrimidines and their complexes. In these

crystal structures, the hydrogen bonding extends beyond the base pairs to other

molecules, and infinite hydrogen bond – π bond networks link the molecules throughout

the crystal structures. In the nucleic acids, base-pairing between purines and pyrimidines

involves the hydrogen bonds which link two conjugated ring systems. Thus, RAHB plays

an important role in increasing the delocalization energy of the molecules involved and in

strengthening the hydrogen bonding. In proteins, the main chain is not an extended

conjugated system, since two peptide units are separated by a single C-C bond. However,

there is RAHB in the pleated-sheet hydrogen bond structures running laterally across the

main chains.[1]

Jeffrey describes that the first evidence of the Polarization Enhanced Hydrogen

Bonding came from the ab initio calculations of del Bene and Pople on cyclic and chain

water polymers.[36] The calculations of the cyclic sequential (H2O)n shows a increase of

hydrogen bond energy per bond from 5.6 kcal/mol for n = 3 to 10.6 kcal/mol for n = 5

and 10.8 kcal/mol to n = 6, simultaneously there was a corresponding shortening of the

calculated H…O bond lengths from 1.57 to 1.45 Å. An example of polarization via a

combination of σ and π bonds are the (HCN)n chains. Desiraju also explains that the

ethynyl group is of particular relevance to the phenomenon of cooperativity because it

can simultaneously form C-H…B and A-H…π hydrogen bonds.[29] Another case of

cooperativity is the steroid danazole where the hydrogen bonds form a cooperative

pattern with infinite chains of C-H…O and O-H…π interactions.[29]

Proton transfer

Hydrogen bonding can ultimately lead to proton transfer, but it is important to stress

the differences between proton transfer and hydrogen bonding.[1] The pyridine –

hydrogen fluoride complexes are one example for the transition of hydrogen bonding to

19

proton transfer. For the 1:1 pyridine – HF complex there is an F-H…N hydrogen bond

and no proton transfer although the F-H distance is long, 1.13 Å, and the H...N distance is

short, 1.32 Å. In the 1:2 and 1:3 complexes, there is proton transfer forming the

pyridinium cation.[1]

The fact that hydrogen bonding facilitates, or restricts, proton transfer is considered as

the most important chemical property of the hydrogen bond by authors like Jeffrey.[1] The

facility of hydrogen bonds to transmit H+(or H3O+) and OH- ions in water or an aqueous

media provides a catalysis mechanism for many reactions. In the field of molecular

biology proton transfer has been recognized as a significant component of enzyme

catalysis and the transmission of ions through membranes.[1]

Methods of studying hydrogen bonds

The following methods are frequently used to study hydrogen bonded systems,

according to Jeffrey:[1]

• Spectroscopy methods: They depend on exciting the vibrational or rotational

energy levels of molecules, resulting in the absorption, or emission of the incident

radiation at specific frequencies. This radiation can be electromagnetic or

neutrons. Spectroscopy methods include infrared and Raman, microwave and

NMR, among others. They provide information relating to structure and processes

on a picosecond time scale (10-10 – 10-15 sec). NMR spectroscopy provides

information at 10 – 10-4 sec.

• Diffraction methods: They depend on the three-dimensional periodicity of the

atoms in crystals to provide a diffraction grating for X-rays or neutrons of wave

lengths comparable to the interatomic distances. Diffraction by liquids gives

much less information, even when X-ray and neutron diffraction results are

combined for simple liquid such as water. As already mentioned, diffraction

methods include X-ray and neutron diffraction. They provide information on a 10

– 103 sec time scale.

20

• Thermochemical methods: Thermodynamic methods involve either direct

calorimetry or using the effect of hydrogen bonding on physical properties at

different concentrations or temperatures to determine the equilibrium constants

for the formation of the hydrogen bond. Thermochemical methods include

calorimetry of heats of mixing or dilution and the determination of enthalpies

directly or through the measurements of equilibrium constants. Like diffraction

methods, thermodynamic methods provide information on a 10 – 103 sec time

scale.

• Theoretical methods: They include ab-initio, density functional, semi-empirical,

and empirical methods.

Here we are presenting some aspects related to spectroscopic and diffraction methods,

according to Jeffrey’s point of view.[1]

Spectroscopy methods

Spectroscopic methods are more general and sensitive than diffraction methods. They

are used to identify hydrogen bonding in all states of matter. For example, the weak C-

H…O hydrogen bonds were identified by spectroscopists long before they were

recognized by the crystallographers.[1]

Infrared spectroscopy is the most used tool for identifying hydrogen bonding. Near

infrared spectroscopy uses the electromagnetic frequency ranges of 10 000 – 4000 cm-1,

middle, 4000–200 cm-1, and far 200–10 cm-1. Raman spectra are recorded in the range of

4000 – 10 cm-1. Most infrared studies of hydrogen bonding are in the mid IR range. With

these methods the hydrogen bonding is investigated by observing the transitions between

the vibrational levels of the molecules involved in hydrogen bonding.

Jeffrey points out some general IR criteria for hydrogen bonding:[1]

• The A – H stretching frequency νs, is shifted to lower frequencies (red shift).

This is accompanied with an increase in intensity and band width compared to

the isolated monomers.

21

• The A – H bending frequencies νb, move to higher frequencies.

• Upon cooling, νs shifts to high frequencies with increase in intensity and decrease

in band width, νb moves to lower frequencies with decrease in band width.

• Isotopic substitution of H by D lowers νs frequencies by a factor of around 0.75.[1]

With the introduction of interferometers in place of dispersive elements in IR

spectroscopy it is possible to record data of all frequencies at the same time. This method

is known as Fourier transform infrared spectroscopy and delivers in addition more

radiation with greater stability.[37]

The correlations between stretching frequencies and hydrogen-bond geometries have

been studied. A linear relationship has been found between νA-H and the A----B bond

distances for the strong O-H…O hydrogen bonds with νs 2700 cm-1 to 750 cm-1 and O----

O from 2.60 to 2.45 Å. For weaker bonds with O----O > 2.6 Å, the relationship is getting

curved, the agreement deteriorated, and the frequency shifts become increasingly

insensitive to changes in O----O distances.

The microwave rotational spectroscopy uses electromagnetic radiation in the frequency

region 109 – 1011 Hz to record the vibrational and rotational spectra of hydrogen-bonded

dimers and 1:1 adducts in gas phase. The analysis of these spectra provides information

about rotational constants, centrifugal distortion constants, nuclear quadrupole and

nuclear spin-nuclear spin coupling constants, and the Stark and Zeeman effects.

Molecular geometries, bond energies, force constants, electric dipole moments, electric

charges distributions, and electric quadrupole moments are derived from these

measurements. This method is sensitive enough to give information about very weak

hydrogen bonds and provides hydrogen bonding information that it is not compromised

by solvent effects or crystal field effects.[1]

There are two experimental methods in gas-phase microwave rotational spectroscopy:

One uses Stark modulated microwave spectroscopy with binary gas mixtures at

temperatures ≥ 175ºC and pressures of 50 mTorr and the second uses Fourier transform

microwave spectroscopy of a pulse of gas mixture diluted in argon and expanded

supersonically into an evacuated Fabry-Perot cavity.[1] Since gas-phase microwave

22

rotational spectroscopy measures the distances between centers of mass of the donor and

acceptors molecules, and hydrogen atoms make only a very small contribution to the

molecular mass, there are ambiguities in the measurements of geometries for weakly

bonded dimers where the A-H…B interaction is not linear.[1]

NMR spectroscopy is another very sensitive method for identifying hydrogen bonding.

It is less widely applied than infrared spectroscopy, because of the complexity of

hydrogen bonding in solution due to the uncertainty in identifying the particular bonds

and the number of molecules involved. NMR spectroscopy measures the degree to which

the proton is shielded by its electronic environment in terms of chemical shifts. These

shifts provide evidence of hydrogen bonding in liquids and solution and their magnitude

is quantitatively proportional to the strength of the hydrogen bond. As with infrared

spectroscopy, the change in chemical shift with concentration or temperature can give

equilibrium constants and therefore thermodynamic data. The sensitivity of 1H NMR to

changes in the electronic environment makes it a useful probe for detecting hydrogen

bonding from weak donors, such as C-H, and weak acceptors, such as multiple bonds and

aromatic rings.[1] With the development of multi-dimensional methods, NMR

spectroscopy has become a powerful tool for elucidating molecular structure in solution.

However, solution NMR spectroscopy has only relative little impact on the study of

hydrogen bonds, because of the complexity of the liquid state. The results of the solid-

state NMR spectroscopy can be correlated with those of crystal structure analysis and

therefore, solid-state NMR spectroscopy has become a tool for studying hydrogen

bonding.[1]

Diffraction methods

Location of the hydrogen atoms is essential to understand the nature of the hydrogen

bond, and crystal structure analysis by means of neutron diffraction is the most definitive

method for locating hydrogen atoms in hydrogen bonds. Together with infrared

spectroscopy, neutron diffraction provides a basis for distinguishing between strong,

moderate, and weak bonds. It also gives information to differentiate between two-, three-,

and four-center bonds. In strong hydrogen bonds the A-H…B bonds are almost collinear

and the covalent A-H bond length elongates to become almost equal to that of the

23

hydrogen bond. In moderate and weak hydrogen bonds the extension of the covalent A-H

bond is small and is marginally observable, but the A-H…B angles may deviate

significantly from 180º.[1]

Some advantages and disadvantages of X-ray vs. neutron diffraction single crystal

analysis are presented by Jeffrey:[1]

• An important aspect to consider is the availability of each method: While X-rays

are available on demand from laboratory instruments; neutron diffraction

equipments are only available in national or international specialized centers.

• For X-ray diffraction the time required for collecting the data is one day or less

for routine work and small crystals (~0.01 mm3, ~0.01 mg) can be used. Neutron

diffraction requires large crystals (~1 mm3, 1 – 2 mg) and a few weeks of data

collection time.

• With the X-ray diffraction method the hydrogen atoms are poorly located,

especially O-H with an accuracy of approximately 0.1 Å. Neutron diffraction

hydrogen positional parameters are comparable in accuracy to C, N, and O

(~0.001 Å)

• In X-ray diffraction experiments it can be difficult to distinguish between thermal

motion and disorder even for nonhydrogen atoms because of a fall-off in intensity

with scattering angle. In neutron diffraction fall-off in intensity with scattering

angle is only due to its thermal motion. Therefore it is easier to distinguish from

disorder.

• For X-ray diffraction, careful absorption corrections are necessary for other than

first-row atoms. In neutron diffraction the absorption is negligible, except for

crystals containing B, Cd, Sm, Li.[1]

One important advantage of both X-ray and neutron diffraction crystal structure analyses

over every other method of structure analysis is that both methods are over-determined.

24

Except for macromolecules such as proteins, the number of observations exceeds the

number of variable parameters, generally by a factor between five and ten.[1]

Matrix isolation

The matrix isolation technique was first introduced in 1954 by Pimentel and co-

workers,[38] who used the technique for systematic studies of free radicals and other

unstable or transient species. Matrix isolation was developed independently by Norman

and Porter.[39] The matrix isolation technique is used for trapping and producing chemical

species and preserving them in solidified inert (or occasionally reactive) gases at low

temperatures between 10 – 40 K. The matrices are formed, most of the time, by a non-

reactive substance like rare gases or solid nitrogen. The low temperatures required are

achieved by cryostats with closed helium-cycles.

Since solidified inert gases are used as matrix, interactions between the reaction

medium and the molecules to be studied are weak. To avoid reactions between isolated

molecules, the samples are highly diluted (1000 : 1) during the preparation of the matrix;

the molecules are thus spatially separated while embedded into the matrix lattice. In most

cases, rearrangements of trapped molecules are ruled out at 10 K by sufficiently high

energy barriers.[40]

For preparing the matrices, an excess of inert gas is condensed simultaneously with the

substance to be examined or a suitable precursor onto a cooled spectroscopic window,

usually CsI for IR spectroscopy and quartz or sapphire windows for UV/Vis

spectroscopy. Solids and liquids should have a sufficient vapor pressure (about 10-6

mbar) at temperatures which will not lead to decomposition. Gases can be mixed with

argon in an appropriate ratio before the deposition is performed.[40]

By codeposition of more than one substance, controlled reactions under matrix

conditions can be achieved. The matrix should have a temperature that is about 30 % of

the melting point of the noble gas (e. g. 30 K for Argon). Under these conditions, smaller

molecules like ozone, carbon monoxide or oxygen are able to diffuse through lattice gaps

to encounter a reaction partner. Another kind of reactions of matrix-isolated molecules

25

are photochemically induced processes by irradiation at a suitable wavelength using

mercury high pressure lamps or lasers.[40]

In order to characterize matrix isolated species, infrared, UV/VIS and EPR

spectroscopy are frequently used. Matrix isolation experiments allow, among other

applications to study unstable molecules generated by photolysis or gas phase

thermolysis, to observe directly reaction intermediates, to generate and to study novel

reactive species, to determine the structures of reactive species and to freeze out and to

study particular molecular conformations. An example for the latter is cyclohexane in its

chair and twist conformations.[41]

Another important application of the matrix isolation technique is the study of weakly

bound systems like H-bonded, charge transfer, and van der Waals complexes that can be

isolated in low temperature matrices, despite they dissociate under normal temperature

conditions due to the weak intermolecular forces. The IR bands of the components of

these complexes are significantly perturbed which provides an insight into the

intermolecular interactions. Therefore, matrix isolation, combined with spectroscopic

methods such as infrared spectroscopy, is a very important tool for the study of hydrogen

bonding.[40, 42]

26

1.3 Quantum mechanical calculations

Computational chemistry has become an important method for understanding hydrogen

bonding. In addition to the global minimum, computational methods can locate secondary

minima and stationary points of higher order. It is also possible to study the

interconversion pathways from one minimum to another and the magnitudes and shapes

of energy barriers along these paths. In addition, quantum chemical methods greatly

improve the understanding of the perturbations in vibrational spectra that accompany the

formation of a hydrogen bond. A frequent problem of experimental studies of hydrogen

bonded complexes is separating the intrinsic properties of the complex from the

perturbations due to interactions with the solvent. In this respect, an advantage of

computational methods is that they are free of complicating solvent effects.[43]

The quantum-mechanical treatment of molecules.

Quantum chemical methods are based on the time-independent Schrödinger equation:

H r R E r RΨ Ψ( , ) ( , )= (2.1)

where Ψ(r,R) is the wave function that represents the “trajectories” of the particles and

should be single-valued, quadratically integrable and continuous. H is the Hamiltonian

operator that returns the system energy, E, as an eigenvalue. For a molecule with n-

electrons and N nuclei the Hamiltonian operator is:

H(r,R)= Tel+ T nucl + V nucl,el + V el,el+ V nucl,nucl (2.2)

where:

Tel operator of the kinetic energy of the electrons.

Tnucl operator of the kinetic energy of the nuclei.

Vnucl,el attraction potential between the electrons and the nuclei.

Vel,el repulsion potential between the electrons.

Vnucl,nucl repulsion potential between the nuclei.

r set of coordinates of the n electrons.

27

R set of coordinates of the N nuclei

The Born-Openheimer approximation is used to simplify the solution of the

Schrödinger equation. Under typical physical conditions, the nuclei of molecular systems

are moving much more slowly than the electrons since the mass of a typical nucleus is

thousands of times greater than that of an electron. Consequently, according to the Born-

Openheimer approximation, the electronic energies are computed for fixed nuclear

positions. Therefore, the nuclear kinetic energy term is taken to be independent of the

electrons, correlation in the attractive electron-nuclear potential energy term is

eliminated, and the repulsive nuclear-nuclear potential energy term becomes a simply

evaluated constant for a given geometry.[44]

Due to the electron-electron repulsion term in the Hamiltonian, an exact solution to the

Schrödinger equation is not possible for systems with more that one electron. However, a

number of simplifying assumptions and procedures do make an approximate solution

possible for a large range of molecules. As Scheiner explains in “Hydrogen bonding: A

Theoretical Perspective”,[43] the usual method is the Hartree-Fock (HF) approximation

where electron 1 is considered to move in the field of the electron cloud associated with

the probability distribution of all other electrons. The same idea is applied to electron 2

which moves in the time-averaged field of electron 1 plus all the others, and so on.

Solution of the 1-electron Hartree-Fock equation for each electron changes its probability

density, thereby altering the field it sets up for the other electrons. Consequently, the

equations are solved iteratively, until the 1-particle wave function and the fields

generated there from no longer change appreciably from one cycle to the next. Because

of this, sometimes the SCF abbreviation of self consistent field is used synonymously

with HF.[43]

The Hartree-Fock approximation neglects the electron correlation. Since the electrons

are constantly aware of each others’ presence via their electrostatic repulsion, they tend to

correlated their motions to avoid one another. The electron correlation lowers the energy

of the system and affects the overall electron density of the system.[43]

28

Electron correlation

There are different approaches to the electron correlation problem. The conceptually

simplest is configuration interaction (CI)[45] which takes the Hartree-Fock solution as a

starting point, or reference configuration.[43] Other configurations are generated by

permitting the excitation of one electron from the subset of occupied molecular orbitals to

the subset of unoccupied or “virtual” MOs. The complete list of single excited

configurations is generated by considering all possible excitations with the same spin

state as the ground state under study. The list is then extended to double excitations,

accounting for all possible combinations of excitations of two electrons from the

occupied to the virtual MOs. A full-CI list is generated by progressing to include triple,

quadruple, and higher excitations, until all n electrons have been excited. The correlated

wave function is then expressed as a linear combination of the reference, Hartree-Fock

configuration, plus small amounts of all the possible excitations. Variational treatment of

this trial wave function leads to the correlation energy by adjusting the relative amount

that each particular configuration contributes to the final correlated wave function.[43]

One problem is that even for small systems, the number of configurations generated by

all possible excitations is out the reach of any computer. For this reason, one of the

common points of termination of the list is after the inclusion of all single and double

excitations (CISD). One problem with termination of the full CI expansion is the size-

consistency problem.[43] This means that the same treatment of a complex is

fundamentally different than that of the subunits of which it is composed. For instance,

for a dimer, the CISD treatment would permit double excitations for each monomer but,

instead of permitting quadruple excitations within the dimer, taking into account

simultaneous double excitations of each of the monomers, CISD terminates the excitation

list at doubles in the complex. Therefore, truncated CI treatments handle poorly with

molecular interactions like hydrogen bonds. Another means of introducing size

consistency is by quadratic approximation, QCISD.[46]The approach achieves this size

consistency by giving up its variational character.[43] It must be pointed out that single

excitations only improve the quality of the total wave function, mostly regarding the so

29

called “empty” Hartree-Fock levels, but, according to the Brillouin theorem, do not affect

the total energy of the system.

Other procedures, like the Coupled pair theories[47, 48] are size consistent but are not

variational. This means that in principle, it is possible to obtain a value of energy lower

than the true energy of the system. In the independent electron-pair approximation

(IEPA), the total correlation energy is partitioned into a sum of contributions from each

occupied pair of spin orbitals. A different correlation wave function is constructed for

each pair, letting their electrons be excited into the virtual MOs of the reference

configuration. The total correlation energy then corresponds to the sum of all pair

energies.

When the IEPA approach is extended to incorporate coupling between different pairs,

becomes a coupled-pair theory. In terms of excitations from the original Hartree-Fock

determinant, the correlation energy depends directly upon the double excitations, but

their contributions involve quadruple excitations in an indirect way, and the latter are

linked to hextuple excitations, and so on. The coupled-cluster approximation expresses

this relationship in a closed set of equations. [49] When the applications of coupled-cluster

theory include only double excitations it is identified as CCD.[50] More general versions

of the theory that include also single and higher excitations are abbreviated as CCSD.[51]

Various approximations have been suggested to coupled-cluster since it is highly

demanding of computer resources. One is the linear coupled-cluster approximation (L-

CCA) which sets certain products equal to zero, and is equivalent to doubly-excited many

body perturbation theory. If instead of ignoring all the product terms set equal to zero in

L-CCA, some of them are retained, ones arrives to the coupled electron pair

approximation (CEPA).[43]

The correlation method mostly used to calculate hydrogen bonded systems is the

Møller−Plesset perturbation theory.[52, 53] This approach considers the true Hamiltonian

as a sum of its Hartree-Fock part plus an operator corresponding to electron correlation.

In other words, the unperturbed Hamiltonian consists of the interaction of the electrons

with the nuclei, plus their kinetic energy, to which is added the Hartree-Fock potential:

the interaction of each electron with the “time-averaged” field generated by the others.

30

The perturbation or “correction” operator therefore becomes the difference between the

expected exact interelectronic repulsion operator, with its instantaneous correlation

between electrons, and the latter Hartree-Fock potential.

The first correction to the Hartree-Fock energy appears as the second-order

perturbation energy. The energy including this correction is known as MP2. The MP3

level involves additional terms, but remains restricted to double substitutions from the

reference configuration. At fourth order, there are contributions from single, triple, and

quadruple excitations, as well as doubles. One strong advantage of the MP theory is that,

in addition to its computational efficiency, it is size consistent, so it is a good choice for

different types of molecular interactions. According to Schneiner,[43] several calculations

indicate that MP2 provides results in excellent agreement with the much more

computationally demanding MP4. Therefore, the literature of correlated calculations of

hydrogen bonds is dominated by Møller−Plesset theory.[43]

In certain cases, a single determinant does not offer an adequate representation of the

electronic structure. In such cases, it is useful to perform a Multi-Configurational SCF

calculation (MCSCF) in which a number of different electron configurations among the

Hartree – Fock orbitals are chosen as important and their adjustable parameters like

orbital coefficients are variationally optimized.[54] This procedure is arbitrary in the

choice of which configurations are considered being important. The calculation can be

more objective by including all excitations between a subset of occupied MOs and a

subset of vacant orbitals. These excitations have some restrictions like multiplicity or

order of excitation. The orbitals selected for the excitations are the active space and the

method is called Complete Active Space Self Consistent Field (CASSCF).[55]

Semiempirical methods

Using semiempirical methods allows to calculate large systems when the ab initio

treatment is too demanding computationally. Semiempirical approaches are developed

under the same formalism than ab initio methods, but the semiempirical methods neglect

many smaller integrals.[56] To compensate for these approximations, empirical

parameters are introduced into the remaining integrals and their values are assigned on

the basis of calculations or experimental data.[56] According to Jensen, the various

31

semiempirical methods are defined by how many integrals are neglected, and how the

parameterization is done.[56]

In “Semiempirical Methods”,[57] Thiel points out that, among other applications,

semiempirical methods are useful as a previous approach to a computational problem

before proceeding with higher-level of theory. Compared with ab initio or density

functional methods, semiempirical calculations are much faster and therefore can be used

for calculating larger systems. But semiempirical methods have the disadvantage of being

less accurate and the errors are less systematic.[57]

According to Thiel, the quantum-chemical semiempirical treatments can be defined

depending on:[57]

• The basic theoretical approach: Most semiempirical methods are based on MO

theory and use a minimal basis set for the valence electrons. Electron

correlation is treated explicitly only when necessary for an appropriate zero-

order description.

• The integral approximation and the types of interactions included: According to

that, there are three levels of integral approximation:[57] CNDO (complete

neglect of differential overlap), INDO (intermediate neglect of differential

overlap), and NDDO (neglect of diatomic differential overlap). Unlike CNDO

and INDO which truncate after the monopole, NDDO keeps the higher

multipoles of charge distributions in the two-center interactions.[57]

• The evaluation of integrals: The integrals can be estimated directly from

experimental data, calculated from analytical formulas or from appropriate

parametric expressions. One-center integrals can be calculated from atomic

spectroscopic data. The selection between analytical formulas or parametric

expressions depends mostly on the consideration of how to model the

interactions.

• The parameterization: the semiempirical MO methods are parameterized to

reproduce experimental reference data (or, possibly, accurate high-level

theoretical predictions as substitutes for experimental data). The reference

32

properties are chosen to be representative for the intended applications. The

quality of semiempirical results depends very much of the parameterization.[57]

As Thiel states, the most popular semiempirical methods for studying ground-state

potential surfaces are based on the MNDO model. MNDO is a valence-electron self-

consistent-field (SCF) MO treatment which uses a minimal basis of atomic orbitals and

the NDDO integral approximation.[57] The total energy of a molecule is the sum of its

electronic energy and the core-core repulsion energies.

The MNDO model includes only one-center and two-center terms, so it is

computationally more efficient. The one-center terms are taken from atomic

spectroscopic data, and slight adjustments are permitted in the optimization to take into

account the differences between free atoms and atoms in a molecule.[57]

In “Hydrogen Bonding by Semiempirical MO Methods”,[58] Hadzi and Koller state that

the MNDO approximation includes the terms of: one-centre one-electron energies, which

parameters are taken from atomic spectroscopic data and are allowed in the optimization

to account for differences between atoms in molecules and free atoms; the one-centre,

two-electron repulsion integrals Coulomb and exchange integrals, which are derived from

spectroscopic data with some adjustments and are smaller than the analytically calculated

to partially consider the electron correlation ; the two-centre one-electron resonance

integrals that represent the electronic kinetic energy and electrostatic core-electron

energies; the two-centre, one-electron integrals representing the core-electron attractions;

the two-centre two electron repulsion integrals which are evaluated by semiempirical

parametric formulas that simulate multipole-multipole interactions; and the two-centre,

core-core repulsion terms composed by an electrostatic and an additional effective part,

the effective term represents the Pauli repulsion and compensate the errors of the

model.[58]

MNDO, AM1 and PM3 methods are standard implementations of the MNDO model

that have been parameterized mainly with respect to ground-state properties, with special

attention on the energies and geometries of organic molecules. AM1 and PM3 give some

improvement in accuracy over the original MNDO method, but the mean absolute errors

remain of the same order of magnitude.[57]

33

As a result of too repulsive interactions in the core-core potential, MNDO

overestimates the repulsion between two atoms 2 - 3 Å apart.[56] As a solution to this, in

the Austin Model 1 (AM1) by Dewar,[59] the core-core function was modified by adding

Gaussian functions, and the whole model was parameterized again. The Gaussian

functions were added somehow as patches onto the basic parameters, which explains why

different number of Gaussians are used for each atom.[56] According to Jensen in

“Introduction to Computational Chemistry”, some improvements and limitations of the

AM1 model are:[56]

• AM1 predicts the strength of hydrogen bonds more or less correctly, but the

geometry is frequently wrong.

• The activation energies are much better than with MNDO.

• Hypervalent molecules are improved compared to MNDO, but still there are

significant errors.

• Alkyl groups are systematically too stable by around 2 kcal/mol per CH2 group.

Nitro compounds are systematically too unstable. The gauche conformation of

ethanol is predicted to be more stable than the trans.

• Peroxide bonds are ~0.17 Å too short.

• When atoms are around 3 Å apart, phosphor compounds show incorrect

geometries.

The Modified Neglect of Diatomic Overlap, Parametric Method Number 3 (MNDO-

PM3)[60] is a reparameterization of the AM1 with all the parameters automatic fully

optimized. The AM1 expression for the core-core repulsion was kept, except that only

two Gaussians were assigned to each atom. These Gaussian parameters are included as an

integral part of the model, and allowed to vary freely.[56]

The PM3 method has been parameterized using the standard heats of formation of a

large set of typical reference molecules. It has been designed to reproduce standard heats

of formation from total energies (after the inclusion of accurate experimental atomization

heats) in the case of molecular geometries corresponding to the minimal SCF value of

34

trial molecules. PM3 is considered to have the best set of parameters for the given set of

experimental data.[56]

Jensen points out some limitations of the PM3 model:[56]

• Almost all sp3-nitrogens are predicted to be pyramidal, contrary to experimental

observation. The charge in nitrogen atoms is frequently of wrong sign and

magnitude.

• Hydrogen bonds are ~0.1 Å too short. Bonds between Si and Cl, Br and I are

also underestimated.

• The gauche conformation of ethanol is predicted to be more stable than the

trans. H2NNH2 is predicted to have a C2h structure, while the experimental is

C2. ClF3 is predicted D3h, while the experimental structure is C2v.[56]

Another point to take into account about the PM3 calculations is that, as described by

Csonka et al.,[61, 62] the PM3 Hamiltonian has a tendency to create wrong geometries with

H-H interactions between 1.8 and 2.0 Å due to parameterizations errors.

Jensen mentions other limitations which are common to MNDO, AM1 and PM3:[56]

• The rotational barriers for bonds with partial double bond character are

significantly too low.

• The bond length to nitrosyl groups is underestimated.

• The parameters for metals which are included are based on only a few

experimental data.

• For weak interactions, like van der Waals complexes or hydrogen bonds, the

minimum geometry is wrong or the interaction is too weak.

However, there are some distinctions about the use of the semiempirical methods in the

calculations of hydrogen bonding and weak interactions. For instance, in “An

Introduction to Hydrogen Bonding”, Jeffrey[1] states that the semi-empirical methods

such as MNDO and AM1 are considered to be inappropriate for simulating moderate or

weak hydrogen bonding due to an overestimation of the exchange repulsion at hydrogen

35

bond distances and PM3 is said to give better results for systems like the water dimer.[1]

The Semi-ab initio Method 1 (SAM1) was developed by Dewar, Jie and Yu[63] and it is

said to correct this deficiency.[1] SAM1 is based on the NDDO approximation, but instead

of replacing all integrals by parameters, the one- and two-centre electron integrals are

calculated directly from the atomic orbitals.[56] However, for systems like the ammonia

dimer the SAM1 calculations do not lead to correct results.[58] For the formic acid dimer

the AM1, PM3 and SAM1 overestimate the association enthalpy and AM1 overestimates

the O…O distances while PM3 gives better values.[58] Turi and Dannenberg[64] found

similar results for the acetic acid dimer, and they found that the calculated semiempirical

vibrational frequencies were in good agreement with the MP2 results.[58]

Other authors like Zheng and Merz[20] in their studies of hydrogen bonding interactions

relevant to biomolecular structures state that the AM1 geometries were in poor agreement

with ab initio structural results and the PM3 method gives geometries similar to the ab

initio ones. On the other hand, Turi and Dannenberg[65] in their molecular orbital studies

of C-H…O bonded complexes found a good agreement between the energies and

structures at the AM1 and ab initio levels, while the PM3 results were erratic. All of that

shows that, so far, there is no conclusive criterion concerning the selection of an AM1 or

PM3 hamiltonian for the semiempirical calculations of a given system.

According to Thiel,[57] although semiempirical methods can be frequently used with

useful accuracy and at very low computational costs, some general limitations should be

taken into account. One of these is the fact that the errors in semiempirical calculations

are less systematic and harder to correct compared to ab initio or DFT methods. Another

point is that the accuracy of the semiempirical results varies with the classes of

compounds and these variations are more pronounced than in high-level ab initio and

DFT calculations. An additional aspect to consider is that, unlike ab initio and DFT

methods, semiempirical methods require reliable experimental or theoretical reference

data for the parameterizations and they can be used only in molecules which elements

have been parameterized.[57]

36

Density functional theory

The Density Functional Theory (DFT) is based in a one-to-one correspondence

between the electron density of a system and the energy.[56] That means that the ground-

state electronic energy is determined completely by the electronic density.[56, 66] The

advantage of this approach is that the electron density is independent of the number of

electrons. That means that, while the complexity of a wave function increases with the

number of electrons, the electron density is independent of the system size. The problem

is to find the appropriate density functional. Therefore, the goal of the research in DFT is

the design of functionals connecting the electron density with the energy.[56]

The use of DFT methods is based on the introduction of the Kohn and Sham (KS)

formalism[67] which splits the kinetic energy functional in two parts, one that can be

calculated exactly and a small correction term. The kinetic energy is calculated assuming

non-interacting electrons and the remaining kinetic energy is included into an exchange-

correlation term.

Therefore, and accordingly to Kohn and Sham, the functionals used by DFT methods

part the electronic energy into several terms:[68]

E = ET + EV + EJ + EXC (2.3)

Where ET is the kinetic energy term, EV includes terms of nuclear-electron attraction

and nuclear-nuclear repulsion. EJ is the electronic repulsion term and EXC is the

exchange-correlation term that includes the remaining part of the electron-electron

interactions.[68] Since the functional form of the exchange-correlation energy is still

unknown, the main problem of DFT is to find the adequate formulas for this exchange-

correlation term. Therefore, the difference between DFT methods is the choice of the

functional form of the exchange-correlation energy.

The Local Density approximation (LDA) assumes that the local density can de treated

as a uniform electron gas and that means that the density is a slowly varying function.[56]

The exchange-correlation energy is very frequently separated into exchange and

correlation parts. In the LDA approximation, the correlation energy of a uniform gas has

been determined using Monte Carlo methods for different densities and an analytic

37

interpolation formula was developed by Vosko, Wilk and Nusair (VMN)[69] to use these

results in DFT calculations.

According to Guo, Sirois et al. in “Density Functional Theory and its Aplications to

Hydrogen-bonded Systems”,[70] the most used LDA functionals use the Slater functional

for exchange and the VMN formula for the correlation energy. LDA provides reliable

results for molecular properties as structures, vibrations and ionization potentials, but it is

not able to provide an adequate description of hydrogen bonding interactions.[70]

The Gradient Corrected or Generalized Gradient Approximation (GGA) methods are

improvements of the LDA approach that consider a non-uniform electron gas. In the

GGA methods the exchange and correlation energies depend not only on the electron

density, but also on derivatives of the density.[56] Very popular GGA exchange

functionals are those of Becke (B)[71], Perdew and Wang (P),[72] among others.

Commonly used GGA correlation functionals are the Perdew (P86), and Lee-Yang-Parr

(LYP)[73] functionals.

Jensen states that the models that include exact exchange are called hybrid methods,[56]

and Guo, Sirois et al.[70] affirm that hybrid functionals are those that include a component

of Hartree-Fock exchange. One example of this is the very accepted GGA exchange

functionals Becke 3 parameter (B3)[74] hybrid functional.

As already mentioned, the advantage of DFT is that only the total density is considered.

In addition, DFT has a computational cost which is similar to HF theory with the

possibility of providing more accurate results. Some DFT methods are very successful in

studying properties of molecules like structural parameters, vibrational frequencies and

electrostatic potentials, among others. For instance, an study of the hydrogen-bonded

formic acid dimer shows that the energies and the barriers for the symmetrical double

proton transfer are well described by the DFT approach with BLYP functionals.[70]

According to Guo, Sirois et al., the GGA and hybrid functionals are very good for the

study of hydrogen-bonded complexes, since they provide reasonably accurate binding

energies, hydrogen bond geometries and thermodynamic properties for small neutral

complexes.[70] DFT methods predict cooperative effects that agree with MP2 calculations

38

for water polymers and neutral complexes containing a peptide linkage. They also give

good dipole moments and polarizabilities for hydrogen-bonded systems like the water

dimer.[70] However, as Jensen states, unlike the mainly electrostatic interactions, the

dispersive weak interactions are poorly described by the current functionals. In addition,

DFT methods are inappropriate for excited states of the same symmetry as the ground

state.[56]

On the other hand, the developing of DFT functionals is a growing field. Recently,

Truhlar and coworkers have developed new DFT methods for the calculation of π

hydrogen bonding systems.[75, 76] They calculated systems like the dimers of benzene with

water and ammonia, among others. They found that their MPW1B95, MPWB1K,

PW6B95, and PWB6K methods predict accurately the energies and geometries of π

hydrogen bonded systems, in cases where the B3LYP functional fails and the PW91 is

less accurate. They also emphasize the application of their PWB6K functional for

calculating large π hydrogen bonded systems and stacking interactions in the DNA base

pairs and amino acid pairs.[75, 76]

Basis sets

Most quantum mechanical treatments describe each molecular orbital as a linear

combination of atomic orbitals (LCAO approximation).[43, 77, 78] In this approximation,

each atom has assigned to it certain functions that resemble the standard s, p, d and so

atomic orbitals that are centered at the nucleus. Whereas the hydrogen-like orbitals die

off as exp(-ζr), where r is the distance from the nucleus and ζ a constant, the integrals

using this form of the orbital are difficult to evaluate. These Slater-type orbitals (STOs)

are usually replaced by a small number of Gaussian functions, where exp(-ζr) is replaced

by exp(-αr2). The quadratic dependence of r in the exponent greatly simplifies the form

of the integrals, particularly those that involve several atomic centers simultaneously. In

fact, it is computationally more efficient to evaluate a large number of integrals involving

Gaussians than a much smaller number of STO integrals. In addition, a series of

Gaussians with progressively larger values of orbital exponent α can fairly closely

reproduce a Slater-type function. Consequently, most modern quantum chemical

39

calculations are performed using basis sets composed exclusively of Gaussian

functions.[43]

The collections of orbitals that are applied to calculations are called basis sets. The

smallest basis sets uses one orbital to represent each of the orbitals of each shell that is

full or partially filled. The STO-3G[79] is one minimal basis set where each Slater-type

orbital is replaced by a contraction of three primitive Gaussian functions.[43]

Minimal basis sets are improved by doubling the number of functions to provide more

flexibility. A “double-ζ” basis set is similar to minimal, except that each atomic orbital is

split into two. The flexibility of a “DZ” basis permits each orbital to expand or contract in

size to conform to the environment in which the atom finds itself. Triple-ζ or TZ basis set

provides even more flexibility. Splitting the valence shell is worthwhile because the inner

shell electrons are little affected by changes in the bonding environment around the atom.

Therefore, there are split valence basis sets like the 6-31G[80] or the 6-311G.[81] In the 6-

31G basis set the 6 refers to the number of Gaussian primitives used to describe the inner

shell, 1s orbital. The 3 and 1 indicate 3 primitives for the inner and 1 primitive for the

outer valence orbitals. The basis set 6-311G is similar except that a third set of functions

are added, by a single Gaussian, to split the valence shell three ways.[43]

The inner and outer s orbitals of an atom in a double-ζ basis set are both spherical. So,

while the presence of two of them permits the orbital to expand, it can do it isotropically

only, with no stretching in any one direction over another. This “polarization” in a given

direction is necessary in many situations. Therefore, the flexibility is added in the form of

basis functions corresponding to one quantum number of higher angular momentum than

the valence orbitals.[44] So, p-orbitals added to hydrogen are called polarization

functions. Analogously, the p-orbitals of C or O can be polarized by a small amount of a

d-orbital of appropriate symmetry.[43]

To indicate when polarization functions have been added to the basis set, various

conventions are used. The addition of the P in the DZP indicates a polarized double-ζ

basis set but does not clarify whether all atoms have had polarization functions added, or

only some of them. In most cases, the P indicates polarization functions on all atoms.

40

Another designation is an asterisk *. In 6-31G*, the single asterisk indicates polarization

functions of d-type added to the non-hydrogen atoms. A second asterisk would inform of

p-functions on hydrogen, too. One more designation is to indicate the numbers of

polarization functions in parenthesis. Thus, the 6-31G** could be also described as 6-

31G(d,p). Doubling the d-functions, but leaving a single set the p-functions of hydrogen,

would be indicated as 6-31G(2d,p). This notation makes possible also the representation

of orbitals with higher angular momentum, as for example 6-311G(3df,2pd).[43]

According to Cramer, an alternative way to introduce polarization is to allow basis

functions to be centered away from atoms. They are called floating Gaussian orbitals

(FLOGOs) and they are rarely employed in modern calculations, since the process of

geometry optimization is considerably more complicated.[44]

Another way to provide flexibility is the use of functions of, for example, s or p

symmetry, with small orbital component. They are called diffuse functions and are

especially useful for describing systems like anions since they permit the overload of

electrons to better avoid one another since they take advantage of the large expansion of

space over which this orbital extends. The + symbol is used to indicate the presence of

such functions. For example, the 6-31+G* includes a diffuse sp-shell on non-hydrogen

atoms; a second + indicates diffuse functions on H as well.[44]

Correlation consistent (cc) basis sets have been designed[82] specifically to be suitable

to calculations involving electron correlation which has been taken into account at the

basis atomic level, because they have been constructed from the results of correlated

atomic calculations.[56] The smallest is the cc-pVDZ (correlation-consisted, polarized

valence double zeta). These basis sets can be augmented (prefix aug) by additional

functions optimized for atomic anions to describe diffuse electronic distributions. This

augmentation consists of adding one extra function with a smaller exponent for each

angular momentum. For example, the aug-cc-pVDZ has additionally 1s-, 1p- and 1d-

functions.

41

Basis set superposition error

If a bimolecular interaction is considered, the AB interaction energy at the same level

of theory for the monomers and the dimer, can be defined as:[44]

ΔEcomplexation = E(A…B)*ab – [E(A)a + E(B)b] (2.4)

The asterisk (*) denotes the geometry of the complex. The basis functions a and b are

associated with the monomers A and B, respectively, and they are both (ab) used in the

calculation of the complex. Consequently, there are more basis functions employed in the

calculation of the complex than in either of the monomers. The greater flexibility of the

basis set for the complex can provide an artificial lowering of the energy when one of the

monomers “borrows” basis functions of the other to improve its own wave function.[44]

This artificial stabilization of the complex, due only to its larger basis set, in comparison

to the smaller sets of the monomers, is called basis set superposition error (BSSE).

One approximate way to correct this phenomenon is the counterpoise (CP) correction

proposed by Boys and Bernardi.[83] To estimate how much of this complexation energy is

due to BSSE, four additional energy calculations are needed. Using basis set a for A and

basis set b for B, the energies of each of the two fragments are calculated with the

geometry they have in the complex. Two additional energy calculations of the fragments

at the complex geometry are then carried out with the full ab basis set. For example, the

energy of A is calculated in the presence of both the normal a basis functions and with

the b basis functions of fragment B located at the corresponding nuclear positions, but

without the B nuclei present. Such basis functions located at fixed points in space are

called ghost orbitals. The fragment energy for A will be lowered owing to these ghost

functions, since the basis becomes more complete. The CP correction is defined as:

ΔECP = E(A)*ab + E(B)*

ab – E(A)*a – E(B)*

b (2.5)

The counterpoise corrected complexation energy is given as:

ΔEcomplexation CP corrected = ΔEcomplexation - ΔECP (2.6)

ΔECP is an approximate correction that gives an estimate of the BSSE effect, but it does

not provide either an upper or lower limit. Another point to take into account is that this

42

is a single point correction. If the potential energy surface where the minima were found

was “BSSE contaminated”, the results will only tell us how bad the reached minimum is,

and the interatomic distances in the dimer will be wrong to an unknown extent.

Algorithms to perform this correction at each step of the optimization procedure are

available but they are computationally demanding.[84-89] Recently, Crespo, Montero et

al.[90] discussed the influence of the BSSE in the geometries and energies of methane –

nitric oxide dimers at the MP2 and DFT levels of theory with various basis sets. They

state that weak interactions determined by dispersive forces cannot be predicted with

standard non BSSE corrected PES and that DFT results are erratic, even within BSSE

corrected PES.[90]

Hobza and Havlas[91] studied the PES of hydrogen bonded-systems such as the water

dimer, hydrogen fluoride dimer, formamide dimer and formic acid dimer. They conclude

that the CP-corrected and standard PES of these complexes differ, depending on the level

of theory and that the optimization on the standard PES using medium basis sets

sometimes leads to completely wrong geometries, whereas CP-corrected PES yields the

correct structure.

Simon et al.[84] studied the effect of the counterpoise correction on the geometries and

vibrational frequencies of 15 hydrogen bonded systems at the DFT and MP2 levels of

theory, using the 6-31++G(d,p) basis set. They found larger changes at the MP2 level of

theory and that, in general, the CP correction increases the hydrogen bond distance,

decreases the intermolecular stretching frequency and decreases the red shifts of the

donor A-H stretching vibrational frequency. In addition, they observed an interesting

relationship between the percentage of BSSE, the relative changes of the hydrogen bond

distances and intermolecular stretching frequencies. This relationship allows for the

estimation of these properties on the CP-corrected PES, from one single point

counterpoise calculation on the standard PES.[84]

As shown, the calculation of BSSE-corrected potential energy surfaces is a rising field

with many perspectives and open questions. However, the main disadvantage of the

current implementations of the algorithms to perform this correction at each step of the

optimization procedure is the high demand of computational time, which makes this type

43

of calculations prohibitive in many cases such as for bigger systems or when using larger

basis sets.

According to Scheiner in “Hydrogen Bonding: A Theoretical Perspective”, there are

other important facts to consider about the BSSE:[43]

• In general, BSSE is reduced as the basis set becomes larger and more flexible.

However, there is no strict correspondence, and the BSSE in some cases can

become larger with the increasing of the size of the basis set.

• The BSSE rises rapidly as the two molecules approach one another. Whereas

SCF BSSE can be reduced to negligible proportions with large basis sets, the

superposition error at correlated levels goes down much more slowly, persisting

at large values, even with very flexible bases.

An alternative to the CP method is the Chemical Hamiltonian Approach (CHA)[92] that

attempts to isolate the superposition error directly in the Hamiltonian operator. That

means that the CHA is based on omitting from the Hamiltonian the intermolecular

contributions that cause the BSSE, while keeping all the physically true interaction terms.

Omitting these terms, a Hamiltonian which leads to the BSSE-free wave functions is

achieved. Thus, the CHA ensures that the description of each monomer or fragment

isolated and within the complex is consistent.[43, 93, 94]

According to Meyer et al.,[95] the CHA method performs a detailed analysis of different

terms in Hamiltonians and distinguishes three different types:

• The terms corresponding to true intramolecular interactions of the monomers

treated in their own molecular basis.

• The terms describing the true intermolecular interactions.

• Some finite basis correction terms responsible of BSSE.

As Bende[93] points out, several different approaches have been developed using the

CHA scheme both at the HF level and using second order perturbation theory.[96-98] Many

calculations have been performed in the last decade by applying this CHA to study the

44

structures and interaction energies for different van der Waals and hydrogen bonded

systems.[86, 93, 94] Different systems were investigated, from the small “bimolecular

complexes” like formamide dimers by Bende et al.[93] to large DNA basis pairs,[99] by

using a variety of different basis sets. It has been concluded that in all cases a remarkable

agreement has been found with the results given by the CP method, despite of the fact

that both schemes are conceptually very different. However, Scheiner states that the CHA

approach has been criticized on the grounds of its inconsistency with the results of

symmetry adapted perturbation theory and because, since the BSSE is a non-physical

phenomenon, this Hamiltonian is non-Hermitian.[43]

According to Salvador et al., the removal of the BSSE in molecular complexes

composed by more than two fragments is not extensively discussed in the literature.[87] A

few years ago, Turi and Dannenberg[100] pointed out the ambiguity of the counterpoise

correction when studying growing chains of hydrogen fluoride. They showed that the

BSSE computed for the addition of a new HF monomer to the (HF)n aggregate depends

upon whether the incoming monomer is added to the H or to the F end of the aggregate.

Therefore, one can obtain different interaction energies for the same chemical process,

which is obviously wrong. They proposed the use of the counterpoise method by defining

as many fragments as there are monomer subunits in the complex, with the BSSE defined

as the difference between the energy of each monomer in its own basis set and that of the

whole aggregate. This method solves the problem of the ambiguity of the CP correction

but is unable to explain all the effects of the incoming monomer on the interaction and

BSSE already present in the molecular aggregate.

As Salvador explains, another way to take into account the high-order BSSE effects

within the counterpoise framework is based on a hierarchical counterpoise scheme[101] for

N-body clusters that treats the basis set extension effects of all the monomers, dimers,

trimers and so on, present in an aggregate. Another counterpoise scheme is the pairwise

additive function counterpoise (PAFC)[102] where the counterpoise correction is carried

out over pairs of fragments.

In short, three counterpoise schemes for trimers are mentioned: the site-site[102] function

counterpoise which includes only the basis set extensions of the monomers in the whole

45

basis set, the Valiron-Mayer hierarchical function counterpoise which includes the

differences, for each dimer in the aggregate, between the dimer interaction energy

computed within the dimer-centered basis set (DCBS) and the basis set of the whole

trimer (TCBS); and the pairwise additive function counterpoise in which the three-body

interaction terms are not corrected according to the counterpoise scheme and instead,

only the two-body interaction energies are corrected by using DCBS.

The N-body cluster generalizations for these three function counterpoise schemes, the

site-site (SSFC), pairwise additive(PACF) and Valiron-Mayer (hierarchical) (VMFC) are

as follows:[87]

Figure 1.3: The N-body cluster generalization for the site-site (SSFC), pairwise

additive(PACF) and Valiron-Mayer (hierarchical) (VMFC) counterpoise schemes.

Figure taken from “Counterpoise-corrected geometries and harmonic frequencies of N-

body clusters: Application to (HF)n(n=3,4)” by Salvador and Szczesniak.[87]

46

2. The Multiple Minima Hypersurface (MMH) Approach

2.1 Introduction

The multiple minima problem

According to Cramer in “Essentials of Computational Chemistry, Theories and

Models”,[44] the first step for making the theory closer to the experiment is to consider not

just one structure for a given chemical formula, but all possible structures. This involves

the full characterization of the Potential Energy Surface (PES) for a given chemical

formula. The PES is a hypersurface defined by the potential energy of a collection of

atoms over all possible arrangements and has 3N - 6 coordinate dimensions (for linear

molecules 3N - 5). Therefore, each structure, which is a point on the PES, can be defined

by a vector X where[44]

X ≡ ( x1, y1 , z1 , x2 , y2 , z2 ,… , xN , yN , zN ) (3.1)

and xi, yi, and zi are the Cartesian coordinates of atom i.

Of particular interest on PES are the local minima, which correspond to optimal

molecular structures and saddle points which are lowest energy barriers on paths

connecting minima. The saddle points are related to the chemical concept of the transition

state. Therefore, a complete PES provides, for a given collection of atoms, complete

information about all possible chemical structures and all isomerization pathways

interconnecting them.[44]

However, the problem of locating the global minimum or an important and

representative set of local minima in polyatomic systems is a very difficult computational

task, taking into account that the number of minima increases very much with the number

of variables. This problem is frequently called the Multiple Minima Problem in the

literature.[56]

In “Introduction to Computational Chemistry” Jensen describes some methods which

are commonly used for conformational sampling in larger systems:[56]

• Stochastic and Monte Carlo methods

• Molecular dynamics

47

• Simulated annealing

• Genetic Algorithms

• Diffusion methods

• Distance Geometry methods

He emphasizes that none of them guaranty to find the global minimum, but in many

cases they provide a local minimum which is close in energy to the global one, but not

necessary close in terms of structure.

The Stochastical and Monte Carlo (MC) methods start from a given geometry, usually

a local minimum, to generate new configurations by adding a random perturbation to one

of more atoms.[56] In Monte Carlo methods, if the energy of the new geometry is lower

than the present, this new geometry is taken as starting point for the next step. If not, the

Boltzmann factor e -ΔE/KbT is calculated and compared to a random number between 0 and

1. If the Boltzmann factor is less than this number the new geometry is accepted,

otherwise the next step is taken from the old geometry. In this way, a sequence of

configurations is generated from which geometries may be selected for following

minimization. To have a fair acceptance radio, the step size must be reasonably small.[56]

In Stochastical methods the random perturbation is usually larger compared to Monte

Carlo and a standard minimization is performed starting at the perturbed geometry. Then,

a new perturbed geometry is generated and minimized, and so on.[56] It is important to

consider the length of the perturbing step, because small perturbations basically return the

geometry to the starting minimum and large perturbations may lead to high energy

structures that minimize to high energy local minima. Another point to consider is that

when the perturbing step is done in Cartesian coordinates, many of the perturbed

geometries are high in energy, since two o more atoms are moved close together by the

perturbation.[56] The perturbation can be done in a selected set of internal coordinates,

too. The perturbation step may be taken from the last minimum found or from all the

previous found minima, weighted by a probability factor. Therefore low energy minima

are more frequently used than the higher energy ones.

Molecular Dynamics (MD) methods solve the Newton equation of motion for atoms on

an energy surface.[56] The energy of the molecule is distributed in potential and kinetic

energy, and therefore, molecules are able to rise above barriers separating minima if the

48

barrier height is less than the total energy minus the potential energy. The energy is

related to the simulation temperature and with energy high enough, the dynamics would

sample the entire surface. However, this will not be practical by means of computational

time. Very small time steps must be used for integrating Newton’s equation, therefore the

simulation time is short. That, combined with the use of temperatures of few hundreds or

thousands of degrees, means that only the local area around the starting point is sampled

and only relatively small barriers of few kcal/mol can be overcome. Different local

minima can be generated by selecting configurations at appropriate intervals during the

simulation and subsequently minimizing these structures.[56]

According to Jensen,[56] the Simulated Annealing (SA) techniques choose high initial

temperatures 2000 – 3000 K to initiate an MD or MC run, during which the temperature

is slowly reduced. The molecule is initially allowed to move over a large area, but with

the decreasing of the temperature it becomes trapped in a minimum. That means that, in

principle, if the cooling is done infinitely slowly at infinite run time, the resulting

minimum would be the global minimum. In practice, the MD or MC run is so short that

only the local area is sampled.[56]

Other methods are the Genetic Algorithms (GA) which are based on biological concepts

and terminology.[56] Accordingly, there is a “population” of structures, each characterized

by a set of “genes”. The parent structures generate the new structures which have a

mixture of the parent genes, and “mutations” are allowed to occur in the process. Their

energies are minimized and the best new and parent structures are selected and used for

the next generation and so on. The population is allowed to develop for hundreds of

generations. The criterion for selecting the best structures in each generation may vary

and can be, for instance, that the good structures are those with lower energy. An

example of mutation can be the random variation of geometrical parameters to produce

conformations outside the restricted range of the current population. The size of the

population, the mutation rate and the amount of structures selected in each generation,

can be adjusted, among other parameters. Due to their reliability and facilities for

implementation, the use of genetic algorithms is becoming very popular.[56]

Diffusion methods are those where the energy function is changed to contain only one

minimum.[56] One change may be the addition of a contribution proportional to the

second derivative of the function (local curvature). Consequently, the minima are

increased in energy and maxima and saddle points are reduced in energy, and finally only

49

one minimum remains. Using the single minimum geometry of the modified potential,

the process can be reversed and a minimum on the original surface is obtained. This

minimum is frequently, but not necessarily the global minimum.[56]

In Distance Geometry methods, test geometries are produced from a set of lower and

upper limits on distances between all pairs of atoms, and therefore many different trial

sets of distances can be generated by assigning random numbers between these limits.

The random distance matrix can be translated into a three-dimensional structure and trial

conformations are generated, which can be optimized using conventional methods.[56]

Chemical similarity searching

As Willett, Barnard and Downs explain in “Chemical Similarity Searching”, the

chemical similarity searching is an alternative and complementary method to the

substructure searching.[103] The substructure searching is a mechanism that involves the

recovery of all those molecules in a database that contain a query substructure defined by

the user, irrespective of the environment in which the query substructure occurs.[103] One

limitation of the substructure searching is that the database structure must include the

entire query substructure. In the similarity searching, a query involves the specification

of an entire molecule or structure as reference which is characterized by one or more

structural descriptors, and this set is compared to the corresponding set of descriptors for

each of the molecules in the database. This comparison allows to measure of similarity

between the reference structure and each of the database structures. Therefore, with the

increasing importance of databases of two-dimensional and three-dimensional molecular

structures in chemical research, the chemical similarity searching is becoming a powerful

source of useful and interesting information for chemists, especially in the areas of

molecular design and computer-aided synthetic planning.[103, 104]

According to Willet et al.,[103] the similarity coefficients provide a quantitative measure

of the degree of structural relatedness between a pair of structural representations. Some

coefficients are measures of the distance or dissimilarity between objects and have value

of 0 for identical objects, whereas others measure similarity and have their maximum

value for identical objects. In most cases the values taken for a coefficient are in the

range from 0 to 1.[103] Since the idea of determining a numerical measure of the similarity

between two atoms is common to different sciences apparently not related to each other,

the most used similarity coefficients have been sometimes reinvented, as shown in Figure

2.1.[103]

50

Figure 2.1: Description of two similarity coefficients commonly used in chemical information. Figure taken from “Chemical Similarity

Searching” by Willett, Barnard and Downs.[103]

51

There are several similarity coefficients that can be used for similarity searching.

However, the Tanimoto index has the advantage of the accessibility of its formula for

dichotomous variables for implementation in different subjects (Figure 2.1). When

coefficients are monotonic to each other, it can be shown analytically that they will

always produce identical similarity rankings of objects against a specified reference, even

though the actual coefficient values are different.[103] The Tanimoto index is monotonic

with the Dice coefficient. The use of the Tanimoto coefficient for similarity searching is

described in several studies in the scientific literature.[103-105]

2.2 The Multiple Minima Hypersurface (MMH) approach

The Multiple Minima Hypersurface (MMH) approach combines the quantum chemical

procedures for calculating the cluster energies in local minima of supermolecules and the

statistical thermodynamics approach for the evaluation of macroscopic properties. It was

first introduced in 1998 with the study of the pH-dependent equilibria of 2,4-diamino-5-

phenylthiazole tautomer molecules in water by Montero, Esteva et al.[5]

This approach was developed by Montero and coworkers with the original purpose of

evaluating the thermodynamic association functions of various molecular clusters using

the partition function.[4-6, 106] Therefore, it is advisable to explore the multiple minima

hypersurface (MMH) of the supermolecular systems. This exploration is made by

generating several sets with initial random geometries and following a gradient pathway

to search the local minimum that could be statistically significant for the partition

function. Therefore, it is of key importance to find the appropriate collection of

supermolecules with their respective energies and geometries that represent a set of the

most important states which are significant to the thermodynamic properties of the whole

system. Their energies can be added to the partition function, and by using statistical

thermodynamics the thermodynamic association functions can be calculated through the

partition function.

Here, the MMH approach is arbitrarily divided in two parts: The statistical

thermodynamic analysis and the structural analysis. The statistical thermodynamic

component of this method is not discussed here, since this work is only based on the

structural analysis. The study of the various possible geometrical arrangements of a given

system and the molecular interactions in the system is called structural analysis. The

generation and calculation of different geometries was originally conceived as a

52

necessary step for the thermodynamical analysis. Here, the structural analysis is further

developed as an independent tool for exploring the potential energy surfaces of a variety

of systems.

Therefore, the MMH approach as a very useful and reliable tool for localizing the

minima of hydrogen-bonded and weakly interacting systems is introduced here. The

MMH procedure provides a way to study the multiple possible molecular arrangements

for a given system via the exploration of large numbers of automatically generated

starting geometries. The MMH approach has the advantage that each structure is treated

and optimized independently, which is especially important in the treatment of very flat

potential energies surfaces without a well defined “global” minimum. The number of

minima increases very much with the number of variables; therefore the “global

optimization” is an extremely demanding task. Many times a “global minimum” is just a

local minimum with low energy.

In many instances the finding of the “global minimum” is not sufficient to describe the

properties of the system. The statistical thermodynamic approach of MMH, which is not

discussed here, shows the importance of considering local minima.[4-6, 106] The structural

analysis shows that many interactions which contribute significantly to the stabilization

of the system appear in various “local” minima. Additionally, sometimes the global

minimum is not clearly defined since it depends on the level of theory or the accuracy of

the calculation. Only by a careful and systematic exploration of the PES of the system it

is possible to determine the set of geometries representative for the interactions that

stabilize the system and provide an adequate description of the association process.

Frequently, the starting geometry for the optimization of the “global” minimum of a

system depends on “chemical intuition”, ignoring other alternatives. This may be less

severe in very simple systems, but for larger and more complicated systems, the use of a

systematic and reliable procedure for searching the minima, like MMH, becomes a

necessity. Even for some small systems, MMH can provide surprising structures that

otherwise might be excluded.

There are several steps in the MMH procedure:

1. Generation of starting geometries

The program for generating random sets was especially written for this purpose and is

called GRANADA. The GRANADA program[4] inputs molecular set data, such as the

53

radius of distribution around the coordinate origin in one monomer X (or more general,

subunit X) and the number of molecules of the other subunit or monomer Y to be taken

into account. As can be seen, the generation of geometries is not only restricted to dimers,

since the amount of molecules of monomer Y can be more than one. The definition of

which monomer is X and which one is Y is arbitrary and depends on the user. The user

also determines the number of randomly generated geometries. In most cases, here,

around 1000 randomly arranged clusters or complexes are generated as starting points.

(Figure 2.2)

Figure 2.2: Selection of random geometries of the FA – FMA dimer, produced with

the GRANADA program

Then, the GRANADA program outputs the desired series of different arrangements of

X and Y molecules as input files for MOPAC. Each Y molecule is situated in a randomly

selected new center of coordinates with respect to the X molecule and is rotated, also

randomly, around the three coordinate axes. All cases where the newly generated

molecules overlap the van der Waals volume of any existing molecule, are discarded.

Seeds for the random number generating routines are taken for each new molecular set

from the product of the seconds and the hundredths of seconds of the computer clock

when the program begins each calculation, to avoid any equivalent random number

series. Randomness has been carefully tested.

54

2. Preliminary calculation of the energy

To calculate the preliminary energies of all the generated molecular arrangements,

semiempirical Hamiltonians are used. Since hundreds of thousands of SCF cycles, with

their respective supermolecular geometries, are computed during the process, the amount

of calculations involved at this level of refinement makes the use of ab initio methods not

possible. Despite of controversial opinions, semiempirical methods have been confirmed

as a good choice for this previous discrimination of geometries. In addition, such

calculations have important advantages: computations are fast which allows for the

exploration of large amounts of possible starting geometries and correlation effects are

implicitly considered during the parameterization procedures with respect to

experimental values. Therefore, the PM3 method which is, for instance, able to reproduce

the hydrogen-bonding patterns between water molecules,[107] is selected. Since the choice

of an adequate semiempirical Hamiltonian is still a controversial field, the geometries are

in addition optimized with the AM1 Hamiltonian and both results are compared.

All the semiempirical calculations are performed with the MOPAC program.[108-110] The

minimization method selected is the Eigenvector following (EF),[111, 112] which is

designed to search for critical final structures, as transition states. The EF is supposed to

avoid unexpected molecular rearrangements in the final steps of the minimization path

due to any possible overestimation of geometry changes when the molecules are too close

together. In fact, this very common problem of MOPAC optimization of nonstandard

structures, with the keyword precise, is avoided in all of the calculations performed.

Therefore, the procedure involves a large number of molecular arrangements during the

gradient-driven path to explore the multiple minima hypersurface (MMH) of such

systems. The followed path leads to local minima of potential energy of the molecular

cluster.

3. First discrimination and similarity analysis

Once the geometries of all the starting structures are optimized at a semiempirical level,

the next step is the geometrical and energetic discrimination of the resulting complexes.

The program for processing the MOPAC output data is called Q3.[4] This program

performs the statistical thermodynamic analysis and discards all optimized structures

which are degenerated. In a first approach, all structures with the same energy compared

to a previous one, are discarded. Later, it is necessary to consider that there are two types

of degeneracy in this kind of potential energy surface minima. The first consists of

55

clusters which are identical, which means that they have both the same energy and

molecular geometry, called similarity degeneracy (SD). The second consists of clusters

with different molecular geometry but the same energy, called valid degeneracy (VD). In

the latter, to follow only an energy criterion for the discrimination between structures

might lead to neglect important contributions.

Therefore, a subroutine called Tanimoto is introduced in the program to analyze the

similarity among molecular arrangements, to discard the SD and keep the VD and

preserving, in general, those degenerate clusters when symmetry provides the same

molecular configuration but molecules themselves are exchanged. The Tanimoto

procedure uses the Tanimoto similarity index to calculate the similarity between

structures pair by pair. For this purpose it first converts internal coordinates to Cartesians

for all atoms in a given structure. Then, it obtains the matrix of the position vector

modules with respect to an origin fixed at atom 1, called [D]:

[ ] [ ]NrrrrD ,...,,, 321≡ (3.2)

where N is the total number of atoms.

The Tanimoto similarity index T corresponding to a comparison of cluster A with B is

determined by the expression:

( )MBAMT −+= (3.3)

where A, B, M are calculated using the following expressions:

Bi

N

i

Ai rrM ∑

=

=1

(3.4)

∑=

=N

i

Ai

Ai rrA

1 (3.5)

∑=

=N

i

Bi

Bi rrB

1

(3.6)

where Air is an element of [DA] and B

ir of [DB] that belongs to A and B structures,

respectively.

The Q3 program compares all clusters with an energy difference of less than 0.096

kJ/mol compared to any previous one. This is an arbitrary limit of 10-3 ev to consider that

56

these clusters have the same energy, by far below the accuracy of semiempirical methods.

Then, a limit value of discrimination to consider the clusters equal or not from a

geometrical point of view is defined. For example, if the calculated value of T is equal or

larger than T' = 0.85 (highest limit value of the Tanimoto index), these molecular

arrangements are equivalent and geometrically SD degenerated clusters. If T is less than

0.85 the clusters are different, even when they have the same energy. The latter case

corresponds to VD. As would be expected, the selection of the most adequate highest

limit value of the Tanimoto index T’ depends on the complexity of the system

investigated. Several values were tested, and in general, for bimolecular and simple

trimolecular complexes, the T' = 0.85 value provides a good reference for the geometric

discrimination.

It is important to include structures with VD to validate the partition function in the

statistical thermodynamic analysis. Given that the 10-3 ev energies degeneracy limit is by

far below the accuracy of semiempirical Hamiltonians, it is expected that keeping the VD

structures would not play a key role for the structural analysis. However, since the

geometries of all these structures are subsequently refined at higher level of theory, there

is a possibility, especially when the PES is very flat, that some of these VD structures

might lead to different minima at other level of theory. In addition, the inclusion of the

similarity analysis in the search of supermolecular geometries is a conceptual advance in

the theoretical development of the MMH approach.

4. Refinement of the geometries

This step is not included in the statistical thermodynamic original MMH approach, in

which the thermodynamic analysis is performed in the third step, using the Q3 program,

as already mentioned. But for the purposes of the structural analysis, the semiempirical

results provide just a preliminary overview of the interactions in the complex. For this

reason, the set of relevant semiempirical local minima are refined using DFT and MP2

methods.

The selection of the next level of theory depends very much on the characteristics of

the system to study. For example, for very weak interacting systems the use of the

currently available popular DFT methods is not recommended. However for moderate-

“strong” hydrogen bonded systems like the formic acid – formamide dimer, DFT

provides results in good agreement with the ab initio calculations. Here, for an initial

discrimination of the semiempirical structures, the MP2 method with a small – medium

57

basis set like the 6-31G(d,p) is used, and the size of the basis set is gradually increased,

including correlation-consistent basis sets. During all the process, including the analysis

of the initial semiempirical geometries, each structure is individually analyzed and

calculated. This point gains special relevance in very weakly interacting systems.

The tight criterion is used in the optimization of geometries and force constants are

calculated when necessary. In weak interacting systems the distortions of the geometries

from the imaginary frequency in transition states are carefully followed. Many times this

leads to new local minima, as expected in very flat potential energy surfaces. The BSSE-

CP corrections during the optimizations of the geometry are included only with the

smaller basis set since these corrections are very demanding when larger basis sets are

used. The utilization of different basis sets, the calculation of vibrational spectra, the

analysis of the individual contributions that stabilize each structure, together with the

comparison of the calculated minima or their properties with experimental results and

similar complexes from the literature, allow for a better characterization of the systems

under study.

58

3. Formic Acid Complexes with Formamide and Dimethyl ether

3.1. Introduction

Hydrogen-bonded complexes have been subject to a large number of studies in

chemistry and life sciences. Oxygen atoms as hydrogen bond acceptors are of particular

interest since they play a key role in biological processes.

Formic acid (FA) and its interactions with other molecules provide a very good model

to understand a large variety of hydrogen bond interactions, from classical and strong to

weaker and non-classical bonds.[20, 113-118] The proton transfer mechanism of the FA

dimer has been extensively described,[119-127] and studies of the complexes of FA with

other molecules like water were carried out in our group using matrix isolation

spectroscopy in combination with ab initio methods.[117]

The formamide (FMA) molecule is the simplest molecule that contains a peptide

linkage. Therefore, it can be used as a simple model of hydrogen bond interactions

involving carboxylic acids and amino groups in biological systems, like protein-protein

and protein- substrate interactions.[21, 115] Complexes involving the interactions of ethers

with molecules of biological interest are also described in literature,[128-130] including

matrix isolation and ab initio studies of dimethyl ether (DME) complexes with water,[131]

methanol,[132, 133] hydrogen peroxide[134] and hydracids.[135]

FA – FMA complexes are very interesting hydrogen bonded systems. In addition to its

biological interest, they provide good models for studying the competition between non-

covalent interactions involving nitrogen and oxygen atoms in the same molecule. The

FMA homodimers, as well as its complexes with other molecules like water and

methanol have been intensively studied.[136-138] However, only few reports are found

about FA – FMA heterodimers. The computational study of the electron-density

dependent properties of FA, FMA and their homo and heterodimers made by Gálvez and

Gómez,[115] the crystallographic structure by Nahringbauer in 1968[139] and the ab initio

calculations of Neuheuser[140] are of special interest.

Here, several minima of the FA – FMA potential energy surface are described and their

geometries, binding energies and vibrational spectra are discussed. In addition, the 1:2

and 1:4 FA – FMA complexes are investigated. The FA – FMA trimers allow for the

59

analysis of the influence of a third molecule on the dimer properties, e. g. an additional

FMA molecule on the FA – FMA dimer or a formic acid molecule on the FMA dimer.

The structures of the FA – FMA dimers and trimers are compared to the FMA – water

and FMA – methanol dimers from literature data and with the reported FA – FMA crystal

structure.

The calculated structures, binding energies and vibrational properties of FA – DME

complexes are also discussed using ab initio calculations. The FA – DME system exhibits

classical as well as weak OH…O, C=O…H, C-O…H and CH…O hydrogen-bonding

interactions, making these complexes very interesting for theoretical and experimental

research. The results obtained with various computational methods and basis sets are

discussed, as well as the influence of the basis set superposition error (BSSE) on the

calculated energies and geometries of the different complexes. The calculated vibrational

spectra of the dimers are compared to experimental matrix isolation spectra.

3.2. Computational methods

The Multiple Minima Hypersurface (MMH) approach is used for searching

configurational minima in the FA – FMA and FA – DME systems. For each case around

one thousand randomly arranged clusters were generated as starting points, and the

resulting geometries were optimized and analyzed using PM3 and AM1 semiempirical

quantum mechanical Hamiltonians. These semiempirical results provided a preliminary

overview of the interactions in the complexes, and the relevant configurations were

further refined using DFT and ab initio methods. The AM1 Hamiltonian did not

contribute with new minima to the PM3 structures which were further refined using ab

initio methods.

In the case of the 1:2 and 1:4 FA – FMA complexes, 250 and 198 random geometries,

respectively, were taken as starting point for the PM3 geometry optimizations. A

selection of the resulting geometries was optimized using the B3LYP density functional

and Dunning’s correlation consistent triple ζ basis set.

The ab initio and DFT computations were performed using the Gaussian 98,[141]

Gaussian 03,[141] and MOLPRO[142] programs. In all cases the equilibrium geometries

and vibrational frequencies were calculated using second order Møller−Plesset

perturbation theory (MP2). Pople’s 6-31G(d,p) and 6-311++G(d,p) basis set as well of

60

augmented and non augmented Dunning’s correlation consistent double and triple ζ basis

sets (cc-pVDZ, aug-cc-pVDZ, cc-pVTZ and aug-cc-pVTZ) were used.

For the FA – FMA complexes the geometries and frequencies were calculated also

using the density functional theory (DFT) with the B3LYP hybrid functional[73, 74] and

single point calculations were done with coupled clusters of single and double

substitutions (with non iterative triples) CCSD(T)/aug-cc-pVTZ.

The stabilization energies were calculated by subtracting the energies of the monomers

from those of the complexes and corrected for the basis set superposition errors (BSSE)

using the counterpoise (CP) scheme of Boys and Bernardi. ZPE corrections were also

included.

To investigate the influence of the basis set superposition errors (BSSE) on the

geometries of the complexes, the most stable FA – FMA and FA – DME dimers were

optimized at the MP2/6-31G(d,p) level of theory using the CP scheme during the

optimization process. In addition, the geometries were optimized without BSSE at the

same level of theory to compare the influence of the BSSE on the binding energies as

well as on the geometries. The small 6-31G(d,p) basis set was selected for this purpose

since the BSSE is expected to be more pronounced with small basis sets, and in addition

the computations are less demanding.

3.3. Formic acid – formamide complexes. Results and discussion

Formic acid – formamide dimers

Geometries and binding energies. Analysis of the intermolecular interactions

Nine FA – FMA complexes A – I were localized after MMH search and refined with

both DFT and MP2 calculations (Figure 3.1). The use of Dunning´s cc-pVDZ, aug-cc-

pVDZ, cc-pVTZ, and aug-cc-pVTZ basis sets (Table 3.1) revealed that the geometries of

the complexes are almost independent of the basis sets used. Therefore, the hydrogen

bond distances and angles are discussed here at the MP2/aug-cc-pVTZ level of theory,

only.

61

Seven basic types of interactions (1) – (7) can be differentiated in the FA – FMA

complexes:

(1) NHFMA…O=CFA interaction between the amide hydrogen atom of FMA and the

carbonyl oxygen atom of FA.

(2) C=OFMA…HOFA interaction between the carbonyl oxygen atom of FMA and the

hydroxyl hydrogen atom of FA.

(3) (O)CHFMA…O=CFA interaction between the aldehyde hydrogen atom of FMA and

the carbonyl oxygen atom of FA.

(4) NHFMA…(H)OCFA interaction between the amide hydrogen atom of FMA and the

hydroxyl oxygen atom of FA.

(5) C=OFMA…HC(O)FA interaction between the carbonyl oxygen atom of FMA and

the aldehyde hydrogen atom of FA.

(6) HN(H)FMA…HOFA interaction between the nitrogen atom of FMA and the

hydroxyl hydrogen atom of FA.

(7) (O)CHFMA…(H)OCFA interaction between the aldehyde hydrogen atom of FMA

and the hydroxyl oxygen atom of FA.

62

TABLE 3.1: Calculated binding energies and ZPE and BSSEa corrected binding

energies (in kcal/mol) of the FA – FMA dimers A – I

MP2 cc-pVDZ aug-cc-pVDZ cc-pVTZ ΔE ΔE (ZPE) ΔE (BSSE) ΔE ΔE (ZPE) ΔE (BSSE) ΔE A -18.37 -15.92 -11.60 -16.53 -14.30 -14.21 -16.90 B -14.56 -12.57 -8.67 -12.87 -11.09 -11.03 -13.21 C -11.38 -9.72 -6.69 -9.93 -8.36 -8.23 -10.16 D -10.06 -8.29 -5.34 -8.97 -7.37 -7.41 -8.89 E -10.21 -8.03 -5.19 -7.54 -5.85 -5.81 -8.04 F -7.58 -6.33 -3.67 -6.38 -5.16 -4.97 -6.12 G -6.57 -4.95 -3.15 -5.42 -4.34 -4.29 -5.58 H -6.26 -5.21 -2.47 -5.50 -4.56 -4.31 -5.34 I -5.37 -4.50 -2.15 -4.64 -3.82 -3.52 -4.37 MP2 B3LYP aug-cc-pVTZ cc-pVTZ

CCSD(T)/cc-pVTZ // MP2/aug-cc-pVTZ

ΔE ΔE (ZPE) b ΔE (BSSE) ΔE ΔE (ZPE) ΔE ΔE (ZPE)

b A -16.64 -14.41 -15.25 -16.03 -13.97 -16.76 -14.53 B -12.96 -11.18 -11.87 -12.2 -10.46 -13.21 -11.43 C -9.95 -8.38 -8.92 -8.97 -7.53 -10.00 -8.42 D -8.71 -7.11 -7.86 -7.77 -6.25 -9.06 -7.45 E -7.32 -5.63 -6.37 -6.22 -4.52 -8.06 -6.37 F -6.05 -4.83 -5.26 -4.88 -3.78 -6.24 -5.02 G -5.2 -4.12 -4.58 -4.63 -3.59 -5.40 -4.32 H -5.17 -4.23 -4.60 -4.37 -3.44 -5.60 -4.66 I -4.29 -3.47 -3.73 -3.37 -2.60 -4.54 -3.71

aBSSE corrected binding energies for the cc-pVDZ, aug-cc-pVDZ and aug-cc-pVTZ

basis sets at the MP2 level of theory. bZPE correction from the MP2/aug-cc-pVDZ

calculations.

63

Figure 3.1. The calculated structures with hydrogen bond lengths (Å) and angles

(degree) of the FA– FMA dimers A - I at the MP2/ aug-cc-pVTZ level of theory.

64

The most stable FA – FMA dimer calculated is complex A with a binding energy of

-14.41 kcal/mol (MP2/aug-cc-pVTZ + ZPE), the ZPE correction is taken from the

MP2/aug-cc-pVDZ calculations. The energies of the other dimers B – I are also discussed

at this level of theory (Table 3.1).

The dimer A is stabilized by interactions (1) and (2) involving both carbonyl groups

and the N-H and O-H hydrogen atoms of the formamide and formic acid molecules. The

binding distances are 1.859 and 1.637 Å, respectively. In dimer B (-11.18 kcal/mol) the

amide hydrogen atoms of the FMA are not involved in the stabilization of the complex.

Instead, the aldehyde hydrogen atom of the FMA interacts with the carbonyl oxygen

atom of the FA at 2.304 Å (interaction 3). Cyclic dimer B is also stabilized by interaction

(2) with a binding distance of 1.663 Å (around 0.025Å longer than interaction (2) in

complex A).

The difference between the binding energies of complexes A and B is more than 3

kcal/mol. This is explained by the different hydrogen bond capabilities of the N-H vs. C-

H hydrogen atoms of FMA. Consequently, the difference between the hydrogen bond

distances of interaction (1) in A and interaction (3) in B is more than 0.4 Å.

Compared to dimer A, the FA – FMA complexes C and D are between 6 - 7 kcal/mol

less stable. The binding energy of complex C is calculated to -8.38 kcal/mol. In dimer C,

as in A and B, the carbonyl oxygen atom of FMA is interacting with the hydroxyl

hydrogen atom of FA (interaction (2)). However, with 1.735Å the hydrogen bond

distance in C is considerably larger than in A and B (Figure 3.1). In dimer C the carbonyl

oxygen atom of FA is not involved in the stabilization of the complex. Instead, one of the

amide hydrogen atoms of FMA interacts with the hydroxyl oxygen atom of FA at a

distance of 2.213 Å (interaction (4)).

Dimer D has a binding energy of -7.11 kcal/mol and it is energetically very close

(about 1 kcal/mol) to dimer C. In dimer D, again both carbonyl groups of the formamide

and formic acid molecules are involved in the stabilization of the complex via interaction

(1) (1.968Å hydrogen bond distance) and interaction (5) between the carbonyl oxygen

atom of FMA and the aldehyde hydrogen atom of FA (hydrogen bond distance 2.258 Å).

In this case, the O-H group of the FA is not involved in the stabilization of the dimer.

65

The cyclic-non planar structure of dimer E is an interesting case. The amide group is

pyramidalized, and thus makes possible interaction (6) between the nitrogen atom of

FMA and the O-H hydrogen atom of FA (2.003 Å). Complex E is also stabilized by

interaction (1) with a 2.133 Å distance. The calculated binding energy of dimer E is -5.63

kcal/mol.

The dimers F – I are weakly bound and energetically very close to each other. The

binding energies vary between -4.83 and -3.47 kcal/mol. With the exception of the non

planar dimer G, stabilized only by interaction (1), all dimers are cyclic. Dimer F is

stabilized by interactions (4) and (5) and dimer H by interactions (3) and (5). In dimer I,

interaction (5) appears together with the weakest interaction (7) between the carbonyl

hydrogen atom of FMA and the hydroxyl oxygen atom of FA resulting in a large distance

of 2.561 Å. In none of the F – I dimers the hydroxyl hydrogen atom of the formic acid

molecule interacts with other groups.

All the complexes discussed here were produced from randomly generated geometries

and not via chemical intuition. It is thus interesting to note that:

(a) The calculated geometries of the FA – FMA dimers A and B are in complete

agreement with the calculated structures of the FA – FMA complexes proposed by

Neuhauser[140] and Galvez[115] in their ab initio and DFT studies. These complexes show

also interesting analogies with the FMA – water and FMA – methanol dimers, which

have been extensively studied. These comparisons are discussed in more detail later.

(b) The most stable dimers A and B are those where both carbonyl groups of FMA and

FA are involved in the stabilization of the complex, together with the hydroxyl hydrogen

atom of FA that interacts with the carbonyl oxygen atom of FMA (interaction (2)).

(c) In the less stable complexes F – I the hydroxyl hydrogen atoms of FA are not

involved in hydrogen bonds.

66

According to the calculated geometries and binding energies of all the FA – FMA

dimers, it is possible to make some preliminary qualitative conclusions about the strength

of the different interactions:

(a) C=OFA ····H-NFMA (interaction 1) › C=OFA ···H-CFMA (interaction 3)

(b) C=OFMA ····H-OFA (interaction 2) › C=OFMA ···H-CFA (interaction 5)

(c) NHFMA…O=CFA(interaction 1) › NHFMA… (H)OCFA (interaction 4)

(d) OHFA…O=CFMA(interaction 2) › OHFA… NHFMA(interaction 6)

(e) CHFMA…O=CFA(interaction 3) › CHFMA… (H)OCFA(interaction 7)

(f) The CH group in the formic acid molecule only interacts with the O=C group of the

formamide molecule (interaction 5)

This allows for the qualitative comparison of the hydrogen-bond acceptors and donors

in the FA – FMA dimers:

(a) Donors: OH › NH › CH FMA› CHFA

(b) Acceptors: C=OFMA› C=OFA

The order of the proton donor ability of the hydrogen atoms linked to C, N and O

heteroatoms corresponds to the increase of the electronegativity from carbon to oxygen.

To compare the hydrogen bond acceptor capability of the carbonyl group of FMA with

that of the carbonyl group of FA is more complicated. In this case, the order is based on

the relative binding energies and distances in the complexes.

It is interesting to mention that for the FMA – water and FMA – methanol dimers the

binding energy to the carbonyl group of FMA is slightly more favorable than to the

amide group.[136, 137, 143] In the most stable FA – FMA dimer A both interactions are

present, and the distance between the carbonyl oxygen atom of FMA and the hydroxyl

hydrogen atom of FA is nearly 0.2 Å shorter than the hydrogen bond distance between

the amid hydrogen atom of FMA and the carbonyl oxygen atom of FA (Figure 3.1).

67

In agreement with Neuheuser’s observations[140], the weakest C-H…O interaction still

contributes significantly to the interaction energy in the FA – FMA system, for example

in dimer B.

Comparison with other dimers

The presence of carbonyl groups in both FMA and FA results in additional

stabilizations which do not exist in the water – formamide (W – FMA) and the methanol

– formamide (M – FMA) complexes. Nevertheless, there are very interesting analogies

among all the FMA complexes with water, methanol and formic acid.[136, 137, 143] (Figure

3.2).

Figure 3.2. B3LYP structures of FMA-Water, FMA-Methanol and selected FMA-

FA dimers.

68

Three stable W – FMA structures were described by Fu et al.[136] using DFT and MP2

methods with large basis sets. In all cases the main interaction is OHw…O=CFMA. FW I

and FW II are the two more stable W – FMA calculated complexes. They are cyclic

dimers with additional NHFMA…O-HW and CHFMA…O-HW interactions, respectively

(Figure 3.2). Their geometries have some similarities with the structures of some of the

FA – FMA dimers.

In the FA – FMA dimer A the carbonyl group of the formamide molecule interacts with

the hydroxyl hydrogen atom of the formic acid resembling the interaction between the

carbonyl group of the formamide and the hydroxyl hydrogen atom of water in the FW I

complex. The amide hydrogen atom of FMA interacts with the carbonyl oxygen atom of

FA in a similar way as the NHFMA…O-HW interaction in the FW I formamide – water

complex. The water – formamide dimer FW II shows the C=OFMA…HOw interaction

(similar to the C=OFMA…HOFA in complexes A and B) and the CHFMA…O-HW (similar to

the CHFMA …O=CFA interaction in complex B).

Four stable formamide – methanol (M – FMA) dimers have been studied by Fu et al.

using DFT and ab initio methods with various basis sets.[137] The two most stable M –

FMA complexes have similar geometries compared to the FA – FMA and W – FMA

dimers.

MF I is a cyclic dimer with OHM…O=CFMA and NHFMA…O-HM interactions. The MF

II dimer shows the OHM…O=CFMA and CHFMA…O-HM interactions. Both structures are

similar to the A and B FA – FMA dimers. MF IV compares very well with the FA –

FMA dimer F. They are both cyclic dimers stabilized by the NHFMA…(H)OCFA(MET)

interaction between the amide hydrogen atom of FMA and the hydroxyl oxygen atom of

the FA or methanol molecules. The second interaction is the C=OFMA…HCFA in the case

of complex F. The MF IV dimer shows the C=OFMA…HCM interaction between the

carbonyl oxygen atom of FMA and one hydrogen atom from the methyl group of the

methanol molecule.

FA – FMA dimers have been studied before using ab initio and DFT methods.

Neuheuser calculated five non cyclic and ring H-bonded FA – FMA structures as a model

for the interactions in supramolecular complexes of dicarboxylic acids and

69

dimethylformamide.[140] Pacios performed ab initio calculations of the most stable FMA

– FA dimer A.[144] Galvez and coworkers studied the variation of electron density

properties with the intermolecular distance for various cyclic dimers, including the most

stable FA – FMA complex.[115] Neuheuser, Pacios and Galvez studies corroborate our FA

– FMA dimers A and B as the most stable calculated geometries in the formic acid –

formamide system. This agreement confirms the reliability of the MMH procedure for

localizing the minima in non covalent complexes.

Methods and basis set influence on the calculated geometries and binding energies

of the FMA – FA dimers

Table 3.2 lists some selected intra- and intermolecular distances and hydrogen bond

angles at various levels of theory for selected FA – FMA dimers. Complex A is discussed

because it is the most stable calculated FA – FMA dimer. The weaker complex D has

been selected due to its very weak C=OFMA···H-CFA interaction. In addition,

intermolecular distances for complex B are presented.

The O-HFA bond lengths in dimer A are not sensitive to the method or the basis set used

for the calculations (Table 3.2). However, the N-HFMA intramolecular distances of the

interacting amide hydrogen atoms vary in both complexes A and D substantially with the

basis set. At the MP2 level of theory with the cc-pVDZ basis set, the calculated N-HFMA

bond lengths are 0.007 Å larger than with the cc-pVTZ basis set, whereas inclusion of

diffuse functions has only a minor influence. The B3LYP/cc-pVTZ calculated bond

lengths are only 0.002-0.003 Å larger than the MP2/cc-pVTZ values.

The C-HFA bond lengths in complex D behave in a similar way. In this case the

difference between the MP2 double and triple zeta basis set is even more pronounced

(0.0103-0.009 Å). The MP2/cc-pVTZ and aug-cc-pVTZ C-HFA distances are basically

the same and very similar to the B3LYP/cc-pVTZ values. The MP2/cc-pVDZ C-HFA

bond length is 0.003 Å larger than the MP2/aug-cc-pVDZ calculated value. The C=O

carbonyl bond lengths of the FMA and FA molecules in dimers A and D show a little

more dependence on the basis sets. At the MP2 level of theory the C=O distances

increase with the addition of diffuse functions (aug) in the double and triple zeta basis

70

sets. This variation is less pronounced with the triple zeta basis set. The B3LYP/cc-pVTZ

carbonyl bond lengths are 0.006- 0.003 Å shorter than the MP2/cc-pVTZ values.

TABLE 3.2: Comparison of selected intramolecular and intermolecular parameters

in the FA – FMA dimers A, B and D at the different levels of theory. Distances are

in Å and angles in degrees

B3LYP MP2 cc-pVTZ cc-pVDZ aug-cc-pVDZ cc-pVTZ aug-cc-pVTZ Monomer O-HFA 0.970 0.975 0.975 0.970 0.971 C-HFA 1.097 1.108 1.103 1.092 1.092 N-HFMA

a 1.006 1.014 1.012 1.004 1.006 C=OFA 1.197 1.209 1.215 1.203 1.205 C=OFMA 1.209 1.220 1.228 1.215 1.218 Dimer A O-HFA 1.006 1.005 1.006 1.004 1.005 N-HFMA

a 1.022 1.027 1.027 1.019 1.020 C=OFA 1.214 1.225 1.231 1.219 1.221 C=OFMA 1.229 1.237 1.245 1.232 1.235

NHFMA…O=CFA 1.879 1.872 1.871 1.853 1.859 C=OFMA…HOFA 1.643 1.661 1.657 1.634 1.637

< NHFMA…OFA 164.77 164.65 165.35 165.48 164.90 < OHFA…OFMA 175.70 173.95 173.98 173.82 174.51 Dimer D C-HFA 1.095 1.104 1.101 1.091 1.092 N-HFMA

a 1.016 1.022 1.021 1.014 1.015 C=OFA 1.207 1.219 1.225 1.213 1.214 C=OFMA 1.218 1.227 1.236 1.222 1.225 NHFMA…O=CFA 2.006 1.992 1.981 1.968 1.968

C=OFMA…HC(O)FA 2.295 2.258 2.267 2.251 2.257 < NHFMA…OFA 163.31 164.40 164.52 164.64 163.99 < CHFA…OFMA 138.65 140.70 139.51 140.57 138.88 Dimer B O-HFA 1.001 1.001 1.002 0.998 1.000 C=OFA 1.209 1.220 1.226 1.214 1.216

71

C=OFMA 1.225 1.235 1.242 1.229 1.231 C=OFMA…HOFA 1.676 1.683 1.682 1.663 1.663

(O)CHFMA ...O=CFA 2.350 2.308 2.317 2.303 2.304 aNH hydrogen atom in cis position relative to the carbonyl oxygen atom of the

formamide.

The NHFMA…O=CFA distances in both A and D dimers are 0.026 and 0.038 Å,

respectively, larger at the B3LYP level of theory compared to the MP2 calculations with

the same basis sets. In the MP2 calculations the NHFMA…O=CFA binding distances

decrease from the double to the triple zeta basis sets. The behavior of the

C=OFMA…HOFA (dimer A) and the C=OFMA ···H-CFA (dimer D) distances is very similar,

in general. The calculated hydrogen bond angles are comparable in all cases.

Intermolecular distances in complex B are basically not dependent on the augmentation

of the basis sets.

The B3LYP calculations show a tendency to give a little larger values for the

intermolecular binding distances, compared to the MP2 values with the same basis set.

However, there is no considerable difference between the B3LYP and the MP2 calculated

geometries for the FA – FMA dimers. Reliable geometries for the weak interacting FA –

FMA dimers are also calculated using the B3LYP density functional. At the MP2 level of

theory, there is basically no change of the geometries when the basis set is augmented by

adding diffuse functions.

In the FA – FMA dimers, the MP2/cc-pVDZ calculations have a tendency to

overestimate the binding energies (Table 3.1). The double zeta energies compare better to

the CCSD(T)/cc-pVTZ calculations when the augmented functions are added. At the

MP2 level, using triple zeta basis sets augmented and non-augmented, the results are very

similar to those of the CCSD(T)/cc-pVTZ single point calculations. The B3LYP/cc-

pVTZ binding energies are smaller than the MP2 and CCSD(T) energies.

The MP2 level of theory with the cc-pVTZ basis set provides a very adequate

description of the FA – FMA system. It is in general not necessary to use the expensive

aug-cc-pVTZ basis set, in agreement with the results of others studies of weakly

interacting complexes.[116, 145, 146]

72

Effect of the BSSE on the calculated geometries and binding energies

BSSE corrections have been calculated for all FA – FMA dimers at the MP2 level of

theory with the cc-pVDZ, aug-cc-pVDZ and aug-cc-pVTZ basis sets. As expected, the

BSSE decreases with increasing size of the basis sets. That can be noticed by comparing

the ΔE (binding energies without corrections) with ΔE (BSSE) (BSSE corrected binding

energies) in Table 3.1. For example, in complex A with MP2/cc-pVDZ the BSSE

correction is 6.77 kcal/mol, compared to only 2.32 and 1.39 kcal/mol at the MP2/ aug-cc-

pVDZ and MP2/ aug-cc-pVTZ levels of theory, respectively.

In addition, FA – FMA dimers A and B were optimized at the MP2/6-31G(d,p) level of

theory using the counterpoise (CP) scheme to evaluate the influence of BSSE on the

calculated energies and geometries.

The intramolecular bond distances in the FA and FMA molecules are almost not

effected by the inclusion of BSSE corrections during the optimization processes (Figure

3.3). In all cases the difference between the bond distances was in the order of 10-3 Å or

less. Only the intermolecular distances C=OFA ····H-NFMA (interaction 1), C=OFMA ····H-OFA

(interaction 2), and C=OFA ···H-CFMA (interaction 3) are significantly influenced by BSSE.

Especially for the weaker interaction 3, the BSSE-optimized distance C=OFA ···H-CFMA is

almost 0.12 Å larger than the non-BSSE optimized distance. Hydrogen bond angles are

less sensitive to BSSE corrections (Figure 3.3). Thus, the geometrical changes introduced

by BSSE corrections are very limited, and the basic geometries and interactions in FA –

FMA complexes do not depend on the inclusion of BSSE during the optimization

process, in accordance with other observations.[146]

73

Figure 3.3. MP2/6-31G(d,p) geometries with inter and intramolecular lengths(Å)

and hydrogen bond angles(degree) of dimers A and B. 1) Optimized without BSSE

corrections. 2) Optimized with BSSE corrections.

It is therefore not surprising that the binding energies of dimers A and B are almost

independent of BSSE corrections during geometry optimization. For complexes A and B

the calculated BSSE corrections are 4 – 5 kcal/mol, and the differences in binding

energies between the BSSE-optimized and the non-BSSE-optimized geometries are in the

range of 0.19 - 0.57 kcal/mol only (Table 3.3).

TABLE 3.3: Comparison of the binding energies of the calculated FA–FMA dimers

A and B at the MP2/6-31G(d,p) level of theory including BSSE corrections in the

optimization processes

MP2/6-31G(d,p) Optimization with BSSE Optimization without BSSE ΔE ΔE (BSSE) ΔE ΔE (BSSE) Dimer A -18.03 -13.39 -18.22 -13.19 Dimer B -14.88 -10.23 -14.31 -10.08

Intramolecular distances and vibrational frequencies. Calculated spectra

The vibrational frequencies of all the FA – FMA dimers have been calculated at the

B3LYP/cc-pVTZ, MP2/cc-pVDZ and MP2/aug-cc-pVDZ levels of theory. The

74

B3LYP/cc-pVTZ vibrational frequencies and selected intermolecular distances for

complexes A and B are discussed here. Based on experimental and B3LYP/cc-pVTZ

calculated vibrational frequencies of the monomers, the frequency shifts and correction

factors for some molecular vibrations of complexes A and B are estimated (Table 3.4) to

accurately match calculated with experimental frequencies (Table 3.5).

TABLE 3.4: Comparison between the experimental (Ar matrix, 10K) and the

calculated B3LYP vibrational frequencies (in cm–1) of formic acid and formamide

monomers, shift and factor of correction

Experimental Computed frequencies B3LYP/ cc-pVTZ

Shifta and factor of correction (exp/B3LYP freq)

3549.9 3722.1 172.2 (0.954) νOHb

3066.0 3043.8 -22.2 (1.007) νCHb

1766.9 1826.2 59.3 (0.967) νC=Ob

1103.5 1125.0 21.5 (0.981) νCOb

1739.1 1803.8 64.7 (0.964) νC=Oc

2882.9 2931.0 48.9 (0.984) νCHc

3547.4 3718.2 170.8 (0.954) νasNH2c

3426.6 3579.9 153.3 (0.957) νsNH2c

aShifts are calculated as the difference between the computed and the experimental

frequencies. bFormic acid fundamental modes. cFormamide fundamental modes.

Compared to the monomers, intramolecular distances and the corresponding vibrational

frequencies in the complexes are perturbed as a consequence of the intermolecular

interactions (Tables 3.8 and 3.11). In complexes A and Β the O-H stretching vibrations of

the FA molecule show the largest red shifts with -677 and -585 cm-1, respectively (Tables

3.8 and 3.11). This demonstrates the strong interaction between the OH hydrogen atom of

FA and the carbonyl oxygen atom of FMA (Figure 3.1) resulting in an elongation of the

OH bonds of 0.036 and 0.031 Å, respectively, for complexes A and B (B3LYP/cc-pVTZ,

Table 3.2).

75

TABLE 3.5: Calculated B3LYP/cc-pVTZ vibrational frequencies (in cm–1) of

dimers A and B and frequency shift in the complex, from the isolated monomer (in

parentheses). Predicted frequencies after scaling.

Monomer Dimer A Dimer B B3LYP/cc-pVTZ Calculated Predicteda Calculated Predicteda

3722.1

2983.7 3012.4 3044.8 (-677.3)

2904.7

3136.7 (-585.4)

2992.4

νOHb, c

1826.2 1779.8 (-46.4) 1721.1 1788.5 (-37.7) 1729.5 νC=Ob

1125.0 1256.5 (+131.5) 1232.6 1230.6 (+105.6) 1207.2 νCOb

1803.8 1729.9 (-73.9) 1667.6 1732.3 (-71.5) 1669.9 νC=Od

3718.2 3675.5 (-42.7) 3506.4 3715.8 (-2.4) 3544.9 νasNH2d

3579.9 3348.5 (-231.4) 3204.5 3579.4 (-0.5) 3425.5 νsNH2d

aPredicted frequencies after scaling the individual frequencies with a scaling factor

obtained by comparing calculated vs experimental frequencies of the corresponding

monomer bands (Table 3.4). bformic acid. cIn the case of complex A, there is a very

strong coupling between the νOH and the νCH vibrations of formic acid and formamide.

The same happen for the νC=O vibrations of formic acid and formamide. dFormamide.

The carbonyl stretching frequencies of the FA molecules in dimers A and B are

calculated to be shifted by -46 and -38 cm-1. The C=OFA bond lengths in A and B

increase by 0.017 and 0.012 Å compared to the monomers. The red shift for dimer B is 8

cm-1 less than for dimer A. This difference is caused by the stronger C=OFA ····H-NFMA

interaction (interaction 1) in dimer A compared to the weaker C=OFA ···H-CFMA interaction

(interaction 3) in dimer B.

The carbonyl stretching frequency shifts of the FMA molecules are -74 and -71 cm-1.

The C=OFMA bond lengths in the dimers A and B increases by 0.020 y 0.016 Å,

respectively. According to the structure of the complex, only in dimer A a significant

shift (-231 and -43 cm-1) respectively, for the symmetrical and anti-symmetrical

vibrations of the N-H group, is predicted. The intramolecular N-H bond distance of the

interacting NH group of FMA in complex A is consequently 0.016 Å larger than in the

FMA monomer.

76

Larger systems

The intermolecular interactions between FMA and FA create a very flat intermolecular

energy surface. That makes the analysis of systems larger than dimers even more

complicated. It would take huge computational efforts to get a complete description of

the possible geometries for trimers and larger aggregates. However, there are very

interesting correlations between the geometries of the FA – FMA dimers and a selection

of structures of larger FA – FMA systems preliminarily calculated at the DFT level of

theory.

1:2 Formic acid – formamide complexes

Figure 3.4 shows a selection of the most stable calculated 1:2 FA – FMA complexes T-

A to T-G and their B3LYP/cc-pVTZ binding energies with and without ZPE corrections.

It is important to remark, once again, that all the trimer structures are found starting from

a large amount of randomly generated geometries calculated with semiempirical

Hamiltonians and later refined at the B3LYP/cc-pVTZ level of theory.

T-A is the most stable calculated trimer with a binding energy of -22.22 kcal/mol. T-B

is energetically very close to T-A with -21.94 kcal/mol (Figure 3.4). It is interesting to

compare the T-A and T-B geometries with the structure of the dimers. The part of trimer

T-A where the FA and FMA molecules interact to each other, is similar to the FA – FMA

dimer B (Figure 3.1). But due to the presence of a second interacting formamide

molecule, the intermolecular C=OFA···H-CFMA and C=OFMA····H-OFA distances are 0.031

and 0.008 Å larger, respectively, compared to dimer B at the same level of theory (Table

3.2, Figure 3.4). The FMA – FMA interactions in trimer T-A reproduce the structure of

the most stable formamide homodimer.

In trimer T-B, the FA interactions with FMA disturb the structure of the FA – FMA

dimer A. The carbonyl oxygen atom of FA shows an additional interaction with one

amide hydrogen atom of the second FMA molecule. This causes an elongation of 0.102 Å

of the C=OFA ····H-NFMA distance compared to dimer A. The C=OFMA ····H-OFA distance in

trimer T-B is also 0.069 Å larger in comparison to dimer A (Table 3.2, Figure 3.4).

77

Figure 3.4. The calculated structures with hydrogen bond lengths (Å) of the 1:2

FA– FMA complexes T-A to T-G at the B3LYP/ cc-pVTZ level of theory. a

B3LYP/cc-pVTZ binding energies. b B3LYP/cc-pVTZ binding energies, ZPE

corrected.

The trimers T-C and T-D are very close energetically to each other with binding

energies of -21.35 and -21.02 kcal/mol, respectively. Again, the main interactions

between FA and FMA in T-C resemble the FA – FMA dimer B, but the intermolecular

78

distances are shorter compared to the dimer (Figure 3.4, Table 3.2). In this case, one

amide hydrogen atom of the second FMA molecule shows an additional interaction with

the carbonyl oxygen atom of the FA molecule.

T-E, T-F and T-G have calculated binding energies of -18.96, -18.59 and -18.25

kcal/mol, respectively. But in the case of complex T-G there is one imaginary out of

plane vibration at -15cm-1 that is related with the repulsive interaction at 2.306 Å between

the aldehyde hydrogen atoms of the two FMA (Figure 3.4). The geometry of the T-G

complex at the B3LYP/6-31++G (d,p) level of theory is very similar, but there is no

imaginary vibration and the distance between the aldehyde hydrogen atoms of the two

FMA molecules is 2.314 Å.

It is interesting to notice that in the trimers T-A, T-C, T-E, and T-G the interactions

between the FA and one FMA molecule resemble the geometry of the FA – FMA dimer

B. In the same way, interactions in trimers T-B, T-D and T-F resemble the structure of

dimer A. Complex T-F is the only with no direct interactions between the two FMA

molecules. Instead, the interactions between FA and the second molecule of FMA

resemble the structure of the FA – FMA dimer F (Figure 3.1). In the T-A, T-D and T-E

trimers the FA molecule interacts with only one molecule of FMA and the system is

additionally stabilized by the FMA-FMA attractions.

Compared to T-A, the trimers T-H and T-I are much less stable with binding energies

of -14.78 and -16.26 kcal/mol, respectively (Figures 3.4 and 3.5). However, they are

considered here in order to compare the geometry of T-H with the crystal structure and,

in both cases, to analyze the pairs of intermolecular interactions in the trimers.

The T-H complex is less stable compared to T-A, since the carbonyl group and the

hydroxyl hydrogen atom of the FA molecule are not directly interacting with the FMA

molecules. The stabilizing FA – FMA interactions in T-H are the same than in the FA –

FMA dimer F, the C=OFMA ···H-CFA (interaction 5) and the NHFMA…(H)OCFA (interaction

4), however in this case the FA interacts with two molecules of FMA, forming the

structure of the most stable cyclic FMA homodimer.

79

Figure 3.5. The calculated structures with hydrogen bond lengths (Å) of the 1:2 FA–

FMA complexes T-H and T-I at the B3LYP/ cc-pVTZ level of theory. aB3LYP/cc-

pVTZ binding energies. bB3LYP/cc-pVTZ binding energies, ZPE corrected.

The trimer T-I stabilizes by the NHFMA…O=CFA (interaction 1) and C=OFMA…HOFA

(interaction 2) interactions; but unlike other complexes, both interactions take place with

different molecules of FMA. This fact makes this system a very interesting case to

analyze the intermolecular interactions in the trimers. In addition, Trimer T-I is stabilized

by FMA-FMA attractions that resemble the geometry of a FMA homodimer

Analysis of the intermolecular interactions in the trimers

To quantify the contributions of intermolecular interactions in the trimers T-A, T-D, T-

F, T-H and T-I; in each of the trimers one of the three monomers (formic acid or one of

the two FMA molecules) is removed subsequently. The energies of the remaining partial

structures (remaining dimers) were calculated (B3LYP/cc-pVTZ) in the geometries of the

parent trimers (Figure 3.6). These partial structures are then compared with the optimized

dimers to analyze the influence of the third molecule in the trimer on the dimer structures.

From that, a detailed picture of the interactions in an aggregate consisting of three

components is achieved.

80

Figure 3.6. Dimers of complexes T-H and T-I. Partial structure (i): Formic acid and

FMA 1. Partial structure (ii): Formic acid and FMA 2. Partial structure (iii): FMA 1

and FMA 2.

Partial structure (i) is formed by the FA and the FMA 1 molecules; partial structure (ii)

by FA and FMA 2, and partial structure (iii) by FMA 1 and FMA 2 (Figures 3.4 and 3.6).

From trimer T-A the first partial structure is formed by removing FMA 2, and thus this

partial structure consists of FA and the remaining FMA 1 (partial structure (i)) stabilizing

via interactions (2) and (3)). Analogously, partial structure (ii) is formed by removing

FMA 1 and consists of the non-interacting FA and FMA 2. Finally, partial structure (iii)

results from removing the FA molecule and represents the most stable FMA – FMA

homodimer A stabilized via two C=OFMA…H-NFMA interactions. The partial structures (i)

– (iii) in the trimers T-D, T-F, T-H and T-I are formed analogously by subsequently

removing FMA 2, FMA 1, and the formic acid molecule (Figure 3.6).

81

In order to compare with the partial structures (iii) from the trimers, the structures of

the FMA – FMA dimers A and B were also optimized at the B3LYP/cc-pVTZ level of

theory (Figure 3.7, Table 3.6). Opposed to FMA – FMA dimer A, in the partial structures

(iii) of trimers T-A and T-H, the N-HFMA…O=CFA hydrogen bonds are not equivalent,

while one N-HFMA…O=CFA hydrogen bond is elongated; the other is shorter compared to

that in FMA – FMA dimer A. That is clearly due to effect of the interactions with the FA

molecule. Therefore, in partial structure (iii) of trimer T-A, the N-HFMA…O=CFA distance

which is more affected (0.069 Å larger) compared to FMA – FMA dimer A, is the one

corresponding to the C=OFMA group which interact in addition with the FA molecule

(Figures 3.4 and 3.7).

TABLE 3.6: Calculated binding energies (in kcal/mol) of some FMA - FMA

dimers and selected partial structures (iii) of the trimers

B3LYP/cc-pVTZ FMA – FMA dimer A -14.25 FMA – FMA partial structure (iii), Trimer T-A -13.73 FMA – FMA partial structure (iii), Trimer T-H -14.19 FMA – FMA dimer B -9.40 FMA – FMA partial structure (iii), Trimer T-D -8.64 FMA – FMA partial structure (iii), Trimer T-I -8.67

In the partial structure (iii) of trimer T-H, one N-HFMA…O=CFMA hydrogen bond

distance is larger than in the FMA – FMA dimer A. This is the one corresponding to the

side of the FMA – FMA complex that interacts with the FA molecule in the trimer. This

elongation of the hydrogen bond can be attributed not only to the attractive interaction

between one NH hydrogen atom of FMA and the OH oxygen atom of FA at 2.669 Å, but

also to the repulsion at 2.763 Å between the other amidic hydrogen atom of FMA and the

aldehyde-type hydrogen atom of FA (Figures 3.5 and 3.7).

82

Figure 3.7. Comparison between FMA – FMA dimers A and B and the partial

structures (iii) of the 1:2 FA – FMA trimers T-A, T-D, T-H, and T-I.

The partial structures (iii) of trimer T-D and T-I are similar in geometry and close in

energy to the FMA – FMA dimer B (Figure 3.7, Table 3.6). The NHFMA2…O=CFMA1

distances in partial structures (iii) are, compared to the FMA – FMA dimer B, more

affected than the C=OFMA2…H-CFMA1 distances. That is expected, since according to the

geometries of trimers T-D and T-I the carbonyl oxygen atom of FMA 1 interacts also

with the FA molecule, unlike the carbonyl group of FMA 2 (Figures 3.4, 3.5 and 3.7). In

addition, compared to FMA – FMA dimer B, the NHFMA2…O=CFMA1 distance is larger in

trimer T-I than in trimer T-D, which is explained by the interaction of one NH hydrogen

atom of FMA 2 with the carbonyl oxygen atom of FA in trimer T-I (Figure 3.5).

The partial structures (i) in trimers T-D and T-F resemble the FA – FMA dimer A with

very similar hydrogen bond distances and angles, as well as binding energies (Figure 3.8

and Table 3.7). In the same way, the partial structure (i) of trimer T-A is similar to the FA

– FMA dimer B and partial structure (ii) of trimer T-F corresponds to the FA – FMA

dimer F.

83

TABLE 3.7: Calculated binding energies (in kcal/mol) of some FMA - FA dimers

and selected partial structures (i) and (ii) of the trimers

B3LYP/cc-pVTZ FA – FMA dimer A -16.03 FA – FMA partial structure (i), Trimer T-D -15.79 FA – FMA partial structure (i), Trimer T-F -15.80 FA – FMA dimer B -12.32 FA – FMA partial structure (i), Trimer T-A -11.68 FA – FMA dimer F -4.88 FA – FMA partial structure (ii), Trimer T-F -2.88

Figure 3.8. Comparison between FA – FMA dimers A, B and F and selected partial

structures (i) y (ii) of the 1:2 FA – FMA trimers T-A, T-D, and T-F.

84

Several non-additive contributions are considered, including basis set superposition

errors and other non-conventional interactions that may contribute to the stabilization of

the trimers (Tables 3.8 and 3.9). These non-additive contributions are obtained by

subtracting the binding energies of all contributions from the trimer binding energy. As it

can be seen from the Tables 3.8 and 3.9 the non-additive contributions represent between

9.6 – 12.6 % of the total binding energy.

TABLE 3.8: B3LYP/cc-pVTZ energies of the trimers T-A, T-D, T-F, and their

partial structures (i), (ii), and (iii) (in kcal/mol). The percents of the energies of each

partial structure, compared to the total energy of the trimers, are in parenthesis.

B3LYP/cc-pVTZ

T-A T-D T-F Trimer, E(t) -25.92 -24.59 -21.75 Partial structure (i), E(i) -11.68 (45.1%) -15.79 (64.2%) -15.80 (72.6%) Partial structure (ii), E(ii) +2.69 (10.4%) +2.93 (11.9%) -2.88 (13.2%) Partial structure (iii), E(iii) -13.73 (53.0%) -8.64 (35.1%) -0.52 (2.4%) E(t) - (E(i)+ E(ii)+ E(iii)) -3.20 (12.3%) -3.09 (12.6%) -2.55 (11.7%)

TABLE 3.9: B3LYP/cc-pVTZ energies of the trimers T-H, T-I and their partial

structures (i), (ii), and (iii) (in kcal/mol)

B3LYP/cc-pVTZ

T-H T-I Trimer, E(t) -17.72 -19.22 Partial structure (i), E(i) +0.15 (0.8%) -6.67 (34.7%) Partial structure (ii), E(ii) -1.97 (11.1%) -1.59 (8.3%) Partial structure (iii), E(iii) -14.19 (80.1%) -8.76 (45.6%) E(t) - (E(i)+ E(ii)+ E(iii)) -1.71 (9.6%) -2.20 (11.4%)

In trimers T-D and T-F, partial structure (i) with 64.2 and 72.6 %, respectively,

contributes most to the trimer energies. The C=OFA…HNFMA and O-HFA…O=CFMA

interactions between the FA and FMA 1 dominate the interaction in the trimer, especially

in T-F, where the two FMA molecules basically do no interact to each other and the FA –

85

FMA 2 interactions are weaker, since neither the carbonyl oxygen atom nor the OH

hydrogen atom of FA are involved. In trimer T-D the FA – FMA 2 interactions are

repulsive, but the FMA 1 – FMA 2 interactions contribute to the stabilization of the

complex by 35.1 % (Table 3.8).

In trimer T-A, partial structure (iii) contributes 53 % to the total binding energy,

whereas the binding energy of partial structure (i) is 45.1 % and the FA – FMA 2

interactions are repulsive. This is expected, in accordance with the interactions involved

in the stabilization of partial structures (i) and (iii).

Trimers T-H and T-I are interesting complexes, since, unlike T-A, T-D and T-F, in both

cases the FA molecule interacts directly with the FMA 1 and FMA 2 molecules; and, in

each case the two FMA molecules provide an additional stabilization to the trimer by

interactions that resemble the structures of very stable FMA homodimers. Subsequently,

the energies of partial structures (iii) represent the largest contribution to the total binding

energies with 80.1 % for trimer T-H and 45.6 % in T-I. Partial structure (i) in T-I

contributes 34.7 % to the total binding energy (Table 3.9).

In trimers T-I and T-D the FMA – FMA stabilization is the same for both cases,

resembling the FMA – FMA dimer B. In addition, the interactions between the FA and

FMA molecules are of the same type ( (1) and (2) ) but, while in T-D they are both

present in the interactions of FA with FMA 1, in trimer T-I, interactions (1) and (2) occur

between the FA with the FMA 1 and FMA 2 molecules (Figures 3.4, 3.5 and 3.7).

Therefore it is interesting to point out the difference of about 20 % between the

contributions to the total energies in both cases. In T-D partial structure (i), which

contains both interactions (1) and (2), represent 64.2 % of the total binding energy; while

in T-I the sum of partial structure (i) (interaction (2)) and partial structure (ii) (interaction

(1)) represent 43 % of the total binding energy. Therefore, partial structure (iii) is the

dominating contribution to the stabilization of the complex T-I, unlike in T-D, where

partial structure (i) clearly dominates.

In T-H, the FMA – FMA interactions, which resemble the structure of the FMA

homodimer A, with 80.1 %, are contributing most to the stabilization of the complex. The

energy of the partial structure (ii) represents the 11.1 % of the total energy, whereas the

86

partial structure (i) is slightly repulsive, probably due to the repulsion at 2.763 Å between

one NH hydrogen atom of FMA 1 and the aldehyde hydrogen atom of the FA molecule

(Table 3.9, Figures 3.5 and 3.7).

1:4 Formic acid – formamide complexes

1:4 FA – FMA complexes have been calculated starting from 198 arbitrary geometries

that were optimized at the semiempirical level. A selection of complexes was refined at

the B3LYP/cc-pVTZ level of theory (Figure 3.9).

P-A is the most stable of the calculated pentamers with a binding energy of -50.54

kcal/mol. Complexes P-B and P-C are energetically close with binding energies of -46.42

and -43.95 kcal/mol, respectively. The binding energies of P-D and P-E are very similar,

with -38.64 and -38.47 kcal/mol, respectively.

The structure of the P-A complex is very interesting. Two pairs of FMA molecules are

forming two FMA cyclic homodimers which are then interacting to each other. The FA

molecule stabilizes the complex with the same type of interaction than in the FA – FMA

dimer A (Figure 3.1). The difference is that in P-A the FA molecule interacts with the

two closest FMA molecules, and the FA carbonyl oxygen atom cooperates with two NH

hydrogen atoms. In all the other complexes (P-B to P-E) the FA molecule interacts with

one FMA molecule (FMA-a) with the C=OFMA ····H-OFA and C=OFA ···H-CFMA interactions,

forming the FA – FMA dimer B (Figure 3.1).

By comparison with the 1:2 FA – FMA complexes it is easy to identify the structure of

the T-C trimer as part of the P-C, P-D and P-E complexes. The P-C pentamer is even

more interesting, since it combines the geometries of both, the T-A and the T-C trimers

(Figure 3.9). Considering the FA molecule, the FMA-a and the FMA molecule at the

right side of FMA-a, we get the geometry of trimer T-A. On a similar way, looking at the

interactions between FA, FMA-a and the FMA molecule at the left side of FMA-a, the

geometry of this subunit is very similar to trimer T-C.

87

Figure 3.9. The calculated structures with hydrogen bond lengths (Å) of the 1:4 FA–

FMA complexes P-A to P-E at the B3LYP/ cc-pVTZ level of theory. aB3LYP/cc-

pVTZ binding energies (kcal/mol) .

88

Comparison of FMA – FA complexes with the crystal structure

The complexity of the FA – FMA crystal structure can not be completely described by

a small set of FA – FMA complexes. However, interesting structural similarities are

noticed. Three sections of the FA – FMA crystal structure are presented in Figure 3.10,

whereas Figure 3.11 shows a large fragment of the FA – FMA crystal structure.[139]

Figure 3.10. Selected sections of the FMA-FA crystal structure.[139]

89

Figure 3.11. FMA-FA crystal structure. Ref [139]

The same type of interactions (1 – 4) that have been discussed above for the FA – FMA

dimers are present in the crystal structure. The geometry of dimer B is clearly reproduced

in the FA – FMA crystal interactions of fragments FA and FMA2 (Figure 3.1, Figure

3.10). In both cases the carbonyl group of the FA interacts preferentially out of plane

with another molecule.

The geometry of the trimer T-A is also very similar to the marked selection in the

FMA2 section (Figure 3.4). Trimer T-H describes the geometry of the interactions

between the FA molecule and the two FMA molecules forming the FMA cyclic

homodimer in section FA (Figure 3.10).

90

Table 3.10 lists some selected intermolecular distances in FA – FMA complexes and two

motifs of the crystal structure. For the dimers, the intermolecular distances shown were

calculated at the MP2/aug-cc-pVTZ level of theory and for trimers and pentamers at the

B3LYP/cc-pVTZ level. The C=OFMA…HOFA calculated distances of dimer B and trimer

T-A agree especially well with distances in the motif 1 of the crystal structure (CI), while

the calculated CHFMA…O=CFA distance for the pentamer P-B is closer to the one of CI

compared to the dimer and trimer. The value of the intermolecular distance for the

NHFMA…O=CFA interaction in the motif 2 of the crystal structure (CII) is very close to

the calculated one in trimers T-B and T-C, whereas the calculated NHFMA…(H)OCFA

distance in trimer T-H is similar to the value of this interaction in CII.

TABLE 3.10: Selected intermolecular distances in FMA-FA complexes and two

motifs of the crystal structure (Å)

C=OFMA…HOFA CHFMA…O=CFA NHFMA…O=CFA NHFMA…(H)OCFA

Dimers 1.663 (B) 2.304 (B) 1.859 (A) 2.213 (C)

Trimers 1.668 (T-A) 2.381 (T-A) 2.008 (T-B) 1.992 (T-C)

2.086 (T-F) 2.669 (T-H)

Pentamers 1.684 (P-A) 2.504 P-B

2.679 (P-A)a 1.987(P-D,P-E)a 2.234 (P-C) 2.228 (P-E)

Crystal structure CI[139]. 1.670 2.670 2.112 2.560 Crystal structure CII[139]. 1.760 3.084 2.057 2.643

aIn the crystal structure the carbonyl group of the FA interacts preferentially out of

plane with another molecule of FMA like in the pentamers.

91

3.4. Formic acid – dimethyl ether dimers. Results and discussion

Geometries and binding energies

Six FA – DME complexes A - F were found at the MP2 level of theory with the 6-

311++G(d,p) and cc-pVTZ basis sets (Figure 3.12). For all complexes the geometries are

almost independent of the basis sets used, therefore only hydrogen bond distances and

angles calculated at the MP2/cc-pVTZ level of theory are discussed. The calculated

binding energies (Table 3.11) predict complexes A and B as the lowest minima, both

being very close in energy. During the optimization process the second enantiomer of

complex B was also found which shows the reliability of the MMH procedure. According

to the calculated binding energies complexes C - F are much less stable.

Four basic types of interactions (1) – (4) can be differentiated in the FA-DME

complexes:

(1) HC(=O)OH…O(CH3)2 interaction between the hydroxyl hydrogen atom of FA

and the ether oxygen atom of DME.

(2) HOC(H)=O…HCH2OCH3 interaction between the carbonyl oxygen atom of FA

and the hydrogen atoms of DME.

(3) HOC(=O)H…O(CH3)2 interaction between the aldehyde H hydrogen atom of FA

and the ether oxygen atom of DME.

(4) O=C(H)O(H)…HCH2OCH3 interaction between the hydroxyl oxygen atom of FA

and the hydrogen atoms of DME.

92

Figure 3.12. The calculated structures with hydrogen bond lengths (Å) and some

bond angles (degree) of the FA – DME complexes A - F at the MP2/cc-pVTZ level of

theory.

93

TABLE 3.11: Calculated binding energies including ZPE and BSSE corrections (in

kcal/mol) of the FA – DME complexes A – F

MP2/6-311G++(d,p) ΔE ZPE BSSE ΔE (ZPE) ΔE (ZPE+BSSE) A -11.41 1.63 3.38 -9.78 -6.40 B -10.87 1.43 2.75 -9.44 -6.69 C -4.19 0.64 1.13 -3.55 -2.42 D -3.90 0.53 1.08 -3.37 -2.29 E -4.30 0.77 1.62 -3.53 -1.91 F -4.35 0.93 2.26 -3.42 -1.16 MP2/cc-pVTZ ΔE ZPE BSSE ΔE (ZPE) ΔE (ZPE+BSSE) A -12.23 1.50 2.76 -10.73 -7.97 B -11.31 1.28 2.22 -10.03 -7.81 C -5.03 0.73 1.43 -4.30 -2.87 D -4.47 0.59 1.22 -3.38 -2.66 E -5.26 0.82 1.78 -4.44 -2.66 F -5.01 0.91 1.84 -4.10 -2.26

Complexes A and B show the same type of interactions (1) and (2). The hydrogen atom

of the OH group of the FA interacts with the ether oxygen atom of DME at hydrogen

bond distances of 1.672 Å and 1.691 Å for the dimers A and B, respectively. The

difference between these complexes is caused by the interaction (2). In dimer A the C=O

group of FA interacts simultaneously with two hydrogen atoms of DME at distances of

2.663 Å, while in dimer B the C=O group of FA is approaching only one hydrogen atom

of the DME with a distance of 2.508 Å.

With the 6-311++G(d,p) and cc-pVTZ basis sets (without including ZPE and BSSE

corrections) the differences between the binding energies of complexes A and B (ΔEB –

ΔEA, ΔΔEBA) are 0.54 kcal/mol and 0.92 kcal/mol, respectively. After including BSSE

and ZPE corrections these differences are reduced to -0.29 kcal/mol (6-311++G(d,p)) and

0.16 kcal/mol(cc-pVTZ) (Table 3.11). With the 6-311++G(d,p) basis set including all

corrections the binding energies for dimers A and B are -6.40 and -6.69 kcal/mol,

94

respectively. This is due to the large BSSE error for complex A with the 6-311++G(d,p)

basis set which considerably lowers the binding energy of A compared to B.

Comparing the structures of the FA – DME complexes A and B with the DME –

methanol complex reported in literature[132, 133] reveals large similarities. Han and

Kim[133] calculated at the MP2/6-31+G** level of theory the hydrogen bond distance

CH3OH…O(CH3)2 to 1.855 Å and the OH...O hydrogen bond angle to 177.2°, in good

agreement with our OH…O hydrogen bond distances of 1.672 Å and 1.691 Å,

respectively, and OH...O bond angles of 177.3° and 176.8°, respectively, for complexes

A and B (MP2/cc-pVTZ).

A cyclic dimer of FA – water has been described experimentally by Astrand[113] and

Priem.[147] It is also interesting to compare the FA – DME complexes with the geometries

of the recently calculated FA – water complexes by Zhou.[148] The most stable FA – water

complex found by this author is a cyclic complex with both FA and water acting as

hydrogen donor and acceptor. The HC(O)OH…OH2 and HOC(H)=O…HOH distances

were calculated to 1.792 Å and 2.144 Å, respectively, at the MP2/6-311++G(d,p) level of

theory. It is remarkable how this cyclic FA – water complex shows the same type (i) and

(ii) interactions as the DME – FA complexes A and B.

Zhou reported two additional more weakly bound FA – water dimers. One of those is a

HOC(H)=O…HOH complex with a calculated C=O...H bond angle of 99.8° and an

O…H distance of 2.053 Å. Its geometry is very similar to the FA – DME dimer C, where

the calculated C=O…H distance is 2.666 Å and the COH hydrogen bond angle is 95.5°

(for comparison with Ref. [148] calculated at the MP2/6-311++G(d,p) level of theory). The

FA – DME complex C is stabilized by both interactions (2) and (3) and the C-H…O

distance is calculated to 2.257 Å.

The other FA – water complex found by Zhou is the OC(H)O(H)…HOH complex,

where the O…Hwater distance is 2.204 Å and the (H)O...Hwater angle is 147.6°. FA – DME

dimer D exhibits a similar geometry, the calculated O=C(H)O(H)…HCH2OCH3 distance

is 2.812 Å and the (H)O...H hydrogen bond angle is 154.5° (MP2/6-311++G(d,p)).

Complex D shows interactions (3) and (4), the O=C(OH)H…O(CH3)2 distance is 2.230

Å. Comparing the MP2/cc-pVTZ geometries of complexes C and D (Figure 3.12) with

95

their MP2/6-311++G(d,p) geometries (as discussed above) reveals that the structures of

these complexes are almost independent of the basis set used in the calculations.

Dimer E is stabilized by interactions (2) and (3) and dimer F by interactions (2), (3)

and (4). The calculated MP2/cc-pVTZ + ZPE + BSSE binding energies of complexes C –

F are very similar and around 5 kcal/mol smaller than the binding energies of complexes

A and B. This can be rationalized by the lack of the strongest interaction (1) in complexes

C – F.

Geometry optimization including BSSE

To investigate the influence of the basis set superposition errors (BSSE) on the

geometries of the complexes, the geometries of complexes A – E were optimized at the

MP2/6-31G(d,p) level of theory using the counterpoise (CP) scheme of Boys and

Bernardi during the optimization process. The small 6-31G(d,p) basis set was selected

since the BSSE is here more pronounced compared to larger basis sets and in addition the

calculations are less demanding. Complexes A – D are discussed here to illustrate

different types of dimers including strong and weak interactions and similar and different

binding energies (Tables 3.12 and 3.13).

The C=O and O-H bond distances in the FA part of the complexes and the C-O

distances in the DME part are almost not effected by the inclusion of BSSE during the

optimization. In all cases the difference in these bond distances was in the order of 10-3 Å

(Table 3.12). The OH…O, C=O…H, CH…O and C- O…H intermolecular distances

(interactions (1), (2), (3) and (4), respectively) are more influenced by BSSE. The weak

interactions (2) and (4), where DME hydrogen atoms are involved, exhibit the largest

changes. Thus, the C=O…H and C-O…H distances increase by about 0.2 – 0.3 Å when

BSSE is considered during the optimization. Despite these variations, the basic

geometries and interactions in the FA – DME complexes do not change.

96

Table 3.12. Comparison of selected intramolecular and intermolecular distances in

the FA – DME complexes A – D at the MP2/6-31G(d,p) level of theory. The results

from geometry optimizations without and including BSSE corrections are

compared.

MP2/6-31G (d,p) optimization with BSSE optimization without BSSE A B C D A B C D Intramolecular distances

r(C=O)FA 1.219 1.218 1.216 1.215 1.222 1.220 1.218 1.214 r(O─H)FA 0.989 0.986 0.972 0.972 0.995 0.991 0.972 0.972 r(C─O)DME 1.425 1.423 1.419 1.417 1.429 1.427 1.421 1.419

r(C─O)e 1.419 1.415 1.416 1.421 1.416 1.416 Intermolecular distances OH…Oa 1.801 1.804 _ _ 1.706 1.725 _ _ C=O…Hb 2.922 2.589 2.751 _ 2.596 2.457 2.526 _ CH…Oc _ _ 2.330 2.315 _ _ 2.260 2.252 C-O…Hd _ _ _ 2.880 _ _ _ 2.638

aInteraction (1). bInteraction (2). cInteraction (3). dInteraction (4) between FA and DME. eC-O bond distance between the DME oxygen atom and the non interacting methyl

group.

The calculated binding energies of A – D (Table 3.13) are also almost independent of

using BSSE corrections during geometry optimization, which clearly demonstrates that

consideration of BSSE during geometry optimization is not mandatory in this case (Table

3.13). The large BSSE in complex A lowers its binding energies below that of complex

B. However, both complexes are still very close in energy and more stabilized than the

other dimers. These results are similar to that obtained with the MP2/6-311++G(d,p)

level of theory described above.

97

Table 3.13. Calculated binding energies including BSSE corrections (in kcal/mol)

of the FA – DME complexes A – D at the MP2/6-31G(d,p) level of theory. The

results from geometry optimizations without and including BSSE corrections are

compared.

MP2/6-31G(d,p) optimization with BSSE optimization without BSSE ΔE BSSE ΔE + BSSE ΔE BSSE ΔE +BSSE

A -12.23 4.05 -8.18 -12.82 5.10 -7.72 B -11.60 3.26 -8.34 -11.92 3.86 -8.06 C -4.81 2.07 -2.74 -4.98 2.39 -2.59 D -4.21 1.72 -2.49 -4.35 1.99 -2.36

Intramolecular distances and vibrational frequencies

Compared to the monomers, intramolecular distances and vibrational frequencies in the

complexes are distorted as a consequence of the intermolecular interactions. For the

monomers the available experimental geometrical data are well reproduced at the

MP2/cc-pVTZ and MP2/6-311++G(d,p) levels of theory (Table 3.14).

The most perturbed vibrational modes in complexes A and Β are the O-H stretching

vibrations of the FA molecule (Tables 3.15 and 3.16). At the MP2/cc-pVTZ level of

theory the frequency shifts in the complexes A and B are -538 and -449 cm-1,

respectively. This reflects the strong interactions between the OH hydrogen atom and the

ether oxygen atom in complexes A and B (Figure 3.12) which results in an elongation of

the OH bonds of approximately 0.025 Å (MP2/cc-pVTZ, Table 3.14).

98

Table 3.14. Comparison of the selected intramolecular distances of FA, DME, and

the FA – DME complexes A and B.

aRef. [149] bRef [131] cC-O bond between the DME oxygen atom and the non interacting

methyl group.

Table 3.15. The experimental (Ar matrix at 35 K – 45 K) and the calculated MP2/cc-

pVTZ unscaled vibrational frequencies (in cm–1) of the FA – DME complex A, along

with the frequency shift in the complex, ∆ν, from the monomer (in parentheses).

Experimental MP2/cc-pVTZ Monomer Complex A Monomer Complex A

Formic Acid 3550.5 3000.3 (-550.2) 3763.4 3225.1 (-538.3) ν (O-H) 1767.3 1735.0 (-32.3) 1818.1 1786.7 (-31.4) ν (C=O) 1103.7 1181.9 (78.2) 1136.7 1225.6 (88.9) ν (C-O) 635.4 961.6 (326.2) 686.2 1029.7 (343.5) γ (O-H) o.o.p. 629.2 679.8 (50.6) 629.2 688.4 (59.2) δ (O-C=O)

Dimethyl ether 1098.3 1077.3 (-21.0) 1138.0 1124.4 (-13.6) νas (O-C-O) 925.9 899.9 (-26.0) 970.6 943.4 (-27.2) νs (O-C-O)

Experiment MP2/cc-pVTZ MP2/6-311++G(d,p) monomer monomer A B monomer A B Formic acid

r(C=O) 1.202(10)a 1.203 1.212 1.210 1.205 1.214 1.212 r(O─H) 0.972 (5)a 0.969 0.996 0.992 0.969 0.993 0.989 r(C─O) 1.343 (10)a 1.346 1.329 1.329 1.348 1.332 1.333

Dimethyl ether r(C─O) 1.410b 1.408 1.423 1.416[c]

1.421 1.411 1.424 1.419[c]

1.424 r(C─H) in plane

1.091b 1.086 1.086 1.084 1.090 1.090 1.089

r(C─H) 1.100b 1.095 1.091 1.092 1.099 1.096 1.096

99

The carbonyl stretching frequencies of the FA molecules in A and B are predicted to be

red-shifted by 31 and 24 cm-1, respectively, again reflecting the strong CO…H interaction

(contribution (2)). The C=O bond lengths in dimers A and B compared to the monomer

increase by nearly 0.010 Å. In complex A, where the carbonyl oxygen atom of FA

interacts with two H atoms of DME, the predicted red-shift of the carbonyl stretching

frequency is 7 cm-1 larger than dimer B with only one contact.

The C-OH stretching modes of the FA molecules in A and B are blue-shifted by

89 cm-1 and 91 cm-1, respectively, and consequently the C-OH bond lengths are 0.017 Å

shorter than in the monomer. The symmetrical and antisymmetrical O-C-O and CH3

stretching modes of the DME moieties are also perturbed in A and B as listed in Tables

3.15 and 3.16

Table 3.16. The experimental (Ar matrix at 25 K – 35 K) and the calculated MP2/cc-

pVTZ unscaled vibrational frequencies (in cm–1) of the FA – DME complex B, along

with the frequency shift in the complex, ∆ν, from the monomer (in parentheses).

Experimental MP2/cc-pVTZ Monomer Complex B Monomer Complex B

Formic Acid 3550.5 3077.8 (-472.7) 3763.4 3314.6 (-448.8) ν (O-H) 1767.3 1741.2 (-26.1) 1818.1 1793.7 (-24.4) ν (C=O) 1103.7 1179.4 (75.7) 1136.7 1228.1 (91.4) ν (C-O) 635.4 852.8 (217.4) 686.2 933.6 (247.4) γ (O-H) o.o.p. 629.2 681.5 (52.3) 629.2 693.1 (63.9) δ (O-C=O) Dimethyl ether 1098.3 1092.0 (-6.3) 1138 1133.1 (-4.9) νas (O-C-O) 925.9 915.2 (-10.7) 970.6 953.2 (-17.4) ν s (O-C-O)

Comparison with matrix isolation spectroscopy results

The IR spectra of complexes A and B were assigned by comparison of the calculations

at the MP2/cc-pVTZ level of theory (Tables 3.15 and 3.16) with matrix isolation

experiments. Characteristic for the formation of complexes of FA is the red shift of the

100

C=O stretching vibration found at 1767.3 cm-1. In the symmetrical doubly bridged dimer

of FA the C=O stretching vibration is shifted by -38.6 cm-1 and in the complex with water

by -30.4 cm-1. In the complexes A and B these shifts are -32.3 and -26.1 cm-1,

respectively, and thus in the expected range. The experimental red-shifts are in excellent

agreement with the MP2 calculations which predict shifts of -31.4 and -24.4 cm-1,

respectively, for the two complexes. The red-shifts in the C=O stretching vibration reflect

the elongation of the C=O bond due to the formation of complexes. Obviously, the

interaction of the carbonyl oxygen atom in A with two DME hydrogen atoms is more

efficient than the interaction with only one hydrogen atom in complex B.

Due to a large number of absorptions in the region of the OH/CH stretching vibrations,

the assignment of bands to complexes A and B is very difficult and only tentative. For

complex A the experimental O-H stretching frequency shift of -550 cm−1 is in excellent

agreement with the calculated shift (-538 cm−1) using MP2/cc-pVTZ (Table 3.15). For

dimer B the calculated frequency shift (-449 cm−1) also agrees very well with the

experimental value (-473 cm−1) (Table 3.16). The large shifts of the O-H stretching

vibrations allows to discard the FA – DME dimers C – F where the O-H…O interaction

is lacking and thus the O-H stretching vibration of the FA molecule is much less

perturbed (Figure 3.12).

The formation of hydrogen-bonded complexes of FA results in a contraction of the C-O

bond and a blue shift of the C-O stretching vibration found at 1104 cm-1 in the

unperturbed molecule. For complex A a blue shift of 89 cm-1 is predicted at the MP2/cc-

pVTZ level of theory which matches the experimental value of 78 cm-1. For complex B a

shift of 91 cm-1 is predicted and 76 cm-1 is observed.

The DME molecule shows less pronounced shifts of IR absorptions due to the complex

formation. The strong absorptions of monomeric DME and other constituents in the

matrix do not allow to identify C-H stretching vibrations of the DME molecule in the

complexes. The asymmetric and symmetric C-O-C stretching vibration of both

complexes A and B could be identified. In complex A the symmetrical C-O-C stretching

vibration exhibits a red-shift of 25 cm-1 while the asymmetrical vibration is shifted by 15

cm-1. This might be rationalized by the smaller distortion of the weak

101

HOC(H)=O…HCH2OCH3 hydrogen bonds during the asymmetrical vibration as

compared to the symmetrical vibration. In complex B the corresponding shifts are with

15 and 6 cm-1 smaller. The experimental red-shifts are in good agreement with the MP2

calculated values (Tables 3.15 and 3.16).

3.5. Conclusion

The geometries of the FA – DME and FA – FMA complexes are calculated starting

from randomly generated molecular arrangements using the MMH procedure. Nine FA –

FMA dimers with binding energies between -2.91 and -13.02 kcal/mol (MP2/aug-cc-

pVTZ + ZPE + BSSE) are identified and seven competitive individual molecular

interactions are discussed based on the calculated geometries and binding energies of the

FA – FMA complexes. The pair contributions of each dimer to the stabilization of the 1:2

FA – FMA trimers are also analyzed.

Six different FA – DME bimolecular 1 : 1 complexes with binding energies between -

2.26 and -7.97 kcal/mol (MP2/cc-pVTZ + ZPE + BSSE) could be identified. They are

classified in two groups. One group consists of the complexes A and B, where the OH

hydrogen atom of FA forms a strong hydrogen bond with the ether oxygen atom of DME.

These two complexes are predicted to be the most stable ones and with -7.81 to -7.97

kcal/mol binding to be almost isoenergetic. Although the OH…O interaction is

dominating in complexes A and B, the secondary interaction between a methyl group

hydrogen atom of DME and the carbonyl oxygen atom of FA leads to an additional

significant stabilization of these complexes.

The second group of FA – DME complexes C – F is defined by the absence of the

strong OH…O hydrogen bond. With -2.3 to -2.9 kcal/mol the binding energy is

considerably smaller and consequently these complexes could not be identified

experimentally. The dominant interaction in these complexes is the interaction between

the aldehyde hydrogen atom of FA and the DME oxygen atom. Again, interactions

between the methyl groups of DME and oxygen atoms of FA form secondary, weak

interactions which, however, determine the geometry of the complexes.

102

The structures of the FA – FMA dimers A and B are in excellent agreement with the

geometries of the FA – FMA dimers reported in the literature. They show also interesting

analogies with the FMA – water and FMA – methanol dimers. Comparing the structures

of the FA – DME complexes A and B with the DME – methanol complex reported in

literature reveals large similarities. There are also interesting analogies between the FA –

DME and the FA – water reported dimers.

The B3LYP density functional with the cc-pVTZ basis set provides reliable geometries

for the FA – FMA complexes. At the MP2 level of theory, basically no change of the

geometries is found when the basis set is augmented by adding diffuse functions. At the

MP2 level cc-pVDZ calculations show a tendency to overestimate the binding energies,

however, triple zeta basis sets either augmented or non-augmented result in binding

energies very similar to those from CCSD(T)/cc-pVTZ single point calculations.

The geometries and energies of the FA – FMA dimers A and B do not change

considerably with the inclusion of BSSE corrections during the optimization process. For

the FA – DME dimers A – D, the intermolecular distances corresponding to weaker

interactions, where the DME hydrogen atoms are involved, are more influenced, showing

an increasing of about 0.2-0.3 Å when BSSE is considered during the geometry

optimization. Despite these variations the basic geometries and interactions in the FA –

DME complexes do not change.

The calculated geometries and binding energies of 1:2 and 1:4 FA – FMA complexes

show very interesting similarities with the FA – FMA dimers and with the FA – FMA

crystal structure. Of special interest are structural motives found in the crystal structure

that are already present in complexes of very few molecules. This could lead to in-depths

knowledge of the complex processes of molecular nucleation and crystal growth.

The distortion of the intramolecular distances and vibrational frequencies in the FA –

FMA dimers A and B compared to the monomers are discussed, and reliable vibrational

frequencies are predicted. The empirically corrected frequencies should allow for the

experimental detection of these complexes in matrix isolation or gas phase studies. For

the FA – DME dimers A and B, comparison of the experimental data from matrix

103

isolation experiments with calculated spectra indicated the formation of these two

complexes.

The identification of the FA – DME and FA – FMA complexes, as well their

comparison with matrix spectroscopy results and crystal structure data, confirms the

quality of the MMH procedure as a very useful tool for reliably localizing minima in

hydrogen bonded complexes without recurring to previous knowledge of the structure of

supramolecular complexes.

104

4. Formic Acid Complexes with π Systems

4.1. Introduction

During the past several years, weak hydrogen bonds involving a hydrogen atom bound

to a carbon atom as hydrogen donor has attracted attention from the scientific

community. It was found that these weak interactions play important roles in many

chemical and biochemical processes.[1-3] In contrast to the conventional strong hydrogen

bonds, which were extensively discussed in literature, the nature and characteristics of

weak interactions do not comprise a well-known field.[3] Thus, experimental and

theoretical studies of CH…O, and C-H…π as well as O-H…π interactions are of key

interest for the understanding of biological and chemical systems.

As already mentioned, the formic acid molecule and its interactions with other

molecules provide a good model for the study of many association processes. From early

spectroscopic studies, acetylene was recognized as a weak hydrogen bond donor and

weak acceptor through the triple bond.[1] Furan is an aromatic heterocyclic compound

with many applications in different fields of chemistry from natural product synthesis to

material science.[150]

A number of studies have been published on the weak interactions in complexes or

dimers of hydrocarbons with π systems like acetylene, benzene and ethylene.[151-157] The

dimer of acetylene was investigated both experimentally and theoretically,[158-165] and a

π–type hydrogen-bonded C2v minimum was found together with several first and second

order saddle points.[166-170] Studies of the 1:1 FA – acetylene complexes were carried out

using matrix isolation and theoretical methods.[116]

Furan, its aggregates, and complexes with small molecules have been subject to many

experimental and theoretical studies.[171-187] The complexes of furan with hydrogen

halides and alkynes were investigated by Ault using matrix isolation spectroscopy.[185, 187]

The equilibrium structures of the furan dimers and the nature of their intermolecular

interactions were studied by Pei using density functional theory and the natural bond

orbital analysis.[171] The π…π interactions in the parallel, sandwich-shaped furan dimer

were analyzed using high level ab initio theory.[188] The excited states of furan and the

105

rotational spectra of some of its complexes with halides were also investigated.[176, 177, 179]

Chan, Del Bene et al. studied the reactions of various acids with furan as part of their

research about the preferred sites of protonation and hydrogen bonding for a set of basic

substrates.[114]

Many studies of the interactions of furan with other molecules were focused on the

furan – hydrogen halide complexes.[174, 177, 185] Legon and Millen[189, 190] proposed some

general rules to predict the geometry of B…HX complexes (X: F, Cl, Br, I, CN, CCH). If

the Lewis base possesses both n pairs and π electrons, the angular geometry is determined

by the n pair rather than by the π electrons, and the Lewis acid lies along the axis of the n

pair. Since the furan molecule contains both an n pair and π electrons, several studies of

furan - HX complexes have been carried out to determine whether or not this rule is

obeyed.[174, 177, 179] Cole, Legon and Ottaviani[177] studied the rotational spectrum of furan

– HCl and HBr and showed that, while the HCl complex obeys Legon’s rule, the HBr

complex behaves differently and interacts with the π system of furan. In the same line,

the furan complexes with hydrogen halides HX were studied by Huang and Wang[174]

using ab initio calculations.

The structures of a variety of FA – furan complexes are identified here and their

binding energies and the influence of the BSSE and basis sets on their geometries are

discussed. The calculated vibrational spectra of the FA – furan dimers are compared to

experimental matrix isolation spectra. The FA – furan system exhibits both n and π

hydrogen bond interactions. Therefore, the study of the FA – furan complexes leads to a

detailed knowledge of the interactions in this type of systems.

The trimers formed by the interaction of one molecule of formic acid with two

molecules of acetylene (1:2 FA – acetylene complexes) are investigated. The

characterization of trimer complexes is a challenge both for theory and experiment. In

addition, these complexes consisting of three molecules allow to analyze in detail the

influence of a third molecule on the dimer properties, e. g. an additional acetylene

molecule on the FA – acetylene dimer or a formic acid molecule on the acetylene dimer.

An interesting feature of this system is the competition between the strongly acidic

carboxyl group, the acetylene group, and the formyl group as hydrogen bridge donors and

106

the carbonyl group, the hydroxyl group, and the acetylene π-system as hydrogen bridge

acceptors. This leads to several complexes with strong OH...O and weak CH...O or

CH...π hydrogen bridges. Due to the amount of complexes with similar binding energies

a careful search for minima on the potential energy surface is mandatory. The calculated

vibrational frequencies of the 1:2 FA – acetylene complexes are compared to data from

matrix isolation spectroscopy.

4.2. Computational methods

The Multiple Minima Hypersurface (MMH) approach was used for searching

configurational minima in the FA – furan and 1:2 FA – acetylene systems. One thousand

randomly arranged FA – furan and1:2 FA – acetylene clusters were generated as starting

point, and the resulting geometries were optimized and analyzed using PM3 and AM1

semiempirical quantum mechanical Hamiltonians. These semiempirical results provided

a preliminary overview of the FA – furan and 1:2 FA – acetylene interactions, and the

relevant configurations were further refined using ab initio methods at various levels of

theory.

After geometrical analysis both PM3 and AM1 Hamiltonians lead to nearly the same

set of minima. Compared to the PM3 results, the AM1 calculation allows for the

identification of another FA – furan minimum at all levels of theory. Another two

additional minima from AM1 starting geometries were found only when non-augmented

double zeta basis sets were used for the geometry optimizations.

The ab initio computations were performed using the Gaussian 98 and Gaussian 03

programs. The equilibrium geometries and vibrational frequencies were calculated at the

SCF level including second order Møller−Plesset perturbation theory, MP2. The tight

convergence criteria was used for the geometries optimizations and the force constants

where calculated when necessary.

The 1:2 FA – acetylene complexes were calculated in addition at the DFT level with

the B3LYP hybrid functional for initial geometry optimizations. For the 1:2 FA –

acetylene complexes A, B and C single point calculations are performed with coupled

clusters[141] of single and double substitutions (with non iterative triples) CCSD(T)/cc-

107

pVTZ, using the MOLPRO program. The DFT results for the 1:2 FA – acetylene

complexes are presented to compare with the MP2 and CCSD(T) data.

Pople’s 6-31G(d,p) basis set, the extended valence triple ζ basis set augmented with

diffuse and polarization functions 6-311++G(d,p) and the cc-pVTZ Dunning’s correlation

consistent triple ζ basis set were used. For the FA – furan dimers, in addition the

augmented and non augmented Dunning’s correlation consistent double ζ basis sets cc-

pVDZ and aug-cc-pVDZ were used. To evaluate the effect of the size of the basis set in

1:2 FA – acetylene complexes, the strongly polarized basis set 6-311++G(3df,3pd) (381

basis functions) is used and the results are compared to MP2 calculations with the cc-

pVTZ (294 basis functions) and 6-311++G(d,p) (196 basis functions) basis sets.

The stabilization energies were calculated by subtracting the energies of the monomers

from those of the complexes and including ZPE corrections. Most of the FA – furan

dimer energies were also corrected for the basis set superposition errors (BSSE) using the

counterpoise (CP) scheme of Boys and Bernardi. In the case of the 1:2 FA – acetylene

complexes, the binding energies for complexes A, B, and C were CP-BSSE corrected at

the MP2/cc-pVTZ level of theory.

To investigate the influence of the basis set superposition errors (BSSE) on the

geometries of the complexes, all FA – furan dimers were optimized at the MP2/6-

31G(d,p) level of theory using the CP corrections during the optimization process.

108

4.3. Formic acid – furan dimers. Results and discussion

Geometries and binding energies

After refining the MMH semiempirical results, nine FA – furan dimers A – I were

localized at the MP2/6-311++G(d,p) level of theory. The FA molecule exhibits two

protons that in principle can act as hydrogen bridge donor: the more acidic OH proton

and the less acidic aldehyde type CH proton. Consequently, the FA – furan dimers are

classified into two types:

• Type (i), where the acidic OH hydrogen atom of FA acts as hydrogen bond

donor (Dimers A, B, C and D1) (Figure 4.1).

• Type (ii), where the less acidic CH hydrogen atom of FA acts as hydrogen bond

donor (Dimers E, F1, G, H and I) (Figure 4.2).

The interactions between the FA and furan molecules in the complexes can be broken

down into the six basic two center interactions (1) – (6):

(1) OHFA…OF interaction between the hydroxyl hydrogen atom of FA and the

oxygen atom of furan.

(2) C=OFA…HF interaction between the carbonyl oxygen atom of FA and the

hydrogen atom of furan.

(3) O-HFA…π interaction between the hydroxyl hydrogen atom of FA and the π

system of furan.

(4) CHFA…OF interaction between the aldehyde hydrogen atom of FA and the

oxygen atom of furan.

(5) H-OFA…HF interaction between the hydroxyl oxygen atom of FA and the

hydrogen atom of furan.

(6) C-HFA…π interaction between the aldehyde hydrogen atom of FA and the π

system of furan.

109

Figure 4.1. The calculated structures with hydrogen bond lengths (Å) of the type (i)

FA – furan complexes A, B, C, and D1 at the MP2/6-311++G(d,p) level of theory.

110

Figure 4.2. The calculated structures with hydrogen bond lengths (Å) of the type

(ii) FA – furan complexes E, F1, G, H, and I at the MP2/6-311++G(d,p) level of

theory.

The complexes A – I were found being minima using double as well as triple zeta basis

sets. In addition, the five complexes D, D2, F, K, and J were localized at the double zeta

without augmentation level of theory only. These complexes do not represent minima at

higher levels of theory using larger basis sets. (Figure 4.3)

111

Figure 4.3. The calculated structures with hydrogen bond lengths (Å) of FA –

furan dimers D, D2, F, J, and K at the MP2/6-31G(d,p) level of theory. 1) Geometry

optimized without BSSE corrections. 2) Geometry optimized with BSSE corrections.

Thus, dimers showing the basic interactions (1) or (3) are classified as type (i)

complexes, while those showing interactions (4) or (6) are classified as type (ii)

complexes. In general, more than one of the basic interactions (1) – (6) contributes to the

stabilization of the complex.

Type (i) complexes

The most stable FA – furan dimer is complex A with a binding energy of -3.91

kcal/mol at the MP2/6-311++G(d,p) + BSSE + ZPE level of theory (Table 4.1). The

binding energies of all dimers are discussed at this level of theory including BSSE and

ZPE corrections unless specified differently. The influence of the basis sets on the

complexes is discussed later.

112

TABLE 4.1: Calculated binding energies of furan-Formic Acid dimers A-I at MP2

level of theory with the 6-31g(d,p) and 6-311++G(d,p) Pople’s basis set including

ZPE and BSSE corrections (in kcal/mol).

MP2 6-31g(d,p) opt with BSSE 6-31g(d,p) ΔE BSSE ZPE ΔE BSSE+ZPE ΔE BSSE ZPE ΔE BSSE+ZPE A -8.37 2.83 1.01 -4.53 -8.54 3.15 1.11 -4.27 B -5.16 1.65 0.61 -2.90 -6.05 3.11 0.88 -2.06 C -5.02 1.80 0.55 -2.67 -5.67 3.00 0.65 -2.02 D -5.11 1.61 0.48 -3.02 -5.76 2.80 0.64 -2.32 D1 D2a Da D2 -5.16 1.66 0.54 -2.96 Ba E -3.30 1.44 0.55 -1.31 -3.73 2.29 0.64 -0.8 F -2.89 1.41 0.40 -1.08 -3.17 1.91 0.65 -0.61 F1 -2.89 1.27 0.48 -1.14 -3.21 1.95 0.58 -0.68 G -2.88 1.27 0.46 -1.15 -3.21 1.96 0.58 -0.67 H -4.26 1.96 0.56 -1.74 -4.38 2.19 0.67 -1.52 I -3.40 1.56 0.47 -1.37 -3.50 1.75 0.51 -1.24 J Ha -3.37 1.99 0.64 -0.74 K —b -2.59 1.59 0.43 -0.57 6-311++G(d,p) ΔE BSSE ZPE ΔE BSSE+ZPE A -7.12 2.03 1.18 -3.91 B -5.51 2.24 1.03 -2.24 C -5.56 2.42 1.02 -2.12 D D1a D1 -5.44 2.12 0.95 -2.37 D2 D1a E -4.19 2.21 1.02 -0.96 F F1a F1 -3.76 2.01 0.93 -0.82 G -3.73 1.99 0.88 -0.86 H -3.55 1.17 0.44 -1.94 I -3.05 1.18 0.52 -1.35

113

J Ea K —b

a Indicates the dimer that it is found after geometry optimization. b Leads to a second

order saddle point.

For comparison, selected intermolecular parameters of the dimers A – D2 are shown in

Table 4.2. The dimer A is stabilized by interaction (1) that involves the OH of the FA

molecule and the oxygen atom of furan, and interaction (2) between the carbonyl oxygen

atom of FA and a α-hydrogen atom of furan. The binding distances and hydrogen bond

angles (in parentheses) for interactions (1) and (2) are 1.877 Å (177.87º) and 2.540 Å

(124.56º), respectively (Table 4.2). If dimer A is forced to Cs symmetry it shows one

imaginary out of plane vibration that after free optimization leads to a slightly distorted

equilibrium geometry with C1 symmetry (Figure 4.1). The intramolecular bond lengths

are not sensitive to this decrease of symmetry in dimer A, and the variation of the

intermolecular hydrogen bond distances and angles are less than 0.008 Å and 1º,

respectively (Table 4.2). As expected, the binding energy is also hardly affected. Since

with other basis sets the Cs symmetrical dimer is a true minimum, the distortion is

probably an artifact of the 6-311++G(d,p) basis set.

Dimers B, C, and D1 are O-H…π complexes (interaction (3)) with very similar binding

energies. Dimer D1 has binding energy -2.37 kcal/mol and the OH group of FA interacts

with the C2-C3 region of furan. The distances of the OH hydrogen atom to C2 and C3 are

2.321 Å and 2.508 Å, respectively (Figure 4.1). In the complexes B (-2.24 kcal/mol) and

C (-2.12 kcal/mol) the OH hydrogen atom of FA interacts with the C1-C2 bond of furan.

In dimer B, the O-H…C2 and O-H…C1 distances are 2.369 Å and 2.427 Å, respectively.

In dimer C the O-H…C2 and O-H…C1 distances are 2.300 Å and 2.551 Å, respectively.

The main difference between dimers B and C is the rotation of the FA molecule around

its O-H axis (Figure 4.1). Compared to dimer B the OH hydrogen atom in dimer C is

closer to atom C2.

114

TABLE 4.2: Comparison of selected intermolecular parameters in the FA – furan dimers A-D2 at various levels of theory.

Distances are in Å and angles in degrees

MP2 6-31G(d,p)BSSE 6-31G(d,p) 6-311G++(d,p) cc-pVDZ aug-cc-pVDZ cc-pVTZ Dimer A OHFA…OF 1.973 1.889 1.877 (1.885)a 1.888 1.859 1.866

C=OFA…H(C1)F 2.551 2.416 2.540 (2.536)a 2.400 2.433 2.447 < OHFA…OF 177.93 176.58 177.87(177.53)a 177.43 177.56 178.04 < C=OFA…H(C1)F 124.66 124.65 124.56(125.26)a 125.25 124.67 124.22 Dimer B OHFA…OF 3.074 3.289 2.756 3.225 2.656 2.698 OHFA…C1F 2.566 2.426 2.427 2.409 2.342 2.350 OHFA…C2F 2.432 2.354 2.369 2.328 2.294 2.305 OHFA…C3F 2.907 3.242 2.684 3.167 2.592 2.640 OHFA…C4F 3.234 3.672 2.881 3.579 2.774 2.835 <OHFA…OF 139.14 155.35 124.33 153.96 125.38 127.47 <OHFA…C2F 168.89 159.38 170.44 159.92 171.13 169.91 Dimer C OHFA…OF 3.103 2.871 2.961 2.895 2.857 2.871 OHFA…C1F 2.679 2.450 2.551 2.495 2.429 2.426 OHFA…C2F 2.441 2.303 2.300 2.297 2.236 2.253

115

MP2 6-31G(d,p)BSSE 6-31G(d,p) 6-311G++(d,p) cc-pVDZ aug-cc-pVDZ cc-pVTZ OHFA…C3F 2.772 2.672 2.610 2.617 2.596 2.638 OHFA…C4F 3.131 2.965 2.970 2.933 2.921 2.957 C=OFA…H(C4)F 3.788 3.219 3.365 3.209 3.197 3.313 <OHFA…OF 154.04 144.60 146.22 144.77 145.56 146.72 <C=OFA…H(C4)F 125.75 133.29 128.06 131.09 131.47 130.26 <OHFA…C2F 158.89 165.43 165.18 165.32 163.45 163.20 Dimer D C=OFA…H(C2)F 3.586 2.929 2.900 3.129 OHFA…OF 3.234 3.359 3.368 3.098 OHFA…C1F 2.967 3.007 3.013 2.816 OHFA…C2F 2.510 2.366 2.359 2.323 OHFA…C3F 2.510 2.366 2.359 2.323 OHFA…C4F 2.967 3.007 3.013 2.816 <C=OFA…H(C2)F 102.05 105.46 105.89 104.83 <OHFA…OF 139.35 156.72 156.57 146.91 <OHFA…C2F 163.45 156.84

D1b 156.76

D1b 159.51

Dimer D1 C=OFA…H(C3)F 3.225 3.122 3.111 OHFA…OF 2.911 2.845 3.014 OHFA…C1F 2.595 2.487 2.637 OHFA…C2F

D2b 2.321

2.230 2.244

116

OHFA…C3F 2.508 2.476 2.440 OHFA…C4F 2.846 2.818 2.890 <C=OFA…H(C3)F 111.40 110.95 108.37 <OHFA…OF 129.68 129.38 141.66 <OHFA…C2F

176.60

178.54 170.14 Dimer D2 C=OFA…H(C2)F 3.255 OHFA…OF 3.206 OHFA…C1F 2.769 OHFA…C2F 2.416 OHFA…C3F 2.704 OHFA…C4F 3.150 <C=OFA…H(C2)F 108.22 <OHFA…OF 138.07 <OHFA…C2F 176.29

Bb

D1b

Db

D1b

—c

a Geometrical parameters in the Cs symmetry dimer A b Indicates the dimer that it is found after geometry optimization. b Leads to a

saddle point structure with two imaginary frequencies. c The geometry was not optimized at this level of theory

117

Huang et al. calculated the electrostatic potential map of furan at the MP2/6-

311++G(d,p) level of theory.[174] Their results indicate that the region with the strongest

negative potential is in the vicinity of the O atom. Along the C2-C3 bond the negative

potential is also noticeable. This analysis based on electrostatic potentials agrees well

with our results of the type (i) FA – furan dimers. For the most stable dimer A the OH

hydrogen atom of FA interacts with the furan oxygen atom at the site of the strongest

negative potential. Dimer A is 1.54 kcal/mol more stable that dimer D1 where the OH

hydrogen atom of FA interacts with the C2-C3 bond of furan, which indicates a stronger

interaction of the acidic hydrogen atom with the furan oxygen atom than with the π

system.

Type (ii) complexes

The FA – furan dimers E – I are very weakly bound complexes (Table 4.1, Figure 4.2).

Dimers H (-1.94 kcal/mol) and I (-1.35kcal/mol) are both C-H…O complexes where the

CH hydrogen atom of FA interacts with the lone pairs of the furan oxygen atom

(interaction (4)) at 2.502 and 2.488 Å distances, respectively. Dimer H in addition is

stabilized by interaction (2) between the carbonyl oxygen atom of FA and the α-hydrogen

atom of furan (2.508 Å). Complex I is stabilized additionally by interaction (5) with an

intermolecular distance of 2.569 Å between the OH oxygen atom of FA and a ring

hydrogen atom of furan.

Dimers E, F1and G are energetically very close with binding energies of -0.96, -0.82

and -0.86 kcal/mol, respectively. All of them show a combination of the C-HFA…π

interaction (6) and an atypical C-H…O interaction in the plane of the π orbitals. The C-

H…O distances are 2.780 Å, 2.694 Å, and 2.682 Å for the E, F1, and G dimers,

respectively. The C-H…C1 distances take values of 2.814 Å, 2.673 Å, and 2.705 Å, in

that order. Similarly to dimers B, C, and D1, the dimers E, F1, and G show the formic

acid molecule “traveling” around the furan molecule by rotating around its CH axis. The

geometries of the type (ii) complexes suggest that not only electrostatic interactions, but

probably also orbital interactions govern the structures of the complexes.

118

Other FA – furan geometries

Two additional type (i) and three type (ii) FA – furan dimers D, D2, F, J and K were

found with smaller basis sets than 6-311++G(d,p) (Figure 4.3). As mentioned before, at

higher levels of theory these complexes do not represent minima. Dimer D shows Cs

symmetry and at the MP2/6-31G(d,p) + ZPE + BSSE level of theory its binding energy is

-2.32 kcal/mol. The O-H…C2 distance is calculated to 2.366 Å and the O-H…C2 bond

angle to 156.84º (Tables 4.1 and 4.2, Figure 4.3). At the MP2/6-31G(d,p)BSSEopt + ZPE

+ BSSE level of theory the binding energy of dimer D is with -3.02 kcal/mol slightly

larger. The O-H…C2 distance is predicted to 2.510 Å, and the hydrogen bond angle to

163.45º.

Dimer D2 could only be localized at the MP2/6-31G(d,p) level of theory if CP-BSSE

corrections were included during the geometry optimization (Tables 4.1 and 4.2). At this

level of theory the binding energy of dimer D2, including the ZPE and BSSE corrections,

is -2.96 kcal/mol. The O-H…C2 distance is 2.416 Å and the corresponding hydrogen

bond angle is 176.29º.

By comparing the structures of dimers D, D1, and D2 (Figures 4.1 and 4.3) it is clear

that D1 and D2 are the C1 symmetrical structures that derive from the Cs symmetrical

dimer D by tilting and slightly shifting the FA molecule. Of these structures, only the D1

complex is a minimum at higher level of theory (6-311++G(d,p) and aug-cc-pVDZ basis

sets).

The type (ii) dimers F, K and J were only localized with the double zeta basis sets

without augmentation. Dimer F is a very weak interacting complex with an MP2/6-

31G(d,p)+ZPE+BSSE binding energy of -0.61 kcal/mol. At the same level of theory the

C-HFA…C2F and C=OFA…HF distances are 2.495 and 2.839 Å, respectively. Dimer F is

mainly stabilized by the weak interaction (2) between the carbonyl oxygen atom of FA

and a ring hydrogen atom of furan. Similarly, dimer J (-0.74 kcal/mol) is stabilized by

interaction (2) with an intermolecular distance of 2.477 Å and by the CH...π interaction

between the CH hydrogen atom of FA and the C1 carbon atom of furan (2.893 Å

intermolecular distance).

119

TABLE 4.3: Comparison of selected intermolecular parameters in the FA – furan dimers E-K at various levels of theory.

MP2 6-31G (d,p) BSSE 6-31G (d,p) 6- 311G ++ (d,p) cc-pVDZ aug-cc-pVDZ cc-pVTZ Dimer E CHFA…OF 2.888 2.809 2.780 2.851 2.717 2.770 CHFA…C1F 2.906 2.839 2.814 2.898 2.714 2.799 CHFA…C2F 2.967 2.934 2.911 3.022 2.751 2.868 CHFA…C3F 2.970 2.946 2.924 3.030 2.756 2.863 CHFA…C4F 2.912 2.859 2.834 2.912 2.723 2.792 <CHFA…OF 103.15 87.25 92.84 84.21 92.53 92.34 <CHFA…C1F 120.61 106.14 111.60 102.50 114.74 111.06 Dimer F

C=OFA…H(C2)F 2.815 2.495 2.430 CHFA…OF 4.227 4.103 4.349 CHFA…C1F 3.614 3.198 3.415 CHFA…C2F 2.911 2.839 2.943 CHFA…C3F 3.228 3.681 3.793 CHFA…C4F 4.010 4.322 4.516

<C=OFA…H(C2)F 94.04 95.02 96.38 <CHFA…OF 148.16 143.51 137.37 <CHFA…C2F 119.33 114.87

F1a 117.79

F1a

—b

Dimer F1 C=OFA…H(C3)F 4.299 3.886 3.850 3.880 4.099 3.742

120

MP2 6-31G (d,p) BSSE 6-31G (d,p) 6- 311G ++ (d,p) cc-pVDZ aug-cc-pVDZ cc-pVTZ CHFA…OF 2.762 2.632 2.694 2.691 2.577 2.651 CHFA…C1F 2.798 2.687 2.673 2.717 2.548 2.633 CHFA…C2F 3.043 2.957 2.857 2.971 2.764 2.844 CHFA…C3F 3.144 3.053 2.973 3.082 2.902 2.972 CHFA…C4F 2.965 2.850 2.863 2.900 2.774 2.842

<C=OFA…H(C3)F 70.61 75.61 77.88 76.83 80.96 79.47 <CHFA…OF 140.60 127.62 126.14 121.88 126.95 127.11 <CHFA…C1F 150.94 140.77 146.69 139.21 144.93 143.98 Dimer G CHFA…OF 2.716 2.637 2.682 2.710 2.561 2.643 CHFA…C1F 2.828 2.704 2.705 2.726 2.568 2.660 CHFA…C2F 3.123 2.982 2.933 2.972 2.817 2.900 CHFA…C3F 3.185 3.073 3.036 3.086 2.941 3.014 CHFA…C4F 2.934 2.859 2.876 2.916 2.775 2.847 <CHFA…OF 139.05 121.55 122.09 114.79 122.00 120.90 <CHFA…C1F 152.15 140.15 142.56 137.06 143.46 142.31 Dimer H CHFA…OF 2.558 2.440 2.502 2.467 2.424 2.462

C=OFA…H(C1)F 2.544 2.437 2.508 2.412 2.413 2.432 <CHFA…OF 131.72 129.90 127.37 129.99 134.06 131.44 <C=OFA…H(C1)F 103.50 105.56 108.39 105.07 101.56 104.16

121

Dimer I CHFA…OF 2.532 2.426 2.488 2.464 2.396 2.460 H-OFA…H(C1)F 2.646 2.533 2.569 2.504 2.505 2.543

<CHFA…OF 143.71 140.90 136.20 140.05 145.70 141.31 < H-OFA…H(C1)F 145.38 142.33 136.42 142.09 146.89 143.00

Dimer J CHFA…C1F 2.893 2.992 CHFA…C2F 3.054 3.261 CHFA…OF 3.834 3.959 C=OFA…H(C1)F 2.477 2.399 <CHFA…C1F 110.82 108.58 <CHFA…OF 119.01 114.49 <C=OFA…H(C1)F Ha 99.21

Ea 99.77

Ea

— b

Dimer K CHFA…OF 4.099 4.361

CHFA…C2F 2.902 3.021

<CHFA…C2F — c

122.16

— c

117.41

— c

— b

a Geometrical parameters in the Cs symmetry dimer A. b Indicates the dimer that it is found after geometry optimization. b The

geometry was not optimized at this level of theory. c Leads to a saddle point structure with two imaginary frequencies.

122

Dimer K (-0.57 kcal/mol) exhibits Cs symmetry and its geometry is much related to the

structure of dimer D (Figure 4.3). In dimer K the CH hydrogen atom of FA interacts with

the C2-C3 bond of furan. The C-H FA …C2 F distance is 2.902 Å. The larger A-HFA

…C2F (A = O, C) bond length of dimer K compared to dimer D and the difference

between their binding energies (dimer D is 1.75 kcal/mol more stable than dimer K) are

clearly due to the difference in the acidities of the OH hydrogen atom and the CH

hydrogen atom of the FA molecule.

It is remarkable that both structures D and K were localized after a procedure based on

a completely random exploration of the FA – furan dimers potential energy surface. The

most stable FA – furan dimer A and all the other type (i) and type (ii) dimers, as well as

their enantiomers were found using the MMH procedure.

Basis set influence on the calculated geometries of the FA – furan dimers

To analyze the influence of the basis set on the geometries of the FA – furan dimers,

the structures of the dimers were optimized using the 6-31G(d,p), 6-311++G(d,p), cc-

pVDZ, aug-cc-pVDZ and cc-pVTZ basis sets. Dimers A, B and C are minima with all

basis sets (Tables 4.1 and 4.4), but only with the 6-311++G(d,p) basis set the equilibrium

geometry of dimer A deviates from the Cs symmetry plane (Figure 4.1). Dimer D1 is

only a minimum using the triple zeta and augmented double zeta basis sets. With the 6-

31G(d,p) basis set the optimization of complex D1 leads to D, whereas with the 6-

31G(d,p) basis set including CP corrections during the geometry optimization dimer D2

is produced (Table 4.1).

The structure D was found as a minimum after geometry optimization at the MP2 level

with the 6-31G(d,p) (with and without BSSE-CP corrections) and cc-pVDZ basis sets.

However, when the basis set is augmented by diffuse functions (aug-cc-pVDZ, 6-

311++G(d,p)) it becomes a transition state that leads to the equilibrium geometry D1

(Tables 4.1 and 4.4). At the MP2/cc-pVTZ level dimer D is apparently a minimum, but

due to the huge computational requirements vibrational spectra were not calculated at this

level. Therefore, it is likely that with the cc-pVTZ basis set D would become a transition

state leading to D1, similar to the aug-cc-pVDZ and 6-311++G(d,p) basis sets.

123

TABLE 4.4: Calculated binding energies of FA – furan dimers A-I at MP2 level of

theory with the cc-pVDZ, aug-cc-pVDZ and cc-pVTZ Dunning’s basis set including

ZPE corrections (in kcal/mol).

MP2 cc-pVDZ aug-cc-pVDZ cc-pVTZ ΔE ZPE ΔE ZPE ΔE BSSE A -8.56 1.13 -7.58 1.06 -7.14 1.61 B -5.93 0.81 -6.85 0.78 -5.73 1.50 C -5.72 0.65 -6.77 0.76 -5.63 1.57 D -5.85 0.68 D1 -5.65 1.43 D1 D -6.77 0.73 -5.65 1.44 D2 D D1 —b E -3.68 0.60 -5.30 0.82 -3.75 1.15 F -3.40 0.64 F1 — b

F1 -3.14 0.57 -5.06 0.80 -3.44 1.07 G -3.17 0.57 -5.09 0.82 -3.46 1.06 H -4.67 0.68 -4.19 0.63 -3.74 1.17 I -3.73 0.52 -3.45 0.52 -2.90 0.92 J -3.61 0.64 E — b K -2.79 0.46 —a — b

a Leads to a saddle point geometry with two imaginary frequencies. b The geometry of

the complex was not optimized at this level of theory.

As already mentioned, dimer D2 could be found only at the MP2/6-31G(d,p) level of

theory when including the CP BSSE corrections during the geometry optimization. With

the augmented basis sets dimer D2 became D1 after geometry optimization. At the

MP2/6-31G(d,p) level without the CP corrections the geometry optimization of D2 lead

to structure B. With the cc-pVDZ basis set dimer D2 is converted into dimer D, which is

a minimum at this level of theory (Tables 4.1 and 4.4).

The dimers E, F1, G, H, and I are minima with all basis sets used (Tables 4.1 and 4.4).

Complex I exhibits Cs symmetry at all levels, except with the 6-311++G(d,p) basis set

where it adopts C1 symmetry (Figure 4.2). Dimer F is found only with the small double

124

zeta basis sets without augmentation. With the 6-311++G(d,p) and aug-cc-pVDZ basis

sets it transforms to dimer F1 after geometry optimization (Table 4.1).

Dimers J and K were localized only using the cc-pVDZ and 6-31G(d,p) basis sets.

When the BSSE correction was included during the geometry optimization (6-31G(d,p)

basis set), dimer J was transformed to dimer H (Table 4.1). In contrast, with the 6-

311++G(d,p) and aug-cc-pVDZ basis sets the geometry optimization of J leads to dimer

E. Geometry optimization of complex K at the 6-31G(d,p)BSSEopt, aug-cc-pVDZ and 6-

311++G(d,p) levels of theory results in a second order stationary point (Tables 4.1 and

4.4).

Tables 4.5 and 4.6 show some selected intramolecular distances at various levels of

theory for the FA – furan dimers, respectively. The C=OFA distances decrease up to 0.012

Å when increasing the size of the Pople’s basis set from the 6-31G(d,p) to 6-311++G(d,p)

whereas the O-HFA and O-CF distances decrease less, up to 0.005Å. The C-HFA and C-HF

distances increase with the size of the basis set by around 0.004 Å, and the C2-C3F

distances increase by about 0.005 Å.

With Dunning’s basis set at the double zeta level of theory the O-HFA, C2-C3F and C-

HF distances are not sensitive to the augmentation. With the augmentation the C=OFA and

O-CF bond lengths increase by only 0.006 and 0.009 Å, respectively, and the C-HFA

distances decrease by about 0.005 Å. These variations are slightly more pronounced if the

intramolecular distances calculated with Dunning’s double or triple zeta basis sets are

compared. The cc-pVTZ calculated bond lengths are in general shorter than the cc-pVDZ

values. Thus, the C2-C3F, C-HFA, and C-HF cc-pVTZ distances are 0.011, 0.015 and

0.013 Å, respectively, shorter than the cc-pVDZ values. The C=OFA, O-HFA, and O-CF

bond lengths calculated with the triple zeta basis set are also around 0.006 Å shorter than

the cc-pVDZ values.

125

TABLE 4.5: Comparison of selected intramolecular parameters in the monomer and the FA – furan dimers A - D2 at

various levels of theory. Distances are in Å and angles in degrees

MP2 Expa,b 6-31G(d,p) 6-311G++ (d,p) cc-pVDZ aug-cc-pVDZ cc-pVTZ Monomer O-HFA 0.972 0.972 0.969 0.975 0.975 0.969 C=OFA 1.202 1.213 1.205 1.209 1.215 1.203 C-OFA 1.343 1.351 1.348 1.350 1.359 1.346 C1- C2F 1.354 1.366 1.370 1.378 1.380 1.364 C2- C3F 1.440 1.427 1.432 1.436 1.438 1.425 C1-H F 1.075 1.075 1.079 1.088 1.087 1.074 O-CF 1.371 1.366 1.360 1.364 1.372 1.359 6-31G (d,p )BSSE 6-31G(d,p) 6-311G++ (d,p) cc-pVDZ aug-cc-pVDZ cc-pVTZ Dimer A O-HFA 0.979 0.980 0.977 0.983 0.984 0.978 C=OFA 1.216 1.218 1.210 1.215 1.221 1.208 C-OFA 1.342 1.339 1.338 1.339 1.347 1.335 C1- C2F 1.364 1.363 1.367 1.375 1.377 1.362 C2- C3F 1.429 1.429 1.434 1.438 1.439 1.427 C1-H F 1.075 1.075 1.079 1.088 1.087 1.075 O-C1F 1.374 1.375 1.370 1.373 1.381 1.368 O-C4F 1.371 1.372 1.367 1.370 1.377 1.364 < C1OC4 106.98 107.15 107.33 107.30 107.38 107.19

126

Dimer B O-HFA 0.977 0.978 0.974 0.981 0.982 0.977 O-C1F 1.365 1.365 1.360 1.363 1.370 1.358 O-C4F 1.366 1.366 1.360 1.364 1.372 1.359 Dimer C O-HFA 0.976 0.977 0.975 0.981 0.983 0.977 O-C4F 1.365 1.364 1.358 1.361 1.369 1.357 Dimer D O-HFA 0.976 0.977 0.980 0.976 C2-C3F 1.430 1.431

D1 1.439

D1 1.428

Dimer D1 O-HFA 0.974 0.982 0.977 C2-C3F D2

D 1.434

D 1.440 1.428

Dimer D2 O-HFA 0.976 C2-C3F 1.429

B D1 D D1 —

a FA reference[149] b Furan reference[191]

127

TABLE 4.6: Comparison of selected intramolecular parameters in the FA – furan dimers E - K at various levels of theory

MP2 6-31G (d,p) BSSE 6-31G (d,p) 6-311G++ (d,p) cc-pVDZ aug-cc-pVDZ cc-pVTZ Dimer E C-HFA 1.091 1.092 1.096 1.107 1.102 1.092 O-C1F 1.366 1.367 1.361 1.364 1.372 1.359 Dimer F C-HFA 1.092 1.092 1.107 C=OFA 1.214 1.216 1.212 C2-HF 1.077 1.077 1.090 C3-HF 1.077 1.077

F1 1.089

F1

Dimer F1 C-HFA 1.091 1.091 1.095 1.106 1.102 1.092 C=OFA 1.214 1.214 1.206 1.210 1.216 1.204 C2-HF 1.076 1.077 1.080 1.089 1.088 1.075 C3-HF 1.076 1.076 1.080 1.089 1.088 1.075 Dimer G O-HFA 0.972 0.972 0.969 0.976 0.975 0.970 C-HFA 1.091 1.091 1.095 1.106 1.102 1.092 O-C1F 1.367 1.367 1.362 1.364 1.373 1.360 O-C4F 1.367 1.367 1.361 1.363 1.372 1.359 Dimer H C=OFA 1.215 1.217 1.205 1.213 1.219 1.207

128

C-HFA 1.092 1.091 1.094 1.105 1.102 1.091 O-C1F 1.370 1.371 1.365 1.369 1.377 1.364 O-C4F 1.367 1.367 1.362 1.365 1.373 1.360 H-C1F 1.076 1.075 1.079 1.088 1.087 1.075

<C1OC4 106.74 106.84 107.08 107.00 107.07 106.91 Dimer I C-HFA 1.091 1.091 1.094 1.106 1.102 1.091 C-OFA 1.356 1.357 1.354 1.357 1.365 1.351 O-C1F 1.369 1.369 1.364 1.367 1.376 1.362 Dimer J C-HFA 1.092 1.107 C=OFA 1.216 1.213 C1-C2F 1.368 1.379 H

E

E

Dimer K C-HFA 1.091 1.107

C2-C3F —

1.428 — 1.437

129

The intermolecular bond distances and angles of dimer A are not very sensitive to the

basis set used (Table 4.2, BSSE effects on the geometries of the complexes are discussed

separately). For instance, the C=OFA…HF distance is 0.12 Å larger in the 6-311++G(d,p)

geometry than in the 6-31G(d,p) calculated structure. However, in this case it should be

taken into account that the 6-311++G(d,p) dimer A has C1 symmetry while the structures

of dimer A calculated with the other basis sets have Cs symmetry.

Most interesting is the case of dimer B where the geometry of the complex shows the

largest dependence on the basis set used (Table 4.2). While the 6-311++G(d,p), aug-cc-

pVDZ and cc-pVTZ structures are similar, the geometries calculated with the smaller 6-

31G(d,p) and cc-pVDZ basis sets show larger deviations, although not resulting in new

minima. In the structures of complex B calculated with the 6-31G(d,p) basis set

compared to that with the 6-311++G(d,p) basis set the OHFA…C2F hydrogen bond angle

is reduced by 11º and the OHFA…OF hydrogen bond angle increases by 31º. The cc-

pVTZ calculated hydrogen bond angles of dimer D1 also change considerably when

compared to the 6-311++G(d,p) and aug-cc-pVDZ calculated values (Table 4.2). Table

4.3 shows some selected intermolecular parameters for the dimers E – K at various levels

of theory.

By comparing calculated binding energies of all complexes with the different basis

sets, no general conclusion is evident (Tables 4.1 and 4.4). However, the aug-cc-pVDZ

calculated binding energies are in general higher than at the other levels of theory, except

for dimer A. The 6-311++G(d,p) and cc-pVTZ energies are in general in good agreement.

Taking into account the compromise between quality of the results and the computational

time requirements, we consider that the 6-311++G(d,p) basis set is appropriate for the

geometry optimization of these systems. In many cases the aug-cc-pVDZ basis set

provides good geometries too and compares well to the cc-pVTZ and 6-311++G(d,p)

results.

Effect of the BSSE on the calculated geometries and binding energies

BSSE corrections of the binding energies were calculated for all FA – furan dimers at

the MP2 level of theory with the cc-pVTZ, 6-311++G(d,p) and 6-31G(d,p) basis sets.

With the cc-pVTZ basis set the BSSE is smaller than with the other basis sets (Tables 4.1

130

and 4.4). With the exception of complexes F1 and G, with Pople’s basis sets the BSSE

decreases when increasing the size of the basis set from 6-31G(d,p) to 6-311++G(d,p).

However, the differences between the BSSE with the 6-311++G(d,p) and the 6-31G(d,p)

basis sets are very small (0.06 – 0.03 kcal/mol, Table 4.1).

To evaluate the influence of the BSSE on the calculated geometries the structures of all

FA – furan dimers were optimized at the MP2/6-31G(d,p) level of theory using the

counterpoise (CP) scheme during optimization (6-31G(d,p)BSSEopt). In contrast to our

previous results for other systems[146, 192] but in agreement with other studies,[90] the

BSSE effects considerably the geometry of the FA – furan complexes. As already

mentioned before, structure D2 could only be found with the 6-31G(d,p)BSSEopt

corrected geometry optimization. Furthermore, the 6-31G(d,p)BSSEopt optimizations of

dimers D1, J and K lead to entirely different structures compared to those calculated

without BSSE corrections (Tables 4.2 and 4.3). For dimer K the 6-31G(d,p)BSSEopt

optimization gave the same structure as with the larger cc-pVTZ, 6-311++G(d,p) and

aug-cc-pVDZ basis sets (Tables 4.1 and 4.4).

Here, the 6-31G(d,p)BSSE corrected geometries of the FA – furan dimers compared to

the non BSSE corrected geometries at the same level of theory are discussed. As

expected, the intramolecular parameters of FA and furan in the complexes were not

affected by the inclusion of the BSSE corrections during the geometry optimizations

(Tables 4.5 and 4.6). For the most stable dimer A, there is no significant change when

including BSSE corrections, only the OHFA…OF and C=OFA…HF hydrogen bonds are

elongated by 0.084Å and 0.135Å respectively (Table 4.2).

The geometries of other complexes, especially some angles in dimers B, C, D, F1 and

G, are more effected by the inclusion of the BSSE corrections during the geometry

optimizations (Tables 4.2 and 4.3). For complexes D and F the intermolecular distances

increase up to 0.32Å (C=OFA…HF interaction (2) of complex F) when the BSSE is

included during the geometry optimization (Figure 4.3). Our results indicate that the

BSSE corrections are more important for the weakly bound π dimers than for the relative

strongly bound dimers such as A, H, and I.

131

Comparison with other furan complexes

All geometries of the FA – furan dimers described here were localized via a random

exploration of the multiple minima hypersurface for the FA – furan system without any

previous chemical assumptions. It is therefore interesting to discuss structural similarities

between these FA – furan dimers and furan homo- and heterodimers found by other

methods described in literature.

Pei and Li[171] found four equilibrium isomers of the furan dimer at the B3LYP/6-

311G(d,p) level of theory. The most stable structure shows two equivalents CH…O

interactions at 2.547Å and resembles the geometry of the most stable FA – furan dimer

A. As expected, in dimer A the CH…O interaction is stronger than in the furan

homodimer. The dimer A is stabilized additionally by the C=OFA…HF interaction

between the α-hydrogen atom of furan and the carbonyl oxygen atom of FA.

Huang and Wang[174] found two basic geometries for the furan – hydrogen halides

dimers. One is the atom-on type where the H atom of HX interacts with the nonbonding

electron pairs of the furan oxygen atom and the HX deviates slightly from the furan ring

plane. The other geometry is the face-on type, with a hydrogen bond between the H atom

of HX and the π system of furan. For the furan – HF complexes only the atom-on

geometry is observed and for the furan –HI only the face-on type was found. Furan –HCl

and furan – HBr dimers exhibit both types of geometries.[174]

Compared to that, the FA – furan complexes A, H and I show atom-on geometries

while dimers B, C and D1 are face-on complexes. The geometries of the furan –

hydrogen halide dimers described by Huang were optimized at the same level of theory

(MP2/6-311++G(d,p)) than the FA – furan dimers. It is remarkable that while with others

basis sets the equilibrium geometries of dimers A and I are of Cs symmetry, at the

MP2/6-311++G(d,p) level of theory the geometries deviate from the Cs symmetry.

Similarly, the furan – hydrogen halide dimers deviate slightly from the C2v plane of the

furan ring.[174] With both smaller and larger basis sets a higher symmetry is found, which

indicates that the deviation from Cs symmetry might be an artifact of the 6-311++G(d,p)

basis set. At the same level of theory, the Cs geometry of the formamide molecule shows

one out of plane imaginary frequency resulting again in a slightly distorted structure.

132

Chan and Del Bene[114] carried out theoretical studies of the reactions of different acids

with furan at the MP2/aug’-cc-pVTZ//MP2/6-31+G(d,p) level of theory. They found that

hydrogen bonding is largely determined by electrostatic interactions, which generally

corresponds to hydrogen bonding to the site with the most localized negative charge.

Consequently, the furan molecule forms two hydrogen-bonded complexes with HF, one

through a lone pair at the oxygen atom, which is favored, and another dimer through the π

system at C2. They found a similar behavior for other weak acids interacting with furan.

Depending on the nature of the donor, the π or the O complex is more stable.[114]

Huang[174] obtained the atom-on geometry as the only minimum for the furan – HF dimer

and the face-on geometry exclusively for the furan – HI complex. This discrepancy

between the results of Huang and Del Bene is an evidence for the difficulties in

calculating the weak interacting complexes of furan.

Comparison with matrix isolation spectroscopy results

The experimental IR spectra of dimer A was assigned by comparison with calculations

at the MP2/aug-cc-pVDZ and MP2/6-311++G(d,p) levels of theory (Tables 4.7 and 4.8).

The red shift of the C=O stretching vibration of monomeric FA is characteristic for the

formation of complexes of FA. The experimental red-shift (-17cm-1) is in excellent

agreement with both the MP2/aug-cc-pVDZ and MP2/6-311++G (d,p) calculations for

dimer A, which predict shifts of -15.2 and -14.6 cm-1, respectively. The red-shift in the

C=O stretching vibration corresponds to an elongation of the C=O bond by 0.005 and

0.006 Å with the 6-311++G (d,p) and aug-cc-pVDZ basis sets, respectively, due to the

formation of the complex (Table 4.5). For dimer B the red shift is predicted to -12.7 cm-1,

for C to -13.2 cm-1, and for D1 to -13.5 cm-1 at the MP2/aug-cc-pVDZ level theory.

133

TABLE 4.7: The experimental (Ar matrix at 30 K – 35 K) and the calculated MP2/aug-cc-pVDZ and MP2/6-311++G(d.p)

vibrational frequencies (unscaled, in cm–1) of the FA – furan dimer A, along with the frequency shift in the complex, ∆ν, from

the monomer (in parentheses).

MP2 /aug-cc-pVDZ MP2 /6-311++G(d,p) M Experimental M Complex A M Complex A

Formic Acid

3550.5 3375.2 3365.3 3364.2

(-175.3) (-185.1) (-186.3)

3726.7 3543.9 (-182.8) 3797.7 3637.7 (-160.0) ν(O-H)

1767.3 1750.3 (-17.0) 1771.0 1755.8 (-15.2) 1807.6 1793.0 (-14.6) ν(C=O)

1103.5 1150.4 1153.0

(46.9) (49.5)

1115.7 1169.2 (53.5) 1142.7 1193.9 (51.2) ν(C-O)

Furan 1177.7 1167.7 (-10.0) 1218.9 1207.9 (-11.0) 1249.9 1236.2 (-13.7) νas(C-O-C) 1065.0 1051.6 (-13.4) 1094.0 1081.4 (-12.6) 1112.5 1095.8 (-16.7) νs(C-C)+ νs(C-O) 993.6 988.3 (-5.3) 1011.4 1006.2 (-5.3) 1025.5 1021.7 (-3.8) δ(CC-H) 869.1 874.8 (5.7) 865.6 873.3 (7.7) 879.9 885.9 (6.0) δs(C-O-C)

744.1 769.7 767.9 764.8

(25.6) (23.8) (20.7)

742.6 758.3 (15.7) 732.5 748.9 (16.4) γ(CC-H)

134

TABLE 4.8: Calculated vibrational frequencies (in cm–1) of FA – furan

complexes B - D1 by MP2/aug-cc-pVDZ level theory, along with the frequency shifts

in the complexes ∆ν, from the monomer (in parentheses).

MP2 /aug-cc-pVDZ M Complex B Complex C Complex D1

Formic Acid 3726.7 3580.8 (-145.9) 3565.7 (-161.0) 3577.4 (-149.3) ν(O-H) 1771.0 1758.2 (-12.8) 1757.8 (-13.2) 1757.4 (-13.6) ν(C=O) 1115.7 1134.8 (19.1) 1140.7 (25.0) 1141.8 (26.1) ν(C-O) Furan 1218.9 1218.4 (-0.5) 1222.8 (3.9) 1217.3 (-1.6) νas(C-O-C)

1094.0 1094.7 (0.7) 1099.2 (5.2) 1094.5 (0.5) νs(C-C)+ νs(C-O)

1011.45 1011.47 (0.02) 1013.3 (1.9) 1011.6 (0.2) δ(CC-H) 865.6 865.5 (0.1) 865.9 (0.3) 865.6 (0) δs(C-O-C) 742.6 746.9 (4.3) 757.7 (15.1) 744.5 (1.9) γ(CC-H) 770.0 27.4

For the OH stretching vibration of FA, the calculations at the MP2/aug-cc-pVDZ and

MP2/6-311++G (d,p) levels of theory for dimer A predict red shifts of -182.7 cm-1 and

-159.9 cm-1, respectively. In particular the results obtained with the aug-cc-pVDZ basis

are in excellent agreement with the experiment (-185 cm-1) (Table 4.7). The formation of

a complex results in an elongation of the O-HFA bond length in dimer A (compared to the

FA monomer) by 0.008 and 0.009 Å with the 6-311++G(d,p) and aug-cc-pVDZ basis

sets, respectively (Table 4.5). For dimer B the red-shift is calculated to -145.8 cm-1, for C

to -160.9 cm-1, and for D1 to -149.4 cm-1 (aug-cc-pVDZ basis, Table 4.8). The large

shifts of the OH stretching vibrations allow to discard the FA – furan complexes type (ii)

where the OH…O and O-H…π interactions are lacking and thus the OH stretching

vibration of the FA molecule is much less perturbed (Table 4.7).

The formation of hydrogen-bonded complexes of FA results in a contraction of the C-

OH bond compared to the monomer (0.010 and 0.012 Å with the 6-311++G(d,p) and

aug-cc-pVDZ basis sets, respectively, Table 4.5 ) and a blue shift of the corresponding C-

135

OH stretching vibration (observed at 1103.5 cm-1 in the unperturbed FA). For complex A

the predicted blue shift of 53.4 cm-1 (aug-cc-pVDZ) nicely matches the experimental

value of 46.9 cm-1. For complex B the shift is calculated to 19.1cm-1, for complex C to

24.9 cm-1, and for complex D1 to 26.1 cm-1 (Table 4.8), in much less agreement with the

experimental value.

The IR spectrum of the furan molecule in dimer A is less affected by the formation of

intermolecular complexes. The symmetrical and asymmetrical C-O-C stretching

vibrations of the furan ring are red-shifted by -13.4 cm-1 and by -10 cm-1, respectively.

These experimental shifts are in excellent agreement with the aug-cc-pVDZ calculated

values (Table 4.7). The CCH deformation mode δ(CCH) of monomeric furan at 993.6

cm-1 is shifted to 988.5 cm-1 in dimer A, which again is in good agreement with the

calculation (Table 4.7). The δ(COC) and γ(CCH) vibration modes are blue-shifted by 5.7

and 23.8 cm-1, respectively. The aug-cc-pVDZ calculations of complex A predict shifts of

7.7 and 15.7, respectively for these vibrations.

136

4.4. 1:2 Formic acid – acetylene complexes. Results and

discussion

Geometries and binding energies

Six complexes (A – G) corresponding to local minima between one molecule of formic

acid and two molecules of acetylene were found at the MP2/cc-pVTZ level of theory.

Four additional structures B1, E1, G1 and H1 were located using DFT theory (B3LYP/6-

311++G(d,p)) or MP2 with a smaller basis set (MP2/6-311++G(d,p)), but are not minima

at higher levels of theory (Figures 4.4 and 4.5). At all levels of theory complex A is

predicted to be the global minimum, followed by complexes B and C (Tables 4.9 – 4.11).

At the CCSD(T)/cc-pVTZ//MP2/cc-pVTZ level of theory after ZPE corrections the

calculated binding energy for complex A is -7.44 kcal/mol and the binding energies for

complexes B and C are -6.85 and -6.47 kcal/mol, respectively. Thus, three stable

complexes with binding energies within 1 kcal/mol are predicted.

The binding in complexes between formic acid and acetylene shows contributions from

the following five basic binding motives:

(1) OH…π interaction between the acidic H atom of formic acid and the π system of

acetylene;

(2) CH… π interaction between the formyl H atom of formic acid and the π system of

acetylene;

(3i) CH… π interaction between one acetylene H atom and the π system of the second

acetylene molecule;

(4) CH…O interaction between one acetylene H atom and the carbonyl oxygen atom

of formic acid;

(5) CH…O interaction between one acetylene H atom and the hydroxyl oxygen atom

of formic acid.

137

Figure 4.4. The calculated structures with hydrogen bond lengths and some

hydrogen bond angles of 1:2 formic acid – acetylene complexes A, B, C, and B1. a)

MP2/cc-pVTZ; b) MP2/6-311++G(3df,3pd); c) MP2/6-311++G(d,p); d) B3LYP/6-

311++G (d,p).

In the following the structure and binding energy of the complexes between one

molecule of formic acid and two molecules of acetylene with respect to these binding

contributions are discussed and quantified. To avoid confusion between both molecules

of acetylene in complexes the acetylene with the π system directly interacting with the O-

H group is named “Acetylene 1” and the second acetylene molecule “Acetylene 2”. In

complexes D, F and G the acetylene molecule with the π system interacting with the C-H

group of the formic acid is called “Acetylene 1”.

138

Figure 4.5. The calculated structures with hydrogen bond lengths and some

hydrogen bond angles of 1:2 formic acid – acetylene complexes: D, F, G, E1, G1 and

H1. a) MP2/cc-pVTZ; c) MP2/6-311++G(d,p); d) B3LYP/6-311++G(d,p).

139

At the MP2 level of theory the calculated binding energies and geometries are not very

dependent of the size of the basis set. For complex A, using the 6-311G++(d,p), cc-

pVTZ, and 6-311G++(3df,3pd) basis sets, the binding energies after ZPE corrections are

-7.64, -7.98, and -8.29 kcal/mol, respectively. B3LYP computations predict larger

binding energies compared to MP2. At all levels of theory and in accordance with

qualitative expectations, complexes D, F and G are less stable than the complexes A, B

and C (Table 4.9 and Figures 4.4 and 4.5). The MP2 geometries and binding energies

calculated with the cc-pVTZ and the 6-311++G(3df,3pd) basis sets are very similar.

However, the geometries obtained with the smaller 6-311++G(d,p) basis set differs

significantly from the cc-pVTZ results. The MP2/cc-pVTZ level of theory has been

established as very adequate for calculating weakly interacting systems as has been

confirmed by calculations of molecular systems with similar types interactions, like the

1:1 formic acid – acetylene complexes[116] and the acetylene dimers studies by

Karpfen.[166] The good performance of the cc-pVTZ basis set in these systems also

suggests to use this more economical basis set instead of the 6-311++G(3df,3pd) basis

set. For complexes A, B, and C the BSSE corrected binding energies at the MP2/cc-

pVTZ level of theory are also listed (Table 4.9).

TABLE 4.9: Calculated binding energies and ZPE corrected values of the 1:2

formic acid–acetylene complexes A – D, F, and G (in kcal/mol)

B3LYP/6-311++G(d,p) MP2/6-311G++(d,p) MP2/cc-pVTZ

ΔE ΔE (ZPE) ΔE ΔE (ZPE) ΔE ΔE (ZPE) ΔE (BSSE) A -6.58 -4.95 -9.23 -7.64 -9.52 -7.98 -7.87 B - -8.77 -7.18 -6.84 C -5.12 -3.80 -8.03 -6.63 -8.42 -7.05 -6.73 D -4.43 -2.98 -7.00 -5.27 -6.79 -5.47 - F -3.06 -1.85 -6.02 -4.33 -5.37 -4.32 - G - - -5.01 -3.65 -5.06 -3.93 -

140

TABLE 4.10: MP2 and CCSD(T) calculated binding energies for the complexes A,

B and Ca

MP2/6-311G++(3df,3pd) CCSD(T)/cc-pVTZ // MP2/cc-pVTZ

ΔE ΔE (ZPE) ΔE ΔE (ZPE)

A -9.83 -8.29 -8.98 -7.44 B -8.62 -7.03 -8.44 -6.85 C -8.53 -7.16 -7.84 -6.47

a ZPE corrections are from the MP2/cc-pVTZ calculations.

TABLE 4.11: Calculated binding energies and ZPE corrected values of 1:2 formic

acid–acetylene complexes B1, E1, G1, and H1 (in kcal/mol)

B3LYP/6-311++G(d,p) MP2/6-311G++(d,p) MP2/cc-pVTZ Complex

ΔE ΔE (ZPE) ΔE ΔE (ZPE) ΔE ΔE (ZPE)

B1 -6.08 -4.55 -7.91 -6.06 B E1 -4.15 -2.92 -6.63 -4.94 C G1 -3.12 -1.98 G H1 -4.48 -3.24 -5.03 -3.00 D

In complex A the acidic hydrogen atom of the formic acid molecule interacts with the π

system of one molecule of acetylene (contribution (1)). The MP2 calculated distance of

the C1 carbon atom of acetylene to the acidic hydrogen atom of formic acid (OH…C1

distance) is 2.293 Å (2.301 Å), and that of C2 (OH…C2 distance) is 2.297 Å (2.307 Å)

using a cc-pVTZ (6-311++G(3df,3pd)) basis set. The carbonyl oxygen atom of formic

acid interacts with one hydrogen atom of the second acetylene molecule (contribution

(4)) with a CH…O distance of 2.173 Å (cc-pVTZ, 2.171 Å with the 6-311++G(3df,3pd)

basis). Additional stabilization of the complex is due to the C-H…π interaction between

the two acetylene molecules (contribution (3), which corresponds to the “T” shape of the

acetylene dimer.[166] In this case, the T is distorted by interactions of both acetylene

molecules with the formic acid molecule. The acetylene – acetylene interaction is

characterized by a CH…C3 distance of 2,505 Å (2.496 Å) and a CH…C4 distances of

2,836 Å (2.815 Å) at the MP2/cc-pVTZ (MP2/6-311++G(3df,3pd)) level of theory. These

141

distances are in agreement with the expectations for weak hydrogen bonds[1] and compare

well with that of the undisturbed “T” shape acetylene dimer.[166, 167]

Complexes B and C provide further evidence for O-H…π, C-H…π and CH…O

interactions (Figure 4.4). In complex B the two molecules of acetylene do not interact

directly (thus, contribution (3) is absent), but both interact with the formic acid molecule

via contribution (4) and contribution (1). A type (4) interaction is found between the C1

H atom and the carbonyl oxygen atom. However, the long C1H…O distance of 2.635 Å

indicates that this interaction is rather weak. Even weaker is, with a distance of 3.418 Å,

the CH… π interaction of type (2) between the CH hydrogen atom of formic acid and the

acetylene π system.

Complex C shows an additional weak CH…O interaction between the hydroxyl oxygen

atom of the formic acid molecule and the hydrogen atom of one of the acetylene

molecules (contribution (5)). The other complexes - D, F and G - show C-H…π and

CH…O interactions; however, here the acidic hydroxyl hydrogen atom of the formic acid

molecule is not involved, resulting in weak overall binding energies.

The structures B1, E1, G1 and H1 are artifacts of the inability of the B3LYP functional

to deal with weak C-H…π dispersive interactions. Geometry optimization at the MP2/cc-

pVTZ level of theory results in the transformation of B1 to B, E1 to C, G1 to G and H1 to

D (Table 4.11, Figure 4.5). Thus, by comparison of complex B with structure B1 it is

evident that weak C-H…π interactions between the CH hydrogen atom of formic acid

and the π system of acetylene are not reproduced by B3LYP. This has been referred in

the study of the 1:1 formic acid- acetylene complexes.[116]

Intramolecular distances and vibrational frequencies

The intermolecular interactions in the complexes result in a distortion of the monomer

intramolecular distances and vibrational frequencies. Table 4.12 lists some selected bond

distances of the formic acid and acetylene monomer and the A, B, and C 1:2 formic acid–

acetylene complexes. For the monomers the experimental values [149, 193] are well

reproduced by calculations at the MP2/cc-pVTZ and MP2/6-311++G(3df,3pd) levels of

theory. The vibrational frequencies of the A, B and C complexes calculated at various

142

levels of theory are listed in Tables 4.13 – 4.15. The Dunning basis sets are expected to

be most reliable, since for the 1:1 complexes at this level an excellent agreement between

theory and experiment was found. [116]

TABLE 4.12: Comparison of selected intramolecular distances of the formic acid

and acetylene monomers (M), and 1:2 formic acid–acetylene complexes (A, B and

C)a

Exp MP2/cc-pVTZ MP2/6-311++G(3df,3pd) M M A B C M A B C

HCOOH r(C=O) 1.202(10)b 1.203 1.209 1.211 1.205 1.202 1.207 1.209 1.204 r(O─H) 0.972 (5)b 0.969 0.978 0.977 0.979 0.968 0.976 0.975 0.977 r(C─O) 1.343(10)b 1.346 1.333 1.332 1.342 1.343 1.331 1.331 1.339

C2H2 (acetylene1) r(C1≡C2) 1.203c 1.211 1.214 1.213 1.214 1.211 1.213 1.213 1.213 r(C1─H) 1.062c 1.061 1.067 1.064 1.064 1.062 1.067 1.064 1.064 r(C2─H) 1.063 1.063 1.064 1.063 1.063 1.064

C2H2 (acetylene2) r (C3≡C4) 1.213 1.213 1.213 1.213 1.212 1.212

r(C3─H) 1.066 1.067 1.064 1.067 1.067 1.065 r(C4─H) 1.062 1.062 1.062 1.062 1.062 1.062

a The distances are given in Å. b Reference[149] c Reference [193]

TABLE 4.13: Calculated vibrational frequencies (in cm–1) of the complex A and

frequency shifts relative to the isolated monomer M (in parentheses)

B3LYP/6-311++G(d,p) MP2/6-311++G(d,p) MP2/cc-pVTZ

M A M A M A 3738.0 3583.2 (-155) 3797.2 3672.7 (–125) 3763.4 3604.9 (-159) νO─Ha

1816.2 1794.0 (-22) 1807.5 1795.3 (–12) 1818.1 1800.1 (-18) νC=Oa 1125.3 1167.3 (+42) 1142.6 1184.0 (+41) 1136.7 1183.9 (+47) νC─Oa

3057.5 3054.3 (-3) 3132.3 3126.4 (-6) 3125.1 3119.2 (-6) νC─Ha 3419.8 3379.2 (-41)

3375.2 (-45) 3455.2 3420.4 (–35)

3429.8 (-25) 3446.1 3397.0 (-49)

3409.3 (-37) νC─Hb

772.7 801.3 (+29) 819.5 (+47) 832.1 (+59) 835.0 (+62)

766.3 789.7 (+23) 795.2 (+29) 809.9 (+44) 828.8 (+63)

753.0 786.9 (+34) 791.8 (+39) 802.0 (+49) 806.3 (+53)

δCCHb

a Formic acid modes in the complex. b Acetylene modes in the complex.

TABLE 4.14: Calculated vibrational frequencies (in cm–1) and frequency shifts

relative to the isolated monomer (in parentheses) of the complexes B and B1

B3LYP/6-311++G(d,p) MP2/6-311++G(d,p) MP2/cc-pVTZ

M B1 M B1 M B 3738.0 3586.4 (-152) 3797.2 3681.6 (–116) 3763.4 3618.4 (-145) νO─Ha

1816.2 1786.1 (-30) 1807.5 1788.6 (–19) 1818.1 1787.3 (-31) νC=Oa 1125.3 1172.8 (+48) 1142.6 1187.6 (+45) 1136.7 1186.7 (+50) νC─Oa

3057.5 3057.7 (+0.2) 3132.3 3128.3 (-4) 3125.1 3121.4 (-4) νC─Ha 3419.8 3370.6 (-49)

3404.1 (-16) 3455.2 3417.9 (–37)

3442.0 (-13) 3446.1 3393.7 (-52)

3429.4 (-17) νC─Hb

772.7 791.7 (+19) 795.8 (+23) 838.5 (+66) 841.0 (+68)

766.3 780.4 (+14) 782.6 (+16) 869.2 (+101) 872.6 (+106)

753.0 767.6 (+15) 785.7 (+33) 796.2 (+43) 816.9 (+64)

δCCHb

a Formic acid modes in the complex. b Acetylene modes in the complex.

144

TABLE 4.15: Calculated vibrational frequencies (in cm–1) of complex C and

frequency shifts relative to the isolated monomer (in parentheses)

B3LYP/6-311++G(d,p) MP2/6-311++G(d,p) MP2/cc-pVTZ

M C M C M C 3738.0 3562.6 (-175) 3797.2 3650.7 (–147) 3763.4 3580.4 (-183) νO─Ha

1816.2 1806.2 (-10) 1807.5 1798.8 (–9) 1818.1 1804.2 (-14) νC=Oa 1125.3 1153.0 (+28) 1142.6 1168.9 (+26) 1136.7 1171.5 (+35) νC─Oa

3057.5 3048.1 (-9) 3132.3 3121.2 (-11) 3125.1 3114.9 (-10) νC─Ha 3419.8 3396.0 (-24)

3400.0 (-20) 3455.2 3434.4 (-21)

3435.7 (–20) 3446.1 3422.6 (-24)

3424.2 (-22) νC─Hb

772.7 790.0 (+17) 797.4 (+25) 805.2 (+33) 813.4 (+41)

766.3 779.9 (+14) 788.2 (+22) 800.9 (+35) 802.2 (+36)

753.0 767.5 (+15) 771.2 (+18) 789.9 (+37) 794.3 (+41)

δCCH b

a Formic acid modes in the complex. b Acetylene modes in the complex.

The most perturbed vibrational modes in complexes A, B and C are the O-H stretching

vibrations of the formic acid molecules (Tables 4.13 – 4.15). At the MP2/cc-pVTZ level

of theory the frequency shifts in the complexes are -159, -145, and -183 cm-1, for the

trimers A, B, and C, respectively. This reflects the structures of complexes A, B and C

(Fig. 4.4) which all exhibit strong interactions between the OH hydrogen atom and the π

system of Acetylene 1. In complex C, with the largest frequency shift (-183 cm-1), an

additional interaction of the OH oxygen atom with one hydrogen atom of Acetylene 2 is

found. The formation of the complexes results in an elongation of the OH bonds of nearly

0.01 Å (MP2/cc-pVTZ, Table 4.12)

The carbonyl stretching frequencies of the formic acid molecules are predicted to be

shifted by -18, -31, and -14 cm-1 in the complexes A, B, and C, respectively. This again

reflects the CO…H interaction between the carbonyl oxygen atom and the acetylene H

atom (contribution (4)). In complex B, with a red shift (-31 cm-1) almost twice as large as

in complexes A and C, there is an additional CO…H interaction of the carbonyl group

with the H atom of Acetylene 2. This interaction results in a larger increase of the C=O

bond length in B compared to A and C.

145

The C-OH stretching modes of formic acid are blue-shifted in the complexes, and the

C-OH formic acid bond lengths are consequently shorter with respect to the monomer.

Again, trimer B shows a larger blue shift (+50 cm-1) than complexes A and C ( +47 cm-1

and +35 cm-1, respectively) which can be associated with the very weak CH…π

interaction between the CH group of the formic acid and the π system of Acetylene 2

(contribution (2)).

The C-H stretching modes of the acetylene moieties are also perturbed in the

complexes. At the MP2/cc-pVTZ level of theory these bands in trimer A are red-shifted

by -49 and -37 cm-1, in complex B by -52 and -17 cm-1, and in complex C by -24 and -22

cm-1.

Comparison with matrix isolation spectroscopy results

A mixture of aggregates was found in the matrix isolation experiments with FA –

acetylene. The main constitutes of these mixtures had been identified previously: the

acetylene dimer,[194],[195],[196] formic acid dimer[197],[7],[198] and 1:1 complex between

acetylene and formic acid.[116] However, the careful analysis of the IR absorptions

revealed the formation of several new bands which could not be assigned to any of the

previously known species (Table 4.16).

At higher acetylene concentrations two new peaks were observed in the O─H stretching

region at 3395.0 and 3384.0 cm−1. Since the intensity of these bands increase with

increasing acetylene concentration more than that of the 1:1 complexes they are

tentatively assigned to the O─H stretching vibration of formic acid in a 1:2 complex of

formic acid and acetylene. The experimental frequency shift (-155cm−1) is in excellent

agreement with the calculated frequency shifts (-159 cm−1 with MP2/cc-pVTZ and -155

with B3LYP/6-311++G(d,p)) for complex A (Table 4.16).

146

TABLE 4.16: Experimental (Ar matrix at 30 K – 10 K) and calculated (MP2/cc-

pVTZ and B3LYP/6-311++G(d,p)) vibrational frequencies (in cm–1) of the formic

acid–acetylene 1:2 complex A and frequency shifts ∆ν in the complex relative to the

isolated monomers (in parentheses)

Calculated frequencies Experimental frequencies MP2/ cc-pVTZ B3LYP/6-

311++G(d,p) M Complex A M Complex A M Complex A

3550.4 3395.0 (-155) 3763.4 3604.9 (-159) 3738.0 3583.2 (-155) νO─Ha

1767.1 1747.1 (-20) 1818.1 1800.1 (-18) 1816.2 1794.0 (-22) νC=O a 1103.5 1145.4 (+42) 1136.7 1183.9 (+47) 1125.5 1167.3(+42) νC─Oa

3288.8 3244.5 (-44) 3236.9 (-52)

3446.1 3409.3 (-37) 3397.0 (-49)

3419.8 3379.2 (-41) 3375.2 (-45)

νC─H b

a Formic acid modes in the complex. b Acetylene modes in the complex.

Figure 4.6. Matrix isolation IR spectra in the O─H stretching region of formic

acid. a: HCOOH : Ar = 1:600. b: HCOOH : C2H2 : Ar = 1:1:600. c: HCOOH :

C2H2 : Ar = 1:2:600. [145]

In the C=O stretching region increasing acetylene concentration results in the increase

of the intensity of the band at 1747 cm−1 assigned to the less stable, unsymmetrical

147

formic acid dimer. In addition, this absorption broadens significantly. This indicates that

in addition to the formic acid dimer a new species assigned to the 1:2 complex of formic

acid and acetylene is formed under these conditions. The intensity of this band depends

on the acetylene concentration, in agreement with the assignment to the C=O stretching

vibration of the 1:2 complex A. The good agreement between the experimental frequency

shift (-20 cm−1) and the calculated values (-18 cm−1 with MP2/cc-pVTZ and -22 with

B3LYP/6-311++G(d,p)) further confirms the formation of complex A.

Figure 4.7. Matrix isolation IR spectra in the C=O stretching region of formic

acid. a: HCOOH : Ar = 1:400. b: HCOOH : C2H2 : Ar = 1:1:600. c: HCOOH :

C2H2–Ar (1/2/600). FAD: non-symmetrical (acyclic) formic acid dimer; FCD: cyclic

formic acid dimer. [145]

In the C-O stretching region a new band appears at 1145.4 cm−1 at higher acetylene

concentration. By comparing with the calculated frequencies this band is also assigned to

the 1 : 2 complex A. The experimental frequency shift (+42 cm−1) is in excellent

agreement with the B3LYP/6-311++G(d,p) value (+42) and the MP2/cc-pVTZ (+47)

value.

148

Figure 4.8. Matrix isolation IR spectra in the C─O stretching region of formic

acid. a: HCOOH : Ar = 1:600. b: HCOOH : C2H2 : Ar = 1:1:600. c: HCOOH :

C2H2 : Ar = 1:2:600.[145]

The experimental frequency shifts for the acetylene streching (-44 and -52 cm−1) are in

reasonable agreement with the calculated shifts for complex A (-37, -49 cm−1 with

MP2/cc-pVTZ and -41, -45 cm−1 with B3LYP/6-311++G(d,p), Table 4.16). In the CCH

bending region of acetylene the 1 : 2 complex absorptions are too weak to be observed.

Analysis of the intermolecular interactions in the trimers

To quantify the contributions of intermolecular interactions in the trimers A, B, and C

in each of the trimers one of the three monomers (formic acid or one of the two acetylene

molecules) is subsequently removed. The energies of the remaining partial structures

(remaining dimers) were calculated (MP2/cc-pVTZ) in the geometries of the parent

trimers (Figure 4.9). These partial structures are then compared with the optimized

dimers to analyze the influence of the third molecule in the trimer on the dimer structures.

From that, a detailed picture of the non-covalent interactions in an aggregate consisting of

three components is achieved.

149

Figure 4.9. Dimers of complex A. Partial structure (i): Formic acid and Acetylene 1.

Partial structure (ii): Formic acid and Acetylene 2. Partial structure (iii): Acetylene1

and Acetylene 2.

In trimer A the first partial structure is formed by removing Acetylene 2 and thus

consists of formic acid and the remaining Acetylene1 (partial structure (i)) interacting via

noncovalent bond contribution (1). Analogously, partial structure (ii) is formed by

removing Acetylene 1 and consists of formic acid and Acetylene 2 interacting via

contribution (4). Finally, partial structure (iii) results from removing the formic acid

molecule and represents the (more or less distorted) “T” shaped acetylene dimer

interacting via contribution (3). The partial structures (i) – (iii) in the trimers B and C are

formed analogously by subsequently removing Acetylene 2, Acetylene 1, and the formic

acid molecule.

Several non-additive contributions have to be taken in to account, including basis set

superposition errors and other non-conventional interactions that may contribute to the

stabilization of the trimers (Table 4.16). These non-additive contributions are obtained by

subtracting all dimer binding energies from the trimer binding energy. As can be seen

150

from the Table 4.16 the non-additive contributions are small and never exceed 5% of the

total binding energy.

TABLE 4.16: MP2/cc-pVTZ energies of the trimers A, B, and C and their partial

structures (i), (ii), and (iii) (in kcal/mol)

MP2/cc-pVTZ

A B C FA – acetylene 2:1 Complex E(t) -9.52 -8.77 -8.42

Partial structure (i), E(i) -4.62 (48.5%) -5.18 (59.1%) -4.97 (59.0%) Partial structure (ii), E(ii) -2.67 (28.1%) -3.33 (38.0%) -1.73 (20.5%) Partial structure (iii), E(iii) -1.39 (14.6%) +0.11 (1.2%) -1.35 (16.0%) E(t) - (E(i)+ E(ii)+ E(iii)) -0.84 (8.8%) -0.37 (4.2%) -0.37 (4.4%)

In all complexes partial structure (i) contributes most to the trimer energies. The

“strongest” O-H…π interaction between the formic acid and Acetylene1 (contribution

(1)) dominates the interaction in the trimer. In trimers B and C, where partial structure (i)

contributes 59% to the total binding energy, an additional CH…O interaction between the

carbonyl oxygen atom and one of the H atoms of Acetylene 1 can be found (contribution

(4)). This increases the binding energy of this partial structure in the trimers B and C

compared to A (48.5%).

Partial structure (ii) contributes differently to each trimer: 28.1% in complex A (only

contribution 4), 38.0% in complex B and 20.5% in complex C. In B the interaction of

type (4) is most important, but in this case it is accompanied by a very weak interaction

of type (2).

In complex B a type (3) interaction between the acetylene molecules is not possible,

only a very small repulsive interaction (1.2%) is observed. This is a consequence of

repulsions between the closest hydrogen atoms in the two acetylene molecules.

For complex A it is interesting to compare the calculated binding energies and

geometries of the acetylene dimer from partial structure (iii) with the well known C2v

acetylene dimer (Table 4.17).[166] The presence of formic acid results in a small

destabilization of the partial structure (iii) compared to the non distorted acetylene dimer

151

by around 0.25 kcal/mol. For the C2v acetylene dimer the CH…C3 (and CH…C4)

distance is 2.727Å at the MP2/cc-pVTZ level of theory, while for the partial structure (iii)

the CH…C3 distance is 2.505Å and the CH…C4 distance 2.836Å.

TABLE 4.17: Calculated binding energies (in kcal/mol) of the acetylene dimer

(partial structure (iii) from complex A) and the C2v acetylene dimer

MP2/6-311++G(3df,3pd) MP2/cc-pVTZ Acetylene dimer (complex A, partial structure (iii))

-1.61 -1.39

C2v acetylene dimera -1.86 -1.63 a Reference[166]

Figure 4.10 shows the FA – acetylene dimers (left hand side) compared to several

partial structures (i) and (ii) (right hand side). A comparison between these structures

reveals interesting similarities between the partial structures and the dimers. Thus, the O-

H…π bidentate 1:1 complex is very similar to the partial structure (i), and the structures

of the CH…O=C bidentate and monodentate 1:1 complexes agree well with the partial

structures (ii) of the B and B1 trimers, respectively.

152

Figure 4.10. Comparison of 1:1 complexes of formic acid – acetylene with the partial

structures of the 1:2 complexes. Left side: 1:1 complexes (MP2/6-311++G(d,p); right

side: Partial structures (i) and (ii) from the 1:2 complex B (MP2/cc-pVTZ) and

partial structure (ii) from the 1:2 complex B1 (MP2/6-311++G(d,p)).

4.5. Conclusion

Furan is a heterocyclic system that is able to accept hydrogen bonds both in the

molecular plane at the basic oxygen atom and perpendicular to the molecular plane at its

electron rich π-system. Formic acid, on the other hand, can act as hydrogen bridge donor

either via its acidic OH group or via the less acidic CH group. In addition, it can serve as

hydrogen bridge acceptor via its two oxygen atoms. The combination of furan and formic

acid thus results in a variety of non-covalently bound dimers which were systematically

studied.

FA – acetylene is an interesting system to study non-covalent interactions, since a

variety of “non-classical” hydrogen bonds can be formed where four acidic hydrogen

atoms (OH and CH at formic acid, two CH at acetylene) compete for the two oxygen

atoms in formic acid and the acetylene π-system as hydrogen bond acceptors.

The MMH method was used to localize and characterize the FA – furan and 1:2 FA –

acetylene structures. Six 1:2 FA – acetylene complexes with binding energies between -

153

3.93 and -7.98 kcal/mol (MP2/cc-pVTZ + ZPE) are identified. Since the introduction of a

further acetylene molecule in the trimer complexes results in a number of energetically

close lying complexes, the three most strongly bound 1:2 FA – acetylene complexes are

found within a range of 1 kcal/mol. The binding interactions in these complexes are O-

H…π, C-H…π and CH…O interactions which can be classified as weak hydrogen bonds.

Nine FA – furan complexes with binding energies between -3.91 and -0.82 kcal/mol

were identified as minima at the MP2/6-311++G(d,p) + ZPE + BSSE level of theory.

Another five weaker bound complexes are minima at lower level of theory only. The FA

– furan dimers are classified into two types: type (i) with the OH hydrogen atom of FA

interacting with the furan molecule and type (ii) where the main interactions are via the

CH hydrogen atom of formic acid. Due to the lower acidity of the CH hydrogen atom,

type (ii) complexes are less stable than type (i) complexes.

The binding energy of the most stable complex A is -3.91 kcal/mol (MP2/6-

311++G(d,p) + ZPE+ BSSE). Although the OH…O interaction (1) is dominating in

dimer A, the secondary C=OFA…HF interaction (2) between the carbonyl oxygen atom of

FA and the hydrogen atom of furan leads to an additional significant stabilization of this

complex. The matrix isolation experiments reveal that dimer A is the major – if not only

– complex formed if the two monomers are allowed to diffuse slowly in solid argon. The

additional IR absorptions that appear in the matrix spectra under these conditions nicely

match the theoretical predictions for complex A.

FA – furan complexes B, C, and D1 are π complexes defined by the absence of the

strong in-plane OH…O hydrogen bond. These dimers are stabilized by the O-HFA…π

interaction (2) between the OH hydrogen atom of FA and the π system of furan. With

-2.24, -2.12 and -2.37 kcal/mol (MP2/6-311++G(d,p) + ZPE+ BSSE, Table 1) the

binding energies are considerably smaller than that of A, and consequently these

complexes are not identified experimentally.

The second group of FA – furan dimers is mainly stabilized by the very weak CH…O

or CH…π interaction. The most stable dimers in this group are H and I with binding

energies of -1.94 and -1.35 kcal/mol, respectively at the MP2/6-311++G(d,p) + ZPE+

BSSE level of theory. The higher stability of dimer H can be attributed to an additional

154

C=OFA…HF interaction between the carbonyl oxygen atom of FA and one of the

hydrogen atoms of furan. Dimer I exhibits the less stabilizing interaction HOFA…HF

between the OH oxygen atom of FA and one of the hydrogen atoms of furan. Dimers E,

F1, and G are very weakly bound CH…π complexes with binding energies between -0.96

and -0.82 kcal/mol (MP2/6-311++G(d,p) + ZPE+ BSSE).

The MP2 level of theory with the 6-311++G(d,p) and aug-cc-pVDZ basis sets provides

reliable geometries for FA – furan complexes. With the small double zeta basis sets

without augmentation the structures of five additional very weak FA – furan complexes

could be localized. Introducing the BSSE corrections during the geometry optimization at

the MP2/6-31G(d,p) level of theory leads to large changes of the calculated geometries of

some of the very weak FA – furan complexes. Here, clearly, BSSE and variations of the

basis sets have the largest effects on the dimers.

1:2 FA – acetylene complexes were analyzed at the B3LYP and MP2 levels of theory

with large basis sets. The deficit of the B3LYP method and MP2 calculations with

smaller basis sets to account for intermediate range dispersive interactions results in

producing artificial minima B1, E1, G1, and H1 which disappear at the higher level of

theory.

In 1:2 FA – acetylene complexes the interaction between the acidic H atom of formic

acid and the acetylene π-system (partial structure (i)) provides 50 – 60% of the binding

energies of the most stable complexes A – C, and thus dominates the non-covalent

interactions in these systems. Partial structures (ii) and (iii) add between 15 and 29% to

the total binding energy. Only in complex B, where no “T-acetylene” interaction is

present, contribution (iii) is slightly repulsive. Other interactions are less important for

the stabilization of the complexes, but might influence the structure. Thus, complex B is

only slightly stabilized by an additional very weak CH...π interaction compared to B1.

This interaction results in a considerable structural change. The calculated vibrational

frequencies of the most stable 1:2 complex A are in good agreement with the

experimental values, indicating that under the conditions of matrix isolation indeed the

complex A is formed.

155

5. Acetylene Complexes with Oxygen Heterocycles. An Outlook

5.1. Introduction

The interactions of the formic acid molecule with acetylene, as well as the structure of

the FA – furan dimers, were already discussed in a previous chapter. The investigation of

the interaction between acetylene and furan is the next logical step to provide a deeper

insight into the characterization of weak interacting complexes. Therefore, the

interactions of acetylene with different types of oxygen heterocycles like furan,

tetrahydrofuran (THF), and 1,4-dioxane are investigated. The study of the acetylene –

furan complexes is a very interesting and challenging task due to the competition

between very weak CH…π and CH…O interactions. The acetylene – furan dimers are

compared to the acetylene – THF and acetylene – 1,4-dioxane complexes.

“Twist” and “envelope” conformers have been identified for the THF molecule[199-201]

(Figure 5.1). The twist conformation is described by Cremer and Pople as the most stable

form with a low barrier to pseudorotation into an envelope form[199] .

Figure 5.2. Twist and envelope conformers of THF [199]

X-ray analysis indicated that THF has a twist conformation in the crystal structure,

according to Luger and Buschmann[200] (Figure 5.2). The microwave spectrum of THF

156

has been studied by Engerholm and coworkers and their results indicated that the twisted

conformation is at lower energy than the bent conformer.[201]

Figure 5.2. Twist conformer of the THF in crystal structure. Figure taken from

“Twist conformation of tetrahydrofuran in the crystal form” by P Luger and J.

Buschmann.[200]

Cadioli et al. studied one twisted (C2), one envelope (Cs), the C2v and two asymmetric

(C1) conformations of THF at the HF and MP2 levels of theory with different basis sets.

Their most reliable computations show the twist conformation being the absolute energy

minimum, the envelope structure being a transition state, only 0.3 kcal/mol higher, and

the C2v, being an energy maximum, 4.7 kcal/mol high.[202] However, in a recent ab initio

study on the pseudorotation motion in tetrahydrofuran, Rayón and Sordo extensively

explored the conformational potential energy surface of tetrahydrofuran at the MP2 level,

using double and triple ζ Dunning´s correlation consistent basis sets. They predict that the

equilibrium conformation of tetrahydrofuran is an envelope Cs structure.[203]

Alonso, Lopez et al. studied seven isotopomers of the hydrogen bonded heterodimer

THF…HF using Fourier transform microwave spectroscopy.[204] They concluded that

there is a pseudorotation of the THF subunit of the complex. The spectroscopic

parameters are interpreted in terms of a geometry in which THF has a conformation close

to the twisted ring form, with HF lying in the plane bisected by the COC ring angle.[204]

157

Ten stationary points as energy minima or transition states for the 1,4-dioxane were

characterized by Chapman and Hester at the HF and DFT levels of theory using the 6-

31G* basis set (Figure 5.3).[205] They found that the chair conformation is the lowest in

energy, followed by the two twist-boats. A half-chair structure, in which four of the ring

atoms lie in a plane, was found to be the transition state connecting the chair and the

twist-boats.

Figure 5.3. Eight conformations of the 1,4-dioxane and their calculated energies.

Figure taken from “Ab initio conformational analysis of 1,4-dioxane” by D.M

Chapman and R.E Hester.[205]

158

5.2. Computational methods

The Multiple Minima Hypersurface (MMH) approach was used for searching

configurational minima in the acetylene – furan, acetylene – THF and acetylene – 1,4-

dioxane dimers. In each case, one thousand randomly arranged structures were generated

as starting points, and the resulting geometries were optimized and analyzed using the

PM3 semiempirical quantum mechanical Hamiltonians. For the acetylene – furan dimers,

the AM1 semiempirical geometries were also analyzed. In all cases, the relevant

configurations were further refined using ab initio methods at various levels of theory.

The ab initio computations were performed using the Gaussian 98 and Gaussian 03

programs. The equilibrium geometries and vibrational frequencies were calculated with

tight convergence criteria at the SCF level including second order Møller−Plesset

perturbation theory, MP2. The force constants where calculated when necessary. Pople’s

6-31G(d,p) and 6-311++G(d,p) basis set as well as augmented and non augmented

Dunning’s correlation consistent double and triple ζ basis sets (cc-pVDZ, aug-cc-pVDZ

and cc-pVTZ) were used. Here, only the results with the 6-31G(d,p), 6-311++G(d,p) and

cc-pVTZ basis sets are discussed. Vibrational frequencies were calculated at all levels of

theory, except at the MP2/cc-pVTZ for the acetylene – THF and acetylene – 1,4-dioxane

dimers.

The stabilization energies were calculated by subtracting the energies of the monomers

from those of the complexes and ZPE corrections are included. The energies were also

corrected for the basis set superposition errors (BSSE) using the counterpoise (CP)

scheme of Boys and Bernardi.

To investigate the influence of the basis set superposition errors (BSSE) on the

geometries of the complexes, the acetylene – furan dimers were optimized at the MP2/6-

31G(d,p) level of theory using CP corrections during the optimization process. In

addition, the geometries were optimized without BSSE at the same level of theory to

compare the influence of the BSSE on the binding energies as well as on the geometries.

159

5.3. Acetylene – furan dimers. Results and discussion

After refining the MMH results, five acetylene – furan dimers A – E were localized at

the MP2/6-311++G(d,p) level of theory with binding energies between -0.75 and -1.77

kcal/mol (MP2/6-311++G(d,p) + BSSE) (Table 5.1). Here the geometries and binding

energies of the complexes are discussed at this level of theory.

TABLE 5.1: Calculated binding energies of the acetylene – furan dimers A – E at

the MP2 level of theory with the 6-311++G(d,p) and cc-pVTZ basis sets, including

BSSE and ZPE corrections (in kcal/mol)

MP2 6-311++G(d,p cc-pVTZ ΔE ΔE BSSE ΔE BSSE+ZPE ΔE ΔE BSSE ΔE BSSE+ZPE

A -3.51 -1.77 -0.55 -3.11 -2.45 -2.00

B -2.71 -1.76 -1.31 -2.60 -1.95 -1.46 B1 -a - a C -2.29 -1.08 -0.51 C1 b C1 C b -1.92 -1.45 -1.09 D -1.98 -0.83 -0.35 -1.39 -1.09 -0.83 E -1.91 -0.75 -0.28 C1 b

a Transition state b Geometry that was found after optimization

Three basic types of interactions (1) – (3) are differentiated in the acetylene – furan

complexes: (Figure 5.4)

(1) C-Hacet…OF interaction between the hydrogen atom of acetylene and the oxygen

atom of furan.

(2) C-Hacet…π interaction between the hydrogen atom of acetylene and the π system

of furan.

(3) C-HF…π interaction between the hydrogen atom of furan and the π system of

acetylene.

160

Figure 5.4. The calculated structures with hydrogen bond lengths (Å) of the

acetylene – furan complexes A – E at the (a) MP2/6-311++G(d,p) and (b) MP2/cc-

pVTZ levels of theory

Dimer A is the most stable complex (binding energy -1.77 kcal/mol (BSSE corrected,

Table 5.1)) and the only structure that was found as a minimum at all levels of theory

(including the cc-pVDZ and aug-cc-pVDZ basis set, which are not discussed here).

Dimer A is stabilized by interaction (2) between one hydrogen atom of acetylene and the

π system of furan (Figure 5.5). This hydrogen atom shows a distance of 2.725 Å to the

oxygen atom of furan (interaction (1)) (Figure 5.4). The geometry of dimer A resembles

the benzene – acetylene dimer described by Steiner el at in their crystallographic,

161

spectroscopic and quantum mechanical studies on the C-H…π hydrogen bonding in

terminal alkynes[206] (Figure 5.5). Dimer A shows an additional stabilization, not present

in the benzene – acetylene system, due to the C-H…O interaction (1). Consequently, the

acetylene molecule is not perpendicular to the aromatic ring.

Figure 5.5. Acetylene – Benzene and acetylene – furan dimers. The acetylene –

benzene figure was taken from “An Introduction to Hydrogen Bonding” by G.A Jeffrey[1]

and the original source is ref[206]. The acetylene – furan dimer parameters are referred to

the furan plane at the MP2/6-311++G(d,p) level of theory.

The binding energy of dimer B is -1.76 kcal/mol and it is basically the same that for

dimer A (at this level of theory and BSSE corrected). Dimer B has a planar Cs geometry

and is stabilized by interaction (1) with C-H…O distance 2.387 Å and interaction (3)

between the α-hydrogen atom of furan and the π system of acetylene at C-HF…Cacet 2.959

Å distance. With the cc-pVTZ basis set, including BSSE corrections, dimer B is 0.5

kcal/mol less stable than dimer A (Table 5.1). The geometry of complex B was localized

after following the imaginary vibration of dimer B1, which is a transition state with both,

the 6-311++G(d,p) and cc-pVTZ basis sets. Compared to B, dimer B1 has C2v symmetry

and is stabilized only by interaction (1) with shorter C-H…O hydrogen bond distances

and a hydrogen bond angle of 180º (Figure 5.4). At a lower level of theory, with the 6-

31G(d,p) basis set, in both, the standard and the BSSE free PES, dimer B could not be

located as a minimum. Instead, complex B1 was identified as the most stable minimum at

this level (Table 5.2).

Dimers C, D and E are stabilized only by interaction (3) between one hydrogen atom of

furan and the π system of acetylene and thus they are less stable. Dimer C (-1.08

162

kcal/mol) was located as a minimum only with the 6-311++g(d,p) basis set. The C-

HF…Cacet distance is large, even for a weak hydrogen bond (3.054 Å) and its geometry

suggests a stabilization via π…π stacking between the furan and the acetylene π systems.

With the cc-pVTZ basis set, dimer C is a transition state which leads to the geometry of

C1, where the acetylene molecule is not in a symmetric position with respect to the C1-

C2 bond of furan (Figure 5.4). The optimization of C1 with the 6-311++G(d,p) basis set

leads to the dimer C. With the 6-31G(d,p) basis set, the optimization of both C and C1

produces the dimer A. When BSSE corrections are introduced in the geometry

optimization, dimer C is a transition state and the optimization of C1 produces dimer A

(Tables 5.1 and 5.2).

TABLE 5.2: Calculated binding energies of the acetylene – furan dimers A – E at

the MP2 level of theory with the 6-31g(d,p) basis set, including (or not) BSSE

corrections during the geometry optimizations

MP2 6-31g(d,p)opt BSSE 6-31g(d,p) ΔE ΔE BSSE ΔE ΔE BSSE

A -2.73 -1.64 -2.97 -1.39

B B1b B1 b B1 -2.99 -1.76 -3.04 -1.71 C - a A b C1 A A b D -1.52 -0.81 -1.62 -0.72 E -1.35 -0.68 -1.45 -0.58

a Transition state b Geometry that was found after optimization

Dimers D and E are very weakly interacting complexes with binding energies of -0.83

and -0.75 kcal/mol, respectively. They show very similar geometries, in dimer D the C-

HF…Cace distance is 2.796 Å and in dimer E, 2.823 Å (Figure 5.4).The main difference

between these two complexes is which hydrogen atom of furan interacts with the

acetylene; in the case of dimer D, it is the α-hydrogen atom of furan; for dimer E, the β-

163

hydrogen of furan. With the cc-pVTZ basis set, the geometry optimization of dimer E

leads to the structure of dimer C1.

The similarity between the two most stable FA – furan complexes and the A and B

acetylene – furan dimers is noticeable. The FA – furan dimer A is stabilized by the

interactions of the oxygen and the α-hydrogen atoms of furan with the FA molecule,

similar to the acetylene – furan dimer B (Figures 5.4 and 5.6). In FA – furan dimer B, the

O-H hydrogen of FA interacts with the π system of furan; in acetylene – furan dimer A it

is the hydrogen atom of acetylene which interacts with the π system of furan.

Figure 5.6. FA – furan dimers A and B.

Therefore, at the same level of theory the comparison between the binding energies and

geometries in both systems provides preliminary conclusions about the relative strengths

of the interactions involved. In the FA – furan system, the O-HFA…OF and the

C=OFA…HF interactions define the dimer A as more stable compared to dimer B,

stabilized by the O-H…π interaction, whereas in the acetylene – furan system, dimer A

(C-Hacet…πF (interaction (2) and C-Hacet…OF ( interaction (1)) is slightly favored over

dimer B, stabilized by the C-Hacet…OF and C-HF…π interactions (1) and (3).

However, when ZPE corrections are also included at the MP2/6-311++G(d,p) level of

theory, the acetylene – furan dimer B is more stable than dimer A. The very low BSSE

and ZPE corrected binding energy of dimer A at this level of theory may be due to the

164

overestimation of the ZPE corrections by the model (harmonic oscillator) used, which

provides an unexpected high value of ZPE for dimer A. At this level of theory, the ZPE

calculated correction for dimer A is 1.75 kcal/mol, which represent 50% of the

uncorrected binding energy. At the MP2/cc-pVTZ level of theory the values of the

ZPE+BSSE corrected binding energies behave “normally” and dimer A is ~0.5 kcal/mol

more stable than dimer B.

The acetylene – furan dimers were optimized at the MP2/6-31G(d,p) level of theory

using CP corrections during the optimization process to investigate the influence of the

basis set superposition errors (BSSE) on the geometries of the complexes. The results

were similar to the standard optimization at the same level of theory (Tables 5.2 and 5.3).

The BSSE corrected binding energies for the BSSE geometry-corrected dimers are 0.05 –

0.25 kcal/mol larger compared to the standard values. The BSSE-corrected hydrogen

bond distances are 0.1 – 0.3 Å larger than the standard ones (Table 5.3)

TABLE 5.3: Selected intermolecular distances and angles for the acetylene – furan

dimers A, B1, D and E at the MP2 level of theory with the 6-31G(d,p) basis set,

including BSSE-CP corrections during the geometry optimizations

MP2/6-31G (d,p) optimization with BSSE optimization without BSSE A B1 D E A B1 D E C-Hacet…OF 2.981 2.358 – – 2.688 2.261 – –

C-Hacet…C1F 2.922 – – – 2.671 – – – Cacet…H-CF – – 3.008 3.073 – – 2.810 2.855 <C-Hacet…OF 134.6 180.0 – – 135.0 180.0 – –

<C-Hacet…C1F 143.5 – – – 143.9 – – – <Cacet…H-CF – – 165.5 168.6 – – 166.2 167.7

5.4. Acetylene – THF dimers. Results and discussion

Two acetylene – THF dimers were found after refining the MMH results (Figure 5.7).

In this case (and for the acetylene – 1, 4-dioxane dimers, too) only the PM3 results were

165

analyzed. With -3.66 and -3.48 kcal/mol binding energies, respectively (MP2/6-

311++G(d,p) + BSSE), acetylene – THF dimers A and B are very close in energy (Table

5.4). The THF subunit in dimer A shows the twist conformation, while in dimer B

resembles a slightly distorted “envelope”. Both dimers A and B are stabilized by the

OTHF…C-Hacet interaction with similar hydrogen bond distances and angles (Figure 5.7).

However, despite the hydrogen atoms which are closer to the acetylene molecule show C-

HTHF…Cacet distances larger than 3 Å, the position of the acetylene subunit in the

complex suggest an additional C-HTHF…π stabilizing interaction.

At the MP2/cc-pVTZ level of theory, the geometries and binding energies of dimers A

and B are very similar to the MP2/6-311++G(d,p) results. In both cases, dimer A is

slightly more stable than dimer B, which is in agreement with previous results that state

that the “twist conformation” of THF is more stable than the “envelope”.[199, 202] By our

calculations, the “envelope” monomer of THF is a transition state at both levels of theory

and is 0.17 and 0.14 kcal/mol less stable than the “twist” form. (MP2/6-311++G(d,p) and

MP2/cc-pVTZ, respectively). Therefore, it is noticeable how the formation of the

complex leads to the stabilization of the “envelope” form in the dimer, compared to the

isolated monomer. Another point to remark is that both geometries were localized after

optimization of the semiempirical geometries, which were produced without any previous

considerations of the conformation equilibrium of the THF monomer.

TABLE 5.4: Calculated binding energies of the acetylene – THF dimers A and B at

MP2 level of theory with the 6-311++G(d,p) and cc-pVTZ basis sets, including BSSE

corrections (in kcal/mol)

MP2 6-311++G(d,p cc-pVTZ ΔE ΔE BSSE ΔE ΔE BSSE

A -5.23 -3.66 -5.23 -3.92

B -5.23 -3.48 -5.14 -3.76

166

Figure 5.7. The calculated structures with hydrogen bond lengths (Å) of the

acetylene – THF complexes A and B at the (a)MP2/6-311++G(d,p) and (b)MP2/cc-

pVTZ levels of theory

TABLE 5.5: Selected intramolecular parameters, at the MP2/cc-pVTZ level of

theory, of the THF as isolated monomer and in the acetylene – THF dimers

MP2/cc-pVTZ THF “twist” THF “envelope” M Dimer A M Dimer B

Bond lengths

O-C1 1.431 1.435 1.420 1.425 C1-C2 1.522 1.519 1.533 1.526 C2-C3 1.526 1.527 1.545 1.543 Angles

<C1OC4 109.3 109.2 104.0 104.5 <C1C2C3 101.0 100.9 103.1 102.9 <C2C3C4 101.0 101.4 103.1 103.5

167

Compared to the monomer at the same level of theory, the “twist” and “envelope”

conformers of THF in the complexes show a distortion of the C2 and Cs symmetry,

respectively (Table 5.5).

For both complexes, vibrational frequencies were calculated at the MP2/6-311++G(d,p)

level of theory. The structures of the acetylene – THF dimers A and B are minima also

with others basis sets. (cc-pVDZ and aug-cc-pVDZ)

5.5. Acetylene – 1,4-dioxane dimers. Results and discussion

Two acetylene – 1,4-dioxane dimers A and B were found at the MP2/cc-pVTZ level of

theory, with BSSE corrected binding energies -3.22 and -3.06 kcal/mol, respectively.

(Figure 5.8, Table 5.6) At the MP2/6-311++G(d,p) level of theory, only the structure of

dimer A was found, since the geometry optimization of B leads to dimer A geometry.

Both dimers stabilize by the OTHF…C-Hacet interaction between the oxygen atom of 1,4-

dioxane and one hydrogen atom of acetylene. The main difference between both

complexes is the orientation of the acetylene molecule.

Figure 5.8. The calculated structures with hydrogen bond lengths (Å) of the

acetylene – 1,4-dioxane complexes A and B at the (a) MP2/6-311++G(d,p) and (b)

MP2/cc-pVTZ levels of theory

168

TABLE 5.6: Calculated binding energies of the acetylene – 1,4-dioxane dimers A

and B at MP2 level of theory with the 6-311++G(d,p) and cc-pVTZ basis sets,

including BSSE corrections (in kcal/mol)

MP2 6-311++G(d,p cc-pVTZ ΔE ΔE BSSE ΔE ΔE BSSE

A -4.54 -2.96 -4.46 -3.22

B A -4.33 -3.06

5.6. Conclusion

The geometries of the acetylene – furan, acetylene – THF and acetylene – 1,4- dioxane

dimers are calculated starting from randomly generated molecular arrangements using the

MMH procedure. Five acetylene – furan dimers with binding energies between -0.75 and

-1.77 kcal/mol are identified (MP2/6-311++G(d,p)+ BSSE). Two dimers of acetylene –

THF were found with binding energies -3.66 and -3.48 kcal/mol, (MP2/6-311++G(d,p)+

BSSE) respectively and the THF subunit in the “twist” and “envelope” conformations. At

the same level of theory, only one acetylene – 1,4-dioxane dimer (-2.96 kcal/mol) was

localized. However with other basis sets (e.g. cc-pVTZ) a second complex with a

different orientation of the acetylene molecule but very close in energy was identified.

In the acetylene – furan dimers two types of interactions are found: the C-H…O and

the C-H…π interaction. The C-H…π interaction appears in two variations, depending on

which molecule provides the hydrogen atom and which molecule the π system. The

results indicate that the C-H…π interaction between one hydrogen atom of acetylene and

the π system of furan is the stronger interaction, also compared to the in-plane C-H…O

interaction of the second more stable acetylene – furan dimer B. Acetylene – THF and

acetylene – 1,4-dioxane dimers are stabilized by the C-H…O interaction between the

oxygen atom of the heterocycle and the hydrogen atom of acetylene, however, the

position of the acetylene subunit in the complex suggests an additional C-HTHF…π

stabilizing interaction.

169

6. General Conclusion

Molecular interactions in formic acid complexes

The complexes formed by the s-trans conformer of the formic acid molecule with

formamide, dimethyl ether, furan and acetylene were studied. The variety of functional

groups shown in these molecules allows for a more detailed analysis of the

intermolecular interactions in formic acid complexes, from the moderate-“strong”

C=OFA… H-N and O-HFA…O=C interactions in FA – FMA dimers to the weak C-

HFA…π in FA – furan and 1:2 FA – acetylene complexes (Table 6.1).

The geometries of the most stable complexes in all cases show that the FA molecule

interacts strongly via its OH hydrogen atom and the carbonyl oxygen via O-HFA…O,

C=OFA…H and the O-HFA…π interaction which gets especial relevance in the

stabilization of the FA – furan and 1:2 FA – acetylene complexes. The O-HFA…O and

C=OFA…H interactions are responsible of the geometries of the most stable FA – FMA,

FA – DME and FA – furan dimers and the different stabilities of these complexes allow

to suggest a preliminary scale of strength of the interactions:

a) O-HFA…O=CFMA > O-HFA…ODME > O-HFA…OF

b) C=OFA...H-NFMA> C=OFA...HDME > C=OFA...HF

In accordance to that, the interactions of the OH hydrogen atom of FA are stronger with

a carbonyl oxygen atom than with an open-chain ether oxygen atom, and the weakest is

the O-H…OF interaction with the oxygen atom of an aromatic heterocycle like furan.

For the C=O…H interactions the most favored is the amide hydrogen atom, followed

by the methyl hydrogen of DME and the hydrogen atom of furan. Therefore, the FA

interacts weaker with the aromatic system than with the others. This can be

corroborated by an outlook into the acetylene – furan, acetylene – THF and acetylene –

1,4-dioxane dimers, which shows that the acetylene molecule also interacts more

strongly with saturated oxygen heterocycles than with the aromatic furan.

170

TABLE 6.1: Calculated binding energies of the most stable FA dimers at the MP2

level of theory with the cc-pVTZ Dunning’s basis set including BSSE corrections (in

kcal/mol).

MP2/cc-pVTZ FA – FMA ΔE ΔE (BSSE)

A -16.90 -14.21

B -13.21 -10.93

FA – DME A -12.23 -9.47

B -11.31 -9.09

FA - Furan A -7.14 -5.53

B -5.73 -4.23

The geometries of all the FA complexes show that not only the OH hydrogen atom and

the carbonyl oxygen atom of FA play an important role in the stabilization of the FA

aggregates. The aldehyde hydrogen and the OH oxygen atoms of FA contribute also to

the geometries of the FA – FMA, FA – DME, FA – furan and 1:2 FA – acetylene

171

complexes with their H-OFA…H, C-HFA…O (clearly not found in acetylene complexes)

and C-HFA…π interactions.

MMH for localizing the minima

The MMH approach, in combination with high level ab initio theory, was used to

localize and characterize the FA – FMA, FA – DME, FA – furan, 1:2 FA – acetylene,

acetylene – furan, acetylene – THF and acetylene – 1,4-dioxane structures. The

comparison with matrix spectroscopy, crystal structure data, and the analysis of studies of

similar complexes from the literature, confirms the quality of the MMH procedure as a

very useful tool for reliably localizing minima in hydrogen bonded complexes without

recurring to previous knowledge of the structure of supramolecular complexes.

One question is the reliability of using a semiempirical Hamiltonian for the preliminary

calculations of geometries, and if starting from semiempirical minima might result in

dropping some structures which are not equilibrium geometries at the semiempirical

level. The refinement requires the calculation of a large number of starting geometries

and to carefully select and analyze the semiempirical geometries. The comparison of

semiempirical an ab initio structures shows that many ab initio minima, which are not

minima at the semiempirical level, are found. In addition, in general, the amount of

minima decreases by increasing the level of theory, since many geometries, which are a

consequence of a poor optimization, converge into a limited set of final structures.

There are interesting similarities between the semiempirical and ab initio structures

(Figure 6.1). In many cases the “global” minimum with ab initio calculations was also

found at the semiempirical level as one of the most stable geometries. In the PM3

structures the artificial stabilization of the complex due to H…H interactions at ~1.7 –

1.75 Ǻ was found.[61, 62] For the FA – FMA dimers, the amide group was also optimized

to an out of plane geometry with the PM3 method. The FA – FMA PI and PII PM3

minima differ by the hydrogen bonding angles, which demonstrates the irregularities of

PM3 in reproducing the Cs symmetry of the complex. In this case, the AM1 method

shows better results, e.g. for the FA – furan complexes AM1 is able to predict the most

stable dimer as the “global” semiempirical minimum and also predicts the geometries of

others ab initio minima. But some hydrogen bond distances are much better described by

172

the PM3 hamiltonian, like in FA – FMA dimers PI and AI. The PM3 method is also able

to find as one of the more stable local minima, the structure of FA – furan dimer D1. All

of this corroborate the fact that there is not a conclusive criterion to select the

semiempirical method, but semiempirical Hamiltonians are adequate for a previous

discrimination of geometries in the MMH approach.

Figure 6.1. Selected PM3, AM1 and ab initio structures of the FA – furan, FA– FMA

and 1:2 FA – acetylene complexes.

173

BSSE effect on the geometry of the complex. Effect of the method and basis set

selection

The geometries of FA – FMA, FA – DME, FA – furan and acetylene – furan dimers

were optimized at the MP2/6-31G(d,p) level of theory using the CP scheme during the

optimization process to investigate the influence of the basis set superposition errors

(BSSE) on the geometries of the complexes. It is found that the geometries of the

“stronger” interacting complexes like FA – FMA are not as sensitive to the BSSE as the

weakly interacting complexes like the FA – furan dimers. For the FA – furan dimers the

BSSE corrected PES was closer to the PES calculated at higher level of theory than the

non-BSSE corrected, since the inclusion of the CP scheme during the optimization

process lead in many cases to geometries different to the non-BSSE ones. However, that

is critical for the weakest interacting complexes, which are also very sensitive to the basis

set used. The most stable complexes are not so affected by the inclusion of the CP

scheme during the optimization process. In general, the BSSE-free complexes show

larger hydrogen bond distances compared to the standard geometries at the same level of

theory.

Weakly interacting complexes are also more dependent on the method and basis set

used. In general, the cc-pVTZ basis set provides very good results, but with increasing

complexity of the system it becomes computationally very demanding. In this cases, the

use and comparison of the results with the aug-cc-pVDZ and the 6-311++G(d,p) basis

sets is an alternative.

174

Summary

To find out a representative and large set of the possible molecular arrangements

(minima) of hydrogen bonded and weakly interacting complexes is quite often one of the

most complicated questions. Therefore, the structural analysis application of the Multiple

Minima Hypersurface (MMH) approach[4-6] as a tool for localizing minima is introduced

here. Randomly arranged clusters are generated as starting points and subsequently

optimized. The results are processed with programs especially written for this purpose

and the geometries are afterward re-optimized at higher level of theory.

There are several steps in the MMH procedure:

• Generation of starting geometries: Random geometries are generated with the

GRANADA program,[4] especially written for this purpose. In most cases,

around 1000 randomly arranged clusters or complexes are generated as starting

points.

• Preliminary calculation of the energy: To calculate the preliminary energies of

all the generated molecular arrangements, PM3 and AM1 semiempirical

Hamiltonians are used.

• First discrimination and similarity analysis: All optimized structures which are

degenerated are discarded. Two types of degeneracy are considered. The first

consists of clusters which are identical, which means that they have both the

same energy and molecular geometry (SD). The second consists of clusters with

different molecular geometry but the same energy (VD). A subroutine called

Tanimoto is introduced in the program to analyze the similarity among

molecular arrangements, to discard the SD and keep the VD. The Tanimoto

procedure uses the Tanimoto similarity index to calculate the similarity between

structures pair by pair.

• Refinement of the geometries: The semiempirical results provide just a

preliminary overview of the interactions in the complex. For this reason, the set

of relevant semiempirical local minima are refined using DFT and MP2

175

methods. The ab initio and DFT computations are performed using the

Gaussian 98,[141] Gaussian 03,[141] and MOLPRO[142] programs. The

equilibrium geometries and vibrational frequencies are calculated using second

order Møller−Plesset perturbation theory (MP2). Pople’s 6-31G(d,p), 6-

311++G(d,p) and 6-311++G(3df,3pd) basis set as well of augmented and non

augmented Dunning’s correlation consistent double and triple ζ basis sets (cc-

pVDZ, aug-cc-pVDZ, cc-pVTZ and aug-cc-pVTZ) are used. In some cases

single point calculations are done with coupled clusters of single and double

substitutions (with non iterative triples) CCSD(T) and Dunning’s correlation

consistent triple ζ basis sets. The B3LYP density functional and Dunning’s

correlation consistent triple ζ basis sets are used also for the DFT calculations.

Formic acid – formamide (FA – FMA)

Nine FA – FMA dimers A – I with binding energies between -2.91 and -13.02 kcal/mol

(MP2/aug-cc-pVTZ + ZPE + BSSE) are identified after MMH search and refinement

with both DFT and MP2 calculations. The B3LYP density functional with the cc-pVTZ

basis set provides reliable geometries for the FA – FMA complexes. At the MP2 level of

theory, basically no change of the geometries when the basis set is augmented by adding

diffuse functions is found. At the MP2 level cc-pVDZ calculations show a tendency to

overestimate the binding energies, however, triple zeta basis sets either augmented or

non-augmented result in binding energies very similar to those from CCSD(T)/cc-pVTZ

single point calculations. The geometries and energies of the FA – FMA dimers A and B

do not change considerably with the inclusion of BSSE corrections during the

optimization process.

Seven basic types of interactions can be differentiated in the FA – FMA complexes.

They are the NHFMA…O=CFA interaction between the amide hydrogen atom of FMA and

the carbonyl oxygen atom of FA; the C=OFMA…HOFA interaction between the carbonyl

oxygen atom of FMA and the hydroxyl hydrogen atom of FA; the (O)CHFMA…O=CFA

interaction between the aldehyde hydrogen atom of FMA and the carbonyl oxygen atom

of FA; the NHFMA…(H)OCFA interaction between the amide hydrogen atom of FMA and

the hydroxyl oxygen atom of FA; the C=OFMA…HC(O)FA interaction between the

176

carbonyl oxygen atom of FMA and the aldehyde hydrogen atom of FA; the

HN(H)FMA…HOFA interaction between the nitrogen atom of FMA and the hydroxyl

hydrogen atom of FA; and the (O)CHFMA…(H)OCFA interaction between the aldehyde

hydrogen atom of FMA and the hydroxyl oxygen atom of FA.

The most stable dimers A and B are those where both carbonyl groups of FMA and FA

are involved in the stabilization of the complex, together with the hydroxyl hydrogen

atom of FA that interacts with the carbonyl oxygen atom of FMA. In the less stable

complexes F – I the hydroxyl hydrogen atoms of FA are not involved in hydrogen bonds.

Since all the geometries of the complexes were produced from randomly generated

geometries and not via chemical intuition, it is interesting to note that the structures of the

FA – FMA dimers A and B are in excellent agreement with the geometries of the FA –

FMA dimers reported in the literature. They show also interesting analogies with the

FMA – water and FMA – methanol dimers. The calculated geometries and binding

energies of 1:2 and 1:4 FA – FMA complexes show very interesting similarities with the

FA – FMA dimers and with the FA – FMA crystal structure. Of special interest are

structural motives found in the crystal structure that are already present in complexes of

very few molecules. The pair contributions of each dimer to the stabilization of the 1:2

FA – FMA trimers are also analyzed.

Formic acid – dimethyl ether (FA – DME)

Six FA – DME complexes with binding energies between -2.26 and -7.97 kcal/mol

(MP2/cc-pVTZ + ZPE+ BSSE) are identified. The two strongest bound complexes are

within a range of 0.3 kcal/mol isoenergetic. The binding in these six dimers can be

described in terms of OH…O, C=O…H, C-O…H and CH…O interactions. In the most

stable complexes A and B, the OH hydrogen atom of FA forms a strong hydrogen bond

with the ether oxygen atom of DME. Although the OH…O interaction is dominating in

complexes A and B, the secondary interaction between a methyl group hydrogen atom of

DME and the carbonyl oxygen atom of FA leads to an additional significant stabilization

of these complexes. The difference between these complexes is that in dimer A the C=O

group of FA interacts simultaneously with two hydrogen atoms of DME, while in dimer

B the C=O group of FA is approaching only one hydrogen atom of the DME. These two

177

complexes are predicted to be the most stable ones and with -7.81 to -7.97 kcal/mol

binding to be almost isoenergetic. During the optimization process the second enantiomer

of complex B was also found which shows the reliability of the MMH procedure.

The second group of FA – DME complexes C – F is defined by the absence of the

strong hydrogen bond. With -2.3 to -2.9 kcal/mol the binding energy is considerably

smaller and consequently these complexes could not be identified experimentally. The

dominant interaction in these complexes is the interaction between the aldehyde

hydrogen atom of FA and the DME oxygen atom. Again, interactions between the methyl

groups of DME and oxygen atoms of FA form secondary, weak interactions which,

however, determine the geometry of the complexes.

For all complexes the geometries are almost independent of the basis sets used. When

BSSE corrections are included in the optimization the intermolecular distances

corresponding to weaker interactions, where the DME hydrogen atoms are involved are

more influenced, showing an increasing of about 0.2-0.3 Å. Despite these variations the

basic geometries and interactions in the FA – DME complexes do not change.

Comparing the structures of the FA – DME complexes A and B with the DME –

methanol complex reported in literature reveals large similarities. There are also

interesting analogies between the FA – DME and the FA – water reported dimers. The

calculated vibrational frequencies are compared to matrix isolation IR spectra and the two

strongest bound complexes are identified.

Formic acid – furan (FA – furan)

Nine FA – furan complexes with binding energies between -3.91 and -0.82 kcal/mol

(MP2/6-311G(d,p) + ZPE+ BSSE) are identified. Another five weaker bound complexes

are localized at lower level of theory only. The binding in the furan – FA dimers can be

described in terms of OH…O, C=O…H, HO…H, CH…O, OH…π and CH…π

interactions. Therefore, the furan – FA complexes are classified in two types: (i) the

dimers where the hydroxyl group of formic acid interacts with the furan molecule and (ii)

the dimers where the main interactions of FA with the furan molecule are via the less

178

acidic C-H group. Due to the lower acidity of the CH hydrogen atom, type (ii) complexes

are less stable than type (i) complexes.

The most stable complex A has binding energy -3.91 kcal/mol (MP2/6-311++G(d,p) +

ZPE+ BSSE). Although the OH…O interaction is dominating in dimer A, the secondary

C=OFA…HF interaction between the carbonyl oxygen atom of FA and the hydrogen atom

of furan leads to an additional significant stabilization of this complex. The experimental

matrix isolation IR vibrational frequencies agree very well with the calculated IR spectra

of dimer A.

FA – furan complexes B, C, and D1 are π complexes defined by the absence of the

strong OH…O hydrogen bond. These dimers are stabilized by the O-HFA…π interaction

between the OH hydrogen atom of FA and the π system of furan. With -2.24, -2.12 and -

2.37 kcal/mol (MP2/6-311++G(d,p) + ZPE+ BSSE, Table 1) the binding energies are

considerably smaller than that of A, and consequently these complexes are not identified

experimentally.

The second group of FA – furan dimers is mainly stabilized by the very weak CH…O

or CH…π interaction. The most stable dimers in this group are H and I with binding

energies of -1.94 and -1.35 kcal/mol, respectively at the MP2/6-311++G(d,p) + ZPE+

BSSE level of theory. The higher stability of dimer H can be attributed to an additional

C=OFA…HF interaction between the carbonyl oxygen atom of FA and one of the

hydrogen atoms of furan. Dimer I exhibits the less stabilizing interaction HOFA…HF

between the OH oxygen atom of FA and one of the hydrogen atoms of furan. Dimers E,

F1, and G are very weakly bound CH…π complexes with binding energies between -0.96

and -0.82 kcal/mol (MP2/6-311++G(d,p) + ZPE+ BSSE).

The MP2 level of theory with the 6-311++G(d,p) and aug-cc-pVDZ basis sets provides

reliable geometries for FA – furan complexes. With the small double zeta basis sets

without augmentation the structures of five additional very weak FA – furan complexes

could be localized. Introducing the BSSE corrections during the geometry optimization at

the MP2/6-31G(d,p) level of theory leads to large changes of the calculated geometries of

some of the very weak FA – furan complexes. Here, clearly, BSSE and variations of the

basis sets have the largest effects on the dimers.

179

1:2 Formic acid – Acetylene Complexes (1:2 FA – acetylene)

An interesting feature of this system is the competition between the strongly acidic

carboxyl group, the acetylene group, and the formyl group as hydrogen bridge donors and

the carbonyl group, the hydroxyl group, and the acetylene π-system as hydrogen bridge

acceptors.

Six complexes with binding energies between -3.93 and -7.98 kcal/mol (MP2/cc-pVTZ

+ ZPE) are identified. Four additional structures B1, E1, G1 and H1 were located using

DFT theory (B3LYP/6-311++G(d,p)) or MP2 with the MP2/6-311++G(d,p) basis set, but

are not minima at higher levels of theory. Three stable complexes with binding energies

within 1 kcal/mol are predicted. The binding in complexes between formic acid and

acetylene shows contributions from the CH...O, OH...π and CH...π interactions.

In 1:2 FA – acetylene complexes the interaction between the acidic H atom of formic

acid and the acetylene π-system (partial structure (i)) provides 50 – 60% of the binding

energies of the most stable complexes A – C, and thus dominates the non-covalent

interactions in these systems. Partial structures (ii) and (iii) add between 15 and 29% to

the total binding energy. Only in complex B, where no “T-acetylene” interaction is

present, contribution (iii) is slightly repulsive. Other interactions are less important for

the stabilization of the complexes, but might influence the structure. Thus, complex B is

only slightly stabilized by an additional very weak CH...π interaction compared to B1.

This interaction results in a considerable structural change. The calculated vibrational

frequencies of the most stable 1:2 complex A are in good agreement with the

experimental values, indicating that under the conditions of matrix isolation indeed the

complex A is formed.

Acetylene complexes with oxygen heterocycles

Five acetylene – furan dimers with binding energies between -0.75 and -1.77 kcal/mol

are identified (MP2/6-311++G(d,p)+ BSSE). Two dimers of acetylene – THF were found

with binding energies -3.66 and -3.48 kcal/mol, (MP2/6-311++G(d,p)+ BSSE)

respectively and the THF subunit in the “twist” and “envelope” conformations. At the

same level of theory, only one acetylene – 1,4-dioxane dimer (-2.96 kcal/mol) was

180

localized. However, with other basis sets (e.g. cc-pVTZ) a second complex with a

different orientation of the acetylene molecule but very close in energy was identified.

In the acetylene – furan dimers two types of interactions are found: the C-H…O and

the C-H…π interaction. The C-H…π interaction appears in two variations, depending on

which molecule provides the hydrogen atom and which molecule the π system. The

results indicate that the C-H…π interaction between one hydrogen atom of acetylene and

the π system of furan is the stronger interaction, also compared to the in-plane C-H…O

interaction found in the second more stable acetylene – furan dimer B. Acetylene – THF

and acetylene – 1,4-dioxane dimers are stabilized by the C-H…O interaction between the

oxygen atom of the heterocycle and the hydrogen atom of acetylene, however, the

position of the acetylene subunit in the complex suggests an additional very weak C-

HTHF(DIOXANE)…π stabilizing interaction.

For the different systems studied, the comparison with matrix spectroscopy, crystal

structure data, and the analysis of studies of similar complexes from the literature,

confirms the quality of the MMH procedure as a very useful tool for reliably localizing

minima in hydrogen bonded complexes without recurring to previous knowledge of the

structure of supramolecular complexes.

181

References

[1] G. A. Jeffrey, An Introduction to Hydrogen Bonding, New York, 1997. [2] G. A. Jeffrey, Journal of Molecular Structure 1994, 322, 21. [3] M. Nishio, M. Hirota, Y. Umezawa, CH/p Interaction: Evidence, Nature, and

Consequences, 1998. [4] L. A. Montero, GRANADA and Q programs for PC computers ed., 1996. [5] L. A. Montero, A. M. Esteva, J. Molina, A. Zapardiel, L. Hernandez, H. Marquez,

A. Acosta, Journal of the American Chemical Society 1998, 120, 12023. [6] L. A. Montero, J. Molina, J. Fabian, International Journal of Quantum Chemistry

2000, 79, 8. [7] M. Gantenberg, M. Halupka, W. Sander, Chemistry--A European Journal 2000, 6,

1865. [8] A. K. Roy, A. J. Thakkar, in Chemical Physics, Vol. 312, 2005, pp. 119. [9] W. H. Hocking, Zeitschrift fuer Naturforschung, Teil A: Astrophysik, Physik und

Physikalische Chemie 1976, 31A, 1113. [10] J. Karle, L. O. Brockway, Journal of the American Chemical Society 1944, 66,

574. [11] R. C. Millikan, K. S. Pitzer, Journal of Chemical Physics 1957, 27, 1305. [12] E. Bjarnov, W. H. Hocking, Zeitschrift fuer Naturforschung, Teil A: Astrophysik,

Physik und Physikalische Chemie 1978, 33A, 610. [13] M. Pettersson, J. Lundell, L. Khriachtchev, M. Raesaenen, J. Am. Chem. Soc.

1997, 119, 11715. [14] M. D. Taylor, J. Bruton, Journal of the American Chemical Society 1952, 74,

4151. [15] I. Nahringbauer, Acta Crystallographica, Section B: Structural Crystallography

and Crystal Chemistry 1978, B34, 315. [16] I. Bako, G. Schubert, T. Megyes, G. Palinkas, G. I. Swan, J. Dore, M.-C.

Bellisent-Funel, Chemical Physics 2004, 306, 241. [17] E. Spinner, Spectrochimica Acta, Part A: Molecular and Biomolecular

Spectroscopy 1999, 55A, 1819. [18] D. Chapman, Journal of the Chemical Society 1956, 225. [19] S. Scheiner, C. W. Kern, in Journal of the American Chemical Society, Vol. 101,

1979, pp. 4081. [20] Y. J. Zheng, K. M. Merz, Jr., Journal of Computational Chemistry 1992, 13,

1151. [21] I. Rozas, I. Alkorta, J. Elguero, Journal of Physical Chemistry B 2004, 108, 3335. [22] J. W. Keller, in Journal of Physical Chemistry A, Vol. 108, 2004, pp. 4610. [23] P. Pfeiffer, G. Birencweig, A. Hofmann, C. Windheuser, Ber. 1914, 47, 1580. [24] A. Werner, Liebigs Annalen der Chemie 1902, 261. [25] T. S. Moore, T. F. Winmill, Journal of the Chemical Society, Transactions 1912,

101, 1635. [26] W. M. Latimer, W. H. Rodebush, Journal of the American Chemical Society

1920, 42, 1419. [27] L. Pauling, The Nature of the Chemical Bond., Cornell University Press, Ithaca,

NY, 1939.

182

[28] G. C. Pimentel, A. L. McClellan, The Hydrogen Bond, Freeman, San Francisco, 1960.

[29] G. R. Desiraju, T. Steiner, The Weak Hydrogen Bond in Structural Chemistry and Biology, Oxford University Press, 1999.

[30] L. Hunter, Ann. Repts. Progress Chem. (Chem. Soc. London) 1947, 43, 141. [31] G. R. Desiraju, Accounts of Chemical Research 2002, 35, 565. [32] T. W. Panunto, Z. Urbanczyk-Lipkowska, R. Johnson, M. C. Etter, Journal of the

American Chemical Society 1987, 109, 7786. [33] W. Saenger, T. Steiner, Conference Proceedings - Italian Physical Society 1993,

43, 55. [34] W. Saenger, Nature (London, United Kingdom) 1979, 279, 343. [35] G. Gilli, F. Bellucci, V. Ferretti, V. Bertolasi, Journal of the American Chemical

Society 1989, 111, 1023. [36] J. Del Bene, J. A. Pople, Journal of Chemical Physics 1970, 52, 4858. [37] J. L. Koenig, Accounts of Chemical Research 1981, 14, 171. [38] E. Whittle, D. A. Dows, G. C. Pimentel, J. Chem. Phys. 1954, 22, 1943. [39] I. Norman, G. Porter, Nature 1954, 174, 508. [40] http://www.orch.ruhr-uni-bochum.de/sander/. [41] M. Squillacote, R. S. Sheridan, O. L. Chapman, F. A. L. Anet, J. Am. Chem. Soc.

1975, 97, 3244. [42] M. J. Almond, K. S. Wiltshire, Annual Reports on the Progress of Chemistry,

Section C: Physical Chemistry 2001, 97, 3. [43] S. Scheiner, Hydrogen Bonding: A Theoretical Perspective, Oxford University

Press, New York, 1997. [44] C. J. Cramer, Essentials of Computational Chemistry : Theories and Models, John

Wiley and Sons, Ltd, 2004. [45] J. M. Foster, S. F. Boys, Reviews of Modern Physics 1960, 32, 300. [46] J. A. Pople, M. Head-Gordon, K. Raghavachari, Journal of Chemical Physics

1987, 87, 5968. [47] R. Ahlrichs, Computer Physics Communications 1979, 17, 31. [48] P. Carsky, I. Hubac, V. Staemmler, Theoretica Chimica Acta 1982, 60, 445. [49] J. Cizek, J. Paldus, Physica Scripta 1980, 21, 251. [50] J. A. Pople, R. Krishnan, H. B. Schlegel, J. S. Binkley, International Journal of

Quantum Chemistry 1978, 14, 545. [51] G. D. Purvis, III, R. J. Bartlett, Journal of Chemical Physics 1982, 76, 1910. [52] J. S. Binkley, J. A. Pople, International Journal of Quantum Chemistry 1975, 9,

229. [53] J. A. Pople, J. S. Binkley, R. Seeger, International Journal of Quantum

Chemistry, Symposium 1976, 10, 1. [54] A. C. Whal, G. Das, in Methods of Electronic Structure Theory, Vol. 3 (Ed.: H. F.

Schaefer), Plenum, New York, 1977. [55] B. O. Roos, P. R. Taylor, E. M. Siegbahn, Chemical Physics 1980, 48, 157. [56] F. Jensen, Introduction to Computational Chemistry, John Wiley and Sons, 1999. [57] W. Thiel, in Modern Methods and Algorithms of Quantum Chemistry (Ed.: J.

Grotendorst), John von Neumann Institute for Computing, Jülich, 2000, pp. 233.

183

[58] D. Hadzi, G. Koller, in Theoretical Treatments of Hydrogen Bonding (Ed.: D. Hadzi), John Wiley and Sons, 1997.

[59] M. J. S. Dewar, E. G. Zoebisch, E. F. Healy, J. J. P. Stewart, Journal of the American Chemical Society 1985, 107, 3902.

[60] J. J. P. Stewart, Journal of Computational Chemistry 1989, 10, 221. [61] G. I. Csonka, Journal of Computational Chemistry 1993, 14, 895. [62] G. I. Csonka, J. G. Angyan, Theochem 1997, 393, 31. [63] M. J. S. Dewar, C. Jie, J. Yu, Tetrahedron 1993, 49, 5003. [64] L. Turi, J. J. Dannenberg, Journal of Physical Chemistry 1993, 97, 12197. [65] L. Turi, J. J. Dannenberg, Journal of Physical Chemistry 1993, 97, 7899. [66] P. Hohenberg, W. Kohn, Phys. Rev. 1964, 136, B864. [67] W. Kohn, L. J. Sham, Phys. Rev. 1965, 140, A1133. [68] J. B. Foresman, A. Frisch, Exploring Chemistry with Electronic Structure

Methods, Second ed., Gaussian, Inc., Pittsburgh, 1996. [69] S. J. Vosko, L. Wilk, M. Nusair, Can. J. Phys 1980, 58, 1200. [70] H. Guo, S. Sirois, E. I. Proynov, D. R. Salahub, in Theoretical Treatments of

Hydrogen Bonding (Ed.: D. Hadzi), John Wiley and Sons, 1997. [71] A. D. Becke, Phys. Rev. A: Gen. Phys. 1988, 38, 3098. [72] J. P. Perdew, Y. Wang, Phys. Rev. 1986, 8800. [73] C. Lee, W. Yang, R. G. Parr, Physical Review B: Condensed Matter and

Materials Physics 1988, 37, 785. [74] A. D. Becke, Journal of Chemical Physics 1993, 98, 5648. [75] Y. Zhao, O. Tishchenko, D. G. Truhlar, Journal of Physical Chemistry B 2005,

109, 19046. [76] Y. Zhao, D. G. Truhlar, Physical Chemistry Chemical Physics 2005, 7, 2701. [77] E. R. Davidson, D. Feller, Chemical Reviews (Washington, DC, United States)

1986, 86, 681. [78] T. H. Dunning, P. J. Hay, in Methods of Electronic Structure Theory, Vol. 3 (Ed.:

H. F. Schaeffer), Plenum, New York, 1977. [79] W. J. Hehre, R. F. Stewart, J. A. Pople, Journal of Chemical Physics 1969, 51,

2657. [80] R. Ditchfield, W. J. Hehre, J. A. Pople, Journal of Chemical Physics 1971, 54,

724. [81] R. Krishnan, J. S. Binkley, R. Seeger, J. A. Pople, in Journal of Chemical

Physics, Vol. 72, 1980, pp. 650. [82] T. H. Dunning, Jr., Journal of Chemical Physics 1989, 90, 1007. [83] S. F. Boys, F. Bernardi, Mol Phys 1970, 19, 553. [84] S. Simon, J. Bertran, M. Sodupe, Journal of Physical Chemistry A 2001, 105,

4359. [85] P. Salvador, S. Simon, M. Duran, J. J. Dannenberg, Journal of Chemical Physics

2000, 113, 5666. [86] P. Salvador, B. Paizs, M. Duran, S. Suhai, Journal of Computational Chemistry

2001, 22, 765. [87] P. Salvador, M. M. Szczesniak, Journal of Chemical Physics 2003, 118, 537. [88] M. C. Daza, J. A. Dobado, J. M. Molina, P. Salvador, M. Duran, J. L. Villaveces,

Journal of Chemical Physics 1999, 110, 11806.

184

[89] S. Simon, M. Duran, J. J. Dannenberg, Journal of Chemical Physics 1996, 105, 11024.

[90] R. Crespo-Otero, L. A. Montero, W.-D. Stohrer, J. M. Garcia de la Vega, Journal of Chemical Physics 2005, 123, 134107/1.

[91] P. Hobza, Z. Havlas, Theoretical Chemistry Accounts 1998, 99, 372. [92] I. Mayer, International Journal of Quantum Chemistry 1983, 23, 341. [93] A. Bende, A. Vibok, G. J. Halasz, S. Suhai, International Journal of Quantum

Chemistry 2001, 84, 617. [94] P. Salvador, M. Duran, X. Fradera, in Journal of Chemical Physics, Vol. 116,

2002, pp. 6443. [95] I. Mayer, A. Vibok, G. Halasz, P. Valiron, International Journal of Quantum

Chemistry 1996, 57, 1049. [96] I. Mayer, P. R. Surjan, International Journal of Quantum Chemistry 1989, 36,

225. [97] I. Mayer, A. Vibok, Chemical Physics Letters 1987, 140, 558. [98] I. Mayer, A. Vibok, Molecular Physics 1997, 92, 503. [99] A. Hamza, A. Vibok, G. J. Halasz, I. Mayer, Theochem 2000, 501-502, 427. [100] L. Turi, J. J. Dannenberg, Journal of Physical Chemistry 1993, 97, 2488. [101] P. Valiron, I. Mayer, Chemical Physics Letters 1997, 275, 46. [102] B. H. Wells, S. Wilson, Chemical Physics Letters 1983, 101, 429. [103] P. Willett, J. M. Barnard, G. M. Downs, Journal of Chemical Information and

Computer Sciences 1998, 38, 983. [104] W. Fisanick, A. H. Lipkus, A. Rusinko, III, Journal of Chemical Information and

Computer Sciences 1994, 34, 130. [105] D. Butina, Journal of Chemical Information and Computer Sciences 1999, 39,

747. [106] E. S. Garcia, L. A. Montero, J. M. Hermida, R. Cruz, G. Gonzalez, Revista

Cubana de Fisica 2000, 17, 41. [107] J. J. P. Stewart, Journal of Computational Chemistry 1989, 10, 209. [108] J. J. P. Stewart, MOPAC v 6.0 ed. [109] J. J. P. Stewart, Journal of Computer-Aided Molecular Design 1990, 4, 1. [110] J. M. Stewart, Frank J. Seiler Res. Lab.,United States Air Force Acad.,Colorado

Springs,CO,USA., 1988, p. 183 pp. [111] P. Culot, G. Dive, V. H. Nguyen, J. M. Ghuysen, Theoretica Chimica Acta 1992,

82, 189. [112] J. Baker, Journal of Computational Chemistry 1986, 7, 385. [113] P. O. Astrand, G. Karlstrom, A. Engdahl, B. Nelander, Journal of Chemical

Physics 1995, 102, 3534. [114] B. Chan, J. E. Del Bene, J. Elguero, L. Radom, Journal of Physical Chemistry A

2005, 109, 5509. [115] O. Galvez, P. C. Gomez, L. F. Pacios, Journal of Chemical Physics 2003, 118,

4878. [116] L. George, E. Sanchez-Garcia, W. Sander, Journal of Physical Chemistry A 2003,

107, 6850. [117] L. George, W. Sander, Spectrochimica Acta, Part A: Molecular and Biomolecular

Spectroscopy 2004, 60A, 3225.

185

[118] I. Agranat, N. V. Riggs, L. Radom, in Journal of the Chemical Society, Chemical Communications, 1991, pp. 80.

[119] P. R. L. Markwick, N. L. Doltsinis, D. Marx, Journal of Chemical Physics 2005, 122, 054112/1.

[120] C. S. Tautermann, M. J. Loferer, A. F. Voegele, K. R. Liedl, Journal of Chemical Physics 2004, 120, 11650.

[121] N. R. Brinkmann, G. S. Tschumper, G. Yan, H. F. Schaefer, Journal of Physical Chemistry A 2003, 107, 10208.

[122] F. Madeja, M. Havenith, Journal of Chemical Physics 2002, 117, 7162. [123] H. Ushiyama, K. Takatsuka, in Journal of Chemical Physics, Vol. 115, 2001, pp.

5903. [124] J. Kohanoff, S. Koval, D. A. Estrin, D. Laria, Y. Abashkin, in Journal of

Chemical Physics, Vol. 112, 2000, pp. 9498. [125] S. J. Grabowski, T. M. Krygowski, in Chemical Physics Letters, Vol. 305, 1999,

pp. 247. [126] T. Loerting, K. R. Liedl, Journal of the American Chemical Society 1998, 120,

12595. [127] Y. Kim, Journal of the American Chemical Society 1996, 118, 1522. [128] C. Zhu, Y. Ling, W. Y. Feng, C. Lifshitz, International Journal of Mass

Spectrometry 2000, 194, 93. [129] G. P. Ayers, A. D. E. Pullin, Spectrochimica Acta, Part A: Molecular and

Biomolecular Spectroscopy 1976, 32A, 1641. [130] M. L. H. Jeng, B. S. Ault, Journal of Molecular Structure 1991, 246, 33. [131] A. Engdahl, B. Nelander, J. Chem. Soc., Faraday Trans. 1992, 88, 177. [132] B. C. Bricknell, T. M. Letcher, T. A. Ford, South African Journal of Chemistry

1995, 48, 142. [133] S. W. Han, K. Kim, Journal of Molecular Structure 1999, 475, 43. [134] J. Goebel, B. S. Ault, J. E. D. Bene, Journal of Physical Chemistry A 2000, 104,

2033. [135] A. Loutellier, L. Schriver, A. Burneau, J. P. Perchard, Journal of Molecular

Structure 1982, 82, 165. [136] A. Fu, D. Du, Z. Zhou, Theochem 2003, 623, 315. [137] A. Fu, D. Du, Z. Zhou, International Journal of Quantum Chemistry 2004, 97,

865. [138] E. L. Coitino, K. Irving, J. Rama, A. Iglesias, M. Paulino, O. N. Ventura,

Theochem 1990, 69, 405. [139] I. Nahringbauer, G. Larsson, Arkiv foer Kemi 1968, 30, 91. [140] T. Neuheuser, B. A. Hess, C. Reutel, E. Weber, Journal of Physical Chemistry

1994, 98, 6459. [141] G. W. T. M. J. Frisch, H. B. Schlegel, G. E. Scuseria, , J. R. C. M. A. Robb, J. A.

Montgomery, Jr., T. Vreven, , J. C. B. K. N. Kudin, J. M. Millam, S. S. Iyengar, J. Tomasi, , B. M. V. Barone, M. Cossi, G. Scalmani, N. Rega, , H. N. G. A. Petersson, M. Hada, M. Ehara, K. Toyota, , J. H. R. Fukuda, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, , M. K. H. Nakai, X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, , J. J. C. Adamo, R. Gomperts, R. E. Stratmann, O. Yazyev, , R. C. A. J. Austin, C. Pomelli, J. W. Ochterski, P. Y. Ayala, , G. A. V.

186

K. Morokuma, P. Salvador, J. J. Dannenberg, , S. D. V. G. Zakrzewski, A. D. Daniels, M. C. Strain, , D. K. M. O. Farkas, A. D. Rabuck, K. Raghavachari, , J. V. O. J. B. Foresman, Q. Cui, A. G. Baboul, S. Clifford, , B. B. S. J. Cioslowski, G. Liu, A. Liashenko, P. Piskorz, , R. L. M. I. Komaromi, D. J. Fox, T. Keith, M. A. Al-Laham, , A. N. C. Y. Peng, M. Challacombe, P. M. W. Gill, , W. C. B. Johnson, M. W. Wong, C. Gonzalez, J. A. Pople., Gaussian 03, Revision B.03, Pittsburgh PA, 2003.

[142] H.-J. K. Werner, P. J.; Schu¨tz, M.; Lindh, R.; Celani, P.;, T. R. Korona, G.; Manby, F. R.; Amos, R. D.; Bernhardsson, A.;, A. C. Berning, D. L.; Deegan, M. J. O.; Dobbyn, A. J.; Eckert, F.;, C. H. Hampel, G.; Lloyd, A. W.; McNicholas, S. J.; Meyer, W.; Mura,, A. P. M. E.; Nicklass, P.; Pitzer, R.; Schumann, U.; Stoll, H.; Stone,, R. T. A. J.; Tarroni, T. MOLPRO; Universita¨t Stuttgart:, G. Stuttgart.

[143] P. G. Jansien, Stevens, WalterJ, J.Chem. Phys 1986, 84, 3271. [144] L. F. Pacios, Journal of Physical Chemistry A 2004, 108, 1177. [145] E. Sanchez-Garcia, L. George, L. A. Montero, W. Sander, Journal of Physical

Chemistry A 2004, 108, 11846. [146] E. Sanchez-Garcia, M. Studentkowski, L. A. Montero, W. Sander,

ChemPhysChem 2005, 6, 618. [147] D. Priem, T.-K. Ha, A. Bauder, Journal of Chemical Physics 2000, 113, 169. [148] Z. Zhou, Y. Shi, X. Zhou, Journal of Physical Chemistry A 2004, 108, 813. [149] M. D. Harmony, V. W. Laurie, R. L. Kuczkowski, R. H. Schwendeman, D. A.

Ramsay, F. J. Lovas, W. J. Lafferty, A. G. Maki, Journal of Physical and Chemical Reference Data 1979, 8, 619.

[150] A. R. Katritzky, J. M. Lagowski, Heterocyclic Chemistry, 1960. [151] M. Hartmann, S. D. Wetmore, L. Radom, Journal of Physical Chemistry A 2001,

105, 4470. [152] P. Hobza, C. Riehn, A. Weichert, B. Brutschy, Chemical Physics 2002, 283, 331. [153] M. L. H. Jeng, B. S. Ault, Journal of Physical Chemistry 1990, 94, 4851. [154] M. L. H. Jeng, B. S. Ault, Journal of Physical Chemistry 1990, 94, 1323. [155] S. Tsuzuki, K. Honda, T. Uchimaru, M. Mikami, K. Tanabe, Journal of Physical

Chemistry A 1999, 103, 8265. [156] P. Hobza, H. L. Selzle, E. W. Schlag, Collection of Czechoslovak Chemical

Communications 1992, 57, 1186. [157] H. Petrusova, Z. Havlas, P. Hobza, R. Zahradnik, Collection of Czechoslovak

Chemical Communications 1988, 53, 2495. [158] G. Maier, C. Lautz, European Journal of Organic Chemistry 1998, 769. [159] V. Lukes, M. Breza, Petroleum and Coal 2002, 44, 51. [160] J. S. Muenter, Journal of Chemical Physics 1991, 94, 2781. [161] J. S. Craw, M. A. C. Nascimento, M. N. Ramos, Spectrochimica Acta, Part A:

Molecular and Biomolecular Spectroscopy 1991, 47A, 69. [162] S. Scheiner, S. J. Grabowski, in Journal of Molecular Structure, Vol. 615, 2002,

pp. 209. [163] S. M. Resende, W. B. De Almeida, Chemical Physics 1996, 206, 1. [164] C. E. Dykstra, Journal of the American Chemical Society 1990, 112, 7540. [165] K. Shuler, C. E. Dykstra, Journal of Physical Chemistry A 2000, 104, 11522. [166] A. Karpfen, Journal of Physical Chemistry A 1999, 103, 11431.

187

[167] D. G. Prichard, R. N. Nandi, J. S. Muenter, Journal of Chemical Physics 1988, 89, 115.

[168] J. S. Craw, W. B. De Almeida, A. Hinchliffe, Theochem 1989, 60, 69. [169] T. Aoyama, O. Matsuoka, N. Nakagawa, Chemical Physics Letters 1979, 67, 508. [170] I. L. Alberts, T. W. Rowlands, N. C. Handy, Journal of Chemical Physics 1988,

88, 3811. [171] K. Pei, H. Li, in Journal of Molecular Structure, Vol. 693, 2004, pp. 141. [172] A. C. Legon, P. Ottaviani, Physical Chemistry Chemical Physics 2004, 6, 488. [173] H. Koppel, E. V. Gromov, A. B. Trofimov, Chemical Physics 2004, 304, 35. [174] D.-M. Huang, Y.-B. Wang, L. M. Visco, F.-M. Tao, in Journal of Physical

Chemistry A, Vol. 108, 2004, pp. 11375. [175] E. V. Gromov, A. B. Trofimov, N. M. Vitkovskaya, H. Koppel, J. Schirmer, H. D.

Meyer, L. S. Cederbaum, Journal of Chemical Physics 2004, 121, 4585. [176] E. V. Gromov, A. B. Trofimov, N. M. Vitkovskaya, J. Schirmer, H. Koppel,

Journal of Chemical Physics 2003, 119, 737. [177] G. C. Cole, A. C. Legon, P. Ottaviani, Journal of Chemical Physics 2002, 117,

2790. [178] S. A. Cooke, G. K. Corlett, A. C. Legon, Chemical Physics Letters 1998, 291,

269. [179] S. A. Cooke, G. K. Corlett, J. H. Holloway, A. C. Legon, Journal of the Chemical

Society, Faraday Transactions 1998, 94, 2675. [180] L. A. Montero, R. Gonzalez-Jonte, L. A. Diaz, J. R. Alvarez-Idaboy, Journal of

Physical Chemistry 1994, 98, 5607. [181] A. C. Legon, Faraday Discussions 1994, 97, 19. [182] J. R. Alvarez-Idaboy, L. A. Montero, Theochem 1992, 85, 243. [183] J. R. Alvarez, L. Montero, R. Martinez, Makromolekulare Chemie 1990, 191,

281. [184] L. Radom, R. H. Nobes, D. J. Underwood, W. K. Li, Pure Appl. Chem. 1986, 58,

75. [185] B. S. Ault, in Journal of Molecular Structure, Vol. 127, 1985, pp. 343. [186] Z. Latajka, H. Ratajczak, W. J. Orville-Thomas, E. Ratajczak, J. Mol. Struct.

(THEOCHEM) 1981, 85, 303. [187] A. M. DeLaat, B. S. Ault, Journal of the American Chemical Society 1987, 109,

4232. [188] B. W. Hopkins, G. S. Tschumper, in Journal of Physical Chemistry A, Vol. 108,

2004, pp. 2941. [189] A. C. Legon, D. J. Millen, Faraday Discussions of the Chemical Society 1982, 73,

71. [190] A. C. Legon, D. J. Millen, Chemical Society Reviews 1987, 16, 467. [191] B. Bak, L. Hansen, J. Rastrup-Andersen, Discussions of the Faraday Society

1955, No. 19, 30. [192] E. Sanchez-Garcia, L. A. Montero, W. Sander, Journal of Physical Chemistry A,

in press. [193] A. Baldacci, S. Ghersetti, S. C. Hurlock, K. N. Rao, Journal of Molecular

Spectroscopy 1976, 59, 116.

188

[194] E. S. Kline, Z. H. Kafafi, R. H. Hauge, J. L. Margrave, Journal of the American Chemical Society 1985, 107, 7559.

[195] S. A. McDonald, G. L. Johnson, B. W. Keelan, L. Andrews, Journal of the American Chemical Society 1980, 102, 2892.

[196] E. D. Jemmis, K. T. Giju, K. Sundararajan, K. Sankaran, V. Vidya, K. S. Viswanathan, J. Leszczynski, Journal of Molecular Structure 1999, 510, 59.

[197] M. Halupka, W. Sander, in Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy, Vol. 54A, 1998, pp. 495.

[198] I. D. Reva, A. M. Plokhotnichenko, E. D. Radchenko, G. G. Sheina, Y. P. Blagoi, Spectrochimica Acta, Part A: Molecular and Biomolecular Spectroscopy 1994, 50A, 1107.

[199] D. Cremer, J. A. Pople, Journal of the American Chemical Society 1975, 97, 1358.

[200] P. Luger, J. Buschmann, Angewandte Chemie 1983, 95, 423. [201] G. G. Engerholm, A. C. Luntz, W. D. Gwinn, D. O. Harris, Journal of Chemical

Physics 1969, 50, 2446. [202] B. Cadioli, E. Gallinella, C. Coulombeau, H. Jobic, G. Berthier, Journal of

Physical Chemistry 1993, 97, 7844. [203] M. Rayon Victor, A. Sordo Jose, The Journal of chemical physics 2005, 122,

204303. [204] J. L. Alonso, J. C. Lopez, S. Blanco, A. Lesarri, F. J. Lorenzo, Journal of

Chemical Physics 2000, 113, 2760. [205] D. M. Chapman, R. E. Hester, Journal of Physical Chemistry A 1997, 101, 3382. [206] T. Steiner, E. B. Starikov, A. M. Amado, J. J. C. Teixeira-Dias, Journal of the

Chemical Society, Perkin Transactions 2: Physical Organic Chemistry 1995, 1321.

189

Acknowledgments

I am very much indebted to Prof Dr. Luis A. Montero for his support and guidance

over all these years and to Prof. Dr. Martina Havenith-Newen for her assistance and

encouragement.

Many thanks to Dr. Holger Bettinger for the help and advices in Computational

Chemistry and to Dr.Lisa George, Arthur Mardyukov and Marc Studentkowski, who did

the matrix isolation experiments. I also wish to thank Herr Torsten Haenschke for

keeping the computers running and his support. Thanks to Dr. Friedrich Scheidt and Frau

Ulrike Steger for their always kind attention, and to all the members of the Organic

Chemistry II research group for the nice working atmosphere and their support. Many

thanks to Dr. Thomas Koch and Frau Gundula Talbot of the Graduate School of

Chemistry and Biochemistry. I thank Prof. Dr Silvia Bravslasky and the members of the

Bravslasky group in the Max Planck Institute of Bioinorganic Chemistry.

I am very grateful to Rachel Crespo of the Computational Chemistry Laboratory in

Havana, for her friendship and invaluable unconditional support since many years. I

specially thank my friends Dr. Victor Martinez and Denise Larrieux, for their total

support. Thanks Loli and Vir for your help.

190

Curriculum Vitae

Name Elsa Sánchez García

Born 15th October 1976, Havana, Cuba

Nationality Cuban, with residence in Germany

Education Since Sept. 2002 Postgraduate studies of Chemistry, University of Bochum, Germany

1999-2002 Postgraduate studies of Chemistry at the Laboratory of Theoretical

and Computational Chemistry, Havana University

1999-2001 Study of English Language, Abraham Lincoln School, Havana

1999 Diploma for Chemistry at the University of Havana. Title: "Efectos

del ambiente molecular sobre procesos y estructuras moleculares en

solventes puros"

1994- 1999 Study of Chemistry at the University of Havana

1991-1994 Pre-University training in Havana

1988-1991 Secondary School in Havana

1981-1988 Primary School in Havana

Professional Career September 2002 - till present

Research Scientist at the Ruhr University of Bochum, Germany, with

Prof. Dr. W. Sander

April 2002 - July 2002

Research Scientist, Max Planck Institute für Strahlenchemie (now

Max Planck Institute für Bioanorganische Chemie), Mülheim an der

Ruhr, Germany, with Prof Dr. S. Bravslasky

1999-2002 Novel Lecturer, Faculty of Chemistry, Havana University

191

Awards 2002 Award of a Scholarship of the DAAD (German Academic Exchange

Service).

2002 Best Novel Lecturer (Reserva Científica), Department of General

Chemistry, Faculty of Chemistry, University of Havana

1999-2000 Award in the category of "Best Novel Lecturer (Reserva Científica)".

Faculty of Chemistry, University of Havana

1998-1999 Best graduated student of the University of Havana.

1999 Suma cum Laude Diploma.

1998-1999 Best graduated student of the Chemistry Faculty

1998 Award in the Student’s Scientific Meeting. Commission of Physical

Chemistry, Faculty of Chemistry, University of Havana

1998 Award in the Student’s Scientific Meeting. Commission of

Computational Sciences, University of Havana

1998 Outstanding student of the Havana University

1993-1994 Bronze Medal in the National Competition of Chemistry for Pre-

Universitary students, Cuba

1992-1993 Gold Medal in the National Competition of Spanish Language and

Literature for Pre-Universitary students.

1992-1993 Bronze Medal in the National Competition of Chemistry for Pre-

Universitary students, Cuba

1991-1992 Gold Medal in the National Competition of Chemistry for Pre-

Universitary students, Cuba