STACKING INTERACTIONS OF RESONANCE-ASSISTED

6 days ago - Stacking interactions of resonance-assisted hydrogen-bridged (RAHB) rings are quite common, as 44% of their crystal structures show mutua...
0 downloads 0 Views 703KB Size
Subscriber access provided by Nottingham Trent University

Article

STACKING INTERACTIONS OF RESONANCE-ASSISTED HYDROGEN-BRIDGED RINGS. A SYSTEMATIC STUDY OF CRYSTAL STRUCTURES AND QUANTUM CHEMICAL CALCULATIONS Jelena Blagojevic, Michael B. Hall, and Snezana D. Zaric Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.9b00589 • Publication Date (Web): 15 Aug 2019 Downloaded from pubs.acs.org on August 15, 2019

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

STACKING INTERACTIONS OF RESONANCE-ASSISTED HYDROGENBRIDGED RINGS. A SYSTEMATIC STUDY OF CRYSTAL STRUCTURES AND QUANTUM CHEMICAL CALCULATIONS

Jelena P. Blagojević Filipović a, Michael B. Hall b, Snežana D. Zarić c,d*

aInnovation

Center of the Faculty of Chemistry, Studentski trg 12-16, Belgrade, Serbia;

bDepartment

of Chemistry, Texas A&M University, College Station, TX 77843-3255, USA;

cFaculty

of Chemistry, University of Belgrade, Studentski trg 12-16, Belgrade, Serbia;

dDepartment

of Chemistry, Texas A&M University at Qatar, P. O. Box 23874, Doha, Qatar

KEYWORDS: resonance-assisted hydrogen-bridged rings, stacking interactions, quantum chemical calculations, Cambridge Structural Database

ABSTRACT

1 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Stacking interactions of resonance-assisted hydrogen-bridged (RAHB) rings are quite common, as 44% of their crystal structures show mutually parallel contacts. High-level quantum chemical calculations by the CCSD(T)/CBS method indicate that these interactions are quite strong, up to -4.7 kcal/mol. This strength is comparable to the stacking interactions of saturated hydrogen-bridged rings (-4.9 kcal/mol), while it is substantially stronger than stacking interaction between two benzene molecules (-2.7 kcal/mol). SAPT energy decomposition analysis shows that the dispersion component makes the major contribution in total interaction energy, but it is mostly cancelled by the exchange-repulsion term in some systems, while electrostatic attraction terms are very significant in all systems. The electrostatic terms can be dominant or similar to the net dispersion term.

Introduction

2 ACS Paragon Plus Environment

Page 2 of 47

Page 3 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Resonance-assisted hydrogen bonds (RAHBs)1 are a special type of hydrogen bonding, where hydrogen donor and acceptor atoms are connected through πconjugated system. The term RAHB highlights the cooperativity between the π-electron delocalization and hydrogen bonds, since there is a synergistic reinforcement between H-bond strengthening and π-delocalization enhancement1–6. Cooperative effects on hydrogen bonding are an extensively studied phenomenon7-10. Resonance stabilization in ring structures containing hydrogen bridges and conjugated double bonds has been observed11 as well as RAHB effects in non-cyclic structures12. X-ray diffraction and spectroscopic data of enol forms of β-diketones, in which the proton-donor and acceptor are linked through a chain of π-conjugated bonds, show shortening of O–H···O bonds, increased enolic d(OH) chemical shifts in 1H NMR spectra, and decreased ν(OH) stretching frequencies in IR spectra3. Similar effect have been observed in o-hydroxy Schiff bases13 and in differently sized rings with RAHB14. Few examples of RAHB systems are shown in Figure 1. The nature of RAHB and the mechanism of synergistic interplay between π-electron delocalization and hydrogen-bonding interactions have been investigated extensively15– 3 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

22.

Page 4 of 47

These studies have concluded that the origin of the stabilization is a charge-

delocalization16, which is

a)

b)

d)

c)

e)

Figure 1. Examples of RAHB systems: a) β-diketone in enol form; b) o-hydroxy Schiff base; c) tenmembered RAHB ring; d) intermolecular RAHB acyclic system; e) intermolecular RAHB cyclic system

solely a consequence of resonance. Namely, the enhanced hydrogen bonding originates mainly from an increased classical dipole-dipole (i.e. electrostatic) attraction, as resonance redistributes the electron density and increases the dipole moments in the monomers, and it is not due to an increase in covalent character of the hydrogen bond, which might have been expected from the electron delocalization15. Delocalization of π-

4 ACS Paragon Plus Environment

Page 5 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

electron density is responsible for the changes in the geometry1, since redistribution of charges changes the structure of σ-skeleton of the system and keeps the HB donor and the HB acceptor coplanar and closer to each other21. The NMR properties do not receive significant contributions directly from resonance, but are a consequence of the σ-skeleton framework22. The fact that RAHB interactions are stronger than classical hydrogen bonds made RAHB systems a very common pattern in crystal architecture. RAHB fragments can be exploited for the rational design of molecular crystals with known packing features and specific physical properties (crystal engineering)23,24. Resonance assisted hydrogen bonding (RAHB) as well as polarization-assisted hydrogen bonding (PAHB), among various types of noncovalent interactions are very frequently used concepts in structural chemistry, features often invoked for stabilized metal-organic frameworks (MOFs)25. RAHB frameworks, organized into infinite chains, are interesting as prototypes of a large family of switching-proton, bistate molecular devices in cases where two tautomeric forms, arising from the proton shift, are present. Another necessary condition for a device is that the shift of the proton must reverse the value of some physical 5 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

property of the system. For example, two tautomeric chains can have opposite dipole moments, which makes them prototypic constituents of ferroelectric or antiferroelectric crystals, as in RAHB chains of β-enolones, β-enaminons, squaric acid26 or dihydrogenphosphates27,28. Another example23 of functional RAHB materials are molecule systems that undergo excited-state proton transfer (ESPT)29–36, where the Hbond assumes tautomeric configurations (X−H∙∙∙Y or X∙∙∙H−Y) according to whether the molecule is in its ground or excited state. If the material is irradiated by light with a suitable wavelength, the equilibrium between the two tautomers is shifted and these systems can be used as energy storage, optical-data storage or luminescent materials29–36. The possibility of forming RAHB can be useful in synthesis, as a driving force for chemical reaction flow, activation of covalent bonds, or as starting moieties in synthetic transformations, as well as synthons in the design of materials23. Stacked arrangement of RAHB rings is observed in crystal structures of salicylideneaniline (SA) derivatives37. RAHB is also present in biomolecules, as in DNA base pairing or secondary protein

6 ACS Paragon Plus Environment

Page 6 of 47

Page 7 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

structures26. It is also observed that intramolecular resonance-assisted hydrogen bonds can be affected by stacking interactions38. Our previous work on the stacking interactions of hydrogen-bridged rings39-41 showed that planar hydrogen-bridged rings, which have only single bonds in the ring (Figure 2a), stack more strongerly (-4.89 kcal/mol) than benzene molecules (-2.73 kcal/mol)39,42. Strength of stacking between a hydrogen-bridged ring and an aromatic ring (-4.38 kcal/mol)41 (Figure 2b) is also stronger than that of benzene molecules. These facts contribute to widening the concept of stacking interactions, since it was shown43,44 that aromaticity is not essential for stacking. It has been also shown that chelates40-55 or non-aromatic hydrocarbons44,61,62 form strong stacking interactions.

(a)

(b)

7 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 2. Examples of stacked dimers of hydrogen-bridged rings; (a) homodimer of planar hydrogen-bridged rings, which have only single bonds in the ring; (b) heterodimer of a hydrogen-bridged ring and benzene

Since the RAHB concept is a very intriguing fundamental scientific topic and has important biological and technological applications, it is important to understand the strength and origin of the noncovalent interactions that are responsible for supramolecular structures in RAHB systems. The goal of this work is to study stacking interactions of RAHB rings in the crystal structures from Cambridge Structural Database (CSD)63, to describe the stacking geometries, and to calculate their interaction energies. Interaction energy potential surfaces were calculated using quantum chemical methods, including CCSD(T)/CBS level that is considered gold standard in quantum chamistry65, while the nature of the interactions was analyzed by the Symmetry Adapted Perturbation Theory (SAPT)69,70.

Methodology 8 ACS Paragon Plus Environment

Page 8 of 47

Page 9 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Crystal structures of six-membered rings formed by intramolecular hydrogen bonds are studied by searching the Cambridge Structural Database (CSD)63. A CSD search is performed by using Con-Quest 1.1864. Geometric constrains applied in the search are shown in Figure 3.

Figure 3. Geometric parameters and atom labeling scheme used for the description of intermolecular interactions of six-membered rings with intramolecular RAHB; Ω1 and Ω2 mark the ring centroids; donor (D) and acceptor (A) atoms include N, O, and S atoms, X, Y and Z are any atoms; R and r mark normal distance and offset value, respectively; θ1 and θ2 angle are torsion angles D1Ω1Ω2D2 and A1Ω1Ω2A2, respectively. Constraints applied in the search to find the RAHB rings were: (1) distances between donor (D) and acceptor (A) atoms within the hydrogen-bridged ring less than 4.0 Å; (2) angles between 9 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

donor (D), hydrogen, and acceptor (A) atoms within the ring from 90° to 180°; (3) absolute torsions AXYZ, XYZD and YZDH within the ring from 0 to 10°; (4) donor (D) and acceptor (A) atoms include N, O, and S atoms; (5) all covalent bonds within rings are set to be acyclic, in order to exclude rigid condensed rings, where stacking arrangement could be the consequence of the condensed ring contacts and not RAHB contacts. The criterion for the intermolecular contacts between rings was that distances between Ω1 and Ω2 id 4.5 Å or less.

The crystallographic R factor is set to be less than 10%, disordered structures are excluded, and coordinates are error-free; according to the criteria used in the CSD, the H-atom positions were normalized using the CSD default X–H bond lengths (O–H = 0.983 Å, C–H = 1.083 Å and N–H = 1.009 Å), and no polymer and powder structures were included. Quantum chemical calculations were performed on three different homodimers, of malonaldehyde (H4C3O2) and its mononitrogen (H5C3NO) and dinitrogen (H4C2N2O) analogs (Figure 4), that were chosen due to the large abundance of their derivatives in

10 ACS Paragon Plus Environment

Page 10 of 47

Page 11 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

CSD. Mutual orientations of the interacting molecules were also based on CSD. Optimizations of monomers were performed at the MP2/cc-pVTZ level. Stacking interaction energies between two rings were calculated at the CCSD(T)/CBS level, obtained by extrapolation by Mackie and DiLabio65, while potential energy curves were obtained by MP2/aug-cc-pVDZ method, since it is in good agreement with CCSD(T)/CBS. In order to find the method suitable for each homodimer, five geometries are chosen for each of the homodimers (a sandwich geometry, two geometries along ΩC direction and two geometries along the orthogonal direction-Figures S5-S7). The interaction energies were calculated for each of them by MP2 and several DFT methods and compared with the energy calculated by CCSD(T)/CBS method (Tables S1-S15). Single-point interaction energy, calculated by MP2 and DFT methods was determined as a difference of the dimer energy and the sum of energies of monomers, having included correction of basis set superposition error (BSSE)66. All calculations were carried out using Gaussian09 series of programs67. Maps of electrostatic potentials for three selected molecules are calculated and visualized from wavefunctions obtained on MP2/cc-pVTZ level of theory, using the Wavefunction Analysis Program (WFA-SAS)68. 11 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 47

Interaction energy decomposition analysis is done on the basis of Symmetry Adapted Perturbation Theory (SAPT2+3)69,70, using PSI4 program package71.

(a)

(b)

(c)

Figure 4. Molecules used for the model systems for quantum chemical calculations of stacking interaction energies; (a) malonaldehyde (H4C3O2) (b) its mononitrogen analog (H5C3NO) and (c) its dinitrogen analog (H4C2N2O)

Results and discussion

Analysis of crystal structures from the CSD

The criteria that define the presence of RAHB rings in crystal structures, as described in Methodology section (Figure 3), are satisfied by 1543 RAHB ring structures. Intermolecular contacts, having distances between two centroids of 4.5 Å or less (Figure 12 ACS Paragon Plus Environment

Page 13 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

3) were found in 678 ring contacts. Parallel orientation, defined by the value of interplanar angle P1/P2 in the range between 0 and 10°, were found in a large number of contacts, namely 617 contacts (91% of all contacts, or 44% of the total number of sixmembered RAHB rings) (Figure 5). Our previous results show that parallel contacts between two hydrogen-bridged rings are also relatively frequent in crystal structures of five-membered hydrogen-bridged rings with only single bonds39,40 and also in crystal structures where C6-aromatic rings and five-membered hydrogen-bridged rings with only single bonds are present41. Hence, the previous results and results from this study indicate that stacked structures are a very common type of supramolecular arrangement in crystals of hydrogen-bridged rings, while RAHB have the largest tendency to form parallel contacts.

13 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Figure 5. Distribution of angle between ring planes in contacts of RAHB rings.

Normal distances of RAHB rings parallel contacts are typical for stacking, since a large group of parallel contacts with offset values smaller than 3.0 Å has interplanar distance between 3.0 Å and 4.0 Å (Figure 6). Similarly, normal distances of parallel contacts between previously studied hydrogen-bridged rings and hydrogen-bridged and C6-aromatic rings, corresponding to horizontal displacements shorter than 3.0 Å, are mostly in the same range39-41. Separate group of parallel contacts can be noticed in crystal structures of RAHB rings at offset values larger than 3.5 Å and normal distances smaller than 2.0 Å (Figure 6).

14 ACS Paragon Plus Environment

Page 14 of 47

Page 15 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 6. Interplanar distance R for RAHB ring contacts having parallel ring planes (P1/P2 < 10°), plotted as a function of offset value r (Figure 3).

Figure 7. Distributions of absolute torsion angles θ1 and θ2 (Figure 3) for parallel contacts of RAHB rings in crystal structures.

Distributions of absolute torsion angles θ1 and θ2, (defined as torsion angles D1Ω1Ω2D2 and A1Ω1Ω2A2, respectively (Figure 3) show that the absolute torsion angles 15 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

are predominantly in the range between 160° and 180°, which implies an antiparallel arrangement of rings (Figure 7). The contribution of contacts corresponding to absolute torsion angles θ1 and θ2 in the range between 0° and 20°, is much smaller than the contribution of antiparallel contacts, but still noticeably larger than for any other angle (Figure 7). These data on the distributions of absolute torsion angles in parallel contacts are similar to our previous work on interactions of two hydrogen-bridged rings with only single bonds39. We analyzed the elemental composition of the RAHB rings that form stacking interactions in crystal structures and the three most numerous groups are shown in Table 1. Molecules that belong to these groups (Figure 4) were chosen for quantum mechanical calculations of stacking energies. In order to find if there are any differences in the interactions of the three groups of rings, analyses of geometric parameters were performed separately for every group (SI). The trends (Figures S1, S2, S3) are similar to overall trends (Figures 5, 6 and 7). Namely, for all three groups most of the rings are in parallel contacts (Figure S1) with anti orientation (Figure S3). These data indicate that the three types of rings, presented in Figure 4 and Table 1, are good representatives for 16 ACS Paragon Plus Environment

Page 16 of 47

Page 17 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

stacking interractions of RAHB rings making them the appropriate model systems for stacking energy calculations.

Table 1. Abundance of the most numerous groups of stacked RAHB rings in the CSD

Number of Stacking Composition of the rings

Atomic Sequence of the Rings (HDZYXA)a

Contacts and Percentage of Total Number of Stacking Contacts

aNotation

H4C3O2

HOCCCO

190 (31%)

H5C3NO

HNCCCO

172 (28%)

H 4C 2N 2O

HNNCCO

107 (17%)

of the atoms is defined in Figure 3.

Phenyl group (regardless of its substitution) is a very common supstituent in structures with RAHB/RAHB contacts (Figure S4); 31% of contacts have at least one phenyl substituent on either D, Z, Y or X atom (Figure 3), while aproximately half of these structures have a parallel alignment of RAHB and phenyl ring planes within the same molecule. The presence of phenyl substituents probably does not govern the

17 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

stacking arrangement of RAHB rings in crystal structures, since, as was mentioned above, the percentage of stacking RAHB contacts is very high (91%).

Quantum chemical calculations of interacting energies

Model systems for quantum chemical calculations are homodimers of the three molecules (Figure 4), chosen because of their frequency in crystal structures (Table 1). In the model systems, the molecules were in the antiparallel arrangement, since that arrangement is preferred in crystal structures (Figure 7). We also calculated the interactions of parallel rings (opposite to antiparallel) (Figure S9, Table S16) and the data show that the interactions in parallel orientation are mostly weaker than the interactions of antiparallel rings (Figures S5-S7, Tables S1-S15), in agreement with results found in CSD. Potential energy curves, calculated at MP2/aug-cc-pVDZ level, along Ω-C direction (where Ω marks the center of the ring and C marks the center of H-A bond, Figure 8) and along the orthogonal direction in the ring plane, are obtained by varying distances

18 ACS Paragon Plus Environment

Page 18 of 47

Page 19 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

between molecular planes, for every particular offset value, in order to find the energy minimum. In these calculations the geometries of the monomers were frozen. The calculated energy minima for each offset value are presented as potential curves in Figure 9, while the geometries corresponding to the strongest calculated interactions for each model system are shown in Figure 10.

(a)

(b)

Figure 8. Representation of directions for the DFT calculation of potential surface; (a) ΩC direction; (b) direction orthogonal to Ω-C.

19 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Potential energy curves along Ω-C direction of H4C3O2 and H5C3NO homodimers have similar shape, with the energy minima at -0.5 Å and -1.0 Å (Figure 9a). The interaction is stronger for H4C3O2 dimer for almost all offset values. The curve for H4C2N2O dimer along Ω-C direction is quite different, with a minimum at -1.8 Å. Interactions in this dimer are the weakest for all horizontal displacements (Figure 9).

(a)

20 ACS Paragon Plus Environment

Page 20 of 47

Page 21 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(b) Figure 9. Potential energy curves for three selected homodimers at MP2/aug-cc-pVDZ level; (a) along Ω-C direction; (b) along the direction orthogonal to Ω-C (Figure 8). The directions are presented in Figure 8.

21 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(a)

(b)

Page 22 of 47

(c)

Figure 10. Two views on geometries corresponding to the strongest calculated stacking interactions of homodimers of (a) H4C3O2 (b) H5C3NO (c) H4C2N2O. All torsion angles are 180°, since the orientation in crystal structures are predominantly antiparallel (Figure 7). The interactions energies are given in Table 2.

The shape of potential energy curves along the direction orthogonal to Ω-C differs for the three homodimers (Figure 9b). The curve for H4C3O2 homodimer shows a minimum 22 ACS Paragon Plus Environment

Page 23 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

corresponding to sandwich geometry, while the curve obtained for H5C3NO homodimer has a minimum that is the strongest stacking interaction energy calculated in this work (CCSD(T)/CBS interactions energy of -4.74 kcal/mol at horizontal displacement of -1.8 Å). The curve corresponding to H4C2N2O dimer along the direction orthogonal to Ω-C has two minima, at -1.9 Å and 1.5 Å. (Figure 9b). Similar to the Ω-C direction (Figure 9a), interactions corresponding to H4C2N2O dimer are the weakest for almost all offset values. (Figure 9b). The calculated interplane distance dependences on horizontal displacements are given in Figure S8. The distances between the planes of stacked RAHB ring dimers are mostly in the range between 3.0 and 3.5 Å, which is in accordance with the data from crystal structures (Figure 6). The most stable geometries calculated in this work for each of the RAHB rings are shown in Figure 10. In the H4C3O2 dimer, the geometry very similar to sandwich geometry (offset value of the two centroids is -0.5 Å along the direction orthogonal to ΩC, Figure 9b) is the most stable.(Figure 10a), while parallel-displaced geometries are favored for the dimers of H5C3NO (offset value of -1.8 Å along the direction orthogonal to Ω-C) and H4C2N2O (offset value of -1.8 Å along the Ω-C direction) (Figures 9, 10b 23 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

and 10c). For these geometries, which represent the most stable geometries for each of the ring dimers, the stacking energies are calculated at the accurate CCSD(T)/CBS level (Table 2). The dimer geometries corresponding to the potential curves minima are optimized (Figure S10). The comparison of hydrogen bond distances in monomer and the optimized dimer shows that the shortening of hydrogen bond occurs as a consequence of stacking, which is already observed in RAHB ring structures38. The strength of stacking interactions of hydrogen-bridged rings containing π-systems, presented in this work (Table 2) are comparable with stacking interactions of hydrogenbridged rings that have only single bonds in the ring39-41. Namely, the strongest calculated interactions of H4C3O2, H5C3NO and H4C2N2O homodimers at CCSD(T)/CBS level are -4.26 kcal/mol, -4.74 kcal/mol and -2.23 kcal/mol, respectively (Table 2), while interaction energy of stacked saturated hydrogen-bridged rings at the same level of theory can be as strong as -4.89 kcal/mol34. Also, the energy of stacking interaction between saturated hydrogen-bridged and C6-aromatic rings are similar (-4.38 kcal/mol)41. All of these interactions are much stronger than stacking of benzene 24 ACS Paragon Plus Environment

Page 24 of 47

Page 25 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

molecules (-2.73 kcal/mol)42. These data indicate that delocalized π-systems do not form stronger stacking interactions than systems without π-bonds.

Table 2. The energies of the most stable stacking interactions of the three homodimers, calculated at CCSD(T)/CBS level, with the corresponding offset values and interplanar distances. Geometries of these minima are given in Figure 10.

model system

direction

ΔE

r

R

(kcal/mol)

(Å)

(Å)

H4C3O2

Ω-C

-4.26

-0.5

3.30

H5C3NO

orthogonal to Ω- -4.74

-1.8

3.10

-1.8

3.35

C H 4C 2N 2O

Ω-C

-2.23

These results are in line with some previous results44,62, which show that the presence of π-systems does not necessarily influence the energies of stacking interactions. For example, energies of stacking interactions of small saturated cycloalkanes and small aromatic hydrocarbons are similar44. Interaction in cyclohexane dimer (-2.62 kcal/mol)61 is slightly weaker than interaction in benzene dimer (-2.73 kcal/mol)42, while stacking

25 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 26 of 47

interaction in heterodimer of benzene and cyclohexane is slightly stronger (-3.27 kcal/mol)62.

Quantum chemical calculations of electrostatic potentials

In order to understand the geometries of minima on the potential curves, as well as the calculated energies, electrostatic potential maps (Figure 11) for the three selected molecules (Figure 4), were calculated with WFA-SAS program68.

(a)

(b)

26 ACS Paragon Plus Environment

(c)

Page 27 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Figure 11. Electrostatic potential maps of (a) H4C3O2 (b) H5C3NO (c) H4C2N2O molecules.

Although the electrostatic potential maps for the three molecules are quite different, they have some common features. The potential above the rings is mainly negative; it is the most negative in H5C3NO molecule, while it is partially positive in the other two molecules. As one can anticipate, the potential around electronegative oxygen and nitrogen atoms are negative, while potential around or on the edges near hydrogen atoms are positive. The potentials around hydrogen atoms forming hydrogen bond have quite small positive area, as was noted in previous work 72,73. The different electrostatic potentials of the three molecules (Figure 11) cause different geometries of the potential curves minima (Figure 9). The geometry similar to the antiparallel sandwich geometry is the most stable in H4C3O2 dimer (Figure 10a) since it enables very efficient overlay of positive and negative regions of the molecule. Namely, the electrostatic potential show that one part of the molecule has dominant positive

27 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

potential, while the opposite part has dominant negative potential (Figure 11a) and the most stable geometry enables overlay of these two parts (Figure 10a). For H5C3NO and H4C2N2O homodimers, parallel-displaced geometries are preferred (Figure 10b and c), since quite different electrostatic potential of these molecules cause that these geometries enable overlay of positive and negative potentials (Figure 11b and c). The antiparallel arrangement of the interacting rings observed in crystal structures (Figure 7), can be explained by these electrostatic potential maps (Figure 11).

SAPT decomposition analyses of interaction energies

SAPT energy decomposition analyses are performed in order to reveal the nature of stacking interactions in RAHB ring dimers. For the analyzes we used geometries corresponding to the strongest stacking interactions calculated in this study (Figure 10). The calculations were performed with SAPT2+3/cc-pVQZ method, since the interaction

28 ACS Paragon Plus Environment

Page 28 of 47

Page 29 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

energies calculated with this method are in good agreement with CCSD(T)/CBS interaction energies (Tables 2 and 3). Energy decompositions obtained with the SAPT2+3/cc-pVQZ method, show similar energy contributions for H4C3O2 and H5C3NO homodimer systems (Table 3), since these two systems have a dominant attractive dispersion contribution, which is, however, almost surpassed by the exchange-repulsion term, leading to slightly negative (attractive) net dispersion in H4C3O2 dimer and slightly positive (repulsive) net dispersion in H5C3NO dimer. The electrostatic contributions are relatively large in these systems (-3.55 kcal/mol and -4.81 kcal/mol), and they are the main contribution to the strength of the interactions. Additional stability is achieved by the induction term that is slightly attractive in both cases (Table 3). Significantly weaker interaction in H4C2N2O dimer is a consequence of the significantly smaller electrostatic contribution (-0.89 kcal/mol, Table 3) than in the other two dimers (Table 2). Unlike the H4C3O2 and H5C3NO dimers, the net dispersion term has the similar contribution to electrostatics (-0.88 kcal/mol, Table 3) in H4C2N2O dimer.

29 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 30 of 47

Table 3. SAPT2+3/cc-pVQZ energy decomposition analysis of the H4C3O2, H5C3NO and H4C2N2O stacked homodimers (Figures 9 and 10). The total SAPT2+3 interaction energy is composed of electrostatic (ELST), exchange-repulsion (EXCH), induction (IND) and dispersion (DISP) terms. Net dispersion (NET DISP) is a sum of dispersion (DISP) and exchange-repulsion (EXCH) terms74,75.

Energy Terms (kcal/mol) Dimer Systems

ELST

EXCH

IND

DISP

NET DISP

Total SAPT2+ 3

H4C3O2

-3.55

5.92

-0.60

-6.10

-0.18

-4.32

H5C3NO

-4.81

7.86

-1.12

-6.74

1.12

-4.81

H 4C 2N 2

-0.89

3.76

-0.33

-4.64

-0.88

-2.09

O

As in the H4C3O2 and H5C3NO RAHB stacked homodimers, the electrostatic term in parallel-displaced benzene/benzene, pyridine/benzene and pyridine/pyridine stacked systems74, is the largest contribution to the total interaction energy, while the net dispersion plus induction terms are much less significant. However, for the H4C3O2 and H5C3NO stacked dimers, the electrostatic terms are larger than that for aromatic 30 ACS Paragon Plus Environment

Page 31 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

molecules74. Results of SAPT2+3 analysis for parallel-displaced orientation of benzene stacking with a hydrogen-bridge ring with single bonds (C2H2N3S) show the interactions to be similar to those RAHB stacked dimers (since, the electrostatics is dominant, while net dispersion term is much less significant)41. Conclusion

Stacking interactions are observed for RAHB rings in their crystal structures and these interactions are quite frequent, 44% of all RAHB rings in the CSD form stacking interactions. Distances between ring planes are typical for staking interactions (3.0-4.0 Å), while the alignment is antiparallel. The most frequent RAHB rings forming stacking in the crystal structures are H4C3O2, H5C3NO, and H4C2N2O, hence we used these molecules as model systems to calculate their interaction energies. The calculated energies of stacking interactions at CCSD(T)/CBS level are -4.26 kcal/mol for the H4C3O2 dimer, -4.74 kcal/mol for the H5C3NO dimer, and -2.23 kcal/mol for the H4C2N2O dimer. The preferred geometry of the H4C3O2 dimer is nearly sandwich geometry, while the preferred geometries of the H5C3NO and H4C2N2O dimers are 31 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

parallel-displaced. The preferred stacking geometries can be explained by the calculated electrostatic potential maps. The SAPT analysis showed that the attractive dispersion contribution is dominant in all three systems studied in this work, however, it is almost cancelled by exchange-repulsion in H4C3O2 and H5C3NO homodimers, while in H4C2N2O homodimer the attractive dispersion term is significantly larger than the exchange-repulsion, contributing to net attraction. The electrostatic contribution is significant in all systems and it is the main contribution to the net attraction in H4C3O2 and H5C3NO homodimers, while the influences of electrostatics and net dispersion are almost equal in H4C2N2O dimer. These stacking interaction energies are comparable with previously published stacking energies between hydrogen-bridged rings which do not possess π-system. That is indication that the presence of a π-system is not necessary for strong stacking and that stacking interaction energies can be similar in saturated and unsaturated systems. Furthermore, the stacking of RAHB rings can be stronger than the stacking of benzene molecules (-2.7 kcal/mol), which makes RAHB species promising building blocks for various supramolecular assemblies. 32 ACS Paragon Plus Environment

Page 32 of 47

Page 33 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

ASSOCIATED CONTENT

Supporting Information

The following file are available free of charge.

Additional analyses of the crystal structures and the dimer structures, evaluation of quantum chemical methods (PDF)

AUTHOR INFORMATION

Corresponding Author *Snežana Zarić ([email protected])

Author Contributions.

Funding Sources

33 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

This work was supported by the Serbian Ministry of Education, Science and Technological Development (Grant 172065) and an NPRP grant from the Qatar National Research Fund (a member of the Qatar Foundation) (grant number NPRP8425-1-087).

Notes The authors declare no competing financial interest.

ACKNOWLEDGMENT

The HPC resources and services used in this work were partially provided by the IT Research Computing group in Texas A&M University at Qatar. IT Research Computing is funded by the Qatar Foundation for Education, Science and Community Development (http://www.qf.org.qa).

ABBREVIATIONS 34 ACS Paragon Plus Environment

Page 34 of 47

Page 35 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

RAHB, Resonance-assisted Hydrogen Bond; 1H NMR, Proton Nuclear Magnetic Resonance; IR, Infrared; HB, Hydrogen Bond; NMR, Nuclear Magnetic Resonance; PAHB, Polarizationassisted Hydrogen Bond MOF, Metal-organic Framework; ESPT, Excited-state Proton Transfer DNA, Deoxyribonucleic Acid; SA, Salicylideneaniline CSD; Cambridge Structural Database; MP2, Møller-Plesset Perturbation Theory of second order; CCSD(T), Coupled-Cluster with Single, Double and Perturbative Triple excitations; CBS, Complete Basis Set; BSSE, Basis Set Superposition Error; WFA-SAS, Wavefunction Analysis-Surface Analysis Suite; DFT, Density Functional Theory; SI, Supporting Information; ESP, Electrostatic Potential, SAPT, Symmetry Adapted Perturbation Theory

REFERENCES

1)

Gilli, G.; Bellucci, F.; Ferretti, V.; Bertolasi, V. Evidence for Resonance-Assisted Hydrogen Bonding from Crystal-Structure Correlations on the Enol Form of the .Beta.Diketone Fragment. J. Am. Chem. Soc. 1989, 111 (3), 1023–1028.

(2)

Gilli, P.; Bertolasi, V.; Ferretti, V.; Gilli, G. Evidence for Intramolecular N-H···O Resonance-Assisted Hydrogen Bonding in β-Enaminones and Related Heterodienes. A Combined Crystal-Structural, IR and NMR Spectroscopic, and Quantum-Mechanical Investigation. J. Am. Chem. Soc. 2000, 122 (42), 10405–10417.

(3)

Gilli, P.; Bertolasi, V.; Pretto, L.; Lyčka, A.; Gilli, G. The Nature of Solid-State NH⋯O/O-H⋯N Tautomeric Competition in Resonant Systems. Intramolecular Proton Transfer in Low-Barrier Hydrogen Bonds Formed by the ⋯O=C-C=N-NH⋯⇄⋯HOC=C-N=N⋯ Ketohydrazone-Azoenol System. A Variable-Temperature x-Ray 35 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystallograp. J. Am. Chem. Soc. 2002, 124 (45), 13554–13567. (4)

Bertolasi, V.; Gilli, P.; Ferretti, V.; Gilli, G. Evidence for Resonance-Assisted Hydrogen Bonding. 2.1Intercorrelation between Crystal Structure and Spectroscopic Parameters in Eight Intramolecularly Hydrogen Bonded 1,3-Diary 1–1,3-Propanedione Enols. J. Am. Chem. Soc. 1991, 113 (13), 4917–4925.

(5)

Gilli, P.; Bertolasi, V.; Pretto, L.; Ferretti, V.; Gilli, G. Covalent versus Electrostatic Nature of the Strong Hydrogen Bond: Discrimination among Single, Double, and Asymmetric Single-Well Hydrogen Bonds by Variable-Temperature X-Ray Crystallographic Methods in β-Diketone Enol RAHB Systems. J. Am. Chem. Soc. 2004, 126 (12), 3845–3855.

(6)

Gilli, P.; Bertolasi, V.; Ferretti, V.; Gilli, G. Covalent Nature of the Strong Homonuclear Hydrogen Bond. Study of the O-H⃨O System by Crystal Structure Correlation Methods1. J. Am. Chem. Soc. 1994, 116 (3), 909–915.

(7)

Stare, J.; Hadži, D. Cooperativity Assisted Shortening of Hydrogen Bonds in Crystalline Oxalic Acid Dihydrate: DFT and NBO Model Studies. J. Chem. Theory Comput. 2014, 10, 1817−1823.

(8)

Vishweshwar, P.; Babu, N. J.; Nangia, A.; Mason, S. A.; Puschmann, H.; Mondal, R.; Howard, J. A. K. Variable Temperature Neutron Diffraction Analysis of a Very Short OH···O Hydrogen Bond in 2,3,5,6-Pyrazinetetracarboxylic Acid Dihydrate: SynthonAssisted Short Oacid-H···Owater Hydrogen Bonds in a Multicenter Array. J. Phys. Chem. A 2004, 108, 9406-9416.

36 ACS Paragon Plus Environment

Page 36 of 47

Page 37 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

(9)

Monteath Robertson, J. ; Ubbelohde, A. R. Structure and thermal properties associated with some hydrogen bonds in crystals. I. The isotope effect. Proc. R. Soc. A 1939, 170, 222-240.

(10) King, B. F.;,Farrar, T. C.;, Weinhold, F. Quadrupole coupling constants in linear (HCN)n clusters: Theoretical and experimental evidence for cooperativity effects in C–H⋅⋅⋅N hydrogen bonding J. Chem. Phys. 1995, 103, 348. (11)

Huggins, M. L. Hydrogen Bridges in Organic Compounds. J. Org. Chem. 1936, 1 (5), 407–456.

(12)

Trujillo, C.; Sánchez-Sanz, G.; Alkorta, I.; Elguero, J.; Mó, O.; Yáñez, M. Resonance Assisted Hydrogen Bonds in Open-Chain and Cyclic Structures of Malonaldehyde Enol: A Theoretical Study. J. Mol. Struct. 2013, 1048, 138–151.

(13)

Krygowski, T. M.; Zachara-Horeglad, J. E.; Palusiak, M.; Pelloni, S.; Lazzeretti, P. Relation between π-Electron Localization/Delocalization and H-Bond Strength in Derivatives of o-Hydroxy-Schiff Bases. J. Org. Chem. 2008, 73 (6), 2138–2145.

(14)

Mohajeri, A. Theoretical Evidences for Resonance-Assisted Hydrogen Bonding. J. Mol. Struc. Theochem 2004, 678 (1–3), 201–205.

(15)

Beck, J. F.; Mo, Y. How Resonance Assists Hydrogen Bonding Interactions: An Energy Decomposition Analysis. J. Comput. Chem. 2007, 28 (1), 455–466.

(16)

Góra, R. W.; Maj, M.; Grabowski, S. J. Resonance-Assisted Hydrogen Bonds Revisited. Resonance Stabilization vs. Charge Delocalization. Phys. Chem. Chem. Phys. 2013, 15 (7), 2514–2522. 37 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(17)

Grabowski, S. J. π-Electron Delocalisation for Intramolecular Resonance Assisted Hydrogen Bonds. J. Phys. Org. Chem. 2003, 16 (10), 797–802.

(18)

Lyssenko, K. A.;Antipin, M. Y. The Nature and Energy Characteristics of Intramolecular Hydrogen Bonds in Crystals. Russ. Chem. Bull., Int. Ed. 2006, 55 (1), 1–15.

(19)

Nowroozi, A.; Roohi, H.; Sadeghi Ghoogheri, M. S.; Sheibaninia, M. The Competition between the Intramolecular Hydrogen Bond and π-Electron Delocalization in Trifluoroacetylacetone-A Theoretical Study. Int. J. Quantum Chem. 2011, 111 (3), 578– 585.

(20)

Sobczyk, L.; Grabowski, S. J.; Krygowski, T. M. Interrelation between H-Bond and πElectron Delocalization. Chem. Rev. 2005, 105 (10), 3513–3560.

(21)

Sanz, P.; Mó, O.; Yáñez, M.; Elguero, J. Resonance-Assisted Hydrogen Bonds: A Critical Examination. Structure and Stability of the Enols of β-Diketones and β-Enaminones. J. Phys. Chem. A 2007, 111 (18), 3585–3591.

(22)

Alkorta, I.; Elguero, J.; Mó, O.; Yáñez, M.; Bene, J. E. D. Are Resonance-Assisted Hydrogen Bonds “Resonance Assisted”? A Theoretical NMR Study. Chem. Phys. Lett. 2005, 411 (4–6), 411–415.

(23)

Mahmudov, K. T.; Pombeiro, A. J. L. Resonance‐Assisted Hydrogen Bonding as a Driving Force in Synthesis and a Synthon in the Design of Materials. Chem. Eur. J. 2016, 22, 16356–16398.

(24)

Bertolasi, V.; Gilli, P.; Ferretti, V.; Gilli, G. Resonance-Assisted O-H⋯O Hydrogen Bonding: Its Role in the Crystalline Self-Recognition of β-Diketone Enols and Its 38 ACS Paragon Plus Environment

Page 38 of 47

Page 39 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Structural and IR Characterization. Chem. - A Eur. J. 1996, 2 (8), 925–934. (25)

Mirzaei, M.; Eshtiagh-Hosseini, H.; Shamsipur, M.; Saeedi, M.; Ardalani, M.; Bauzá, A.; Mague, J. T.; Frontera, A.; Habibi, M. Importance of Polarization Assisted/Resonance Assisted Hydrogen Bonding Interactions and Unconventional Interactions in Crystal Formations of Five New Complexes Bearing Chelidamic Acid through a Proton Transfer Mechanism. RSC Adv. 2015, 5 (89), 72923–72936.

(26)

Gilli, G.; Bertolasi, V.; Ferretti, V.; Gilli, P. Resonance‐assisted Hydrogen Bonding. III. Formation of Intermolecular Hydrogen‐bonded Chains in Crystals of Β‐diketone Enols and Its Relevance to Molecular Association. Acta Crystallogr. Sect. B 1993, 49 (3), 564– 576.

(27)

Lasave, J.; Abufager, P.; Koval, S. Ab Initio Study of the One-Dimensional H-Bonded Ferroelectric CsH2PO4. Phys. Rev. B 2016, 93 (13), 134112.

(28)

Lines, M. E.; Glass, A. M. Principles and Applications of Ferroelectrics and Related Materials; Oxford University Press, 2001.

(29)

Ernsting, N. P.; Arthen-Engeland, T.; Rodriguez, M. A.; Thiel, W. State-Selectivity of Excited-State Intramolecular Proton Transfer in a “Double” Benzoxazole: Jet Spectroscopy and Semiempirical Calculations. J. Chem. Phys. 1992, 97 (6), 3914–3919.

(30)

Abou-Zied, O. K.; Jimenez, R.; Thompson, E. H. Z.; Millar, D. P.; Romesberg, F. E. Solvent-Dependent Photoinduced Tautomerization of 2-(2′-Hydroxyphenyl)Benzoxazole. J. Phys. Chem. A 2002, 106 (15), 3665–3672.

(31)

Padalkar, V. S.; Seki, S. Excited-State Intramolecular Proton-Transfer (ESIPT)-Inspired 39 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Solid State Emitters. Chem. Soc. Rev. 2016, 45 (1) 169–202. (32)

Hammes-Schiffer, S.; Stuchebrukhov, A. A. Theory of Coupled Electron and Proton Transfer Reactions. Chem. Rev. 2010, 110 (12), 6939–6960.

(33)

Hammes-Schiffer, S. Introduction: Proton-Coupled Electron Transfer. Chem. Rev. 2010, 110 (12) 6937–6938.

(34)

Elsaesser, T.; Bakker, H. J. Ultrafast Dynamics of Hydrogen Bonding and Proton Transfer in the Condensed Phase. In Ultrafast Hydrogen Bonding Dynamics and Proton Transfer Prosesses in the Condensed Phase; Springer Netherlands: Dordrecht, 2002.

(35)

Kwon, J. E.; Park, S. Y. Advanced Organic Optoelectronic Materials: Harnessing ExcitedState Intramolecular Proton Transfer (ESIPT) Process. Adv. Mater. 2011, 23 (32) 3615– 3642.

(36)

Dick, B. Fluorescence Studies of Excited-State Intramolecular Proton-Transfer (ESIPT) in Molecules Isolated in Solid Argon - 3-hydroxyflavone and 2,5-bis(2benzoxazolyl)hydroquinone. Ber. Bunsen Phys. Chem. 1987, 91 (11), 1205–1209.

(37)

Karabyk, H.; Karabyk, H.; Ocak Skeleli, N. Hydrogen-Bridged Chelate Ring-Assisted πStacking Interactions. Acta Crystallogr. Sect. B Struct. Sci. 2012, 68 (1), 71–79.

(38)

Panek, J.; Stare, J.; Hadži, D. From the Isolated Molecule to Oligomers and the Crystal: A Static Density Functional Theory and Car-Parrinello Molecular Dynamics Study of Geometry and Potential Function Modifications of the Short Intramolecular Hydrogen Bond in Picolinic Acid N-Oxide. J. Phys. Chem. A 2004, 108, 7417-7423.

(39)

Blagojević, J. P.; Zarić, S. D. Stacking Interactions of Hydrogen-Bridged Rings-Stronger 40 ACS Paragon Plus Environment

Page 40 of 47

Page 41 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

than the Stacking of Benzene Molecules. Chem. Commun. 2015, 51 (65), 12989–12991. (40)

Blagojević, J. P.; Janjić, G. V.; Zarić, S. D. Very Strong Parallel Interactions Between Two Saturated Acyclic Groups Closed with Intramolecular Hydrogen Bonds Forming Hydrogen-Bridged Rings. Crystals 2016, 6 (4), 34.

(41)

Blagojević, J. P.; Veljković, D. Ž.; Zarić, S. D. Stacking Interactions between HydrogenBridged and Aromatic Rings: Study of Crystal Structures and Quantum Chemical Calculations. CrystEngComm 2017, 19 (1), 40–46.

(42)

Lee, E. C.; Kim, D.; Jurecka, P.; Tarakeshwar, P.; Hobza, P.; Kim, K. S. Understanding of Assembly Phenomena by Aromatic−Aromatic Interactions:  Benzene Dimer and the Substituted Systems. J. Phys. Chem. A, 2007, 111 (18), 3446 -3457.

(43)

Bloom, J. W. G.; Wheeler, S. E. Taking the Aromaticity out of Aromatic Interactions. Angew. Chemie - Int. Ed. 2011, 50 (34), 7847–7849.

(44)

Grimme, S. Do Special Noncovalent π-π Stacking Interactions Really Exist? Angew. Chemie - Int. Ed. 2008, 47 (18), 3430–3434.

(45)

Gui, L. C.; Wang, X. J.; Yang, K. G.; Bi, X. S. Synthesis, Structures and Characterization of a Series of Cu(i)-Diimine Complexes with Labile N,N′Bis((Diphenylphosphino)Methyl)Naphthalene-1,5- Diamine: Diverse Structures Directed by π-π Stacking Interactions. New J. Chem. 2011, 35 (11), 2471–2476.

(46)

Malenov, D. P., Janjić, G. V ; Medaković, V. B.; Hall, M. B.; Zarić, S. D. Noncovalent bonding:Stacking interactions of chelate rings of transition metal complexes. Coord. Chem. Rev. 2017, 345, 318-341. 41 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(47)

Petrović, P. V.; Janjić, G. V.; Zarić, S. D. Cryst. Growt Des. 2014, 14, 3880–3889.

(48)

Malenov, D.P.; Hall, M. B.; Zarić, S. D. Influence of metal ion on chelate–aryl stacking interactions. Int J Quantum Chem. 2018, 118 (16), e25629.

(49)

Malenov, D. P.; Zarić, S. D. Chelated metal ions modulate the strength and geometry of stacking interactions: energies and potential energy surfaces for chelate–chelate stacking Phys.Chem.Chem.Phys. 2018, 20, 14053-14060.

(50)

Basu Baul, T. S.; Kundu, S.; Höpfl, H.; Tiekink, E. R. T.; Linden, A. Synthesis, Structures, and Spectroscopic Properties of Hg(II) Complexes of Bidentate NN and Tridentate NNO Schiff-Base Ligands. J. Coord. Chem. 2014, 67 (6), 1061–1078.

(51)

McKay, A. P.; Lo, W. K. C.; Preston, D.; Giles, G. I.; Crowley, J. D.; Barnsley, J. E.; Gordon, K. C.; McMorran, D. A. Palladium(II) and Platinum(II) Complexes of ((2Pyridyl)Pyrazol-1-Ylmethyl)Benzoic Acids: Synthesis, Solid State Characterisation and Biological Cytotoxicity. Inorganica Chim. Acta 2016, 446, 41–53.

(52)

Basu Baul, T. S.; Kundu, S.; Ng, S. W.; Guchhait, N.; Tiekink, E. R. T. Synthesis, Characterization, Photoluminescent Properties and Supramolecular Aggregations in Diimine Chelated Cadmium Dihalides. J. Coord. Chem. 2014, 67 (1), 96–119.

(53)

Basu Baul, T. S.; Kundu, S.; Höpfl, H.; Tiekink, E. R. T.; Linden, A. Synthesis, Crystal Structures and Optical Properties of Mercury(II) Halide Compounds with (E)-N-(Pyridin2-Ylmethylidene)Arylamines: Effect of Ligand R-Group upon Structure. Polyhedron 2013, 55, 270–282.

(54)

Wang, X. J.; Jian, H. X.; Liu, Z. P.; Ni, Q. L.; Gui, L. C.; Tang, L. H. Assembly 42 ACS Paragon Plus Environment

Page 42 of 47

Page 43 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

Molecular Architectures Based on Structural Variation of Metalloligand [Cu(PySal)2] (PySal = 3-Pyridylmethylsalicylideneimino). Polyhedron 2008, 27 (12), 2634–2642. (55) Malenov, D. P.; Veljković, D. Ž.; Hall, M. B.; Brothers, E. N.; Zarić, S. D. Influence of Chelate Ring Type on Chelate-Chelate and Chelate-Aryl Stacking: The Case of Nickel Bis(Dithiolene), Phys Chem Chem Phys, 2018. DOI: 10.1039/C8CP06312E. (56)

Chakrabarty, P. P.; Biswas, D.; García-Granda, S.; Jana, A. D.; Saha, S. Sodium Ion Assisted Molecular Self-Assembly in a Class of Schiff-Base Copper(II) Complexes. Polyhedron 2012, 35 (1), 108–115.

(57)

Biswas, D.; Chakrabarty, P. P.; Saha, S.; Jana, A. D.; Schollmeyer, D.; García-Granda, S. Ligand Mediated Structural Diversity and Role of Different Weak Interactions in Molecular Self-Assembly of a Series of Copper(II)-Sodium(I) Schiff-Base Heterometallic Complexes. Inorg. Chim. Acta 2013, 408, 172–180.

(58)

Jana, S.; Khan, S.; Bauzá, A.; Frontera, A.; Chattopadhyay, S. The Crucial Role of Chelate-Chelate Stacking Interactions in the Crystal Structure of a Square Planar Copper(II) Complex. J. Mol. Struct. 2017, 1127, 355–360.

(59)

Bhattacharya, S.; Roy, S.; Harms, K.; Bauza, A.; Frontera, A.; Chattopadhyay, S. Importance of Chelate-Chelate Stacking Interactions in Crystal Structures of Square Pyramidal Copper(II) Complexes with Two Distinct Chelating Bidentate Ligands. Inorganica Chim. Acta 2016, 442, 16–23.

(60)

Mahmoudi, G.; Castiñeiras, A.; Garczarek, P.; Bauzá, A.; Rheingold, A. L.; Kinzhybalo, V.; Frontera, A. Synthesis, X-Ray Characterization, DFT Calculations and Hirshfeld

43 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Surface Analysis of Thiosemicarbazone Complexes of M N+ Ions (n = 2, 3; M = Ni, Cd, Mn, Co and Cu). CrystEngComm 2016, 18 (6), 1009–1023. (61)

Kim, K. S.; Karthikeyan, S.; Singh, N. J. How Different Are Aromatic π Interactions from Aliphatic π Interactions and Non-π Stacking Interactions? J. Chem. Theory Comput. 2011, 7 (11), 3471–3477.

(62)

Ninković, D. B.; Vojislavljević-Vasilev, D. Z.; Medaković, V. B.; Hall, M. B.; Brothers, E. N.; Zarić, S. D. Aliphatic-Aromatic Stacking Interactions in Cyclohexane-Benzene Are Stronger than Aromatic-Aromatic Interaction in the Benzene Dimer. Phys. Chem. Chem. Phys. 2016, 18 (37), 25791–25795.

(63)

Allen, F. H. The Cambridge Structural Database: a quarter of a million crystal structures and rising. Acta Cryst. 2002, B58, 380-388.

(64)

Bruno, I. J.; Cole, J. C.; Edgington, P. R.; Kessler, M.; Macrae, C. F.; McCabe, P.; Pearson, J.; Taylor, R. New software for searching the Cambridge Structural Database and visualizing crystal structures. Acta Cryst. Sect. B-Struct. Sci. 2002, 58, 389–397.

(65)

Mackie, I. D.; DiLabio, G. A. Approximations to Complete Basis Set-Extrapolated, Highly Correlated Non-Covalent Interaction Energies. J. Chem. Phys. 2011, 135 (13), 134318.

(66)

Boys, S. F.; Bernardi, F. The Calculation of Small Molecular Interactions by the Differences of Separate Total Energies. Some Procedures with Reduced Errors. Mol. Phys. 1970, 19 (4), 553–566.

(67)

Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. 44 ACS Paragon Plus Environment

Page 44 of 47

Page 45 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; et al. Gaussian 09, Revision D.01. 2016. (68)

Bulat, F. A.; Toro-Labbé, A.; Brinck, T.; Murray, J. S.; Politzer, P. Quantitative Analysis of Molecular Surfaces: Areas, Volumes, Electrostatic Potentials and Average Local Ionization Energies. J. Mol. Model. 2010, 16 (11), 1679–1691.

(69)

Jeziorski, B.; Moszynski, R.; Szalewicz, K. Perturbation Theory Approach to Intermolecular Potential Energy Surfaces of van der Waals Complexes. Chem. Rev.1994, 94 (7), 1887-1930.

(70)

Hohenstein, E. G.; Sherrill, C. D. Density fitting of intramonomer correlation effects in symmetry-adapted perturbation theory. J. Chem. Phys.2010, 133 (1), 014101.

(71)

Parrish, R. M.; Burns, L. A.; Smith, D. G. A.; Simmonett, A. C.; DePrince, A. E.; Hohenstein, E. G.; Bozkaya, U.; Sokolov, A. Y.; Di Remigio, R.; Richard, R. M.; Gonthier, J. F.; James, A. M.; McAlexander, H. R.; Kumar, A.; Saitow, M.; Wang, X.; Pritchard, B. P.; Verma, P.; Schaefer, H. F.; Patkowski, K.; King, R. A.; Valeev, E. F.; Evangelista, F. A.; Turney, J. M.; Crawford, T. D.; Sherrill, C. D. PSI4 1.1: An OpenSource Electronic Structure Program Emphasizing Automation, Advanced Libraries, and Interoperability. J. Chem. Theory Comput. 2017, 13 (7), 3185–3197.

(72) Politzer, P.; Murray, J. S. Quantitative Analyses of Molecular Surface Electrostatic Potentials in Relation to Hydrogen Bonding and Co-Crystallization. Cryst. Growth Des. 2015, 15, 3767−3774. 45 ACS Paragon Plus Environment

Crystal Growth & Design 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(73) Hussein, W.; Walker, C. G.; Peralta-Inga, Z.; Murray, J. S. Computed Electrostatic Potentials andAverage Local Ionization Energies on theMolecular Surfaces of Some Tetracyclines. Int J Quantum Chem.. 2001, 82, 160-169. (74) Hohenstein, E. G.; Sherrill, C. D. Effects of Heteroatoms on Aromatic π-π Interactions: Benzene-Pyridine and Pyridine Dimer. J. Phys. Chem. A 2009, 113 (5), 878-886. (75) Sherrill, C. D. Energy Component Analysis of π Interactions. Acc. Chem. Res. 2013, 46 (4), 1020–1028.

46 ACS Paragon Plus Environment

Page 46 of 47

Page 47 of 47 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Crystal Growth & Design

For Table of Contents Use Only

STACKING INTERACTIONS OF RESONANCE-ASSISTED HYDROGENBRIDGED RINGS. A SYSTEMATIC STUDY OF CRYSTAL STRUCTURES AND QUANTUM CHEMICAL CALCULATIONS

Jelena P. Blagojević Filipović, Michael B. Hall, Snežana D. Zarić

Analysis of crystal structures from Cambridge Structural Database reveals that structures with stacked RAHB rings are quite common, while high-level interaction energy calculations on dimer model systems show that interaction energies depend on the composition of the ring and can be quite strong, much stronger than stacking interaction between two benzene molecules (2.73 kcal/mol).

47 ACS Paragon Plus Environment