Structure–Spectra Correlations in Anilate Complexes with Picolines

Aug 26, 2016 - Baroni , S.; Dal Corso , A.; de Gironcoli , S.; Giannozzi , P. Rev. Mod. Phys. 2001, 73, 515– 561 DOI: 10.1103/RevModPhys.73.515. [Cr...
0 downloads 0 Views 3MB Size
Subscriber access provided by United Arab Emirates University | Libraries Deanship

Article

Structure-Spectra Correlations in Anilate Complexes with Picolines Katarzyna #uczy#ska, Kacper Dru#bicki, Krzysztof Lyczko, and Jan Cz. Dobrowolski Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.6b01114 • Publication Date (Web): 26 Aug 2016 Downloaded from http://pubs.acs.org on September 1, 2016

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 free 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 accessible to all readers and 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.

Crystal Growth & Design 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 38

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

Structure-Spectra Correlations in Anilate Complexes with Picolines Katarzyna Łuczyńska,†,‡, * Kacper Drużbicki,‡,¶ Krzysztof Lyczko,† Jan Cz. Dobrowolski,†,§



Institute of Nuclear Chemistry and Technology, Dorodna 16, 03-195, Warsaw, Poland;



Frank Laboratory of Neutron Physics, Joint Institute for Nuclear Research, 141980, Dubna, Russian Federation;



Department of Radiospectroscopy, Faculty of Physics, Adam Mickiewicz University, Umultowska 85, 61-614, Poznan,

Poland; §

National Medicines Institute, 30/34 Chełmska Street, 00-725, Warsaw, Poland.

*Corresponding author. e-mail: [email protected]

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

Page 2 of 38

I. INTRODUCTION In the era of rapid progress in supramolecular chemistry, low-weight organic molecular crystals have attracted much attention in the design of new functional materials, where donoracceptor type molecular complexes are particularly interesting. Since they reveal proton and electron transfer phenomena, they exhibit polar and semiconductive properties. From a macroscopic perspective, proton affected by heavy atoms reflects a considerable temperature-dependent dynamics, which may break the system centrosymmetry.1-3 Donor-acceptor properties can also be associated with electron transfer. In such a case, the molecule of an organic acid in conjunction with an electron-rich system can behave as a charge acceptor, forming mixed, sandwich-type chargetransfer stacks. The collective electron transfer can also result in dimerization into electron donoracceptor pairs, which can further break the centrosymmetry. Such a mechanism was found in tetrathiafulvalene chloranilate (TTF : ClA), a complex which shows fascinating ferroelectric properties at low temperatures.4 The aforementioned phenomena are, however, still barely understood and concern the most-exciting issues in modern condensed matter physics, including quantum ferroelectricity or nuclear quantum effects .5-7 There is tremendous interest in complexes of organic aromatic acids with heterocyclic aromatic amines, from which the family of anilic acids (2,5-dihalogen-1,4-benzoquinone; hereafter XA) was found to be of vital importance. Anilic acids have been used in combination with various secondary molecular components to produce chains, sheets and three-dimensional structures with a reasonable degree of design intent.8 Low-weight anilate synthons utilize a structural synergy between hydrogenbonded networks and charge-transfer stacks, which self-assemble into well-ordered architectures. While solid-state spectroscopy is the method of choice for characterization of supramolecular objects, there have still been very few attempts to correlate crystal structures with their spectral responses. Modern solid-state first-principles calculations have become a powerful tool for such analyses.9-13 In this regard, several systems are analyzed here from both the experimental and theore-

2 ACS Paragon Plus Environment

Page 3 of 38

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 1. Common structural motifs observed in complexes of anilic acids (X = F; Cl; Br; I) with heterocyclic aromatic amines: a. :B:XA:B:XA; b. :(B:XA:B):A; c. B:XA:XA:B; d. B:XA:B, where XA – stands for anilic acid and B – represents the organic base. Two possible arrangements of proton in each structure is given. tical perspectives as utilizing different types of possible structural arrangements. To this end, we focus on model anilate complexes with picolines (methyl pyridines or MP). Picoline chloranilates have been characterized in part by Adam et al.,14 while a structural analysis of the bromanilate analogues has been delivered by Thomas et al.15 From inspection of the crystallographic data retrieved from the Cambridge Structural Database, one can find that several types of supramolecular synthons are formed by XA as summarized in Fig. 1.16 Type a., :B:XA:B:XA, was found to form in 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

Page 4 of 38

the complexes composed of imidazole,17 pyrazine18 and dipyridyl-type molecules.19-21Amine and XA moieties are alternately arranged via O-H···N or N-H···O hydrogen bonds to form molecular chains or tapes. This type is specific to the phenazine : ClA (1:1) complex, introduced as the prime example of an organic ferroelectric.22-23 Recently, we have also reported such an arrangement in the case of the BrA (1:1) complex.12 Type b., :(B:XA:B):XA, has been observed in the case of 3-MP : ClA (studied here) and has earlier beenreported by Adam et al.14 Form c., B:XA:XA:B, can commonly be found in the compounds composed of pyridine-type molecules, such as lutidines16 or picolines as well as in complexes of primary and secondary amines. We have also reported such a structure for a sterically-hindered tetramethylpyrazine bromanilate.13 In this type, XA molecules form a centrosymmetric dimer through the formation of O-H···O H-bonds. Finally, type d., B:XA:B, is specific to complexes crystallizing in a 1:2 ratio, where the structural arrangement is much less diverse. By referring to the representative structural motifs, this study supplements the literature findings and seeks to provide a sound description of the structures and spectral responses in a series of MP complexes with BrA and ClA. Retaining consistency with modern solid-state oriented quantum chemical calculations, we provide a reliable analysis of structural features and spectral signatures which may be helpful in further design, screening and recognition of supramolecular architectures and synthons formed. To this end, we extend our analysis to intermolecular interactions, which are then characterized by a computationally-supported combination of highresolution solid-state NMR spectroscopy with optical (IR, RS, THz-TDS) and neutron (INS) vibrational techniques. By employing ab initio MD simulations we provide an insight into the proton displacements and underpin the role of anharmonicity in describing spectral features. This procedure comprises the necessary first step in gathering combined structure-dynamics data on these lowweight supramolecular synthons. II. RESULTS AND DISCUSSION

4 ACS Paragon Plus Environment

Page 5 of 38

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

Crystallographic Structure and Forces The simple fragments of all the crystal structures studied here, containing XA (ClA or BrA) and one of the two picolines (2-MP (α) or 3-MP (β)), are presented in Fig. 2. For experimental details please see the NOTES; while the full crystallographic data and structure refinement parameters are collected in the SI (Table S1 therein). All the complexes of a 1:1 stoichiometry crystallize in the triclinic P-1 space group. Two polymorphic modifications of BrA : 3-MP in 1:2 stoichiometry were found to crystallize as monoclinic P21/c structures, differing in their internal conformation as discussed below. The crystal structures of ClA : 2-MP (1:1) (Fig. 2a.) and ClA : 3-MP (1:1) (Fig. 2c.) have earlier been reported by Adam et al.,(14) while the structure of BrA : 3-MP (1:2) (Fig. 2d.) has been solved by Thomas et al.15 (hereafter form I, Fig. 2d). We focus here on its polymorphic modification (hereafter form II, see Fig. 2e). The structure of BrA : 2-MP (1:1) (Fig. 2b.) had never been reported. We found it to be isostructural with the ClA analog (Fig. 2a.). Picolines preferentially crystallize with BrA in a 1:2 ratio,15 however, we were able to synthesize the BrA : 2-MP (1:1) structure by avoiding an excess of 2-MP added (see NOTES for more details). All the studied motifs are classified as in Fig. 1. As seen in Fig. 2., XA molecules readily donate H atoms to basic N atoms and the consequent protonation state has a significant effect on the formation of all structures. The 2-MP-anilates (1:1) form the B:XA:XA:B motifs, in which each XA molecule is singularly deprotonated (Fig. 2a,b). In these two structures, two XA molecules are linked in a cyclic dimer by a pair of resonant-assisted O3-H3···O2 H-bonds (2.686(2) and 2.675(2) Å for ClA and BrA, respectively). On the other hand, XA forms an N1-H1···O1 H-bond with protonated picoline (2.689(2) and 2.675(2) Å for ClA and BrA, respectively). The ring planes of MP are twisted over the acid planes by about 30o, which is seen in the C6-N1-C12-C11 torsion angles of 29.9(2) and 31.5(2)o for the chloranilate and bromanilate, respectively.

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

Page 6 of 38

The ClA : 3-MP (1:1) forms an unusual structure, with infinite H-bonded (O3-H3···O1 2.542(2) Å) chains expanded through alternately deprotonated and non-deprotonated ClA2- and ClA molecules (Fig. 2c). However, as is shown below, the protonation state is further affected by temperature. Each doubly-deprotonated ClA2- unit is flanked on both sides by two protonated 3-MP molecules which form characteristic bifurcated H-bonds of 2.729(3) (N1-H1···O1) and 2.812(2) Å (N1-H1···O2). These interactions cause the rings of the co-molecules to be almost co-planar with only a slight twist of about 1.8(2)o (N1-C6-C8-C9 torsion angle). In turn, the ClA and ClA2- moieties are twisted by 66.6(2)o (C9-C7-C12-C11).

Figure 2. Structural motifs formed by the studied set of picoline-anilic acid complexes, according to single-crystal X-Ray diffraction at 100K. The thermal ellipsoids are given with 80% probability. Finally, two polymorphic modifications of the BrA : 3-MP (1:2) complex (Fig. 2d,e) are of the same B:XA:B type, which is the most typical for this stoichiometry. At the molecular level, the structures differ in the mutual orientation of the doubly-deprotonated BrA2- molecules and the 3-MP

6 ACS Paragon Plus Environment

Page 7 of 38

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

attached on both sides. In the case of Form I, the alternating molecular rings are twisted to each other by 75.6(2)o (C2-N1-C9-C8). In turn, the neighboring molecules in Form II are co-planar with a twisting angle of 1.0(2)o (C2-N1-C9-C8) and are stabilized by bifurcated H-bonds (2.671(3) (N1H1···O1) and 2.921(3) Å (N1-H1···O2)). Form I shows a lower degree of this bifurcation with both H-bonds of 2.596(2) (N1-H1···O1) and 3.036(2) Å (N1-H1···O2).

Figure 3. NCI surfaces calculated for the title compounds, namely: a.) ClA : 2-MP (1:1); b.) BrA : 2MP (1:1) c.) ClA : 3-MP (1:1) and d.) BrA : 3-MP (1:2), form II. The isosurfaces are plotted in the color range of +0.0175/-0.0175, where the significantly attracting forces are colored in blue; the repulsive interactions are shown in red; weak vdW forces are marked in green and yellow, respectively, as following their repulsive nature. In order to shed more light on the intermolecular interactions present, we applied NCI24-26 analysis. NCI originates from RDG analysis at low densities. The analysis of the sign of the second density Hessian eigenvalue times the density shows the difference between non-covalent forces. Such a value allows one to characterize the interactions by means of the density strength and its curvature, where the results can be conveniently represented as a high-resolution isosurface (Fig. 3). The RDG analysis was performed using pro-molecular densities with a density cutoff of 0.25. 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

Page 8 of 38

Table 1. The most important interactions (with energy > 5 kJ/mol) for the series of studied complexes as recognized with PIXEL approach based on the MP2/aug-cc-PVTZ clevel of theory. The crystal structures taken from X-Ray at 100K, with hydrogens optimized with standard average neutron diffraction lengths.

I. II. III. IV. V. VI. I. II. III. IV. V. VI. I. II. III. IV. V. VI. VII VII IX X I. II. III

a

Energy Decomposition [kJ/mol] ECoul. EPol. EDisp. ERep. ETot. ClA : 2MP (1:1) O-H∙∙∙O 7.243 -82.3 -41.0 -20.0 119.2 -24.1 ClA∙∙∙2-MP (S) 4.468 -17.3 -7.2 -25.9 36.1 -14.3 ClA∙∙∙ClA (S) 5.389 -16.4 -9.3 -24.6 36.0 -14.3 2-MP∙∙∙2-MP (S) 4.053 -3.5 -5.7 -28.4 24.7 -12.9 2-MP∙∙∙2-MP (↺) 6.374 -4.9 -1.4 -9.1 5.3 -10.1 N-H∙∙∙O 6.070 -58.7 -27.1 -23.5 103.1 -6.1 BrA : 2MP (1:1) O-H∙∙∙O 7.225 -84.9 -38.1 -20.2 119.1 -24.1 2-MP∙∙∙2-MP (S) 4.122 -12.3 -8.2 -32.3 38.1 -14.6 2-MP∙∙∙2-MP (↺) 6.398 -4.6 -1.2 -8.6 5.1 -9.3 BrA∙∙∙2-MP (S) 4.583 -16.2 -6.5 -24.3 38.4 -8.5 BrA∙∙∙ BrA (S) 5.397 -22.9 -8.9 -24.6 48.9 -7.5 N-H∙∙∙O 6.056 -59.3 -27.3 -23.0 102.3 -7.3 ClA : 3MP (1:1) O-H∙∙∙O 6.667 -74.3 -31.4 -12.8 89.8 -28.7 N-H∙∙∙O 6.832 -44.4 -18.9 -20.1 64.0 -19.3 N-H∙∙∙O 6.590 -38.3 -20.7 -25.2 68.6 -15.5 3-MP∙∙∙3-MP (S) 5.031 -7.0 -3.0 -17.8 13.4 -14.4 ClA∙∙∙3-MP (T) 6.365 -8.3 -11.9 -24.3 31.5 -13.0 ClA∙∙∙ClA (T) 6.763 -10.9 -1.1 -0.8 0.2 -12.6 ClA∙∙∙ClA (S) 7.471 -9.8 -1.0 -2.6 4.7 -8.6 ClA∙∙∙3-MP (S) 5.208 -8.9 -7.8 -25.3 33.7 -8.3 C-H∙∙∙O 5.963 -5.5 -3.0 -8.9 9.7 -7.6 C-H∙∙∙O 5.058 -5.2 -2.8 -9.7 10.5 -7.2 BrA : 3MP (1:2) Form II BrA ∙∙∙3-MP (S) 3.837 -31.3 -14.1 -33.3 51.7 -26.9 3-MP ∙∙∙3-MP (T) 5.876 -6.9 -2.7 -9.4 6.9 -12.1 N-H∙∙∙O 6.221 -66.0 -29.2 -23.9 107.9 -11.1

Important Interactions

d (Å)

Table legend: ECoul. – Coulomb; EPol. – Polarisation; EDisp. – Dispersion; ERep. – Repulsion; ETot. – Total; S – sandwich, face-to-face stacking; ↺ - Twisted ring stacking; T(-shaped) – edge-to-face stacking.

The intermolecular forces revealed by NCI were further analyzed quantitatively by employing PIXEL analysis27 as focusing on the electron density of individual species calculated at the MP2/augcc-pVTZ level of theory.28-29 The leading interactions were identified within the spatial resolution of the PIXEL method (defined in the vector manner) and further decomposed into their Coulombic, polarization, dispersion and repulsion contributions (Table 1.).The presented analysis has been supplemented with Hirshfeld 3D surface Both Fig. 3 and Table 1. show that the ClA : 2-MP (1:1) 8 ACS Paragon Plus Environment

Page 9 of 38

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

and projection, as given in the SI (Figs. S1-S5 therein), which was, however, found to be much less informative. BrA : 2-MP (1:1) systems are perfectly isostructural and are driven by non-covalent forces of similar strengths. Most prominent forces are exerted by moderate-strength O-H···O bonds, which stabilize the synthons with contributions of about 24 kJ/mol. Contrary to Adam’s assumption,14 the N-H···O bonds are two-center rather than bifurcated. The PIXEL analysis clearly shows that the NH···O ones bonds are much weaker than O-H···O ones, with total energy contributions estimated as

ca. 6-7 kJ/mol in each case. According to Table 1, in such a closely packed ensemble large stabilizing contributions from Coulombic terms are compensated by destabilizing repulsive energy from a large charge density overlap. Interestingly, most of the stabilizing contributions come from stacking interactions rather than H-bonds. As a result, B:XA:XA:B synthons are packed into stacks, propagated toward the a-axes, mutually displaced and oriented anti-parallelly. Note that these are pairwise stackings, as they are not long-range correlated. Both homo- and hetero-molecular stackings are present in both anilates, while according to Table 1, BrA contributions are much weaker than ClA. Nevertheless, in both cases the stabilizing energy coming from N-H···O bonds is much smaller than that coming from the stacking forces. The parallel stacking interactions with 2-MP molecules are accompanied by a ~30o twist of 2MP.14 According to NCI, some trails of weak interactions between adjacent methyl groups, separated by 2.774 (ClA) and 2.900 Å (BrA), can also be noticed. These methyl groups are visibly separated from halogen atoms. The mutual X···X interactions30-31 may also be of importance due to a relatively close separation of halogen atoms, which are located at distances of 3.3240(7) (Cl1···Cl1), 3.4766(4) Å (Br1···Br1) and 3.5949(4) (Br1···Br2) from each other. The PIXEL analysis, however, does not indicate their contributions to be greater than 5kJ/mol. The ClA : 3-MP (1:1) structure is quite unique because of the presence of some intriguing Hbond features. In this system, the anti-parallelly oriented ClA and ClA2- molecules form infinite

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

Page 10 of 38

chains. In addition, a bifurcated pair of H-bonds is formed at each site of the ClA2- sub- units, involving all the carbonyl groups. However, half of these carbonyls are shared to form the infinite ClA chains. Therefore, two protons are in close contact with three oxygen acceptors, making the structure resonant. According to PIXEL, these resonant H-bonds dominate the interactions, where the bifurcated N-H···O bond is two times stronger than in all the remaining systems. In addition to this unique proton arrangement, several other features were found . According to PIXEL analysis, the total stabilizing contributions from a group of homo- and hetero-molecular stackings, where the molecules interact both in a face-to-face (S) and in an edge-to-face (T) manner, are nearly the same (~55kJ/mol in sum). Furthermore, there is a unique stabilization of 3-MP, not only due to Hbondings and stackings, but also due to short CH···O contacts. It is quite exceptional that each C-H group is involved in such contacts. The remaining methyl groups are considerably in contact with chlorine atoms (~2.9Å) and significantly influence the energetics of the system (Table 1). Finally, the NCI analysis clearly reveals dihydrogen close-contacts, H···H, at a distance of ~2.1Å. Nevertheless, the PIXEL analysis predicts their strength is only ca. 2.5 kJ/mol. Next we focus on the BrA : 3-MP (1:2) complex, with the shortest centrosymmetric synthons of a B:XA:B type. Because of the absence of OH···O bonds and higher synthon mobility, the pre sence of alternative polymorphs is not surprising. Here we focus on the Form II, which was using complementary solid-state spectroscopy. Nevertheless, the difference between these two structures deserves a comment. Both forms crystallize in the same P21/c space group, driven by a relatively small variety of non-covalent forces. In both structures, BrA molecules are doubly deprotonated, but 3-MP can follow two alternative orientations: one nearly perpendicularly twisted (Form I) and the other co-planar w.r.t. the BrA plane (form II). The co-planar orientation in the latter is stabilized by the formation of bifurcated NH···O bonds, with an additional contribution from the CH···O interactions. The main difference in the crystal packing of both polymorphs is due to the formation of homo- or hetero-molecular ensembles. In the case of Form I, we found only homo-molecular, anti-

10 ACS Paragon Plus Environment

Page 11 of 38

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

parallel sandwich stackings of picoline pairs and T-shaped interactions between both bromine atoms in each BrA molecule and BrA rings from its neighbors. By contrast, there are no such contacts in Form II, where very specific forces are present. In Form II, BrA is bound in-plane to the 3-MP moieties, forming three-center bifurcated H-bonds. Most interestingly, the BrA molecule is also coordinated out-of-plane at each site, leading to the formation of a heteromolecular stacking complex of a 1:2 stoichiometry. According to Table 1, the BrA ···3-MP (S) interaction has an enormous stabilizing strength of -27kJ/mol, with a close centroid distance of 3.837 Å, constraining the planar geometry of the synthons. According to PIXEL, such interactions are much greater than that of bifurcated N-H···O bond, which stabilize the system by ~11 kJ/mol. Such a finding should be, however, treated with caution. The resulting charge-transfer in solid-state may lead to a significant flow of electron density, making PIXEL estimation questionable as based on the molecule in vacuo reference Nevertheless, PIXEL analysis clearly indicates that parallel, heteromolecular pairwise stackings are here of great importance. Proton Migration Since temperature-affected proton positions cannot be precisely elucidated with standard single-crystal X-Ray diffraction, we employed AIMD simulations at several temperatures. The most important interatomic distances were derived by assuming they follow a Gaussian-like distribution for a given ion. These results are provided in the SI (Table S2 therein). The limited proton refinement with X-Ray leads to a distance discrepancy of up to 0.2 Å in each case. For H-bonding, the simulations clearly show that in each compound, a proton is localized at the nitrogen atom and no proton transfer within N···O was found across the studied temperature range. The same holds true for the OH···O bonds except for those of the ClA : 3-MP (1:1) crystal. The OH···O dynamics was analyzed here with the use of RDFs, calculated for constituted elements using the following relation:

[

g i , j ( r ) = ni , j (r ) / ρ j 4π 2 dr

]

(1)

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

where gi, j (r ) is the radial distribution function of species i and j,

ni, j

Page 12 of 38

is the corresponding histogram, r is

the distance between two species, dr is the histogram shell width and

ρj

is the average density of

species j. In the current calculations, dr = 1×10-3 Å. Fig. 4 shows the RDFs for O···O, H···H and O···H distances in ClA : 3-MP (1:1), calculated at 75, 225 and 375K, respectively. The RDFs for the ClA species are further divided into their interand intra-molecular terms. To facilitate discussion, the supercell employed in the simulations is visualized along with its analyzed close contacts, including the most interesting CH···O ones. These are spread over the range ~2.3-2.5 (75K) and ~2.4-2.6 Å (375K), evincing well-defined features in the RDF projection. By analyzing Fig. 4, one can clearly note that the :(B:XA:B):A structure undergoes a temperature-induced disorder, manifesting itself in a broadening of the curves.

Figure 4. Radial distribution functions (RDF), calculated the ClA : 3-MP (1:1) structure with NVT BOMD at 75, 225 and 375K, respectively(left; see the NOTES for more details) The supercell {3×2×1} employed in the calculations is given (right) with the analyzed interatomic distances colored in line with the RDF sketch. In Table S2 one can note that the sharp peak found at 1.045 Å (r(OH) at the lowest temperature) is due to hydroxyl groups (the black line in Fig. 4). The second peak at ~2.3 Å defines intramolecular O···H contacts. However, even at the lowest temperature, a small fraction of transferred protons can be found. It corresponds to the peaks at around 1.5 and 1.0 in the intra- and intermolecular O···H curves, respectively.

12 ACS Paragon Plus Environment

Page 13 of 38

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

Proton mobility is evident from higher-temperature simulations, which result in a visible increase of proton-transfer peaks and a broadening of curves. Mobility is also evident from the broadening of the intramolecular feature at ~2.3 Å, which further maps the proton delocalization. By comparing the intermolecular peaks for r(OH···O) and r(NH···O), one can find their overlap at higher temperatures, which suggests that these H-bonds are coupled, and possibly, resonant-assisted. One can also note a similar broadening of the intermolecular O···O distances, which shows that proton migration results in a redistribution of C=O lengths. However, as the intensity of the intramolecular O···H feature is higher, it shows that the proton is still more localized on one of the donors. Such a proton disorder indicates that related potential energy surface is very shallow. This is important because it breaks the symmetry and can be the source of potential polar properties (to be further verified). While beyond the scope of this work, one should note that a similar mechanism permitted to discovery of organic ferroelectrics, with a prime example of the ClA complex with phenazine.22-23 These results motivative an examination of the influences of crystal packing and non-covalent interactions on the spectral response of each system.

Solid-State Spectroscopy 13

C CP/MAS NMR The solid-state

13

C CP/MAS NMR spectra are shown in Fig. 5 along with the results of

theoretical calculations (the NOTES for details), yielding unambiguous evidence of a crystal structure in each powder sample. For full spectral assignment we refer the reader to the SI (Table S3 therein). The calculations were based on the aforementioned crystallographic models, including both polymorphs of BrA : 3-MP (1:2). In addition, the hypothetical structure of ClA : 3-MP (1:1) was used, where the proton was transferred onto the ClA2- moieties (Fig. 1 b). The break of the centrosymmetry is accompanied here by a unit cell transformation from Z = 2 to Z = 4 (P1). The NMR study clearly confirms that both 2-MP complexes are highly isostructural and also proves that the BrA : 3-MP (1:2) powder consists of Form II rather than Form I. Finally, A comparison of

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

Page 14 of 38

experimental and calculated NMR spectra for ClA : 3-MP (1:1) favors a centrosymmetric structure in the non-hydrogen skeleton, despite the expectedness of the proton-transfer events. The upper spectral range can be used to identify the various synthons formed. All signals above 155 ppm are due to carbon-chalcogen fragments. The occurence of resonance at 177 ppm is due to a proton-accepting carbonyl group, which is only slightly perturbed by CH···O contacts. This is an upper deshielding limit, which has been previously reported for other complexes of BrA.12-13 A downward shift from this value can be treated as an H-bond marker. Accordingly, in the Form II of the BrA : 3-MP complex there is no free carbonyl group and each non-equivalent one is engaged in the bifurcated NH···O bond. Hence, a doublet is observed at 175 and 171 ppm, respectively. In the case of the 2-MP complexes, two sets of signals were found at around 170 ppm. Finally, in ClA : 3MP (1:1), one can observe a set of coupled H-bonds.

Figure 5. Room temperature experimental 13C CP/MAS NMR spectra for the series of studied complexes, presented along with chemical shifts calculated based on the crystallographic and hypothetical structures (sticks; see the NOTES for more details). The raw experimental data are presented as green lines, while the Voigt profile fits are given as black curves.

14 ACS Paragon Plus Environment

Page 15 of 38

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

The GIPAW calculations favors a centrosymmetric structure, but improperly suggest the presence of a degenerated resonance at 175 ppm (see Fig. 5). Calculations performed for the P1 structure properly predict the splitting of the C=O resonances, but fail to describe the remaining signals. The loss of this degeneracy can be linked to the temperature , which slightly moves the proton toward the middle of the O···O bridge. As a result the O(2)H···N and O(1)H···O(3) contacts strengthen and the related C=O signals shift downward. Perfect agreement with all the remaining signals suggests that such a proton delocalization is a local feature, one without a great effect on the molecular architecture. It is also probable that the resulting configuration is short-lived, and the structural response averaged in the time-window of NMR experiment supports the P-1 symmetry. In turn, the OH-linked carbons are found to be more shielded. In the singularly deprotonated anilate complexes there is a signal at around 158 ppm, assigned to the OH-linked carbon. Additional XA signals are found at around 100 ppm; they are, however, completely obscured by residual dipolar coupling between the carbon atoms and quadrupolar halogen nuclides, where the relativistic effects exceed the limitations of the applied computational methodology.12-13 To better understand the influence of the crystal environment on α- and β-picoline, one can refer to their native NMR spectra reported by Pajzderski et al.32 By inspection of the

13

C CP/MAS

NMR data, one can note a tremendous influence of the complexation on magnetic response, where the change in isotropic shielding can be expressed as ∆δ13C = δ13CComplex ̶ δ13CNeutral (ppm). The picoline complexes give a virtually identical response as in the neat form. Upon reception of the proton, the nitrogen-linked, C(2) and C(6) resonances shifts downward by ∆δ13C = ̶ 4 and ̶ 8 ppm, manifesting themselves at 154 and 141 ppm, respectively. While C(6) is in close contact with the oxygen’s electron lone pair, it becomes further shielded upon formation of the synthon. In the case of

β-picoline, the C(2) and C(6) are much more similar. For the ClA : 3-MP complex, the bifurcated NH···O bond results in the resonances found at 142 and 138 ppm (∆δ13C C(2) = ̶ 8 ppm and ∆δ13C C(6) = ̶ 9 ppm). These signals are even more shifted in Form II of the BrA : 3-MP complex, where 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

Page 16 of 38

∆δ13C = ̶ 11 ppm for both nuclei and the resulting resonances are found at 140 and 136 ppm, respectively. Such increased shielding can be explained by the aforementioned formation of heteromolecular stacks. The set of C(3), C(4) and C(5) nuclei are deshielded to the same extent in each compound with ∆δ13C of +6 (C(3)); +11 (C(4)) and + 5 ppm (C(5)) w.r.t. the

13

C NMR signals of

neat α- and β-picolines reported at 123; 136; 121 and 133; 136; 123 ppm, respectively. The highest NMR frequency is exhibited by the CH3 resonances, which for neat picolines are observed at 24 (α) and 18 ppm (β).32 For the 2-MP complexes there is an influx of electron density, resulting in a shielding effect, visibly moving the signals down to 18 ppm. Such a prominent effect can be attributed to the anti-parallel homo-molecular stackings of MP in the crystal lattice. In line with the intermolecular interaction analysis, the methyl groups in BrA : 3-MP (1:2) are barely affected by the crystal field, manifesting themselves at 17 ppm, while a slight upward shift was noted in the Cla : 3-MP complex, attributed to weak CH3 ··· Cl contacts.

Middle-Wavenumber Vibrational Spectroscopy In this study we mainly refer to the room-temperature spectra, explored with the help of IR, RS (>110 cm-1) and THz-TDS ( 98.0%) were used as purchased. The compounds were dissolved in methanol and magnetically stirred at 30 °C. The single-crystals were grown by slow evaporation of the mixture in ambient conditions, taking up to two weeks in each case. All the powder samples were synthesized in a multi-gram amount and their thermal stability and polymorphism were tested with TG and DSC (see SI). The complexes in 1:1 ratios, namely: BrA : 2MP; ClA : 2-MP and ClA : 3-MP, were synthesized by dissolving equimolar amounts of the substrates. Despite several attempts, we were unable to synthesize pure BrA : 3-MP in a 1:1 stoichiometry. We followed the recipe given by O’Neill49 and used an equimolar ratio of both substrates. However, the resulting single-crystals were of a 1:2 stoichiometry. The related crystal structure (denoted here as Form II) was slightly different from that reported by Thomas et al.15 (Form I). We repeated the synthesis using a threefold excess of volatile 3-MP as trying to synthesize the Form I.15 We avoided the much larger, twentyfold excess of β-picoline suggested in Thomas’ recipe, as we intended to obtain a high-quality polycrystalline material. Single-crystal diffraction confirmed the presence of Form I, however, TG and solid-state spectroscopy revealed contamination with 3-MP

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

Page 32 of 38

occluded in the powder. It should be noted that Thomas et al.15 have only studied single crystals with X-Ray, without reference to any other technique.

Single-Crystal X-Ray Diffraction The measurements were done at 100K (Agilent SuperNova diffractometer; EOS CCD detector and a mirror-monochromated CuKα radiation source λ = 1.54184 Å) The structure was solved by a direct methods and refined by full-matrix least squares method on F2 data using SHELXTL programs.50 All non-hydrogen atoms were refined using anisotropic displacement parameters. Hydrogens bonded to carbons were inserted in calculated positions with C-H = 0.98 (methyl) or 0.95 Å (aromatic) and refined isotropically as riding model with Uiso(H) equal to 1.5Ueq(C) or 1.2Ueq(C) for methyl and aromatic H atoms, respectively. The H atoms of the OH and NH pairs were located in a difference map and their positions were freely refined with Uiso(H) values established as free or set to 1.5 and 1.2 of those of O and N, respectively. The five crystal structures presented in this paper were deposited at the Cambridge Crystallographic Data Center (CCDC 1456515-1456519).

Solid-State Spectroscopy The

13

C CP/MAS NMR spectra were acquired at room temperature (Bruker Avance III

spectrometer 16.5 T). The 3 mm diameter zirconia rotor was spun at a frequency of 18 kHz. About 600 scans were accumulated in each case, with a cross-polarization contact time of 2.5 µs and a recycle delay of 6 s. The 13C chemical shifts were referenced with respect to tetramethylsilane. Room-temperature optical vibrational spectroscopy was employed with a spectral resolution of -1

2 cm . FT-MIR spectra were acquired using a KBr pellet technique (Bruker Equinox 55 FT-IR spectrometer). The FT-FIR measurements were carried out by suspending the samples in Apiezon N grease on a HDPE substrate (Bruker Vertex v70). FT-RS measurements were performed over the effective range of 4000–100 cm-1 (Bruker FRA 106/S; Nd:YAG : 1064 nm). THz-TDS measurements with Teraview TPS 3000 were done for mixtures containing 10% of a sample and 90% of HDPE to obtain 400 mg pellets. In the case of the ClA : 3-MP (1:1) complex, its content was reduced to 5% due to a large dielectric response. 1800 scans were accumulated with a pure HDPE sample used as a reference. INS measurements were performed with the inverted-geometry spectrometer NERA51 set at the high flux pulsed nuclear reactor IBR-2 at JINR Dubna, Russia. The incident neutron energies were determined by measuring the neutron time-of-flight across the 110 m distance from the water moderator to the sample. The INS spectra were recorded at the final energy of the scattered neutrons of Ef = 4.65 meV, fixed by crystal analyzers and beryllium filters. In each case the ~10g sample closed in a flat aluminium container was measured for 20 hours at 10K.

32 ACS Paragon Plus Environment

Page 33 of 38

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

Computational Details Theoretical calculations in periodic boundary conditions were carried out with fixed-cell approach (cell constants from 100K). We used the previously optimized numerical methodology, presented in more detail in Refs. 12-13. In brief, PW-DFT calculations were employed using the CASTEP code.52-53 A PBE functional54 was used throughout. Core electrons were represented by designed NCPPs55 and the valence states were expanded using a PW basis set with a kinetic-energy cutoff of 1050 eV. The k-point grid was adjusted to maintain a constant spacing in a reciprocal space of 0.07 Å−1, and the SCF energy converged with a tolerance of 10−12 eV/atom. The convergence criteria of the variation in maximum force and displacement were defined as as: 1×10−5 eV/Å, 1×10−6 Å, respectively. The HLD calculations were performed using a DFPT approach.56-62 The IR spectra were calculated analytically, using the computed permittivity and Born charge tensors. The nonanalytic contribution, generating the long-range dipole coupling was added in multiple directions w.r.t. the Γ−point. The Raman activity tensors were calculated numerically using hybrid finite displacement/DFPT method in the presence of an external field63 and transformed into Raman intensities.64 The INS spectra were modeled with the help of the aClimax program, using HLD results as the input.65 BOMD simulations were performed at finite classical temperatures. The supercells were constructed to allow for a description of the electronic structure of each system with a single k-point. The same numerical conditions have been used except for the SCF convergence, which was reduced down to 5×10−8 eV/atom. An NVT ensemble with a Nose–Hoover thermostat was used, with a MD time step of 0.5 fs. The production runs of 8 ps were collected after 1 ps equilibration of each system. Using the 300 × 2.5GHz processors of the high-performance supercomputer PROMETHEUS, these calculations took about two weeks per ps. The

13

C NMR calculations were performed with the GIPAW method.66 The all-electron

information was reconstructed for the NCPP equilibrium geometries, using on-the-fly generated scalar relativistic USPPs.67

ASSOCIATED CONTENT Supporting Information Available crystallographic data and structure refinement parameters; interatomic distances analysis from AIMD; Hirshfeld 3D surface analysis; experimental and theoretical spectra with detailed labeling; tables with the IR, RS, INS and THz-TDS data along with detailed band assignment; TG and DSC curves. This material is available free of charge via the Internet at http://pubs.acs.org

ABBREVIATIONS 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

Page 34 of 38

CP/MAS – cross-polarization / magic angle spinning; NMR –nuclear magnetic resonance; FTFourier transform; IR – infrared; RS – Raman spectroscopy; THz- terahertz; TDS – time-domain spectroscopy; INS – inelastic neutron scattering; TG – thermogravimetry; DSC – differential scanning calorimetry; MD – molecular dynamics; NCI – non-covalent interactions analysis; RDG – reduced density gradient; RDF – radial distribution function; DF(P)T– density functional (perturbation) theory; HLD – harmonic lattice dynamics; PES – potential energy surface; TO transverse-optical; LO – longitudinal optical; MD – molecular dynamics; BOMD - Born-

(ω) Oppenheimer MD; AIMD – ab initio MD; VDOS – vibrational density of states; G(ω (ω – generalized VDOS; VACF - velocity autocorrelation function; 2D-IR – two-dimensional ultrafast IR; ZPVE – zero-point vibrational energy; HDPE - high-density polyethylene; SCF - self-consistent field; PW – plane-wave; GIPAW – gauge-included projected augmented wave; NCPP / USPP – normconserving / ultrasoft pseudopotential.

ACKNOWLEDGEMENTS This research was supported in part by PL-Grid Infrastructure (Grant IDs: latticedynamics). K. Łuczyńska and K. Drużbicki gratefully acknowledge the financial support of the Polish Government Plenipotentiary for JINR in Dubna (Grant No. № 46/9/22.01.2016) as well as the OMUS scholarship for outstanding young scientists at FLNP, JINR (Grant IDs: 16-402-06; 16-401-01). Research support from Dr Norbert Pałka (Military University of Technology, Warsaw); Dr Jerzy Antonowicz (Warsaw University of Technology) Ms. Zofia Huppenthal (Nicolaus Copernicus University, Torun) and Ms. Halina Thiel-Pawlicka (Adam Mickiewicz University, Poznan) is gratefully acknowledged.

REFERENCES (1) Horiuchi, S.; Tokura, Y. Nature Materials, 2008, 7, 357−366. (2) Horiuchi, S.; Kumai, R.; Tokura, Y. Angew. Chem., 2007, 46, 3497−501. (3) Horiuchi, S.; Noda, Y.; Hasegawa, T.; Kagawa, F.; Ishibashi, S. Chem. Mater., 2015, 27, 6193−6197. (4) Kobayashi, K.; Horiuchi, S.; Kumai, R.; Kagawa, F.; Murakami, Y.; Tokura, Y. Phys. Rev. Lett.,

2012, 108, 237601−237606. (5) Wikfeldt, K.T.; Michaelides, A. J. Chem. Phys., 2014, 140, 041103−041108. (6) Mukhopadhyay, S.; Gutmann, M.; Fernandez-Alonso, F. Phys. Chem. Chem. Phys., 2014, 16, 26234−26239. (7) Horiuchi, S.; Kobayashi, K.; Kumai, R.; Minami, N.; Kagawa, F.; Tokura, Y. Nat Commun.

2015, 16; 7469−7477. (8) Molčanova, K.; Kojić-Prodić, B. CrystEngComm, 2010, 12, 925−939. 34 ACS Paragon Plus Environment

Page 35 of 38

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) Fernandez-Alonso, F.; Gutmann, M.J.; Mukhopadhyay, S.; Jochym, D.B.; Refson, K.; Jura, M.; Krzystyniak, M.; Jimenez-Ruiz, M.; Wagner, A. J. Phys. Soc. Jpn., 2013, 82, SA001−SA10. (10)

Mukhopadhyay, S.; Gutmann, M.J.; Jura, M.; Jochym, D.B.; Jimenez-Ruiz, M.; Sturniolo,

S.; Refson, K.; Fernandez-Alonso, F. Chem. Phys., 2013, 427, 95−100. (11)

Drużbicki, K.; Mikuli, E.; Pałka, N.; Zalewski, S.; Ossowska-Chrusciel, M. D. J. Phys.

Chem. B, 2015, 119, 1681−1695. (12)

Łuczyńska, K.; Drużbicki, K.; Lyczko, K.; Dobrowolski, J.Cz. J. Phys. Chem. B, 2015, 119,

6852–6872. (13)

Łuczyńska, K.; Drużbicki, K.; Lyczko, K.; Starosta, W. Vib. Spectrosc., 2014, 75, 26−38.

(14)

Adam, M.S.; Parkin, A.; Thomas, L.H.; Wilson, C.C. CrystEngComm, 2010, 12, 917–924.

(15)

Thomas, L.H.; Adam, M.S.; O'Neill, A.; Wilson, C.C. Acta Cryst., 2013, C69, 1279–1288.

(16)

Ishida, H.; Kashino, S. Z. Naturforsch. 2002, 57a, 829–836.

(17)

Ishida, H.; Kashino, S. Acta Cryst., 2001, C57, 476–479.

(18)

Ishida, H.; Kashino, S. Acta Cryst., 1999, C55, 1923–1926.

(19)

Zaman, Md.B.; Tomura, M.; Yamashita, Y. Chem. Commun. 1999, 999–1000.

(20)

Zaman, Md.B.; Tomura, M.; Yamashita, Y. Org. Lett. 2000, 2, 273–275.

(21)

Akhtaruzzaman, Md.; Tomura, M.; Yamashita, Y. Acta Cryst. 2001, E57, o353–o355.

(22)

Gotoh, K.; Asaji, T.; Ishida, H. Acta Cryst., 2007, C63, o17–o20.

(23)

Horiuchi, S.; Kumaia, R.; Tokura, Y. J. Mater. Chem., 2009, 19, 4421–4434.

(24)

Hanson, R.M. J. Appl. Crystallogr., 2010, 43, 1250−1260.

(25)

Johnson, E.R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A. J.; Yang, W. J.

Am. Chem. Soc., 2010, 132, 6498−6506. (26)

Contreras-García, J.; Johnson, E.R.; Keinan, S.; Chaudret, R.; Piquemal, J.P.; Beratan, D.N.,

Yang, W. J. Appl. Cryst., 2010, 43, 1250−1260. (27)

Gavezzotti, A. J. Phys. Chem. B 2003, 107, 2344−2353.

(28)

Møller, C; Plesset, M. S. Phys. Rev., 1934, 46, 0618-0622.

(29)

Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheesman, J. R.;

Montgomery, J. A.; Vreven, T.; Kudin, K. N.; Burant, J. C.; Millam, J. M.; Iyengar, S. S.; Tomasi, J.; Barone, V.; et al. GAUSSIAN03; Gaussian Inc.: Wallingford, CT, 2004. (30)

Capdevila-Cortada, M.; Castellóa, J.; Novoa, J.J. CrystEngComm, 2014, 16, 8232–8242.

(31)

Vener, M.V.; Shishkina, A.V.; Rykounov, A.A.; Tsirelson, V.G. J. Phys. Chem. A 2013. 117,

8459–8467. (32)

Pajzderski, L.; Tousek, J.; Sitkowski, J.; Malinakova, K.; Kozerski, L.; Szłyk, E. Magn.

Reson. Chem. 2009, 47, 228–238. 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

Page 36 of 38

(33)

Bowman J. M. J. Chem. Phys. 1978, 68, 608–610.

(34)

Monserrat, B.; Drummond, N.D.; Needs, R.J. Phys. Rev. B. 2013, 87, 144302–144312.

(35)

Thomas, M.; Brehm, M.; Fligg, R.; Vöhringer, P.; Kirchner, B. Phys. Chem. Chem. Phys.

2013, 15, 6608–6622. (36)

Williams,, R.W., Schlücker, S., Hudson, B.S. Chem. Phys. 2008, 343, 1–18.

(37)

Denisov, G.S.; Mavri, J.; Sobczyk, L. Potential Energy Shape for the Proton Motion in

Hydrogen Bonds Reflected in Infrared and NMR Spectra. in Hydrogen Bonding—New Insights. S. J. Grabowski ed., Springer 2006, 377–416. (38)

Witkowski, A. J. Chem. Phys. 1967, 47, 3645–3648.

(39)

Marechal, Y., Witkowski, A. Infrared Spectra of H-Bonded Systems. J. Chem. Phys. 1968,

48, 3697–3705. (40)

Witkowski, A., Wójcik, M. Infrared Spectra of Hydrogen Bond a General Theoretical Model.

Chem. Phys. 1973, 1, 9–16. (41)

Wójcik, M. J. Theory of the Infrared Spectra of the Hydrogen Bond in Molecular Crystals.

Int. J. Quant. Chem. 1976, 10, 747–760. (42)

Cao, J.; Voth, G. A. J. Chem. Phys. 1994, 100, 5106-5117.

(43)

Jang, S.; Voth, G. A. J. Chem. Phys. 1999, 111, 2371-2384.

(44)

Rossi, M.; Liu, H.; Paesani, F.; Bowman, J.; Ceriotti, M. J. Chem. Phys. 2014, 141,

181101−181106. (45)

Rossi, M.; Ceriotti, M.; Manolopaulos, D.E. J. Chem. Phys. 2014, 140, 234116−181128.

(46)

Petersen, P.B.; Roberts, Krupa Ramasesha, S.T.; Nocera, D.G.; Tokmakoff, A. J. Phys.

Chem. B Lett., 2008, 112, 13167–13171. (47)

Stingel, A.M.; Calabrese, C.; Petersen, P.B. J. Phys. Chem. B, 2013, 117, 15714−15719.

(48)

Van Hoozen, B.L.; Petersen, P.B. J. Chem. Phys. 2015, 142, 104308−104316.

(49)

Pröpper, D.; Yaresko, A.N.; Larkin, T.I.; Stanislavchuk, T.N.; Sirenko, A.A.; Takayama, T.;

Matsumoto, A.; Takagi, H.; Keimer, B.; Boris, A.V. Phys. Rev. Lett., 2014, 112, 087401−087405. (50)

O'Neill, A. PhD Thesis, 2010, University of Glasgow, http://theses.gla.ac.uk/id/eprint/1709

(51)

Sheldrick, G.M. Acta Cryst., 2008, A64, 112–122.

(52)

Natkaniec, I.; Chudoba, D.; Hetmanczyk, L.; Kazimirov, V.Y.; Krawczyk, J.I.; Sashin, I.;

Zalewski, S. J. Phys.: Conf. Series, 2014, 554, 01–12 (53)

Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne,

M.C. J. Phys.: Condens. Matter, 2002, 14, 2717−2744. (54)

Clark, S.J.; Segall, M.D.; Pickard, C.J.; Hasnip, P.J.; Probert, M.J.; Refson, K.; Payne, M.C.

Z. Kristallogr., 2005, 220, 567−570. 36 ACS Paragon Plus Environment

Page 37 of 38

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

(55)

Perdew, J.P.; Burke, K.; Wang, Y. Phys. Rev. B, 1996, 54, 16533−16539.

(56)

Rappe, A.M.; Rabe, K.M.; Kaxiras, E.; Joannopoulos, J.D. Phys. Rev. B, 1990, 41,

1227−1230. (57)

Refson, K.; Clark, S.J.; Tulip, P.R. Phys. Rev. B, 2006, 73, 155114−155136.

(58)

Baroni, S.; Giannozzi, P.; Testa, A. Phys. Rev. Lett., 1987, 58, 1861−1864.

(59)

Gonze, X.; Allan, D. C.; Teter, M.P. Phys. Rev. Lett., 1994, 68, 3603−3606.

(60)

Gonze, X. Phys. Rev. A, 1995, 52, 1086−1095.

(61)

Gonze, X. Phys. Rev. A, 1995, 52, 1096−1114.

(62)

Gonze, X. Phys. Rev. B, 1997, 55, 10337−10354.

(63)

Baroni, S.; Dal Corso, A.; de Gironcoli, S.; Giannozzi, P. Rev. Mod. Phys., 2001, 73,

515−561. (64)

Milman, V.; Perlov, A.; Refson, K.; Clark, S. J.; Gavartin, J.; Winkler, B. J. Phys.: Condens.

Matter., 2009, 21, 485404−485416. (65)

Polavarapu, P.L. J. Phys. Chem., 1990, 94, 8106−8112.

(66)

Ramirez-Cuesta, A. J. Comp. Phys. Commun., 2004, 157, 226−238.

(67)

Pickard, C.; Mauri, F. Phys. Rev. B, 2001, 63, 245101–245115.

(68)

Yates. J.R.; Pickard, C.; Mauri, F. Phys. Rev. B, 2007, 76, 024401–024412.

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

Page 38 of 38

For Table of Contents Use Only Structure-Spectra Correlations in Anilate Complexes with Picolines Katarzyna Łuczyńska, Kacper Drużbicki, Krzysztof Lyczko, Jan Cz. Dobrowolski

Synopsis: A sound description of structure and spectral response in a series of strongly hydrogenbonded complexes of anilic acids with picolines is given. To this end, X-Ray crystallography was extended toward computationally-supported solid-state spectroscopy, providing a comprehensive analysis of the spectral signatures to facilitate design and recognition of supramolecular architectures formed by this kind of components.

38 ACS Paragon Plus Environment