Structural and Energetic Analysis of Molecular Assemblies in a Series

Aug 1, 2017 - PDF. cg7b00868_si_001.pdf (2.99 MB). Citing Articles; Related Content. Citation data is made available by participants in Crossref's Cit...
0 downloads 10 Views 2MB Size
Subscriber access provided by Warwick University Library

Article

Structural and energetic analysis of molecular assemblies in the series of nicotinamide and pyrazinamide cocrystals with dihydroxybenzoic acids Katarzyna N. Jarzembska, Anna A. Hoser, SUNIL VARUGHESE, PhD, Rados#aw Kami#ski, Maura Malinska, Marcin Stachowicz, Venkateswara R. Pedireddi, and Krzysztof Wozniak Cryst. Growth Des., Just Accepted Manuscript • DOI: 10.1021/acs.cgd.7b00868 • Publication Date (Web): 01 Aug 2017 Downloaded from http://pubs.acs.org on August 4, 2017

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 37

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 energetic analysis of molecular assemblies in the series of nicotinamide and pyrazinamide cocrystals with dihydroxybenzoic acids# Katarzyna N. Jarzembska,a*¶ Anna A. Hoser,b¶ Sunil Varughese,c,d Radosław Kamiński,a Maura Malinska,b Marcin Stachowicz,b Venkatesvara R. Pedireddi,e Krzysztof Woźniak,b*

a

Department of Chemistry, University of Warsaw, Żwirki i Wigury 101, 02-089 Warsaw, Poland

b

Biological and Chemical Research Centre, Department of Chemistry, University of Warsaw, Żwirki i Wigury 101, 02-089 Warsaw, Poland

c

Chemical Science and Technology Division, CSIR-National Institute for Interdisciplinary Science and Technology, Trivandrum 695 019, Kerala, India

d

Academy of Scientific and Innovative Research, Delhi-Mathura Road, New Delhi 110 025, India

e

Indian Institute of Technology, Bhubaneswar 751 013, India

* Corresponding authors: Katarzyna N. Jarzembska ([email protected]) Krzysztof Woźniak ([email protected])

#

Honouring Professor William Jones for his outstanding contributions to organic solid-state chemistry



Both authors contributed equally

Abstract: Four new cocrystals of pharmaceutically active N-donor compounds, pyrazinamide (P) and nicotinamide (N), with a series of dihydroxybenzoic acids, i.e. 2,3dihydroxybenzoic acid (23DHB), 2,4-dihydroxybenzoic acid (24DHB), and 2,6dihydroxybenzoic acid (26DHB), were synthesised and structurally evaluated in order to study basic recognition patterns and crystal lattice energetic features. The literaturereported structures of this kind, i.e., N:24DHB, N:25DHB and N:26DHB (the last two were crystallised and remeasured by us at 100 K) and P:25DHB, completed the series. The analysis of interaction networks in the examined cocrystals reflects the relative affinity of the COOH and OH groups towards N-donor compounds. A major factor that governs the primary synthon formation is the basic character of the proton acceptors in the heterocyclic compounds. In a crystal lattice, the more rigid pyrazinamide tends to form its primary structural motifs, and hence is less influenced by the molecular

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

surrounding than nicotinamide. Consequently, crystal lattice stabilisation energy values for the cocrystals of nicotinamide are more advantageous, whereas the patterns created by pyrazinamide are more predictable. Nicotinamide cocrystals are also characterised by crystal lattices being more energetically uniform in all directions than the pyrazinamide equivalents. Importantly, cocrystal cohesive energies are more favourable than that of the respective single component crystal structures, which supports the cocrystal formation when both coformers are dissolved and mixed together. Although classical hydrogen bonds are majorly responsible for synthon formation, weak dispersive forces cannot be neglected either as far as the structure stabilisation is concerned.

2 ACS Paragon Plus Environment

Page 2 of 37

Page 3 of 37

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

1. Introduction Crystal engineering deals with the design and synthesis of solid state assemblies with desired chemical and physical properties.1-7 It has been recently shown that physicochemical properties of a given compound in the solid state can be tuned by its cocrystalization with an appropriately selected coformer. In order to find the most suitable coformers, it is indispensable to understand how a given molecule may interact with the compound of interest and what kind of 3-dimensional (3D) network is expected. Therefore, it is of high importance to study the nature and strength of intermolecular interactions and their directing role in the self-assembly processes.8-13 Considering the organic assemblies, hydrogen bonds constitute a crucial class of noncovalent interactions4, 10, 14 being weaker and more flexible than conventional covalent bonds, and, on the other hand, more directional than other intermolecular interactions. Such properties of hydrogen bonds allow for fine-tuning and optimising the selfassembly phenomena.15 There are several methods of analysis and description of mutual orientation of the crystal structure components in a crystal lattice in terms of recognition patterns.1, 10, 1618

For instance, structural motifs and the graph set notation describe features of existing

crystal structures, whereas the synthon concept indicates various ways, in which complementary molecules may approach one another. A thorough understanding of supramolecular synthons, their preferred geometries, competitive hydrogen bonds, interaction strength, etc., should result in rational design and non-covalent synthesis of novel assemblies.19-24 Such knowledge provides also some greater insight into various biological recognition occurrences, such as, ligand-protein binding.25 A systematic quantitative evaluation of binding strength and nature of synthons provides additional dimension to our understanding of the intermolecular interactions and their stabilizing role in the structure formation. In the current study, we continue our structural and energetic investigations of hydrogen bonds leading to hydroxylgroup⋯pyridine-fragment O−H⋯N heterosynthons in the presence of competitive hydrogen-bonding

functional

groups,

extending

our

previous

analysis

of

dihydroxybenzoic acids (DHBs) cocrystals with bipyridyls26 by those with nicotinamide (N) and pyrazinamide (P).27-29 Dihydroxybenzoic acids constitute a very interesting group of compounds in terms of crystal engineering, as they can exist as six positional isomers and form complex 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

hydrogen bond networks.30 Among others, DHBs can create various interaction patterns with N-donor compounds due to the variably arranged hydroxyl and carboxylic functionalities with competent ability to form robust hydrogen bonds.31-36 In turn, pyrazinamide is a compound of significant pharmaceutical interest, being a frontline anti-tuberculosis drug.37-42 Up to now, 30 different organic pyrazinamidecontaining cocrystal structures and one salt-like complex structure are known. Nevertheless, most of the reports concerning pyrazinamide are related to its pharmaceutical properties,37-42 or encompass its polymorphic behaviour.43-47 On the other hand, nicotinamide, being a GRAS (generally recognized as safe) compound, has been more extensively studied by crystal engineering methods,48-51 which resulted in a vast number of cocrystal structures (101 different cocrystals and 20 different salt-like structures) available in the Cambridge Structural Database (CSD).52 Both P and N have been previously thoroughly analysed by us in terms of charge density distribution.53 In this contribution we present four new cocrystal structures of DHBs with nicotinamide or pyrazinamide (Scheme 1), i.e., N:23DHB, P:23DHB, P:24DHB and P:26DHB. The series was supplemented with the analogous cocrystals already present in the CSD, namely P:25DHB54 (REFCODE: XAQQOW), N:25DHB55 and N:26DHB56 (REFCODEs: PEKRUU and LAGTOF & LAGTOF01, respectively; both synthesised and remeasured by us at 100 K), two polymorphs of N:24DHB57 (REFCODEs: DINRUP & DINRUP01; RT data; here denoted as N:24DHB-I and N:24DHB-II), and the respective representative literature-available single-component systems (for full information see Table 4). In the case of P and N which appear as various polymorphic forms, we have selected the most commonly crystallized polymorphs for comparison purposes. The analysed crystal structures were characterised in terms of crystal packing and structural motifs. For all the investigated cocrystals cohesive energy calculations in the CRYSTAL58-59 and PIXEL60-61 packages were performed, while the interaction energies were additionally derived for selected supramolecular motifs. Moreover, as apart from the strength of intermolecular interactions, their directionality is crucial in determining the crystal network overall stability and its physicochemical properties, energy frameworks62 were also calculated and examined.

4 ACS Paragon Plus Environment

Page 4 of 37

Page 5 of 37

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

Scheme 1. Components of the analysed cocrystals: nicotinamide (N), pyrazinamide (P), 2,3-dihydroxybenzoic acid (23DHB), 2,4-dihydroxybenzoic acid (24DHB), 2,5dihydroxybenzoic acid (25DHB), 2,6-dihydroxybenzoic acid (26DHB).

2. Experimental section 2.1. Materials and crystallisation. All substrates were purchased from commercial sources. Cocrystals were obtained from equimolar substrate methanol solution via standard solvent evaporation method. 2.2. X-ray data collection and refinement. Single-crystal X-ray measurements of all compounds were carried out on a classical diffractometer (Rigaku Oxford Diffraction, formerly KUMA Diffraction), equipped with the CCD detector and low-temperature open-flow device to keep the samples at the temperature of 100 K. Data collection strategies determination and optimisation, unit cell determination, raw diffraction image integration and data scaling were all performed using the appropriate algorithms implemented in the diffractometer software.63 All structures were solved using direct methods as implemented in the SHELXS program,64 and refined with the SHELXL program64 within the independent atom model (IAM) approximation. CIF files for each refinement are also present in the Supporting Information, or can be retrieved from the CSD (for deposition numbers see CCDC 1554966-1554971). 2.3. Hirshfeld surface analysis. This method allows for fast identification of the shortest possible intermolecular contacts and, subsequently, for their quantification. Hirshfeld surfaces (HSs) and fingerprint plots (FPs) for all the studied systems, i.e. for the nicotinamide molecule in N:23DHB, N:24DHB-I, N:24DHB-II, N:25DHB and 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

N:26DHB, and for the pyrazinamide molecule in P:23DHB, P:24DHB, P:25DHB, P:26DHB, were generated using the CRYSTALEXPLORER program,65 and are available in the Supporting Information. 2.4. Computational studies. The geometry of all crystal structures was optimised at the DFT(B3LYP)/6-31G** level of theory66-69 in the CRYSTAL program (CRYSTAL09 version)58-59 prior to further computational analyses. During the optimisation procedure cell parameters were kept fixed while the atom positions varied. Crystal cohesive energies for the studied crystal systems were calculated at the same level of theory. The results were corrected for dispersion70-72 and the basis set superposition error (BSSE).73 Ghost atoms used for the BSSE estimation were selected up to 5 Å distance from the considered molecule in the crystal lattice. The evaluation of Coulomb and exchange series was controlled by five thresholds, set to the values of 10−7, 10−7, 10−7, 10−7, 10−25. The general methodology was analogous to that described in our previous papers.74-76 The intermolecular interaction energies evaluation was conducted using the GAUSSIAN package77 (GAUSSIAN09 version). In this case the DFT(B3LYP)/aug-ccpVDZ66-68,

78-79

method was employed with the Grimme empirical dispersion

correction71-72 modified by the Becke-Johnson damping function,80-81 and correction for BSSE.73,

82

The GAUSSIAN package was also employed to calculate nicotinamide and

pyrazinamide amide fragment rotation barrier. Atomic partial charges were determined using the Merz-Kollman-Singh fit to electrostatic potentials.83-84 Geometry optimisations were performed at each rotational step (every 2°) keeping the restricted O−C−C−C torsion angle value ( ). All input files for either CRYSTAL or GAUSSIAN programs were prepared using the CLUSTERGEN program.85 Additionally, for comparison purposes and estimation of different total energy components, PIXEL60-61 energy calculations were performed. The optimised crystal structures were used to calculate the molecular electron density by standard quantum chemical methods with the GAUSSIAN program (GAUSSIAN98 version) at the MP2/631G** level of theory.69, 86 The electron density model of the molecule was then analysed using the PIXEL program package,60-61 which allows for calculation of dimer and cohesive energies. A cluster of molecules of radius 18 Å was used for lattice energy evaluation. PIXEL provides crystal cohesive energy and also its breakdown into electrostatic, polarisation, dispersion, and repulsion components. 6 ACS Paragon Plus Environment

Page 6 of 37

Page 7 of 37

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 CRYSTALEXPLORER program65 (CRYSTALEXPLORER17 version) was also used to evaluate interaction energies for selected dimers present in the studied cocrystals. Similarly as in PIXEL, the total energy of a given dimer constitutes the sum of electrostatic, polarisation, dispersion and exchange-repulsion components. The molecular electron density was calculated at the DFT(B3LYP)/6-31G** level of theory using the above-described CRYSTAL-optimised molecular geometries. The dimer interaction energies were further utilized to generate the so-called energy frameworks.62

3. Results and discussion 3.1. General remarks. As nicotinamide is a GRAS substance and a stated co-former, a few database analyses of its cocrystals with carboxylic acids have already been reported.49-50 Hence, here we just recall the robust synthons formed between the N molecule and carboxylic acids. The 71 different nicotinamide:carboxylic-acid cocrystals available in the CSD follow the Etter’s empirical rule16 that intermolecular hydrogen bonds are formed between the best proton donors and acceptors remaining after formation of intramolecular hydrogen-bonds. Thus, in the reported complexes the most acidic proton from the carboxylic fragment, which is the best hydrogen bonding donor, should interact with the best hydrogen bonding acceptor, i.e. the basic lone pair of the pyridine nitrogen atom. Indeed, such an interaction (synthon I; see Figure 1 for all synthon schemes) is present in a vast majority of the nicotinamide cocrystals, i.e., 57 which constitutes 80% of the analysed population. In certain cases, characterised by a sufficient difference of the relative acidity and basicity of the two components, formation of synthon I can even induce a proton transfer between carboxylic and pyridine centers. In turn, the most prevailing motif involving the N amide group is synthon V.

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 1. Selected hetero- (I, I’, IST, II, II’, III, III’, IV) and homosynthons (V, VI, VII, VIII), which appear in the available cocrystals of nicotinamide and pyrazinamide with aromatic carboxylic acids (either from the CSD-derived, or our newly synthesised structures; (N) atom means the dimer is present in cocrystals of both nicotinamide and pyrazinamide). Please note that some dimers may in fact exhibit various conformations and/or be supported by a variety of secondary interactions in the solid state, and here are presented in an idealised way.

In the case of pyrazinamide, the literature-available set of cocrystals is much more limited than that for nicotinamide. Up to now, 18 different pyrazinamide structures with carboxylic acids have been reported. As the P molecule contains additional nitrogen atom in the aromatic ring when compared to N, namely N1 at the closest vicinity to the amide fragment (Figure 2), it can be noticed that P is more robust and conformationally predictable. This is reflected in the usually lower number of various synthons encountered in its known cocrystals. However, at the same time an additional 8 ACS Paragon Plus Environment

Page 8 of 37

Page 9 of 37

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

synthon VII can be found. Concerning the reported cocrystals with carboxylic acids, in 6 of them the mixed P:carboxylic-acid synthon II is formed. Otherwise, synthon V engages the amide group, whereas the carboxylic moieties may either interact with one another, or with the N2 atom. Considering the character of nicotinamide and pyrazinamide molecules and that of dihydroxybenzoic acids, it is apparent that the mutual recognition of such crystal components is based mainly on the interaction of −OH and −COOH with the heteroatoms of N-donor compounds. In this respect, the chemical nature of heteroatoms in N-donor molecules is crucial to understand preferences in synthon formation in the studied series of cocrystals, and to explain the trends observed when analysing the CSD entries. As already mentioned before, an obvious difference between pyrazinamide and nicotinamide species is the presence of additional nitrogen in the aromatic ring of P. This feature significantly changes the conformational preferences of the amide substituent and the mutual basicity of the pyrimidine nitrogen atom and carbonyl group in P when compared to N. The density functional theory (DFT) calculation of atomic partial charges fitted to the electrostatic potential (Figure 2) confirms that structural differences between these two moieties have a significant effect on the charge density distribution on the heteroatoms. The same conclusion can be drawn when comparing the QTAIM-derived charges from our previous work.53 Furthermore, computations show a minor, though noticeable, effect of the amide group conformation on the charge distribution (which is naturally more visible in the case of P due to the formation of the intramolecular hydrogen bond in the anti conformation).

(a)

(b)

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

(c)

(d)

Figure 2. Partial charges fitted to electrostatic potential (DFT(B3LYP)/aug-cc-pVDZ level of theory) for both nicotinamide (a,b) and pyrazinamide (c,d) in the energy minima corresponding to their syn (a,c) and anti (b,d) conformations. Please note that for (a), (b) and (c) figures the amide-group-to-aromatic-ring torsion angle deviates significantly from 180°, whereas in (d) the molecule is planar (due to the presence of intramolecular hydrogen bond) (see also Figure 3).

In the case of nicotinamide, the pyridine N-atom (N1, Figure 2a,b) is the best nucleophilic center, and thus should preferably interact with the −COOH group of DHBs leading to synthon I. In turn, the carboxamide fragment, being both a proton donor and acceptor, should form hydrogen bonds with the −OH groups of DHBs. The assignment of the pyridine nitrogen of N as the most nucleophilic site is supported by its most negative charge, c.a. −0.70 e, followed by the amide carbonyl oxygen (−0.53 e). In the case of P, N1 influences the molecular charge distribution and, consequently, pyridine nitrogen N2 has reduced nucleophilic character (Figure 2d). Thus, different synthons involving pyrimidine N atoms and the amide group are favoured. The O-atom from the amide group constitutes the best proton acceptor and, hence, should bind with the carboxyl group, the most acidic site, leading to the synthon II. Since the N2 is more accessible and constitutes a better nucleophile than N1, it is preferably involved in intermolecular hydrogen bonds, binding with the −OH groups of DHBs (synthon I’). N1 may participate in weaker intermolecular interactions, due to its close vicinity to the amide fragment with which N1 preferably forms intramolecular hydrogen bond in accordance with Etter’s rules. A visible change in charge density is also observed at the carbon and hydrogen atoms in the aromatic ring fragment. However, these protons are 10 ACS Paragon Plus Environment

Page 10 of 37

Page 11 of 37

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

significantly less acidic than those from the amide moiety and such differences will only affect the formation of C−H⋯π interactions and other secondary contacts. Another important issue, when comparing P and N molecules, is the rotational liability of the amide substituent, thus the conformational energy analysis concerning the O−C−C−C torsion angle has been performed. In the case of nicotinamide, there are two main energy minima, i.e. a global minimum around ±162° (syn amide group conformation, Figure 2) and a local minimum at about ±22° (anti amide group conformation, Figure 2), as the energy curve has a mirror plane at 0° (these results are in agreement with similar calculations for nicotinamide performed earlier by Bathori et al.49 and Lemmerer et al.).50 The energy difference between the two orientations of the amide group amounts to only about 4 kJ·mol−1 in favour of the syn arrangement, whereas the energy barrier between the syn and anti conformations reaches 16 kJ·mol−1. Therefore, the amide group in N can exhibit either syn or anti orientation according to the N-pyridine atom. This is well reflected in the conformation histogram based on the nicotinamide containing structures in the CSD (Figure 3a). There are two well-defined conformational peaks corresponding to the two energy minima, among which the more stable syn conformation is indeed more populated.

(a)

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

(b)

Figure 3. Graphs for nicotinamide (a) and pyrazinamide (b) showing plots of relative molecular energy ( ) vs. the O−C−C−C torsion angle (), and the CSD-derived histograms for the same angle. Next to the -constrained optimisation energy minima the respective idealised-geometry schemes are presented.

In turn, from the respective histogram and the energy curve (Figure 3b) it is evident that the P molecule is conformationally more restricted. The presence of N1 in the aromatic ring hampers the rotation of the amide fragment. N1 interacts with the H atom from the amino group, forming a weak intramolecular hydrogen bond, which stabilises the anti conformation of the amide moiety according to the N2 atom. The anti conformation is strongly favoured (over 30 kJ·mol−1 difference between the global (0°, anti) and local (±138°, syn) energy minima). Indeed, the amide group in all the measured (Table 1) and literature-available structures is characterized by the O−C−C−C torsion angle around 0°. When slightly rotated, it is just due to the crystal packing effects.

Table 1. O−C−C−C torsion angle () values exhibited by nicotinamide and pyrazinamide in the analysed experimental and the literature-available supplementary crystal structures. (Every value is followed by the description of the molecule conformation; N/Pα/β means either α or β phase for nicotinamide or pyrazinamide).

12 ACS Paragon Plus Environment

Page 12 of 37

Page 13 of 37

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 structure X-ray geom. Optimised geom. N:23DHB +147.2(2) (syn) +146.25 (syn) a,b N:24DHB-I +10.6 (anti) +16.22 (anti) −j +18.54 (anti) N:24DHB-I a,c N:24DHB-II a,d +179.3 (syn) +170.04 (syn) N:25DHB +159.2(1) (syn) +157.91 (syn) N:25DHBRT a,e −158.5 (syn) −157.36 (syn) −11.5(2) (anti) −12.37 (anti) N:26DHB f P:23DHB +7.9(2) (anti) +9.35 (anti) P:24DHB −3.9(3) (anti) −4.39 (anti) P:25DHB +0.3(3) (anti) +1.08 (anti) P:26DHB +11.0(2) (anti) +13.02 (anti) Nα a,g −157.5 (syn) −156.81 (syn) Pα a,h −4.7 (anti) −7.12 (anti) Pβ a,i −2.3 (anti) −3.53 (anti) a Structure retrieved from CSD and optimised with the CRYSTAL program (additionally underlined). b The first polymorph of the N:24DHB structure (REFCODE: DINRUP) with the OH group disorder removed (the ortho-OH group is farther away from the protonated oxygen from the carboxylic group). c Same as point (b) (REFCODE: DINRUP) but with the OH group located at the second ortho-position. d The second polymorph of the N:24DHB cocrystal (REFCODE: DINRUP01). e Same as N:25DHB but measured at room temperature (REFCODE: PEKRUU). Note change in sign due to the different enantiomer. f Ionic structure. g REFCODE: NICOAM01. h REFCODE: PYRZIN15. i REFCODE: PYRZIN18. j Same experimental structure as in point (b) – only optimised structures are different.

3.2. Cocrystals of nicotinamide with dihydroxybenzoic acids. The aforesaid features of the molecular charge density distribution and, thus, properties of nicotinamide and pyrazinamide are reflected in the favourable structural motifs present in their respective cocrystals. As the nitrogen atom from the heteroaromatic ring in N is more negative than the corresponding one in P (Figure 2), nicotinamide forms preferably synthon I with DHBs rather than synthon II, which is predominant in the P equivalents. In the case of synthon I the carboxylic group is involved in the relatively strong O−H⋯N hydrogen bond with the pyridine N atom. The DFT-estimated interaction energy characterizing such a synthon varies between −60 and −70 kJ·mol−1 (Table 2). The only nicotinamide cocrystal structure, in which synthon I is not observed, is the N:24DHB-II polymorph. Instead, synthons I’ and II are encountered in N:24DHB13 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

II, among which the latter one is characterised by the most stabilising interaction energy (−86.7 kJ·mol−1) in the N series. The flexibility of the amide group from N and the fact that it is rarely involved in synthon II, facilitate creation of a greater variety and number of hydrogen bonds when compared to P. Consequently, this function assists formation of heterosynthons II’, III, III’, IV and homosynthon VI. Among them, the well-stabilised synthon II’ (−43.5 kJ·mol−1), which can be described as a ring motif R 9, is particularly distinctive and formed by N solely with the 23DHB acid thanks to its neighbouring hydroxyl groups. Due to multiple hydrogen bond donor and acceptor centers in the coformers, hydrogen bonding interactions in the studied cocrystals extend the described synthons into 1-3D architectures stabilised further by π⋯π secondary interactions and other dispersive contacts. Hence, in the case of N:23DHB one can observe molecular planes formed by hydrogen-bond-connected ladder-like motifs (Figure 4a), which then interact one with another by C−H⋯π contacts. In turn, hydrogen-bonded chain motifs are present in both literature-reported polymorphs of N:24DHB. In the case of the N:24DHB-II form (REFCODE: DINRUP01) flat chains are created by synthons II (R 8 – carboxyl-amide group interaction) and I’ (interaction of the para-hydroxyl group with the N pyridine atom). The adjacent chains are connected via the N−H⋯O hydrogen bonds forming molecular tapes (Figure 4c). Various C−H⋯O contacts link such tapes into molecular layers, while π⋯π stacking interactions stabilise further such a layered architecture. In N:24DHB-I chain motifs resemble more closely these encountered in the remaining nicotinamide cocrystal structures, i.e., synthon I is the major dimer, whereas the amide group is involved in three heteromeric interactions via hydrogen bonding (Figure 4b). The chains are inter-connected one with another forming an efficient 3D hydrogen-bonded network. Similar chain motifs and 3D hydrogen-bonded architecture based on synthons I, III’ and IV (Table 2, Figure 4d), is found in N:25DHB. The only structure in the studied series, in which a proton transfer takes place, is N:26DHB (Figure 4e). Alike in the majority of other nicotinamide cocrystals, synthon I is primarily formed here. As expected, the two ortho-hydroxyl groups from 26DHB are involved in intramolecular hydrogen bonding. Moreover, oxygen atoms from these hydroxyl groups are acceptors for protons from the amide fragment of N. Interestingly, carbonyl oxygen atom from the amide function is not engaged in any significant interactions, and mediates only some weak C−H⋯O contact. Synthon I and two types of 14 ACS Paragon Plus Environment

Page 14 of 37

Page 15 of 37

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

synthon III’ expand the hydrogen-bonded network into flat molecular layers, interacting further via π⋯π stacking stabilising the 3D crystal architecture.

Table 2. Interaction energies for selected hydrogen-bonded synthons present in the  ¶ analysed cocrystals ( are the energies computed with the GAUSSIAN program).

 ¶ Dimer  motif / kJ·mol−1 N:23DHB I −67.9 II’ −43.5 VI −42.0 N:24DHB-I a,b I −58.8 III −20.3 III' −14.9 IV −49.3 I −64.9 N:24DHB-I a,c III −25.2 III' −14.6 IV −51.1 N:24DHB-II a,d I' −50.1 II −86.7 III' −23.2 N:25DHB I −66.7 III' g −19.4 g III' −17.0 IV −44.0 I −66.3 N:25DHBRT a,e III' g −19.5 III' g −17.5 IV −44.0 N:26DHB f IST −410.7 III' g −292.3 III' g −223.8 P:23DHB I' −31.7 II −82.1 P:24DHB I' −43.7 II −81.8 VII −23.6 P:25DHB I' −41.8 II −84.4 III' −17.6 P:26DHB I' −21.0 II −96.2 III' −12.1 ¶ DFT(B3LYP)/aug-cc-pVDZ method with the Grimme empirical dispersion correction modified by the BeckeJohnson damping function, and BSSE-corrected. a

Cocrystal

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 37

Structure retrieved from CSD and optimised with the CRYSTAL program (additionally underlined). b The first polymorph of the N:24DHB structure (REFCODE: DINRUP) with the OH group disorder removed (the ortho-OH group is farther away from the protonated oxygen from the carboxylic group). c Same as point (b) (REFCODE: DINRUP) but with the disordered OH group placed at the second ortho-position. d The second polymorph of the N:24DHB cocrystal (REFCODE: DINRUP01). e Same as N:25DHB but measured at room temperature (REFCODE: PEKRUU). f Ionic structure. g Two different dimers of this type present in the structure.

(a)

(b)

(c)

(d)

(e)

16 ACS Paragon Plus Environment

Page 17 of 37

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 4. Structural motifs in cocrystals with nicotinamide: (a) N:23DHB, (b) N:24DHB-I

b

(see Table 2 for explanation of the upper index b) (c) N:24DHB-II, (d)

N:25DHB, and (e) N:26DHB.

3.3. Cocrystals of pyrazinamide with dihydroxybenzoic acids. As mentioned before, a general observation which can be made for pyrazinamide cocrystals with dihydroxybenzoic acids is that in all the studied cases synthon II is preferably formed. It involves the carboxylic group from the acid molecule and the amide fragment from P. Such created molecular complexes are stabilised via two hydrogen bonds leading to the R 8 ring motif according to Etter’s notation. The corresponding interaction energy varies from −81.8 kJ·mol−1 for P:24DHB up to −96.2 kJ·mol−1 for P:26DHB (Table 2). Due to the two relatively strong hydrogen bonds this synthon is the strongest-stabilised one among the studied cocrystals. The other synthon present in all the pyrazinamide cocrystals is I’, which also constitutes the second motif in the P:acid set in terms of interaction energy. In the case of P:24DHB and P:25DHB the derived interaction energy exceeds −40.0 kJ·mol−1, whereas it is weaker in the P:23DHB and P:26DHB cocrystals (Table 2) due to the close vicinity of an additional oxygen atom to the interacting OH moiety. In the case of 23- and 24DHBs, the hydroxyl group in positions 3 or 4 facilitates chain motif formation (Figure 5a,b). Furthermore, in P:24DHB the neighbouring chains are interconnected via hydrogen bond interactions between inversion-center-related pyrazinamide molecules (synthon VII, see Figures 1 and 5b, dimer interaction energy equals −23.6 kJ·mol−1). This leads to a more energetically efficient packing than it is in the case of P:23DHB, in which only more distant weaker interactions are observed. In the P:25DHB and P:26DHB cocrystals the DHB hydroxyl groups at positions 5 and 6 stimulate ring motif rather than chain pattern formation (Figure 5c,d). Additionally, in P:25DHB the oxygen atom from the hydroxyl group, beside being a proton donor in hydrogen bonding with N2 from pyrazinamide, acts as an acceptor for the proton from the amide group of pyrazinamide (synthon III’, dimer interaction energy equals −17.6 kJ·mol−1). Ring motifs form further undulated hydrogen-bonded layers interacting one with another by dispersive interactions. In turn, in P:26DHB such molecular rings lead to infinite tapes (Figure 5d). Such tapes interact via π⋯π stacking and some other secondary contacts. In general, no 3D hydrogen-bonded network is observed in the P 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

Page 18 of 37

cocrystal series. It is also worth mentioning that in P:26DHB the heterosynthon III’ is weaker-stabilised (dimer interaction energy equals −12.1 kJ·mol−1) than the P⋯P homosynthon present in this crystal structure (the respective dimer interaction energy amounts to −16.4 kJ·mol−1).

(a)

(b)

(c)

(d)

Figure 5. Structural motifs in cocrystals with pyrazinamide: (a) P:23DHB (b) P:24DHB (c) P:25DHB, and (d) P:26DHB.

3.4. Molecular motifs. The interaction energies of the above mentioned dimeric motifs were calculated using three basic approaches. The classical DFT (performed with the GAUSSIAN program) and PIXEL methods were supplemented with a new fast and efficient approach recently proposed by Turner et al.62 Figure 6 shows comparison of the GAUSSIAN, PIXEL and CRYSTALEXPLORER results obtained for the hydrogen-bonded complexes in the analysed series of cocrystals (for computational details see the Experimental Section). It occurs that all used computational methods provide the same energy trends ranking dimers in a given cocrystal structure. Accordingly, a motif strongest-stabilised in terms of interaction energy is synthon II and next I. In turn, IV, 18 ACS Paragon Plus Environment

Page 19 of 37

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

VI and VII are characterised by medium energy values, while synthons III and III’ are weakest-stabilised. Regarding the strongest-bound dimers the most advantageous energies were obtained using GAUSSIAN, then with the CRYSTALEXPLORER approach, while the least favourable values were derived from the PIXEL computations. The differences between the three approaches are less pronounced for medium-strength interactions. A comparison of relative contributions of electrostatic, dispersive, polarization and repulsion terms between PIXEL and CRYSTALEXPLORER (see Tables 3S & 4S in the Supporting Information) reveals that, while the electrostatic energy contributions are very much alike, other terms differ significantly.

Figure 6. Interaction energy values for selected hydrogen-bonded molecular complexes present in the analysed cocrystal structures. Colour coding: blue (left) – DFT(B3LYP)/aug-cc-pVDZ (GAUSSIAN), red (middle) – MP2/6-31G** (PIXEL), green (right) – DFT(B3LYP)/6-31G** (CRYSTALEXPLORER).

Importantly, in the analysed cocrystals a heterodimer always constitutes the strongestbound molecular complex, which is one of the primary factors supporting cocrystal formation. In the case of synthons I and II the respective synthon interaction energies are relatively robust, varying only within 10 kJ·mol−1 between different cocrystals. In contrary, the interaction strength describing synthon I’ is more variable, as it depends noticeably on the location of the hydroxyl group involved in the hydrogen bond. Thus, dimer interaction energies may differ here by up to 30 kJ·mol−1. Among the P cocrystals this interaction is strongest when the OH group is located at positions 4 or 5 at the acid 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

aromatic ring, i.e. possibly far from the second OH group and the COOH function. In the case of N cocrystals I’ is present only in the second polymorph of N:24DHB and the respective interaction energy outnumbers that calculated for the P equivalent by about 6 kJ·mol−1. Interestingly, N:24DHB-II resembles P:25DHB in terms of synthons and their stabilisation energies. In turn, the synthons encountered in the N:24DHB-I crystal structure (i.e. I, III, III’, IV) are alike those present in N:25DHB and are described by similar interaction energy values. All synthons can also be easily identified and compared among the studied series through the analysis of the closest contacts contribution to the Hirshfeld surface constructed either for N or P (see Figure 1S in the Supporting Information). The examination of partial contributions of particular contact types to HS (Figure 7) led to a number of remarkable conclusions. The percentage of the N⋯H contacts vs. O⋯H is greater in the molecular complexes of P, whereas the opposite is observed for the N cocrystals. Obviously this is the result of two N-atoms available for non-covalent interactions in P when compared to only one in N. Yet, the sum of hydrogen bondinglike interactions (i.e., the above-mentioned N⋯H and O⋯H contacts) is similar throughout the whole series. However, the structural analysis indicates that more complex hydrogen-bonded patterns were found for the nicotinamide-containing cocrystals due to its greater flexibility than P, which was described earlier in the text. Finally, the contribution of the N⋯H contacts is lowest in N:26DHB because of the proton transfer present in its structure.

(a)

20 ACS Paragon Plus Environment

Page 20 of 37

Page 21 of 37

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 7. Percentage contributions to Hirshfeld surface of (a) nicotinamide and (b) pyrazinamide in the analysed cocrystals with DHBs.

Naturally, not only hydrogen bonding interactions contribute to the crystal network stability. As all used coformers are aromatic molecules, various dispersive interactions are also noticeable and essential to understand the formed crystal architectures. This is confirmed by the HS analysis indicating that, for instance, the C⋯C contribution to the respective HS area for N:24DHB-II, N:25DHB, P:23DHB and N:26DHB, reflecting majorly π⋯π stacking, is higher than 10%. For the first three systems dimer interaction energies for molecules involved in π⋯π stacking range from −10 to −25 kJ·mol−1 (according to the CRYSTALEXPLORER results; see Figure 8 and Supporting Information). As both N and P possess nitrogen atoms in the aromatic rings, high contribution of C⋯N to the HS area may also indicate some π⋯π stacking interactions. This is the case for N:24DHB-I, in which the C⋯C interactions’ contribution to HS does not exceed 5%. However, when these are considered together with the C⋯N and C⋯O contacts, it gives more than 15% overall contribution to the HS area. Indeed, the analysis of dimer interaction energies reveals that for this particular structure the most energetically favourable stacking contacts are observed (e.g., −42 kJ·mol−1 between two acid molecules, see Figure 8). In turn, in the case of P:26DHB and N:23DHB, even though the molecules of P or N and acid form layered architecture, the offset between aromatic rings from the adjacent layers is high enough to reduce significantly contributions from the C⋯C contacts. Thus, in N:23DHB these are substituted by the greater contribution from the C⋯H contacts, which reaches 20%, being around two times higher than that 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

Page 22 of 37

derived for the remaining structures (Figure 7). It means that here the hydrogenbonded network is further stabilized by C−H⋯π and C−H⋯C interactions (from −10 to −15 kJ·mol−1, see Figure 8d) rather than π⋯π stacking. In the case of P:26DHB both C⋯H and C⋯N contacts support additionally the crystal lattice stability.

(a)

(b)

(c)

(d)

Figure 8. Total interaction energy calculated for molecular complexes stabilised by π⋯π stacking (a) N:24DHB-I, (b) N:24DHB-II, (c) P:26DHB, and C−H⋯π interactions (d) N:23DHB (energy values are given in kJ·mol−1).

3.5. General discussion of the studied cocrystal structures and related cohesive energies. The structural studies indicate that, while nicotinamide or pyrazinamide molecules may define the formation of the main synthons, the overall structural network strongly depends on the dihydroxybenzoic acid type. This is because the orientation and the accessibility of the OH groups are critical factors determining the interaction geometry, thus the hydrogen bond network type and also the stability of 22 ACS Paragon Plus Environment

Page 23 of 37

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 crystal architecture. Furthermore, in all the acids the ortho-OH group is locked in a 6-membered intramolecular hydrogen bond motif, which leads to the decreased stabilisation of other hydrogen bond interactions it is involved in. In the case of 26DHB, both OH groups are engaged in intramolecular contacts. Similarly, an intramolecular interaction is formed between the N1 atom and the amine group in the pyrazinamide molecule, which makes the amide group less flexible than that of nicotinamide (see Figure 3). Considering such restrictions and the above structural motif analysis, it appears that 24DHB, 25DHB and the mobile amide group of N provide greatest flexibility in terms of molecular inter-connections. First general conclusions regarding crystal network stability can be made on the basis of the so-called energy frameworks analysis (Figure 9). Such frameworks indicate the directionality and strength of main interactions present in a crystal lattice. In the case of the studied series the analysis shows that for the P cocrystals, as well as for N:24DHB-II, synthon II is clearly the most energetically advantageous one, constituting, major subunit building chains and layers. As can be seen in Figure 9 the presence of such a strongly-bound synthon imposes less isotropic distribution of interactions in different directions. In contrast, in the cocrystals containing nicotinamide (excluding N:24DHB-II), in which the synthon II does not appear, more significant interactions are of similar strength and are quite uniformly distributed in space. This observation is well illustrated by the example of the N:24DHB polymorphs. In the case of N:24DHB-I interactions are in general relatively isotropically distributed along the ,  and  directions (especially along  and ), whereas for N:24DHB-II strongly-stabilised molecular chains lead to the more pronounced directionality of interactions in the crystal lattice. Importantly, energy frameworks reflect to some extent crystal mechanical properties.62 Diagrams showing separately electrostatic and dispersive energy frameworks are available from the Supporting Information.

N:23DHB

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

N:24DHB-I

N:24DHB-II

N:25DHB

P:23DHB

P:24DHB

P:25DHB

P:26DHB

24 ACS Paragon Plus Environment

Page 24 of 37

Page 25 of 37

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 9. Total energy frameworks generated for the studied cocrystals. Line thickness indicates the interaction energy value (the thicker the line the greater the energy). Views are presented along the  (2nd column),  (3rd column) and  (4th column) axes.

Additionally, the cocrystal cohesive energies and the corresponding cohesive energies of cocrystal component structures were evaluated. In all the cases, cocrystal cohesive energies are more advantageous than the sum of the respective single component crystal cohesive energies (Tables 3 & 4), which supports the cocrystal formation.

Table 3. Cohesive energy values ( ) computed for optimised crystal structures using either CRYSTAL or PIXEL programs. Note the energies are given per two ASU contents together. Packing coefficient () was estimated with the OLEX2 program.87 $%&! "# ! '% / kJ·(2mol)−1 / kJ·(2mol)−1 N:23DHB −254.8 −226.4 71 N:24DHB-I a,b −273.7 -230.0 70 N:24DHB-I a,c −295.5 -250.2 70 −270.2 -234.6 70 N:24DHB-II a,d N:25DHB −274.4 −240.2 71 N:25DHBRT a,e −271.5 −235.0 71 N:26DHB f −653.1 − 75 P:23DHB −237.0 −210.0 73 P:24DHB −256.7 −230.1 72 P:25DHB −253.0 −236.4 73 P:26DHB −240.8 −219.4 75 a Structure retrieved from CSD and optimised with the CRYSTAL program (additionally underlined). b The first polymorph of N:24DHB structure (REFCODE: DINRUP) with the OH group disorder removed (the ortho-OH group is farther away from the protonated oxygen from the carboxylic group). c Same as point (b) (REFCODE: DINRUP) but with the disordered OH group at the second ortho-position d The second polymorph of N:24DHB cocrystal (REFCODE: DINRUP01). e Same as N:25DHB but measured at room temperature (REFCODE: PEKRUU). f Ionic structure. Cocrystal

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 37

Table 4. Crystal cohesive energies calculated for CRYSTAL-optimised single cocrystal component structures on the basis of structural data provided by CSD (computations done with CRYSTAL and PIXEL programs). (N/Pα/β means either α or β phase for nicotinamide or pyrazinamide; ( – measurement temperature as reported). Crystal CSD Space $%&! "# ! )/K −1 structure REFCODE group / kJ·mol / kJ·mol−1 23DHB CACDAM '1+ RT c −118.6 d −113.8 d '2- // 24DHB ZZZEEU02 100 −132.6 −116.2 '2- // 25DHB BESKAL RT −131.0 −122.0 '/0226DHB-I a LEZJAB RT −129.1 −113.5 '2- /1 26DHB-II b LEZJAB01 RT −113.6 −103.0 '2- /1 NICOAM01 150 −121.3 −115.4 Nα '2- // Pα PYRZIN15 100 −106.7 −88.60 '2- /1 PYRZIN18 100 −109.0 −103.7 Pβ a First polymorph. b Second polymorph. c RT – room temperature. communication.

Reference Okabe & Kyoyama, 200134 Parkin et al., 200731 Haisa et al., 198235 Gdaniec et al., 199432 MacGillivray & Zaworotko, 199433 Miwa et al., 199948 Nangia & Srinivasulu, 2006 e Cherukuvada et al., 201046 d Averaged energy. e CSD private

As presumed, pyrazinamide cocrystals are less energetically favoured than their nicotinamide equivalents. The least stabilising cohesive energy among the studied systems is observed for P:23DHB. Interestingly, the 23DHB acid also forms the least energetically favourable cocrystal with N. This is attributed to the arrangement of the hydroxyl groups in the acid molecule, which reduces the number of beneficial intermolecular interaction patterns. For instance, the only hydrogen-bonded motif that exists in the structure of P:23DHB is a single ribbon (based on synthons II and I’), while the N1 atom is not involved in any significant intermolecular contacts. In the case of P:26DHB, a hydrogen-bonded tape motif is formed, which involves three types of synthons and results in only slightly better stabilised crystal lattice (by about 4 kJ·mol−1). It should be noted, however, that the 26DHB acid is a specific case, being the most acidic species among the chosen DHB series (Table 5). On the other hand, N is much more basic than P. These two facts result in a ∆p45 significantly higher for N:26DHB than for all other cocrystal component pairs in the studied set.

Table 5. ∆p45 values characterising cocrystal components as referenced to 6/7

nicotinamide (p45 = 3.35) or pyrazinamide (p45 = 0.50) (∆p45 = p45

26 ACS Paragon Plus Environment

− p459:; ).

Page 27 of 37

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

DHB component 23DHB 24DHB 25DHB 26DHB

p45 2.96 3.32 3.01 1.30

Cocrystals with N ∆p45 Type +0.39 neutral +0.03 neutral +0.34 neutral +2.05 ionic

Cocrystals with P ∆p45 Type −2.46 neutral −2.82 neutral −2.51 neutral −0.80 neutral

∆p45 can generally be considered as a useful indicator for the formation of cocrystals vs salts when acids and bases are the reactant constituents.88 According to that concept, ∆p45 value (as computed in Table 5) greater than 2 or 3 is a typical threshold for salt formation. Yet, there are also other criteria based on experimental observations. For instance, Johnson & Rumon89 proposed a higher limit, with ∆p45 > 3.75 for possible proton transfer. Although the ∆p45 values are usually reliable indicators of salt formation when ∆p45 > 3, we stress that the ∆p45 range 0 to 3 is rather ambiguous. Thus, cocrystal formation is expected when ∆p45 < 0, but both neutral-type and salttype crystals can be formed between 0 and 3. In the studied series ∆p45 = 2.05 occurred to be sufficient to indicate a proton transfer observed in N:26DHB. All other values are close to 0, or negative, therefore, they support the observed neutral cocrystal forms. Due to the presence of charged species, the cohesive energy of N:26DHB is particularly high in amplitude. The nicotinamide molecule and 2,4- and 2,5-dihydroxybenzoic acids are the components which have potential to form the best stabilised cocrystals among the studied compounds. The nicotinamide molecule is more flexible than pyrazinamide, as the rotation of the amide group is less constrained. It may adopt different structural conformations and, therefore, more effectively utilise the intermolecular contacts when compared with pyrazinamide. 24DHB’s and 25DHB’s advantage over the other acid molecules is the aforementioned arrangement of the hydroxyl substituents which guarantees the accessibility of these two hydrogen-bond-donor sites. Indeed, the P:24DHB and P:25DHB crystal structures are characterised by comparable cohesive energy and are the most stable ones among the P cocrystals, whereas both polymorphs of N:24DHB, as well as, N:25DHB constitute the most energetically favoured crystal structures in the whole series. Furthermore, the flexibility of N, and the mentioned accessibility of the hydroxyl groups of 24DHB, resulted in the only one pair of polymorphs observed for the studied cocrystals.

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

4. Summary In the studied systems the major factor governing the main synthons formation is the mutual basic character of proton acceptors in pyrazinamide and nicotinamide, i.e. the carbonyl fragment and aromatic nitrogen atoms. Regarding the acid molecules, certainly the competitive affinities of the COOH and OH groups towards the N-donor compounds are crucial, including the relative location of all functions. The carboxyl group, being the most acidic moiety, binds preferably to the strongest base, thus either to the amide group in the case of pyrazinamide forming the most energetically favoured synthon II, or to the pyridine N atom of the nicotinamide molecule leading most often to motif I. In general, pyrazinamide, characterised by greater rigidity than nicotinamide, creates more preserved basic structural motifs and less complex crystal architectures. This is because of the hampered rotation of the amide group involved in the intramolecular hydrogen bonding interaction, which makes P less capable to adjust to the other molecules constituting a crystal lattice. In turn, in the case of nicotinamide there are two distinct energy minima in the amide group rotation energy scan, which differ of about 4.6 kJ·mol−1. Thus, N can adopt almost equally well both syn and anti conformation in a crystal. Consequently, crystal lattice stabilisation energy values for the cocrystals formed with the corresponding dihydroxybenzoic acid moieties are more advantageous for nicotinamide, whereas patterns created by pyrazinamide are more predictable. In the latter case, as illustrated by the energy frameworks, preferred formation of synthon II, significantly outweighing other motifs in terms of interaction energy (exceeding −80 kJ·mol−1), leads to molecular patterns of certain directionality of interactions in a crystal structure, which is opposite to more energetically uniform motifs, and thus more isotropic crystal lattice, in the case of N cocrystals. Another important factor determining crystal architecture concerns the accessibility of hydroxyl groups in the aromatic ring of the benzoic acid. Although the lattice energies of the analysed acid mono-component crystals are quite alike, the energetic differences are clearly visible in their cocrystal structures. The most complex crystal architectures are formed by 2,4- and 2,5-dihydroxybenzoic acids, while 2,3dihydroxybenzoic acid limits the crystal network most notably. Therefore, cocrystals containing N and 24DHB or 25DHB exhibit the most advantageous cohesive energy values (excluding the salt-like structure of N:26DHB), and also much greater variety of 28 ACS Paragon Plus Environment

Page 28 of 37

Page 29 of 37

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 patterns and energetically significant intermolecular contacts than the remaining structures. Furthermore, N:24DHB is the only case when polymorphic structures are encountered. The mutual location of hydroxyl groups also determines the acidity of a given acid molecule. In the case of N:26DHB the difference of the p45 values of both components is high enough to stimulate proton transfer between N and 26DHB leading to the only salt type structure among the series. Despite the fact that in all studied cocrystals main structural motifs are based on hydrogen bonds being in accord with the Etter’s rules, the dispersive forces surely cannot be neglected. This is especially true for P:23DHB, N:24DHB-II, N:25DHB and N:26DHB cocrystals, where there is a notable contribution of π⋯π stacking interactions to the total lattice energy value. Finally, it should be noted that in all the cases the best stabilised molecular synthon is a heterodimer, while cocrystal cohesive energies are more favourable than that of the respective single component crystal structures, which supports the cocrystal formation.

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

Associated content Supporting information: The Supporting Information is available free of charge on the ACS Publications website at DOI: [to be filled by the editorial office]. Experimental details regarding crystallographic data, interaction energy values calculated for selected hydrogenbonded synthons, details on thermal analysis, energy contributions to the total interaction energy computed with the CRYSTALEXPLORER and PIXEL programs, Hirshfeld surfaces and Hirshfeld fingerprint plots generated for the pyrazinamide cocrystals, dispersive energy frameworks generated for the studied cocrystals and a comment on the energy frameworks for the studied cocrystals to all the following figures. Accession codes: CCDC 1554966-1554971 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk /data_request/cif, or by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033. Author information Corresponding authors: Katarzyna N. Jarzembska ([email protected]) Krzysztof Woźniak ([email protected]) ORCID: Katarzyna N. Jarzembska: 0000-0003-4026-1849 Radosław Kamiński: 0000-0002-8450-0955 Notes: The authors declare no competing financial interest. Acknowledgements K.N.J. and A.A.H. would like to thank the PRELUDIUM grant (2011/03/N/ST4/02943) of the National Science Centre in Poland for financial support. The Wrocław Centre for Networking and Supercomputing (grant No. 285) and the Interdisciplinary Centre for Mathematical and Computational Modelling in Warsaw (G33-14) are gratefully acknowledged for providing computational facilities. S.V. thanks DST, New Delhi for the start-up research grant. The Authors would like to thank Mark A. Spackman (Perth, Australia) for his substantial help with the newest version of the CRYSTALEXPLORER program.

30 ACS Paragon Plus Environment

Page 30 of 37

Page 31 of 37

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

References 1.

Desiraju, G. R., Supramolecular synthons in crystal engineering - a new organic synthesis. Angew. Chem. Int. Ed. 1995, 34, 2311-2327.

2.

Basavoju, S.; Boström, D.; Velaga, S. P., Pharmaceutical cocrystal and salts of norfloxacin. Cryst. Growth Des. 2006, 6, 2699-2708.

3.

Schultheiss, N.; Newman, A., Pharmaceutical cocrystals and their physicochemical properties. Cryst. Growth Des. 2009, 9, 2950-2967.

4.

Burrows, A. D., Crystal engineering using multiple hydrogen bonds. In Supramolecular Assembly via Hydrogen Bonds I, Mingos, D. M. P., Ed. Springer-Verlag Berlin Heidelberg: 2004.

5.

Thakuria, R.; Delori, A.; Jones, W.; Lipert, M. P.; Roy, L.; Rodríguez-Hornedo, N., Pharmaceutical cocrystals and poorly soluble drugs. Int. J. Pharm. 2013, 453, 101-125.

6.

Eddleston, M. D.; Jones, W., Formation of tubular crystals of pharmaceutical compounds. Cryst. Growth Des. 2010, 10, 365-370.

7.

Madusanka, N.; Eddleston, M. D.; Arhangelskis, M.; Jones, W., Polymorphs, hydrates and solvates of a co-crystal of caffeine with anthranilic acid. Acta Cryst. Sect. B 2014, 70, 72-80.

8.

Desiraju, G. R., Cryptic crystallography. Nat. Mater. 2002, 1, 77-79.

9.

Desiraju, G. R., Chemistry beyond the molecule. Nature 2001, 412, 397-400.

10. Desiraju, G. R., Designer crystals: intermolecular interactions, network structures and supramolecular synthons. Chem. Commun. 1997, 1475-1482. 11. Braga, D., Crystal engineering, where from? Where to? Chem. Commun. 2003, 2751-2754. 12. Eddleston, M. D.; Sivachelvam, S.; Jones, W., Screening for polymorphs of cocrystals: a case study. CrystEngComm 2013, 15, 175-181. 13. Batchelor, E.; Klinowski, J.; Jones, W., Crystal engineering using co-crystallisation of phenazine with dicarboxylic acids. J. Mater. Chem. 2000, 10, 839-848. 14. Soldatov, D. V.; Terekhova, I. S., Supramolecular chemistry and crystal engineering. J. Struct. Chem. 2005, 46, S1-S8. 15. Aoyama, Y.; Endo, K.; Anzai, T.; Yamaguchi, Y.; Sawaki, T.; Kobayashi, K.; Kanehisa, N.; Hashimoto, H.; Kai, Y.; Masuda, H., Crystal engineering of stacked aromatic columns. three-dimensional control of the alignment of orthogonal aromatic triads and guest quinones via self-assembly of hydrogenbonded networks. J. Am. Chem. Soc. 1996, 118, 5562-5571. 16. Etter, M. C., Encoding and decoding hydrogen-bond patterns of organic compounds. Acc. Chem. Res. 1990, 23, 120-126. 17. Nangia, A.; Desiraju, G. R., Supramolecular structures - reason and imagination. Acta Cryst. Sect. A 1998, 54, 934-944. 18. Nangia, A.; Desiraju, G. R., Supramolecular synthons and pattern recognition. In Design of Organic Solids, Weber, E., Ed. Springer-Verlag Berlin Heidelberg: 1998; pp 57-95. 19. Lemmerer,

A.;

Michael,

J.

P.,

Hydrogen

bonding

patterns

arylcycloalkanecarboxamides. CrystEngComm 2008, 10, 95-102.

31 ACS Paragon Plus Environment

in

a

series

of

1-

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

20. Desiraju, G. R., The C-H···O hydrogen bond:  structural implications and supramolecular design. Acc. Chem. Res. 1996, 29, 441-449. 21. Carlucci, L.; Gavezzotti, A., Molecular recognition and crystal energy landscapes: an X-ray and computational study of caffeine and other methylxanthines. Chem. Eur. J. 2005, 11, 271-279. 22. Bernstein, J.; Davis, R. E.; Shimoni, L.; Chang, N. L., Patterns in hydrogen bonding: functionality and graph set analysis in crystals. Angew. Chem. Int. Ed. 1995, 34, 1555-1573. 23. Biradha, K., Crystal engineering: from weak hydrogen bonds to co-ordination bonds. CrystEngComm 2003, 5, 374-384. 24. Bis, J. A.; Vishweshwar, P.; Weyna, D.; Zaworotko, M. J., Hierarchy of supramolecular synthons:  persistent hydroxyl···pyridine hydrogen bonds in cocrystals that contain a cyano acceptor. Mol. Pharm. 2007, 4, 401-416. 25. Chen, D.; Oezguen, N.; Urvil, P.; Ferguson, C.; Dann, S. M.; Savidge, T. C., Regulation of protein-ligand binding affinity by hydrogen bond pairing. Sci. Adv. 2016, 2, e1501240. 26. Varughese, S.; Hoser, A. A.; Jarzembska, K. N.; Pedireddi, V. R.; Woźniak, K., Positional isomerism and conformational flexibility directed structural variations in the molecular complexes of dihydroxybenzoic acids. Cryst. Growth Des. 2015, 15, 3832-3841. 27. Zhang, J.; Wu, L. X.; Fan, Y. G., Heterosynthons in molecular complexes of azopyridine and 1,2-bis(4pyridyl)ethylene with dicarboxylic acids. J. Mol. Struct. 2003, 660, 119-129. 28. Vishweshwar, P.; Nangia, A.; Lynch, V. M., Recurrence of carboxylic acid-pyridine supramolecular synthon in the crystal structures of some pyrazinecarboxylic acids. J. Org. Chem. 2002, 67, 556-565. 29. Shattock, T. R.; Arora, K. K.; Vishweshwar, P.; Zaworotko, M. J., Hierarchy of supramolecular synthons: persistent carboxylic acid···pyridine hydrogen bonds in cocrystals that also contain a hydroxyl moiety. Cryst. Growth Des. 2008, 8, 4533-4545. 30. Sarma, B.; Sanphui, P.; Nangia, A., Polymorphism in isomeric dihydroxybenzoic acids. Cryst. Growth Des. 2010, 10, 2388-2399. 31. Parkin, A.; Adam, M.; Cooper, R. I.; Middlemiss, D. S.; Wilson, C. C., Structure and hydrogen bonding in 2,4-dihydroxybenzoic acid at 90, 100, 110 and 150 K; a theoretical and single-crystal X-ray diffraction study. Acta Cryst. Sect. B 2007, 63, 303. 32. Gdaniec, M.; Gilski, M.; Denisov, G. S., γ-Resorcylic acid, its monohydrate and its pyridinium complex. Acta Cryst. Sect. C 1994, 50, 1622. 33. MacGillivray, L. R.; Zaworotko, M. J., Crystal and molecular structure of 2,6-dihydroxybenzoic acid. J. Chem. Cryst. 1994, 24, 703. 34. Okabe, N.; Kyoyama, H., 2,3-Dihydroxybenzoic acid. Acta Cryst. Sect. E 2001, 57, o1224. 35. Haisa, M.; Kashino, S.; Hanada, S.-I.; Tanaka, K.; Okazaki, S.; Shibagaki, M., The structures of 2hydroxy-5-methylbenzoic acid and dimorphs of 2,5-dihydroxybenzoic acid. Acta Cryst. Sect. B 1982, 38, 1480. 36. Cohen, D. E.; Benedict, J. B.; Morlan, B.; Chiu, D. T.; Kahr, B., Dyeing polymorphs:  the MALDI host 2,5dihydroxybenzoic acid. Cryst. Growth Des. 2007, 7, 492-495. 37. Perriot, J.; Chambonnet, E.; Eschalier, A., Managing the adverse events of antitubercular agents. Rev. Mal. Respir. 2011, 28, 542-555.

32 ACS Paragon Plus Environment

Page 32 of 37

Page 33 of 37

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

38. Tarshis, M. S.; Weed, W. A., Lack of significant in vitro sensitivity of Mycobacterium tuberculosis to pyrazinamide on three different solid media. Am. Rev. Tuberc. 1953, 67, 391-395. 39. McDermott, W.; Tompsett, R., Activation of pyrazinamide and nicotinamide in acidic environments in vitro. Am. Rev. Tuberc. 1954, 70, 748-754. 40. Yeager, R. L.; Munroe, W. G. C.; Dessau, F. I., Pyrazinamide (aldinamide*) in the treatment of pulmonary tuberculosis. Trans. Annu. Meet. Natl. Tuberc. Assoc. 1952, 48, 178-201. 41. Cardoso, S. H.; Barreto, M. B.; Lourenco, M. C. S.; Henriques, M.; P.Candea, A. L.; Kaiser, C. R.; de Souza, M. V. N., Antitubercular activity of new coumarins. Chem. Biol. Drug. Des. 2011, 77, 489-493. 42. Zhang, Y.; Mitchison, D., The curious characteristics of pyrazinamide: a review. Int. J. Tuberc. Lung Dis. 2003, I 7, 6-21. 43. Tamura, C.; Sasada, Y.; Kuwano, H., Crystallographic data of carboxylic acids and carboxyamides of picoline and pyrazine derivatives. Acta Cryst. 1961, 14, 693. 44. Rø, G.; Sørum, H., The crystal and molecular structure of β-pyrazincarboxamide. Acta Cryst. Sect. B 1972, 28, 991-998. 45. Rø, G.; Sørum, H., The crystal and molecular structure of δ-pyrazincarboxamide. Acta Cryst. Sect. B 1972, 28, 1677-1684. 46. Cherukuvada, S.; Thakuria, R.; Nangia, A., Pyrazinamide polymorphs: relative stability and vibrational spectroscopy. Cryst. Growth Des. 2010, 10, 3931-3941. 47. Takaki, Y.; Sasada, Y.; Watanabe, T., The crystal structure of α-pyrazinamide. Acta Cryst. 1960, 13, 693-702. 48. Miwa, Y.; Mizuno, T.; Tsuchida, K.; Tagaa, T.; Iwata, Y., Experimental charge density and electrostatic potential in nicotinamide. Acta Cryst. Sect. B 1999, 55, 78-84. 49. Báthori, N. B.; Lemmerer, A.; Venter, G. A.; Bourne, S. A.; Caira, M. R., Pharmaceutical co-crystals with isonicotinamide - vitamin B3, clofibric acid, and diclofenac - and two isonicotinamide hydrates. Cryst. Growth Des. 2011, 11, 75-87. 50. Lemmerer, A.; C, E.; Bernstein, J., Synthesis, characterization, and molecular modeling of a pharmaceutical co-crystal: (2-chloro-4-nitrobenzoic acid):(nicotinamide). J. Pharm. Sci. 2010, 99, 4054-4071. 51. Berry, D. J.; Seaton, C. C.; Clegg, W.; Harrington, R. W.; Coles, S. J.; Horton, P. N.; Hursthouse, M. B.; Storey, R.; Jones, W.; Friščić, T.; Blagden, N., Applying hot-stage microscopy to co-crystal screening: a study of nicotinamide with seven active pharmaceutical ingredients. Cryst. Growth Des. 2008, 8, 1697-1717. 52. Allen, F. H., The Cambridge Structural Database: a quarter of a million crystal structures and rising. Acta Cryst. Sect. B 2002, 58, 380-388. 53. Jarzembska, K. N.; Hoser, A. A.; Kamiński, R.; Madsen, A. Ø.; Durka, K.; Woźniak, K., Combined experimental and computational studies of pyrazinamide and nicotinamide in the context of crystal engineering and thermodynamics. Cryst. Growth Des. 2014, 14, 3453-3465. 54. McMahon, J. A.; Bis, J. A.; Vishweshwar, P.; Shattock, T. R.; McLaughlin, O. L.; Zaworotko, M. J., Crystal engineering of the composition of pharmaceutical phases. 3. Primary amide supramolecular

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 37

heterosynthons and their role in the design of pharmaceutical co-crystals. Z. Kristallogr. 2005, 220, 340-350. 55. Stilinovic, V.; Kaitner, B., Salts and co-crystals of gentisic acid with pyridine derivatives: the effect of proton transfer on the crystal packing (and vice versa). Cryst. Growth Des. 2012, 12, 5763-5772. 56. Solomon, K. A.; Blacque, O.; Venkatnarayan, R., Molecular salts of 2,6-dihydroxybenzoic acid (2,6DHB) with N-heterocycles: crystal structures, spectral properties and Hirshfeld surface analysis. J. Mol. Struct. 2017, 1134, 190-198. 57. Lemmerer, A.; Adsmond, D. A.; Esterhuysen, C.; Bernstein, J., Polymorphic co-crystals from polymorphic

co-crystal

formers:

competition

between

carboxylic

acid···pyridine

and

phenol···pyridine hydrogen bonds. Cryst. Growth Des. 2013, 13, 3935. 58. Dovesi, R.; Orlando, R.; Civalleri, B.; Roetti, C.; Saunders, V. R.; Zicovich-Wilson, C. M., CRYSTAL: a computational tool for the ab initio study of the electronic properties of crystals. Z. Kristallogr. 2005, 220, 571-573. 59. Dovesi, R.; Saunders, V. R.; Roetti, C.; Orlando, R.; Zicovich-Wilson, C. M.; Pascale, F.; Civalleri, B.; Doll, K.; Harrison, N. M.; Bush, I. J.; D’Arco, P.; Llunell, M. CRYSTAL09, University of Torino, Torino, Italy, 2009. 60. Gavezzotti, A., Calculation of intermolecular interaction energies by direct numerical integration over electron densities. 2. An improved polarization model and the evaluation of dispersion and repulsion energies. J. Phys. Chem. B 2003, 107, 2344-2353. 61. Gavezzotti, A., Non-conventional bonding between organic molecules. The 'halogen bond' in crystalline systems. Mol. Phys. 2008, 106, 1473-1485. 62. Turner, M. J.; Thomas, S. P.; Shi, M. W.; Jayatilaka, D.; Spackman, M. A., Energy frameworks: insights into interaction anisotropy and the mechanical properties of molecular crystals. Chem. Commun. 2015, 51, 3735-3738 63. Rigaku Oxford Diffraction CRYSALIS PRO, Yarnton, Oxfordshire, England, UK, 2017. 64. Sheldrick, G. M., A short history of SHELX. Acta Cryst. Sect. A 2008, 64, 112-122. 65. Turner, M. J.; McKinnon, J. J.; Wolff, S. K.; Grimwood, D. J.; Spackman, P. R.; Jayatilaka, D.; Spackman, M. A. CRYSTALEXPLORER17, University of Western Australia, Crawley, Western Australia, Australia, 2017. 66. Becke, A. D., Density-functional exchange-energy approximation with correct asymptotic behavior. Phys. Rev. A 1988, 38, 3098-3100. 67. Becke, A. D., Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648-5652. 68. Lee, C.; Yang, W.; Parr, R. G., Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785-789. 69. Krishnan, R.; Binkley, J. S.; Seeger, R.; Pople, J. A., Self-consistent molecular orbital methods. XX. A basis set for correlated wave functions. J. Chem. Phys. 1980, 72, 650-654. 70. Civalleri, B.; Zicovich-Wilson, C. M.; Valenzano, L.; Ugliengo, P., B3LYP augmented with an empirical dispersion term (B3LYP-D*) as applied to molecular crystals. CrystEngComm 2008, 10, 405-410.

34 ACS Paragon Plus Environment

Page 35 of 37

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

71. Grimme, S., Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787-1799. 72. Grimme, S., Accurate description of van der Waals complexes by density functional theory including empirical corrections. J. Comput. Chem. 2004, 25, 1463-1473. 73. 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, 553-566. 74. Mazur, L.; Jarzembska, K. N.; Kamiński, R.; Hoser, A. A.; Madsen, A. Ø.; Pindelska, E.; Zielińska-Pisklak, M., Crystal structures and thermodynamic properties of polymorphs and hydrates of selected 2pyridinecarboxaldehyde hydrazones. Cryst. Growth Des. 2016, 16, 3101-3112. 75. Jarzembska, K. N.; Kamiński, R.; Wenger, E.; Lecomte, C.; Dominiak, P. M., Interplay between charge density distribution, crystal structure energetic features, and crystal morphology of 6-methyl-2thiouracil. J. Phys. Chem. C 2013, 117, 7764-7775. 76. Jarzembska, K. N.; Kubsik, M.; Kamiń ski, R.; Woź niak, K.; Dominiak, P. M., From a single molecule to molecular crystal architectures: structural and energetic studies of selected uracil derivatives. Cryst. Growth Des. 2012, 12, 2508-2524. 77. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegaw, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; J. A. Montgomery, J.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, J. M.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. GAUSSIAN09, Gaussian Inc.: Wallingford, Connecticut, United States, 2009. 78. Dunning, T. H., Gaussian basis sets for use in correlated molecular calculations. I. The atoms boron through neon and hydrogen. J. Chem. Phys. 1989, 90, 1007-1023. 79. Kendall, R. A.; Dunning, T. H.; Harrison, R. J., Electron affinities of the first-row atoms revisited. Systematic basis sets and wave functions. J. Chem. Phys. 1992, 96, 6796-6806. 80. Grimme, S.; Ehrlich, S.; Goerigk, L., Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456-1465. 81. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H., A consistent and accurate ab initio parameterization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. 82. Simon, S.; Duran, M.; Dannenberg, J. J., How does basis set superposition error change the potential surfaces for hydrogen bonded dimers? J. Chem. Phys. 1996, 105, 11024-11031. 83. Singh, U. C.; Kollman, P. A., An approach to computing electrostatic charges for molecules. J. Comput. Chem. 1984, 5, 129-145.

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

84. Besler, B. H.; Merz Jr., K. M.; Kollman, P. A., Atomic charges derived from semiempirical methods. J. Comput. Chem. 1990, 11, 431-439. 85. Kamiński, R.; Jarzembska, K. N.; Domagała, S., CLUSTERGEN: a program for molecular cluster generation from crystallographic data. J. Appl. Cryst. 2013, 46, 540-534. 86. Møller, C.; Plesset, M. S., Note on an approximation treatment for many-electron systems. Phys. Rev. 1934, 46, , 618-622. 87. Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H., OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Cryst. 2009, 42, 339-341. 88. Childs, S. L.; Stahly, G. P.; Park, A., The salt-cocrystal continuum: the influence of crystal structure on ionization state. Mol. Pharm. 2007, 4, 323-338. 89. Johnson, S. L.; Rumon, K. A., Infrared spectra of solid 1:1 pyridine-benzoic scid vomplexes; the nature of the hydrogen bond as a function of the scid-base levels in the complex. J. Phys. Chem. 1965, 69, 7486.

36 ACS Paragon Plus Environment

Page 36 of 37

Page 37 of 37

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 Title: Structural and energetic analysis of molecular assemblies in the series of nicotinamide and pyrazinamide cocrystals with dihydroxybenzoic acids

Authors: Katarzyna N. Jarzembska, Anna A. Hoser, Sunil Varughese, Radosław Kamiński, Maura Malinska, Marcin Stachowicz, Venkatesvara R. Pedireddi, Krzysztof Woźniak

TOC Graphic:

Synopsis: Four new cocrystals of pharmaceutically active N-donor compounds, pyrazinamide and nicotinamide, with a series of dihydroxybenzoic acids were synthesised and structurally characterised in the context of analogous literaturereported systems and the respective single-component crystals. Main structural motifs and relative crystal stabilities were analysed computationally.

37 ACS Paragon Plus Environment