Predicting Supramolecular Connectivity of Metal-Containing Solid

3 days ago - Predicting Supramolecular Connectivity of Metal-Containing Solid-State Assemblies using Calculated Molecular Electrostatic Potential ...
10 downloads 0 Views 5MB Size
Article Cite This: Cryst. Growth Des. XXXX, XXX, XXX−XXX

pubs.acs.org/crystal

Predicting Supramolecular Connectivity of Metal-Containing SolidState Assemblies using Calculated Molecular Electrostatic Potential Surfaces Mladen Borovina, Ivan Kodrin, and Marijana Đaković* Department of Chemistry, Faculty of Science, University of Zagreb, Horvatovac 102a, HR-10000 Zagreb, Croatia

Downloaded via WEBSTER UNIV on February 21, 2019 at 09:01:05 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

S Supporting Information *

ABSTRACT: We demonstrated that the primary supramolecular features of metal-based solid-state systems can be reliably predicted based solely on the relative strengths of competing hydrogen-bond acceptor sites. The predictive protocol utilizes a simple electrostatic view of the hydrogen bond and ranks the multiple acceptor sites according to calculated molecular electrostatic potential (MEP) surface values. The MEP was calculated for competing acceptors on 12 zero-dimensional (0-D) 2,4-pentanedionate (acac)-based complexes (Ni(II), Co(II), Cu(II)), equipped with the lactam moiety, and the structural outcome was successfully predicted in 10 of 12 compounds by comparing the MEP difference between two acceptors, namely, the lactam and acac-based oxygen atoms. The two acceptor sites displayed structural selectivity as long as there was a substantial difference (ΔE > ΔEcutoff) between their relative hydrogen-bond acceptor capabilities. In the remaining two cases, the expected coordination geometry around the metal center did not materialize, which meant that a prediction of the supramolecular details could not be done. The working cutoff value (ΔEcutoff ≈ 30 kJ/mol) proved to be a valid and decisive criterion for predicting the supramolecular connectivity in these 0-D systems. The results further indicate that the ΔEcutoff is likely to be primarily dependent on the supramolecular functionality itself rather than on external “packing forces”.



the structural outcome is often unexpected or undesired.22,23 The reason is that metal-based building units regularly (i) exhibit a greater degree of structural complexity compared to many metal-free systems and (ii) present multiple acceptor sites. In addition to a set of (hydrogen-/halogen-bond) donors and acceptors already present at the supramolecular functionality, new acceptor sites are involved into the system as chargebalancing entities or a part thereof. They readily compete for the same donors, giving rise to “synthon crossover”24 instead of intended “transferability” of synthons form the organic solid state. Therefore, it is essential to improve our understanding of the fine balance between the intermolecular interactions in the systems with multiple hydrogen-bond donor/acceptor sites to successfully control supramolecular synthesis and assembly of metal-based building blocks. For organic solids containing a number of different competing donor and/or acceptor sites, “ranking” hydrogen/ halogen-bond donors and acceptors using calculated molecular electrostatic potential (MEP) values proved to be successful for rationalization of observed interactions.25−30

INTRODUCTION Crystal engineering1−4 has delivered many efficient synthetic strategies based upon noncovalent interactions for the assembly of desired architectures within the organic solidstate arena.5−7 Systematic experimental and theoretical studies of the structural landscape and binding preferences of key functional groups have been refined to such an extent that numerous synthons capable of producing robust avenues for supramolecular synthesis have been identified.8−12 However, there is a relative shortage of reliable strategies for the crystal engineering of specific constructions and motifs built around coordination complexes and other metal ion-containing building blocks. This is particularly true for the synthesis of assemblies of lower dimensionality. Given the fact that successful “inorganic” crystal engineering13,14 can ultimately provide access to a range of tunable properties (e.g., catalytic,15 optical,16 luminescent,17,18 magnetic,19 or photochemical20,21), it is important to develop additional tools and strategies for directed assembly of metal complexes using noncovalent interactions as synthetic vectors. It has been shown that, when adapting approaches that work very well in organic crystal engineering to coordination chemistry (i.e., equipping metal complexes with bifunctional ligands that can simultaneously coordinate to metal cations and engage in self-complementary noncovalent interactions), © XXXX American Chemical Society

Received: December 29, 2018 Revised: January 31, 2019

A

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

In contrast, there is only a limited number of reports using calculated MEP values for rationalizing key structural features in coordination compounds.31,32,32−34 Recently, we presented that self-complementary synthons can be transferred from organic to one-dimensional (1-D) metal−organic system and rationalized the supramolecular outcome on the basis of calculated MEP values. Our results showed that assembling 1D metal-based crystalline solids via the same supramolecular links (and displaying the same supramolecular motifs) as in organic solids is possible as long as the metal cations and their charge-balancing anions do not disrupt the relative importance and ranking of different hydrogen-bond donors and acceptors.33 Also, for crystalline zero-dimensional (0-D) solids we showed that, by manipulating the electrostatic effects, the reactants could be adjusted so that the difference in the MEP values of the competing acceptor sites allows formation of the targeted supramolecular outcome.34 In the current study, we want to establish if the same, relatively simple, electrostatic view of hydrogen bonding can be translated into practical “inorganic” synthetic strategies. We want to ascertain if the supramolecular synthetic outcome of metal-based solid-state systems comprising multiple acceptor sites can be reliably predicted based solely on the relative strength of the competing acceptor sites. If this is the case, we also intend to map more carefully what the synthetic boundaries are to deliver a specific supramolecular product. To address these challenges, we opted for 0-D metalcontaining building units equipped with the same selfcomplementary supramolecular functionality that proved to be successful for assembling 1-D metal complexes and to retain the octahedral geometry around the metal cation.33 To construct our 0-D metal complexes we use three different constituents that allow for tuning the electrostatic effects around the competing acceptor sites either through inductive or resonance effects of the two sets of ligands or via introduction of different metal ions in the core of the building unit.35 Here we employ (a) a set of divalent metal cations (Co(II), Ni(II), and Cu(II)), (b) acac-based ligands to compensate for charge (2,4-pentanedione (acac) and 1,1,1,5,5,5,-hexafluoro-2,4-pentanedione (hfac)), and (c) a small heterocyclic organic ligands bearing the lactam functionality (4-pyrimidinone (4-pym) and 4-quinazolinone (4-quz)), Scheme 1. Combining the three constituents will ideally allow us to deliver 12 (3 metal cations × 2 acac-based ligands × 2 lactam bearing ligands) complex 0-D building units with a trans-orientation of the ligands, all presenting only one (good) hydrogen-bond donor and two competing acceptor sites, that are intended (by changing only one parameter at the time) to display a range of hydrogen-bond donating and accepting abilities, Figure 1. By combining Etter’s best donor−best acceptor rule36 and MEP-based guidelines derived for purely organic solids,37 we hypothesize that three outcomes are possible for our metal− organic systems: (i) if the acceptor residing at the lactam moiety is a better acceptor (i.e., the carbonyl oxygen atom) and the difference in the MEP values between the two acceptor sites is substantial (A1 > A2), we will be able to assemble our metal-based building blocks via the same supramolecular links as in the metal-free systems (R22(8) and C(4)); (ii) if the acceptor at the acac-based ligand is a substantially better acceptor (A1 < A2), a supramolecular link involving the oxygen atom originated from the acac ligand will be an exclusive

Scheme 1. Targeted Octahedral 0-D Metal-Containing Building Units with trans-Orientation of the Ligandsa and Supramolecular Motifsb Observed for the Lactam Functionality in Organic Solids

a

M = Co(II), Ni(II), Cu(II); chelating ligands: 2,4-pantanedionate (acac) and 1,1,1,5,5,5,-hexafluoro-2,4-pentanedionate (hfac); L = 4(3H)-pyrimidinone (4-pym), 4(3H)-quinazolinone (4-quz). b 2 R 2(8) and C(4).

Figure 1. Metal-containing 0-D building unit, equipped with the lactam functionalities as the desired supramolecular “drivers”, displaying two different acceptor sites (A1 and A2) and only one donor (D).

supramolecular outcome (N−H···Oacac); (iii) if the MEP difference between the two acceptors is relatively small (A1 ≈ A2) (smaller than a certain cutoff value, either in favor of the lactam or acac oxygen atom), there might be a diffuse region with no statistically significant trend to be identified (“no selectivity”), Figure 2. To test our hypothesis and to examine if it is possible to predict the supramolecular synthetic outcome of metalcontaining systems, we calculated MEP surface values for the hydrogen-bond donors and acceptors and performed the structural work.



RESULTS AND DISCUSSION Computational Study. Twelve metal complexes used in this study contain two acceptor types (two oxygen atoms Oacac and Olac residing on the acac-based and lactam ligands, respectively) and only one good hydrogen-bond donor site (the lactam hydrogen atom). The ranking of the two acceptor sites is based upon calculated MEP surface values, and the B

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

metal centers (Co(II) for Ni(II), Cu(II)). A detailed description of the procedure used to generate all the starting geometries is given in the Supporting Information. All optimizations were performed in Gaussian16 with B3LYP in combination with def2-TZVP basis set and D3 Grimme’s dispersion corrections. Analysis of the calculated MEP values confirmed that the lactam hydrogen atom and the two oxygen atoms (Oacac and Olac) are indeed the sites with the most positive and the most negative MEP values, respectively, in all 12 instances. The MEP values for the donor and acceptor sites are provided in Figure 3 (and Table S4). The listed MEP values show that the oxygen atom residing at the lactam moiety (Olac) is presenting a lower MEP value than the acac-based oxygen atom (Oacac), thus being categorized as the best hydrogen-bond acceptor in all 12 instances. Having Olac as the best acceptor in all cases, unfortunately, does not allow us to test the third hypothesized outcome (A1 < A2, Figure 2), that is, to examine if linking the metal complexes via N−H···Oacac is the exclusive outcome for the MEP(Olac) < MEP(Oacac) case. Therefore, we can rule out the third hypothesis and from now on focus only on the first two hypothesized outcomes (A1 > A2 and A1 ≈ A2, Figure 2). To better understand possible hydrogen-bond selectivity, we introduce a ΔE value as the molecular electrostatic potential difference between the two acceptor sites, ΔE = |MEP(Olac) − MEP(Oacac)|, similarly to what was previously proposed for purely organic halogen-bonded systems39 and proved efficient for rationalizing connectivity in metal−organic supramolecular assemblies.32 Also, to differentiate between the two outcomes (A1 > A2 and A1 ≈ A2), based on our previous findings for the lactam functionality, we are using ∼30 kJ/mol as a working cutoff value (ΔEcutoff).33 In all the studied 1-D metal complexes equipped with the lactam moieties, the lactam oxygen atom was a better acceptor than the bridging halide anion (i.e., charge-balancing entity), with the smallest MEP difference between the two acceptors ΔE = 33 kJ/mol, and in all the instances building units were assembled via the supramolecular link transferred from the organic systems, that is, N−H···Olac. Therefore, to establish if we can predict the supramolecular outcome we decided to take ΔEcuf off of ∼30 kJ/mol as a value against which to test our hypothesis. We classify our 12 0-D metal complexes into two groups, having ΔE > ΔEcuf off and ΔE < ΔEcuf off, Figure 4. For the former we expect the transferability of the supramolecular motifs (dimers R22(8) or catemers C(4)) from the organic solid-state systems, and “no selectivity” (or no statistically significant trend to be observed) for the latter one. To test the validity of our prediction of the supramolecular synthetic outcome, we performed synthetic and structural work. Structural Considerations. To produce intended metal complexes (Figure 3), we allowed starting acac-based complexes (1−6) to react with 4(3H)-pyrimidone (4-pym) and 4(3H)-quinazolinone (4-quz), and 10 of 12 intended compounds were formed, Scheme 2. To avoid any solvent−solute bias throughout the experiments,40 all the crystallizations were conducted by a slow evaporation from the same solvent or solvent mixture. Eight of ten X-ray quality crystals were obtained by slow evaporation from the ethanol, while a recrystallization (of all compounds) from a 1:1 ethanol−chloroform mixture yielded one more crystalline sample (1a) but of a different polymorphic form (1a-II). Regardless of any solvent used, 1a remained the

Figure 2. Possible outcomes of the assembly process of metalcontaining solids, i.e., after introduction of metal cations and charge compensating entities into purely organic environment.

better hydrogen-bond acceptor ability is ascribed to acceptor displaying the more negative MEP value. A Cambridge Crystallographic Database (CSD)38 survey (ConQuest, Version 1.23) was performed to determine preferred coordination geometries and cis/trans propensities of the intended bis-β-diketonato complexes and to extract starting geometries for density functional theory (DFT) calculations. The survey (details provided in Supporting Information) revealed that, in general, bis-β-diketonato complexes of Co(II), Ni(II), and Cu(II) have a strong tendency to display octahedral coordination (that being the least pronounced for Cu(II); only ∼64% of Cu(II) bis-βdiketonato complexes exhibit octahedral coordination, while it is ∼98% and ∼92% for Co(II) and Ni(II), respectively). If the survey is restricted to only the class of complexes related to this study, that is, comprising at least one N-heterocyclic ligand ([M(L1)2(L2)] or [M(L1)2(L2)2]; M: Co(II), Ni(II), Cu(II); L1: β-diketonato ligands; L2: pyridine, pyrazine, pyrimidine, quinoline, quinazoline, and quinoxaline-based ligands), the octahedral geometry becomes even more pronounced, with trans-orientation of the ligands being more than 93%, Table 1. Table 1. CSD Survey of cis/trans Preferences of Octahedral Bis-β-diketonato Complexes of Co(II), Ni(II), and Cu(II) with Small N-Heterocyclic Ligands [M(L1)2(L2)]a

[M(L1)2(L2)2]a cis-

Co(II)

1

Ni(II)

0

Cu(II)

26

trans54

2/54

4%

52/54

96%

67/68

98%

52/56

93%

68 1/68

2% 56

4/56

7%

L1: β-diketonato ligands; L2: pyridine, pyrazine, pyrimidine, quinoline, quinazoline, and quinoxaline-based ligands (for details see Supporting Information).

a

Therefore, to calculate the MEP surface values for the 12 metal complexes, we opted for starting geometries with a trans arrangement of the ligands. Molecular geometries of 1a−6b were generated starting from the diaqua complex [Co(acac)2(H2O)2] (1) extracted from the CSD (CODAAC03), by optimizing its molecular geometry and substituting axial H2O molecules for the lactam ligands (4-pym, 4-quz), terminal −CH3 groups for −CF3 (acac → hfac), and/or C

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 3. Calculated MEP values (mapped on the 0.002 au isodensity surface) at the best donor and two acceptor sites for each of the 12 potential metal-based building units. The corresponding ΔE values for the two acceptor sites are given (in kJ/mol); ΔE = |MEP(Olac) − MEP(Oacac)|. In cases where the atoms of the same acceptor type (i.e., two Olac or four Oacac) are residing at conformationally different surroundings, the lowest MEP values are listed.

isolated case of the solvent-dependent polymorphism of the studied class. Structure determination for all 11 cases confirmed the expected 0-D metal complexes as primary building units, with metal centers octahedrally coordinated to two sets of transoriented ligands, Figure 5. The Cu(II) building units differ only slightly from their Co(II) and Ni(II) analogues due to the Jahn−Teller distortion of the coordination sphere; for Co(II) and Ni(II) the acac-based ligands occupy the equatorial plane, and two lactam-type ligands reside in the axial positions, while

for Cu(II) the lactam-type ligands and one hfac oxygen atom from each hfac ligand define the equatorial plane, and the other hfac oxygen atoms occupy the axial positions, Figures S1−S6. Crystallographic data, a full summary of the intensity data collection and structure refinement as well as selected bond distances and angles for all the complexes, are reported in Tables S1−S3 in the Supporting Information, while the hydrogen-bond geometries are listed in Table 2. For the complex [Co(acac)2(4-pym)2] (1a) (ΔE = 15 kJ/ mol), two polymorphic forms were isolated, form I (1a-I) and D

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 4. Expected outcome of the supramolecular synthetic efforts. Transferability of the lactam synthons (R22(8) and C(4)) from the organic to metal-containing system is expected in 8/12, while “noselectivity” is expected in 4/12 instances.

Scheme 2. Preparation of Intended Metal Complexes (1a− 6b)a

Figure 5. Metal-containing building units of Ni(II) (ORTEP-style plots) equipped with 4-pyrimidinone, 2a and 5a (left), and 4quinazolinone ligands, 2b and 5b (right). The Co(II)- and Cu(II)containing building units are essentially the same as those of Ni(II).

a

The X-ray quality crystals were obtained by slow evaporation from ethanol solution (except for 1a-I and 4a (*), where no X-ray quality crystals were obtained). The recrystallization of 1a-I from 1:1 ethanol−chloroform mixture yielded X-ray quality crystals but of a different polymorphic form 1a-II (**).

11 kJ/mol) revealed substantially different supramolecular connectivity from that observed in 1a-II. The crystal structure of 2a comprises two symmetrically independent molecules, both displaying identical supramolecular connectivity. The adjacent symmetrically independent metal-containing building blocks are linked via two single-point N−H···Oacac hydrogen bonds between the lactam moieties and the acac oxygen atoms, forming R22(12) motifs, which in turn produced infinite 1-D supramolecular chains (Figure 8, top). The carbonyl oxygen atoms, as not being involved in formation of one of the lactam motifs, are engaged into single-point C−H···O hydrogen bonds resulting in catemeric C(4) motifs, resembling the one that lactam moiety forms in pure organic settings, (Figure 8, bottom). Overall, it gives rise to 2-D layers. Unfortunately, all of our synthetic attempts (solution synthesis, as well as the mechanochemical and solvothermal synthetic procedures) failed to produce the complex 3a (ΔE = 13 kJ/mol), Figure 3. Regardless of the conditions used, reactions of starting complex [Cu(acac)2] (3) with the 4pyrimidinone did not result in coordination of the 4-pym ligand but instead gave the mixture of starting compounds (3 and 4-pym). Exchanging acac for hfac ligand and reacting the starting hfac complexes (4−6) with the 4-pym yielded two new crystal structures, [Ni(hfac)2(4-pym)2] (5a) and [Cu(hfac)2(4pym)2] (6a) with the difference in the MEP values between the two acceptor sites of 77 and 64 kJ/mol, respectively. In both structures only the best acceptor was engaged, and the lactam catemeric C(4) and head-to-head R22(8) motifs were

form II (1a-II). Polycrystalline form I (1a-I) was harvested from the ethanol solution (through a slow evaporation), while the single crystals of form II (1a-II) were obtained via recrystallization of 1a-I from a 1:1 ethanol−chloroform mixture. These observations clearly demonstrate the role of solvent polarity41 in controlling the intermolecular interactions and determining the supramolecular synthetic outcome.38 Although it has so far remained the isolated observation of solvent-driven polymorphic control, it is worth mentioning that it is observed for one of the complexes with the smallest ΔE values. In the crystal structure of 1a-II the nearest neighboring building units are linked via two N−H···O hydrogen bonds between adjacent pyrimidinone groups (Table 2), resulting in R22(8) motifs (Figure 6). In addition, the pyrimidinone hydrogen atom attached to the endocyclic nitrogen atom adjacent to the carbonyl group is also engaged in an interaction to the acac oxygen atom N−H···Oacac,35 thus acting as a bifurcated donor and linking primary 1-D supramolecular chains (formed by R22(8) rings) into a two-dimensional (2-D) architecture. Although the crystal structure of the form 1a-I could not be obtained, a comparison of the powder X-ray diffraction (PXRD) traces of 1a-I and 2a (Figure 7) strongly suggests that they are in fact isostructural. Moving from 1a to 2a and exchanging only the metal ion in the core of the building unit ([Ni(acac)2(4-pym)2] (2a), ΔE = E

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Similarly to 1a-I, the crystal structure of 4a could not be obtained, but a comparison of the powder X-ray diffraction (PXRD) traces of 4a and 5a again strongly suggests that they are isostructural (Figure S13). Getting back to the starting acac complexes (1−3) and reacting them with the 4-quz ligand yielded two new crystal structures, [Co(acac)2(4-pym)2] (1b) (ΔE = 46 kJ/mol) and [Ni(acac)2(4-pym)2] (2b) (ΔE = 49 kJ/mol). In both structures 1b (for both symmetrically independent molecules) and 2b, the best acceptor was engaged in formation of the lactam head-to-head R22(8) motifs. Here again, for the Cu(II) complex no desired octahedral building block (3b: ΔE = 32 kJ/mol) was delivered regardless of the synthetic procedure used. The observed synthetic “outcome” clearly demonstrated the necessity of enhanced Lewis acidity of the β-diketonato ligands (e.g., hfac instead of acac) for a delivery of octahedral Cu(II) acac-based complexes.40 And finally, the reactions of the starting hfac complexes (4− 6) with the 4-quz ligand yielded three new complexes, all with the desired octahedral building units, [Co(hfac)2(4-quz)2] (4b) (ΔE = 98 kJ/mol), [Ni(hfac)2(4-quz)2] (5b) (ΔE = 97 kJ/mol), and [Cu(hfac)2(4-quz)2] (6b) (ΔE = 77 kJ/mol). In all the three structures the best donor binds with the best acceptor, revealing the lactam head-to-head R22(8) motif (Figure 10). As a summary, we are listing all the structurally described complexes along with their ΔE values (in a drop-down manner) and the primary supramolecular motifs observed (Table 3). It can be seen that they fall into two distinct categories, with ΔE > 40 kJ/mol and ΔE < 20 kJ/mol. The former displays only lactam motifs R22(8) and C(4), while the latter presents no significant trend. It lists hydrogen bonds toward both acceptors without any signs of “selectivity” (Figure 11). Our experimental data indicate that the cutoff value (ΔEcutoff), as a boundary that differentiates between locations at the structural landscape comprising different supramolecular products, is indeed quite close to that which we used as a working model (ΔEcutoff ≈ 30 kJ/mol). Where it should actually be drawn is yet difficult to say, but given the fact that our existing data form two clearly distinct classes of results, we believe it to be in the 20−40 kJ/mol range. Ideally, to augment the data in this study and to test our findings, we could complement our data with the relevant structures form the CSD. Unfortunately, the survey of the database did not reveal any hit on the class of the compounds used in this study, that is, octahedral bis-β-diketonato 0-D complexes with transoriented pyridine-/quinoline-based ligands decorated with the carbonyl moiety and additional heterocyclic nitrogen atom(s) in its close proximity. The data also point at a few interesting features. Although the working cutoff value was actually “borrowed” from a substantially different system (a system based on more “rigid” 1-D building units, displaying a lower degree of freedom), it proved to be successful for prediction of the supramolecular outcome of the 0-D system in hand. This fact might indicate its potential independence of the dimensionality of the building unit (i.e., 0-D or 1-D) and its adherence only to the organic supramolecular functionality used as an intended supramolecular vector.

Table 2. Hydrogen-Bond Geometries for Compounds 1a− 6b D−H···Aa 1a-II N2−H2···O1b N2−H2···O2c 1b N2−H2···O4d N4−H4···O1d 2a N2−H2···O3e N4−H4···O6f C3−H3···O1g C12−H12···O4h 2b N2−H2···O1i 4b N2−H2···O1j 5a N2−H2···O1j N4−H4a···O1k 5b N2−H2···O1l 6a N2−H2···O1m 6b N2−H2···O1m

d(D···A)/Å

d(H···A)/Å

∠(D−H···A)/deg

3.050(2) 3.082(2)

2.29(2) 2.56(2)

151(2) 122(2)

2.820(6) 2.779(6)

1.95 1.92(2)

173 177(7)

2.857(8) 2.853(7) 3.152(10) 3.145(10)

2.05 2.05 2.24 2.23

155 155 165 169

2.799(2)

1.93(2)

174(3)

2.813(4)

1.98(2)

168(4)

2.699(10) 2.709(10)

1.90 1.92

155 152

2.804(7)

1.94

178

2.768(2)

1.93(2)

169(3)

2.788(3)

1.93(2)

175(4)

Symmetry operators. b2 − x, −y, 1−z. c1 − x, 1 − y, 1−z. d−x, −y, 1−z. e−x, 1 − y, −z. fx, y + 1, z. g1/2 − x, y − 1/2, 1/2 − z. h3/2 − x, y − 1/2, 1/2 − z. i1 − x, y, 1/2 − z. j1/2 − x, y + 1/2, 3/2 − z. k1/2 − x, y − 1/2, 1/2 − z. l−x, −y, −z. m−x, 1 − y, 1 − z. a

Figure 6. Neighboring metal-containing building units in the crystal structure of 1a-II linked via two N−H···O hydrogen bonds between lactam moieties and forming head-to-head R22(8) motifs (top). The lactam hydrogen atom additionally participates in N−H···Oacac link. The hydrogen atoms at the acac ligands (bottom) were omitted for clarity.

formed in 5a (for both symmetrically independent molecules) and 6a, respectively (Figure 9). This is, in fact, an example of synthon crossover,24 which, although statistically supported for metal-organic setting (by a substantially limited data set, see Supporting Information), cannot be further rationalized based on the available data. F

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

Figure 7. Overlay of the experimental PXRD traces of [Co(acac)2(4-pym)2] (1a-I) (orange) and [Ni(acac)2(4-pym)2] (2a) (green).

Figure 9. Catemeric C(4) in 5a (top) and head-to-head R22(8) lactam synthons in 6a (bottom).

Figure 10. R22(8) motifs linking molecules in 1b and producing the 1D supramolecular chains. The same connectivity was also found in 2b, 4b−6b (Figure S7).

Figure 8. Connectivity patterns observed in the crystal structure of 2a. The metal-containing building units are linked via two single-point N−H···O hydrogen bonds between the lactam nitrogen and acac oxygen atoms resulting in R22(12) motifs and producing 1-D supramolecular chains. The lactam oxygen atom is involved into Car−H···O hydrogen-bond-forming catemeric motifs similar to the lactam ones. Hydrogen atoms from the acac ligands are omitted for clarity.

83%), while for the other two cases, unfortunately, the coordination chemistry failed, thus preventing us from testing all 12 predicted events. Upon the basis of the data produced in this study, a confident conclusion could be made that, for assembling metalbased building units via the same link as in related purely organic solids, two criteria must be met: (i) the introduction of the metal cation and charge-balancing entity must not disrupt the relative importance of the hydrogen-bond donors and acceptors (i.e., the acceptor at the organic functionality should even after introduction of the metal center and charge balancing entity remain the best acceptor on the building unit) and (ii) the difference in the MEP values between competing acceptor sites (organic functionality and the charge-



CONCLUSIONS By comparing differences in calculated MEP values between competing acceptor sites residing on the same molecule (ΔE) with a certain working cutoff value (as a boundary between two potential outcomes; ΔEcutoff), we have successfully predicted the supramolecular outcome for a series of multitopic 0-D complexes. The correct supramolecular link was correctly assigned to 10 of 12 intended complexes (10/12; G

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

X-ray powder diffraction experiments were performed on a Philips PW 1850 diffractometer, Cu Kα radiation, voltage 40 kV, and current 40 mA. The patterns were collected in the angle region between 5° and 50° (2θ) with a step size of 0.02°. Synthesis. Starting complexes [Co(acac)2(H2O)2] (1), [Ni(acac)2(H2O)2] (2), [Cu(acac)2] (3), [Co(hfac)2(H2O)2] (4), [Ni(hfac)2(H2O)2] (5), and [Cu(hfac)2(H2O)2]·H2O (6) were synthesized following a modified procedure of a literature reports.42,43 The complexes 1a−6b were prepared by mixing an ethanol solution (10 mL) of starting metal complex (1−6) with the ethanol solution (10 mL) of ligand (4(3H)-pyrimidinone, 4-pym, or 4(3H)quinazolinone, 4-quz). The resulting mixture was continuously stirred and heated under reflux for 2 h. Solution was then cooled and left standing at room temperature to slowly evaporate. X-ray quality crystals for compounds 1b, 2a, 2b, 4b, 5a, 5b, 6a, and 6b were harvested in a period of 5−10 d. Crystalline products of 1a-I and 4a were also obtained, but the crystal quality was not satisfactory for a single-crystal X-ray diffraction (SCXRD) experiment. Compositional purity of all the obtained products was confirmed by elemental analysis, and phase purity of the final product was examined by analysis of powder X-ray diffraction (PXRD) patterns. For that purpose, the PXRD of the bulk crystals was performed and compared with calculated powder patterns (Figures S8−S19). Recrystallization Experiments. Products 1a−6b were dissolved in an ethanol−chloroform mixture (1:1) at room temperature. The resulting solutions were left to slowly evaporate at room temperature, and X-ray quality crystals of 1a-II were harvested in a period of 5 d. For all other products (1b, 2a−b, 4a−b, 5a−b, 6a−b) the original metal complexes were harvested in complete phase purity, which was proven by a comparison of PXRD patterns of original and recrystallized bulk samples. Single crystals of 4a were still proven to be of inadequate quality for a SCXRD experiment. [Co(acac)2(4-pym)2], 1a-I. Used: [Co(acac)2(H2O)2] (1) (58.6 mg, 0.20 mmol) and 4(3H)-pyrimidinone (40.3 mg, 0.42 mmol). Yield: 53% (47.6 mg, 0,11 mmol). Microanalysis. Calcd for C18H22N4O6Co (Mr = 449.33): C, 48.12%; H, 4.94%; N, 12.47%. Found: C, 48.01%; H, 4.99%; N, 12.72%. ATR-FTIR (cm−1): 3429, 3254 ν(N−H); 1706, 1677 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the experimental PXRD pattern of 2a (Figure S9). [Co(acac)2(4-pym)2], 1a-II. Used: [Co(acac)2(4-pym)2] (1a-I) (30.1 mg, 0.07 mmol). Yield: 60% (18.0 mg, 0.04 mmol). Microanalysis. Calcd for C18H22N4O6Co (Mr = 449.33): C, 48.12%; H, 4.94%; N, 12.47%. Found: C, 48.27%; H, 4.85%; N, 12.51%. ATRFTIR (cm−1): 3063 ν(N−H); 1704, 1669 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figure S8). [Co(acac)2(4-quz)2], 1b. Used: [Co(acac)2(H2O)2] (1) (58.5 mg, 0.20 mmol) and 4(3H)-quinazolinone (61.4 mg, 0.42 mmol). Yield: 57% (62.6 mg, 0,11 mmol). Microanalysis. Calcd for C26H26N4O6Co (Mr = 549.45): C, 56.84%; H, 4.77%; N, 10.20%. Found: C, 56.79%; H, 4.81%; N, 10.24%. ATR-FTIR (cm−1): 3186 ν(N−H); 1675 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figure S10). [Ni(acac)2(4-pym)2], 2a. Used: [Ni(acac)2(H2O)2] (2) (58.7 mg, 0.2 mmol) and 4(3H)-pyrimidinone (40.4 mg, 0.42 mmol). Yield: 64% (57.5 mg, 0,13 mmol). Microanalysis. Calcd for C18H22N4O6Ni (Mr = 449.09): C, 48.14%; H, 4.94%; N, 12.48%. Found: C, 48.32%; H, 4.78%; N, 12.35%. ATR-FTIR (cm−1): 3063 ν(N−H); 1705, 1665 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figure S11). [Ni(acac)2(4-quz)2], 2b. Used: [Ni(acac)2(H2O)2] (2) (58.7 mg, 0.20 mmol) and 4(3H)-quinazolinone (61.4 mg, 0.42 mmol). Yield: 63% (69.2 mg, 0.13 mmol). Microanalysis. Calcd for C26H26N4O6Ni (Mr = 549.22): C, 56.84%; H, 4.77%; N, 10.20%. Found: C, 56.49%; H, 4.80%; N, 10.15%. ATR-FTIR (cm−1): 3185 ν(N−H); 1671 ν(CO). The powder diffraction pattern (bulk sample) was

Table 3. Summary of the Structural Results compound

ΔE/ kJ mol−1

motif

[Ni(acac)2(4-pym)2] [Co(acac)2(4-pym)2]-I [Co(acac)2(4-pym)2]-II [Co(acac)2(4-quz)2]

11 15 15 46

N−H···Oacac N−H···Oacac R22(8) R22(8)

[Ni(acac)2(4-quz)2] [Cu(hfac)2(4-pym)2] [Co(hfac)2(4-pym)2] [Ni(hfac)2(4-pym)2] [Cu(hfac)2(4-quz)2] [Ni(hfac)2(4-quz)2] [Co(hfac)2(4-quz)2]

49 64 71 77 77 97 98

R22(8) R22(8) C(4) C(4) R22(8) R22(8) R22(8)

supramolecular outcome no selectivity

transferability of the synthons from the organic solid state

Figure 11. Summary of the structural results. The data indicate that the ΔEcutoff value, as a boundary between two possible outcomes, should be between 20 and 40 kJ/mol.

compensation entity; ΔE) should be greater than a certain cutoff value (ΔEcutoff). The working cutoff value (ΔEcutoff ≈ 30 kJ/mol), although adopted from a system of different dimensionality (i.e., 1-D) but involving the same supramolecular functionality as intended supramolecular synthetic vector, proved to be a valid decisive criterion for the 0-D system, too. Despite the fact that the working ΔEcutoff value worked well for the tested system, we expect that it might undergo certain changes (tuning) with the availability of more relevant data. The data also indicate that the ΔEcutoff is likely to be primarily dependent on the supramolecular functionality itself rather than the other factors (e.g., metal cation employed, dimensionality of the building unit, or steric requirements). In this work we have shown that simple electrostatic view of hydrogen bond could be employed for predicting the supramolecular outcome of metal-based architectures, and the results produced in this study could be readily employed in practical “inorganic” crystal engineering when designing systems with desired supramolecular connectivity.



EXPERIMENTAL SECTION

General Considerations. Materials and methods. All metal salts, precursors, and solvents were purchased from commercial suppliers and used without further purification. CHN analyses were performed with a PerkinElmer 2400 Series II CHNS analyzer in the Analytical Services Laboratories of the Ruđer Bošković Institute. IR analyses were performed on a PerkinElmer Spectrum Two spectrometer equipped with Diamond UATR accessory. FT-IR spectra were measured in ATR mode in the range of 4000−450 cm−1 with resolution 4 cm−1. H

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

consistent with the pattern calculated from single-crystal data (Figure S12). [Co(hfac)2(4-pym)2], 4a. Used: [Co(hfac)2(H2O)2] (4) (101.9 mg, 0.2 mmol) and 4(3H)-pyrimidinone (40.5, 0.42 mmol). Yield: 71% (94.5 mg, 0.14 mmol). Microanalysis. Calcd for C18H10N4O6F12Co (Mr = 665.23): C, 32.50%; H, 1.52%; N, 8.42%. Found: C, 32.60%; H, 1.52%; N, 8.47%. ATR-FTIR (cm−1): 3401, 3140 ν(N−H); 1725, 1699 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the experimental PXRD pattern of 4b (Figure S13). [Co(hfac)2(4-quz)2], 4b. Used: [Co(hfac)2(H2O)2] (4) (101.9 mg, 0.20 mmol) and 4(3H)-quinazolinone (61.5, 0.42 mmol). Yield: 72% (110.2 mg, 0.14 mmol). Microanalysis. Calcd for C26H14N4O6F12Co (Mr = 765.35): C, 40.80%; H, 1.84%; N, 7.32%. Found: C, 40.74%; H, 1.85%; N, 7.27%. ATR-FTIR (cm−1): 3202 ν(N−H); 1688 ν(C O). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figure S14). [Ni(hfac)2(4-pym)2], 5a. Used: [Ni(hfac)2(H2O)2] (5) (101.7 mg, 0.20 mmol) and 4(3H)-pyrimidinone (40.1, 0.42 mmol). Yield: 74% (98.4 mg, 0.15 mmol). Microanalysis. Calcd for C18H10N4O6F12Ni (Mr = 664.99): C, 32.51%; H, 1.52%; N, 8.43%. Found: C, 32.55%; H, 1.52%; N, 8.37%. ATR-FTIR (cm−1): 3133 ν(N−H); 1724, 1699 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figure S15). [Ni(hfac)2(4-quz)2], 5b. Used: [Ni(hfac)2(H2O)2] (5) (101.8 mg, 0.20 mmol) and 4(3H)-quinazolinone (61.5 mg, 0.42 mmol). Yield: 74% (113.2 mg, 0.15 mmol). Microanalysis. Calcd for C26H14N4O6F12Ni (Mr = 765.11): C, 40.82%; H, 1.84%; N, 7.32%. Found: C, 41.08%; H, 1.82%; N, 7.26%. ATR-FTIR (cm−1): 3202 ν(N−H); 1690 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figure S16). [Cu(hfac)2(4-pym)2], 6a. Used: [Cu(hfac)2(H2O)2]·H2O (6) (106.4 mg, 0.20 mmol) and 4(3H)-pyrimidinone (40.4, 0.42 mmol). Yield: 59% (79.0 mg, 0.12 mmol). Microanalysis. Calcd for C18H10N4O6F12Cu (Mr = 669.78): C, 32.28%; H, 1.50%; N, 8.36%. Found: C, 32.55%; H, 1.50%; N, 8.39%. ATR-FTIR (cm−1): 3148 ν(N−H); 1721, 1673 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figure S17). [Cu(hfac)2(4-quz)2], 6b. Used: [Cu(hfac)2(H2O)2]·H2O (6) (106.3 mg, 0.20 mmol) and 4(3H)-quinazolinone (61.4 mg, 0.42 mmol). Yield: 54% (83.2 mg, 0.11 mmol). Microanalysis. Calcd for C26H14N4O6F12Cu (Mr = 769.92): C, 40.56%; H, 1.83%; N, 7.28%. Found: C, 40.38%; H, 1.84%; N, 7.30%. ATR-FTIR (cm−1): 3204 ν(N−H); 1674 ν(CO). The powder diffraction pattern (bulk sample) was consistent with the pattern calculated from single-crystal data (Figures S18 and S19). Crystallographic Data Collection and Structure Determination. Single crystals were mounted in a random orientation on a glass fibber using mineral oil. Data collections were performed on an Oxford Diffraction Xcalibur four-circle kappa geometry single-crystal diffractometer with Sapphire 3 CCD detector, using a graphite monochromated Mo Kα (λ = 0.710 73 Å) radiation and applying the CrysAlisPro Software system44 at 296(2) K (1a−6a) and 200(2) K (6b). Data reduction, including absorption correction, was done by CrysAlisPro program. The structures were solved by SHELXT program.45 The coordinates and the anisotropic thermal parameters for all non-hydrogen atoms were refined by full-matrix least-squares methods based on F2 using the SHELXL program. Hydrogen atoms were generated geometrically using the riding model with the isotropic factor set at 1.2 Ueq of the parent atom. Graphical work was performed by Mercury 3.9.46 The thermal ellipsoids were drawn at the 20% probability level. General and crystal data with the summary of intensity data collection and structure refinement for compounds 1a−7a are given in Table S1 (Supporting Information). CCDC Nos. 1887879−1887887 contain the supplementary crystallographic data for this paper. Compounds 2a and 5a have a twin component that was present in all of the measured crystals,

which results in higher residual electron density peaks. After the twinning was treated using a twin law (0 0−1 0−1 0−1 0 0), the residual electron density peaks drops significantly for both structures. Computational Details. Molecular geometries of 1a−6b were generated starting from the complex [Co(acac)2(H2O)2] (1) extracted from the CSD, optimizing its molecular geometry and substituting axial H2O molecules for the lactam ligands (4-pym, 4quz), terminal −CH3 groups for −CF3, and/or metal centers (Co(II), Ni(II), and Cu(II)). Bond lengths (C−H, N−H) were normalized to values obtained from neutron diffraction experiments. All optimizations were performed in Gaussian1647 with B3LYP in combination with def2-TZVP basis set and D3 Grimme’s dispersion corrections.48 The MEP maps were visualized in GaussView 6.0.49 The MEP at a specific point on the 0.002 au isodensity surface is given by the electrostatic potential energy (in kJ/mol) that a positive unit charge would experience at that point. A continuous color spectrum is used to assign different values of electrostatic potential energy values, the most negative values being red and the most positive values colored blue. A detailed description of the process used to generate all of the starting geometries is given in the Supporting Information.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.8b01930. Single-crystal X-ray crystallographic data (SCXRD), power X-ray crystallographic data (PXRD), molecular electrostatic potential (MEP) calculation, CSD survey (PDF) Accession Codes

CCDC 1887879−1887887 contain 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 Author

*E-mail: [email protected]. ORCID

Marijana Đaković: 0000-0001-6789-6399 Author Contributions

M.B. and M.Đ. conceived and designed the experiments; M.B. performed the experiments, I.K. contributed with computational work and data analysis, while M.B. and M.Đ. analyzed the data and wrote the paper. All authors have given approval to the final version of the manuscript. Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work has been fully supported by the Croatian Science Foundation under Project No. UIP-11-2013-1809. Computational resources were provided by the Croatian National Grid Infrastructure (http://www.cro-ngi.hr) at Zagreb University Computing Centre.



REFERENCES

(1) Desiraju, G. R. Crystal engineering: a holistic view. Angew. Chem., Int. Ed. 2007, 46, 8342−8356. (2) Tiekink, E. R. T.; Vittal, J.; Zaworotko, M. Organic Crystal Engineering: Frontiers in Crystal Engineering; Wiley & Sons, 2010.

I

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

(24) Aakeröy, C. B.; Chopade, P. D.; Desper, J. Avoiding “synthon crossover” in crystal engineering with halogen bonds and hydrogen bonds. Cryst. Growth Des. 2011, 11, 5333−5336. (25) Aakeröy, C. B.; Epa, K. Controlling supramolecular assembly using electronic effects. Top. Curr. Chem. 2011, 351, 125−147. (26) Aakeröy, C. B.; Chopade, P. D.; Ganser, C.; Desper, J. Facile synthesis and supramolecular chemistry of hydrogen bond/halogen bond-driven multi-tasking tectons. Chem. Commun. 2011, 47, 4688− 4690. (27) Aakeröy, C. B.; Epa, K.; Forbes, S.; Schultheiss, N.; Desper, J. Ranking relative hydrogen-bond strengths in hydroxybenzoic acids for crystal-engineering purposes. Chem. - Eur. J. 2013, 19, 14998−15003. (28) Aakeröy, C. B.; Wijethunga, T. K.; Desper, J. Molecular electrostatic potential dependent selectivity of hydrogen bonding. New J. Chem. 2015, 39, 822−828. (29) Perera, M. D.; Desper, J.; Sinha, A. S.; Aakeröy, C. B. Impact and importance of electrostatic potential calculations for predicting structural patterns of hydrogen and halogen bonding. CrystEngComm 2016, 18, 8631−8636. (30) Aakeröy, C. B.; Spartz, C. L.; Dembowski, S.; Dwyre, S.; Desper, J. A systematic structural study of halogen bonding versus hydrogen bonding within competitive supramolecular systems. IUCrJ 2015, 2, 498−510. (31) Zordan, F.; Brammer, L.; Sherwood, P. Supramolecular Chemistry of Halogens: Complementary Features of Inorganic (MX) and Organic (C-X′) Halogens Applied to M-X···X′-C Halogen Bond Formation. J. Am. Chem. Soc. 2005, 127, 5979−5989. (32) Chakraborty, P.; Adhikary, J.; Samanta, S.; Majumder, I.; Massera, C.; Escudero, D.; Ghosh, S.; Bauza, A.; Frontera, A.; Das, D. Influence of para substituents in controlling photophysical behavior and different non-covalent weak interactions in zinc complexes of a phenol based “end-off” compartmental ligand. Dalton Trans 2015, 44, 20032−20044. (33) Đaković, M.; Soldin, Ž .; Kukovec, B.-M.; Kodrin, I.; Aakeröy, C. B.; Baus, N.; Rinkovec, T. Building inorganic supramolecular architectures using principles adopted from the organic solid state. IUCrJ 2018, 5, 13−21. (34) Kukovec, B.-M.; Malik, M.; Kodrin, I.; Aakeröy, C. B.; Đaković, M. Directed assembly of acac-based complexes by deliberately finetuning electrostatic molecular-recognition events. Cryst. Growth Des. 2016, 16, 7308−7317. (35) Borovina, M.; Kodrin, I.; Đaković, M. Testing the limits of halogen bonding in coordination chemistry. CrystEngComm 2018, 20, 539−549. (36) Etter, M. C. Encoding and decoding hydrogen-bond patterns of organic compounds. Acc. Chem. Res. 1990, 23, 120−126. (37) Aakeröy, C. B.; Wijethunga, T. K.; Desper, J.; Đaković, M. Crystal engineering with iodoethynylnitrobenzenes: A group of highly effective halogen-bond donors. Cryst. Growth Des. 2015, 15, 3853− 3861. (38) Groom, C. R.; Bruno, I. J.; Lightfoot, M. P.; Ward, S. C. The Cambridge Structural Database. Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater. 2016, B72, 171−179. (39) Aakeröy, C. B.; Wijethunga, T. K.; Desper, J.; Đaković, M. Electrostatic potential differences and halogen-bond selectivity. Cryst. Growth Des. 2016, 16, 2662−2670. (40) Robertson, C. C.; Wright, J. S.; Carrington, E. J.; Perutz, R. N.; Hunter, C. A.; Brammer, L. Hydrogen bonding vs. halogen bonding: the solvent decides. Chem. Sci. 2017, 8, 5392−5398. (41) Reichardt, C. Solvents and Solvent Effects in Organic Chemistry, 3rd ed.; Wiley-VCH Publishers, 2003. (42) Soldatov, D. V.; Ripmeester, J. A. 4-Vinylpyridine-modified nickel and cobalt dibenzoylmethanates as new hosts: Inclusions with carbon tetrachloride and chlorobenzene. Supramol. Chem. 2001, 12, 357−368. (43) Goff, H. M.; Hines, J.; Griesel, J.; Mossman, C. J. Synthesis, characterization, and use of a cobalt(II) complex as an NMR shift reagent: An integrated laboratory experiment. J. Chem. Educ. 1982, 59, 422−423.

(3) Braga, D.; Grepioni, F.; Maini, L.; d’Agostino, S. Making crystals with a purpose; a journey in crystal engineering at the University of Bologna. IUCrJ 2017, 4, 369−379. (4) Nangia, A. K.; Desiraju, G. R. Crystal Engineering. An outlook for the future. Angew. Chem., Int. Ed. 2018, DOI: 10.1002/ anie.201811313. (5) Aakeröy, C. B.; Forbes, S.; Desper, J. Using cocrystals to systematically modulate aqueous solubility and melting behavior of an anticancer drug. J. Am. Chem. Soc. 2009, 131, 17048−17049. (6) Aakeröy, C. B.; Forbes, S.; Desper, J. Altering physical properties of pharmaceutical co-crystals in a systematic manner. CrystEngComm 2014, 16, 5870−5877. (7) Sandhu, B.; Sinha, A. S.; Desper, J.; Aakeröy, C. B. Modulating the physical properties of solid forms of urea using co-crystallization technology. Chem. Commun. 2018, 54, 4657−4660. (8) Dunitz, J. D.; Gavezzotti, A. Supramolecular synthons: Validation and ranking of intermolecular interaction energies. Cryst. Growth Des. 2012, 12, 5873−5877. (9) Bolla, G.; Mittapalli, S.; Nangia, A. Modularity and threedimensional isostructurality of novel synthons in sulfonamide-lactam cocrystals. IUCrJ 2015, 2, 389−401. (10) Vishweshwar, P.; Nangia, A.; Lynch, V. M. Supramolecular synthons in phenol−isonicotinamide adducts. CrystEngComm 2003, 5, 164−168. (11) Aakeröy, C. B.; Schultheiss, N.; Rajbanshi, A.; Desper, J.; Moore, C. Supramolecular synthesis based on a combination of hydrogen- and halogen bonds. Cryst. Growth Des. 2009, 9, 432−441. (12) Corpinot, M. K.; Stratford, S. A.; Arhangelskis, M.; AnkaLufford, J.; Halasz, I.; Judaš, N.; Jones, W.; Bučar, D.-K. On the predictability of supramolecular interactions in molecular cocrystals − the view from the bench. CrystEngComm 2016, 18, 5434−5439. (13) Brammer, L. Developments in inorganic crystal engineering. Chem. Soc. Rev. 2004, 33, 476−489. (14) Blake, A. J.; Champness, N.; Hubberstey, P.; Li, W.-S.; Withersby, M. A.; Schröder, M. Inorganic crystal engineering using self-assembly of tailored building-blocks. Coord. Chem. Rev. 1999, 183, 117−138. (15) Raynal, M.; Ballester, P.; Vidal-Ferran, A.; van Leeuwen, P. W. N. M. Supramolecular catalysis. Part 1: non-covalent interactions as a tool for building and modifying homogeneous catalysts. Chem. Soc. Rev. 2014, 43, 1660−1733. (16) Christopherson, J.-C.; Potts, K. P.; Bushuyev, O. S.; Topić, F.; Huskić, I.; Rissanen, K.; Barrett, C. J.; Frišcǐ ć, T. Assembly and dichroism of a four-component halogen-bonded metal-organic cocrystal salt solvate involving dicyanoaurate(I) acceptors. Faraday Discuss. 2017, 203, 441−457. (17) Atzori, M.; Artizzu, F.; Marchiò, L.; Loche, D.; Caneschi, A.; Serpe, A.; Deplano, P.; Avarvari, N.; Mercuri, M. L. Switching-on luminescence in anilate-based molecular materials. Dalton Trans 2015, 44, 15786−15802. (18) Visbal, R.; Gimeno, M. C. N-heterocyclic carbene metal complexes: photoluminescence and applications. Chem. Soc. Rev. 2014, 43, 3551−3574. (19) Wang, X. Y.; Avendano, C.; Dunbar, R. K. Molecular magnetic materials based on 4d and 5d transition metals. Chem. Soc. Rev. 2011, 40, 3213−3238. (20) Vittal, J. J.; Quah, H. S. Photochemical reactions of metal complexes in the solid state. Dalton Trans 2017, 46, 7120−7140. (21) Tuikka, M.; Hirva, P.; Rissanen, K.; Korppi-Tommola, J.; Haukka, M. Halogen bondinga key step in charge recombination of the dye-sensitized solar cell. Chem. Commun. 2011, 47, 4499−4501. (22) Đaković, M.; Vila-Viçosa, D.; Calhorda, M. J.; Popović, Z. Coordination-driven self-assembly of thiocyanate complexes of Co(II), Ni(II) and Cu(II) with picolinamide: a structural and DFT study. CrystEngComm 2011, 13, 5863−5871. (23) Đaković, M.; Vila-Viçosa, D.; Bandeira, N. A. G.; Calhorda, M. J.; Kozlevčar, B.; Jagličić, Z.; Popović, Z. Can self-assembly of copper(II) picolinamide building blocks be controlled? CrystEngComm 2013, 15, 8074−8087. J

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Article

(44) CrysAlisPRO; Oxford Diffraction/Agilent Technologies UK Ltd: Yarnton, England, 2019. (45) Sheldrick, G. M. SHELXT - Integrated space-group and crystalstructure determination. Acta Crystallogr., Sect. A: Found. Adv. 2015, A71, 3−8. (46) Macrae, C. F.; Bruno, I. J.; Chisholm, J. A.; Edgington, P. R.; McCabe, P.; Pidcock, E.; Rodriguez-Monge, L.; Taylor, R.; van de Streek, J.; Wood, P. A. New features for the vsualization and investigation of crystal structures. J. Appl. Crystallogr. 2008, 41, 466− 470. (47) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson, G. A.; Nakatsuji, H.; Li, X.; Caricato, M.; Marenich, A. V.; Bloino, J.; Janesko, B. G.; Gomperts, R.; Mennucci, B.; Hratchian, H. P.; Ortiz, J. V.; Izmaylov, A. F.; Sonnenberg, J. L.; Williams-Young, D.; Ding, F.; Lipparini, F.; Egidi, F.; Goings, J.; Peng, B.; Petrone, A.; Henderson, T.; Ranasinghe, D.; Zakrzewski, V. G.; Gao, J.; Rega, N.; Zheng, G.; Liang, W.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Throssell, K.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M. J.; Heyd, J. J.; Brothers, E. N.; Kudin, K. N.; Staroverov, V. N.; Keith, T. A.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A. P.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Millam, J. M.; Klene, M.; Adamo, C.; Cammi, R.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Farkas, O.; Foresman, J. B.; Fox, D. J. Gaussian 16, Revision B.01; Gaussian, Inc.: Wallingford, CT, 2016. (48) Grimme, S.; Antony, J.; Ehrlich, J. S.; Krieg, H. J. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. (49) Dennington, R.; Keith, T. A.; Millam, J. M. GaussView, Version 6; Semichem Inc.: Shawnee Mission, KS, 2016.

K

DOI: 10.1021/acs.cgd.8b01930 Cryst. Growth Des. XXXX, XXX, XXX−XXX