Crystal Structures, Photoluminescence, and ... - ACS Publications

Jul 16, 2019 - different molecules is a possible way of intentionally influencing the .... in both pyridyl and phenyl rings are omitted for clarity). ...
1 downloads 0 Views 3MB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2019, 4, 12230−12237

http://pubs.acs.org/journal/acsodf

Crystal Structures, Photoluminescence, and Magnetism of Two Novel Transition-Metal Complex Cocrystals with Three-Dimensional H‑Bonding Organic Framework or Alternating Noncovalent Anionic and Cationic Layers Xu-Sheng Gao,*,† Hai-Jie Dai,† Yuerou Tang,§ Mei-Juan Ding,† Wen-Bo Pei,†,‡ and Xiao-Ming Ren*,†,§ Downloaded via 193.56.65.96 on July 18, 2019 at 06:55:57 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



State Key Laboratory of Materials-Oriented Chemical Engineering and College of Chemistry & Molecular Engineering and College of Materials Science and Engineering, Nanjing Tech University, Nanjing 211816, P. R. China § American Division, Nanjing Jinling High School, Nanjing 210005, P. R. China ‡

S Supporting Information *

ABSTRACT: Cocrystallization may alter material physicochemical properties; thus, the strategy of forming a cocrystal is generally used to improve the material performance for practical applications. In this study, two transition-metal complex cocrystals [Zn(bpy)3]H0.5BDC·H1.5BDC·0.5bpy· 3H2O (1) and [Cu2(BDC)(bpy)4]BDC·bpy·2H2O (2) have been achieved using a hydrothermal reaction, where bpy and H2BDC represent 2,2′-bipyridine and benzene-1,3-dicarboxylic acid, respectively. Cocrystals were characterized by microanalysis, infrared spectroscopy, and UV−visible spectroscopy. Cocrystal 1 contains five components and crystallizes in a monoclinic space group P21/n. The H0.5BDC1.5−, H1.5BDC0.5−, and H2O molecules construct three-dimensional H-bonding organic framework; the [Zn(bpy)3]2+ coordination cations and uncoordinated bpy molecules reside in channels, where two coordinated bpy ligands in [Zn(bpy)3]2+ and one uncoordinated bpy adopt sandwich-type alignment via π···π stacking interactions. Cocrystal 2 with four components crystallizes in a triclinic space group P-1 to form alternating layers; the binuclear [Cu2(bpy)4(BDC)]2+ cations and uncoordinated bpy molecules build the cationic layers, and the BDC2− species with disordered lattice water molecules form the anionic layers. Cocrystal 1 shows intense photoluminescence at an ambient condition with a quantum yield of 14.96% and decay time of 0.48 ns, attributed to the π* → π electron transition within phenyl/pyridyl rings, and 2 exhibits magnetic behavior of an almost isolated spin system with rather weak antiferromagnetic coupling in the [Cu2(bpy)4(BDC)]2+ cation.



INTRODUCTION

influencing the position of molecules in a crystal lattice and allows for the investigation of newly generated macroscopic properties.14−18 It has been known that the material properties, including magnetism,19 photoluminescence,15,20−23 mechanical strength,14 etc., are modifiable using a cocrystallization strategy, and the desired physical and chemical properties are delivered by means of selecting a suitable cocrystal precursor to form a cocrystal with the active compound.2 The organic cocrystals have been widely reported hitherto,14,15,19,24−31 and in this context, the multicomponents generally have similar molecular structures and connect together via noncovalent interactions, such as π···π stacking and C−H···π or H-bonding interactions. For example, a survey of the Cambridge Structural Database32 reveals that many cocrystal compounds of carboxylate salts contain the neutral

Cocrystals, multicomponent molecular complexes defined by Bond,1 are at the forefront of the quest for novel crystal forms.2 Design and synthesis of cocrystals have gained significant interest in recent years, and this is because that the cocrystals may alter the material physicochemical properties and further improve the material performance for practical applications. In pharmaceutical materials, cocrystals can enhance solubility and dissolution by forming a crystal of drug and other benign molecule or coformer with specific stoichiometric compositions. Meanwhile, the cocrystallization process does not affect the chemical integrity of molecule compounds.3−12 Pharmaceutical cocrystals, formed between an active pharmaceutical ingredient (API) and a cocrystal precursor that is a solid under ambient conditions, represent a new paradigm in API formulation that might address important intellectual and physical property issues in the context of drug development and delivery.13 In other material science, the cocrystal of two different molecules is a possible way of intentionally © 2019 American Chemical Society

Received: May 30, 2019 Accepted: July 2, 2019 Published: July 16, 2019 12230

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237

ACS Omega

Article

carboxylic acid molecules and numerous neutral pyridyl derivative residues in the cocrystals of corresponding pyridinium derivatives. In contrast, the cocrystals containing metal coordination compounds are infrequently reported, and this situation is due to the fact that, commonly, the metal complexes have different coordination modes, and the compounds with different coordination modes rarely possess similar lattice packing forces and exhibit similar crystallization kinetics.33−38 In this study, we have prepared two metal complex cocrystals of [Zn(bpy)3]H0.5BDC·H1.5BDC·0.5bpy·3H2O (1) and [Cu2(BDC)(bpy)4]BDC·bpy·2H2O (2) utilizing the selfassembly strategy of auxiliary ligands with transition-metal zinc(II) and copper(II) ions, respectively. We found that the uncoordinated ligands form a cocrystal with the transitionmetal complex cations, and the H-bonding and π···π stacking interactions play a critical role in the formation of two metal complex cocrystals. Herein, we present the crystal structure analyses for both 1 and 2, the solid state photoluminescence at an ambient condition for 1 and the magnetic property in 2− 300 K for 2.

Figure 2. (a) Distorted octahedral coordination geometry of Zn1 center, (b) centrosymmetric bpy molecule, (c) dimer of [H(BDC)2]3−, and (d) dimer of [H3(BDC)2]− (the hydrogen atoms in both pyridyl and phenyl rings are omitted for clarity).



RESULTS AND DISCUSSION Crystal Structures. Cocrystal 1 crystallizes in a monoclinic space group P21/n. An asymmetry unit of 1, as shown in Figure 1, includes one coordination cation [Zn(bpy)3]2+, one H0.5BDC1.5−, one H1.5BDC0.5−, one half uncoordinated bpy molecule, and three lattice H2O molecules.

As shown in Figure 3a, the neighboring [H(BDC)2]3− and [H3(BDC)2]− acidic dimers are connected together via strong H-bonds to form a zigzag H-bond chain along the c-axis direction, where the phenyl rings in the neighboring [H(BDC)2]3− and [H3(BDC)2]− dimers make a dihedral angle of 69.6°. Besides the strong H-bonds between different acidic dimers as well as within each type of acidic dimer, as displayed in Figure 3b, the strong O−H···O H-bond interactions also appear between different water molecules and between water molecules and each of two types of acidic dimers in which the typical distances between the atoms of the H-bond donor and acceptor (O···O), as listed in Table 1, fall in the ranges of 2.801(3)−2.856(3) Å. As shown in Table 1, the H-bond parameters in 1 are comparable to those in other O−H···O bond systems.44−46 The zigzag H-bond chains formed by two types of acidic dimers are further developed into a 3D H-bonding organic framework (HOF) through the H-bonding interactions between lattice water molecules and two types of acidic dimers, and the HOF shows channels along the b-axis direction, which is illustrated in Figure 3c. Both [Zn(bpy)3]2+ coordination cations and lattice bpy molecules are accommodated in the channels (Figure 3d). As shown in Figure 3e, each uncoordinated bpy molecule is sandwiched between two bpy molecules from two neighboring [Zn(bpy)3]2+ coordination cations, with a mean plane distance of 2.981 Å between the neighboring bpy rings and a closest interatomic distance of 2.868 Å. Obviously, the five different components cocrystallized in 1 are due to multiple strong H-bonding47,48 and π···π stacking interactions.49,50 Cocrystal 2 belongs to a triclinic space group P-1, and its asymmetric unit consists of one binuclear Cu2+ coordination cation, one disordered BDC2−, one lattice bpy molecule, and two heavily disordered lattice water molecules, forming a cocrystal with a 1:1:1:2 stoichiometry, which is shown in Figure 4a. Two crystallographically inequivalent Cu2+ ions in the binuclear unit show a distorted square pyramid coordination sphere, as displayed in Figure 4b, and each coordination square pyramid is comprised of four N atoms from two different bpy ligands and one O atom from one BDC2− ligand, where each carboxylate adopts the mono-

Figure 1. Asymmetry unit of 1 (the hydrogen atoms except those of the carboxylic groups are omitted for clarity, and the thermal ellipsoids are drawn at the 50% probability level).

The Zn2+ ion in [Zn(bpy)3]2+ shows the distortedly octahedral coordination sphere, which is comprised of six nitrogen atoms (labeled as N1−N6) from three bpy ligands (Figure 2a). The bond lengths of Zn−N range from 2.149(1) to 2.181(1) Å, and three bite angles of N−Zn−N show the similar values of 75.25(7)°, 76.32(7)°, and 76.49(7)°; the bond angles of N−Zn−N in which two N atoms adopt in a trans manner arrangement are equal to 167.99(7), 164.54(7), and 165.59(7), respectively. These geometric parameters in [Zn(bpy)3]2+ are comparable to those in other analogues.39−43 The uncoordinated bpy molecule shows a centrosymmetric conformation, as shown in Figure 2b, the inversion center locates at the midpoint of C43-C43#7 (symmetry code: #7 = 2 − x, −y, −z), and the inversion symmetry constrains two pyridyl rings being coplanar. The partially deprotonated H0.5BDC1.5− and H1.5BDC0.5− anions form two types of centrosymmetrical supramolecular dimers through H-bonds. In a dimer, two monomers share one proton, which occupies at an inversion center, and two phenyl rings are coplanar (Figure 2c,d). 12231

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237

ACS Omega

Article

Figure 3. (a) 1D zigzag H-bond chain formed between [H(BDC)2]3− (magenta color) and [H3(BDC)2]− (blue color) dimers running along c-axis direction, (b) H-bonds between water molecules, between acids, and between water molecules and acids, (c) 3D HOF with channels along b axis, (d) crystal packing diagram showing HOF and components in channels, and (e) π···π stacking between coordinated and lattice bpy molecules in channels of HOF in the crystal structure of 1.

Table 1. Geometric Parameters of H-Bonds in the Crystal Structure of 1 DHA O12−H12A···O11 O9−H9A···O1#3 O11−H11A···O8#3 O5−H5A···O2#2 O9−H9B···O2#1

symmetry

d(D-H) (Å)

d(H···A) (Å)

d(D···A) (Å)

∠D-H···A (°)

x, y + 1, z x, y + 1, z x − 1/2, −y + 1/2, z + 1/2 −x + 3/2, y + 1/2, −z + 1/2

0.795(18) 0.86(5) 0.86(4) 0.86(4) 0.92(5)

2.09(2) 1.97(5) 1.95(4) 1.73(4) 1.94(5)

2.853(4) 2.826(3) 2.801(3) 2.565(2) 2.856(3)

160(4) 172(4) 169(3) 162(4) 177(4)

bonds show the distances of Cu1−O1 = 1.991(2) Å and Cu2− O3 = 2.029(2) Å. These bond lengths are similar to those reported in the literature.51−54 It is worth noting that two different types of metal coordination cations [Zn(bpy)3]2+ and [Cu2(bpy)4(BDC)]2+ appear in 1 and 2, respectively, although almost the same reaction conditions were used for preparation of them. The formation of distinct metal coordination cations in 1 and 2 may be relevant to two reasons: (1) the bpy is a chelating ligand, and the entropy effect promotes the Zn2+/ Cu2+ ion to easily form the species of the metal ion with bpy compared to carboxylate using a monodentate binding manner and (2) the Cu2+ ion favors to form quadrilateral or square pyramidal coordination geometry owing to the Jahn−Teller effect arising from its electron configuration of d9. In the crystal of 2, the uncoordinated bpy molecule and deprotonated BDC2− coexist in the lattice, and the bpy molecule shows a transoid conformation, which is similar to that in 1. One carboxylate shows disordered, and the oxygen atoms have two possible positions (Figure 4a). As shown in Figure 5a, 2 shows lamellar packing fashion, and the layers are parallel to the crystallographic (0 1 −1) plane. The anionic layer contains the deprotonated BDC2− ions and heavily disordered lattice water molecules (removed in the refined crystal structure, see Figure S3), and the cationic

Figure 4. (a) Asymmetry unit (all hydrogen atoms are omitted for clarity, and the thermal ellipsoids are drawn at the 50% probability level) and (b) distorted coordination square pyramids of two crystallographically different Cu2+ ions in the crystal structure of 2.

dentate binding manner in the BDC2− ligand. The Cu−N distances range from 1.975(2) to 2.177(2) Å, and the Cu−O 12232

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237

ACS Omega

Article

PXRD patterns indicate the high phase purity of the crystalline samples. UV−vis spectra of cocrystals 1 and 2 in the solid state are shown in Figure 6a. Two compounds show similar ultraviolet spectra in a region of 250−400 nm, with two intense absorption bands located at ∼287 and 285 nm and three shoulders; these absorption bands correspond to the π → π* electron transitions within the aromatic rings, including the pyridyl rings in bpy and phenyl rings in BDC species.52,55 Cocrystals 1 and 2 display different spectra in the visible spectral regime (400−800 nm). Cocrystal 2 exhibits a broad absorption band spanning from 550 to 800 nm, attributed to the d−d electron transition in the Cu2+ ion. In the spectrum of 1, there is no sizable absorption in a spectral regime of 400− 800 nm, this is because that the Zn2+ ion has d10 electron configuration, and no d−d transition is possible. The solidstate excitation and emission spectra of 1 were investigated at an ambient condition. Cocrystal 1 was irradiated by ultraviolet light with λex = 335 nm, giving rise to an emission band with the maximum λem = 368 nm (Figure 6b), which is attributed to the π* → π transitions within the aromatic rings. A similar emission band was also observed in other compounds containing [Zn(bpy)3]2+ coordination cations.56,57 The quantum yield of 1 was determined to be 14.96% under an ambient condition, and the curves of fluorescence decay are shown for 1 in Figure S5. The process of fluorescence decay follows a single exponential decay law. The best fit of the timedependent fluorescence intensity to eq 1 led to a fluorescence lifetime τ of 0.48 ns, which falls within the nanosecond range and shows fluorescence character.

Figure 5. (a) Crystal packing diagram showing a lamellar structure, and the layers are parallel to the (0 1 −1) plane, (b) a cationic layer in 2 (all H atoms are omitted for clarity) and π···π stacking between bpy molecules in 2, (c) ladder-like motif in a manner of ···(coordinated bpy)2(uncoordinated bpy)··· along the b + c direction, and (d) two neighboring coordinated bpy molecules along the a + b + c direction.

layer is comprised of binuclear [Cu2(bpy)4(BDC)]2+ units and uncoordinated bpy molecules. As displayed in Figure 5b,c, in a cationic layer, the uncoordinated and coordinated bpy molecules form a ladder-like motif in a manner of ··· (coordinated bpy)2(uncoordinated bpy)··· along the b + c direction, where the bpy molecules act as the rungs of a ladder and interact with each other via π···π interactions. The mean plane of molecules between the neighboring uncoordinated bpy and coordinated bpy molecules shows a distance of 3.493/ 3.657/3.405/3.523 Å, and the closest interatomic distance is 3.268/3.288/3.323/3.329 Å. The mean plane of molecules between the neighboring coordinated bpy molecules shows a distance of 3.429/3.433 Å, and the shortest interatomic distance is 3.382/3.380 Å. These distances fall within the typical separation range of π···π interactions. In addition, two neighboring coordinated bpy molecules show the π···π interaction along the a + b + c direction (Figure 5b,d) with a mean plane distance of 3.354/3.496 Å and a shortest interatomic separation of 3.292/3.476 Å. PXRD, UV−Vis Absorption Spectra of 1 and 2, and Photoluminescence Spectrum of 1. The synthesized and simulated PXRD patterns are shown in Figure S4 for 1 and 2, respectively. The well-matched synthesized and simulated

y = A*exp( −t /τ )

(1)

Magnetic Property of 2. Temperature-dependent magnetic susceptibility is shown in Figure 7a, where χm denotes the molar magnetic susceptibility of 2 with two Cu2+ ions per formula unit, and the diamagnetism was not subtracted from χm. The plot of χm−T shows the typical Curie−Weiss magnetic behavior; thus, first, the Curie−Weiss law was used for the fit of temperature-dependent magnetic susceptibility data, χm =

C + χ0 T−θ

(2)

In eq 2, C and θ represent the Curie and Weiss constants, respectively, and χ0 denotes the temperature-independent magnetic susceptibility term, including the diamagnetism and possible van Vleck paramagnetism. The best fit gave the parameters of C = 0.773 emu·K·mol−1, θ = −0.35 K, and χ0 = −1.8 × 10−6 emu·mol−1. The fitted Curie constant is slightly

Figure 6. (a) UV−vis spectra of 1 and 2 and (b) excitation and emission spectra of 1 in solid state. 12233

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237

ACS Omega

Article

Figure 7. Plots of (a) χm vs T (the black squares denote the experimental magnetic susceptibility; the red line is theoretically reproduced using fitted parameters) and (b) χm(dimer)T vs T for 2.

larger than a spin-only value of 0.75 emu·K·mol−1 for two uncoupled S = 1/2 spins with g = 2.0. Too small θ value indicates that the magnetic coupling is quite weak between the neighboring magnetic centers, and this result is in agreement with the crystal structure analysis (refer to the Crystal Structures description section). Next, the Bleaney−Bowers equation was further used for the magnetic susceptibility fit to estimate the magnetic coupling interaction within a dimer. The function of molar magnetic susceptibility against temperature is expressed in eq 3 (Bleaney−Bowers equation) for an S = 1/2 spin dimer, which is deduced from the spin Hamiltonian H = −2JS1S2. χm (dimer) =

Ng 2μB2 kBT



heavily disordered water molecules. The formation of cocrystals 1 and 2 is attributed to the strong H-bonding and π···π stacking interactions between multicomponents. Cocrystal 1 emits intense photoluminescence in the solid state at an ambient condition with a quantum yield of 14.96% and decay time of 0.48 ns, assigned to the π−π* electron transition within the aromatic rings, and 2 shows the magnetic characterization of an almost isolated magnetic system with rather weak antiferromagnetic coupling in the Cu2(bpy)4(BDC)2+ dimer.



2

( ) 2J

3 + exp − k T B

(3)

In eq 3, the symbols of N, g, μB, and J have normal meanings, and then, the total experimental magnetic susceptibility follows the eq 4, χm = χm (dimer) + χ0

EXPERIMENTAL SECTION

Materials and Physical Measurements. Chemicals, including Zn(NO3)2·6H2O, Cu(NO3)2·3H2O, H2BDC, and bpy, and all solvents were commercially purchased and directly used without further purification. Elemental analyses (EA) for carbon, hydrogen, and nitrogen were performed on a PerkinElmer CHN-2400 elemental analyzer. Infrared spectra (IR) were recorded in the wavenumber ranges of 400−4000 cm−1 on an AVATAR-360 spectrophotometer using KBr pellets. The UV−vis spectra were recorded using a PerkinElmer Lambda 950 UV−vis spectrometer. The 1H NMR spectra were recorded in DMSOd6 on a Bruker 400 MHz. The photoluminescence (PL) spectra were measured with an Edinburgh Instruments FLS920P fluorescence spectrometer at room temperature. The emission decay lifetime and quantum yield were recorded at room temperature in an Edinburgh FLS980 fluorescence spectrometer. Magnetic susceptibility was measured for the polycrystalline sample in the temperature ranges of 2−300 K using a Quantum Design MPMS-5S superconducting quantum interference device (SQUID) magnetometer, and the diamagnetism arising from the core of atoms was not removed from the experimental magnetic susceptibility. Preparation and Characterization of 1 and 2. Preparation of [Zn(bpy)3]H0.5BDC·H1.5BDC·0.5bpy·3H2O (1). A mixture of Zn(NO3)2·6H2O (0.15 g, 0.50 mmol), H2BDC (0.17 g, 1.0 mmol), bpy (0.080 g, 0.50 mmol), and NaOH (0.060 g, 1.5 mmol) was dissolved in 15 mL of distilled water. The mixture was stirred for 0.5 h, sealed in a 23 mL Teflonlined stainless steel autoclave, heated at 170 °C for 3 days, and then slowly cooled to ambient temperature. The solutions were filtered, and the filtration was allowed to slowly evaporate solvent. Pink crystals suitable for single-crystal X-ray structure analysis were obtained after 3 days and dried in air yielding 0.025 g (∼50% yield calculated using the reactant Zn(NO3)2· 6H2O). Anal. calcd for C51H44N7O11Zn (Mr = 996.32): C, 61.43, H, 4.45, and N, 9.84%. Found: C, 60.97, H, 4.54, and N, 9.63%. IR spectrum (KBr disc, cm−1): 3419(m, br), 3076(m),

(4)

In eq 4, the symbol χ0 represents the temperature-independent magnetic susceptibility and is fixed using the value obtained from the fit by eq 2. The best fit using eq 3 and eq 4 produced a g of 2.02 and J of −0.52 K. The g factor value is quite reasonable for the Cu2+ ion, and the small negative J value also indicates the existence of weak AFM coupling interaction within a dimer. The corresponding χm(dimer)T versus T is plotted in Figure 7b, which also shows the presence of weak antiferromagnetic interaction within the dimer of Cu2(bpy)4(BDC)2+.



CONCLUSIONS In summary, we have achieved two novel cocrystal compounds, which contain correspondingly Zn2+/Cu2+ coordination cations with the uncoordinated ligands, using a hydrothermal reaction. In cocrystal 1, the partially deprotonated H2BDC dimers of [H(BDC)2]3− and [H3(BDC)2]− form 3D HOF with channels, and the [Zn(bpy)3]2+ cations together the uncoordinated bpy reside in channels; there are strong π···π stacking interactions between the uncoordinated and coordinated bpy molecules and between the coordinated bpy molecules. Cocrystal 2 shows the layered structure with alternating anionic and cationic layers, the dimers of Cu2(bpy)4(BDC)2+ are comprised of the cationic layers together with the uncoordinated bpy molecules, and the BDC2− ions are consistent of anionic layers together with 12234

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237

ACS Omega

Article

Table 2. Crystallographic Data and Structure Refinement Details of 1 and 2 compound

1

2

empirical formula CCDC deposit no. Fw (g·mol−1) crystal system space group crystal color crystal size temperature (K) a (Å) b (Å) c (Å) α (°) β (°) γ (°) V (Å3) Z Dc (g·cm−3) F(000) reflections collected/unique (Rint) observed reflections refinement method data/restraints/parameters goodness-of-fit on F2 final R factor wR2

C51H44N7O11Zn 1525050 996.32 monoclinic P21/n pink 0.30 × 0.20 × 0.10 mm 296(2) 12.5732(17) 13.1278(18) 28.584(4) 90 98.849(2) 90 4661.9(11) 4 1.419 2068 41116/10647 Rint = 0.0551 8140 full-matrix least squares on F2 10647/6/656 1.054 0.0444 0.1158

C66H48N10O8Cu2 1525309 1236.26 triclinic P-1 green 0.30 × 0.20 × 0.10 mm 293(2) 14.5367(16) 16.3208(18) 17.3434(19) 94.183(3) 104.754(3) 114.463(2) 3546.7(7) 2 1.158 1272 43305/16138 Rint = 0.0375 11529 full-matrix least squares on F2 16138/125/794 1.030 0.0583 0.1466



1698(s), 1605(m), 1562(m), 1474(m), 1442(m), 775(s), and 736(s). 1H NMR data (400 MHz, DMSO, 25 °C, TMS, δ): =7.47−8.70(m, 38H, Hpheyl‑ring and Hpyrid‑ring). Preparation of [Cu2(BDC)(bpy)4]BDC·bpy·2H2O (2). This cocrystal compound was prepared using the similar procedure to that of 1, instead of Zn(NO3)2·6H2O by Cu(NO3)2·3H2O (0.18 g, 0.75 mmol). Dark green crystals suitable for singlecrystal X-ray structure analysis were obtained after 3 days and dried in air yielding 0.20 g (65% yield calculated using the reactant Cu(NO 3 ) 2 ·3H 2 O) of 2. Anal. calcd for C66H52Cu2N10O10 (Mr = 1272.26): C, 62.30, H, 4.12, and N, 11.01%. Found: C, 61.99, H, 4.52, and N, 11.63%. IR spectrum (KBr disc, cm−1): 3431(m, br), 3074(m), 1606(s), 1560(m), 1473(m), 1444(m),765(s), and 718(s). X-ray Crystallography. Single-crystal X-ray diffraction data of 1 and 2 were collected on a Siemens SMART-CCD diffractometer with graphite monochromatic Mo Kα radiation (λ = 0.71073 Å). The structures were solved by direct methods and refined on F2 using full-matrix least-squares methods with SHELXTL package.58 All nonhydrogen atoms were refined anisotropically, and hydrogen atoms were theoretically added, riding on their parent atoms. It is worth mentioning that two heavily disordered and diffused outer-sphere water molecules were found in the structure of 2, which were removed by the program SQUEEZE in Platon during the structural refinements.59 Crystallographic data and structural refinement details of 1 and 2 are summarized in Table 2. Selected bond lengths and angles are tabulated in the Supporting Information (Table S1). The Diamond 3.1 program was used for the creation of figures and analysis of hydrogen bonding interactions in the crystal lattice.60

ASSOCIATED CONTENT

* Supporting Information S

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.9b01584. Selected bond length (Å) and bond angle (°) for 1 and 2; IR spectra of 1 and 2; 1H NMR spectrum of 1; cationic layer of [Cu2(bpy)4(BDC)]2+ ions with uncoordinated bpy molecules and anionic layer containing BDC2− anions and heavily disordered lattice water molecules, which are removed from the anionic layer in 2; the powder X-ray diffraction patterns for 1 and 2; emission decay of 1; plots of temperature-dependent molar magnetic susceptibility of 2 (PDF) Accession Codes

CCDC 1525050 and 1525309 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, by emailing [email protected], or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, U.K.; Fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*E-mail: [email protected]. Tel: +86 25 58139527. Fax: +86 25 58139988 (X.-S.G.). *E-mail: [email protected] (X.-M.R.). ORCID

Xu-Sheng Gao: 0000-0001-7184-759X Xiao-Ming Ren: 0000-0003-0848-6503 Notes

The authors declare no competing financial interest. 12235

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237

ACS Omega



Article

and Co-Crystals that Tune Inter-Electron Exchange. CrystEngComm 2012, 14, 1515−1526. (20) Li, S.; Yan, D. Two-Component Aggregation-Induced Emission Materials: Tunable One/Two-Photon Luminescence and StimuliResponsive Switches by Co-Crystal Formation. Adv. Optical Mater. 2018, 6, 1800445. (21) Fang, X.; Yang, X.; Yan, D. Vapor-phase π−π Molecular Recognition: a Fast and Solvent-free Strategy towards the Formation of Co-Crystalline Hollow Microtube with 1D Optical Waveguide and up-Conversion Emission. J. Mater. Chem. C 2017, 5, 1632−1637. (22) Zhai, X.; Li, S.; Ding, Y.; Pan, L.; Yang, H.; Jiang, B.; Yan, D.; Meng, Q. Fabrication and Investigation of Two-Component Film of 2,5-Diphenyloxazole and Octafluoronaphthalene Exhibiting Tunable Blue/Bluish Violet Fluorescence Based on Low Vacuum Physical Vapor Deposition Method. J. Nanomater. 2016, 1−8. (23) Yan, D.; Evans, D. G. Molecular Crystalline Materials with Tunable Luminescent Properties: from Polymorphs to MultiComponent Solids. Mater. Horiz. 2014, 1, 46−57. (24) Arenas-García, J. I.; Herrera-Ruiz, D.; Mondragón-Vásquez, K.; Morales-Rojas, H.; Höpfl, H. Modification of the Supramolecular Hydrogen-Bonding Patterns of Acetazolamide in the Presence of Different Cocrystal Formers: 3:1, 2:1, 1:1, and 1:2 Cocrystals from Screening with the Structural Isomers of Hydroxybenzoic Acids, Aminobenzoic Acids, Hydroxybenzamides, Aminobenzamides, Nicotinic Acids, Nicotinamides, and 2,3-Dihydroxybenzoic Acids. Cryst. Growth Des. 2012, 12, 811−824. (25) Koshima, H.; Miyamoto, H.; Yagi, I.; Uosaki, K. Preparation of Co-Crystals of 2-Amino-3-Nitropyridine with Benzenesulfonic Acids for Second-Order Nonlinear Optical Materials. Cryst. Growth Des. 2004, 4, 807−811. (26) Koshima, H.; Masaki, N.; Asahi, T. Optical Activity Induced by Helical Arrangements of Tryptamine and 4-Chlorobenzoic Acid in Their Co-Crystal. J. Am. Chem. Soc. 2005, 127, 2455−2463. (27) Loehlin, J. H.; Etter, M. C.; Gendreau, C.; Cervasio, E. Hydrogen-Bond Patterns in Several 2:1 Amine-Phenol Cocrystals. Chem. Mater. 1994, 6, 1218−1221. (28) Remenar, J. F.; Morissette, S. L.; Peterson, M. L.; Moulton, B.; Macphee, J. M.; Guzmán, H. R.; Almarsson, Ö . Crystal Engineering of Novel Co-Crystals of a Triazole Drug with 1,4-Dicarboxylic Acids. J. Am. Chem. Soc. 2003, 125, 8456−8457. (29) Puigjaner, C.; Barbas, R.; Portell, A.; Valverde, I.; Vila, X.; Alcobé, X.; Font-Bardia, M.; Prohens, R. A Co-Crystal Is the Key Intermediates for the Production of a New Polymorph of Vorinostat. CrystEngComm 2012, 14, 362−365. (30) Ohba, S.; Hosomi, H.; Ito, Y. In Situ X-ray Observation of Pedal-like Conformational Change and Dimerization of transCinnamamide in Cocrystals with Phthalic Acid. J. Am. Chem. Soc. 2001, 123, 6349−6352. (31) Keyes, T. E.; Forster, R. J.; Bond, A. M.; Miao, W. Electron Self-Exchange in the Solid-State: Co-Crystals of Hydroquinone and Bipyridyl Triazole. J. Am. Chem. Soc. 2001, 123, 2877−2884. (32) Allen, F. H. The Cambridge Structural Database: a Quarter of a Million Crystal Structures and Rising. Acta Cryst. 2002, B58, 380− 388. (33) Biswas, A.; Ghosh, M.; Lemoine, P.; Sarkar, S.; Hazra, S.; Mohanta, S. Syntheses, Structures and Magnetic Properties of Trinuclear CuIIMIICuII (M = Cu, Ni, Co and Fe) and Tetranuclear [2×1+1×2] CuIIMnII-2CuII Complexes Derived from a Compartmental Ligand: The Schiff Base 3-Methoxysalicylaldehyde Diamine Can also Stabilize a. Eur. J. Inorg. Chem. 2010, 3125−3134. (34) Das, L. K.; Biswas, A.; Gómez-García, C. J.; Drew, M. G. B.; Ghosh, A. Isolation of Two Different Ni2Zn Complexes with an Unprecedented Co-Crystal Formed by One of Them and a ″Coordination Positional Isomer″ of the Other. Inorg. Chem. 2014, 53, 434−445. (35) Golbedaghi, R.; Salehzadeh, S.; Khavasi, H. R.; Blackman, A. G. Mn(II) Complexes of Three [2+2] Macrocyclic Schiff Base Ligands. Synthesis and X-ray Crystal Structure of the First Binuclear− Di(binuclear) Co-Crystal. Polyhedron 2014, 68, 151−156.

ACKNOWLEDGMENTS Authors thank the Priority Academic Program Development of Jiangsu Higher Education Institutions and the National Nature Science Foundation of China (grant no. 21601084) for financial support.



REFERENCES

(1) Bond, A. D. What Is a Co-Crystal? CrystEngComm 2007, 9, 833−834. (2) Aitipamula, S.; Chow, P. S.; Tan, R. B. H. Polymorphism in CoCrystals: a Review and Assessment of Its Significance. CrystEngComm 2014, 16, 3451−3465. (3) Aitipamula, S.; Chow, P. S.; Tan, R. B. H. Polymorphs and Solvates of a Co-Crystal Involving an Analgesic Drug, Ethenzamide, and 3, 5-Dinitrobenzoic Acid. Cryst. Growth Des. 2010, 10, 2229− 2238. (4) Babu, N. J.; Nangia, A. Solubility Advantage of Amorphous Drugs and Pharmaceutical Co-Crystals. Cryst. Growth Des. 2011, 11, 2662−2679. (5) Braga, D.; Grepioni, F.; Lampronti, G. I.; Maini, L.; Turrina, A. Ionic Co-Crystals of Organic Molecules with Metal Halides: A New Prospect in the Solid Formulation of Active Pharmaceutical Ingredients. Cryst. Growth Des. 2011, 11, 5621−5627. (6) Brittain, H. G. Cocrystal Systems of Pharmaceutical Interest: 2010. Cryst. Growth Des. 2012, 12, 1046−1054. (7) Brittain, H. G. Co-Crystal Systems of Pharmaceutical Interest: 2011. Cryst. Growth Des. 2012, 12, 5823−5832. (8) Prohens, R.; Barbas, R.; Portell, A.; Font-Bardia, M.; Alcobé, X.; Puigjaner, C. Polymorphism of Co-Crystals: The Promiscuous Behavior of Agomelatine. Cryst. Growth Des. 2016, 16, 1063−1070. (9) Surov, A. O.; Solanko, K. A.; Bond, A. D.; Perlovich, G. L.; Bauer-Brandl, A. Crystallization and Polymorphism of Felodipine. Cryst. Growth Des. 2012, 12, 4022−4030. (10) Thakuria, R.; Delori, A.; Jones, W.; Lipert, M. P.; Roy, L.; Rodríguez-Hornedo, N. Pharmaceutical Co-Crystals and Poorly Soluble Drugs. Int. J. Pharm. 2013, 453, 101−125. (11) Dohle, W.; Prota, A. E.; Menchon, G.; Hamel, E.; Steinmetz, M. O.; Potter, B. V. L. Tetrahydroisoquinoline Sulfamates as Potent Microtubule Disruptors: Synthesis, Antiproliferative and Antitubulin Activity of Dichlorobenzyl-Based Derivatives, and a Tubulin Cocrystal Structure. ACS Omega 2019, 4, 755−764. (12) Joshi, M.; Choudhury, A. R. Salts of Amoxapine with Improved Solubility for Enhanced Pharmaceutical Applicability. ACS Omega 2018, 3, 2406−2416. (13) Almarsson, Ö .; Zaworotko, M. J. Crystal Engineering of the Composition of Pharmaceutical Phases. Do Pharmaceutical CoCrystals Represent a New Path to Improved Medicines? Chem. Commun. 2004, 1889−1896. (14) Karki, S.; Frišcǐ ć, T.; Fabian, L.; Laity, P. R.; Day, G. M.; Jones, W. Improving Mechanical Properties of Crystalline Solids by CoCrystal Formation: New Compressible Forms of Paracetamol. Adv. Mater. 2009, 21, 3905−3909. (15) Yan, D.; Delori, A.; Lloyd, G. O.; Frišcǐ ć, T.; Day, G. M.; Jones, W.; Lu, J.; Wei, M.; Evans, D. G.; Duan, X. A Co-Crystal Strategy to Tune the Luminescent Properties of Stilbene-Type Organic SolidState Materials. Angew. Chem., Int. Ed. 2011, 50, 12483−12486. (16) Sonoda, Y.; Goto, M.; Tsuzuki, S.; Tamaoki, N. Fluorinated Diphenylpolyenes: Crystal Structures and Emission Properties. J. Phys. Chem. A 2007, 111, 13441−13451. (17) Hori, A.; Takatani, S.; Miyamotoa, T. K.; Hasegawa, M. Luminescence from π-π Stacked Bipyridines through Arene− Perfluoroarene Interactions. CrystEngComm 2009, 11, 567−569. (18) Tarai, A.; Baruah, J. B. Conformation and Visual Distinction between Urea and Thiourea Derivatives by an Acetate Ion and a Hexafluorosilicate Co-Crystal of the Urea Derivative in the Detection of Water in Dimethylsulfoxide. ACS Omega 2017, 2, 6991−7001. (19) Akpinar, H.; Mague, J. T.; Novak, M. A.; Friedman, J. R.; Lahti, P. M. Radicals Organized by Disk Shaped Aromatics − Polymorphism 12236

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237

ACS Omega

Article

(36) Hadadzadeh, H.; Mansouri, G.; Rezvani, A. R.; Khavasi, H. R. A Novel 1:1 Co-Crystal of Bis(1,10-Phenanthroline)(1,10-Phenanthroline-5,6-Dione)Nickel(II) Hexafluorophosphate and Tris(1,10Phenanthroline)Nickel(II) Hexafluorophosphate Complexes, [Ni(phen)2(phendione)][Ni(phen)3](PF6)4. J. Chem. Crystallogr. 2012, 42, 486−493. (37) Wang, X. L.; Guo, Z. C.; Liu, G. C.; Qu, Y.; Yang, S.; Lin, H. Y.; Zhang, J. W. Tuning the Lead Complexes Based on a Double 1,10Phenanthroline Derivative with Versatile Coordination Behavior by Dicarboxylates: from 0D Nano-Ring to an Unprecedented 0D + 3D Cocrystal. CrystEngComm 2013, 15, 551−559. (38) Mukherjee, P.; Drew, M. G. B.; Gomez-Garcia, C. J.; Ghosh, A. (Ni2), (Ni3), and (Ni2 + Ni3): a Unique Example of Isolated and Cocrystallized Ni2 and Ni3 Complexes. Inorg. Chem. 2009, 48, 4817− 4827. (39) Lee, Y. M.; Hong, S. J.; Kim, H. J.; Lee, S. H.; Kwak, H.; Kim, C.; Kim, S. J.; Kim, Y. Anion Effect on Construction of Zinc(II) Coordination Polymer with a Chelating Ligand 2,2′-Dipyridylamine (Hdpa): Novel Heterogeneous Catalytic Activities. Inorg. Chem. Commun. 2007, 10, 287−291. (40) Massoud, S. S.; Ledet, C. C.; Junk, T.; Bosch, S.; Comba, P.; Herchel, R.; Hošek, J.; Trávníček, Z.; Fischer, R. C.; Mautner, F. A. Dinuclear Metal(II)-Acetato Complexes Based on Bicompartmental 4-Chlorophenolate: Syntheses, Structures, Magnetic Properties, DNA Interactions and Phosphodiester Hydrolysis. Dalton Trans. 2016, 45, 12933−12950. (41) Shi, Q.; Wu, S. Y.; Qiu, X. T.; Sun, Y. Q.; Zheng, S. T. Three Viologen-Derived Zn-Organic Materials: Photochromism, Photomodulated Fluorescence, and Inkless and Erasable prints. Dalton Trans. 2019, 48, 954−963. (42) Qi, Y. J.; Wang, Y. J.; Li, X. X.; Zhao, D.; Sun, Y. Q.; Zheng, S. T. Two d10 Metal−Organic Frameworks as Low-Temperature Luminescent Molecular Thermometers. Cryst. Growth Des. 2018, 18, 7383−7390. (43) Chen, X.-M.; Wang, R.-Q.; Yu, X.-L. Tris(2,2′-bipyridine)zinc(II) Perchlorate. Acta Cryst. 1995, C51, 1545−1547. (44) Zentner, C. A.; Lai, H. W. H.; Greenfield, J. T.; Wiscons, R. A.; Zeller, M.; Campana, C. F.; Talu, O.; Fitzgerald, S. A.; Rowsell, J. L. C. High Surface Area and Z’ in a Thermally Stable 8-Fold Polycatenated Hydrogen-Bonded Framework. Chem. Commun. 2015, 51, 11642−11645. (45) Shanmugan, S.; Cani, D.; Pescarmona, P. P. The Design and Synthesis of an Innovative Octacarboxy-Silsesquioxane Building Block. Chem. Commun. 2014, 50, 11008−11011. (46) Luo, J.; Wang, J. W.; Zhang, J. H.; Lai, S.; Zhong, D. C. Hydrogen-Bonded Organic Frameworks: Design, Structures and Potential Applications. CrystEngComm 2018, 20, 5884−5898. (47) Huang-Fu, Y. J.; Chen, X. Y.; Yang, W.; Bai, Y.; Dang, D. B. 1,4Naphthalenedicarboxylic Acid Functionalized Phosphomolybdate: Synthesis, Crystal Structure and Optical Properties. Mater. Lett. 2015, 155, 48−50. (48) Hu, C.; Noll, B. C.; Piccoli, P. M. B.; Schultz, A. J.; Schulz, C. E.; Scheidt, W. R. Hydrogen Bonding Effects on the Electronic Configuration of Five-Coordinate High-Spin Iron(II) Porphyrinates. J. Am. Chem. Soc. 2008, 130, 3127−3136. (49) Qi, B.; Guo, X.; Gao, Y.; Li, D.; Luo, J.; Li, H.; Eghtesadi, S. A.; He, C.; Duan, C.; Liu, T. Strong Co-Ion Effect via Cation−π Interaction on the Self-Assembly of Metal-Organic Cationic Macrocycles. J. Am. Chem. Soc. 2017, 139, 12020−12026. (50) Zhuo, C.-C.; Li, L.; Hu, C.-J.; Lang, J.-P. Host-Guest Assembly for Highly Sensitive Probing of a Chiral Mono-Alcohol with a Zinc Trisporphyrinate. Sci. Rep. 2017, 7, 3829−3835. (51) Keller, S.; Prescimone, A.; Constable, E. C.; Housecroft, C. E. Dinuclear [Cu2(N^N)(P^P)2][PF6]2 Complexes Containing Bridging 2,3,5,6-tetra(pyridin-2-yl)pyrazine or 2,4,6-Tri(pyridin-2-yl)-1,3,5Triazine Ligands. Polyhedron 2016, 116, 3−11. (52) Wałęsa-Chorab, M.; Marcinkowski, D.; Kubicki, M.; Hnatejko, Z.; Patroniak, V. The Formation of Mononuclear Iron(II) and

Zinc(II) Complexes and Dinuclear Mesocates of Copper(II) with Pyrazine-Bis(bipyridine) Ligand. Polyhedron 2016, 118, 1−5. (53) Laachir, A.; Guesmi, S.; Saadi, M.; Ammari, L. E.; Mentré, O.; Vezin, H.; Colis, S.; Bentiss, F. Copper(II) Coordination Chain Complex with the 2,5-Bis(2-Pyridyl)-1,3,4-Thiadiazole Ligand and an Asymmetric μ2-1,1-Azido double-bridged: Synthesis, Crystal Structure and Magnetic Properties. J. Mol. Struct. 2016, 1123, 400−406. (54) Noro, S.; Akutagawa, T.; Nakamura, T. Selective Separation of Larger Molecules from a Lewis-Base Mixture by Flexible OneDimensional Cu(II) Coordination Polymer with Shape-Recognizing Space. Chem. Commun. 2010, 46, 3134−3136. (55) Wałęsa-Chorab, M.; Stefankiewicz, A. R.; Ciesielski, D.; Hnatejko, Z.; Kubicki, M.; Kłak, J.; Korabik, M. J.; Patroniak, V. New Mononuclear Manganese(II) and Zinc(II) Complexes with a Terpyridine Ligand: Structural, Magnetic and Spectroscopic properties. Polyhedron 2011, 30, 730−737. (56) Eom, G. H.; Park, H. M.; Hyun, M. Y.; Jang, S. P.; Kim, C.; Lee, J. H.; Lee, S. J.; Kim, S. J.; Kim, Y. Anion Effects on the Crystal Structures of ZnII Complexes Containing 2,2′-bipyridine: Their Photoluminescence and Catalytic Activities. Polyhedron 2011, 30, 1555−1564. (57) Krausz, E.; Riesen, H.; Rae, A. E. Crystal Structures of and Polarized Absorption Spectroscopy in Racemic and Resolved [Ru(bpy)3](ClO4)2 and [Zn(bpy)3] (ClO4)2: bpy = 2,2’-Bipyridine. Aust. J. Chem. 1995, 48, 929−954. (58) Sheldrick, G. M. A Short History of SHELX. Acta Crystallogr., Sect. A: Found. Crystallogr. 2008, 64, 112−122. (59) Spek, A. L. The Netherlands, PLATON; Utrecht University: Utrecht, 2010. (60) Brandenburg, K. Diamond, version 3.2k; Crystal Impact: Bonn, Germany 2014.

12237

DOI: 10.1021/acsomega.9b01584 ACS Omega 2019, 4, 12230−12237