Photosalient Behavior of Photoreactive Zn(II ... - ACS Publications

Apr 22, 2019 - Khushboo Yadava and Jagadese J. Vittal*. Department of Chemistry, National University of Singapore, Singapore 117543. •W Web-Enhanced...
0 downloads 0 Views 2MB Size
Communication pubs.acs.org/crystal

Cite This: Cryst. Growth Des. XXXX, XXX, XXX−XXX

Photosalient Behavior of Photoreactive Zn(II) Complexes Published as part of a Crystal Growth and Design virtual special issue Remembering the Contributions and Life of Prof. Joel Bernstein Khushboo Yadava and Jagadese J. Vittal* Department of Chemistry, National University of Singapore, Singapore 117543

Crystal Growth & Design Downloaded from pubs.acs.org by UNIV PARIS-SUD on 04/23/19. For personal use only.

W Web-Enhanced Feature * S Supporting Information *

ABSTRACT: Solids showing unusual and unexpected properties are the key toward the development of new advanced functional materials. Here, we describe two different types of Zn(II) complexes obtained from three different derivatives of 4-styrylpyridine under the same experimental conditions: [Zn3(cin)6(2F-4spy)2·Zn2(cin)4(2F-4spy)2] (1), [Zn2(cin)4(4spy)2] (2), and [Zn2(cin)4(3F-4spy)2] (3), where cin = cinnamate, 2F-4spy = 2-fluoro-4′styrylpyridine and 3F-4spy = 3-fluoro-4′-styrylpyridine. Of these 1 is a photostable inorganic cocrystal comprising a trinuclear Zn(II) complex and a paddlewheel Zn(II) complex, whereas 2 and 3 are photoreactive Zn(II) complexes with paddlewheel structure. The neighboring olefin bonds in the styrylpyridine ligands in 2 and 3 are aligned in a head-to-tail manner, but not the olefin bonds in the cinnamate ligands. The distance between the pairs of olefin bonds are 4.348(2) Å in 2 and 4.045(2) Å in 3. However, they not only undergo solid state photochemical [2 + 2] cycloaddition reaction under UV light but also show the photosalient behavior during the photoreaction. An overall increase in the percentage cell volume of 11.93% and 10.13% for 2 and 3, respectively, during the photoreaction appears to indicate a significant anisotropic expansion is responsible for the photosalient effect. ome crystalline materials, known as “dynamic crystals”, have been found to respond to external stimuli with mechanical movements ranging from changing its shape like curling, bending, and twisting and/or undergoing locomotion such as jumping, flipping, and rotation.1−10 Such crystals are potential candidates for the development of artificial muscles, actuators, and biomimetic and technomimetic smart materials.11−18 The dynamic single crystals that can drive themselves over distances several times larger than their own size under light are known as photosalient crystals.1−3 Solid-state photochemical [2 + 2] cycloaddition reactions have been recognized as an environmentally benign, green method to synthesize cyclobutane derivatives.19−21 The crystal engineering principles have been successfully employed in controlling the crystal packing to align the olefin bond pairs present in organic, metal−organic, and coordination polymeric compounds in parallel to satisfy the Schmidt’s topochemical criteria.22−28 Of these, some photoreactive metal complexes and organic solids have been found not only to undergo photochemical reaction under UV light but also are accompanied by propulsion resembling popping of corn on a hot pan, known as photosalient (PS) effect.29−37 In other cases, the PS effect has been found to accompany intramolecular linkage isomerization,38 ring opening and closing reaction,39−42 phase transition to a polymorph,43 and dynamic changes in pseudorotaxane molecules.44 We and others have been interested in developing metal complexes exhibiting the hetero [2 + 2] cycloaddition reaction

S

© XXXX American Chemical Society

between two different olefin containing ligands to explore their synthetic sustainability of cyclobutane derivatives in the solid state.45−47 Chloro-methyl group exchange has been cleverly used to bring out the “hetero” or “cross” cycloaddition reaction in organic solids.48−50 Incorporation of two different olefin containing ligands may also promote a “double-dimerization” reaction but is not observed. As a part of this ongoing program, we have employed both carboxylate and N-containing neutral ligands having olefin bonds in the backbone to maximize the hetero [2 + 2] cycloaddition reaction or “double-dimerization” reaction. For this purpose, the cinnamate anion with three different derivatives of 4-styrylpyridine have been used to synthesize three Zn(II) complexes. Of these [Zn3(cin)6(2F4spy)2·Zn2(cin)4(2F-4spy)2] (1) (where cin = cinnamate, 2F4spy = 2-fluoro-4′-styrylpyridine and 3F-4spy = 3-fluoro-4′styrylpyridine) is a photostable inorganic cocrystal comprising a linear-trinuclear Zn(II) complex and a dimeric Zn(II) complex with a paddlewheel structure. Under the same experimental conditions, when 4-styrylpyridine (4spy) or 3fluoro-4′-styrylpyridine (3F-4spy) was used, photoreactive complexes [Zn2(cin)4(4spy)2] (2) and [Zn2(cin)4(3F-4spy)2] (3) with a paddlewheel structure were isolated. These photoreactive complexes exhibit the unexpected photosalient property while undergoing a photochemical [2 + 2] cycloReceived: February 27, 2019 Revised: April 2, 2019

A

DOI: 10.1021/acs.cgd.9b00260 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

Figure 1. (a) A view of the trinuclear Zn(II) complex in the inorganic cocrystals, 1 showing selected labeling of atoms. (b) A view of the paddlewheel structure in 1. (c) The packing of the two different Zn(II) complexes present in 1 highlighting the nonalignment of the olefin pairs. The symmetry label A is −x, 2 − y, 1 − z.

Figure 2. Packing of the paddlewheel structure in 2 showing the head-to-tail alignment of the 4-spy ligands.

structure is shown in Figure 1b. The central Zn(II) atom in the Zn3 cluster is at (1/2, 1/2, 0) and the center of the paddlewheel is at (0, 0, 1/2) positions in the unit cell. Hence the Zn3 clusters are packed in the ab plane, and Zn2 dimers are packed in between the Zn3 layers (Figure 1c). The paddlewheel compounds are under the confined space as shown in Figure 1b, and hence the olefin pairs of 2F-4spy ligands are not aligned. In fact, none of the olefin pairs have been found to satisfy Schmidt’s criteria for [2 + 2] cycloaddition reaction, and hence 1 has been found to be photostable. Cocrystals are single phase crystalline compounds having two or more different molecular or ionic compounds in a stoichiometric ratio which are neither solvates nor simple salts.51−53 These are usually observed in organic compounds and are very important in pharmaceutical industries and high

addition reaction under UV light. The details are described below. The asymmetric unit of 1 contains a half formula unit. The crystallographic center of inversion produces a trinuclear [Zn3(cin)6(2F-4spy)2] complex and a dinuclear [Zn2(cin)4(2F-4spy)2] complex in the solid. In the Zn3 trimer, all three Zn(II) atoms are arranged in a linear fashion by virtue of the crystallographic inversion symmetry, in which the central Zn2 is surrounded by six oxygen atoms from different cinnamate ligands to have a highly distorted octahedral environment, while the terminal Zn3 atoms have a distorted tetrahedral ZnO3N core from three cinnamate ligands and a 2F-4spy ligand. Of the three cinnamate ligands between two Zn(II) atoms, two are bridged via two oxygen atoms (κ2-O1,O2 bonding mode), while one is bridged through one oxygen atom (μ2-O mode) as shown in Figure 1a. The paddlewheel B

DOI: 10.1021/acs.cgd.9b00260 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

reaction are anticipated roughly in the (101) plane. When the MAX-150 Xenon Light Source 150W was used to irradiate a big, flat single crystal, cracks developed roughly in the (101) plane. When we collected the data from the single crystal portion of the cracked crystal, we found that no significant photoreaction occurred. However, a comparison of the crystal data revealed that there was an increase in the cell lengths of aand c-axes of about 0.3%, but the b-axis length decreased by 0.3%. This corroborated our previous observation of cracks in the irradiated single crystal (Table 1).

energy solids. A few inorganic cocrystals have been reported in the literature.54−65 Its implications and potential applications are not obvious at present and may be realized when more examples have been discovered. Both [Zn2(cin)4(4spy)2] (2) and [Zn2(cin)4(3F-4spy)2] (3) are isotypical, and hence their structures are described together. Both are centrosymmetric dimer crystallized in the triclinic space group P1̅ with Z = 1 with a half formula unit present in the asymmetric unit. The 4-styrylpyridine ligand is not planar. The pyridine and the phenyl rings are twisted by 34.9(1)° in 2 and 16.4(1)° in 3. In both the structure π···π interactions between pyridyl and phenyl groups are not evident. Further, the paddlewheel structures are spanned approximately along the body diagonal of the unit cell. The styrylpyridine ligands from the adjacent dimers are aligned in a head-to-tail (HT) manner (Figure 2). Nevertheless, the distance between the centroids of the olefin bonds is 4.348(2) Å in 2 due to the nonplanarity of the two sixmembered rings in the 4-spy ligand. As a result, the centroid of the pyridine ring is closer to the C10 of the adjacent phenyl ring by 3.391(2) Å. Because of the lesser twisting of the 3F4spy ligand, the distance between the centroids of the olefin bonds is 4.045(2) Å in 3. These conformational differences observed in the isotypical structures may be attributed to the presence of fluorine atom in 3. The Schmidt’s criteria for the solid-state [2 + 2] cycloaddition reaction is not strictly followed in 3 and very scarcely in 2.69 However, when we tested them under UV light, both of them were found to be photoreactive. The photoreactivity of powdered 2 under UV light (351 nm) in a LUZCHEM LZC4 photoreactor was investigated by taking out the UV-irradiated samples at regular intervals of time and dissolved in DMSO-d6 solvent to record 1H NMR spectra (Figure S11 in SI). A quantitative photoconversion was observed in 4 h (Figure S13a in SI). The disappearance of the olefin protons signals at 7.51 and 7.11 ppm, the shift of the signal for the pyridyl protons from 8.37 and 7.85 to 8.27 and 7.59 ppm respectively, and the appearance of the cyclobutane protons at 4.56 ppm confirmed 100% conversion of 4spy ligands to rctt-ppcb (where rctt-ppcb = 1,2-bis(4′-pyridyl)-3,4bis(phenyl)cyclobutane) cyclobutane rings in 2a. The photoreactivity of 3 was quite analogous to 2 (Figure S12 in SI), and a quantitative photoconversion of 3F-4spy was observed in less than 2 h (Figure S13b in SI). The photoproduct 3a is expected to have rctt-3F-ppcb (where rctt-3F-ppcb = 1,2-bis(4′-pyridyl)3,4-bis(3′-fluoro-phenyl)cyclobutane). It is not very surprising that 2 and 3 were found to be photoreactive in the solid state, although the Schmidt’s rules were not strictly followed. Small rotation66−68 and movements69−72 of the molecules can cause these olefin pairs to come closer to satisfy the Schmidt’s criteria for [2 + 2] photoreaction in the solid state. What is really surprising is their photosalient behavior. When the single crystals 2 and 3 were irradiated with UV light using MAX-150 Xenon Light Source 150 W, they started spurting violently in pieces while undergoing [2 + 2] cycloaddition reaction, displaying photosalient behavior (Videos SV1 and SV2). As a result, the crystals turned into microcrystalline powders. Such similar behavior during in the solid-state photoreaction has been reported by us before.29−31 A close examination of the crystal packing of 3 (Figure S6) reveals that the olefin pairs approximately run along the body diagonal of the unit cell, and the formation of cyclobutane rings and the conformational changes associated with this

Table 1. Changes in Cell Lattice Parameters of 3 upon UVIrradiation 3

before UV (100 K)

before UV (RT)

after UV irradiation until cracks (RT)

% increase or decrease after UV (RT)

a (Å) b (Å) c (Å) α (deg) β (deg) γ (deg) V (Å3)

10.110(1) 11.825(1) 12.664(1) 67.456(4) 71.279(4) 72.420(4) 1296.3(3)

10.130(2) 11.988(2) 12.752(2) 67.307(2) 72.789(2) 72.084(5) 1331(1)

10.1614(4) 11.9521(5) 12.7878(5) 67.243(1) 72.633(1) 71.827(1) 1332.4(1)

0.31% increase 0.29% decrease 0.28% increase 0.065° decrease 0.156° decrease 0.256° decrease 0.08% increase

We have determined the densities of the photoproducts 2a and 3a by the flotation method, and from these values the cell volumes were calculated (Table S2). We found a considerable increase of cell volume of 11.93% in 2a and 10.13% in 3a. A rapid change of phase of 2 and 3 into photoproducts 2a and 3a facilitates the conversion of the accumulated strain energy into kinetic energy. The percentage increase in cell volumes observed here is similar to the previous photosalient reactions which varies from 7.6% to 17.6%.29,30 It is likely that a sudden increase of cell lengths is responsible for such increased percentage volume to account for the photosalient effect. Hence, a sudden anisotropic volume expansion is an important factor for popping of crystals under UV light. The anisotropic expansion of the crystal unit cell accompanied by the nucleation and propagation of a new phase within the bulk crystal may be responsible for the photosalient behavior.38 It appears that the self-actuation results from the buildup of strain at the phase interface, if larger than the cohesive interactions present within the crystal lattice. In conclusion, we isolated a single phase crystalline inorganic cocrystal, 1 and two paddlewheel Zn(II) complexes, 2 and 3. The cocrystal 1 is photostable, while 2 and 3 are photoreactive as well as photosalient. Despite the presence of olefin groups on the cinnamate and 4spy ligands, only the olefin groups in the 4spy ligands were found to be photoreactive. Photoreaction of 2 and 3 led to the formation of one-dimensional coordination polymers comprising cyclobutane linkages. The photosalient nature appears to be due to the strain created by the sudden anisotropic volume expansion during [2 + 2] photo-cycloaddition reaction. Upon UV irradiation on single crystals, the cracks developed in the (101) plane. Upon crystal structure analysis, the development of cracks was correlated to the alignment of the olefin pairs which run approximately along the body diagonal of the unit cell, and hence, the formation of cyclobutane rings and the conformational changes associated with this reaction are anticipated roughly in the (101) plane. The determination of densities of the photoproducts confirms a significant increase in cell volume during C

DOI: 10.1021/acs.cgd.9b00260 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

(8) Krishna, G. R.; Devarapalli, R.; Lal, G.; Reddy, C. M. Mechanically Flexible Organic Crystals Achieved by Introducing Weak Interactions in Structure: Supramolecular Shape Synthons. J. Am. Chem. Soc. 2016, 138, 13561−13567. (9) Hayashi, S.; Koizumi, T. Elastic Organic Crystals of a Fluorescent π-Conjugated Molecule. Angew. Chem., Int. Ed. 2016, 55, 2701−2704. (10) Ghosh, S.; Reddy, C. M. Elastic and bendable caffeine cocrystals: Implications for the design of flexible organic materials. Angew. Chem., Int. Ed. 2012, 51, 10319−10323. (11) Abendroth, J. M.; Bushuyev, O. S.; Weiss, P. S.; Barrett, C. J. Controlling Motion at the Nanoscale: Rise of the Molecular Machines. ACS Nano 2015, 9, 7746−7768. (12) Seki, T.; Ito, H. Molecular-Level Understanding of Structural Changes of Organic Crystals Induced by Macroscopic Mechanical Stimulation. Chem. - Eur. J. 2016, 22, 4322−4329. (13) Terao, F.; Morimoto, M.; Irie, M. Light-Driven MolecularCrystal Actuators: Rapid and Reversible Bending of Rodlike Mixed Crystals of Diarylethene Derivatives. Angew. Chem., Int. Ed. 2012, 51, 901−904. (14) Morimoto, M.; Irie, M. A Diarylethene Cocrystal that Converts Light into Mechanical Work. J. Am. Chem. Soc. 2010, 132, 14172− 14178. (15) Kitagawa, D.; Kobatake, S. Photoreversible current ON/OFF switching by the photoinduced bending of gold-coated diarylethene crystals. Chem. Commun. 2015, 51, 4421−4424. (16) Rai, R.; Krishnan, B. P.; Sureshan, K. M. Chirality-controlled spontaneous twisting of crystals due to thermal topochemical reaction. Proc. Natl. Acad. Sci. U. S. A. 2018, 115, 2896−2901. (17) MacGillivray, L. R.; Papaefstathiou, G. S.; Frišcǐ ć, T.; Hamilton, T. D.; Bučar, D.-K.; Chu, Q.; Varshney, D. B.; Georgiev, I. G. Supramolecular Control of Reactivity in the Solid State: From Templates to Ladderanes to Metal-Organic Frameworks. Acc. Chem. Res. 2008, 41, 280−291. (18) Cheng, S. C.; Chen, K. J.; Suzaki, Y.; Tsuchido, Y.; Kuo, T. S.; Osakada, K.; Horie, M. Reversible Laser-Induced Bending of Pseudorotaxane Crystals. J. Am. Chem. Soc. 2018, 140, 90−93. (19) Ramamurthy, V.; Sivaguru, J. Supramolecular Photochemistry as a Potential Synthetic Tool: Photocycloaddition. Chem. Rev. 2016, 116, 9914−9993. (20) Elacqua, E.; Frišcǐ ć, T.; MacGillivray, L. R. [2.2]Paracyclophane as a Target of the Organic Solid State: Emergent Properties via Supramolecular Construction. Isr. J. Chem. 2012, 52, 53−59. (21) Dutta, S., Georgiev, I. G.; MacGillivray, L. R. Metal-Organic Frameworks with Photochemical Building Units in Metal-Organic Frameworks: Design and Application. In Metal-Organic Frameworks; MacGillivray, L. R., Ed.; John Wiley & Sons: New Jersey, 2010; Chapter 10. (22) Biradha, K.; Santra, R. Crystal engineering of topochemical solid state reactions. Chem. Soc. Rev. 2013, 42, 950−967. (23) Georgiev, I. G.; MacGillivray, L. R. Metal-mediated reactivity in the organic solid state: from self-assembled complexes to metalorganic frameworks. Chem. Soc. Rev. 2007, 36, 1239−1248. (24) Desiraju, G. R. Organic Solid State Chemistry; Elsevier: Amsterdam; New York, 1987. (25) Ramamurthy, V.; Venkatesan, K. Photochemical reactions of organic crystals. Chem. Rev. 1987, 87, 433−481. (26) Contents: Macromol. Rapid Commun. 14/2006. Macromol. Rapid Commun. 2006, 27 1083−1090. (27) Nagarathinam, M.; Peedikakkal, A. M. P.; Vittal, J. J. Stacking of double bonds for photochemical [2 + 2] cycloaddition reactions in the solid state. Chem. Commun. 2008, 42, 5277−5288. (28) Vittal, J. J.; Quah, H. S. Photochemical reactions of metal complexes in the solid state. Dalton Trans. 2017, 46, 7120−7140. (29) Medishetty, R.; Husain, A.; Bai, Z.; Runcevski, T.; Dinnebier, R. E.; Naumov, P.; Vittal, J. J. Single Crystals Popping Under UV Light: A Photosalient Effect Triggered by a [2 + 2] Cycloaddition Reaction. Angew. Chem., Int. Ed. 2014, 53, 5907−5911.

photosalient reactions. Such dynamic crystals convert light into mechanical work during a solid-state chemical reaction. Understanding the underlining mechanism involved in both photoreactive and photosalient properties is useful in developing smart advanced materials.



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.9b00260. Materials and general methods, crystallographic data, additional figures, NMR spectral data for the timedependent UV experiments, face-indexed single crystals, cell parameters of UV-irradiated crystals, density measurements, PXRD patterns, TG plots, and solidstate photoluminescent spectra (PDF) W Web-Enhanced Features *

Videos in AVI format are available. Accession Codes

CCDC 1892462−1892464 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

Jagadese J. Vittal: 0000-0001-8302-0733 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS J.J.V. thanks the Ministry of Education, Singapore, for funding this project through NUS FRC Grant No. R-143-000-A12-114. We also thank Ms. Geok Kheng Tan for the collection of X-ray crystallographic data.



REFERENCES

(1) Naumov, P.; Chizhik, S.; Panda, M. K.; Nath, N. K.; Boldyreva, E. Mechanically Responsive Molecular Crystals. Chem. Rev. 2015, 115, 12440−12490. (2) Commins, P.; Desta, I. T.; Karothu, D. P.; Panda, M. K.; Naumov, P. Crystals on the move: mechanical effects in dynamic solids. Chem. Commun. 2016, 52, 13941−13954. (3) Nath, N. K.; Panda, M. K.; Sahoo, S. C.; Naumov, P. Thermally induced and photoinduced mechanical effects in molecular single crystalsa revival. CrystEngComm 2014, 16, 1850−1858. (4) Gupta, P.; Karothu, D. P.; Ahmed, E.; Naumov, P.; Nath, N. K. Thermally Twistable, Photobendable, Elastically Deformable, and Self-Healable Soft Crystals. Angew. Chem., Int. Ed. 2018, 57, 8498− 8502. (5) Ahmed, E.; Karothu, D. P.; Naumov, P. Crystal Adaptronics: Mechanically Reconfigurable Elastic and Superelastic Molecular Crystals. Angew. Chem., Int. Ed. 2018, 57, 8837−8846. (6) Ghosh, S.; Mishra, M. K.; Kadambi, S. B.; Ramamurty, U.; Desiraju, G. R. Designing Elastic Organic Crystals: Highly Flexible Polyhalogenated N-Benzylideneanilines. Angew. Chem., Int. Ed. 2015, 54, 2674−2678. (7) Saha, S.; Desiraju, G. R. Crystal Engineering of Hand-Twisted Helical Crystals. J. Am. Chem. Soc. 2017, 139, 1975−1983. D

DOI: 10.1021/acs.cgd.9b00260 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

a 3D Zn(ii) coordination polymer. Chem. Commun. 2018, 54, 13861− 13864. (48) Jones, W.; Ramdas, S.; Theocharis, C. R.; Thomas, J. M.; Thomas, N. W. Crystal engineering of photodimerizable cyclopentanones. Comparison of chloro- and methyl- substitution as solidstate steering groups. J. Phys. Chem. 1981, 85, 2594−2597. (49) Theocharis, C. R.; Desiraju, G. R.; Jones, W. The use of mixed crystals for engineering organic solid-state reactions: application to benzylbenzylidenecyclopentanones. J. Am. Chem. Soc. 1984, 106, 3606−3609. (50) Bučar, D.-K.; Sen, A.; Mariappan, S. V. S.; MacGillivray, L. R. A [2 + 2] cross-photodimerisation of photostable olefinsvia a threecomponent cocrystal solid solution. Chem. Commun. 2012, 48, 1790− 1792. (51) Aitipamula, S.; Banerjee, R.; Bansal, A. K.; Biradha, K.; Cheney, M. L.; Choudhury, A. R.; Desiraju, G. R.; Dikundwar, A. G.; Dubey, R.; Duggirala, N.; Ghogale, P. P.; Ghosh, S.; Goswami, P. K.; Goud, N. R.; Jetti, R. R. K. R.; Karpinski, P.; Kaushik, P.; Kumar, D.; Kumar, V.; Moulton, B.; Mukherjee, A.; Mukherjee, G.; Myerson, A. S.; Puri, V.; Ramanan, A.; Rajamannar, T.; Reddy, C. M.; Rodriguez-Hornedo, N.; Rogers, R. D.; Row, T. N. G.; Sanphui, P.; Shan, N.; Shete, G.; Singh, A.; Sun, C. C.; Swift, J. A.; Thaimattam, R.; Thakur, T. S.; Kumar Thaper, R.; Thomas, S. P.; Tothadi, S.; Vangala, V. R.; Variankaval, N.; Vishweshwar, P.; Weyna, D. R.; Zaworotko, M. J. Polymorphs, Salts, and Cocrystals: What’s in a Name? Cryst. Growth Des. 2012, 12, 2147−2152. (52) Almarsson, Ö .; Zaworotko, M. J. Crystal engineering of the composition of pharmaceutical phases. Do pharmaceutical co-crystals represent a new path to improved medicines? Chem. Commun. 2004, 17, 1889−1896. (53) Schultheiss, N.; Newman, A. Pharmaceutical Cocrystals and Their Physicochemical Properties. Cryst. Growth Des. 2009, 9, 2950− 2967. (54) Jones, P.; Williams, P. A.; Vagg, R. S. Chiral metal complexes’. 17 cocrystallization of four diastereoisomers: A second example. Inorg. Chem. 1984, 23, 4110−4111. (55) Evans, W. J.; Boyle, T. J.; Ziller, J. W. tructural Diversity in Bis(Pentamethylcyclopentadienyl)Yttrium Chloride Complexes: Cocrystallization of [(C5Me5)2Y(Y-Cl)2Li(THF)2] and [(C5Me5)2YCl(YCl)Li(THF)3]. Inorg. Chem. 1992, 31, 1120−1122. (56) Nayak, M.; Jana, A.; Fleck, M.; Hazra, S.; Mohanta, S. A unique example of a three component cocrystal of metal complexes. CrystEngComm 2010, 12, 1416−1421. (57) Dorazco-González, A.; Martinez-Vargas, S.; Hernández-Ortega, S.; Valdés-Martínez, J. Directed self-assembly of mono and dinuclear copper(ii) isophthalates into 1D polymeric structures. Design and an unusual cocrystallization. CrystEngComm 2013, 15 (31), 5961−5968. (58) Holz, R. C.; Thompson, L. C. Spectroscopically distinct geometrical isomers in a single crystal. Characterization of the eightcoordinate adducts of tris(dipivaloylmethanato)lanthanide(III) with 2,9-dimethyl-1,10-phenanthroline. Inorg. Chem. 1993, 32, 5251−5256. (59) Dean, P. A. W.; Vittal, J. J.; Craig, D. C.; Scudder, M. L. Polytopal Isomerism of the [Cd(S{O}CPh)3]− Anion. Inorg. Chem. 1998, 37, 1661−1664. (60) Palaniandavar, M.; Butcher, R. J.; Addison, A. W. Dipicolylamine Complexes of Copper(II): Two Different Coordination Geometries in the Same Unit Cell of Cu(Dipica)2(BF4)2. Inorg. Chem. 1996, 35, 467−471. (61) Dyson, P. J.; Steed, J. W.; Suman, P. Reproducible Growth of a Neutral Inorganic Co-Crystal: Pd{PPh 2 (C 16 H 15 )} 2 Cl 2 .[Pd{PPh2(C16H15)}Cl2 ]2. Solvate (Solvate = CH 2Cl2 or Et2O). CrystEngComm 1999, 1, 5−7. (62) Hazra, S.; Koner, R.; Nayak, M.; Sparkes, H. A.; Howard, J. A. K.; Dutta, S.; Mohanta, S. Role of Water and Solvent in the Formation of Three Mononuclear Copper(II) Crystals: A New Type of Hydrate Isomerism in Coordination Chemistry. Eur. J. Inorg. Chem. 2009, 2009, 4887−4894. (63) Sarkar, S.; Nayak, M.; Fleck, M.; Dutta, S.; Flörke, U.; Koner, R.; Mohanta, S. Syntheses, Crystal Structures and Mass Spectrometry

(30) Medishetty, R.; Sahoo, S. C.; Mulijanto, C. E.; Naumov, P.; Vittal, J. J. Photosalient Behavior of Photoreactive Crystals. Chem. Mater. 2015, 27, 1821−1829. (31) Mulijanto, C. E.; Quah, H. S.; Tan, G. K.; Donnadieu, B.; Vittal, J. J. Curved crystal morphology, photoreactivity and photosalient behaviour of mononuclear Zn(II) complexes. IUCrJ 2017, 4, 65−71. (32) Chandra Sahoo, S.; Nath, N. K.; Zhang, L.; Semreen, M. H.; AlTel, T. H.; Naumov, P. Actuation based on thermo/photosalient effect: a biogenic smart hybrid driven by light and heat. RSC Adv. 2014, 4, 7640−7647. (33) Commins, P.; Natarajan, A.; Tsai, C.-K.; Khan, S. I.; Nath, N. K.; Naumov, P.; Garcia-Garibay, M. A. Structure-Reactivity Correlations and Mechanistic Understanding of the Photorearrangement and Photosalient Effect of α-Santonin and Its Derivatives in Solutions, Crystals, and Nanocrystalline Suspensions. Cryst. Growth Des. 2015, 15, 1983−1990. (34) Nath, N. K.; Runcevski, T.; Lai, C.-Y.; Chiesa, M.; Dinnebier, R. E.; Naumov, P. Surface and Bulk Effects in Photochemical Reactions and Photomechanical Effects in Dynamic Molecular Crystals. J. Am. Chem. Soc. 2015, 137, 13866−13875. (35) Seki, T.; Ito, H. Photoinduced crystalline structure change and photosalient effect through strengthening metallophilic interaction. Nippon Kessho Gakkaishi 2015, 57, 226−232. (36) Wang, H.; Chen, P.; Wu, Z.; Zhao, Z.; Sun, J.; Lu, R. Bending, Curling, Rolling, and Salient Behavior of Molecular Crystals Driven by [2 + 2] Cycloaddition of a Styrylbenzoxazole Derivative. Angew. Chem., Int. Ed. 2017, 56, 9463−9467. (37) Ravi, A.; Sureshan, K. M. Tunable Mechanical Response from a Crystal Undergoing Topochemical Dimerization: Instant Explosion at a Faster Rate and Chemical Storage of a Harvestable Explosion at a Slower Rate. Angew. Chem., Int. Ed. 2018, 57, 9362−9366. (38) Naumov, P.; Sahoo, S. C.; Zakharov, B. A.; Boldyreva, E. V. Dynamic Single Crystals: Kinematic Analysis of Photoinduced Crystal Jumping (The Photosalient Effect). Angew. Chem., Int. Ed. 2013, 52, 10126−10126. (39) Colombier, I.; Spagnoli, S.; Corval, A.; Baldeck, P. L.; Giraud, M.; Leaustic, A.; Yu, P.; Irie, M. Diarylethene microcrystals make directional jumps upon ultraviolet irradiation. J. Chem. Phys. 2007, 126, 011101. (40) Uchida, K.; Nishimura, R.; Hatano, E.; Mayama, H.; Yokojima, S. Photochromic Crystalline Systems Mimicking Bio-Functions. Chem. - Eur. J. 2018, 24, 8491−8506. (41) Hatano, E.; Morimoto, M.; Hyodo, K.; Yasuda, N.; Yokojima, S.; Nakamura, S.; Uchida, K. Photosalient Effect of a Diarylethene with a Perfluorocyclohexene Ring. Chem. - Eur. J. 2016, 22, 12680− 12683. (42) Hatano, E.; Morimoto, M.; Imai, T.; Hyodo, K.; Fujimoto, A.; Nishimura, R.; Sekine, A.; Yasuda, N.; Yokojima, S.; Nakamura, S.; Uchida, K. Photosalient Phenomena that Mimic Impatiens Are Observed in Hollow Crystals of Diarylethene with a Perfluorocyclohexene Ring. Angew. Chem., Int. Ed. 2017, 56, 12576−12580. (43) Seki, T.; Sakurada, K.; Muromoto, M.; Ito, H. Photoinduced single-crystal-to-single-crystal phase transition and photosalient effect of a gold(I) isocyanide complex with shortening of intermolecular aurophilic bonds. Chem. Sci. 2015, 6, 1491−1497. (44) Chen, K. J.; Chen, P. L.; Horie, M. Dynamic Pseudorotaxane Crystals Containing Metallocene Complexes. Sci. Rep. 2017, 7, 1−12. (45) Quah, H. S.; Yap, J. L. X.; Sambasivam, U.; Chanthapally, A.; Donnadieu, B.; Vittal, J. J. Contrast Solid-State Photoreactive Behavior of Two Two-Dimensional Zn(II) Coordination Polymers. Cryst. Growth Des. 2018, 18, 3693−3696. (46) Medishetty, R.; Bai, Z.; Yang, H.; Wong, M. W.; Vittal, J. J. Influence of Fluorine Substitution on the Unusual Solid-State [2 + 2] Photo-Cycloaddition Reaction between an Olefin and an Aromatic Ring. Cryst. Growth Des. 2015, 15, 4055−4061. (47) Volodin, A. D.; Korlyukov, A. A.; Zorina-Tikhonova, E. N.; Chistyakov, A. S.; Sidorov, A. A.; Eremenko, I. L.; Vologzhanina, A. V. Diastereoselective solid-state crossed photocycloaddition of olefins in E

DOI: 10.1021/acs.cgd.9b00260 Cryst. Growth Des. XXXX, XXX, XXX−XXX

Crystal Growth & Design

Communication

of Mononuclear NiII Inclusion Product and Self-Assembled [2 × 1+1 × 2] NiII3MII (M = Cu, Ni, Co, Fe or Mn) Cocrystals Derived from N,N’-Ethylenebis(3-Ethoxysalicylaldimine). Eur. J. Inorg. Chem. 2010, 2010, 735−743. (64) Nayak, M.; Hazra, S.; Lemoine, P.; Koner, R.; Lucas, C. R.; Mohanta, S. Self-Assembled [2 × 1+1 × 2] Heterotetranuclear CuII3MnII/CuII3Co II and [2 × 2+1 × 3] Heptanuclear Cu II7 Compounds Derived from N,N’-o-Phenylenebis(3-Ethoxysalicylaldimine): Structures and Magnetic Properties. Polyhedron 2008, 27, 1201−1213. (65) Hazra, S.; Sasmal, S.; Nayak, M.; Sparkes, H. A.; Howard, J. A. K.; Mohanta, S. Syntheses and Crystal Structures of CuIIBiIII, CuIIBaIICuII, [CuIIPbII]2 and Cocrystallized (UVIO2)2.4CuII Complexes: Structural Diversity of the Coordination Compounds Derived from N,N’-Ethylenebis(3-Ethoxysalicylaldiimine). CrystEngComm 2010, 12, 470−477. (66) Harada, J.; Ogawa, K. Pedal motion in crystals. Chem. Soc. Rev. 2009, 38, 2244−2252. (67) Harada, J.; Ogawa, K. What molecules are likely or unlikely to undergo pedal motions in crystals? Cryst. Growth Des. 2014, 14, 5182−5188. (68) Harada, J.; Ogawa, K. Torsional Motion of Stilbene-type Molecules in Crystals. Topics in Stereochemistry 2006, 31−47. (69) Schmidt, G. M. J. Photodimerization in the solid state. Pure Appl. Chem. 1971, 27, 647−678. (70) Kaupp, G.; Naimi-Jamal, M. R. Mechanically induced molecular migrations in molecular crystals. CrystEngComm 2005, 7 (66), 402− 410. (71) Kaupp, G. Organic Solid-State Reactions with 100% Yield. In Organic Solid State Reactions; Toda, F., Ed.; Springer: Berlin, 2005; pp 95−183. (72) Kaupp, G.; Schmeyers, J.; Boy, J. Waste-free solid-state syntheses with quantitative yield. Chemosphere 2001, 43, 55−61.

F

DOI: 10.1021/acs.cgd.9b00260 Cryst. Growth Des. XXXX, XXX, XXX−XXX