Structural Transformation of Photoreactive Helical Coordination

Aug 28, 2018 - Department of Chemistry and Research Institute, SRM Institute of Science and Technology , Kattankulathur , Tamil Nadu 603203 , India. C...
0 downloads 0 Views 4MB Size
Article Cite This: Cryst. Growth Des. 2018, 18, 6221−6226

pubs.acs.org/crystal

Structural Transformation of Photoreactive Helical Coordination Polymers to Two-Dimensional Structures Published as part of the Crystal Growth and Design Israel Goldberg Memorial virtual special issue Bibhuti Bhusan Rath,† Goutam Kumar Kole,*,‡ and Jagadese J. Vittal*,† †

Department of Chemistry, National University of Singapore, Singapore 117543 Department of Chemistry and Research Institute, SRM Institute of Science and Technology, Kattankulathur, Tamil Nadu 603203, India

Downloaded via KAOHSIUNG MEDICAL UNIV on October 10, 2018 at 08:43:07 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.



S Supporting Information *

ABSTRACT: Four helical one-dimensional coordination polymers (CPs) have been synthesized using trans-5styrylpyrimidine (5-Spym), triphenylphosphine, and silver(I) salts of NO3− (1), BF4− (2), ClO4− (3), and F3CCO2− (4) anions. The styryl group attached to the pyrimidine ligand makes all the helical CPs photoreactive. The helical chains are aligned in parallel in the solid state such that the 5-Spym arms of the adjacent helical chains are juxtaposed in a head-to-tail manner readily to undergo [2 + 2] photo-cycloaddition reaction. The loss of guest solvents in the solids during the photoreaction prevents the single-crystal-to-single-crystal reaction. However, all four helical chains undergo [2 + 2] cycloaddition under UV light quantitatively to two-dimensional (2D) coordination polymers. The photodimer obtained from the nitrate salt (1) was recrystallized for single crystal X-ray crystallography to confirm the photoproduct (5) structurally. Compound 5 has the 2D coordination polymeric structure comprising helical rods connecting the cyclobutane rings to form a honeycomb topology.



+ 2] cycloaddition reaction forming a helical thread.36 However, no photoreactive helical 1D CPs has been known, to the best of our knowledge. Here we report a simple strategy to synthesize photoreactive helical 1D CPs that can undergo facile solid state [2 + 2] photo-cycloaddition reaction. The formation of cyclobutane rings stitches the helical 1D CPs to 2D sheets. The formation of the 2D sheet structure was proven by single crystal X-ray crystallography (SCXRD) after recrystallizing one of the photoproducts to obtain single crystals.

INTRODUCTION Metal−ligand coordination bonds have been successfully exploited in organizing molecular building blocks into various interesting and fascinating architectures in the field of coordination polymers (CPs) and metal−organic frameworks (MOFs) compounds.1−3 These materials with desirable structures, properties, and functionality can be readily made by a careful selection of appropriate ligands and metal ions for various applications.4−9 Among these, one-dimensional (1D) coordination polymers are the simplest type comprising linear, zigzag, ribbon, ladder, and helical structures.10 Of these, the helical structural motif has generated immense interest due to its similarities to biological moieties such as DNA, α-helixes, and proteins.11−13 Helical 1D CPs can be assembled from either chiral or achiral building blocks.14 The presence of anions has a great influence on the conformation of the helical structure and pitch.15−17 Ever since works of Schmidt and co-workers on the solid [2 + 2] photocycloaddition reaction appeared about 50 years ago,18 solid state photochemistry has blossomed into the forefront research of solid state chemistry,19−23 and later this was extended to metal complexes, coordination polymers, and metal organic frameworks.24−33 Although helical polymers are known, they were rarely found to be photoreactive, especially for the [2 + 2] cycloaddition reaction.24,28−35 An interesting organic molecular nanoaggregate has been cross-linked via a [2 © 2018 American Chemical Society



RESULTS AND DISCUSSION Helical 1D CPs are commonly designed using bis(pyridyl) ligands with a helical twist at the backbone.14−17 Here we have used a pyrimidine derivative, where the two “N” donor centers are predisposed at an angle of 120° providing a helical twist (Scheme 1), instead of a bis(pyridyl) moiety. The bulkier triphenylphosphine ligand is expected to provide steric crowdedness around Ag(I), which in turn can twist the two Spym ligands relative to each other to facilitate the helical turn of the 1D strand. The photoreactive styryl group is attached to the pyrimidine ligand to enable head-to-tail alignment of the Received: July 18, 2018 Revised: August 19, 2018 Published: August 28, 2018 6221

DOI: 10.1021/acs.cgd.8b01087 Cryst. Growth Des. 2018, 18, 6221−6226

Crystal Growth & Design

Article

Compound [(Ph3P)Ag(5-Spym)(NO3)]·1.5MeOH, 1, crystallized in the space group C2/c with Z = 8. Crystals 2−4 were found to have 0.5 THF, 0.5 MeOH, and 0.5 H2O respectively as lattice solvents in their asymmetric units. The asymmetric unit contains the formula of the compound. The silver(I) atom has distorted tetrahedral geometry from a phosphorus atom of triphenylphosphine ligand, an oxygen atom of nitrate anion, and nitrogen atoms from two different 5-Spym ligands (Figure 1a). The connectivity of the 5-Spym ligands to Ag(I) atoms generates helical chains, which propagate along the b-axis (Figure 1b). The pitch of the [Ag(5-Spym)] helical chain is the length of the b-axis that varies from 9.044 Å for 1 to 9.228 Å for 4. However, there is no correlation between the size of the anions to the pitch length that could be found in this work as found elsewhere.17 The 5-Spym arms spread along the c-axis, and the adjacent helical chains are aligned in parallel such that the 5-Spym ligands from adjacent chains are aligned in a head-to-tail manner as shown in Figure 1c,d. The centers of the pyrimidine and phenyl groups are separated by 3.866 Å, indicating strong π···π interaction, which enforces the olefin bonds to align parallel with a distance of 3.632 Å between the centers of the olefin bonds. Details of the similar geometrical parameters of the head-to-tail aligned 5-Spym ligands in 2−4 are given in Table 1. As these geometrical parameters satisfy the Schmidt’s criteria, the CPs 1−4 are expected to be photoreactive. The formation of cyclobutane rings is expected to stitch the helical chains yielding a 2D coordination polymer.

Scheme 1. : Designing Helical Coordination Polymers Using Pyrimidyl Functionality and Bulky Auxiliary Ligand PPh3a

a

Here X = NO3, BF4, ClO4, and F3CCO2 anions.

resultant 5-styrylpyrimidine ligands through π···π interactions. The following 1D CPs with different semicoordinating anions have been synthesized by the slow evaporation method in moderate to good yield: [(Ph3P)Ag(5-Spym)(X)] (where 5Spym = trans-5-styrylpyrimidine; X = NO3 (1), BF4 (2), ClO4 (3), and F3CCO2 (4)). All of these complexes show broad resonance, in the 31P NMR spectra in CDCl3 solution, in the range of 15.11−11.56 ppm. The broad signals indicate the dynamic behavior due to the labile nature of the Ag−P bond in solution, and the solid state structures of these CPs may not be maintained in solution.37−39 The solid state structures of all the compounds have been determined by SCXRD techniques. All the crystals are isomorphous and isostructural, and hence the structure of 1 will be described here. The structural details of 2−4 can be found in Supporting Information, Figures S1−S5.

Figure 1. (a) A view of 1 showing the coordination geometry of silver(I) atom and partial numbering scheme. Symmetry code: (A) 1/2 − x, 1/2 + y, 11/2 − z. (b) Ag-pyrimidine chain showing the helicity. (c) Packing of 1 viewed along the b-axis showing the helical chains and the relative alignments of the 5-Spym ligands. Anions, hydrogen atoms, and guest solvents are omitted for clarity. (d) A portion of the packing highlighting the head-to-tail alignments of the adjacent 5-Spym ligands from neighboring helical polymers. Only relevant atoms are shown. 6222

DOI: 10.1021/acs.cgd.8b01087 Cryst. Growth Des. 2018, 18, 6221−6226

Crystal Growth & Design

Article

formations from 1D to 2D CPs have been reported40,41 in the literature, and a non-SCSC photochemical solid state structural transformation by [2 + 2] cycloaddition reaction has also been described.42 Hence, 1D CP to 2D CP structural transformations are still not well-explored.

Table 1. Some Selected Geometrical Parameters Related to Photoreactivity

Xtl

distance between centers of the olefin pairs, Å

distance between the centers of 5-Spym and phenyl rings, Å

1 2 3 4

3.632 3.919 3.763 3.977

3.866 3.921 3.836 4.181

symmetry operator 0.5 0.5 0.5 0.5

− x, 0.5 − y, 1 − z − x, 0.5 − y, 1 − z − x, 0.5 − y, 1 − z + x, 0.5 − y, 1.5 − z

interplanar angles between pyrimidine and phenyl rings, °



CONCLUSION In summary, several photoreactive helical 1D CPs have been synthesized and structurally characterized, for the first time, to the best of our knowledge. The combination of 5-pyrimidine derivative and [Ag(I)-PPh3]+ cation has been used to generate helical chains, and the styryl group attached to the pyrimidine ligand makes it photoreactive. It may be noted that the helices were not resolved, and both right- (P) and left-handed (M) helices are present in all the crystal structures. These photoreactive chains are aligned in parallel such that the 5Spym arms on the helical chains are aligned in a head-to-tail manner. A limited number of anions have been employed to make these crystals without any guest solvents to cleanly undergo SCSC reactions, but were not successful. However, all the four helical chains undergo [2 + 2] cycloaddition under UV light quantitatively leading to the formation of 2D coordination polymer as depicted in Figure 3. The photodimer from 1 was recrystallized to obtain single crystals to confirm the photoproducts structurally. Compound 5 has the 2D coordination polymeric structure comprising helical rods connecting the cyclobutane rings to form a honeycomb topology. Future work will focus on making chiral 2D CPs as catalysts for enantioselective synthesis of small organic molecules.

15.6 11.0 8.7 10.0

The photoreactivity of all the compounds in the powdered form was carried out under LUZCHEM UV light. The course of the photoreaction was monitored using 1H NMR spectroscopy by taking out the UV-irradiated samples at regular intervals of time and dissolved in DMSO-d6 solvent (Figures S18−21, Supporting Information). The disappearance of the proton signals of the pyrimidine group at 9.1 ppm, and the appearance of the pyrimidine proton peaks at 8.8 and 8.6 as well as the cyclobutane peaks at ca 4.76 ppm were monitored to follow the percentage of photoconversion with time. Indeed, quantitative photoconversion was observed in less than 8 min for 3 to 50 h for 2 (Figures S19 and 20, Supporting Information), and all the 5-Spym ligands were converted to rctt-1,3-bis(5′-pyrimidyl)-2,4-bis(phenyl)cyclo-butane (rcttbpcb). From the alignment of the 5-Spym ligands in the crystal structures of 1−4, it is likely that the photoreaction occurs in a head-to-tail fashion. Furthermore, we would like to gather the solid state structural evidence from X-ray crystallography. All the crystals 1−4 we tried to grow had guest solvent in the lattice. These solvents were lost easily, and the single crystals cracked severely, when we attempted to carry out singlecrystal-to-single-crystal (SCSC) photoreaction. Hence we have recrystallized the photoproduct to confirm the formation of the cyclobutane in the solid state and to prove the formation of a two-dimensional (2D) coordination polymeric structure in the solid state [2 + 2] photo-cycloaddition reaction. We were successful in obtaining suitable single crystals of the photodimer of 1 (now 5) from DMSO solution. The photodimer [{(Ph3P)Ag(NO3)}2(rctt-bpcb)], 5 (rctt-bpcb = 1,3-bis(5′pyrimidyl)-2,4-bis(phenyl)cyclobutane) crystallized in the monoclinic space group P21/n with Z = 4. The asymmetric unit has one formula unit as shown in Figure 2a. The formation of rctt-bpcb confirms the head-to-tail photoproduct, the regio-isomer as expected from the alignment of the 5-Spym ligands (Figure 2). Further, the formation of a 2D sheet structure resulting from the cyclobutane ligand rctt-bpcb confirms the structural transformation of the helical 1D chain to 2D structures upon photoreaction. In 5, the [Ag(pyrimidine)] helical chain is extended along the b-axis. The pitch of the helix, 10.088 Å, which is the length of the b-axis, is elongated as compared to that in 1, probably due to the strained cyclobutane rings flanking the helix. This is also reflected in the reduction in area and elongation in shape under the helix. As a result, the helix is less pronounced as compared to 1−4. The rctt-bpcb ring is now a four-connected node. Four such rctt-bpcb ligands connect Ag(I) to form a honeycomb-like sheets with (6,3) net topology as shown in Figure 2c. These sheets are packed parallel to the ab plane. The distance between Ag1···Ag1 or Ag2···Ag2 is equal to the length of a-axis. Two examples of solid state structural trans-



EXPERIMENTAL SECTION

Preparation of [(Ph3P)Ag(5-Spym)(NO3)]·1.5CH3OH, 1. Compound 1 was prepared by reacting AgNO3 (8.5 mg, 0.05 mmol), 5SPym (9.1 mg, 0.05 mmol), and PPh3 (13.1 mg, 0.05 mmol) in methanol at room temperature for 1 h. Diffraction-quality, colorless, rod-shaped single crystals were obtained within a few days by slow evaporation of the reaction mixture, which were filtered and air-dried. Yield (43%). Elemental analysis for 1 (%) Calculated: C 57.11, H 4.72, N 6.34; Found: C 56.67, H 4.36, N 6.77; FT-IR (KBr pellet, cm−1): 3450, 3050, 1965, 1636, 1573, 1479, 1435, 1385, 1182, 1095, 1027, 964, 824, 751, 695, 520, 491; 1H NMR (300 MHz, 298 K, DMSO-d6): δH = 9.05 (m, 3H, Pym-H of 5-Spym), 7.64 (m, 2H, PhH of 5-Spym), 7.56 (m, 9H, Ph-H of PPh3), 7.55 (d, 1H, HCCH of 5-Spym), 7.46 (m, 6H, Ph-H of PPh3), 7.43 (m, 2H, Ph-H of 5Spym), 7.34 (t, 1H, Ph-H of 5-Spym), 7.26 (d, 1H, HCCH of 5Spym), 3.17 (s, 3H, methanol); 31P NMR (161.92 MHz, 298 K, CDCl3): δP = 11.83 (broad, Ag-PPh3). Preparation of [(Ph3P)Ag(5-Spym)(BF4)]·0.5THF, 2. Compound 2 was prepared by reacting AgBF4 (9.7 mg, 0.05 mmol), 5SPym (9.1 mg, 0.05 mmol), and PPh3 (13.1 mg, 0.05 mmol) in tetrahydrofuran at room temperature for 1 h. Diffraction-quality, colorless, plate-shaped single crystals were obtained within few days by slow evaporation of the reaction mixture, which were filtered and air-dried. Yield (36%). Elemental analysis for 2 (%) Calculated: C 56.31, H 4.12, N 4.24; Found: C 55.69, H 4.01, N 4.62; FT-IR (KBr pellet, cm−1): 3557, 3504, 1638, 1574, 1480, 1436, 1184, 1065, 967,850, 751, 694, 662, 555, 521, 499, 439; 1H NMR (300 MHz, 298 K, DMSO-d6):): δH = 9.05 (m, 3H, Pym-H of 5-Spym), 7.64 (m, 2H, Ph-H of 5-Spym), 7.56 (m, 9H, Ph-H of PPh3), 7.55 (d, 1H, HC CH of 5-Spym), 7.46 (m, 6H, Ph-H of PPh3), 7.43 (m, 2H, Ph-H of 5-Spym), 7.34 (t, 1H, Ph-H of 5-Spym), 7.26 (d, 1H, HCCH of 5Spym); 31P NMR (161.92 MHz, 298 K, CDCl3): δP = 12.04 (broad, Ag-PPh3) Preparation of [(Ph3P)Ag(5-Spym)(ClO4)]·0.5CH3OH, 3. Compound 3 was prepared by reacting AgClO4 (10.3 mg, 0.05 mmol), 56223

DOI: 10.1021/acs.cgd.8b01087 Cryst. Growth Des. 2018, 18, 6221−6226

Crystal Growth & Design

Article

Figure 2. (a) A ball and stick diagram showing the coordination environment of Ag(I) atoms and selected labeling in 5. (b) A portion of the helical chain flanked by the cyclobutane rings in 5, viewed down from the b-axis. (c) A view of the 2D structure of 5 showing the connectivity. Only selected atoms are shown for clarity. (d) A ball and stick figure of the 2D CP of 5 is shown without hydrogen atoms and disorders. Ph-H of 5-Spym), 7.56 (m, 9H, Ph-H of PPh3), 7.55 (d, 1H, HC CH of 5-Spym), 7.46 (m, 6H, Ph-H of PPh3), 7.43 (m, 2H, Ph-H of 5-Spym), 7.34 (t, 1H, Ph-H of 5-Spym), 7.26 (d, 1H, HCCH of 5Spym); 31P NMR (161.92 MHz, 298 K, CDCl3): δP = 15.11 (broad, Ag-PPh3) Preparation of [(Ph3P)Ag(5-Spym)(CF3CO2)]·0.5 H2O, 4. Compound 4 was prepared by reacting Ag(CF3CO2) (11.0 mg, 0.05 mmol), 5-SPym (9.1 mg, 0.05 mmol), and PPh3 (13.1 mg, 0.05 mmol) in methanol at room temperature for 1 h. Diffraction-quality,

Spym (9.1 mg, 0.05 mmol), and PPh3 (13.1 mg, 0.05 mmol) in methanol at room temperature for 1 h. Diffraction-quality, colorless, prism-shaped single crystals were obtained within a few days by slow evaporation of the reaction mixture, which were filtered and air-dried. Yield (44%). Elemental analysis for 3 (%) Calculated: C 54.85, H 4.07, N 4.19; Found: C 55.38, H 3.99, N 3.45; FT-IR (KBr pellet, cm−1): 3450, 3055, 1893, 1637, 1573, 1480, 1435, 1309, 1181, 1093, 998, 966, 849, 748, 695, 622, 542, 501, 439; 1H NMR (300 MHz, 298 K, DMSO-d6):): δH = 9.05 (m, 3H, pym-H of 5-Spym), 7.64 (m, 2H, 6224

DOI: 10.1021/acs.cgd.8b01087 Cryst. Growth Des. 2018, 18, 6221−6226

Crystal Growth & Design

Article

20.9171(10), b = 10.0188(4), c = 26.2834(12) Å; α = 90, β = 104.727(2), γ = 90°; V = 5327.1(4) Å3; Z = 4; ρcalc = 1.532 g·cm−3; μ = 0.854 mm−1; GOF = 1.098; final R1 = 0.0869; wR2 = 0.2179 [for 9761 data I > 2σ(I)].



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.cgd.8b01087. Materials and general methods, crystallographic data, single crystal XRD structures, PXRD patterns, TGA curves, 1H NMR spectra, 31P NMR spectrum, time dimerization plots, solid state UV−visible absorption spectra, solid state photoluminescence spectra (PDF)

Figure 3. A non-SCSC solid state structural transformation of a photoreactive 1D helical coordination polymer with well-aligned olefin pairs to a coordination polymeric sheet comprising helical strands under UV light via [2 + 2] photo-cycloaddition reaction is depicted. Only relevant atoms are shown for clarity.

Accession Codes colorless, needle-shaped single crystals were obtained within a few days by slow evaporation of the reaction mixture, which were filtered and air-dried. Yield (41%). Elemental analysis for 4 (%) Calculated: C 56.99, H 3.89, N 4.15; Found: C 57.47, H 3.77, N 3.46; FT-IR (KBr pellet, cm−1): 3441, 3507, 1891, 1684, 1569, 1479, 1435, 1309, 1198, 1135, 1096, 1026, 997, 965, 827, 798, 748, 694, 511, 502; 1H NMR (300 MHz, 298 K, DMSO-d6):): δH = 9.05 (m, 3H, Pym-H of 5Spym), 7.64 (m, 2H, Ph-H of 5-Spym), 7.56 (m, 9H, Ph-H of PPh3), 7.55 (d, 1H, HCCH of 5-Spym), 7.46 (m, 6H, Ph-H of PPh3), 7.43 (m, 2H, Ph-H of 5-Spym), 7.34 (t, 1H, Ph-H of 5-Spym), 7.26 (d, 1H, HCCH of 5-Spym); 31P NMR (161.92 MHz, 298 K, CDCl3): δP = 11.56 (broad, Ag-PPh3). [{(Ph3P)Ag(NO3)}2(rctt-bpcb)], 5. Powdered 1 was irradiated inside a LUZCHEM UV box for 15 min for complete photodimerization. The photoproduct was dissolved in DMSO and recrystallized by slow evaporation method at room temperature. Colorless, block-shaped crystals of 5 were obtained within a few days. Yield (87%). Elemental analysis for 5 (%) Calculated: C 58.65, H 4.10, N 6.84; Found: C 58.05, H 3.89, N 6.45; FT-IR (KBr pellet, cm−1): 3455, 3050, 1961, 1890, 1572, 1479, 1435, 1384, 1308, 1179, 1097, 997, 825, 752, 695, 657, 542, 501, 438; 1H NMR (300 MHz, 298 K, DMSO-d6): δ = 8.86 (s, 2H, Pym-H of rctt-bpcb), 8.66 (s, 4H, Pym-H of rctt-bpcb), 7.62 (m, 4H, Ph-H of of rctt-bpcb), 7.56 (m, 18H, Ph-H of PPh3), 7.46 (m, 12H, Ph-H of PPh3), 7.24 (m, 4H, PhH of rctt-bpcb), 7.09 (t, 2H, Ph-H of rctt-bpcb), 4.70 (dd, 4H, cyclobutane protons of rctt-bpcb); 31P NMR (161.92 MHz, 298 K, CDCl3): δP = 13.63 (broad, Ag-PPh3). Crystal Data for 1 at 100(2) K (CCDC 1850455). C 31.5 H 31 AgN 3 O 4.5 P, M = 662.43; monoclinic, C2/c; a = 30.2037(18), b = 9.0444(5), c = 22.8688(14) Å; α = 90, β = 113.614(2), γ = 90°; V = 5724.1(6) Å3; Z = 8; ρcalc = 1.537 g·cm−3; μ = 0.805 mm−1; GOF = 1.068; final R1 = 0.0327; wR2 = 0.0744 [for 7310 data I > 2σ(I)]. Crystal Data for 2 at 298(2) K (CCDC 1850456). C 31 H 27 AgBF 4 N2 O 0.5 P, M = 661.19; monoclinic, C2/c; a = 30.000(4), b = 9.206(3), c = 23.927(4) Å; α = 90, β = 113.237(11), γ = 90°; V = 6072(2) Å3; Z = 8; ρcalc = 1.447 g· cm−3; μ = 0.766 mm−1; GOF = 1.021; final R1 = 0.0663; wR2 = 0.1589 [for 3344 data I > 2σ(I)]. Crystal Data for 3 at 100(2) K (CCDC 1850457). C 30.5 H 27 AgClN 2 O 4.5 P, M = 667.83; monoclinic,C2/c; a = 30.1894(12), b = 9.1675(4), c = 23.1425(9) Å; α = 90, β = 113.7890(10), γ = 90°; V = 5860.8(4) Å3; Z = 8; ρcalc = 1.514 g·cm−3; μ = 0.874 mm−1; GOF = 1.092; final R1 = 0.0487; wR2 = 0.1089 [for 7105 data I > 2σ(I)]. Crystal Data for 4 at 298(2) K (CCDC 1850458). C32H26AgF3N2O2.5P, M = 674.39; monoclinic, C2/c; a = 30.400(6), b = 9.2280(19), c = 24.320(5) Å; α = 90, β = 115.17(3), γ = 90°; V = 6175(3) Å3; Z = 8; ρcalc = 1.451 g·cm−3; μ = 0.755 mm−1; GOF = 0.954; final R1 = 0.0544; wR2 = 0.1141 [for 3246 data I > 2σ(I)]. Crystal Data for 5 at 100(2) K (CCDC 1850459). C 60 H 50 Ag 2 N 6 O 6 P 2 , M = 1228.74; monoclinic, P2 1 /n; a =

CCDC 1850455−1850459 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 Authors

*E-mail: [email protected] (G.K.K.). *E-mail: [email protected] (J.J.V.). Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We thank the Ministry of Education, Singapore, for financial support through NUS FRC grant Tier 1 No. R-143-000-678114 and R-143-000-A12-114. We thank Ms. Geok Kheng Tan of X-ray diffraction laboratory, CMMAC at NUS as well as Mr. Amey P. Wadawale of Babha Atomic Research Center, Mumbai, India, for the X-ray intensity data collection and data processing. We also thank Dr. In-Hyeok Park for generating some figures. G.K.K. is currently at the Julius Maximilians Universität Würzburg, Germany, and gratefully acknowledges the fellowship from the Alexander von Humboldt Foundation.



REFERENCES

(1) Batten, S. R.; Neville, S. M.; Turner, D. R. Coordination Polymers: Design, Analysis and Application; Royal Society of Chemistry, 2009. (2) Hong, M.-C.; Chen, L. Design and Construction of Coordination Polymers; John Wiley & Sons, 2009. (3) MacGillivray, L. R. Metal-Organic Frameworks: Design and Application; John Wiley & Sons, 2010. (4) Schröder, M. Functional Metal-Organic Frameworks: Gas Storage, Separation and Catalysis; Springer, 2010; Vol. 293. (5) Farrusseng, D. Metal-Organic Frameworks: Applications from Catalysis to Gas Storage; John Wiley & Sons, 2011. (6) Ortiz, O. L.; Ramírez, L. D. Coordination Polymers and Metal Organic Frameworks: Properties, Types and Applications; Nova Science Publishers, 2012. (7) i Xamena, F. L.; Gascon, J. Metal Organic Frameworks as Heterogeneous Catalysts; Royal Society of Chemistry, 2013. (8) Zhou, H.-C.; Long, J. R.; Yaghi, O. M. Introduction to Metal− Organic Frameworks; ACS Publications, 2012. (9) Long, J. R.; Yaghi, O. M. The pervasive chemistry of metal− organic frameworks. Chem. Soc. Rev. 2009, 38, 1213−1214. 6225

DOI: 10.1021/acs.cgd.8b01087 Cryst. Growth Des. 2018, 18, 6221−6226

Crystal Growth & Design

Article

(32) MacGillivray, L. R.; Papaefstathiou, G. S.; Frišcǐ ć, T.; Varshney, D. B.; Hamilton, T. D. Template-controlled synthesis in the solidstate. In Templates in Chemistry I; Springer, 2005; pp 201−221. (33) Georgiev, I. G.; MacGillivray, L. R. Metal-mediated reactivity in the organic solid state: from self-assembled complexes to metal− organic frameworks. Chem. Soc. Rev. 2007, 36, 1239−1248. (34) Garai, A.; Sasmal, S.; Biradha, K. Diversity in the Coordination Polymers of 2-(2-(Pyridin-4/3-yl) vinyl)-1 H-benzimidazole and Dicarboxylates/Disulfonates: Photochemical Reactivity and Luminescence Studies. Cryst. Growth Des. 2016, 16, 4457−4466. (35) Garai, M.; Biradha, K. Water-Resistant and Transparent Plastic Films from Functionalizable Organic Polymers: Coordination Polymers as Templates for Solid-State [2+ 2]-Photopolymerization. Chem. - Eur. J. 2017, 23, 273−277. (36) Yamauchi, M.; Ohba, T.; Karatsu, T.; Yagai, S. Photoreactive helical nanoaggregates exhibiting morphology transition on thermal reconstruction. Nat. Commun. 2015, 6, 8936. (37) Lin, Y.-Y.; Lai, S.-W.; Che, C.-M.; Fu, W.-F.; Zhou, Z.-Y.; Zhu, N. Structural Variations and Spectroscopic Properties of Luminescent Mono- and Multinuclear Silver(I) and Copper(I) Complexes Bearing Phosphine and Cyanide Ligands. Inorg. Chem. 2005, 44, 1511−1524. (38) Gardinier, J. R.; Hewage, J. S.; Lindeman, S. V. Isomer Dependence in the Assembly and Lability of Silver(I) Trifluoromethanesulfonate Complexes of the Heteroditopic Ligands, 2-, 3-, and 4-[Di(1H-pyrazolyl)methyl]phenyl(di-p-tolyl)phosphine. Inorg. Chem. 2014, 53, 12108−12121. (39) Kole, G. K.; Vivekananda, K. V.; Kumar, M.; Ganguly, R.; Dey, S.; Jain, V. K. Hemilabile silver(I) complexes containing pyridyl chalcogenolate (S, Se) ligands and their utility as molecular precursors for silver chalcogenides. CrystEngComm 2015, 17, 4367−4376. (40) Aslani, A.; Morsali, A. Crystal-to-crystal Transformation from a Chain Polymer to a Two-Dimensional Network by Thermal Desolvation. Chem. Commun. 2008, 3402−3404. (41) Coronado, E.; Giménez-Marqués, M.; Mínguez Espallargas, G. Combination of Magnetic Susceptibility and Electron Paramagnetic Resonance to Monitor the 1D to 2D Solid State Transformation in Flexible Metal−Organic Frameworks of Co(II) and Zn(II) with 1,4Bis(triazol-1-ylmethyl)benzene. Inorg. Chem. 2012, 51, 4403−4410. (42) Peedikakkal, A. M. P.; Vittal, J. J. Solid-State Photochemical Behavior of a Triple-Stranded Ladder Coordination Polymer. Inorg. Chem. 2010, 49, 10−12.

(10) Leong, W. L.; Vittal, J. J. One-dimensional coordination polymers: complexity and diversity in structures, properties, and applications. Chem. Rev. 2011, 111, 688−764. (11) Yashima, E.; Ousaka, N.; Taura, D.; Shimomura, K.; Ikai, T.; Maeda, K. Supramolecular Helical Systems: Helical Assemblies of Small Molecules, Foldamers, and Polymers with Chiral Amplification and Their Functions. Chem. Rev. 2016, 116, 13752−13990. (12) Yashima, E.; Maeda, K.; Iida, H.; Furusho, Y.; Nagai, K. Helical polymers: synthesis, structures, and functions. Chem. Rev. 2009, 109, 6102−6211. (13) Kim, H.-J.; Zin, W.-C.; Lee, M. Anion-directed self-assembly of coordination polymer into tunable secondary structure. J. Am. Chem. Soc. 2004, 126, 7009−7014. (14) Han, L.; Hong, M. Recent advances in the design and construction of helical coordination polymers. Inorg. Chem. Commun. 2005, 8, 406−419. (15) Yuan, G.; Zhu, C.; Liu, Y.; Xuan, W.; Cui, Y. Anion-driven conformational polymorphism in homochiral helical coordination polymers. J. Am. Chem. Soc. 2009, 131, 10452−10460. (16) Chen, X.-D.; Mak, T. C. Single-strand helical complexes constructed from 2-pyridinyl-3-pyridinylmethanone: tuning the helical pitch length by variation of metal cation and/or counter anion. Dalton Trans. 2005, 3646−3652. (17) Jung, O.-S.; Kim, Y. J.; Lee, Y.-A.; Park, J. K.; Chae, H. K. Smart molecular helical springs as tunable receptors. J. Am. Chem. Soc. 2000, 122, 9921−9925. (18) Schmidt, G. Photodimerization in the solid state. Pure Appl. Chem. 1971, 27, 647−678. (19) Ramamurthy, V.; Venkatesan, K. Photochemical reactions of organic crystals. Chem. Rev. 1987, 87, 433−481. (20) Desiraju, G. R. Organic Solid State Chemistry; Elsevier, 1987. (21) Ramamurthy, V.; Mondal, B. Supramolecular photochemistry concepts highlighted with select examples. J. Photochem. Photobiol., C 2015, 23, 68−102. (22) Gan, M.-M.; Yu, J.-G.; Wang, Y.-Y.; Han, Y.-F. TemplateDirected Photochemical [2+ 2] Cycloaddition in Crystalline Materials: A Useful Tool to Access Cyclobutane Derivatives. Cryst. Growth Des. 2018, 18, 553−565. (23) Ramamurthy, V.; Sivaguru, J. Supramolecular photochemistry as a potential synthetic tool: photocycloaddition. Chem. Rev. 2016, 116, 9914−9993. (24) 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. (25) Biradha, K.; Santra, R. Crystal engineering of topochemical solid state reactions. Chem. Soc. Rev. 2013, 42, 950−967. (26) Nagarathinam, M.; Vittal, J. J. A rational approach to crosslinking of coordination polymers using the photochemical [2+ 2] cycloaddition reaction. Macromol. Rapid Commun. 2006, 27, 1091−1099. (27) Kole, G. K.; Vittal, J. J. Solid-state reactivity and structural transformations involving coordination polymers. Chem. Soc. Rev. 2013, 42, 1755−1775. (28) Medishetty, R.; Vittal, J. J. Metal-organic frameworks for photochemical reactions. In Metal-Organic Frameworks for Photonics Applications; Springer, 2013; pp 105−144. (29) Huang, S.-L.; Hor, T. A.; Jin, G.-X. Photodriven single-crystalto-single-crystal transformation. Coord. Chem. Rev. 2017, 346, 112− 122. (30) Chanthapally, A.; Vittal, J. J. Metal-Organic Framework: Photoreactive Frameworks. In Metal-Organic Framework Materials; MacGillivray, L. R., Lukehart, C. M., Eds.; John Wiley & Sons, Ltd: Chichester, UK, 2014; pp 135−157. (31) Vittal, J. J.; Quah, H. S. Photochemical reactions of metal complexes in the solid state. Dalton Trans. 2017, 46, 7120−7140. 6226

DOI: 10.1021/acs.cgd.8b01087 Cryst. Growth Des. 2018, 18, 6221−6226