Hydrogen Bond-driven Assembly of Stilbazolium Dye with Zinc Chloride: X-aggregation Induced Photoluminescence and Dual Stimuliresponsive Performance

Zhao-Hua CHEN Cai-Hua GUO Zhen-Hua LU Qin-Tian YANG

Citation:  Zhao-Hua CHEN, Cai-Hua GUO, Zhen-Hua LU, Qin-Tian YANG. Hydrogen Bond-driven Assembly of Stilbazolium Dye with Zinc Chloride: X-aggregation Induced Photoluminescence and Dual Stimuliresponsive Performance[J]. Chinese Journal of Structural Chemistry, 2020, 39(10): 1841-1848. doi: 10.14102/j.cnki.0254–5861.2011–2714 shu

Hydrogen Bond-driven Assembly of Stilbazolium Dye with Zinc Chloride: X-aggregation Induced Photoluminescence and Dual Stimuliresponsive Performance

English

  • The development of functional organic-inorganic hybrids has attracted great interest due to the potential of combining the distinct properties of organic and inorganic components into a single molecular composite[1, 2]. In this field, more attention was paid to the organic functional materials because their functions can be easily modified by introducing functional groups, and consequently, fascinating physical properties such as optical signal processing, THz generation and luminescence chromisms can be achieved[3-5]. Among various organic functional molecules, particular interest was paid to push-pull π-delocalized stilbazolium-type dyes[6]. Their unique D-π-A structures contain an electron acceptor linked to an electron donor group via a π-conjugated bridge. Consequently, fascinating characters including the large molecular hyperpolarizability (β) and low dielectric constants will be generated[7-9]. So far, by changing electron-donating/withdrawing groups on D/A terminals, many stilbazolium derivatives have been investigated[10-12]. The aggregation modes of stilbazolium-type dyes can determine their optical properties (for example, electronic intramolecular charge transfer, ICT) which can also be modulated by introducing different substituents on the amine groups[13]. Inorganic moieties are also important in achieving function promotion. To our knowledge, the following types of inorganic components have been used to incorporate with stilbazolium-type dyes: (a) metal complexes, for example, Au(CN)2- and MPS3 (M = Mn, Cd, Zn)[10, 14, 15], (b) polymeric metal halides[2, 16] and (c) polyoxometalates[11, 17]. Among them, d10 metal halides have captured special attention due to the stimulus-responsive performances like the X-ray-induced photochromism of [ZnLBr2][18]. To the best of our knowledge, the study about the combination of stilbazolium-type dyes with zinc halides/pseudohalide is rare[19]. In this work, D-π-A stilbazolium-type dye chromophore DANP (DANP = trans-4-(4΄-(N, N-diethylaminostyryl))-N-methyl-pyridinium) was hybridized with zinc chloride to give a new hybrid (DANP-H)(ZnCl4) (1), and its unquenched photoluminescence was induced by X-aggregation mode of (DANP-H)2+ cations. Furthermore, dual stimuli-responsive performance and reversible photo/thermochromic behaviors can be found, which were explained by theoretical calculation.

    All the reactants were of reagent grade and used as purchased. Elemental analyses (C, H, N) were performed on a Perkin-Elemer 240C instrument. IR spectrum was recorded as KBr discs on a Shimadzu IR-408 infrared spectrophotometer in the 400~4000 cm-1 region. The electronic spectrum was taken on a Shimadzu UV-2101 PC UV-Vis scanning spectrophotomete equipped with an integrating sphere at ambient temperature. 1H NMR spectrum was recorded at a Bruker AVIII 400 spectrometer, whose chemical shifts were reported in parts per million (ppm) downfield from tetramethylsilane. The photoluminescence measurement was carried out on an FLS920 fluorescence spectrophotometer.

    2.2.1   Synthesis of (DANP)I

    Trans-4-(4΄-(N, N-diethylaminostyryl))-N-methyl-pyridinium) (DANP)I was synthesized by Knoevenagel reaction (Scheme 1)[20]. 4-Methyl-pyridine (0.9300 g, 10 mmol) and methyl iodide (1.4200 g, 10 mmol) were refluxed in 30 mL methanol for 5 hours to give 1, 4-dimethylpyridin-1-ium iodide. 4-N, N-diethylamino-benzaldehyde (0.1490 g, 10 mmol) was added into the above solution and then refluxed for 5 hours using piperidine (5 drops) as catalyst. The product was washed with methanol for three times. 1H NMR (400 MHz, DMSO, Fig. 1) δ 8.66 (d, J = 6.4 Hz, 2H), 8.03 (d, J = 6.4 Hz, 2H), 7.89 (d, J = 16.1 Hz, 1H), 7.57 (d, J = 8.5 Hz, 2H), 7.12 (d, J = 16.0 Hz, 1H), 6.75 (d, J = 8.6 Hz, 2H), 4.16 (s, 3H), 3.43 (q, J = 6.8 Hz, 4H), 1.13 (t, J = 6.9 Hz, 6H).

    Scheme S1

    Scheme S1.  Synthesis route of (DANP)I

    Figure 1

    Figure 1.  NMR spectrum of (DANP)I
    2.2.1   Synthesis of (DANP-H)(ZnCl4) (1)

    1 was prepared in a solution method. ZnCl2 (0.0681 g, 0.5 mmol) and (DANP)I (0.1996 g, 0.5 mmol) were dissolved in 10 mL methanol, and the solution was stirred for 2 h. Afterwards, the pH of the mixture was adjusted as 2.0 by adding HCl, which was further filtered and kept at room temperature for slow evaporation. Although the I- has stronger coordinated ability than Cl-, the amount of Cl- in the solution is much larger than that of I-. As a result, the chloride was obtained in the final product. Yellow block crystals were obtained after one week (0.1020 g, yield 49.2% based on ZnCl2). Anal. Calcd. for C36H47Cl8N4Zn2 (950.18): C, 45.51; H, 4.98; N, 5.90 %. Found: C, 46.04; H, 4.82; N, 5.81%. IR(cm-1): 3056(w), 2978(m), 2909(w), 1630(s), 1516(s), 1473(s), 1387(m), 1331(m), 1187(s), 1153(s), 1015(m), 984(s), 837(s), 566(s), 517(s).

    The electronic structure calculation based on density function theory (DFT) was conducted without further optimization[21]. During the calculation, wave functions were explained in a plane wave basis set and the spin polarized version of PBE GGA was employed for the exchangecorrelation functional in the CASTEP code[22]. In this case, a 3×3×2 Monkhorst-Pack grid with a total number of 6 k points in the irreducible Brillouin zone was used for the simulation of primitive cell, and the separation is set to 0.05 Å-1 in the calculations. The number of plane waves included in the basis was determined by a cutoff energy Ec of 435.4 eV. The pseudoatomic calculations on Zn-3d104s2, Cl-3s23p5, C-2s22p2 and N-2s22p3 were conducted. The convergent criterion of total energy is set to 2.0 × 10-5 eV/atom.

    A yellow crystal with dimensions of 0.40 × 0.30 × 0.20 mm3 was placed on an APEX II CCD area detector equipped with a graphite-monochromatic Mo radiation (λ = 0.71073 Å) at 296(2) K. A total of 14874 reflections were collected by a ϕ-ω scan mode at room temperature in the range of 1.71≤θ≤27.51º with index ranges of –30≤h≤23, –18≤ k≤17 and –18≤l≤17 including 8272 independent ones (Rint = 0.0252), of which 5552 were observed with I > 2σ(I). The empirical absorption corrections by SADABS were carried out. The structure was solved by direct methods using SHELXS-97 program[23] and refined with SHELXL-97[24] by full-matrix least-squares techniques on F2. All non-hydrogen atoms were refined anisotropically, while the hydrogen atoms were located geometrically and refined isotropically. The final R = 0.0408 and wR = 0.1061 (w = 1/[σ2(Fo2) + (0.0646P)2 + 0.7674P], where P = (Fo2 + 2Fc2)/3), S = 1.016, (Δ/σ)max = 0.000, (Δρ)max = 0.583 and (Δρ)min = –0.337 e/Å3. Selected bond lengths and bond angles are given in Table 1.

    Table 1

    Table 1.  Selected Bond Lengths (Å) and Bond Angles (º)
    DownLoad: CSV
    Bond Dist. Bond Dist. Bond Dist.
    Zn(1)–Cl(1) 2.230(3) Zn(1)–Cl(2) 2.316(3) Zn(1)–Cl(3) 2.258(3)
    Zn(1)–Cl(4) 2.263(3) Zn(2)–Cl(5) 2.316(3) Zn(2)–Cl(6) 2.288(3)
    Zn(2)–Cl(7) 2.251(3) Zn(2)–Cl(8) 2.259(3)
    Angle (°) Angle (°) Angle (°)
    Cl(1)–Zn(1)–Cl(3) 110.89(12) Cl(1)–Zn(1)–Cl(4) 112.77(15) Cl(3)–Zn(1)–Cl(4) 106.44(13)
    Cl(1)–Zn(1)–Cl(2) 108.20(17) Cl(3)–Zn(1)–Cl(2) 111.71(13) Cl(4)–Zn(1)–Cl(2) 106.80(13)
    Cl(7)–Zn(2)–Cl(8) 110.15(13) Cl(7)–Zn(2)–Cl(6) 109.15(14) Cl(8)–Zn(2)–Cl(6) 106.20(13)
    Cl(7)–Zn(2)–Cl(5) 112.38(14) Cl(8)–Zn(2)–Cl(5) 111.76(13) Cl(6)–Zn(2)–Cl(5) 106.93(12)

    1 contains ZnCl42- mononuclear anions and stilbazoliumtype dye chromophore DANP-H2+ cations with centrosymmetric space group of Cc. C–H···Cl hydrogen bonds and C–H···π interactions contribute to the formation of a quasi-3D network. The ZnCl42- mononuclear anion was commonly observed in zinc halide system, in which Zn(II) was in a tetrahedronal environment (Fig. 2a)[25, 26]. The Zn–Cl lengths of 2.288(3)~2.316(3) Ǻ and Cl–Zn–Cl angles of 106.20(13)~112.38(14)° hint a slightly distorted ZnCl4 tetrahedron (Table 1). Two independent DANP-H2+ cations exhibit different configurations. In the first DANP (defined by N(1)), the distance of C=C double bond is 1.233 Å. And those around C=C bond are 1.471 and 1.512 Å, which are consistent with the sp3-hybrid C–C bond lengths (Fig. 2b). These values indicate that the conjugated system has been broken. The C–C and C–N bonds of another dependent DANP (defined by N(3)) are in normal ranges (Fig. 2c), which are consistent with those in literatures[7-9]. The dihedral angles between pyridine and benzene rings is 7.90°, indicating that two planes (pyridine and benzene rings) are approximately coplanar. The distances around C=C double bonds (1.461, 1.330, 1.471 Å) hint the typical conjugated configuration, which is favourable to molecular hyperpolarizability according to the idea of Marder[27]. In the pyridine and amine groups, the sum of the three C–N–C angles with N atom as centers is all close to 360.0°, suggesting that the outstretched methyl and carbon atoms of N-substituents are generally coplanar[2]. Besides, the trans E-configuration with respect to the ethenyl double bonds has the torsion angles of 173.86 and 179.80°. The protonization of amine is deduced from the charge balance. Since there are two crystallographically independent ZnCl42- and DANP moieties, for the sake of keeping charge balance, each of DANP must have two positive charges, so another positive charge should locate on the amine. The independent DANP-H2+ cation (defined by N(1)) is linked into a 1D chain along the b-axis via C(3)– H(3)···Cl(3) and C(17)–H(17B)···Cl(4) hydrogen bonds (Fig. 3a), and the second DANP-H2+ (defined by N(3)) is connected into a 1D chain along b-axis via C(21)–H(21)···Cl(8) and C(35)–H(35A)···Cl(6) hydrogen bonds. These two independent [(DANP-H)(ZnCl4)]n chains are connected into a double chain via C(19)–H(19B)···π (benzene ring) (H···π distance: 2.89 Ǻ, Fig. 3b). Furthermore, neighboring doublechains are linked into a 2-D layer via C(1)–H(1C)···Cl(6) and C(19)–H(19C)···Cl(4) hydrogen bonds. Finally, adjacent layers are extended into a 3D network via C(4)–H(4)···Cl(4) hydrogen bond (Fig. 3c). The shortest centroid distance among benzene/pyridine rings is 4.624(6) Ǻ, hinting the absence of π-π stacking interaction. But the distance of neighboring ethenyl double bonds is 3.60 Ǻ, and the crossing angle of two DANP-H2+ cations is 58.57°, suggesting the presence of X-aggregation (Fig. 2d)[28]. In all, the hydrogenbonding interactions, C–H···π interactions and static electricity effect in the lattice solidify the crystal packing of 1.

    Figure 2

    Figure 2.  (a) Structure of ZnCl42- tetrahedron; (b, c) Structures of two independent DANP; (d) X-aggregation of DANP (H atoms were omitted for clarity)

    Figure 3

    Figure 3.  (a) 1D [(DANP-H)(ZnCl4)]n chain based on C–H···Cl hydrogen bonds among b-axis; (b) Double chain C–H···π interaction; (c) 3D network constructed from C–H···Cl hydrogen bonds

    Table 2

    Table 2.  Hydrogen Bridging Details of 1 (Lengths in Å and Angles in º)
    DownLoad: CSV
    D–H···A d(D–H) d(H···A) d(D···A) ∠DHA Symmetry codes
    C(1)–H(1C)···Cl(6) 0.96 2.66 3.567(13) 157
    C(3)–H(3)···Cl(3) 0.93 2.63 3.543(12) 165
    C(4)–H(4)···Cl(4) 0.93 2.81 3.649(12) 150 –1/2 + x, 1/2 – y, –1/2 + z
    C(17)–H(17B)···Cl(4) 0.97 2.69 3.477(12) 139 x, 1 + y, z
    C(19)–H(19C)···Cl(4) 0.96 2.75 3.657(12) 158
    C(21)–H(21)···Cl(8) 0.93 2.61 3.520(12) 164
    C(35)–H(35A)···Cl(6) 0.97 2.68 3.486(10) 141 x, 1 + y, z

    Solid-state unpolarized UV-Vis absorption spectra of 1 and (DANP)I were recorded from powder samples at room temperature (Fig. 4a). (DANP)I exhibits broad adsorption zones from ultra-violet (250~600 nm) to near-infrared (650~800 nm), which is in agreement with other stilbazolium-type dyes and can be attributed to CT (charge transfer)[11, 29]. But after hybridization with zinc halide, clear blue-shift has occurred, while the near-infrared (600~800 nm) adsorption has vanished. The sharp peaks at 288 and 362 nm correspond to the π-π* transitions of benzene/pyridine rings, which is consistent with the structural analysis that the conjugated system in one DANP has broken. The adsorption zone in 400~600 nm stems from another conjugationmaintained DANP. Besides, a great blue-shift compared with (DANP)I has happened, which might also be led by its X-aggregation mode[28]. One-photon photoluminescence spectra of 1 and (DANP)I measured at room temperature in the solid state are shown in Fig. 4b. Upon excitation, (DANP)I presents a broad red emission at 620 nm (λex = 570 nm), which is in agreement with literatures[2]. After the formation of (DANP-H)(ZnCl4) hybrid, red-shifted emission to 667 nm can be observed, which has also been documented in other stilbazolium-type dyes and their metal complexes[30]. The maintenance of red emission can be ascribed to the X-aggregation of DANP in lattice, which could not quench the fluorescence. The clear red-shifted emission might be led by the hydrogen-bonding formation[31].

    Figure 4

    Figure 4.  (a) Optical diffuse-reflection spectra and (b) solid-state emission spectra of 1 (λex = 625 nm) and (DANP)I (λex = 570 nm)

    Due to the presence of mixed configurations of DANP in lattice, dual stimuli-responsive performance will be expected. To our interest, 1 responses to both photo and thermal stimuli. When it is irradiated by UV light (250 nm) for 3 h, its color changes from yellow to red, and then retunes to the initial color after removing the UV light. Similarly, it becomes dark red when heating to 130 ℃. After cooling to room temperature, it again retunes to the original color (Fig. 5). Therefore, 1 exhibits reversible photo/thermochromisms. The relative low thermochromic temperature of 130 ℃ is lower than those of viologen-based chromic materials (for example, 150 ℃ for {(MV)2[Pb7Br18]}n)[32], which might be led by the presence of non-covalent interactions (C–H···Cl H-bonds and C–H···π interactions) in the lattice. The photochromism could be induced by the conjugation-broken or conjugationmaintained DANPs. In order to disclose which DANP dominates the photochromism behavior, theoretical calculations were executed. The chromic behaviors of viologen-based halometallates have been extensively studied, and the mechanism can be explained as the electrons transfer from p orbital of halides in metal halides to π* antibonding orbitals of organic molecules[33]. The chromic behavior of 1 could also be explained by this mechanism, because DANP is also electronically poor. DFT calculation was conducted to verify this process. The band structure and partial density of states (PDOS) of 1 were given (Fig. 6). The band structure shows a direct band gap at the Brillouin zone center with an energy value of 1.72 eV. PDOS indicates that the valence-band maximum mainly results from the nonbonding states of Cl-3p and π bonding orbitals of DANP, while the π* anti-bonding orbitals of DANP (conjugation-maintained one) account for the in-plane dispersion of the conduction-band minimum. Comparably, the π* anti-bonding orbital conjugation-broken DANP does not appear in the frontier orbitals (right of Fig. 5). Consequently, the photo/thermochromisms can be attributed to the electron transfer from Cl-3p/DANP-π components to π* antibonding orbitals of the conjugation-maintained DANP cations, which is similar to that in viologen-based halometallate systems[34].

    Figure 5

    Figure 5.  Photo/thermochromisms of 1

    Figure 6

    Figure 6.  Band structure (left) and projected density of states (pDOS, right) for 1

    In summary, combination of D-π-A stilbazolium-type dye with zinc chloride results in a new hybrid (DANP-H)(ZnCl4) (1), in which conjugation-broken and conjugation-maintained DANP-H2+ co-exist in the lattice. Furthermore, strong C–H···Cl hydrogen bonds and C–H···π interactions contribute to the formation of a 3D network, in which X-aggregation mode of (DANP-H)2+ can be observed. Consequently, its adsorption spectrum exhibits a blue-shift, and unquenched red photoluminescence can be monitored. Interestingly, dual stimuli-responsive performance and reversible photo/thermochromic behaviors are found, with the mechanism disclosed by theoretical calculation.


    1. [1]

      Cariati, E.; Lucenti, E.; Botta, C.; Giovanella, U.; Marinotto, D.; Righetto, S. Cu(I) hybrid inorga-organic materials with intriguing stimuli responsive and optoelectronic properties. Coord. Chem. Rev. 2016, 306, 566−614. doi: 10.1016/j.ccr.2015.03.004

    2. [2]

      Wang, Y. K.; Zhao, L. M.; Fu, Y. Q.; Chen, Z.; Lin, X. Y.; Wang, D. H.; Li, Y.; Li, H. H.; Chen, Z. R. Combination of N-arylstilbazolium organic nonlinear optical chromophores with iodoargentates: structural diversities and optical properties. Cryst. Growth Des. 2018, 18, 3827−3840. doi: 10.1021/acs.cgd.8b00033

    3. [3]

      Sun, Z. H.; Luo, J. H.; Chen, T. L.; Li, L. N.; Xiong, R. G.; Tong, M. L.; Hong, M. C. Distinct molecular motions in a switchable chromophore dielectric 4-N, N-dimethylamino-4΄-N΄-methylstilbazolium trifluoromethanesulfonate. Adv. Funct. Mater. 2012, 22, 4855−4861. doi: 10.1002/adfm.201201770

    4. [4]

      Pan, F.; Wong, M. S.; Bosshard, C.; Gunter, P. Crystal growth and characterization of the organic salt 4-N, N-dimethylamino-4΄-N-methyl-stilbazolium tosylate (dast). Adv. Mater. 1996, 8, 592−595.

    5. [5]

      Zhao, L. M.; Shen, X. Q.; Zhang, W. T.; Song, K. Y.; Li, H. H.; Chen, Z. R. Encapsulation of stilbazolium-type dyes into layered metal-organic frameworks: solvent-dependent luminescence chromisms and their mechanisms. Inorg. Chem. Front. 2019, 6, 1195–1208. doi: 10.1039/C8QI01332B

    6. [6]

      Uda, K.; Tsuda, Y.; Okada, S.; Yamakado, R.; Sun, L.; Suzuri, Y.; White, M. S.; Furis, M.; Stadler, P.; Dimitriev, O.; Yoshida, T. Concerted photoluminescence of electrochemically self-assembled CuSCN/stilbazolium dye hybrid thin films. ACS Omega. 2019, 4, 4056−4062. doi: 10.1021/acsomega.8b03632

    7. [7]

      Wolf, T.; Niazov-Elkan, A.; Sui, X. M.; Weissman, H.; Bronshtein, I.; Raphael, M.; Wagner, H. D.; Rybtchinski, B. Free-standing nanocrystalline materials assembled from small molecules. J. Am. Chem. Soc. 2018, 140, 4761−4764. doi: 10.1021/jacs.7b13638

    8. [8]

      Karuppanan, N.; Kalainathan, S. A new nonlinear optical stilbazolium family crystal of (E)-1-ethyl-2-(4-nitrostyryl)pyridin-1-ium iodide: synthesis, crystal structure, and its third-order nonlinear optical properties. J. Phys. Chem. C 2018, 122, 4572−4582. doi: 10.1021/acs.jpcc.7b11884

    9. [9]

      Yang, Z.; Mutter, L.; Stillhart, M.; Ruiz, B.; Aravazhi, S.; Jazbinsek, M.; Schneider, A.; Gramlich, V.; Günter, P. Large-size bulk and thin-film stilbazolium-salt single crystals for nonlinear optics and THz generation. Adv. Funct. Mater. 2007, 17, 2018−2023. doi: 10.1002/adfm.200601117

    10. [10]

      Li, D.; Sun, X.; Shao, N.; Wu, J.; Tian, Y. Self-assembly of organic chromosphere cations with inorganic lanthanide(III) complex counterions. Z. Anorg. Allg. Chem. 2014, 640, 2283–2286. doi: 10.1002/zaac.201400116

    11. [11]

      Wu, J.; Hua, G.; Wang, P.; Hao, F.; Zhou, H.; Zhou, A.; Tian, Y.; Jin, B. Organic/polyoxometalate hybridization dyes: crystal structure and enhanced two-photon absorption. Dyes Pigm. 2011, 88, 174–179. doi: 10.1016/j.dyepig.2010.06.004

    12. [12]

      Li, D.; Zhang, Q.; Sun, X.; Shao, N.; Li, R.; Li, S.; Zhou, H.; Wu, J.; Tian, Y. Hydrosoluble two-photon absorbing materials: a series of sulfonated organic inner salts in biological imaging application. Dyes Pigm. 2014, 102, 79–87. doi: 10.1016/j.dyepig.2013.10.028

    13. [13]

      Hestand, N. J.; Spano, F. C. Expanded theory of H- and J-molecular aggregates: the effects of vibronic coupling and intermolecular charge transfer. Chem. Rev. 2018, 118, 7069–7163. doi: 10.1021/acs.chemrev.7b00581

    14. [14]

      Cariati, E.; Macchi, R.; Roberto, D.; Ugo, R.; Galli, S.; Casati, N.; Macchi, P.; Sironi, A.; Bogani, L.; Caneshi, A.; Gatteschi, D. Polyfunctional inorganic-organic hybrid materials: an unusual kind of NLO active layered mixed metal oxalates with tunable magnetic properties and very large second harmonic generation. J. Am. Chem. Soc. 2007, 129, 9410−9420. doi: 10.1021/ja0710712

    15. [15]

      Lacroix, P. G.; Munoz, M. C.; Gaspar, A. B.; Real, J. A.; Bonhommeau, S.; Rodriguez, V.; Nakatani, K. Synthesis, crystal structures, and solid state quadratic nonlinear optical properties of a series of stilbazolium cations combined with gold cyanide counter-ion. J. Mater. Chem. 2011, 21, 15940−15949. doi: 10.1039/c1jm12105g

    16. [16]

      Cariati, E.; Ugo, R.; Cariati, F.; Roberto, D.; Masciocchi, M.; Galli, S.; Sironi, A. J-aggregates granting giant second-order NLO responses in self-assembled hybrid inorganic-organic materials. Adv. Mater. 2001, 13, 1665−1668. doi: 10.1002/1521-4095(200111)13:22<1665::AID-ADMA1665>3.0.CO;2-X

    17. [17]

      Compain, J. D.; Mialane, P.; Dolbecq, A.; Marrot, J.; Proust, A.; Nakatani, K.; Yu, P.; Secheresse, F. Second-order nonlinear optical properties of polyoxometalate Ssalts of a chiral stilbazolium derivative. Inorg. Chem. 2009, 48, 6222−6228. doi: 10.1021/ic900519p

    18. [18]

      Wang, M.; Yang, C.; Wang, G. E.; Xu, G.; Lv, X. Y.; Xu, Z. N.; Lin, R. G.; Cai, L. Z.; Guo, G. C. A room-temperature X-ray-induced photochromic material for X-ray detection. Angew. Chem. Int. Ed. 2012, 124, 3488–3491. doi: 10.1002/ange.201108220

    19. [19]

      Tian, Y.; Li, F.; Zhang, G. C.; Zhou, H. P.; Wu, J. Y.; Tian, Y. P. Two solid fluorescent organic-inorganic hybrids: synthesis, crystal structures and strong red fluorescence emissions. Chin. J. Inorg. Chem. 2017, 33, 664–672. (20) Lee, S. H.; Kang, B. J.; Yoo, B. W.; Lee, S. C.; Lee, S. J.; Jazbinsek, M.; Yun, H.; Rotermund, B.; Kwon, O. Terahertz phonon mode engineering of highly efficient organic terahertz generators. Adv. Funct. Mater. 2017, 27, 1605583(1–9).

    20. [20]

      Perew, J. P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phy. Rev. Lett. 1996, 77, 3865–3868. doi: 10.1103/PhysRevLett.77.3865

    21. [21]

      Segall, M.; Probert, M.; Pickard, C.; Hasnip, P.; Clark, S.; Refson, K.; Payne, M. First principles methods using CASTEP. Z. Krystallogr. 2005, 220, 567–570.

    22. [22]

      Sheldrick, G. M. SHELXS-97: Program for the Solution of Crystal Structures. University of Gottingen, Germany 1997.

    23. [23]

      Sheldrick, G. M. SHELXL-97: Program for the Refinement of Crystal Structures. University of Gottingen, Germany 1997.

    24. [24]

      Chen, W. T.; Hu, R. H.; Luo, Z. G.; Chen, H. L.; Zhang, X. N.; Liu, J. [N-ethyl-4, 4΄-bipyridinium][ZnX4] (X = Cl or Br) with N-ethyl-4, 4΄-bipyridinium generated in situ: syntheses, structures, fluorescence and TDDFT calculations. Chin. J. Struct. Chem. 2014, 33, 1141–1146.

    25. [25]

      Chen, C. H.; Xu, G. C. Synthesis, characterization and disorder-order phase transition of inorganic-organic hybrid materials (H2dabco-CH2-Cl)MIICl4] (M = Co, Zn). CrystEngComm. 2016, 18, 550–557. doi: 10.1039/C5CE02059J

    26. [26]

      Marder, S. R.; Perry, J. W.; Bourhill, G.; Gorman, C. B.; Tiemann, B. G.; Mansour, K. Relationbetween bond-length alternation and second electronic hyperpolarizability of conjugated organic molecules. Science 1993, 261, 186–189. doi: 10.1126/science.261.5118.186

    27. [27]

      Xie, Z.; Yang, B.; Li, F.; Cheng, G.; Liu, L.; Yang, G.; Xu, H.; Ye, L.; Hanif, M.; Liu, S.; Ma, D.; Ma, Y. Cross dipole stacking in the crystal of distyrylbenzene derivative: the approach toward high solid-state luminescence efficiency. J. Am. Chem. Soc. 2005, 127, 14152–14153. doi: 10.1021/ja054661d

    28. [28]

      Cariati, E.; Macchi, R.; Roberto, D.; Ugo, R.; Galli, S.; Masciocchi, N.; Sironi, A. Sequential self-organization of silver(I) layered materials with strong SHG by J-aggregation and intercalation of organic nonlinear optical chromophores through mechanochemical synthesis. Chem. Mater. 2007, 19, 3704–3711. doi: 10.1021/cm071003e

    29. [29]

      Xu, G.; Li, Y.; Zhou, W. W.; Wang, G. J.; Long, X. F.; Cai, L. Z.; Wang, M. S.; Guo, G. C.; Huang, J. S.; Batorb, G.; Jakubas, R. A ferroelectric inorganic-organic hybrid based on NLO-phore stilbazolium. J. Mater. Chem. 2009, 19, 2179–2183. doi: 10.1039/b819473d

    30. [30]

      Kolev, T.; Koleva, B. B.; Seidel, R. W.; Spiteller, M.; Sheldrick, W. S. New aspects on the origin of color in the solid state. Coherently shifting of the protons in violurate crystals. Cryst. Growth Des. 2009, 9, 3348–3352. doi: 10.1021/cg900188k

    31. [31]

      Cai, S.; Wang, M. S.; Li, P. X.; Guo, G. C. Conductance switch of a bromoplumbate bistable semiconductor by electron-transfer thermochromism. Angew. Chem. Int. Ed. 2017, 56, 554–558. doi: 10.1002/anie.201610180

    32. [32]

      Lin, R. G.; Xu, G.; Wang, M. S.; Lu, G.; Li, P. X.; Guo, G. C. Improved photochromic properties on viologen-based inorganic-organic hybrids by using π-conjugated substituents as electron donors and stabilizers. Inorg. Chem. 2013, 52, 1199−1205. doi: 10.1021/ic301181b

    33. [33]

      Xu, G.; Guo, G. C.; Wang, M. S.; Zhang, Z. J.; Chen, W. T.; Huang, J. S. Photochromism of a methyl viologen bismuth(III) chloride: structural variation before and after UV irradiation. Angew. Chem. Int. Ed. 2007, 46, 3249–3251. doi: 10.1002/anie.200700122

  • Scheme S1  Synthesis route of (DANP)I

    Figure 1  NMR spectrum of (DANP)I

    Figure 2  (a) Structure of ZnCl42- tetrahedron; (b, c) Structures of two independent DANP; (d) X-aggregation of DANP (H atoms were omitted for clarity)

    Figure 3  (a) 1D [(DANP-H)(ZnCl4)]n chain based on C–H···Cl hydrogen bonds among b-axis; (b) Double chain C–H···π interaction; (c) 3D network constructed from C–H···Cl hydrogen bonds

    Figure 4  (a) Optical diffuse-reflection spectra and (b) solid-state emission spectra of 1 (λex = 625 nm) and (DANP)I (λex = 570 nm)

    Figure 5  Photo/thermochromisms of 1

    Figure 6  Band structure (left) and projected density of states (pDOS, right) for 1

    Table 1.  Selected Bond Lengths (Å) and Bond Angles (º)

    Bond Dist. Bond Dist. Bond Dist.
    Zn(1)–Cl(1) 2.230(3) Zn(1)–Cl(2) 2.316(3) Zn(1)–Cl(3) 2.258(3)
    Zn(1)–Cl(4) 2.263(3) Zn(2)–Cl(5) 2.316(3) Zn(2)–Cl(6) 2.288(3)
    Zn(2)–Cl(7) 2.251(3) Zn(2)–Cl(8) 2.259(3)
    Angle (°) Angle (°) Angle (°)
    Cl(1)–Zn(1)–Cl(3) 110.89(12) Cl(1)–Zn(1)–Cl(4) 112.77(15) Cl(3)–Zn(1)–Cl(4) 106.44(13)
    Cl(1)–Zn(1)–Cl(2) 108.20(17) Cl(3)–Zn(1)–Cl(2) 111.71(13) Cl(4)–Zn(1)–Cl(2) 106.80(13)
    Cl(7)–Zn(2)–Cl(8) 110.15(13) Cl(7)–Zn(2)–Cl(6) 109.15(14) Cl(8)–Zn(2)–Cl(6) 106.20(13)
    Cl(7)–Zn(2)–Cl(5) 112.38(14) Cl(8)–Zn(2)–Cl(5) 111.76(13) Cl(6)–Zn(2)–Cl(5) 106.93(12)
    下载: 导出CSV

    Table 2.  Hydrogen Bridging Details of 1 (Lengths in Å and Angles in º)

    D–H···A d(D–H) d(H···A) d(D···A) ∠DHA Symmetry codes
    C(1)–H(1C)···Cl(6) 0.96 2.66 3.567(13) 157
    C(3)–H(3)···Cl(3) 0.93 2.63 3.543(12) 165
    C(4)–H(4)···Cl(4) 0.93 2.81 3.649(12) 150 –1/2 + x, 1/2 – y, –1/2 + z
    C(17)–H(17B)···Cl(4) 0.97 2.69 3.477(12) 139 x, 1 + y, z
    C(19)–H(19C)···Cl(4) 0.96 2.75 3.657(12) 158
    C(21)–H(21)···Cl(8) 0.93 2.61 3.520(12) 164
    C(35)–H(35A)···Cl(6) 0.97 2.68 3.486(10) 141 x, 1 + y, z
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  1
  • 文章访问数:  614
  • HTML全文浏览量:  11
文章相关
  • 发布日期:  2020-10-01
  • 收稿日期:  2019-12-23
  • 接受日期:  2020-02-28
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章