Structures and magnetism of dinuclear Co complexes based on imine derivatives

Yadan SUN Xinfeng LI Qiang LIU Oshio Hiroki Yinshan MENG

Citation:  Yadan SUN, Xinfeng LI, Qiang LIU, Oshio Hiroki, Yinshan MENG. Structures and magnetism of dinuclear Co complexes based on imine derivatives[J]. Chinese Journal of Inorganic Chemistry, 2024, 40(11): 2212-2220. doi: 10.11862/CJIC.20240131 shu

基于亚胺衍生物的双核Co配合物的结构与磁性

    通讯作者: 孟银杉, mengys@dlut.edu.cn
  • 基金项目:

    国家自然科学基金 22222103

    国家自然科学基金 22173015

    国家自然科学基金 21801037

    国家自然科学基金 22071017

    中央高校基本科研业务费专项资金 DUT22LAB606

摘要: 本研究基于3种亚胺衍生物bis-[4-(2-pyridylmethyleneamino)-phenyl]thioether (L1)、bis-[4-(2-pyridylmethyleneamino)-phenyl]ether (L2)和bis-[4-(2-pyridylmethyleneamino)-phenyl]methane (L3)合成了3例双核钴配合物。单晶X射线衍射表明配合物[Co2(L1)3](ClO4)4·2CH3CN (1)、[Co2(L2)3](ClO4)4·2CH3OH (2)和[Co2(L3)3](ClO4)4·2CH3OH (3)均为双核结构。磁性测试结果表明, 化合物3在300 K时表现出由失溶剂诱导的不可逆的SCO行为, Co—N的平均键长从0.213 9(3) nm增加到0.215 3(3) nm。失溶剂后的化合物3表现出与化合物1和化合物2相似的变温磁化率行为。变温紫外可见光谱研究也表明化合物3发生失溶剂诱导的自旋态转变。

English

  • Metal-organic cages are a class of molecular containers formed via coordination-driven self-assembly[1]. Upon external stimulation[2-3], the binding and release of guest molecules inside the cavity due to structure changes, have been widely used for molecular recognition[4], chirality sensing[5], and catalysis[6]. Spin crossover (SCO) phenomena[7-10], not only enable the reversibly interconverted between high-spin (HS) and low-spin (LS) states in complexes but also lead to changes in color[11], volume[12], and electrical conductivity[13-14]. Given this characteristic, incorporating SCO units into coordination cages, since the spin state change will be accompanied by a huge volume change, the spin state can be regulated through external stimulation to achieve the binding and release of guest molecules[15]. Therefore, the development of metal-organic cages based on SCO has become a research hotspot[16-18].

    To date, numerous studies have reported on different structural configurations, predominantly focusing on the SCO behavior of iron[19-22]. For instance, McConnell′s group[23] achieved the transition from high to low spin states in the Fe centers of Fe4L4 cages through subcomponent exchange. Li′s group[24] successfully prepared a mixed-spin [Fe4L6]8+ tetrahedral SCO cage, with VT-XPS (variable temperature X-ray photoelectron spectroscopy), reproducing a regular SCO dependence. Recently, Sun′s group[25] demonstrated that the spin transition temperature (T1/2) of the SCO host can be adjusted through the encapsulation of multiple guests. Despite many reports on various SCO complexes about irons, however, the construction of Co-based SCO cages remains unreported. Since organic cages based on cobalt have demonstrated significant research value in the field of electrocatalysis, therefore, the construction of SCO-type Co cage will help further explore the control of electrocatalytic performance by regulating spin states.

    In this regard, we designed three bidentate pyridylamine ligands and synthesized three new complexes with cobalt perchlorate: [Co2(L1)3](ClO4)4·2CH3CN (1), [Co2(L2)3](ClO4)4·2CH3OH (2), and [Co2(L3)3](ClO4)4·2CH3OH (3), where L1: bis-[4-(2-pyridylmethyleneamino)-phenyl]thioether, L2: bis-[4-(2-pyridylmethyleneamino)-phenyl]ether, L3: bis-[4-(2-pyridylmethyleneamino)-phenyl]methane. Under solvated conditions, complex 3 exhibited SCO behavior that was different from complexes 1 and 2. At the same time, we investigated the impact of different ligand fields on the SCO behavior of the complexes[9, 17-18]. Electron-donating groups, such as —CH2, enhanced the ligand field strength; in contrast, electron-withdrawing groups, like —O and —S, favored the reverse bonding capability between Co ions and ligands, thereby weakening the ligand field strength and predisposing the complexes towards a high-spin state[16, 26-27]. This dynamic illustrates the nuanced interplay between ligand structure and electronic properties in determining the spin states of coordination complexes, highlighting the crucial role of ligand design in the modulation of magnetic properties in metal complexes[28-30].

    The chemical reagents used are all analytical reagents sold on the market. If there is no special explanation in this paper, it is directly used. The ligands L1, L2, and L3 were synthesized according to the procedure described in the literature[18, 31-32]. Detailed synthesis procedures of the ligands are described in Scheme S1 (Supporting information).

    Complex 1 was synthesized by the solution volatilization method. At room temperature, ligand L1 (0.06 mmol) was dissolved with 5 mL of acetonitrile solution, and 10 mL of a methanol solution containing Co(ClO4)2·6H2O (0.04 mmol) was added to the former solution. The massive red crystals were obtained after three weeks. Based on Co(ClO4)2·6H2O, its yield was about 40%. Anal. Calcd. for C76H60Cl4Co2N14O16S3(%): C, 51.20; H, 3.37; N, 11.00. Found(%): C, 51.13; H, 3.32; N, 10.88.

    The synthesis method of complex 2 was the same as complex 1. Red block crystals were obtained after three weeks. Based on Co(ClO4)2·6H2O, its yield was about 36%. Anal. Calcd. for C75H66Cl4Co2N12O22(%): C, 51.52; H, 3.78; N, 9.62. Found(%): C, 51.47; H, 3.73; N, 9.57.

    The synthesis method of complex 3 was the same as complex 1. Red block crystals were obtained after four weeks. Yield: 30% based on Co(ClO4)2·6H2O. Anal. Calcd. for C77H68Cl4Co2N12O18(%): C, 54.06; H, 3.98; N, 9.83. Found(%): C, 53.98; H, 3.92; N, 9.76.

    Crystals suitable for single-crystal X-ray diffraction, covered with a thin layer of hydrocarbon oil, were placed under a cold nitrogen stream using glass fibers connected to a copper needle. Single-crystal X-ray diffraction data were collected on a Bruker D8 VENTURE CMOS-based diffractometer (Bruker AXS Company, Karlsruhe, Germany) (Mo radiation, λ=0.071 073 nm) using the SMART and SAINT programs. Final unit cell parameters were based on all observed reflections from integrating all frame data. Through the ShelXT structure parsing program implanted in Olex2, the crystal structure was analyzed by the direct method, and the data was refined by the full matrix least square method using the ShelXL program. For all complexes, all non-hydrogen atoms were treated with anisotropy, and hydrogen atoms were determined utilizing electron cloud or theoretical hydrogenation.

    Thermogravimetric analysis (TGA) was performed using TG/DTA Q600 (Mettler Toledo) instruments in a nitrogen atmosphere at a heating rate of 10 K·min-1 from ambient temperature to 1 085 K. Elemental analysis was performed on a Vario EL Ⅲ element analyzer (Elementar Company, Hanau, Germany). Powder X-ray diffraction (PXRD) data were obtained in a Bruker AXS D8 Advance X-ray powder diffractometer using the voltage of 45 kV and the test current of 200 mA in a range of 5°-50° (Cu radiation, λ=0.154 18 nm). Magnetic measurements were carried out using the polycrystalline sample with the Quantum Design PPMS. The variable-temperature magnetization data were corrected for diamagnetic contribution from the sample holder and diamagnetic contributions from the molecule using Pascal′s constants. The UV-Vis absorption spectra of the complexes were measured by HITACHI HU4150 spectrophotometer.

    The synthesis and growth of red crystals of complexes 1, 2, and 3 were conducted by the slow diffusion of the methanol solution of Co(ClO4)2·6H2O into the acetonitrile solution of ligands in test tubes. The solid structures of three dinuclear complexes 1, 2, and 3 were determined by single-crystal X-ray diffraction. Detailed crystallographic data are listed in Table 1.

    Table 1

    Table 1.  Crystallographic data for complexes 1-3 at different temperatures
    下载: 导出CSV
    Parameter 1 2 3
    Temperature / K 120 120 120 400
    Formula C76H60Cl4Co2N14O16S3 C75H66Cl4Co2N12O22 C77H68Cl4Co2N12O18 C75H60Cl4Co2N12O16
    Formula weight 1 781.22 1 747.05 1 709.09 1 645.01
    Crystal system Monoclinic Monoclinic Triclinic Monoclinic
    Space group Pn C2/c P1 P21/n
    a / nm 1.358 57(15) 2.951 75(12) 1.050 30(16) 2.515 35(10)
    b / nm 1.330 72(14) 1.026 54(4) 1.536 2(2) 1.050 94(5)
    c / nm 2.191 3(2) 2.480 22(11) 2.473 5(4) 2.854 46(10)
    α / (°) 87.248(5)
    β / (°) 107.335(4) 91.102 0(10) 88.417(5) 92.009(3)
    γ / (°) 70.555(5)
    V / nm3 3.781 7(7) 7.513 9(5) 3.758 6(10) 7.541 1(5)
    Z 2 4 2 4
    Dc / (g·cm-3) 1.564 1.544 1.510 1.449
    F(000) 1 824.0 3 592.0 1 760.0 3 376.0
    Rint 0.096 9 0.036 9 0.083 9 0.027 0
    GOF on F2 1.059 1.034 1.054 1.028
    R1 [I≥2σ(I)]a 0.061 7 0.053 6 0.078 2 0.074 5
    wR2 [I≥2σ(I)]b 0.155 9 0.143 8 0.214 1 0.201 8
    a R1=∑||Fo|-|Fc||/∑|Fo|; b wR2=[∑w(Fo2-Fc2)2/∑w(Fo2)2]1/2.

    All complexes are isostructural whose assembly mode was characterized by an [M2L3] pattern (Fig. 1), incorporating a dinuclear structure and four isolated perchlorate anions to maintain electrical neutrality. The two Co centers adopt a distorted octahedral coordination environment, bound to three imine and three pyridine nitrogen donors from different ligands.

    Figure 1

    Figure 1.  Major structure of complexes 1 (a), 2 (b), and 3 (c)

    The H atoms and anions are omitted for clarity; Color code: Co, purple; O, red; C, gray; S, yellow; N, blue.

    The asymmetric unit of complex 1 contains two CH3CN solvent molecules and two types of metal Co, while both complexes 2 and 3 contain two CH3OH solvent molecules. Complex 2 has only one type of metal Co compared to the two types of metal Co of complexes 1 and 3. TGA shows that complex 1 slowly lost solvent at 414 K and stabilized at 481 K, followed by a sharp decrease in the curve due to the decomposition of the complex after 600 K. The lost mass of 5.4% is approximately equal to that of two CH3CN molecules (Fig.S1). Compared with complex 1, complex 2 was very stable below 600 K (Fig.S2). For complex 3, as the temperature increased, complex 3 began to slowly lose a small amount of solvent, which lost 2% of its mass, approximately equivalent to the mass of two CH3OH molecules (1.87%) and then decomposed at 600 K (Fig.S3).

    Complexes 1 and 2 crystallize in the space group of Pn and C2/c, respectively. At the measured temperature of 120 K, the average bond length of Co1—N and Co2—N is 0.214 7(4) and 0.214 8(5) nm for complex 1, and 0.214 3(3) nm for complex 2, indicating that the Co centers are in the HS state (Table 2). Owing to the loss of the solvent in complex 3, the space group has changed from P1 at 120 K to P21/n at 400 K. At the same time, the main skeleton remains unchanged, and the average bond lengths for Co1—N and Co2—N increase from 0.213 7(3) and 0.214 1(3) nm to 0.215 2(3) and 0.215 4(3) nm, respectively (Table 2), which represents the occurrence of the SCO process. In addition, the hydrogen bonding interactions between perchlorate ions and methanol molecules were found in the lattice with hydrogen bonding distances of 0.197 4 nm for complex 2 (Fig.S4), 0.198 8 and 0.204 6 nm for complex 3 (Fig.S5). The hydrogen bonding has no significant effect on the environment for the formation of the unique dinuclear structure.

    Table 2

    Table 2.  Structural parameters for complexes 1-3
    下载: 导出CSV
    Complex dCo—Na / nm Σb / (°) CShMc dCo…Cod / nm Φe / (°)
    1-120 K Co1 0.214 7(4) 94.20 1.509 1.133 102.563
    Co2 0.214 8(5) 90.45 1.353
    2-120 K Co 0.214 3(3) 80.75 1.444 1.135 116.427
    3-120 K Co1 0.213 7(3) 83.49 1.342 1.150 114.745
    Co2 0.214 1(3) 79.01 1.364
    3-400 K Co1 0.215 4(3) 85.37 1.452 1.139 114.516
    Co2 0.215 2(3) 82.26 1.487
    a The average Co—N bond length; b Defined as the sum of the deviation from 90° of the 12 cis N—Co—N angles of the [CoN6] octahedron; c The CShM value represents the deviation degree of the ideal octahedron; d The distance between the central cobalt atom; e Refers to the angle on the ligand that is centered on the center of symmetry of the ligand′s center-connecting unit.

    To further investigate the electronic structure, the solid-state variable-temperature UV-Vis absorption spectra of complex 3 were examined within a range of 300-1 500 nm. As shown in Fig. 2, two broad absorption bands near 350 and 950 nm observed at 250 K can be attributed to the metal-to-ligand charge transfer (MLCT) or d-d transitions of LS-Co ions (Fig. 2). The peak was around 950 nm which may be attributed to d-d transitions for high spin Co in the pseudo-octahedral environment. The intensity of these bands gradually decreased with the temperature increased, which was consistent with the thermally induced spin crossover from LS-Co to HS-Co states.

    Figure 2

    Figure 2.  Temperature-dependent UV-Vis absorption spectra of complex 3 in the solid state

    The phase purities of all three complexes were confirmed by PXRD patterns (Fig.S6-S8). The variable-temperature magnetic susceptibility of the above complexes was measured at a scanning speed of 3 K·min-1 and under a DC magnetic field of 1 kOe. Upon warming to 395 K, the χMT values of complexes 1 and 2 continuously increased to 6.41 and 6.16 cm3·mol-1·K (Fig. 3a and 3b), significantly larger than the saturated value of 5.0 cm3·mol-1·K for two Co in the HS state (S=3/2). Curie-Weiss fits the magnetization curves for the temperature interval 2-400 K yielded a Curie constant C of 6.63 cm3·mol-1·K, and a Weiss temperature θ of -16.82 K were obtained for complex 1 (Fig.S9). A Curie constant C of 6.37 cm3·mol-1·K and a Weiss temperature θ of -12.72 K were obtained for complex 2 (Fig.S10). Negative Weiss temperatures provide further evidence that complexes 1 and 2 exhibit antiferromagnetic interactions.

    Figure 3

    Figure 3.  Plots of the temperature dependence of χMT under 1 kOe DC field for complexes 1 (a), 2 (b), and 3 (c)

    Black: heating mode; Red: cooling mode.

    For complex 3, as the temperature increased to 300 K, the χMT value gradually increased and stabilized at 6.13 cm3·mol-1·K. Upon further heating to 395 K, the χMT value rose to 6.76 cm3·mol-1·K, indicating the occurrence of incomplete SCO behavior. After the sample was continuously maintained at 395 K for 1 h to ensure complete desolvation, the magnetic data of the 3-desolvated phase were recorded over the temperature range of 2-395 K. With the reduction of temperature, it was observed that the χMT value decreased to 3.9 cm3·mol-1·K at 2 K similar to complexes 1 and 2. Curie-Weiss fits the magnetization curves for complex 3 gave a Curie constant C of 7.08 cm3·mol-1·K and a Weiss temperature θ of -17.31 K (Fig.S11), demonstrating that the desolvated phase of complex 3 undergoes antiferromagnetic interactions similar to complexes 1 and 2. The above results suggest that complex 3 undergoes an irreversible SCO behavior induced by the loss of solvent.

    Electron-withdrawing groups facilitate the reverse bonding ability between the Co ion and the ligand, thereby weakening the ligand field strength and tending the complex to a high spin state, so complexes 1 and 2 are in the high spin state. However, electron-donating groups, such as —CH2, enhanced the ligand field strength. Consequently, complex 3 possesses SCO behaviors that are different from complex 1 and complex 2. Therefore, it is proved different ligand fields have a great influence on the SCO behavior of complexes. To further analyze the structural distortions among the SCO behaviors, the structural parameters are listed including dCo—N (nm), Σ (°), CShM, Co…Co distance (nm), and Φ (°), which are shown in Table 2. The values of these parameters indicate that there are different degrees of geometrical distortion in the coordination spheres of the three complexes. The Co…Co distance between complexes 1 and 2 is 1.133 and 1.135 nm, respectively, allowing the dinuclear structure to compress horizontally, resulting in a large distortion of the N6 environment involved in the coordination of metal Co, and with distortion parameters of 94.20°, 90.45°, and 80.75°, respectively, which makes the central metal cobalt more inclined to the HS. At the same time, the Co…Co distance of complex 3 decreases from 1.150 nm at 120 K to 1.139 nm at 400 K close to complexes 1 and 2 due to solvent loss, which exacerbates the compression of the crystal cell in the horizontal direction and leads to an increase in the distortion of the center metal from 83.49°, 79.01° to 85.37°, 82.26°. Therefore, the Co…Co distance not only reflects the degree of distortion of the center metal but also indirectly indicates the SCO behavior of complex 3. Besides this, we also explored the symmetry angles of different ligands which are 102.56° and 116.427° for complexes 1 and 2 at 120 K respectively, the symmetry angle of complex 3 decreases from 114.745° to 114.516° with increasing temperature. Since the ligand configuration has no obvious change during the transformation process, so we speculate that the ligand intercalation angle has a weak effect on the SCO of the complexes.

    In addition to the first coordination sphere of the metal center, subtle crystal packing effects also affect the SCO behavior in the solid state. Close examination of intermolecular interactions reveals additional insights into structure-function relationships. As shown in Fig. 4, complex 1 exhibits the ππ intermolecular interaction between the benzene rings of the Schiff base ligands in different units, characterized by a distance of 0.376 9 nm. Besides, there are two types of C—H…π intermolecular interactions between the Schiff base ligands in complex 1, with a distance of 0.348 7 and 0.353 0 nm (Fig. 5a). These multiple intermolecular interactions facilitate a denser spatial arrangement of complex 1, leading to an increased rigidity in the coordination environment of the central metal. This, in turn, enhances the stability of the complex in the HS state. Unlike complex 1 which possesses intermolecular interactions, complexes 2 and 3 only exhibit two kinds of C—H…π intramolecular interactions (Fig. 5b-5d). The two different C—H…π intramolecular interactions originate from the benzene or pyridine rings in the ligand and the C—H of the same ring on the adjacent ligand, whereas the distance of 0.360 5 and 0.368 1 nm for complex 2, with 0.377 7 and 0.378 2 nm for complex 3, respectively. Although complexes 2 and 3 exhibit the same type of intramolecular interactions, the difference in the strength of these interactions leads to a vertically more compressed dinuclear structure in complex 2 compared to complex 3, thus increasing the rigidity within the molecule stabilizes complex 2 in the HS state. For complex 3, the C—H…π intramolecular interaction distances decrease from 0.377 7 to 0.374 0 nm with the temperature increase. And the enhancement of intramolecular forces keeps the desolvated phase of complex 3 in the HS state which is consistent with complex 2. At the same time, this trend aligns with the results from magnetization measurements. Therefore, it can be concluded that intermolecular or intramolecular forces exert a significant influence on the magnetic properties of the complex.

    Figure 4

    Figure 4.  Illustration of intramolecular ππ stacking within the complex 1

    The red dashed lines represent ππ interaction.

    Figure 5

    Figure 5.  C—H…π interaction for the complexes: (a) 1 at 120 K, (b) 2 at 120 K, (c) 3 at 120 K, and (d) 3 at 400 K

    The orange dashed lines represent the C—H…π interaction.

    In summary, three dinuclear Co complexes [Co2(L1)3](ClO4)4·2CH3CN (1), [Co2(L2)3](ClO4)4·2CH3OH (2), and [Co2(L3)3](ClO4)4·2CH3OH (3) were synthesized utilizing three imine-based ligands and cobalt perchlorate. Among these, complex 3, bearing the electron-donating —CH2 substituent, demonstrated solvent-induced SCO behavior, unlike complexes 1 and 2 with electron-withdrawing substituents —S and —O, which exhibit antiferromagnetic behaviors. Through magneto-structure relationship analysis, substituent-mediated ligand field and intramolecular or intermolecular stacking interactions can significantly influence the SCO behavior in dinuclear complexes. In turn, this allows for the adjustment of cavity volume within metal-organic cages to accommodate guest molecules. These insights will contribute to the development of more Co-based cages exhibiting SCO behavior.


    Supporting information is available at http://www.wjhxxb.cn
    1. [1]

      Espinosa C F, Ronson T K, Nitschke J R. Secondary bracing ligands drive heteroleptic cuboctahedral Pd12 cage formation[J]. J. Am. Chem. Soc., 2023, 145(18):  9965-9969. doi: 10.1021/jacs.3c00661

    2. [2]

      刘丹, 赵亮, 邵震, 孟银杉, 刘涛. 一例具有可逆热诱导电荷转移行为的二维氰基桥联W-Co配合物[J]. 无机化学学报, 2023,39,(2): 367-374. LIU D, ZHANG L, SHAO Z, MENG Y S, LIU T. A 2D cyano-bridged W-Co coordination network exhibiting reversible thermal-induced charge transfer[J]. Chinese J. Inorg. Chem., 2023, 39(2):  367-374.

    3. [3]

      杨蕊, 张舒雅, 王润国, 孟银杉, 刘涛, 朱元元. 三联吡啶异配体对构筑单核钴(Ⅱ)自旋交叉配合物的合成和磁性[J]. 无机化学学报, 2023,38,(8): 1477-1486. YANG R, ZHANG S Y, WANG R G, MENG Y S, LIU T, ZHU Y Y. Synthesis and magnetic properties of mononuclear cobalt(Ⅱ) spin crossover complexes from complementary terpyridine ligand pairing[J]. Chinese J. Inorg. Chem., 2023, 38(8):  1477-1486.

    4. [4]

      Moree L K, Faulkner L A V, Crowley J D. Heterometallic cages: Synthesis and applications[J]. Chem. Soc. Rev., 2024, 53(1):  25-46. doi: 10.1039/D3CS00690E

    5. [5]

      Gural'skiy I A, Kucheriv O I, Shylin S I, Ksenofontov V, Polunin R A, Fritsky I O. Enantioselective guest effect on the spin state of a chiral coordination framework[J]. Chem.-Eur. J., 2015, 21(50):  18076-18079. doi: 10.1002/chem.201503365

    6. [6]

      He P, Hu M Y, Li J H, Qiao T Z, Lu Y L, Zhu S F. Spin effect on redox acceleration and regioselectivity in Fe-catalyzed alkyne hydrosilylation[J]. Natl. Sci. Rev., 2024, 11(2):  nwad324. doi: 10.1093/nsr/nwad324

    7. [7]

      Bousseksou A, Molnár G, Real J A, Tanaka K. Spin crossover and photomagnetism in dinuclear iron compounds[J]. Coord. Chem. Rev., 2007, 251(13/14):  1822-1833.

    8. [8]

      Hogue R W, Singh S, Brooker S. Spin crossover in discrete polynuclear iron(Ⅱ) complexes[J]. Chem. Soc. Rev., 2018, 47(19):  7303-7338. doi: 10.1039/C7CS00835J

    9. [9]

      Scott H S, Staniland R W, Kruger P E. Spin crossover in homoleptic Fe(Ⅱ) imidazolylimine complexes[J]. Coord. Chem. Rev., 2018, 362:  24-43. doi: 10.1016/j.ccr.2018.02.001

    10. [10]

      Gütlich P, Gaspar A B, Garcia Y. Spin state switching in iron coordination compounds[J]. Beilstein J. Org. Chem., 2013, 9:  342-391. doi: 10.3762/bjoc.9.39

    11. [11]

      Kusumoto S, Inaba K, Suda H, Nakaya M, Tokunaga R, Thuéry P, Haruki R, Kanazawa T, Nozawa S, Kim Y, Hayami S, Koide Y. Cooperative spin-state switching and vapochromism of mononuclear Ni(Ⅱ) complexes by pyridine coordination/decoordination[J]. Inorg. Chem., 2023, 62(39):  16222-16227. doi: 10.1021/acs.inorgchem.3c02776

    12. [12]

      Guo W, Daro N, Pillet S, Marchivie M, Bendeif E E, Tailleur E, Chainok K, Denux D, Chastanet G, Guionneau P. Unprecedented reverse volume expansion in spin-transition crystals[J]. Chem.-Eur. J., 2020, 26(57):  12927-12930. doi: 10.1002/chem.202001821

    13. [13]

      Ishikawa R, Ueno S, Nifuku S, Horii Y, Iguchi H, Miyazaki Y, Nakano M, Hayami S, Kumagai S, Katoh K, Li Z Y, Yamashita M, Kawata S. Simultaneous spin-crossover transition and conductivity switching in a dinuclear iron(Ⅱ) coordination compound based on 7, 7', 8, 8'-tetracyano-p-quinodimethane[J]. Chem.-Eur. J., 2019, 26(6):  1278-1285.

    14. [14]

      Üngör Ö, Choi E S, Shatruk M. Optimization of crystal packing in semiconducting spin-crossover materials with fractionally charged TCNQδ- anions (0 < δ < 1)[J]. Chem. Sci., 2021, 12(32):  10765-10779. doi: 10.1039/D1SC02843J

    15. [15]

      Yang Y, Ronson T K, Zheng J, Mihara N, Nitschke J R. Fluoride up- and down-regulates guest encapsulation for Zn6L4 and Zn4L4 cages[J]. Chem, 2023, 9(7):  1972-1982. doi: 10.1016/j.chempr.2023.03.027

    16. [16]

      Singh S, Hogue R W, Feltham H L C, Brooker S. Dinuclear helicate and tetranuclear cage assembly using appropriately designed ditopic triazole-azine ligands[J]. Dalton Trans., 2019, 48(41):  15435-15444. doi: 10.1039/C9DT01890E

    17. [17]

      Li L, Craze A R, Akiyoshi R, Tsukiashi A, Hayami S, Mustonen O, Bhadbhade M M, Bhattacharyya S, Marjo C E, Wang Y, Lindoy L F, Aldrich-Wright J R, Li F. Direct monitoring of spin transitions in a dinuclear triple-stranded helicate iron(Ⅱ) complex through X-ray photoelectron spectroscopy[J]. Dalton Trans., 2018, 47(8):  2543-2548. doi: 10.1039/C7DT04190J

    18. [18]

      Parajó Y, Malina J, Meistermann I, Clarkson G J, Pascu M, Rodger A, Hannon M J, Lincoln P. Effect of bridging ligand structure on the thermal stability and DNA binding properties of iron(Ⅱ) triple helicates[J]. Dalton Trans., 2009, (25):  4868-4874. doi: 10.1039/b822039e

    19. [19]

      Archer R J, Scott H S, Polson M I J, Williamson B E, Mathonière C, Rouzières M, Clérac R, Kruger P E. Varied spin crossover behaviour in a family of dinuclear Fe(Ⅱ) triple helicate complexes[J]. Dalton Trans., 2018, 47(24):  7965-7974. doi: 10.1039/C8DT01567H

    20. [20]

      Athira S, Mondal D J, Shome S, Dey B, Konar S. Effect of intermolecular anionic interactions on spin crossover of two triple-stranded dinuclear Fe(Ⅱ) complexes showing above room temperature spin transition[J]. Dalton Trans., 2022, 51(43):  16706-16713. doi: 10.1039/D2DT02115C

    21. [21]

      He C, Duan C Y, Fang C J, Meng Q J. Self-assembled dinuclear molecular box [Ag2L2]2+ and triple helicates [Co2L3]4+, [Ni2L3]4+ {L=bis[4-(2-pyridylmethyleneamino)phenyl] ether}[J]. J. Chem. Soc. Dalton Trans., 2000, (14):  2419-2424. doi: 10.1039/b001820l

    22. [22]

      Craze A, Bhadbhade M, Kepert C, Lindoy L, Marjo C, Li F. Solvent effects on the spin-transition in a series of Fe(Ⅱ) dinuclear triple helicate compounds[J]. Crystals, 2018, 8(10):  376. doi: 10.3390/cryst8100376

    23. [23]

      McConnell A J, Aitchison C M, Grommet A B, Nitschke J R. Subcomponent exchange transforms an Fe4L4 cage from high- to low-Spin, switching guest release in a two-cage system[J]. J. Am. Chem. Soc., 2017, 139(18):  6294-6297. doi: 10.1021/jacs.7b01478

    24. [24]

      Li L, Craze A R, Mustonen O, Zenno H, Whittaker J J, Hayami S, Lindoy L F, Marjo C E, Clegg J K, Aldrich-Wright J R, Li F. A mixed-spin spin-crossover thiozolylimine [Fe4L6]8+ cage[J]. Dalton Trans., 2019, 48(27):  9935-9938. doi: 10.1039/C9DT01947B

    25. [25]

      Yin F, Yang J, Zhou L P, Meng X, Tian C B, Sun Q F. 54 K spin transition temperature shift in a Fe6L4 octahedral cage induced by optimal fitted multiple guests[J]. J. Am. Chem. Soc., 2024, 146(11):  7811-7821. doi: 10.1021/jacs.4c00705

    26. [26]

      Singh S, Brooker S. Correlations between ligand field Δo, spin crossover T1/2 and redox potential Epa in a family of five dinuclear helicates[J]. Chem. Sci., 2021, 12(32):  10919-10929. doi: 10.1039/D1SC01458G

    27. [27]

      Thomas J A. Metal ion directed self-assembly of sensors for ions, molecules and biomolecules[J]. Dalton Trans., 2011, 40:  12368-12373. doi: 10.1039/c1dt11381j

    28. [28]

      Craze A R, Bhadbhade M M, Komatsumaru Y, Marjo C E, Hayami S, Li F. A rare example of a complete, incomplete, and non-occurring spin transition in a [Fe2L3]X4 series driven by a combination of solvent-and halide-anion-mediated steric factors[J]. Inorg. Chem., 2020, 59(9):  1274-1283.

    29. [29]

      Craze A R, Zenno H, Pfrunder M C, McMurtrie J C, Hayami S, Clegg J K, Li F. Supramolecular modulation of spin crossover in an Fe(Ⅱ) dinuclear triple helicate[J]. Inorg. Chem., 2021, 60(9):  6731-6738. doi: 10.1021/acs.inorgchem.1c00553

    30. [30]

      Garcia Y, Grunert C M, Reiman S, Van Campenhoudt O, Gütlich P. The two-step spin conversion in a supramolecular triple helicate dinuclear iron(Ⅱ) complex studied by mössbauer spectroscopy[J]. Eur. J. Inorg. Chem., 2006, (17):  3333-3339.

    31. [31]

      Tanh Jeazet H B, Gloe K, Doert T, Kataeva O N, Jäger A, Geipel G, Bernhard G, Büchner B, Gloe K. Self-assembly of neutral hexanuclear circular copper(Ⅱ) meso-helicates: Topological control by sulfate ions[J]. Chem. Commun., 2010, 46(14):  2373-2375. doi: 10.1039/b925469b

    32. [32]

      Hotze A C G, Hodges N J, Hayden R E, Sanchez-Cano C, Paines C, Male N, Tse M K, Bunce C M, Chipman J K, Hannon M J. Supramolecular iron cylinder with unprecedented DNA binding is a potent cytostatic and apoptotic agent without exhibiting genotoxicity[J]. Chem. Biol., 2008, 15(12):  1258-1267. doi: 10.1016/j.chembiol.2008.10.016

  • Figure 1  Major structure of complexes 1 (a), 2 (b), and 3 (c)

    The H atoms and anions are omitted for clarity; Color code: Co, purple; O, red; C, gray; S, yellow; N, blue.

    Figure 2  Temperature-dependent UV-Vis absorption spectra of complex 3 in the solid state

    Figure 3  Plots of the temperature dependence of χMT under 1 kOe DC field for complexes 1 (a), 2 (b), and 3 (c)

    Black: heating mode; Red: cooling mode.

    Figure 4  Illustration of intramolecular ππ stacking within the complex 1

    The red dashed lines represent ππ interaction.

    Figure 5  C—H…π interaction for the complexes: (a) 1 at 120 K, (b) 2 at 120 K, (c) 3 at 120 K, and (d) 3 at 400 K

    The orange dashed lines represent the C—H…π interaction.

    Table 1.  Crystallographic data for complexes 1-3 at different temperatures

    Parameter 1 2 3
    Temperature / K 120 120 120 400
    Formula C76H60Cl4Co2N14O16S3 C75H66Cl4Co2N12O22 C77H68Cl4Co2N12O18 C75H60Cl4Co2N12O16
    Formula weight 1 781.22 1 747.05 1 709.09 1 645.01
    Crystal system Monoclinic Monoclinic Triclinic Monoclinic
    Space group Pn C2/c P1 P21/n
    a / nm 1.358 57(15) 2.951 75(12) 1.050 30(16) 2.515 35(10)
    b / nm 1.330 72(14) 1.026 54(4) 1.536 2(2) 1.050 94(5)
    c / nm 2.191 3(2) 2.480 22(11) 2.473 5(4) 2.854 46(10)
    α / (°) 87.248(5)
    β / (°) 107.335(4) 91.102 0(10) 88.417(5) 92.009(3)
    γ / (°) 70.555(5)
    V / nm3 3.781 7(7) 7.513 9(5) 3.758 6(10) 7.541 1(5)
    Z 2 4 2 4
    Dc / (g·cm-3) 1.564 1.544 1.510 1.449
    F(000) 1 824.0 3 592.0 1 760.0 3 376.0
    Rint 0.096 9 0.036 9 0.083 9 0.027 0
    GOF on F2 1.059 1.034 1.054 1.028
    R1 [I≥2σ(I)]a 0.061 7 0.053 6 0.078 2 0.074 5
    wR2 [I≥2σ(I)]b 0.155 9 0.143 8 0.214 1 0.201 8
    a R1=∑||Fo|-|Fc||/∑|Fo|; b wR2=[∑w(Fo2-Fc2)2/∑w(Fo2)2]1/2.
    下载: 导出CSV

    Table 2.  Structural parameters for complexes 1-3

    Complex dCo—Na / nm Σb / (°) CShMc dCo…Cod / nm Φe / (°)
    1-120 K Co1 0.214 7(4) 94.20 1.509 1.133 102.563
    Co2 0.214 8(5) 90.45 1.353
    2-120 K Co 0.214 3(3) 80.75 1.444 1.135 116.427
    3-120 K Co1 0.213 7(3) 83.49 1.342 1.150 114.745
    Co2 0.214 1(3) 79.01 1.364
    3-400 K Co1 0.215 4(3) 85.37 1.452 1.139 114.516
    Co2 0.215 2(3) 82.26 1.487
    a The average Co—N bond length; b Defined as the sum of the deviation from 90° of the 12 cis N—Co—N angles of the [CoN6] octahedron; c The CShM value represents the deviation degree of the ideal octahedron; d The distance between the central cobalt atom; e Refers to the angle on the ligand that is centered on the center of symmetry of the ligand′s center-connecting unit.
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  1
  • 文章访问数:  38
  • HTML全文浏览量:  9
文章相关
  • 发布日期:  2024-11-10
  • 收稿日期:  2024-04-13
  • 修回日期:  2024-09-10
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