Coordination equilibrium between cyclometalated Pt(Ⅱ) complexs [Pt(κ3-N^C^N′)(CNXyl)]Cl and [Pt(κ2-N^C^N′)(CNXyl)Cl]

Huahong ZHANG Yang ZHAO Rui NING Shuixing WU Xiaopeng ZHANG

Citation:  Huahong ZHANG, Yang ZHAO, Rui NING, Shuixing WU, Xiaopeng ZHANG. Coordination equilibrium between cyclometalated Pt(Ⅱ) complexs [Pt(κ3-N^C^N′)(CNXyl)]Cl and [Pt(κ2-N^C^N′)(CNXyl)Cl][J]. Chinese Journal of Inorganic Chemistry, 2025, 41(9): 1840-1850. doi: 10.11862/CJIC.20250136 shu

环金属铂配合物[Pt(κ3-N^C^N′)(CNXyl)]Cl与[Pt(κ2-N^C^N′)(CNXyl)Cl]的配位平衡

    通讯作者: 吴水星, shxwu@aliyun.com
    张小朋, zxp_inorganic@126.com
  • 基金项目:

    国家自然科学基金 22461019

    海南省重点研发项目 ZDYF2023SHFZ106

摘要: 以含修饰的不对称蒎烯(N^C^N′)的配合物[Pt(N^C^N′)Cl]与芳基异氰化物2,6-二甲基苯基异腈(CNXyl)反应, 产物是互为异构体的离子型配合物[Pt(κ3-N^C^N′)(CNXyl)]Cl ([A]Cl)和分子型配合物[Pt(κ2-N^C^N′)(CNXyl)Cl] (B)的混合物。异构体B占混合产物中的绝大部分。蒎烯的吡啶环上取代基为—CF3和—Cl的异构体B(分别为5B7B)的结构通过单晶X射线衍射进行了确认, 其呈现出扭曲的平面四配位trans-NN^C^N′, CNR构型(κ2-N^C^N′配体的末端N原子与异腈配体CNXyl处于反式位置); 并且通过理论计算验证了异构体B的热力学稳定性更高。

English

  • Cyclometalated cationic Pt(Ⅱ)-isocyanide complexes have received great interest due to their planar coordination geometry, which can self-organize into various nanostructures and chromonic mesophases through intermolecular Pt…Pt and/or ligand-ligand interactions, demonstrating versatile potential in waveguiding, semiconducting, electroluminescent devices, and DNA-binding agents[1-10]. Due to the large ionic polarizability of chloride and sulfate anions in the Hofmeister anion series, most of the cationic Pt(Ⅱ)-isocyanide complexes with chloride and sulfate anions could dissolve well in water up to 20%[11]. Therefore, an aqueous/organic phase-transfer reaction was employed to prepare the cationic Pt(Ⅱ)-isocyanide complexes, and the targeted cationic products were enriched in the aqueous phase with yields exceeding 80%[11-12]. In addition, other anions, such as perchlorate and hexafluorophosphate, were used to replace chloride, and pure products as precipitates in methanol could be obtained[13-14].

    Previously, Che et al. attempted to prepare κ3-[(N^C^N)Pt(CNXyl)]Cl complexes by a simple ligand-substitution reaction in homogeneous solution (N^C^N=1,3-bis(2′-pyridyl)benzene, CNXyl=2,6-dimethylphenyl isocyanide)[15]. However, they had not obtained the targeted cationic product. And they figured out that the reactant [(N^C^N)PtCl] and the desired product κ3-[(N^C^N)Pt(CNXyl)]Cl were coexisting due to reversible ligand exchange reactions. Similarly, we found that pure cyclometalated κ3-[(N^C^N′)Pt(CNXyl)]Cl complexes with the functionalized-pinene ligand were also difficult to separate, and the reaction showed a fairly low yield[16]. Pinenes are accessible natural products with rigid lipophilic groups, and can be easily modified on pyridine planes through Kröhnke-type reaction[17-18]. In addition, the introduction of aliphatic pinene groups would improve the solubility of square-planar Pt(Ⅱ) complexes in nonpolar solvents[12, 14, 16]. Recently, we prepared two different coordination configurations of difluoro-substituted, pinene-appended Pt(Ⅱ)-isocyanide complexes, namely, the κ3-N^C^N′-coordinated [1A]OTf ([1A]OTf is referred to as [1-isomer A]OTf in the reference[19].) and the κ2-N^C^N′-coordinated 1B (1B is referred to as 1-isomer B in the reference[19].) by controlling the chloride source (Fig.1). Accordingly, our tentative hypothesis is whether these two coordination configurations are concurrent in the direct reaction of the [Pt(N^C^N′)Cl] precursor with CNXyl.

    Figure 1

    Figure 1.  Chemical structures of reported difluoro-substituted N^C^N′-coordinated Pt(Ⅱ)-isocyanide complexes[19]

    All reagents were purchased from commercial suppliers and used as received. High-resolution ESI (HR-ESI) mass spectrometry spectra were acquired on a Thermo Fisher Scientific Q Exactive Focus Mass Spectrometer. The 1H, 13C, and 195Pt NMR spectra were obtained on a Bruker DRX-400 spectrometer. Photoluminescence (PL) spectra were measured by a Hitachi F-4600 PL spectrophotometer (λex=420 nm).

    The non‑substituted/substituted unsymmetric pinene ligands (N^C^N′) and [Pt(N^C^N′)Cl] precursors with different substituents were prepared according to the reported literature[14, 19]. The mixture of [Pt(N^C^N′)Cl] precursor and CNXyl was stirred in CH2Cl2 for 30 min at RT. Then the solvent was evaporated in a vacuum, and the obtained solids were used in NMR and photoluminescence tests.

    [1A]OTf. 1H NMR (400 MHz, CD2Cl2, RT): δ 8.66 (dd, J1=5.2 Hz, J2=1.2 Hz, 1H), 8.15 (s, 1H), 7.88 (d, J1=9.6 Hz, J2=1.2 Hz, 1H), 7.64 (dd, J1=8.8 Hz, J2=3.6 Hz, 1H), 7.61 (s, 1H), 7.58-7.50 (m, 1H), 7.46 (t, J=7.6 Hz, 1H), 7.32 (d, J=7.6 Hz, 2H), 7.02 (dd, J1=12.4 Hz, J2=8.4 Hz, 1H), 3.22 (s, 2H), 2.81 (s, 1H), 2.79 (s, 1H), 2.60 (s, 6H), 2.43-2.37 (m, 1H), 1.43 (s, 3H), 1.27 (s, 1H), 0.72 (s, 3H). 13C NMR (100 MHz, CD2Cl2, RT): δ 169.8, 166.1, 165.3, 158.4 (d, J1=268 Hz, C—F), 156.5 (d, J1=250 Hz, C—F), 155.1, 152.8, 152.3, 149.5, 145.6, 140.6, 140.5, 136.4, 131.5, 130.3 (d, J2=22 Hz, C—C—F), 129.1, 128.1, 128.0, 121.1, 119.8, 115.8 (d, J2=27 Hz, C—C—F), 45.0, 39.9, 39.6, 33.9, 31.7, 25.7, 21.6, 19.4 (Fig.S1-S7, Supporting information).

    1B. 1H NMR (400 MHz, CD2Cl2, RT): δ 9.08 (s, 1H), 8.26-8.29 (m, 1H), 7.70 (dd, J1=8.4 Hz, J2 =6.0 Hz, 1H), 7.58 (s, 1H), 7.20 (t, J=7.6 Hz, 1H), 7.09-7.14 (m, 1H), 7.07 (d, J=8.0 Hz, 2H), 7.00 (t, J=8.8 Hz, 1H), 6.63-6.68 (m, 1H), 3.14 (s, 2H), 2.96 (t, J=5.2 Hz, 1H), 2.72-2.78 (m, 1H), 2.37-2.31 (m, 1H), 2.27 (s, 6H), 1.43 (s, 3H), 1.25 (dd, J1=9.6 Hz, J2 =3.2 Hz, 1H), 0.69 (s, 3H). 13C NMR (100 MHz, CD2Cl2, RT): δ 164.0, 161.3 (d, J1=251 Hz, C—F), 158.2 (d, J1=254 Hz, C—F), 151.09, 151.05, 148.0, 146.1 (d, J3=10 Hz, C—C—C—F), 145.4 (d, J4=4 Hz, C—C—C—C—F), 144.2, 142.7, 142.6, 142.4, 134.8, 129.5, 128.2, 127.9, 126.0 (d, J3=10 Hz, C—C—C—F), 124.7, 122.9 (d, J2=20 Hz, C—C—F), 118.8, 111.8 (d, J2=24 Hz, C—C—F), 45.0, 40.2, 39.6, 33.9, 31.9, 25.9, 21.6, 18.7 (Fig.S8-S14).

    CN-substituted N^C^N′ ligand (substituted at 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CDCl3): δ 8.91 (dd, J1=5.2 Hz, J2=1.6 Hz, 1H), 8.50 (t, J=1.6 Hz, 1H), 8.28 (s, 1H), 8.22 (d, J=8.0 Hz, 1H), 8.11 (dd, J1=8.0 Hz, J2=1.6 Hz, 1H), 7.99 (d, J=7.6 Hz, 1H), 7.66 (t, J=7.6 Hz, 2H), 7.41 (dd, J1=7.6 Hz, J2=4.8 Hz, 1H), 3.08 (d, J=2.4 Hz, 2H), 2.90 (t, J=5.6 Hz, 1H), 2.71-2.77 (m, 1H), 2.35 (m, 1H), 1.44 (s, 3H), 1.26 (d, J=9.6 Hz, 1H), 0.68 (s, 3H) (Fig.S15). HR ESI‑MS (m/z): [M+H]+ Calcd. for C24H22N3+, 352.180 82; Found, 352.361 77.

    CH3O-substituted N^C^N′ ligand (substituted at 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CDCl3): δ 8.45 (t, J=1.6 Hz, 1H), 8.33 (dd, J1=4.8 Hz, J2=1.6 Hz, 1H), 8.22 (s, 1H), 8.04 (dt, J1=8.0 Hz, J2=1.6 Hz, 1H), 7.93 (dt, J1=8.0 Hz, J2=1.6 Hz, 1H), 7.60 (s, 1H), 7.53 (t, J=7.6 Hz, 1H), 7.30 (dd, J1=8.4 Hz, J2=1.2 Hz, 1H), 7.22-7.25 (m, 1H), 3.86 (s, 1H), 3.02 (d, J=2.8 Hz, 2H), 2.85 (t, J=5.6 Hz, 1H), 2.68-2.73 (m, 1H), 2.31 (m, 1H), 1.42 (s, 3H), 1.25 (d, J=9.6 Hz, 1H), 0.66 (s, 3H) (Fig.S16). HR ESI-MS (m/z): [M+H]+ Calcd. for C24H25N2O+, 357.196 14; Found, 357.302 05.

    CF3-substituted N^C^N′ ligand (substituted at 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CDCl3): δ 8.86 (d, J=4.0 Hz, 1H), 8.22 (s, 1H), 8.08-8.12 (m, 3H), 7.53-7.57 (m, 3H), 7.44 (dd, J1=8.0 Hz, J2=4.8 Hz, 1H), 3.01 (d, J=2.4 Hz, 2H), 2.85 (t, J=5.6 Hz, 1H), 2.68-2.73 (m, 1H), 2.31 (m, 1H), 1.42 (s, 3H), 1.24 (d, J=9.6 Hz, 1H), 0.66 (s, 3H) (Fig.S17). HR ESI-MS (m/z): [M+H]+ Calcd. for C24H22F3N2+, 395.172 96; Found, 395.286 75.

    Cl-substituted N^C^N′ ligand (substituted at 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CDCl3): δ 8.61 (dd, J1=4.8 Hz, J2=1.6 Hz, 1H), 8.29 (t, J=1.6 Hz, 1H), 8.23 (s, 1H), 8.07 (d, J=8.0 Hz, 1H), 7.82 (dd, J1=8.0 Hz, J2=1.2 Hz, 1H), 7.75 (d, J=7.6 Hz, 1H), 7.58 (s, 1H), 7.55 (d, J=7.6 Hz, 1H), 7.24 (t, J=5.2 Hz, 1H), 3.02 (d, J=2.4 Hz, 2H), 2.85 (t, J=5.6 Hz, 1H), 2.68-2.73 (m, 1H), 2.32 (m, 1H), 1.42 (s, 3H), 1.25 (d, J=9.6 Hz, 1H), 0.66 (s, 3H) (Fig.S18). HR ESI-MS (m/z): [M+H]+ Calcd. for C23H22ClN2+, 361.146 60; Found, 361.221 58.

    CH3-substituted N^C^N′ ligand (substituted at 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CDCl3): δ 8.45 (d, J=4.8 Hz, 1H), 8.21 (s, 1H), 8.17 (s, 1H), 8.09 (m, 1H), 7.99-8.02 (m, 1H), 7.60 (d, J=8.0 Hz, 1H), 7.49 (t, J=2.0 Hz, 1H), 7.35 (d, J=8.0 Hz, 1H), 7.20 (t, J=4.8 Hz, 1H), 2.99 (d, J=2.8 Hz, 2H), 2.84 (t, J=5.6 Hz, 1H), 2.67-2.72 (m, 1H), 2.38 (s, 3H), 2.30 (m, 1H), 1.41 (s, 3H), 1.24 (d, J=9.6 Hz, 1H), 0.65 (s, 3H) (Fig.S19). HR ESI-MS (m/z): [M+H]+ Calcd. for C24H25N2+, 341.201 23; Found, 341.363 31.

    CN-substituted [Pt(N^C^N′)Cl] precursor (substituted at the 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CD2Cl2): δ 9.53 (dd, J1=6.0 Hz, J2=1.6 Hz, 1H), 8.62 (s, 1H), 8.11-8.15 (m, 2H), 7.46 (s, 1H), 7.42 (d, J=7.2 Hz, 1H), 7.29 (dd, J1=7.6 Hz, J2=6.0 Hz, 1H), 7.16 (t, J=7.6 Hz, 1H), 3.07 (d, J=2.4 Hz, 2H), 2.88 (t, J=5.6 Hz, 1H), 2.66-2.71 (m, 1H), 2.28 (m, 1H), 1.36 (s, 3H), 1.19 (d, J=10.0 Hz, 1H), 0.64 (s, 3H) (Fig.S20). HR ESI-MS (m/z): [M]+ Calcd. for C24H20ClN3Pt+, 580.099 35; Found, 580.201 56.

    CH3O-substituted [Pt(N^C^N′)Cl] precursor (substituted at 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CD2Cl2): δ 8.92 (dd, J1=5.6 Hz, J2=1.2 Hz, 1H), 8.70 (s, 1H), 7.93 (dd, J1=7.6 Hz, J2=0.8 Hz, 1H), 7.43-7.46 (m, 2H), 7.33 (dd, J1=7.6 Hz, J2=0.8 Hz, 1H), 7.18 (dd, J1=8.8 Hz, J2=5.6 Hz, 1H), 7.11 (t, J=7.6 Hz, 1H), 3.97 (s, 3H), 3.03 (d, J=2.4 Hz, 2H), 2.86 (t, J=5.6 Hz, 1H), 2.63-2.69 (m, 1H), 2.25 (m, 1H), 1.34 (s, 3H), 1.16 (d, J=9.6 Hz, 1H), 0.62 (s, 3H) (Fig.S21). HR ESI-MS (m/z): [M]+ Calcd. for C24H23ClN2OPt+, 585.114 67; Found, 585.265 63.

    CF3-substituted [Pt(N^C^N′)Cl] precursor (substituted at 3‑position of the lateral pyridine ring). 1H NMR (400 MHz, CD2Cl2): δ 8.80 (dd, J1=5.6 Hz, J2=1.2 Hz, 1H), 8.79 (s, 1H), 8.31 (d, J=7.6 Hz, 1H), 7.88 (d, J=6.8 Hz, 1H), 7.56 (s, 1H), 7.52 (d, J=7.6 Hz, 1H), 7.42 (dd, J1=8.0 Hz, J2=5.6 Hz, 1H), 7.27 (t, J=7.6 Hz, 1H), 3.14 (d, J=2.4 Hz, 2H), 2.96 (t, J=5.6 Hz, 1H), 2.72-2.78 (m, 1H), 2.35 (m, 1H), 1.43 (s, 3H), 1.25 (d, J=10.0 Hz, 1H), 0.71 (s, 3H) (Fig.S22). HR ESI-MS (m/z): [M]+ Calcd. for C24H20ClF3N2Pt+, 623.091 48; Found, 623.233 71.

    Cl-substituted [Pt(N^C^N′)Cl] precursor (substituted at the 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CD2Cl2): δ 9.56 (dd, J1=5.6 Hz, J2=1.6 Hz, 1H), 8.85 (s, 1H), 7.30 (d, J=7.6 Hz, 1H), 7.95 (dd, J1=8.0 Hz, J2=1.6 Hz, 1H), 7.48 (s, 1H), 7.45 (d, J=7.2 Hz, 1H), 7.19-7.24 (m, 2H), 3.11 (d, J=2.0 Hz, 2H), 2.95 (t, J=5.6 Hz, 1H), 2.70-2.76 (m, 1H), 2.34 (m, 1H), 1.42 (s, 3H), 1.23 (d, J=10.0 Hz, 1H), 0.69 (s, 3H) (Fig.S23). HR ESI-MS (m/z): [M]+ Calcd. for C23H20Cl2N2Pt+, 589.065 13; Found, 589.232 87.

    CH3-substituted [Pt(N^C^N′)Cl] precursor (substituted at 3-position of the lateral pyridine ring). 1H NMR (400 MHz, CD2Cl2): δ 9.43 (dd, J1=5.6 Hz, J2=1.2 Hz, 1H), 8.89 (s, 1H), 7.71 (d, J=6.8 Hz, 1H), 7.67 (d, J=8.0 Hz, 1H), 7.47 (s, 1H), 7.40 (d, J=7.2 Hz, 1H), 7.21 (t, J=7.6 Hz, 1H), 7.16 (dd, J1=7.6 Hz, J2=5.6 Hz, 1H), 3.10 (d, J=2.4 Hz, 2H), 2.95 (t, J=5.6 Hz, 1H), 2.76 (s, 3H), 2.70-2.74 (m, 1H), 2.34 (m, 1H), 1.42 (s, 3H), 1.23 (d, J=10.0 Hz, 1H), 0.69 (s, 3H) (Fig.S24). HR ESI-MS (m/z): [M]+ Calcd. for C24H23ClN2Pt+, 569.119 75; Found, 569.275 16.

    The crystals of 5B and 7B, suitable for X-ray diffraction, have been obtained in mixed CH2Cl2/methanol (1∶2, V/V) solutions. Single-crystal X-ray diffraction measurements were performed on a Xcalibur, Atlas, and Gemini Ultra. Intensities were collected with graphite monochromatized Mo radiation (λ= 0.071 073 nm) operating at 50 kV and 30 mA, using ω/2θ scan mode. The data reduction was made with the Bruker SAINT package. Absorption corrections were performed using the SADABS program. The structures were solved by direct methods and refined on F2 by full-matrix least-squares using SHELXL‑2018/3 with anisotropic displacement parameters for all non‑hydrogen atoms in the two structures. Hydrogen atoms bonded to carbon atoms were placed in calculated positions and refined as riding mode, with C—H length of 0.093 nm (methane) or 0.096 nm (methyl) and Uiso(H)=1.2Ueq(Cmethane) or Uiso(H)=1.5Ueq(Cmethyl). All computations were carried out using the SHELXL-2018/3 program package[20].

    The calculations of structure optimization and transition state searching were carried out using the Gaussian program suite (Gaussian 16 Revision B.01). The PBE0 hybrid functional with the assistance of GD3BJ dispersion-correction scheme was introduced, and the def2-TZVP basis set and the corresponding ECP for Pt were used[21-24]. The transition state structures were followed by the intrinsic reaction coordinate calculations[25].

    Here, we continue to utilize the non-substituted/substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] precursors and CNXyl as starting materials (Scheme 1). For all [Pt(N^C^N′)Cl] precursors, after reacting with CNXyl in CH2Cl2 for 30 min at RT, the reactants [Pt(N^C^N′)Cl] reacted well and could not be detected through TLC analysis. Taking non-substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] as the reactant, we examined the ESI-MS spectrum of the reacted solution. As revealed in the ESI-MS spectrum (Fig.2), positive m/z peaks regarding [Pt(N^C^N′)(CNXyl)]+ and [Pt(N^C^N′)(CNXyl)Cl+H]+ were found at 651.207 86 and 687.184 13, respectively. Therefore, we assumed that the two isomers ([2A]Cl and 2B) should coexist in the reaction equilibrium (Scheme 1). Furthermore, concerning our previously obtained pure [1A]OTf and 1B[19], the 195Pt NMR spectrum of the above-reacted solution was also consistent with the presence of two isomers in solution (Fig.3). Two chemical shifts of 195Pt were distinguished in the reaction product[26-28]. The 195Pt signal of positive [Pt(κ3-N^C^N′)(CNXyl)]+ ions ([2A]+) resided at δ=-3 636.70. In contrast, the 195Pt signal of neutral [Pt(κ2-N^C-N′)(CNXyl)Cl] complex (2B) was located at δ=-3 789.04. Therefore, we are convinced by the presence of a coordination equilibrium between the κ3‑N^C^N′ coordinated [Pt(N^C^N′)(CNXyl)]+ and the κ2-N^C^N′ coordinated [Pt(N^C-N′)(CNXyl)Cl] in the reaction of [Pt(N^C^N′)Cl] and CNXyl.

    Scheme 1

    Scheme 1.  Reaction between the non-substituted/substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] precursor and CNXyl

    Figure 2

    Figure 2.  ESI-MS spectrum of the reacted solution between the non-substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] and CNXyl

    Figure 3

    Figure 3.  195Pt NMR spectra of the reacted solution between the non-substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] and CNXyl

    In addition, we still used the previously obtained pure [1A]+ and 1B as references[19], and their 1H signals were assigned by measuring 2D NMR spectra (Fig.S1-S14). The complex [1A]+ in CD2Cl2 showed a characteristic doublet peak (δ=8.66) in the low field of 1H NMR spectrum, which came from the H15 signal next to the N atom of the pyridine ring without pinene groups (Fig.4). In contrast, the 1H NMR spectrum featured a singlet peak (δ=9.08) in the low field for 1B. The peak was caused by a signal from the H12 atom next to the N atom of the pyridine ring with pinene groups (Fig.4). Not controlling the chloride source, we just performed the direct reaction of difluoro-substituted [Pt(N^C^N′)Cl] and CNXyl in CH2Cl2 for 30 min at RT. It was found that the 1H NMR spectrum of the reaction product showed only a singlet at δ 9.08 with no signals detected at 8.66. So, we believe that only 1B was obtained completely, and 1B surpassed [1A]+ overwhelmingly in the coordination equilibrium.

    Figure 4

    Figure 4.  Characteristic 1H NMR signals of [1A]+ and 1B in the lowest field

    Whether isomer B always prevails over isomer [A]Cl in the coordination equilibrium, the reactions between CNXyl and [Pt(N^C^N′)Cl] with different substituents at the 3-position of the lateral pyridine ring have been inspected (Scheme 1). In principle, the molar ratio between the isomers would vary with the introduction of electron-withdrawing and electron-donating groups. However, isomer B dominated the product composition (Table 1), and only an insignificant amount of [A]+ was present, which could be neglected according to the NMR spectra (Fig.S25-S31). For all the substituents at the 3-position of the lateral pyridine ring, a prominent singlet at ca. 9.0 was observed in 1H NMR spectra (Fig.S25-S31). Accordingly, the variance of substituents may have a small impact on the equilibrium between two isomers.

    Table 1

    Table 1.  Molar ratios of [A]+ and B with different substituents at the 3-position of the lateral pyridine ring
    下载: 导出CSV
    Substituents at the 3-position of the lateral pyridine ring $ {n}_{{\left[{\rm{A}}\right]}^{+}} $nB
    —H 42∶58
    —CH3 8∶92
    —OCH3 9∶91
    —CF3 0∶100
    —F 0∶100
    —Cl 3∶97
    —CN 2∶98

    Apart from the influence of substituent types[29], the repulsion between the substituents of the pyridine ring and the H atom of the benzene ring may be an essential factor in affecting coordination equilibrium and generating a specific product[30-32]. As shown in Table 1, the isomer A/B molar ratios in solutions of different substituted precursors reacting with CNXyl have been obtained (Table 1). To eliminate repulsion as much as possible, we examined the reaction between the non-substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] precursor and CNXyl. Through monitoring 1H signals in the low field (Fig.S25), it can be found that the A/B molar ratio is 42∶58. According to our previous studies, cationic [1A]+ exhibited stronger emission than neutral 1B in CH2Cl2 solution[19]. As revealed in Fig.S32, the reacted solution between non‑substituted [Pt(N^C^N′)Cl] and CNXyl displayed a more intense emission than those containing substituted mixtures, further confirming a distinct component of [2A]+ in the reacted mixture between non-substituted [Pt(N^C^N′)Cl] and CNXyl. Therefore, the proportion of [A]+ increased significantly as any substituent was removed to reduce the repulsion. The repulsion may play a more substantial role in tuning the reaction equilibrium. Eliminating steric hindrance should favor the formation of tridentate N^C^N-coordinated [A]+. In addition, the variation of solvent and temperature would exert an influence on the product[33], and increasing the polarity of the reaction solution could increase the molar ratio of [A]+. However, the change of solvent and temperature would induce the shift or broadening of 1H NMR signals. Solvent and temperature effects on reaction equilibrium are complex and are being investigated.

    The structures of CF3- and Cl-substituted isomer B (5B and 7B) have been confirmed by single-crystal X-ray diffraction (Table 2 and Table S1-S2), revealing a slightly distorted square planar geometry with trans-NN^C^N′, CNR configuration (Fig.5)[34-37]. Inspecting the structural characteristics of [A]+ and B, we also found that B is superior to [A]+ in the coordination equilibrium, revealing the difference in structural stability. According to reported crystal structural data[19, 34-37], for cationic tridentate [A]+, the trans N—Pt—N angles (158°-163°) deviate severely from linearity, and this configuration suffers from a significantly chelating ring strain. In addition, strong Pt—C(aryl) and Pt—C(isocyanide) bonds occupy the trans-position, and in this case, the central Pt atom is pulled tightly by two strong Pt—C bonds, possibly resulting in the instability of tridentate N^C^N-coordinated [A]+. However, neutral bidentate isomer B keeps a trans-NN^C^N, CNR configuration of Pt—N(pyridine) and Pt—C(isocyanide) (Fig.5). The chloride ion resides at the trans position relative to the C atom of the cyclometalated ligand. The trans C—Pt—Cl and C—Pt—N angles (174°-179°) are close to linearity. The central Pt atom is pulled by a strong Pt—C bond and a weak Pt—Cl(N) bond with one tension and one relaxation in this trans position. The coordination configuration of isomer B seems much stable than that of [A]Cl.

    Table 2

    Table 2.  Crystallographic data of 5B and 7B
    下载: 导出CSV
    Parameter 5B 7B
    Formula C33H29ClF3N3Pt C32H29Cl2N3Pt
    Formula weight 755.13 721.57
    Crystal system Monoclinic Orthorhombic
    Space group P21 P212121
    a / nm 1.334 82(8) 1.315 94(8)
    b / nm 1.565 91(9) 1.444 34(8)
    c / nm 1.433 93(7) 1.510 39(11)
    β / (°) 90.544(5)
    V / nm3 2.997 1(3) 2.870 7(3)
    Z 4 4
    T / K 293(2) 293(2)
    Dc / (g·cm-3) 1.674 1.670
    μ / mm-1 4.817 5.100
    F(000) 1 480 1 416
    θ range / (°) 3.318-26.000 3.402-25.998
    Reflection measured 17 929 11 355
    Unique reflection 9 535 5 145
    Rint 0.075 8 0.046 1
    Reflection with I > 2σ(I) 6 308 3 658
    Number of parameters 747 335
    Goodness-of-fit on F2 1.024 1.076
    R1 [I > 2σ(I)] 0.069 1 0.059 6
    wR2 (all data) 0.185 2 0.180 3
    ρ)max, (Δρ)min / (e·nm-3) 2 842, -2 178 3 728, -1 539
    Flack parameter -0.041(13) -0.025(9)

    Figure 5

    Figure 5.  Crystal structures of CF3- and Cl-substituted isomers B with ellipsoid probability levels of 30%

    The structural superiority of isomer B to isomer [A]Cl in the reaction equilibrium has also been verified by the theoretical calculation[38]. As shown in Fig.6 and 7, relative to the starting reactant, the F-substitution stabilizes the neutral product type but destabilizes the cationic type. Both for R=H and R=F, the reaction route for the cationic product has lower forward and reverse barriers than the reaction route for the neutral product, so the neutral products are thermodynamically more stable. Before the Cl- ion is precipitated from the solution phase, the reverse reaction of route B is more difficult to occur for R=F, due to the much higher reverse barrier (126.8 kJ·mol-1 for R=F vs 107.3 kJ·mol-1 for R=H). And, when the Cl- ion is to be precipitated, the reverse reaction of route A is more unlikely to occur for R=H, thanks to the slightly lower reverse barrier (53.8 kJ·mol-1 for R=H vs 50.8 kJ·mol-1 for R=F).

    Figure 6

    Figure 6.  Free energy profiles of two reaction routes, respectively for R=H and R=F, calculated using PBE0(D3BJ)/def2-TZVP with the solvation effect of CH2Cl2

    Routes A and B are for the reactions towards the ionic and neutral type of products, respectively.

    Figure 7

    Figure 7.  Starting reactants, transition state structures, and product structures involved in routes A and B for R=H and R=F

    In summary, the non‑substituted/substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] precursors reacted with CNXyl by simple substitution to give two isomeric products, cationic [Pt(κ3-N^C^N′)(CNXyl)]Cl complex ([A]Cl) and neutral [Pt(κ2-N^C-N′)(CNXyl)Cl] complex (B). The reactions produce predominantly neutral isomer B regardless of substituents on the lateral pyridine. A theoretical study indicates that neutral products are thermodynamically more stable. In addition, F-substitution could stabilize the neutral product type, but destabilize the cationic type. The study guides separating cyclometalated Pt(Ⅱ) complexes with specific configurations.


    Acknowledgments: This work is supported by the National Natural Science Foundation of China (Grant No.22461019) and the Natural Science Foundation of Hainan Province (Grant No.ZDYF2023SHFZ106). We gratefully acknowledge Prof. Yang Xiaoliang (Nanjing University) for 195Pt NMR measurements. Supporting information is available at http://www.wjhxxb.cn
    1. [1]

      YUEN M Y, ROY V A L, LU W, KUI S C F, TONG G S M, SO M H, CHUI S S Y, MUCCINI M, NING J Q, XU S J, CHE C M. Semiconducting and electroluminescent nanowires self-assembled from organoplatinum(Ⅱ) complexes[J]. Angew. Chem.‒Int. Edit., 2008, 47: 9895-9899 doi: 10.1002/anie.200802981

    2. [2]

      WANG P, LEUNG C H, MA D L, SUN R W Y, YAN S C, CHEN Q S, CHE C M. Specific blocking of CREB/DNA binding by cyclometalated platinum(Ⅱ) complexes[J]. Angew. Chem.‒Int. Edit., 2011, 50: 2554-2558 doi: 10.1002/anie.201006887

    3. [3]

      LIU J, SUN R W Y, LEUNG C H, LOK C N, CHE C M. Inhibition of TNF-α stimulated nuclear factor-kappa B (NF-κB) activation by cyclometalated platinum(Ⅱ) complexes[J]. Chem. Commun., 2012, 48: 230-232 doi: 10.1039/C1CC15317J

    4. [4]

      XIAO X S, LU W, CHE C M. Phosphorescent nematic hydrogels and chromonic mesophases driven by intra- and intermolecular interactions of bridged dinuclear cyclometalated platinum(Ⅱ) complexes[J]. Chem. Sci., 2014, 5: 2482-2488 doi: 10.1039/c4sc00143e

    5. [5]

      CHEN Y, CHE C M, LU W. Phosphorescent organoplatinum(Ⅱ) complexes with a lipophilic anion: Supramolecular soft nanomaterials through ionic self-assembly and metallophilicity[J]. Chem. Commun., 2015, 51: 5371-5374 doi: 10.1039/C4CC08569H

    6. [6]

      KUWABARA J, YAMAGUCHI K, YAMAWAKI K, YASUDA T, NISHIMURA Y, KANBARA T. Modulation of the emission mode of a Pt(Ⅱ) complex via intermolecular interactions[J]. Inorg. Chem., 2017, 56: 8726-8729 doi: 10.1021/acs.inorgchem.7b00880

    7. [7]

      LIU Y, FENG J G, ZHANG B, WU Y C, CHEN Y, JIANG L. Regular aligned 1D single-crystalline supramolecular arrays for photodetectors[J]. Small, 2018, 14: 1701861 doi: 10.1002/smll.201701861

    8. [8]

      WAN Q Y, TO W P, CHANG X, CHE C M. Controlled synthesis of Pd and Pt supramolecular copolymer with sequential multiblock and amplified phosphorescence[J]. Chem, 2020, 6: 945-967 doi: 10.1016/j.chempr.2020.01.021

    9. [9]

      LIU M Y, HAN Y F, ZHONG H, ZHANG X L, WANG F. Supramolecular chirogenesis induced by platinum(Ⅱ) tweezers with excellent environmental tolerance[J]. Angew. Chem.‒Int. Edit., 2021, 60: 3498-3503 doi: 10.1002/anie.202012901

    10. [10]

      YANG F, LI H Y, LI H J, HE X. Manipulation of 1D and 2D self-assembly via geometry modulation of adamantane isocyanide Pt(Ⅱ) complexes[J]. Chem. Commun., 2024, 60: 8605-8608 doi: 10.1039/D4CC02400A

    11. [11]

      LU W, CHEN Y, ROY V A L, CHUI S S Y, CHE C M. Supramolecular polymers and chromonic mesophases self-organized from phosphorescent cationic organoplatinum(Ⅱ) complexes in water[J]. Angew. Chem.‒Int. Edit., 2009, 48: 7621-7625 doi: 10.1002/anie.200903109

    12. [12]

      ZHANG X P, CHANG V Y, LIU J, YANG X L, HUANG W, LI Y, LI C H, MULLER G, YOU X Z. Potential switchable circularly polarized luminescence from chiral cyclometalated platinum(Ⅱ) complexes[J]. Inorg. Chem., 2015, 54: 143-152 doi: 10.1021/ic5019136

    13. [13]

      LAI S W, LAM H W, LU W, CHEUNG K K, CHE C M. Observation of low-energy metal-metal-to-ligand charge transfer absorption and emission: Electronic spectroscopy of cyclometalated platinum(Ⅱ) complexes with isocyanide ligands[J]. Organometallics, 2002, 21: 226-234 doi: 10.1021/om0106276

    14. [14]

      ZHANG H H, JING J, WU S X, LUO Y P, SUN S S, ZHANG D S, SHI Z F, ZHANG X P. Tunable circularly polarized luminescence of sol-gels based on chiral cyclometalated platinum(Ⅱ) complexes[J]. Dyes Pigment., 2022, 201: 110228 doi: 10.1016/j.dyepig.2022.110228

    15. [15]

      CHEN Y, LU W, CHE C M. Luminescent pincer-type cyclometalated platinum(Ⅱ) complexes with auxiliary isocyanide ligands: Phase-transfer preparation, solvatomorphism, and self‑aggregation[J]. Organometallics, 2013, 32: 350-353 doi: 10.1021/om300965b

    16. [16]

      JING J, XU G, ZHANG H H, CHEN X H, ZHANG D S, HAN L Z, QI X W, SHI Z F, ZHANG X P. Enhanced circularly polarized luminescence in fluoro-substituted N^C^N-coordinating platinum(Ⅱ) complexes[J]. Inorg. Chim. Acta, 2022, 541: 121067 doi: 10.1016/j.ica.2022.121067

    17. [17]

      KNOF U, ZELEWSKY A V. Predetermined chirality at metal centers[J]. Angew. Chem.‒Int. Edit., 1999, 38: 302-322 doi: 10.1002/(SICI)1521-3773(19990201)38:3<302::AID-ANIE302>3.0.CO;2-G

    18. [18]

      MAMULA O, ZELEWSKY A V. Supramolecular coordination compounds with chiral pyridine and polypyridine ligands derived from terpenes[J]. Coord. Chem. Rev., 2003, 242: 87-95 doi: 10.1016/S0010-8545(03)00062-6

    19. [19]

      ZHANG H H, XU G, LV X S, WU S X, JING J, ZHAO Y, LIU J, CHRN X H, ZHANG X P. Mechanochromism and circularly polarized luminescence of chiral cyclometalated platinum(Ⅱ) complexes with different configurations and aggregated forms[J]. Dyes Pigment., 2024, 221: 111833 doi: 10.1016/j.dyepig.2023.111833

    20. [20]

      SHELDRICK G M. A short history of SHELX[J]. Acta Crystallogr. Sect. A, 2008, A64: 112-122

    21. [21]

      PERDEW J P, BURKE K, ERBZERHOF M. Generalized gradient approximation made simple[J]. Phys. Rev. Lett., 1996, 77: 3865-3868 doi: 10.1103/PhysRevLett.77.3865

    22. [22]

      ADAMO C, BARONE V. Toward reliable density functional methods without adjustable parameters: The PBE0 model[J]. J. Chem. Phys., 1999, 110: 6158-6170 doi: 10.1063/1.478522

    23. [23]

      GRIMME S, EHRLICH S, GOERIGK L. Effect of the damping function in dispersion corrected density functional theory[J]. J. Comput. Chem., 2011, 32: 1456-1465 doi: 10.1002/jcc.21759

    24. [24]

      WEIGEND F, AHLRICHS R. Balanced basis sets of split valence, triple zeta valence and quadruple zeta valence quality for H to Rn: Design and assessment of accuracy[J]. Phys. Chem. Chem. Phys., 2005, 7: 3297-3305 doi: 10.1039/b508541a

    25. [25]

      FUKUI K. The path of chemical reactions‒the IRC approach[J]. Accounts Chem. Res., 1981, 14: 363-368 doi: 10.1021/ar00072a001

    26. [26]

      PAZDERSKI L, PAWLAK T, SITKOWSKI J, KOZERSKI L, SZLYK E. 1H, 13C, 15N and 195Pt NMR studies of Au(Ⅲ) and Pt(Ⅱ) chloride organometallics with 2-phenylpyridine[J]. Magn. Reson. Chem., 2009, 47: 932-941 doi: 10.1002/mrc.2491

    27. [27]

      HARPER B W J, ALDRICH-WRIGHT J R. The synthesis, characterisation and cytotoxicity of bisintercalating (2,2′∶6′,2″-terpyridine)platinum(Ⅱ) complexes[J]. Dalton Trans., 2015, 44: 87-96 doi: 10.1039/C4DT02773F

    28. [28]

      EVSTIGNEEV M P, LANTUSHENKO A O, YAKOVLEVA Y A, SULEYMANOVA A F, ELTSOV O S, KOZHEVNIKOV V N. Tuning the aggregation of N^N^C Pt(Ⅱ) complexes by varying the aliphatic side chain and auxiliary halide ligand: 1H and 195Pt NMR investigation[J]. Eur. J. Inorg. Chem., 2019: 4122-4128

    29. [29]

      ESTERUELAS M A, MORENO-BLÁZQUEZ S, OLIVÁN M, OÑATE E. Competition between N,C,N-pincer and N,N-chelate ligands in platinum(Ⅱ)[J]. Inorg. Chem., 2023, 62: 10152-10170 doi: 10.1021/acs.inorgchem.3c00694

    30. [30]

      CHEN J L, LIN C H, CHEN J H, CHI Y, CHIU Y C, CHOU P T, LAI C H, LEE G H, CARTY A J. Reactions of the (2-pyridyl) pyrrolide platinum(Ⅱ) complex driven by sterically encumbered chelation: A model for the reversible attack of alcohol at the coordinated carbon monoxide[J]. Inorg. Chem., 2008, 47: 5154-5161 doi: 10.1021/ic800117v

    31. [31]

      NAKATA N, IKEDA T, ISHII A. Syntheses of selenolato-bridged dinuclear hydridoplatinum complexes [Pt2H2(μ-SetBu)2(PPh3)2] and [Pt2H(SetBu)(μ-SetBu)2(PPh3)2]: Unusual thermal reaction of hydrido(1, 1-dimethylethaneselenolato) platinum complex cis-[PtH(SetBu)(PPh3)2][J]. Inorg. Chem., 2010, 49: 8112-8116 doi: 10.1021/ic1011742

    32. [32]

      ONGOMA P O, JAGANYI D. Mechanistic elucidation of linker and ancillary ligand substitution reactions in Pt(Ⅱ) dinuclear complexes[J]. Dalton Trans., 2013, 42: 2724-2734 doi: 10.1039/C2DT31956J

    33. [33]

      ZUCCA A, CORDESCHI D, STOCCORO S, CINELLU M A, MINGHRTTI G, CHELUCCI G, MANASSERO M. Platinum(Ⅱ)-cyclometalated "roll-over" complexes with a chiral pinene-derived 2, 2′-bipyridine[J]. Organometallics, 2011, 30: 3064-3074 doi: 10.1021/om200172h

    34. [34]

      DÍEZ Á, FORNIÉS J, LARRAZ C, LALINDE E, LÓPEZ J A, MARTÍN A, MORENO M T, SICILIA V. Structural and luminescence studies on ππ and Pt…Pt interactions in mixed chloro-isocyanide cyclometalated platinum(Ⅱ) complexes[J]. Inorg. Chem., 2010, 49: 3239-3251 doi: 10.1021/ic902094c

    35. [35]

      MARTÍNEZ-JUNQUERA M, LARA R, LALINDE E, MORENO M T. Isomerism, aggregation-induced emission and mechanochromism of isocyanide cycloplatinated(Ⅱ) complexes[J]. J. Mater. Chem. C, 2020, 8: 7221-7233 doi: 10.1039/D0TC01163K

    36. [36]

      MARTÍNEZ-JUNQUERA M, LALINDE E, MORENO M T. Multistimuli-responsive properties of aggregated isocyanide cycloplatinated(Ⅱ) complexes[J]. Inorg. Chem., 2022, 61: 10898-10914 doi: 10.1021/acs.inorgchem.2c01400

    37. [37]

      SOKOLOVA E V, KINZHALOV M A, SMIRNOV A S, CHERANYOVA A M, IANOV D M, KUKUSHKIN V Y, BOKACH N A. Polymorph-dependent phosphorescence of cyclometalated platinum(Ⅱ) complexes and its relation to non-covalent interactions[J]. ACS Omega, 2022, 7: 34454-34462 doi: 10.1021/acsomega.2c04110

    38. [38]

      BUTSCHKE B, SCHWARZ H. "Rollover" cyclometalation—Early history, recent developments, mechanistic insights and application aspects[J]. Chem. Sci., 2012, 3: 308-326 doi: 10.1039/C1SC00651G

  • Figure 1  Chemical structures of reported difluoro-substituted N^C^N′-coordinated Pt(Ⅱ)-isocyanide complexes[19]

    Scheme 1  Reaction between the non-substituted/substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] precursor and CNXyl

    Figure 2  ESI-MS spectrum of the reacted solution between the non-substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] and CNXyl

    Figure 3  195Pt NMR spectra of the reacted solution between the non-substituted unsymmetric pinene-derived [Pt(N^C^N′)Cl] and CNXyl

    Figure 4  Characteristic 1H NMR signals of [1A]+ and 1B in the lowest field

    Figure 5  Crystal structures of CF3- and Cl-substituted isomers B with ellipsoid probability levels of 30%

    Figure 6  Free energy profiles of two reaction routes, respectively for R=H and R=F, calculated using PBE0(D3BJ)/def2-TZVP with the solvation effect of CH2Cl2

    Routes A and B are for the reactions towards the ionic and neutral type of products, respectively.

    Figure 7  Starting reactants, transition state structures, and product structures involved in routes A and B for R=H and R=F

    Table 1.  Molar ratios of [A]+ and B with different substituents at the 3-position of the lateral pyridine ring

    Substituents at the 3-position of the lateral pyridine ring $ {n}_{{\left[{\rm{A}}\right]}^{+}} $nB
    —H 42∶58
    —CH3 8∶92
    —OCH3 9∶91
    —CF3 0∶100
    —F 0∶100
    —Cl 3∶97
    —CN 2∶98
    下载: 导出CSV

    Table 2.  Crystallographic data of 5B and 7B

    Parameter 5B 7B
    Formula C33H29ClF3N3Pt C32H29Cl2N3Pt
    Formula weight 755.13 721.57
    Crystal system Monoclinic Orthorhombic
    Space group P21 P212121
    a / nm 1.334 82(8) 1.315 94(8)
    b / nm 1.565 91(9) 1.444 34(8)
    c / nm 1.433 93(7) 1.510 39(11)
    β / (°) 90.544(5)
    V / nm3 2.997 1(3) 2.870 7(3)
    Z 4 4
    T / K 293(2) 293(2)
    Dc / (g·cm-3) 1.674 1.670
    μ / mm-1 4.817 5.100
    F(000) 1 480 1 416
    θ range / (°) 3.318-26.000 3.402-25.998
    Reflection measured 17 929 11 355
    Unique reflection 9 535 5 145
    Rint 0.075 8 0.046 1
    Reflection with I > 2σ(I) 6 308 3 658
    Number of parameters 747 335
    Goodness-of-fit on F2 1.024 1.076
    R1 [I > 2σ(I)] 0.069 1 0.059 6
    wR2 (all data) 0.185 2 0.180 3
    ρ)max, (Δρ)min / (e·nm-3) 2 842, -2 178 3 728, -1 539
    Flack parameter -0.041(13) -0.025(9)
    下载: 导出CSV
  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  58
  • HTML全文浏览量:  4
文章相关
  • 发布日期:  2025-09-10
  • 收稿日期:  2025-04-20
  • 修回日期:  2025-06-17
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