Next Article in Journal
Insights into Pd-Nb@In2Se3 Electrocatalyst for High-Performance and Selective CO2 Reduction Reaction from DFT
Next Article in Special Issue
A Pt(II) Complex with a PNN Type Ligand Dppmaphen Exhibits Selective, Reversible Vapor-Chromic Photoluminescence
Previous Article in Journal
Novel Trimethoprim-Based Metal Complexes and Nanoparticle Functionalization: Synthesis, Structural Analysis, and Anticancer Properties
Previous Article in Special Issue
Synthesis and Molecular Structure of Iron(III) Diaryl-Dithiocarbamate Complexes, [Fe(S2CNAr2)3], and a Preliminary Study Exploring Their Potential as Single-Source Precursors for Nanoscale Iron Sulfides
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cadmium Complexes—A Novel Family in the Coordination Chemistry of 1,2-bis(arylimino)acenaphthenes

by
Egor V. Chekhov
1,2,
Ivan V. Bakaev
1,
Alisa K. Gukova
1,2,
Nikolay O. Shaposhnikov
1,2,
Veronika I. Komlyagina
1,
Saltanat Appazova
3,
Banu Diyarova
4,*,
Klara Darmagambet
3,
Nurbol Appazov
3,
Nikolai F. Romashev
1 and
Artem L. Gushchin
1,*
1
Nikolaev Institute of Inorganic Chemistry SB RAS, Novosibirsk 630090, Russia
2
Department of Natural Sciences, Novosibirsk State University, Novosibirsk 630090, Russia
3
Department of Engineering and Technology, Institute of Engineering and Technology, Korkyt Ata Kyzylorda University, Ayteke bi Street 29A, Kyzylorda 120014, Kazakhstan
4
Department of Chemistry, Institute of Natural Sciences, Kazakh National Women’s Teacher Training University, Gogol Street 114/8, Almaty 050000, Kazakhstan
*
Authors to whom correspondence should be addressed.
Inorganics 2025, 13(5), 145; https://doi.org/10.3390/inorganics13050145
Submission received: 1 April 2025 / Revised: 28 April 2025 / Accepted: 29 April 2025 / Published: 2 May 2025

Abstract

:
This work presents the synthesis routes for the first representatives of cadmium complexes based on 1,2-bis(arylimino)acenaphthene (Ar-bian). The reaction of CdCl2 with bis-(2,4,6-trimethylphenylimino)acenaphthene (tmp-bian) in a 1-to-1 molar ratio led to a dimeric complex [Cd2(tmp-bian)2Cl2(µ-Cl)2] (1). Further treatment of complex 1 with silver triflate as a chloride-eliminating agent, followed by the addition of one equivalent of tmp-bian, resulted in the formation of a mixture consisting of [Cd2(tmp-bian)2(H2O)4(µ-Cl)2](OTf)2 (2) and [Cd(tmp-bian)2(OTf)2] (3). To obtain complex 3 in its individual form, a reaction of Cd(OTf)2 with two equivalents of tmp-bian was carried out. The characterization of the complexes was conducted through a range of analytical methods, including X-ray diffraction analysis, elemental analysis, as well as IR and 1H NMR-spectroscopies. Redox properties of 1 and 3 were investigated by means of cyclic voltammetry. Cyclic voltammograms revealed irreversible reduction processes centered on the tmp-bian ligand, which were confirmed by quantum chemical calculations.

1. Introduction

1,2-Bis(arylimino)acenaphthene (Ar-bian) is a well-studied class of aromatic acceptor diimines with versatile coordination and redox chemistry [1,2,3,4]. Ar-bian is composed of a central 1,4-diazabutadiene fragment, which is complemented by a naphthalene backbone. The combination of these elements gives rise to strong σ-donor and π-acceptor properties, thereby ensuring the stabilization of metal ions in both high and low oxidation states. Furthermore, the rigid naphthalene moiety stabilizes the anti-anti conformation of the α-diimine fragment, facilitating strong chelation to the metal center.
To date, metal complexes based on Ar-bian with almost all d-block metals have been described in the literature [3,5,6,7,8,9]. However, despite the large number of publications, such complexes have been the focus of research in a highly uneven manner. For example, an impressive number of publications are devoted to copper, nickel, and palladium complexes [10,11,12,13,14,15]. This is mainly due to their demand for various catalytic applications [16,17]. A similar trend continues for the metals of the 12th group of the Periodic Table. At present, approx. 60 structurally characterized zinc complexes are known, most of which are mononuclear complexes of the type [Zn(Ar-bian)Cl2]. They are formed as intermediates in the synthesis of free Ar-bian ligands, which determines their demand [18]. This contrasts with the complete absence of structurally characterized cadmium analogues. Meanwhile, mercury complexes based on Ar-bian are confined to only three examples. Ferreira et al. demonstrated the potential for utilizing HgX2 (X = Cl, SCN) (rather than ZnCl2) in the template synthesis of Ar-bian. Respectively, the reaction of HgX2 with acenaphthenequinone and 4-tert-butylaniline resulted in the formation of [Hg(2-tBu-bian)Cl2] or [Hg(2-tBu-bian)(SCN)2] [19]. In a similar manner, El Ayaan obtained the complex [Hg(dpp-bian)Cl2] [20].
In the present study, we report the synthesis and crystal structures of the first cadmium/BIAN complexes: dimeric [Cd2(tmp-bian)2Cl2(µ-Cl)2] (1), [Cd2(tmp-bian)2(H2O)4(µ-Cl)2](OTf)2 (2) and monomeric [Cd(tmp-bian)2(OTf)2] (3) (tmp-bian = 1,2-bis[(2,4,6-trimethylphenyl)imino]acenatephene).

2. Results and Discussion

2.1. Synthesis and Characterization

The reaction of CdCl2 with 1 equivalent of tmp-bian in ethanol afforded the binuclear complex [Cd2(tmp-bian)2Cl2(μ-Cl)2] (1) in high yield (90%). Evans et al. demonstrated the possibility of the formation of analogous dimeric zinc(II) complexes [Zn2(4-OCF3-C6H4-bian)2Cl2(μ-Cl)2] and [Zn2(4-CF3-C6H4-bian)2Cl2(μ-Cl)2] [21]. Treatment of complex 1 with four equivalents of AgOTf led to a mixture consisting of [Cd2(tmp-bian)2(H2O)4(μ-Cl)2](OTf)2 (2) and [Cd(tmp-bian)2(OTf)2] (3) in approximately equal proportions. Unfortunately, we were unable to isolate the complexes individually from this mixture, even using column chromatography. Alternatively, complex 3 was synthesized in 85% yield by the reaction of Cd(OTf)2 with tmp-bian in a 1:2 molar ratio. Attempts to obtain complex 2 in pure form by reacting complex 1 with two equivalents of AgOTf and tmp-bian were unsuccessful. The general synthetic scheme is shown in Figure 1. The purity of compounds 1 and 3 was confirmed by elemental analysis.
The IR spectrum of complex 1 showed typical ν(C-H) vibration bands in the region of 3078–2817 cm−1. ν(C=N) and ν(C-C) vibration bands were detected at 1660–1590 cm−1 and 1477 cm−1, respectively, which is consistent with the neutral state of tmp-bian. In the IR spectrum of complex 3, ν(C-H) vibration bands were observed in the region of 3070–2831 cm−1. ν(C=N) and ν(C-C) bands were detected at 1658–1585 cm−1 and 1477 cm−1, respectively. In addition, typical vibrations of the triflate group were observed at 1238, 1028, and 637 cm−1 (Figures S1 and S2).
The 1H NMR spectrum of complex 1 revealed characteristic signals of acenaphthene moiety of Ar-bian, appearing as two doublets at 8.10 and 6.55 ppm and one triplet at 7.55 ppm. The signals of the aryl rings appeared as a singlet at 7.05 ppm, as well as two upfield singlets at 2.25 and 2.45 ppm. Note that two tmp-bian ligands in 1 are equivalent to each other. The 1H NMR spectrum of complex 3 has a similar pattern. The signals of the acenaphthene fragment appeared as two doublets and one triplet at 8.14, 6.91, and 7.59 ppm. The signals of the aryl rings appeared as a singlet at 7.09, 2.12, and 2.09 ppm (Figures S3 and S4).

2.2. X-Ray Structure Description

Single crystals of 1•2(C2H5)2O and 3•1.2CH2Cl2 suitable for X-ray diffraction analysis, were obtained by slow diffusion of diethyl ether into a solution of 1 and 3 in CH2Cl2. Single crystals of 2, suitable for XRD, were obtained by slow evaporation of a mixture of complexes 2 and 3 formed after the corresponding reaction (see Figure 1).
The molecular structures of neutral complexes 1 and 3, as well as cationic complex 2, are shown in Figure 2. The values of selected geometric parameters are given in Table 1.
In complex 1, the cadmium atoms have a highly distorted square-pyramidal coordination environment (τ = 0.20), consisting of two nitrogen atoms of tmp-bian, located in the equatorial plane, a chlorine atom located at the apex of the pyramid, and two bridging chlorine atoms that connect the {Cd(tmp-bian)Cl} fragments into a binuclear structure. Each cadmium ion is located above the equatorial plane at a distance of about 0.8 Å. Tmp-bian is coordinated to the Cd(II) ion asymmetrically: the Cd–N1 and Cd–N2 distances are 2.354(2) Å and 2.427(2) Å, respectively. The C–N1 (1.280(3) Å), C–N2 (1.269(3) Å), and C-C (1.514(3) Å) bond lengths indicate the neutral state of the tmp-bian ligand. An asymmetric coordination of the bridging ligands is also observed: the Cd–(μ-Cl) distances are 2.5944(8) Å and 2.5100(7) Å, respectively. The Cd–Cl distance is 2.4211(8) Å.
It is noteworthy that in the literature, there are examples of dimeric centrosymmetric Ni(II) and Zn(II) complexes with Ar-bian ligands of the composition [{M(tmp-bian)(X)}2(μ-X)2] (X = Cl, Br, I). Two variants of the coordination environment of the metal ion are possible. In the first case, the metal ion has a distorted square-pyramidal environment, characteristic of nickel complexes [22,23,24,25,26,27]. In the second case, the coordination polyhedron is a distorted trigonal bipyramid, which is observed in both zinc and nickel complexes [21,28].
In complex 2, the cadmium atoms have a distorted octahedral coordination environment, consisting of two nitrogen atoms of tmp-bian, two oxygen atoms of water, and two bridging chlorine atoms that connect the {Cd(tmp-bian)(H2O)2} fragments into a binuclear structure. In the outer sphere, there are two triflate anions. Unlike complex 1, in complex 2, tmp-bian is coordinated almost symmetrically: the Cd–N1 and Cd–N2 bond lengths are 2.375(3) and 2.362(2) Å, respectively. Cd–O and Cd–Cl distances are 2.329(2), 2.349(3) Å and 2.559(1), 2.6000(9) Å, respectively. The C–N1 (1.278(4) Å), C–N2 (1.271(5) Å), and C-C (1.528(5) Å) distances indicate the neutral state of the tmp-bian ligand [29,30].
Complex 3 is a mononuclear Cd(II) complex in which the metal ion has an octahedral coordination geometry. In the equatorial plane, there are four nitrogen atoms of two tmp-bian ligands, and the axial positions are occupied by oxygen atoms of the coordinated triflate anion. The tmp-bian molecules are coordinated asymmetrically: the Cd–N distances are 2.314(2) and 2.391(3) Å. Cd–O distances are 2.381(2) Å. The C–N and C-C bond lengths are 1.275(4) Å, 1.288(4) Å and 1.515(5) Å, respectively. A similar octahedral bis-chelated Mn(II) complex [Mn(tmp-bian)2(ClO4)2] has been reported [31].

2.3. Electrochemical Properties

The electrochemical properties of complexes 1 and 3 were investigated using cyclic voltammetry (Figure 3). The electrochemical properties of complex 2 were not studied because it was not isolated in individual form. The CV of complex 1 in dichloromethane showed a cathodic peak at Ec = −1.01 V (vs. Ag/AgCl) with the corresponding anodic counter peak at Ea = −0.80 V. The large difference between these peaks (0.21 V) indicates the irreversibility of the redox process. For complex 3, two reduction peaks were detected at Ec = −0.56 V and Ec = −0.73 V (vs. Ag/AgCl); an anodic counter peak appeared at −0.57 V. Oxidative processes in the range of 0–2 V were not detected for either complex. The reduction processes found for 1 and 3 are in good agreement with the electron-accepting capacity of Ar-bian. Irreversible reduction at −1.69 V (vs. Ag/AgCl) was previously shown for the free tmp-bian ligand [15].
To confirm the ligand-centered character of redox processes, quantum chemical calculations were carried out within the framework of density functional theory. The X-ray diffraction data were utilized as the initial geometry.Optimized geometry coordinates of 13 in the gas phase are presented in Tables S2–S4. The identification of the minimum point on the potential energy surface was substantiated by the absence of imaginary vibrational frequencies.
The optimized geometries of 1, the cation of 2, and 3 are consistent with the X-ray diffraction data. However, in the case of complex 1, the structure of the {Cd2(μ-Cl)2} moiety differed from the initial geometry. In the optimized geometry, the angle between [μ-Cl1 Cd1 μ-Cl2] and [μ-Cl1 Cd2 μ-Cl2] planes is 40.2°, whereas, according to X-ray diffraction data, the Cd and μ-Cl atoms lie in the same plane. An attempt to rectify the inversion center during geometry optimization resulted in the appearance of an imaginary frequency.
The composition of frontier molecular orbitals for 1, the cation of 2, and 3 are presented in Table 2. The views of HOMOs and LUMOs of 1–3 are shown in Figure 4. Remarkably, HOMOs and LUMOs are almost completely localized on the tmp-bian ligand; however, the contribution of cadmium and chlorine atomic orbitals to the HOMO is approx. 19% for 1. It is worth noting that the LUMO is mainly localized on the acenaphthene and diimine fragments of tmp-bian, whereas the HOMO is localized on the trimethylphenyl rings. The ligand-centered nature of the LUMO indicates that the observed redox processes are most likely consistent with the reduction of tmp-bian to the anion radical. The presence of two closely located reduction peaks in the cyclic voltammogram of complex 3 correlates with the presence of two tmp-bian ligands in its structure. For complex 1, these peaks seem to merge into one broad cathodic peak. This can be explained by the structural differences between the two complexes. In complex 1, two tmp-bian molecules are coordinated to different cadmium ions. In contrast, in structure 3, tmp-bian ligands are bound to the same Cd(II) ion.

3. Experimental

3.1. Materials and Methods

1,2-Bis[(2,4,6-trimethylphenyl)imino]acenaphthene (tmp-bian) was synthesized according to the literature [32]. The remaining reagents were purchased from commercial sources: acenaphthenequinone (99%, Sigma Aldrich, St. Louis, MI, USA), 2,4,6-trimethylaniline (99%, Sigma Aldrich), CdCl2 (98%, Sigma Aldrich), and Cd(OTf)2 (Sigma Aldrich, 98%). All solvents used were purified according to standard procedures.
Elemental analysis of C, H, N, and S was conducted using a EuroEA3000 Eurovector analyzer. The IR spectra were recorded within the 4000–300 cm−1 range with a Perkin-Elmer System 2000 FTIR spectrometer, with the samples prepared as KBr pellets. The 1H NMR spectra (500 MHz) were acquired on a Bruker Avance-500 spectrometer equipped with a 5 mm PABBO-PLUS probe at ambient temperature. The chemical shifts were expressed in parts per million (ppm) from tetramethylsilane. The cyclic voltammograms (CV) were recorded with a 797 VA Computrace system (Metrohm, Herisau, Switzerland). All measurements were conducted utilizing a conventional three-electrode configuration, comprising a glassy carbon working electrode, a platinum auxiliary electrode, and Ag/AgCl/KCl reference electrodes. The solvent used in all experiments was dichloromethane, which was deoxygenated before use. Tetra-n-butylammonium hexafluorophosphate (0.1 M solution) was used as a supporting electrolyte. The concentration of the complexes was 10−3 M. The half-wave potential (E1/2) was calculated as the mean of the cathodic and anodic peak potentials. Ferrocene was utilized as an internal standard, with the Fc/Fc+ potential measuring 0.49 V.

3.2. X-Ray Diffraction Analysis

Single-crystal X-ray diffraction data were collected on a Bruker D8 Venture diffractometer, equipped with a CMOS PHOTON III detector and an IμS 3.0 microfocus source (collimating Montel mirrors). All experiments were conducted at a temperature of 150 K using Mo Kα (λ = 0.71073 Å) radiation. Absorption correction was applied using SADABS [33]. The structures were solved by a dual space algorithm with SHELXT-2018/2 [34] and refined by full-matrix least-squares treatment against |F|2 with SHELXL-2019/3 [35] using the ShelXle GUI [36]. The atomic displacement parameters for non-hydrogen atoms were refined anisotropically. For compound 3, the presence of highly disordered solvent molecules inside the molecular structure was discovered. Due to the complexity of the modeling, the electron density of disordered solvent molecules was excluded from consideration using the PLATON/SQUEEZE program [37,38]; 52e per formula unit was found, which corresponds to 1.2 CH2Cl2 molecules per formula unit (3·1.2 CH2Cl2). Crystallographic data and details of diffraction experiments for complexes 1•2(C2H5)2O), 2 and 3•1.2CH2Cl2 are presented in Table S1, which contain(s) the supplementary crystallographic data for this paper. Deposition numbers CCDC 2435400 (for 1•2(C2H5)2O), 2435401 (for 2), and 2435402 (for 3•1.2 CH2Cl2) contain the supplementary crystallographic data for this paper. These data are provided free of charge by the joint Cambridge Crystallographic Data Centre http://www.ccdc.cam.ac.uk/structures, accessed on 31 March 2025.

3.3. Quantum Chemical Calculations

Quantum chemical calculations were performed using the ORCA 6.0.0 software package [39,40]. The crystal structure data for 13 were utilized as the initial points for the geometry optimization process. Calculations were conducted within the framework of density functional theory (DFT), employing the non-empirical PBE functional [41] in conjunction with empirical corrections for dispersion interactions, D4 [42]. All calculations used the Karlsruhe family of Gaussian-type orbital basis sets [43]. To describe Cd atoms, the def2-TZVPP basis set was used, which is a triple-zeta basis set augmented with two sets of polarization functions. The effective nuclear potential (Def2-ECP) was used to describe the inner electrons (up to 3d) [44]. To describe H atoms, the double-zeta basis set was used, augmented with one set of polarization functions. To describe all other atoms, the def2-TZVP basis sets were used. All calculations were performed using the RI approximation (RI-J) [45] with an auxiliary def2/J basis set [46].

3.4. Synthesis of [Cd2(tmp-bian)2Cl2(μ-Cl)2] (1)

CdCl2 (50 mg, 0.272 mmol) and tmp-bian (111 mg, 0.272 mmol) were dissolved in 10 mL of ethanol. The resulting solution was refluxed for 24 h. An orange precipitate was filtered using a glass filter and washed with n-hexane. Single crystals suitable for X-ray diffraction analysis were obtained by slow diffusion of diethyl ether into a solution of 1 in dichloromethane. Yield: 145 mg (90%).
Calculated for C60H56N4Cl4Cd2: С 60.8; H 3.40; N 4.7%. Found: C 61.1, H 3.63, N 4.8%.
IR (KBr, cm−1): 2966 (w), 2913 (w), 2850 (w), 1667 (m), 1634 (s), 1587 (s), 1478 (s), 1433 (m), 1420 (m), 1379 (w), 1289 (m), 1240 (s), 1225 (m), 1206 (w), 1159 (w), 1146 (w), 1115 (w), 1053 (w), 1032 (w), 961 (w), 937 (w), 893 (w), 853 (m), 847 (m), 831 (m), 806 (w), 795 (w), 779 (s), 656 (w), 574 (w), 561 (w), 554 (w), 517 (w), 449 (w).
1H NMR (CDCl3, δ ppm): 2.25 (s, 12H), 2.42 (s, 6H), 6.85 (d, 2H), 7.05 (s, 4H), 7.55 (t, 2H), 8.10 (d, 2H).

3.5. Synthesis of [Cd(tmp-bian)2(OTf)2] (3)

Cd(OTf)2 (50 mg, 0.12 mmol) and tmp-bian (99 mg, 0.024 mmol) were dissolved in 5 mL of CH2Cl2. The resulting solution was stirred at room temperature for 24 h, then evaporated under vacuum. The solid residue was washed with cold ethanol and n-hexane, and dried on a glass filter. Single crystals suitable for X-ray diffraction analysis were obtained by slow diffusion of diethyl ether into a solution of 3 in dichloromethane. Yield: 136 mg (90%).
Calculated for C62H56F6N4O6Cd: С 59.8; H 4.5; N 4.50%. Found: C 60.0, H 4.67, N 4.7%.
IR (KBr, cm−1): 2991 (w), 2914 (w), 2858 (w), 1658 (m), 1624 (m), 1585 (m), 1477 (m), 1437 (w), 1421 (w), 1385 (w), 1300 (s), 1282 (m), 1238 (s), 1221 (m), 1163 (s), 1107 (m), 1028 (s), 852 (w), 837 (w), 783 (m), 637 (s), 578 (w), 559 (w), 518 (w), 511 (w).
1H NMR (CDCl3, δ ppm): 2.09 (s, 24H), 2.12 (s, 12H), 6.91 (d, 4H), 7.09 (s, 8H), 7.59 (t, 4H), 8.14 (d, 4H).

4. Conclusions

Despite numerous examples of zinc complexes bearing bis(imino)acenaphthenes (Ar-bians), cadmium analogues were unknown until this work. Herein, we describe the first representatives of a new family of cadmium complexes with Ar-bian. These are monomeric and dimeric Cd(II) complexes based on bis-(2,4,6-trimethylphenylimino)acenaphthene (tmp-bian), for which crystal structures were determined by X-ray diffraction analysis. The coordination number of the Cd(II) ion in these complexes varies from five (square-pyramidal) to six (octahedral). The complexes exhibit ligand-centered reduction, typical of metal complexes with redox-active Ar-bian ligands.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics13050145/s1. Table S1: SCXRD Experimental details; Table S2: Optimized geometry coordinates for 1; Table S3: Optimized geometry coordinates for 2; Table S4: Optimized geometry coordinates for 3; Figure S1: IR spectrum of 1; Figure S2: IR spectrum of 3; Figure S3: 1H NMR spectrum of 1; Figure S4: 1H NMR spectrum of 3. Electronic supplementary information (ESI) available: Crystal data of complexes 1–3, 1H NMR and IR spectra of 1 and 3, optimized geometry coordinates of 1–3, CCDC no. 2435400 (for 1•2(C2H5)2O), 2435401 (for 2), and 2435402 (for 3•1.2 CH2Cl2).

Author Contributions

E.V.C., A.K.G., and N.O.S.—Investigation (synthesis, spectroscopy); I.V.B.—Investigation (SCXRD); V.I.K.—Investigation (DFT); S.A.—Funding acquisition, Resources, Visualization; B.D.—Investigation (electrochemistry); K.D.—Writing—review and editing, Funding acquisition; N.A.—Writing—review and editing, Conceptualization, Supervision, Funding acquisition; N.F.R.—Methodology, Project administration, Writing—original draft preparation; A.L.G.—Supervision, Writing—review and editing, Conceptualization. All authors have read and agreed to the published version of the manuscript.

Funding

The work was financed by the non-profit joint-stock company Korkyt Ata Kyzylorda University. The authors thank the Ministry of Science and Higher Education of the Russian Federation and the Centre of Collective Usage of NIIC SB RAS.

Data Availability Statement

The data supporting this article have been included as part of the Supplementary Materials.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fomenko, I.S.; Romashev, N.F.; Gushchin, A.L. Advances in the chemistry of redox-active bis(imino)acenaphthenes (BIAN): A case of transition metal complexes. Coord. Chem. Rev. 2024, 514, 215845. [Google Scholar] [CrossRef]
  2. Hill, N.J.; Vargas-Baca, I.; Cowley, A.H. Recent developments in the coordination chemistry of bis(imino)acenaphthene (BIAN) ligands with s- and p-block elements. J. Chem. Soc. Dalt. Trans. 2009, 240–253. [Google Scholar] [CrossRef]
  3. Bernauer, J.; Pölker, J.; Jacobi von Wangelin, A. Redox-active BIAN-based Diimine Ligands in Metal-Catalyzed Small Molecule Syntheses. ChemCatChem 2022, 14, e202101182. [Google Scholar] [CrossRef]
  4. Abakumov, G.A.; Piskunov, A.V.; Cherkasov, V.K.; Fedushkin, I.L.; Ananikov, V.P.; Eremin, D.B.; Gordeev, E.G.; Beletskaya, I.P.; Averin, A.D.; Bochkarev, M.N.; et al. Organoelement chemistry: Promising growth areas and challenges. Russ. Chem. Rev. 2018, 87, 393–507. [Google Scholar] [CrossRef]
  5. Yambulatov, D.S.; Nikolaevskii, S.A.; Kiskin, M.A.; Kholin, K.V.; Khrizanforov, M.N.; Budnikova, Y.G.; Babeshkin, K.A.; Efimov, N.N.; Goloveshkin, A.S.; Imshennik, V.K.; et al. Generation of a hetero spin complex from iron(Ii) iodide with redox active acenaphthene-1,2-diimine. Molecules 2021, 26, 2998. [Google Scholar] [CrossRef] [PubMed]
  6. Komlyagina, V.I.; Romashev, N.F.; Kokovkin, V.V.; Gushchin, A.L.; Benassi, E.; Sokolov, M.N.; Abramov, P.A. Trapping of Ag+ into a Perfect Six-Coordinated Environment: Structural Analysis, Quantum Chemical Calculations and Electrochemistry. Molecules 2022, 27, 6961. [Google Scholar] [CrossRef]
  7. Romashev, N.F.; Mirzaeva, I.V.; Bakaev, I.V.; Komlyagina, V.I.; Komarov, V.Y.; Fomenko, I.S.; Gushchin, A.L. Structure of a Binuclear Rhodium(I) Complex With the Acenaphthene- 1,2-Diimine Ligand. J. Struct. Chem. 2022, 63, 242–251. [Google Scholar] [CrossRef]
  8. Komlyagina, V.I.; Bakaev, I.V.; Romashev, N.F.; Gushchin, A.L. Unusual Coordination Type of the Halogenated Bian Ligands in Silver(I) Complexes. Chem. Asian J. 2025, 20, e202401455. [Google Scholar] [CrossRef]
  9. Rosa, V.; Laronha, H.; Gomes, C.S.B.; Cordas, C.M.; Brinco, J.; Freitas, F.; Gomes da Silva, M.D.R.; Avilés, T. Aerobic oxidation of benzylic alcohols catalysed by new (aryl-BIAN)copper(I) complexes: Their synthesis and structural characterization. Appl. Organomet. Chem. 2023, 37, e7193. [Google Scholar] [CrossRef]
  10. Padilla-Vélez, O.; O’Connor, K.S.; Lapointe, A.M.; Macmillan, S.N.; Coates, G.W. Switchable living nickel(ii) α-diimine catalyst for ethylene polymerisation. Chem. Commun. 2019, 55, 7607–7610. [Google Scholar] [CrossRef]
  11. Xu, S.; Chen, X.; Luo, G.; Gao, W. Nickel complexes based on BIAN ligands: Transformation and catalysis on ethylene polymerization. Dalt. Trans. 2021, 50, 7356–7363. [Google Scholar] [CrossRef] [PubMed]
  12. Chen, S.; Huang, Y.; Fang, X.; Li, H.; Zhang, Z.; Hor, T.S.A.; Weng, Z. Aryl-BIAN-ligated silver(i) trifluoromethoxide complex. Dalt. Trans. 2015, 44, 19682–19686. [Google Scholar] [CrossRef]
  13. Romashev, N.F.; Bakaev, I.V.; Komlyagina, V.I.; Sokolov, M.N.; Gushchin, A.L. Synthesis and Structure of Palladacyclopentadienyl Complex with Acenaphthene-1,2-Diimine Ligand. J. Struct. Chem. 2022, 63, 1304–1312. [Google Scholar] [CrossRef]
  14. Chen, C.; Jordan, R.F. Palladium-catalyzed dimerization of vinyl ethers to acetals. J. Am. Chem. Soc. 2010, 132, 10254–10255. [Google Scholar] [CrossRef] [PubMed]
  15. Komlyagina, V.I.; Romashev, N.F.; Besprozvannykh, V.K.; Arakelyan, J.; Wu, C.; Chubarov, A.S.; Bakaev, I.V.; Soh, Y.K.; Abramov, P.A.; Cheung, K.L.; et al. Effects of Bis(imino)acenaphthene (Bian)-Derived Ligands on the Cytotoxicity, DNA Interactions, and Redox Activity of Palladium(II) Bipyridine Complexes. Inorg. Chem. 2023, 62, 11541–11553. [Google Scholar] [CrossRef]
  16. Kluwer, A.M.; Koblenz, T.S.; Jonischkeit, T.; Woelk, K.; Elsevier, C.J. Kinetic and spectroscopic studies of the [palladium(Ar-bian)]-catalyzed semi-hydrogenation of 4-octyne. J. Am. Chem. Soc. 2005, 127, 15470–15480. [Google Scholar] [CrossRef]
  17. Chacon-Teran, M.A.; Findlater, M. Redox-Active BIAN-Based Iron Complexes in Catalysis. Eur. J. Inorg. Chem. 2022, 2022, e202200363. [Google Scholar] [CrossRef]
  18. Moore, J.A.; Vasudevan, K.; Hill, N.J.; Reeske, G.; Cowley, A.H. Facile routes to Alkyl-BIAN ligands. Chem. Commun. 2006, 2913–2915. [Google Scholar] [CrossRef]
  19. Ferreira, L.C.; Filgueiras, C.A.L.; Visentin, L.C.; Bordinhão, J.; Hörner, M. One-pot preparation, spectroscopic and structural characterization of mercury(II) complexes of bulky diimines with halides and pseudohalides. Zeitschrift Anorg. Allg. Chem. 2008, 634, 1896–1900. [Google Scholar] [CrossRef]
  20. El-Ayaan, U. Synthesis and structural studies of group 2B transition metal complexes with the bulky nitrogen ligand bis[N-(2,6-diisopropylphenyl) imino]acenaphthene. Monatshefte Chem. 2004, 135, 919–925. [Google Scholar] [CrossRef]
  21. Evans, D.A.; Lee, L.M.; Vargas-Baca, I.; Cowley, A.H. Photophysical tuning of the aggregation-induced emission of a series of para-substituted aryl bis(imino)acenaphthene zinc complexes. Dalt. Trans. 2015, 44, 11984–11996. [Google Scholar] [CrossRef]
  22. Zou, W.; Chen, C. Influence of Backbone Substituents on the Ethylene (Co)polymerization Properties of α-diimine Pd(II) and Ni(II) Catalysts. Organometallics 2016, 35, 1794–1801. [Google Scholar] [CrossRef]
  23. Mundil, R.; Sokolohorskyj, A.; Hošek, J.; Cvačka, J.; Císařová, I.; Kvíčala, J.; Merna, J. Nickel and palladium complexes with fluorinated alkyl substituted α-diimine ligands for living/controlled olefin polymerization. Polym. Chem. 2018, 9, 1234–1248. [Google Scholar] [CrossRef]
  24. Vignesh, A.; Zhang, Q.; Ma, Y.; Liang, T.; Sun, W.H. Attaining highly branched polyethylene elastomers by employing modified α-diiminonickel(II) catalysts: Probing the effects of enhancing fluorine atom on the ligand framework towards mechanical properties of polyethylene. Polymer 2020, 187, 122089. [Google Scholar] [CrossRef]
  25. Liu, H.R.; Gomes, P.T.; Costa, S.I.; Duarte, M.T.; Branquinho, R.; Fernandes, A.C.; Chien, J.C.W.; Singh, R.P.; Marques, M.M. Highly active new α-diimine nickel catalyst for the polymerization of α-olefins. J. Organomet. Chem. 2005, 690, 1314–1323. [Google Scholar] [CrossRef]
  26. Doerr, A.M.; Curry, M.R.; Chapleski, R.C.; Burroughs, J.M.; Lander, E.K.; Roy, S.; Long, B.K. Redox Potential as a Predictor of Polyethylene Branching Using Nickel α-Diimine Catalysts. ACS Catal. 2022, 12, 73–81. [Google Scholar] [CrossRef]
  27. Du, S.; Kong, S.; Shi, Q.; Mao, J.; Guo, C.; Yi, J.; Liang, T.; Sun, W.H. Enhancing the activity and thermal stability of nickel complex precatalysts using 1-[2,6-bis(bis(4-fluorophenyl)methyl)-4-methyl phenylimino]-2-aryliminoacenaphthylene derivatives. Organometallics 2015, 34, 582–590. [Google Scholar] [CrossRef]
  28. Maldanis, R.J.; Wood, J.S.; Chandrasekaran, A.; Rausch, M.D.; Chien, J.C.W. The formation and polymerization behavior of Ni(II) α-diimine complexes using various aluminum activators. J. Organometalic Chem. 2002, 645, 158–167. [Google Scholar] [CrossRef]
  29. Khrizanforova, V.V.; Fayzullin, R.R.; Musina, E.I.; Karasik, A.A.; Budnikova, Y.H. Electrochemical and catalytic properties of nickel(II) complexes with bis(imino)acenaphthene and diazadiphosphacyclooctane ligands. Mendeleev Commun. 2020, 30, 302–304. [Google Scholar] [CrossRef]
  30. Li, L.; Lopes, P.S.; Figueira, C.A.; Gomes, C.S.B.; Duarte, M.T.; Rosa, V.; Fliedel, C.; Avilés, T.; Gomes, P.T. Cationic and neutral (Ar-BIAN)copper(I) complexes containing phosphane and arsane ancillary ligands: Synthesis, molecular structure and catalytic behaviour in cycloaddition reactions of azides and alkynes. Eur. J. Inorg. Chem. 2013, 2013, 1404–1417. [Google Scholar] [CrossRef]
  31. Laricheva, Y.A.; Guan, C.; Kuratieva, N.V.; Romashev, N.F.; Gushchin, A.L. Synthesis and Structure of Manganese Complexes with N,N′-bis[(2,4,6-trimethylphenyl)imino]acenaphthene. Russ. J. Coord. Chem. Khimiya 2024, 50, 662–672. [Google Scholar] [CrossRef]
  32. El-Ayaan, U.; Murata, F.; El-Derby, S.; Fukuda, Y. Synthesis, structural and solvent influence studies on solvatochromic mixed-ligand copper(II) complexes with the rigid nitrogen ligand: Bis[N-(2,4,6-trimethylphenyl)imino]acenaphthene. J. Mol. Struct. 2004, 692, 209–216. [Google Scholar] [CrossRef]
  33. Sheldrick, G.M. SADABS Program for Scaling and Correction of Area Detector Data; University of Göttingen: Göttingen, Germany, 1996. [Google Scholar]
  34. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Crystallogr. Sect. A Found. Crystallogr. 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
  35. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 3–8. [Google Scholar] [CrossRef]
  36. Hübschle, C.B.; Sheldrick, G.M.; Dittrich, B. ShelXle: A Qt graphical user interface for SHELXL. J. Appl. Crystallogr. 2011, 44, 1281–1284. [Google Scholar] [CrossRef] [PubMed]
  37. Spek, A.L. PLATON SQUEEZE: A tool for the calculation of the disordered solvent contribution to the calculated structure factors. Acta Crystallogr. Sect. C Struct. Chem. 2015, 71, 9–18. [Google Scholar] [CrossRef]
  38. Spek, A.L. Single-crystal structure validation with the program PLATON. J. Appl. Crystallogr. 2003, 36, 7–13. [Google Scholar] [CrossRef]
  39. Neese, F.; Wennmohs, F.; Becker, U.; Riplinger, C. The ORCA quantum chemistry program package. J. Chem. Phys. 2020, 152, 224108. [Google Scholar] [CrossRef]
  40. Neese, F. Software update: The ORCA program system—Version 5.0. Wiley Interdiscip. Rev. Comput. Mol. Sci. 2022, 12, e1606. [Google Scholar] [CrossRef]
  41. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865–3868. [Google Scholar] [CrossRef]
  42. Caldeweyher, E.; Ehlert, S.; Hansen, A.; Neugebauer, H.; Spicher, S.; Bannwarth, C.; Grimme, S. A generally applicable atomic-charge dependent London dispersion correction. J. Chem. Phys. 2019, 150, 154122. [Google Scholar] [CrossRef] [PubMed]
  43. 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. Phys. Chem. Chem. Phys. 2005, 7, 3297–3305. [Google Scholar] [CrossRef] [PubMed]
  44. Andrae, D.; Häußermann, U.; Dolg, M.; Stoll, H.; Preuß, H. Energy-adjusted ab initio pseudopotentials for the second and third row transition elements. Theor. Chim. Acta 1990, 77, 123–141. [Google Scholar] [CrossRef]
  45. Neese, F. An Improvement of the Resolution of the Identity Approximation for the Formation of the Coulomb Matrix. J. Comput. Chem. 2003, 24, 1740–1747. [Google Scholar] [CrossRef]
  46. Weigend, F. Accurate Coulomb-fitting basis sets for H to Rn. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065. [Google Scholar] [CrossRef]
Figure 1. Synthetic routes for complexes 13.
Figure 1. Synthetic routes for complexes 13.
Inorganics 13 00145 g001
Figure 2. Molecular structures of complex 1•2(C2H5)2O, the cation of 2, and complex 3•1.2CH2Cl2. Hydrogen atoms and solvent molecules are omitted for clarity.
Figure 2. Molecular structures of complex 1•2(C2H5)2O, the cation of 2, and complex 3•1.2CH2Cl2. Hydrogen atoms and solvent molecules are omitted for clarity.
Inorganics 13 00145 g002
Figure 3. Cyclic voltammograms of complexes 1 (left) and 3 (right) in dichloromethane.
Figure 3. Cyclic voltammograms of complexes 1 (left) and 3 (right) in dichloromethane.
Inorganics 13 00145 g003
Figure 4. View of the frontier molecular orbitals of 1, the cation of 2, and 3.
Figure 4. View of the frontier molecular orbitals of 1, the cation of 2, and 3.
Inorganics 13 00145 g004
Table 1. Selected geometric parameters for 1•2(C2H5)2O, 2, and 3•1.2CH2Cl2.
Table 1. Selected geometric parameters for 1•2(C2H5)2O, 2, and 3•1.2CH2Cl2.
1•2(C2H5)2ODistance, Å2Distance, Å3•1.2CH2Cl2Distance, Å
Cd–N2.354(2)
2.427(2)
Cd–N2.375(3)
2.362(2)
Cd–N2.314(2)
2.391(3)
Cd–(μ-Cl)2.5944(8)
2.5100(7)
Cd–(μ-Cl)2.559(1)
2.6000(9)
Cd–O2.381(2)
Cd–Cl2.4211(8)Cd–O2.329(2)
2.349(3)
С–N1.275(4)
1.288(4)
С–N1.280(3)
1.269(3)
С–N1.278(4)
1.271(5)
С–С1.515(5)
С–С1.514(3)С–C1.528(5)
Table 2. Contributions of fragments (tmp-bian, Cd, Cl) to the frontier molecular orbitals of complexes 13.
Table 2. Contributions of fragments (tmp-bian, Cd, Cl) to the frontier molecular orbitals of complexes 13.
123
HOMO
tmp-bian74.8%86.7%89.4%
Cd9.2%1.0%3.0%
Cl9.9%0.60%-
LUMO
tmp-bian93.1%91.4%88.8%
Cd1.0%0.5%1.3%
Cl0.5%0.1%-
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Chekhov, E.V.; Bakaev, I.V.; Gukova, A.K.; Shaposhnikov, N.O.; Komlyagina, V.I.; Appazova, S.; Diyarova, B.; Darmagambet, K.; Appazov, N.; Romashev, N.F.; et al. Cadmium Complexes—A Novel Family in the Coordination Chemistry of 1,2-bis(arylimino)acenaphthenes. Inorganics 2025, 13, 145. https://doi.org/10.3390/inorganics13050145

AMA Style

Chekhov EV, Bakaev IV, Gukova AK, Shaposhnikov NO, Komlyagina VI, Appazova S, Diyarova B, Darmagambet K, Appazov N, Romashev NF, et al. Cadmium Complexes—A Novel Family in the Coordination Chemistry of 1,2-bis(arylimino)acenaphthenes. Inorganics. 2025; 13(5):145. https://doi.org/10.3390/inorganics13050145

Chicago/Turabian Style

Chekhov, Egor V., Ivan V. Bakaev, Alisa K. Gukova, Nikolay O. Shaposhnikov, Veronika I. Komlyagina, Saltanat Appazova, Banu Diyarova, Klara Darmagambet, Nurbol Appazov, Nikolai F. Romashev, and et al. 2025. "Cadmium Complexes—A Novel Family in the Coordination Chemistry of 1,2-bis(arylimino)acenaphthenes" Inorganics 13, no. 5: 145. https://doi.org/10.3390/inorganics13050145

APA Style

Chekhov, E. V., Bakaev, I. V., Gukova, A. K., Shaposhnikov, N. O., Komlyagina, V. I., Appazova, S., Diyarova, B., Darmagambet, K., Appazov, N., Romashev, N. F., & Gushchin, A. L. (2025). Cadmium Complexes—A Novel Family in the Coordination Chemistry of 1,2-bis(arylimino)acenaphthenes. Inorganics, 13(5), 145. https://doi.org/10.3390/inorganics13050145

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop