Next Article in Journal
Recovery of Naringin-Rich Flavonoid Extracts from Agroresidues with Anxiolytic- and Antidepressant-like Effects in Mice
Next Article in Special Issue
Metal-Free Aerobic C–N Bond Formation of Styrene and Arylamines via Photoactivated Electron Donor–Acceptor Complexation
Previous Article in Journal
Multicolor Emissive Phosphorescent Iridium(III) Complexes Containing L-Alanine Ligands: Photophysical and Electrochemical Properties, DFT Calculations, and Selective Recognition of Cu(II) Ions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Hybrid Silsesquioxane/Benzoate Cu7-Complexes: Synthesis, Unique Cage Structure, and Catalytic Activity

by
Alexey N. Bilyachenko
1,2,*,
Victor N. Khrustalev
2,3,
Evgenii I. Gutsul
1,
Anna Y. Zueva
2,
Alexander A. Korlyukov
1,4,
Lidia S. Shul’pina
1,
Nikolay S. Ikonnikov
1,
Pavel V. Dorovatovskii
5,
Dmitri Gelman
6,
Elena S. Shubina
1 and
Georgiy B. Shul’pin
7,8,*
1
A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, Vavilov Str. 28, 119991 Moscow, Russia
2
Peoples’ Friendship University of Russia (RUDN University), Miklukho-Maklay Str. 6, 117198 Moscow, Russia
3
Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences (RAS), Leninsky Prospect 47, 119991 Moscow, Russia
4
Pirogov Russian National Research Medical University, Ostrovitianov Str. 1, 117997 Moscow, Russia
5
National Research Center “Kurchatov Institute”, Akademika Kurchatova pl. 1, 123182 Moscow, Russia
6
Institute of Chemistry, Edmond J. Safra Campus, The Hebrew University of Jerusalem, Jerusalem 91904, Israel
7
Semenov Federal Research Center for Chemical Physics, Russian Academy of Sciences, Ulitsa Kosygina 4, 119991 Moscow, Russia
8
Chair of Chemistry and Physics, Plekhanov Russian University of Economics, Stremyannyi Pereulok, Dom 36, 117997 Moscow, Russia
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(23), 8505; https://doi.org/10.3390/molecules27238505
Submission received: 29 October 2022 / Revised: 24 November 2022 / Accepted: 27 November 2022 / Published: 3 December 2022
(This article belongs to the Special Issue Catalysis of Radical Reactions: Recent Progress and Emerging Fields)

Abstract

:
A series of phenylsilsesquioxane-benzoate heptacopper complexes 13 were synthesized and characterized by X-ray crystallography. Two parallel routes of toluene spontaneous oxidation (into benzyl alcohol and benzoate) assisted the formation of the cagelike structure 1. A unique multi-ligation of copper ions (from (i) silsesquioxane, (ii) benzoate, (iii) benzyl alcohol, (iv) pyridine, (v) dimethyl-formamide and (vi) water ligands) was found in 1. Directed self-assembly using benzoic acid as a reactant afforded complexes 23 with the same main structural features as for 1, namely heptanuclear core coordinated by (i) two distorted pentameric cyclic silsesquioxane and (ii) four benzoate ligands, but featuring other solvate surroundings. Complex 3 was evaluated as a catalyst for the oxidation of alkanes to alkyl hydroperoxides and alcohols to ketones with hydrogen peroxide and tert-butyl hydroperoxide, respectively, at 50 °C in acetonitrile. The maximum yield of cyclohexane oxidation products as high as 32% was attained. The oxidation reaction results in a mixture of cyclohexyl hydroperoxide, cyclohexanol, and cyclohexanone. Upon the addition of triphenylphosphine, the cyclohexyl hydroperoxide is completely converted to cyclohexanol. The specific regio- and chemoselectivity in the oxidation of n-heptane and methylcyclohexane, respectively, indicate the involvement of of hydroxyl radicals. Complex 3 exhibits a high activity in the oxidation of alcohols.

1. Introduction

Cagelike metallasilsesquioxanes (CLMSs) are an intensely studied class of metallacomplexes, providing a kaleidoscopic variety of molecular architectures [1,2,3,4,5,6,7,8,9,10,11] as well as demonstrating significant potential for various applications. The latter includes the design of molecular magnets (including single-molecule magnets and spin glasses) [12,13,14,15,16,17,18] and objects with photophysical properties (including the first example of a CLMS-based luminescent thermometer) [19,20,21,22]. For materials science, CLMSs were successfully applied for the production of functional coordination polymers [23,24,25,26], anode materials [27,28], biocompatible agents [29], and flame retardants [30]. Another important area of CLMSs investigations concerns their catalytic properties towards numerous important industrial processes [31,32,33,34,35,36]. Among recent results, we could mention CO2 cycloaddition [37], Chan–Evans–Lam coupling [38], biomass transformations [39], and oxidative amidation [40]. Considering the significant potential of copper-based CLMSs as catalysts [41,42,43], we pursue designing new types of these compounds. Numerous results pointed out the significant influence of systems used for synthesis/crystallization on the structural features of the CLMSs being formed. We could cite effects of “nuclearity shift”, e.g., Nb2 vs. Nb4 [44], Zn7 vs. Zn8 [45], Ni5 vs. Ni6 [46], Cu4Na4 vs. Cu5 [47]. Here a special versatility of copper-CLMSs should be emphasized as the most abundant representatives of this class of metallacomplexes. The kaleidoscopic variety of Cu-CLMSs included many attractive types of molecular geometry, e.g., Cubane [48], Cooling Tower [23], Fullerene-like [49], Tower of Pisa [50], Castel del Monte [51] and others [11]. Moreover, their nuclearity spanned from Cu2 [23,52] to Cu24 [49]. Our most recent results on the Cu-CLMS synthesis, structure determination, and catalytic evaluation are presented below.

2. Experimental

2.1. General Experimental Considerations

All reagents were purchased from the usual suppliers (Sigma, Fluka, St. Louis, MO, USA) and used without further purification. Elemental analyses were carried out with an XRF spectrometer VRA-30. IR spectra of the compounds (KBr pellets) were measured on a Shimadzu IR Prestige 21 FT-IR Spectrophotometer equipped with an MCT detector using a Miracle single reflection ATR unit by Pike. Signals observed (Figures S2–S4): 1600–1400 cm−1 (νC=C, νC=N), 1120 cm−1 (νPh–Si), 940–1100 cm−1 (νasSi–O, νasSi–O–Si), 900 cm−1 (νasSi–O in Si–O–M fragment), 720–680 cm−1 (σC–H of mono-substituted phenyl group). The rest of the signals are attributed to the presence of additional solvating ligands (pyridine/DMF in 1, DMSO in 2). UV-Vis spectra (Figures S5–S7, 10 mm optical path length) were recorded on a Cary 50 spectrophotometer. The d–d transitions are better resolved for 1 (Py solution) than for 2 and 3 (benzonitrile and acetone solutions, respectively) due to higher solubility in pyridine.

2.1.1. Synthesis of 1

1.00 g (5.05 mmol) of PhSi(OMe)3 and 0.28 g (7 mmol) of NaOH were heated at reflux in 35 mL of ethanol for 2 h. Then, 0.47 g (3.50 mmol) of CuCl2 was added, and the resulting mixture was stirred without heating for 24 h. Solution was separated from insoluble part by centrifugation and dried in vacuum. Resulting solid product was mixed with 45 mL of toluene and was heated at reflux for 28 h. Solution was left in contact with air for one week for crystallization, which was, however, unsuccessful. Then the solution was mixed with 15 mL of pyridine and dimethylformamide (1:1, v/v). Crystallization gave a crystalline material in 4–5 days, including single crystals that were used for X-ray diffraction analysis. The remaining part of the crystalline material was dried in vacuum to calculate yield.
Complex 1. Anal. Calcd for (Ph5Si5O10)2Cu7(PhCOO)4: Cu, 19.33; Si, 12.20.
Found: Cu, 18.47; Si, 12.01. Yield: 0.11 g (10%).

2.1.2. Syntheses of 23

In a typical procedure, 1.00 g (5.05 mmol) of PhSi (OMe)3 and 0.28 g (7 mmol) of NaOH were heated at reflux in 30 mL of ethanol for 2 h. Then, 0.47 g (3.50 mmol) of CuCl2 was added, and the resulting mixture was kept at reflux for 12 h, cooled to room temperature, and mixed with 0.24 g (2 mmol) of benzoic acid. Solution was stirred without heating for 6 h, followed by centrifugation of precipitate. Filtrate of 2 was mixed with 15 mL of dimethylsulfoxide. Crystallization gave (in ~10 days for 2 and in 3–4 days for 3) a crystalline material, including single crystals that were used for X-ray diffraction analysis. The remaining part of the crystalline material was dried in a vacuum to calculate yield.
Complex 2. Anal. Calcd for (Ph5Si5O10)2Cu7(PhCOO)4: Cu, 19.33; Si, 12.20.
Found: Cu, 19.02; Si, 12.06. Yield: 0.41 g (36%).
Complex 3. Anal. Calcd for (Ph5Si5O10)2Cu7(PhCOO)4: Cu, 19.33; Si, 12.20.
Found: Cu, 19.20; Si, 12.09. Yield: 0.62 g (54%).

2.2. X-ray Crystal Structure Determination

The single-crystal X-ray diffraction data for 1 were collected on a three-circle Bruker APEX-II CCD diffractometer (T = 100 K, λ(MoKα)-radiation, graphite monochromator, ω- and φ-scanning mode). The data were indexed and integrated using the SAINT program [53] and then scaled and corrected for absorption using the SADABS program [54]. The single-crystal X-ray diffraction data for 2 were collected on a four-circle Rigaku Synergy S diffractometer equipped with a HyPix6000HE area-detector (T = 100 K, λ(CuKα)-radiation, graphite monochromator, shutterless ω-scanning mode). The data were integrated and corrected for absorption by the CrysAlisPro program [55]. The single-crystal X-ray diffraction study of 3 was carried out on the ‘RSA’ beamline (T = 100 K, λ = 0.74500 Å) of the Kurchatov Synchrotron Radiation Source. In total, 720 frames were collected with an oscillation range of 1.0° in the φ scanning mode using two different orientations of the crystal. The semi-empirical correction for absorption was applied using the Scala program [56]. The data were indexed and integrated using the utility iMOSFLM from the CCP4 software suite [57,58]. For details, see Table S1.
The structures were solved by intrinsic phasing modification of direct methods [59] and refined by a full-matrix least-squares technique on F2 with anisotropic displacement parameters for all non-hydrogen atoms. One out of the three coordinated DMF ligands in 1 is disordered over two sites with the occupancies of 0.70:0.30. Two of the fourteen phenyl groups and all six coordinated ethanol ligands in 3 are disordered over two sites each with the occupancies of 0.60:0.40, 0.50:0.50, 0.60:0.40, 0.65:0.35, 0.50:0.50, 0.65:0.35, 0.55:0.45, and 0.65:0.35, respectively. The hydrogen atoms of the hydroxyl groups in 13, as well as the coordinated water molecules in 1, were objectively localized in the difference-Fourier maps and refined isotropically. The other hydrogen atoms were placed in calculated positions and refined within the riding model with fixed isotropic displacement parameters [Uiso(H) = 1.5Ueq(C) for the CH3 groups and 1.2Ueq(C) for the other groups]. All calculations were carried out using the SHELXTL program [60,61].
Crystallographic data for 13 have been deposited with the Cambridge Crystallographic Data Center, CCDC 2216232-2216234, respectively. Copies of this information may be obtained free of charge from the Director, CCDC, 12 Union Road, Cambridge CB2 1EZ, UK (Fax: +44-1223-336033; e-mail: [email protected] or www.ccdc.cam.ac.uk).

3. Results and Discussion

3.1. Synthesis and Structure

For the synthesis, a convenient scheme (Scheme 1) that included alkaline hydrolysis of RSi(OMe)3 [62,63] and subsequent reaction of the intermediate siloxanolate [PhSi(O)ONa]x building blocks with copper(II) chloride in ethanol solution was applied. Toluene was chosen as an additional system component for the synthesis/crystallization. Noteworthy, despite several toluene-assisted syntheses of CLMSs, reported earlier [5,64,65], CLMS structures, including toluene as solvating ligand, remain quite rare [66,67,68]. Crystallization of potential CLMS single crystals in the native system failed; thus, an additional portion of solvents with a high coordination ability (pyridine and dimethylformamide) was used (Scheme 1). This method allowed us to successfully isolate (in a 10% yield) a product 1 of [(Ph5Si5O10)2][PhCOO]4Cu7(PhCH2OH)(Py)(DMF)3(H2O)2 · DMF composition (Figure 1).
Complex 1 exhibits a set of unprecedented structural features. First of all, to the best of our knowledge, 1 is only the second observation of heptanuclear CLMS [69]. Second, the self-assembly of 1 is assisted by two distinct routes of spontaneous oxidation of toluene—(i) to benzyl alcohol and (ii) to benzoate. Both benzyl alcohol and benzoate favor the formation of 1 by playing the role of additional ligands. Several examples of spontaneous alcohol oxidations, giving unexpected carboxylic components of cage metallasilsesquioxanes, were reported for Cu-CLMSs [43,51,70]. To the best of our knowledge, the only example of methyl group hydroxylation-assisted CLMS self-assembly was reported by us for the oxidation of neocuproine into the corresponding diol [71]. Thus, the oxidation of toluene to benzyl alcohol is quite an unusual observation for CLMS chemistry. In sum, both toluene oxygenates (benzoates and benzyl alcohol) coordinate copper ions in 1. To the best of our knowledge, such a combination of oxidized toluene derivatives has never been found for any metallacomplexes to date. More than that, six different ligand types (silsesquioxane, benzoate, benzyl alcohol, pyridine, dimethylformamide, and water) in 1 is a record high multiligation of metallacomplex ever observed. Such a diversity of ligands provoked two additional features of the cage structure of 1. First of all, the location of copper ions is unusual for CLMSs. It represents (i) three copper ions in an almost linear (178.04(2)º) fragment and (ii) four copper ions in a circular arc fragment (Figure 2 top). Second, pentameric ‘heart-like’ (Ph5Si5O10) silsesquioxane ligands in 1 (Figure 2 bottom) are significantly distorted as compared to earlier reported circle-like geometry [16,18,46,72]. We suppose that an equilibrium mixture of various cyclic silanolate (x = 3, 4, 6, 8…) is initially generated in the hydrolysis of [PhSi(O)ONa]x as intermediates, whereas the pentameric form of complex 1 is organized via their concerted coordination to copper centers.
On the other hand, the multiplicity of the ligand set (as well as the necessity of preliminary toluene oxidations) prevents a high yield of complex 1. To overcome this issue, we performed additional reaction runs with a precise stoichiometric ratio of key components according to the as-received composition of 1 (10Si:7Cu:4benzoates). Using benzoic acid as a reagent (Scheme 1), two analogs of complex 1 were successfully prepared. More precisely, complex 2 (Figure 3) of [(Ph5Si5O10)2][PhCOO]4Cu7(DMSO)2(EtOH)4 composition was isolated in a 36% yield from ethanol/dimethylsulfoxide solvate system. Furthermore, complex 3 (Figure S1) of [(Ph5Si5O10)2][PhCOO]4Cu7(EtOH)6 composition was obtained in a 54% yield from an ethanol-only solvate system.
Surprisingly, despite the different ligands coordinated to the copper atoms, compounds 2 and 3 are isostructural. A change in solvating surroundings in compounds 23 compared to complex 1 does not influence the general type of cage fragment based on the heptanuclear core bound by two distorted pentameric cyclic silsesquioxane and four benzoate ligands. Nevertheless, some structural deviations could be mentioned. First, due to steric reasons, the three linearly disposed copper atoms are capable of coordinating only two bulk ligands. Hence, in molecules 2 and 3, these copper atoms coordinate the dimethylsulfoxide and ethanol and two ethanol solvent molecules, respectively. It is important to point out the fact that, both in 2 and 3, the tetraccoordinated “naked” copper atom forms the intramolecular non-covalent Cu⋅⋅⋅H interaction with the hydrogen atom of the methyl group of the neighboring coordinated ethanol molecule at the distance of 2.987 and 3.044/3.090 Å, respectively. Second, the distance between opposite carboxylate carbon atoms in compound 1 is equal to 10.291(6) Å and 10.313(5) Å, in compound 2—10.1291(14) Å and 10.3211(13) Å, in compound 3—10.1292(19) Å and 10.3447(15) Å.
In the context of functional properties, a representative compound 3 (available in a higher yield) has been tested as a homogeneous catalyst toward the oxidative functionalization of hydrocarbons (cyclohexane, n-heptane, methylcyclohexane) and alcohols.

3.2. Complex 3 as a Catalyst in Oxidations with Peroxides

In recent years, the search for fundamentally new ways of involving hydrocarbons in chemical transformations has become a topic of great importance. There is growing interest in new catalytic systems capable of efficiently and selectively functionalizing alkanes under mild conditions. Certain metal complexes are known to catalyze reactions of hydrocarbons and alcohols occurring with the functionalization of C-H bonds [73,74,75,76,77,78,79]. Hydrogen peroxide and organic hydroperoxides are often used as oxidants in these processes. Copper complexes have systematically demonstrated high activity in these oxidation processes [80,81,82,83,84,85]. Recently, we also have found several copper complexes exhibited high efficiency in the oxidation of organic compounds with peroxides [86,87,88,89,90,91,92].

3.3. Oxidation of Alkanes

We discovered that complex 3 exhibits high catalytic activity in the oxidation of saturated hydrocarbons. The data obtained in the oxidation of cyclohexane with hydrogen peroxide for complex 3 showed that in the presence of HNO3, the reactions proceed much faster, which is consistent with our previous studies of oxidative processes using copper complexes as catalysts [68,69,70,86,88,89,90,91]. Nitric acid is a co-catalyst in these reactions. If the reaction is carried out in the absence of nitric acid, only the decomposition of hydrogen peroxide (catalase activity) is observed, and the yield of the reaction products was about 0.003 M (0.6%)
Oxidation of cyclohexane with hydrogen peroxide according to gas chromatography results in the formation of a mixture of cyclohexanol and cyclohexanone. Accumulations of the products are presented in Figure 4A,B. The ratio of cyclohexanone to cyclohexanol for complex 3 is about 1 (0.017 M/0.016 M after 60 min.). Adding triphenylphosphine to the reaction mixture according to the method proposed previously by one of us [93,94] gives rise to a sharp increase in the concentration of cyclohexanol and a noticeable decrease in the amount of cyclohexanone, as shown in Figure 4B. (The ratio of cyclohexanone to cyclohexanol after adding triphenylphosphine is about 0.04). Since the main product of the oxidation reaction is alkyl hydroperoxide, this result can be explained by the fact that in the absence of triphenylphosphine, the alkyl hydroperoxide present in the reaction mixture decomposes to form ketone and alcohol in comparable amounts both in the reaction solution and in the GC injector. Triphenylphosphine reduces an alkyl hydroperoxide to corresponding alcohol; therefore, we see a sharp increase in the amount of alcohol and a decrease in the amount of ketone on the chromatogram. The total yield of alcohol and ketone in the oxidation of cyclohexane catalyzed by complex 3 after 2 h is 32%, and TON is 290.
Dependences of initial rates of cyclohexane oxidation with hydrogen peroxide on initial concentrations of cyclohexane and catalyst are presented in Figure 5 and Figure 6, respectively. To assess the reactivity of a species arising from the decomposition of hydrogen peroxide under the action of complex 3 and inducing alkane oxidation, the influence of the alkane concentration on the initial rate of its oxidation was studied (see Figure 5). The form of the dependence is consistent with the assumption of the competitive interaction of the solvent acetonitrile and alkane (RH) with the intermediate species of an oxidizing nature. As the alkane concentration increases, the rate of its oxidation reaches its maximum limiting value.
The oxidation of n-heptane with hydrogen peroxide catalyzed by complex 3 was investigated. The results are given below (initial concentrations: n-heptane (0.5 M; 0.12 mL; H2O2 1.5 M, 0.32 mL 50% aqueous; catalyst 5 × 10−4 M, 3 mg; HNO3 0.05 M, 0.05 mL stock solution; CH3CN up to 2.5 mL). After two hours, the following concentrations of (M) alcohol isomers were obtained (after reduction with triphenylphosphine, the yield of oxidation products—17%; TON = 170):
C (1) 0.007M; C (2) 0.026M; C (3) 0.026M: C (4) 0.012M.
These values give a series of selectivity: C (1):C (2):C (3):C (4) = 1.0:5.4:5.4:4.8.
The selectivity parameters for the oxidation of methylcyclohexane were also obtained: 1°:2°:3° = 1.0:6.4:18.6 (under the same conditions, the yield of oxidation products are 18%; TOH = 180 (mainly alcohols, ketones in trace amounts). The selectivity parameters measured in the oxidation methylcyclohexane are close to the parameters typical for the reactions of alkanes with hydroxyl radicals [95,96]. However, the selectivity parameters obtained for n-heptane somewhat exceed the typical selectivity parameters for the oxidation of n-heptane with the participation of hydroxyl radicals. This suggests that other oxidizing species also take part in the process.
We have included data from other publications to show that our new Cu complex is a very efficient catalyst in the oxidation of alkanes and is on par with other copper-containing catalysts (Table 1). The difference between them is noticeable—the maximum yields of oxidation products are obtained in a shorter time since the reaction rate is higher, apparently due to the influence of ligands

3.4. Oxidation of Alcohols

The accumulation of products in the reactions of aliphatic and aromatic alcohols with tert-butyl hydroperoxide is shown in Figure 7. They give ketones as oxidation products in high yields. It should be noted that these reactions do not require the addition of nitric acid. However, such reactions with hydrogen peroxide have very low efficiency.

4. Conclusions

The synthesis of copperphenylsilsesquioxane mediated by an unusual spontaneous two-way oxidation of toluene led to an unprecedented cage structure 1, including two specific ligands. These are derivatives of toluene oxidative transformation, i.e., mild one (benzylic alcohol) and harsh one (benzoate). Directed synthesis from ethanol/dimethylsulfoxide (complex 2) or ethanol (complex 3) media using benzoic acid as reactants afforded two complexes with cage structures analogous to compound 1. Features of these cagelike compounds include unusual heptacopper nuclearity supported by the presence of two distorted (heart-like) pentameric silsesquioxane and four benzoate ligands. Such mixed ligand surrounding allowed us to observe two types of copper ions’ location: three ions as a linear fragment while the rest four in a circular arc unit. Compound 3 is a powerful catalyst for the efficient oxidation of alkanes with peroxides. Data on the selectivity of oxidation and the nature of the dependence of the initial rate of cyclohexane oxidation on the initial concentration of hydrocarbon indicate that the reaction proceeds with the participation of hydroxyl radicals and alkyl hydroperoxides are formed as the main primary products. Further investigations of the implementation of carboxylic reagents for the synthesis of metallasilsesquioxane cages in the design of catalytically active compounds are underway in our groups.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27238505/s1. Details of syntheses, catalyzed reactions, and X-ray diffraction studies (CCDC 2216232-2216234). Figure S1: Side view of molecular structure 3. Figures S2–S4: IR spectra of 13. Figures S5–S7: UVvis spectra of 13. Table S1: Crystal data and structure refinement for 13

Author Contributions

Conceptualization, A.N.B. and G.B.S.; investigation, V.N.K., A.Y.Z., E.I.G., P.V.D., A.A.K., L.S.S., E.S.S., N.S.I. and D.G.; data curation, A.N.B., L.S.S. and G.B.S.; writing—review and editing, A.N.B., D.G. and G.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation, grant number 22-13-00250. This work was performed within the framework of the Program of Fundamental Research of the Russian Federation. Reg. No. 122040500068-0. Elemental and GC analyses were performed with financial support from the Ministry of Education and Science of the Russian Federation using equipment from the Center for molecular composition studies of INEOS RAS. D.G. (oxidation of alcohols) acknowledges the ISF (Israel Science Foundation) Grant No. 370/20, and Esther K. and M. Mark Watkins, Chair for Synthetic Organic Chemistry.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not available.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Murugavel, R.; Voigt, A.; Walawalkar, M.G.; Roesky, H.W. Hetero- and Metallasiloxanes Derived from Silanediols, Disilanols, Silanetriols, and Trisilanols. Chem. Rev. 1996, 96, 2205–2223. [Google Scholar] [CrossRef] [PubMed]
  2. Lorenz, V.; Fischer, A.; Gießmann, S.; Gilje, J.W.; Gun’ko, Y.; Jacob, K.; Edelmann, F.T. Disiloxanediolates and polyhedral metallasilsesquioxanes of the early transition metals and f-elements. Coord. Chem. Rev. 2000, 206–207, 321–368. [Google Scholar] [CrossRef]
  3. Hanssen, R.W.J.M.; van Santen, R.A.; Abbenhuis, H.C.L. The Dynamic Status Quo of Polyhedral Silsesquioxane Coordination Chemistry. Eur. J. Inorg. Chem. 2004, 2004, 675–683. [Google Scholar] [CrossRef]
  4. Roesky, H.W.; Anantharaman, G.; Chandrasekhar, V.; Jancik, V.; Singh, S. Control of Molecular Topology and Metal Nuclearity in Multimetallic Assemblies: Designer Metallosiloxanes Derived from Silanetriols. Chem.—Eur. J. 2004, 10, 4106–4114. [Google Scholar] [CrossRef] [PubMed]
  5. Lorenz, V.; Edelmann, F.T. Metallasilsesquioxanes. Adv. Organomet. Chem. 2005, 53, 101–153. [Google Scholar]
  6. Levitsky, M.M.; Zavin, B.G.; Bilyachenko, A. Chemistry of metallasiloxanes. Current trends and new concepts. Russ. Chem. Rev. 2007, 76, 847–866. [Google Scholar] [CrossRef]
  7. Edelmann, F.T. Metallasilsesquioxanes—Synthetic and Structural Studies. In Silicon Chemistry: From the Atom to Extended Systems; Jutzi, P., Schubert, U., Eds.; Wiley: Darmstadt, Germany, 2003; pp. 383–394. [Google Scholar]
  8. Levitsky, M.M.; Bilyachenko, A.N. Modern concepts and methods in the chemistry of polyhedral metallasiloxanes. Coord. Chem. Rev. 2016, 306, 235–269. [Google Scholar] [CrossRef]
  9. Zheng, Y.; Gao, Z.; Han, J. Current Chemistry of Cyclic Oligomeric Silsesquioxanes. Curr. Org. Chem. 2018, 21, 2814–2828. [Google Scholar] [CrossRef]
  10. Ouyang, J.; Haotian, S.; Liang, Y.; Commisso, A.; Li, D.; Xu, R.; Yu, D. Recent Progress in Metal-containing Silsesquioxanes: Preparation and Application. Curr. Org. Chem. 2018, 21, 2829–2848. [Google Scholar] [CrossRef]
  11. Levitsky, M.M.; Zubavichus, Y.V.; Korlyukov, A.A.; Khrustalev, V.N.; Shubina, E.S.; Bilyachenko, A.N. Silicon and Germanium-Based Sesquioxanes as Versatile Building Blocks for Cage Metallacomplexes. A Review. J. Clust. Sci. 2019, 30, 1283–1316. [Google Scholar] [CrossRef]
  12. Levitsky, M.M.; Bilyachenko, A.N.; Shubina, E.S.; Long, J.; Guari, Y.; Larionova, J. Magnetic cagelike metallasilsesquioxanes. Coord. Chem. Rev. 2019, 398, 213015. [Google Scholar] [CrossRef]
  13. Liu, Y.-N.; Hou, J.-L.; Wang, Z.; Gupta, R.K.; Jagličić, Z.; Jagodič, M.; Wang, W.-G.; Tung, C.-H.; Sun, D. An Octanuclear Cobalt Cluster Protected by Macrocyclic Ligand: In Situ Ligand-Transformation-Assisted Assembly and Single-Molecule Magnet Behavior. Inorg. Chem. 2020, 59, 5683–5693. [Google Scholar] [CrossRef] [PubMed]
  14. Kulakova, A.N.; Nigoghossian, K.; Félix, G.; Khrustalev, V.N.; Shubina, E.S.; Long, J.; Guari, Y.; Carlos, L.D.; Bilyachenko, A.N.; Larionova, J. New Magnetic and Luminescent Dy(III) and Dy(III)/Y(III) Based Tetranuclear Silsesquioxane Cages. Eur. J. Inorg. Chem. 2021, 2021, 2696–2701. [Google Scholar] [CrossRef]
  15. Sheng, K.; Wang, R.; Tang, X.; Jagodič, M.; Jagličić, Z.; Pang, L.; Dou, J.-M.; Gao, Z.-Y.; Feng, H.-Y.; Tung, C.-H.; et al. A Carbonate-Templated Decanuclear Mn Nanocage with Two Different Silsesquioxane Ligands. Inorg. Chem. 2021, 60, 14866–14871. [Google Scholar] [CrossRef]
  16. Bilyachenko, A.N.; Yalymov, A.; Dronova, M.; Korlyukov, A.A.; Vologzhanina, A.V.; Es’Kova, M.A.; Long, J.; Larionova, J.; Guari, Y.; Dorovatovskii, P.V.; et al. Family of Polynuclear Nickel Cagelike Phenylsilsesquioxanes; Features of Periodic Networks and Magnetic Properties. Inorg. Chem. 2017, 56, 12751–12763. [Google Scholar] [CrossRef] [PubMed]
  17. Sheng, K.; Wang, R.; Bilyachenko, A.; Khrustalev, V.; Jagodič, M.; Jagličić, Z.; Li, Z.; Wang, L.; Tung, C.; Sun, D. Tridecanuclear Gd(III)-silsesquioxane: Synthesis, structure, and magnetic property. ChemPhysMater 2022, 1, 247–251. [Google Scholar] [CrossRef]
  18. Bilyachenko, A.N.; Yalymov, A.I.; Levitsky, M.M.; Korlyukov, A.A.; Es’kova, M.A.; Long, J.; Larionova, J.; Guari, Y.; Shul’pina, L.S.; Ikonnikov, N.S.; et al. First cagelike pentanuclear Co(II)-silsesquioxane. Dalton Trans. 2016, 45, 13663–13666. [Google Scholar] [CrossRef] [PubMed]
  19. Marchesi, S.; Bisio, C.; Carniato, F. Synthesis of Novel Luminescent Double-Decker Silsesquioxanes Based on Partially Condensed TetraSilanolPhenyl POSS and Tb3+/Eu3+ Lanthanide Ions. Processes 2022, 10, 758. [Google Scholar] [CrossRef]
  20. Kulakova, A.N.; Bilyachenko, A.N.; Levitsky, M.M.; Khrustalev, V.N.; Shubina, E.S.; Felix, G.; Mamontova, E.; Long, J.; Guari, Y.; Larionova, J. New Luminescent Tetranuclear Lanthanide-Based Silsesquioxane Cage-Like Architectures. Chem.—Eur. J. 2020, 26, 16594–16598. [Google Scholar] [CrossRef]
  21. Sheng, K.; Liu, Y.-N.; Gupta, R.K.; Kurmoo, M.; Sun, D. Arylazopyrazole-functionalized photoswitchable octanuclear Zn(II)-silsesquioxane nanocage. Sci. China Ser. B Chem. 2020, 64, 419–425. [Google Scholar] [CrossRef]
  22. Nigoghossian, K.; Kulakova, A.N.; Félix, G.; Khrustalev, V.N.; Shubina, E.S.; Long, J.; Guari, Y.; Sene, S.; Carlos, L.D.; Bilyachenko, A.N.; et al. Temperature sensing in Tb3+/Eu3+-based tetranuclear silsesquioxane cages with tunable emission. RSC Adv. 2021, 11, 34735–34741. [Google Scholar] [CrossRef] [PubMed]
  23. Bilyachenko, A.N.; Korlyukov, A.A.; Vologzhanina, A.V.; Khrustalev, V.N.; Kulakova, A.N.; Long, J.; Larionova, J.; Guari, Y.; Dronova, M.S.; Tsareva, U.S.; et al. Tuning linkage isomerism and magnetic properties of bi- and tri-metallic cage silsesquioxanes by cation and solvent effects. Dalton Trans. 2017, 46, 12935–12949. [Google Scholar] [CrossRef] [PubMed]
  24. Kulakova, A.N.; Bilyachenko, A.N.; Korlyukov, A.A.; Shul’Pina, L.S.; Bantreil, X.; Lamaty, F.; Shubina, E.S.; Levitsky, M.M.; Ikonnikov, N.S.; Shul’Pin, G.B. A new “bicycle helmet”-like copper(ii),sodiumphenylsilsesquioxane. Synthesis, structure and catalytic activity. Dalton Trans. 2018, 47, 15666–15669. [Google Scholar] [CrossRef] [PubMed]
  25. Bilyachenko, A.N.; Astakhov, G.S.; Kulakova, A.N.; Korlyukov, A.A.; Zubavichus, Y.V.; Dorovatovskii, P.V.; Shul’Pina, L.S.; Shubina, E.S.; Ikonnikov, N.S.; Kirillova, M.V.; et al. Exploring Cagelike Silsesquioxane Building Blocks for the Design of Heterometallic Cu4/M4 Architectures. Cryst. Growth Des. 2022, 22, 2146–2157. [Google Scholar] [CrossRef]
  26. Astakhov, G.S.; Bilyachenko, A.N.; Korlyukov, A.A.; Levitsky, M.M.; Shul’pina, L.S.; Bantreil, X.; Lamaty, F.; Vologzhanina, A.V.; Shubina, E.S.; Dorovatovskii, P.V.; et al. High-Cluster Cu9 Cage Silsesquioxanes: Synthesis, Structure, and Catalytic Activity. Inorg. Chem. 2018, 57, 11524–11529. [Google Scholar] [CrossRef]
  27. Lin, X.; Dong, Y.; Chen, X.; Liu, H.; Liu, Z.; Xing, T.; Li, A.; Song, H. Metallasilsesquioxane-derived ultrathin porous carbon nanosheet 3D architectures via an “in situ dual templating” strategy for ultrafast sodium storage. J. Mater. Chem. A 2021, 9, 6423–6431. [Google Scholar] [CrossRef]
  28. Lin, X.; Dong, Y.; Liu, X.; Chen, X.; Li, A.; Song, H. In-situ pre-lithiated onion-like SiOC/C anode materials based on metallasilsesquioxanes for Li-ion batteries. Chem. Eng. J. 2021, 428, 132125. [Google Scholar] [CrossRef]
  29. Yahyaei, H.; Mohseni, M.; Ghanbari, H.; Messori, M. Synthesis and characterization of polyhedral oligomeric titanized silsesquioxane: A new biocompatible cage like molecule for biomedical application. Mater. Sci. Eng. C 2016, 61, 293–300. [Google Scholar] [CrossRef]
  30. Zeng, B.; He, K.; Wu, H.; Ye, J.; Zheng, X.; Luo, W.; Xu, Y.; Yuan, C.; Dai, L. Zirconium-Embedded Polyhedral Oligomeric Silsesquioxane Containing Phosphaphenanthrene-Substituent Group Used as Flame Retardants for Epoxy Resin Composites. Macromol. Mater. Eng. 2021, 306, 2100012. [Google Scholar] [CrossRef]
  31. Abbenhuis, H.C.L. Advances in Homogeneous and Heterogeneous Catalysis with Metal-Containing Silsesquioxanes. Chem. Eur. J. 2000, 6, 25–32. [Google Scholar] [CrossRef]
  32. Levitskii, M.M.; Smirnov, V.V.; Zavin, B.G.; Bilyachenko, A.N.; Rabkina, A.Y. Metalasiloxanes: New structure formation methods and catalytic properties. Kinet. Catal. 2009, 50, 490–507. [Google Scholar] [CrossRef]
  33. Ward, A.J.; Masters, A.F.; Maschmeyer, T. Metallasilsesquioxanes: Molecular Analogues of Heterogeneous Catalysts. In Applications of Polyhedral Oligomeric Silsesquioxanes; Springer: New York, NY, USA, 2011; Volume 3, pp. 135–166. [Google Scholar]
  34. Levitsky, M.M.; Yalymov, A.I.; Kulakova, A.N.; Petrov, A.A.; Bilyachenko, A.N. Cage-like metallasilsesquioxanes in catalysis: A review. J. Mol. Catal. Chem. 2017, 426, 297–304. [Google Scholar] [CrossRef]
  35. Quadrelli, E.A.; Basset, J.M. On silsesquioxanes’ accuracy as molecular models for silica-grafted complexes in heterogeneous catalysis. Coord. Chem. Rev. 2010, 254, 707–728. [Google Scholar] [CrossRef]
  36. Levitsky, M.M.; Bilyachenko, A.N.; Shul’Pin, G.B. Oxidation of C-H compounds with peroxides catalyzed by polynuclear transition metal complexes in Si- or Ge-sesquioxane frameworks: A review. J. Organomet. Chem. 2017, 849–850, 201–218. [Google Scholar] [CrossRef]
  37. Sheng, K.; Si, W.-D.; Wang, R.; Wang, W.-Z.; Dou, J.; Gao, Z.-Y.; Wang, L.-K.; Tung, C.-H.; Sun, D. Keggin-Type Tridecanuclear Europium-Oxo Nanocluster Protected by Silsesquioxanes. Chem. Mater. 2022, 34, 4186–4194. [Google Scholar] [CrossRef]
  38. Astakhov, G.; Levitsky, M.; Bantreil, X.; Lamaty, F.; Khrustalev, V.; Zubavichus, Y.; Dorovatovskii, P.; Shubina, E.; Bilyachenko, A. Cu(II)-silsesquioxanes as efficient precatalysts for Chan-Evans-Lam coupling. J. Organomet. Chem. 2019, 906, 121022. [Google Scholar] [CrossRef]
  39. Garg, S.; Unruh, D.K.; Krempner, C. Zirconium and hafnium polyhedral oligosilsesquioxane complexes–green homogeneous catalysts in the formation of bio-derived ethers via a MPV/etherification reaction cascade. Catal. Sci. Technol. 2020, 11, 211–218. [Google Scholar] [CrossRef]
  40. Astakhov, G.S.; Khrustalev, V.N.; Dronova, M.S.; Gutsul, E.I.; Korlyukov, A.A.; Gelman, D.; Zubavichus, Y.V.; Novichkov, D.A.; Trigub, A.L.; Shubina, E.S.; et al. Cage-like manganesesilsesquioxanes: Features of their synthesis, unique structure, and catalytic activity in oxidative amidations. Inorg. Chem. Front. 2022, 9, 4525–4537. [Google Scholar] [CrossRef]
  41. Bilyachenko, A.N.; Khrustalev, V.N.; Zueva, A.Y.; Titova, E.M.; Astakhov, G.S.; Zubavichus, Y.V.; Dorovatovskii, P.V.; Korlyukov, A.A.; Shul’pina, L.S.; Shubina, E.S.; et al. A Novel Family of Cage-like (CuLi, CuNa, CuK)-phenylsilsesquioxane Complexes with 8-Hydroxyquinoline Ligands: Synthesis, Structure, and Catalytic Activity. Molecules 2022, 27, 6205. [Google Scholar] [CrossRef]
  42. Bilyachenko, A.N.; Gutsul, E.I.; Khrustalev, V.N.; Astakhov, G.S.; Zueva, A.Y.; Zubavichus, Y.V.; Kirillova, M.V.; Shul’pina, L.S.; Ikonnikov, N.S.; Dorovatovskii, P.V.; et al. Acetone Factor in the Design of Cu4-, Cu6-, and Cu9-Based Cage Coppersilsesquioxanes: Synthesis, Structural Features, and Catalytic Functionalization of Alkanes. Inorg. Chem. 2022, 61, 14800–14814. [Google Scholar] [CrossRef]
  43. Astakhov, G.S.; Levitsky, M.M.; Zubavichus, Y.V.; Khrustalev, V.N.; Titov, A.A.; Dorovatovskii, P.V.; Smol’yakov, A.F.; Shubina, E.S.; Kirillova, M.V.; Kirillov, A.M.; et al. Cu6- and Cu8-Cage Sil- and Germsesquioxanes: Synthetic and Structural Features, Oxidative Rearrangements, and Catalytic Activity. Inorg. Chem. 2021, 60, 8062–8074. [Google Scholar] [CrossRef] [PubMed]
  44. Lorenz, V.; Blaurock, S.; Görls, H.; Edelmann, F.T. The First Niobasilsesquioxanes. Organometallics 2006, 25, 5922–5926. [Google Scholar] [CrossRef]
  45. Anantharaman, G.; Chandrasekhar, V.; Nehete, U.N.; Roesky, H.W.; Vidovic, D.; Magull, J. New Polyhedral Zinc Siloxanes: Synthesis and X-ray Crystal Structures of Zn8Me7(dioxane)2(O3SiR)3 and [Zn7Me2(THF)5(O3SiR)4][R = (2,6-i-Pr2C6H3)N(SiMe3)]. Organometallics 2004, 23, 2251–2256. [Google Scholar] [CrossRef]
  46. Bilyachenko, A.N.; Yalymov, A.I.; Korlyukov, A.A.; Long, J.; Larionova, J.; Guari, Y.; Vologzhanina, A.V.; Eskova, M.A.; Shubina, E.S.; Levitsky, M.M. Unusual penta- and hexanuclear Ni(II)-based silsesquioxane polynuclear complexes. Dalton Trans. 2016, 45, 7320–7327. [Google Scholar] [CrossRef]
  47. Astakhov, G.S.; Levitsky, M.M.; Korlyukov, A.A.; Shul’Pina, L.S.; Shubina, E.S.; Ikonnikov, N.S.; Vologzhanina, A.V.; Bilyachenko, A.N.; Dorovatovskii, P.V.; Kozlov, Y.N.; et al. New Cu4Na4- and Cu5-Based Phenylsilsesquioxanes. Synthesis via Complexation with 1,10-Phenanthroline, Structures and High Catalytic Activity in Alkane Oxidations with Peroxides in Acetonitrile. Catalysts 2019, 9, 701. [Google Scholar] [CrossRef] [Green Version]
  48. Edelmann, F.T.; Gießmann, S.; Fischer, A. Silsesquioxane Chemistry, 5. Retention of the Cu O core upon formation of the first copper(I) silsesquioxane from tetrameric copper(I) -t-butoxide. Inorg. Chem. Commun. 2000, 3, 658–661. [Google Scholar] [CrossRef]
  49. Tan, G.; Yang, Y.; Chu, C.; Zhu, H.; Roesky, H.W. Cu24O24Si8R8: Organic Soluble 56-Membered Copper(I) Siloxane Cage and Its Use in Homogeneous Catalysis. J. Am. Chem. Soc. 2010, 132, 12231–12233. [Google Scholar] [CrossRef]
  50. Astakhov, G.S.; Bilyachenko, A.N.; Levitsky, M.M.; Korlyukov, A.A.; Zubavichus, Y.V.; Dorovatovskii, P.V.; Khrustalev, V.N.; Vologzhanina, A.V.; Shubina, E.S. Tridecanuclear CuII11Na2 Cagelike Silsesquioxanes. Cryst. Growth Des. 2018, 18, 5377–5384. [Google Scholar] [CrossRef]
  51. Bilyachenko, A.N.; Khrustalev, V.N.; Zubavichus, Y.V.; Vologzhanina, A.V.; Astakhov, G.S.; Gutsul, E.I.; Shubina, E.S.; Levitsky, M.M. High-nuclearity (Cu8-based) cage silsesquioxanes: Synthesis and structural study. Cryst Growth Des. 2018, 18, 2452–2457. [Google Scholar] [CrossRef]
  52. Schax, F.; Braun, B.; Limberg, C. A Tripodal Trisilanol Ligand and Its Complexation Behavior towards CuI, CuII, and ZnII. Eur. J. Inorg. Chem. 2014, 2014, 2124–2130. [Google Scholar] [CrossRef]
  53. Bruker. SAINT, v. 8.34A; Bruker AXS Inc.: Madison, WI, USA, 2014. [Google Scholar]
  54. Rigaku. CrysAlisPro Software System, v. 1.171.41.106a; Rigaku Oxford Diffraction: Oxford, UK, 2021. [Google Scholar]
  55. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Cryst. 2015, 48, 3–10. [Google Scholar] [CrossRef] [PubMed]
  56. Evans, P. Scaling and assessment of data quality. Acta Crystallogr. 2006, D62, 72–82. [Google Scholar] [CrossRef] [PubMed]
  57. Battye, T.G.G.; Kontogiannis, L.; Johnson, O.; Powell, H.R.; Leslie, A.G.W. iMOSFLM: A new graphical interface for diffraction-image processing with MOSFLM. Acta Crystallogr. 2011, D67, 271–281. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Winn, M.D.; Ballard, C.C.; Cowtan, K.D.; Dodson, E.J.; Emsley, P.; Evans, P.R.; Keegan, R.M.; Krissinel, E.B.; Leslie, A.G.W.; McCoy, A.; et al. Overview of the CCP4 suite and current developments. Acta Crystallogr. 2011, D67, 235–242. [Google Scholar]
  59. Sheldrick, G.M. SHELXT-Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  60. Sheldrick, G.M. A short history of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  61. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Crystallogr. 2015, C71, 3–8. [Google Scholar]
  62. Prigyai, N.; Chanmungkalakul, S.; Ervithayasuporn, V.; Yodsin, N.; Jungsuttiwong, S.; Takeda, N.; Unno, M.; Boonmak, J.; Kiatkamjornwong, S. Lithium-Templated Formation of Polyhedral Oligomeric Silsesquioxanes (POSS). Inorg. Chem. 2019, 58, 15110–15117. [Google Scholar] [CrossRef]
  63. Laird, M.; Totée, C.; Gaveau, P.; Silly, G.; van der Lee, A.; Carcel, C.; Unno, M.; Bartlett, J.M.; Wong Chi Man, M. Functionalised polyhedral oligomeric silsesquioxane with encapsulated fluoride—First observation of fluxional Si⋯F interactions in POSS. Dalton Trans. 2021, 50, 81–89. [Google Scholar] [CrossRef]
  64. Ehle, S.; Lorenz, V.; Liebing, P.; Hilfert, L.; Edelmann, F.T. Synthesis and structural characterization of two complex tantalum(V) siloxides. Inorg. Chem. Commun. 2016, 74, 82–85. [Google Scholar] [CrossRef]
  65. Gießmann, S.; Lorenz, V.; Liebing, P.; Hilfert, L.; Fischer, A.; Edelmann, F.T. Synthesis and structural study of new metallasilsesquioxanes of potassium and uranium. Dalton Trans. 2017, 46, 2415–2419. [Google Scholar] [CrossRef] [PubMed]
  66. Bilyachenko, A.N.; Korlyukov, A.A.; Levitskii, M.M.; Antipin, M.Y.; Zavin, B.G. New cagelike metallasiloxane containing FeIII ions in different coordination spheres. Russ. Chem. Bull. 2007, 56, 543–545. [Google Scholar] [CrossRef]
  67. Bilyachenko, A.N.; Kulakova, A.N.; Levitsky, M.M.; Korlyukov, A.A.; Khrustalev, V.N.; Vologzhanina, A.V.; Titov, A.A.; Dorovatovskii, P.V.; Shul’pina, L.S.; Lamaty, F.; et al. Ionic Complexes of Tetra- and Nonanuclear Cage Cu(II)-phenylsilsesquioxanes: Synthesis and High Activity in Oxidative Catalysis. ChemCatChem 2017, 9, 4437–4447. [Google Scholar] [CrossRef]
  68. Kulakova, A.N.; Khrustalev, V.N.; Zubavichus, Y.V.; Shul’pina, L.S.; Shubina, E.S.; Levitsky, M.M.; Ikonnikov, N.S.; Bilyachenko, A.N.; Kozlov, Y.N.; Shul’pin, G.B. Palanquin-Like Cu4Na4 Silsesquioxane Synthesis (via Oxidation of 1,1-bis(Diphenylphosphino)methane), Structure and Catalytic Activity in Alkane or Alcohol Oxidation with Peroxides. Catalysts 2019, 9, 154. [Google Scholar] [CrossRef] [Green Version]
  69. Bilyachenko, A.N.; Levitsky, M.M.; Korlyukov, A.A.; Khrustalev, V.N.; Zubavichus, Y.V.; Shul’pina, L.S.; Shubina, E.S.; Vologzhanina, A.V.; Shul’pin, G.B. Heptanuclear Cage Cu(II)-Silsesquioxanes. Features of Synthesis, Structure and Catalytic Activity. Eur. J. Inorg. Chem. 2018, 2018, 2505–2511. [Google Scholar] [CrossRef]
  70. Bilyachenko, A.N.; Kulakova, A.N.; Levitsky, M.M.; Petrov, A.A.; Korlyukov, A.A.; Shul’pina, L.S.; Khrustalev, V.N.; Dorovatovskii, P.V.; Vologzhanina, A.V.; Tsareva, U.S.; et al. Unusual tri-, hexa- and nonanuclear Cu(II) cage methylsilsesquioxanes: Synthesis, Structures and Catalytic Activity in Oxidations with Peroxides. Inorg. Chem. 2017, 56, 4093–4103. [Google Scholar] [CrossRef]
  71. Bilyachenko, A.N.; Levitsky, M.M.; Khrustalev, V.N.; Zubavichus, Y.V.; Shul’pina, L.S.; Shubina, E.S.; Shul’pin, G.B. Mild and Regioselective Hydroxylation of Methyl Group in Neocuproine: Approach to an N,O-Ligated Cu6 Cage Phenylsilsesquioxane. Organometallics 2018, 37, 168–171. [Google Scholar] [CrossRef]
  72. Kulakova, A.N.; Sedykh, E.E.; Levitsky, M.M.; Dorovatovskii, P.V.; Khrustalev, V.N.; Shul’pina, L.S.; Shubina, E.S.; Kozlov, Y.N.; Ikonnikov, N.S.; Bilyachenko, A.N.; et al. The first tris-heteroleptic copper cage, ligated by germsesquioxanes, 2,2′-bipyridines and 3,5-dimethylpyrazolates. Synthesis, structure and unique catalytic activity in oxidation of alkanes and alcohols with peroxides. J. Organometal. Chem. 2019, 899, 120911. [Google Scholar] [CrossRef]
  73. Bryliakov, K. (Ed.) Frontiers of Green Catalytic Selective Oxidations; Springer: Singapore, 2019; Chapter 1; pp. 1–35. ISBN 978-981-32-9750-0. [Google Scholar] [CrossRef]
  74. Shul’pin, G.B. New Trends in Oxidative Functionalization of Carbon–Hydrogen Bonds: A Review. Catalysts 2016, 6, 50. [Google Scholar] [CrossRef] [Green Version]
  75. Shul’pin, G.B.; Kozlov, Y.N.; Shul’pina, L.S. Metal Complexes Containing Redox-active Ligands in Oxidation of Hydrocarbons and Alcohols: A Review. Catalysts 2019, 9, 1046. [Google Scholar] [CrossRef] [Green Version]
  76. Shul’pin, G.B.; Shul’pina, L.S. Oxidation of Organic Compounds with Peroxides Catalyzed by Polynuclear Metal Compounds, A Review. Catalysts 2021, 11, 186. [Google Scholar] [CrossRef]
  77. Wójtowicz-Mlochowska, H. Synthetic utility of metal catalyzed hydrogen peroxide oxidation of C-H, C-C and C=C bonds in alkanes, arenes and alkenes: Recent advances. Arkivoc 2017, 2017, 12–58. [Google Scholar] [CrossRef]
  78. Rachmilovich-Calis, S.; Masarwa, A.; Meyerstein, N.; Meyerstein, D.; van Eldik, R. New mechanistic aspects of the Fenton reaction. Chem.—Eur. J. 2009, 15, 8303–8309. [Google Scholar] [CrossRef] [PubMed]
  79. Zhang, Q.; Goldsmith, C.R. Kinetic Analysis of the Formation and Decay of a Non-Heme Ferric Hydroperoxide Species Susceptible to O–O Bond Homolysis. Inorg. Chem. 2014, 53, 5206–5211. [Google Scholar] [CrossRef]
  80. Sabbatini, A.; Martins, L.M.D.R.S.; Mahmudov, K.T.; Kopylovich, M.N.; Drew, M.G.B.; Pettinari, C.; Pombeiro, A.J.L. Microwave-assisted and solvent-free peroxidative oxidation of 1-phenylethanol to acetophenone with a CuII-TEMPO catalytic system. Catal. Commun. 2014, 48, 69–72. [Google Scholar] [CrossRef]
  81. Perraud, O.; Sorokin, A.B.; Dutasta, J.-P.; Martinez, A. Oxidation of cycloalkanes by H2O2 using a copper–hemicryptophane complex as a catalyst. Chem. Commun. 2013, 49, 1288–1290. [Google Scholar] [CrossRef]
  82. Ünver, H.; Kani, I. Homogeneous oxidation of alcohols in water catalyzed with Cu(II)-triphenyl acetate/bipyridyl complex. Polyhedron 2017, 134, 257–262. [Google Scholar] [CrossRef]
  83. Calero, R.; Vega, A.; María García, A.; Spodine, E.; Manzur, J. Oxidation and Catalytic Properties of a Binuclear Copper(I) Complex with a Meta-Xylyl Spacer Ligand. J. Chil. Chem. Soc. 2003, 48, 85–88. [Google Scholar] [CrossRef]
  84. Sutradhar, M.; Alegria, C.B.A.E.; Guedes da Silva, M.F.C.; Liu, C.-M.; Pombeiro, A.J.L. Peroxidative Oxidation of Alkanes and Alcohols under Mild Conditions by Di- and Tetranuclear Copper (II) Complexes of Bis (2-Hydroxybenzylidene) Isophthalohydrazide. Molecules 2018, 23, 2699. [Google Scholar] [CrossRef] [Green Version]
  85. Nandi, M.; Roy, P. Peroxidative oxidation of cycloalkane by di-, tetra- and polynuclear copper (II) complexes. Indian J. Chem. 2013, 52, 1263–1268. [Google Scholar]
  86. Choroba, K.; Machura, B.; Kula, S.; Raposo, L.R.; Fernandes, A.R.; Kruszynski, R.; Erfurt, K.; Shul’pina, L.S.; Kozlov, Y.N.; Shul’pin, G.B. Copper(II) complexes with 2,2′:6′,2″-terpyridine, 2,6-di(thiazol-2-yl)pyridine and 2,6-di(pyrazin-2-yl)pyridine substituted with quinolines. Synthesis, structure, antiproliferative activity, and catalytic activity in oxidation of alkanes and alcohols with peroxides. Dalton Trans. 2019, 48, 12656–12673. [Google Scholar] [PubMed]
  87. Shul’pina, L.S.; Vinogradov, M.M.; Kozlov, Y.N.; Nelyubina, Y.V.; Ikonnikov, N.S.; Shul’pin, G.B. Copper complexes with 1,10-phenanthrolines as efficient catalysts for oxidation of alkanes by hydrogen peroxide. Inorg. Chim. Acta 2020, 51, 2119889. [Google Scholar]
  88. Kulakova, A.N.; Bilyachenko, A.N.; Khrustalev, V.N.; Zubavichus, Y.V.; Dorovatovskii, P.V.; Shul’pina, L.S.; Bantreil, X.; Lamaty, F.; Shubina, E.S.; Levitsky, M.M.; et al. Cu42Ge24Na4—A Giant Trimetallic Sesquioxane Cage: Synthesis, Structure, and Catalytic Activity. Catalysts 2018, 8, 484. [Google Scholar] [CrossRef] [Green Version]
  89. Bilyachenko, A.N.; Khrustalev, V.N.; Zubavichus, Y.V.; Shul’Pina, L.S.; Kulakova, A.N.; Bantreil, X.; Lamaty, F.; Levitsky, M.M.; Gutsul, E.I.; Shubina, E.S.; et al. Heptanuclear Fe5Cu2-Phenylgermsesquioxane containing 2,2′-Bipyridine: Synthesis, Structure, and Catalytic Activity in Oxidation of C–H Compounds. Inorg. Chem. 2017, 57, 528–534. [Google Scholar] [CrossRef] [PubMed]
  90. Kulakova, A.N.; Korlyukov, A.A.; Zubavichus, Y.V.; Khrustalev, V.N.; Bantreil, X.; Shul’Pina, L.S.; Levitsky, M.M.; Ikonnikov, N.S.; Shubina, E.S.; Lamaty, F.; et al. Hexacoppergermsesquioxanes as complexes with N-ligands: Synthesis, structure and catalytic properties. J. Organomet. Chem. 2019, 884, 17–28. [Google Scholar] [CrossRef]
  91. Kokorin, A.I.; Golubeva, E.N.; Vinogradov, M.M.; Kozlov, Y.N.; Shul’pin, G.B. EPR and DFT Study of Copper(II) Complex with 1,10-Phenanthroline as Catalyst for Oxidation of Cyclohexanol with tert-Butyl Hydroperoxide. Appl. Magn. Reason. 2022, 53, 887–894. [Google Scholar] [CrossRef]
  92. Fomenko, I.S.; Afewerki, M.; Gongola, M.I.; Vasilyev, E.S.; Shul’pina, L.S.; Ikonnikov, N.S.; Shul’pin, G.B.; Samsonenko, D.G.; Yanshole, V.V.; Nadolinny, V.A.; et al. Novel Copper(II) Complexes with Dipinodiazafluorene Ligands: Synthesis, Structure, Magnetic and Catalytic Properties. Molecules 2022, 27, 4072. [Google Scholar] [CrossRef]
  93. Shul’pin, G.B. Metal-catalysed hydrocarbon oxygenations in solutions: The dramatic role of additives: A review. J. Mol. Catal. A Chem. 2002, 189, 39–66. [Google Scholar] [CrossRef]
  94. Shul’pin, G.B.; Kozlov, Y.N.; Shul’pina, L.S.; Petrovskiy, P.V. Oxidation of alkanes and alcohols with hydrogen peroxide catalyzed by complex Os3(CO)10(mu-H)2. Appl. Organometal. Chem. 2010, 24, 464–472. [Google Scholar]
  95. Shilov, A.E.; Shul’pin, G.B. Activation and Catalytic Reactions of Saturated Hydrocarbons in the Presence of Metal Complexes; Kluwer Academic Publishers: New York, NY, USA; Boston, MA, USA; Dordrecht, The Netherlands; London, UK; Moscow, Russia, 2002. [Google Scholar]
  96. Shul’pin, G.B.; Nesterov, D.S.; Shul’pina, L.S.; Pombeiro, A.J.L. A hydroperoxo-rebound mechanism of alkane oxidation with hydrogen peroxide catalyzed by binuclear manganese(IV) complex in the presence of an acid with involvement of atmospheric dioxygen Part 14 from the series “Oxidations by the system hydrogen peroxide–[Mn2L2O3]2+ (L = 1,4,7-trimethyl-1,4,7-triazacyclononane)–carboxylic acid”. Inorg. Chim. Acta 2017, 455, 666–676. [Google Scholar]
  97. Shul’pin, G.B.; Gradinaru, J.; Kozlov, Y.N. Alkane hydroperoxidation with hydroperoxides catalysed by copper complexes. Org. Biomol. Chem. 2003, 1, 3611–3617. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. A general method for preparing copperphenylsilsesquioxane/benzoate complexes 13.
Scheme 1. A general method for preparing copperphenylsilsesquioxane/benzoate complexes 13.
Molecules 27 08505 sch001
Figure 1. Side (top) and top (bottom) views of the molecular structure of 1.
Figure 1. Side (top) and top (bottom) views of the molecular structure of 1.
Molecules 27 08505 g001
Figure 2. (Top) Representation of copper ions location and their multi-ligation style in 1. (Bottom) Distorted pentameric siloxane ligand in 1. The color code is the same as in Figure 1.
Figure 2. (Top) Representation of copper ions location and their multi-ligation style in 1. (Bottom) Distorted pentameric siloxane ligand in 1. The color code is the same as in Figure 1.
Molecules 27 08505 g002aMolecules 27 08505 g002b
Figure 3. Side view of the molecular structure of 2.
Figure 3. Side view of the molecular structure of 2.
Molecules 27 08505 g003
Figure 4. Accumulation of cyclohexanol and cyclohexanone in the oxidation of cyclohexane (0.46 M; 0.125 mL in 2.5 mL total volume) with hydrogen peroxide (2.0 M, 0.32 mL 50% aqueous in 2.5 mL total volume) catalyzed by compound 3 (5 × 10−4 M; 3 mg in 2.5 mL total volume) in the presence of HNO3 (0.05 M; 0.5 mL stock solution in 2.5 mL total volume) in MeCN at 50 °C. Concentrations of cyclohexanone and cyclohexanol were determined before the reduction of the solid PPh3 (A). The same reaction after reducing the aliquots with solid PPh3 (B).
Figure 4. Accumulation of cyclohexanol and cyclohexanone in the oxidation of cyclohexane (0.46 M; 0.125 mL in 2.5 mL total volume) with hydrogen peroxide (2.0 M, 0.32 mL 50% aqueous in 2.5 mL total volume) catalyzed by compound 3 (5 × 10−4 M; 3 mg in 2.5 mL total volume) in the presence of HNO3 (0.05 M; 0.5 mL stock solution in 2.5 mL total volume) in MeCN at 50 °C. Concentrations of cyclohexanone and cyclohexanol were determined before the reduction of the solid PPh3 (A). The same reaction after reducing the aliquots with solid PPh3 (B).
Molecules 27 08505 g004
Figure 5. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 3 on the initial concentration of cyclohexane.
Figure 5. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 3 on the initial concentration of cyclohexane.
Molecules 27 08505 g005
Figure 6. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 3 on the initial concentration of catalyst.
Figure 6. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 3 on the initial concentration of catalyst.
Molecules 27 08505 g006
Figure 7. Accumulation of cyclohexanone (yield = 40%, TON = 400) in oxidation of cyclohexanol (0.5 M; 0.14 mL in 2.5 mL total volume; blue—curve 1), acetophenone (yield = 90%, TON = 900) in oxidation of phenylethanol (0.5 M; 0.16 mL in 2.5 mL total volume; green—curve 2), 2-heptanone (yield = 40%, TON = 400) in oxidation of 2-heptanol (0.5 M; 0.14 mL in 2.5 mL total volume red—curve 3) and 2-hexanone(yield = 40%, TON = 400) in oxidation of 2-hexanol (0.5 M; 0.16 mL in 2.5 mL total volume; black—curve 4) with tert-butyl hydroperoxide (1.5 M; 0.36 mL 70% aqueous in 2.5 mL total volume) catalyzed by compound 3 (5 × 10−4 M; 3 mg) in the absence of HNO3 in MeCN at 50 °C.
Figure 7. Accumulation of cyclohexanone (yield = 40%, TON = 400) in oxidation of cyclohexanol (0.5 M; 0.14 mL in 2.5 mL total volume; blue—curve 1), acetophenone (yield = 90%, TON = 900) in oxidation of phenylethanol (0.5 M; 0.16 mL in 2.5 mL total volume; green—curve 2), 2-heptanone (yield = 40%, TON = 400) in oxidation of 2-heptanol (0.5 M; 0.14 mL in 2.5 mL total volume red—curve 3) and 2-hexanone(yield = 40%, TON = 400) in oxidation of 2-hexanol (0.5 M; 0.16 mL in 2.5 mL total volume; black—curve 4) with tert-butyl hydroperoxide (1.5 M; 0.36 mL 70% aqueous in 2.5 mL total volume) catalyzed by compound 3 (5 × 10−4 M; 3 mg) in the absence of HNO3 in MeCN at 50 °C.
Molecules 27 08505 g007
Table 1. Comparison of catalytic efficiency of copper-based catalysts including complex 3.
Table 1. Comparison of catalytic efficiency of copper-based catalysts including complex 3.
Catalytic SystemTotal Yield of Oxidation Products (%)Time (Hours)TON
Complex 3 (5 × 10−4 M) this work321290
Cu2Fe5-germsesquioxane [89]281260
Cu9Na6-silsesquioxane [26]215.0193
Cu(OAc)2 [97]7.62.0
Cu(NO3)2•3H2O + HNO3 (results of the authors)92.080
Cu(NO3)2•3H2O (results of the authors)224.0
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Bilyachenko, A.N.; Khrustalev, V.N.; Gutsul, E.I.; Zueva, A.Y.; Korlyukov, A.A.; Shul’pina, L.S.; Ikonnikov, N.S.; Dorovatovskii, P.V.; Gelman, D.; Shubina, E.S.; et al. Hybrid Silsesquioxane/Benzoate Cu7-Complexes: Synthesis, Unique Cage Structure, and Catalytic Activity. Molecules 2022, 27, 8505. https://doi.org/10.3390/molecules27238505

AMA Style

Bilyachenko AN, Khrustalev VN, Gutsul EI, Zueva AY, Korlyukov AA, Shul’pina LS, Ikonnikov NS, Dorovatovskii PV, Gelman D, Shubina ES, et al. Hybrid Silsesquioxane/Benzoate Cu7-Complexes: Synthesis, Unique Cage Structure, and Catalytic Activity. Molecules. 2022; 27(23):8505. https://doi.org/10.3390/molecules27238505

Chicago/Turabian Style

Bilyachenko, Alexey N., Victor N. Khrustalev, Evgenii I. Gutsul, Anna Y. Zueva, Alexander A. Korlyukov, Lidia S. Shul’pina, Nikolay S. Ikonnikov, Pavel V. Dorovatovskii, Dmitri Gelman, Elena S. Shubina, and et al. 2022. "Hybrid Silsesquioxane/Benzoate Cu7-Complexes: Synthesis, Unique Cage Structure, and Catalytic Activity" Molecules 27, no. 23: 8505. https://doi.org/10.3390/molecules27238505

Article Metrics

Back to TopTop