Next Article in Journal
Applications of MALDI-MS/MS-Based Proteomics in Biomedical Research
Previous Article in Journal
Discovery of New VEGFR-2 Inhibitors: Design, Synthesis, Anti-Proliferative Evaluation, Docking, and MD Simulation Studies
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel Family of Cage-like (CuLi, CuNa, CuK)-phenylsilsesquioxane Complexes with 8-Hydroxyquinoline Ligands: Synthesis, Structure, and Catalytic Activity

by
Alexey N. Bilyachenko
1,2,*,
Victor N. Khrustalev
2,3,
Anna Y. Zueva
1,2,
Ekaterina M. Titova
1,
Grigorii S. Astakhov
1,2,
Yan V. Zubavichus
4,
Pavel V. Dorovatovskii
5,
Alexander A. Korlyukov
1,6,
Lidia S. Shul’pina
1,
Elena S. Shubina
1,
Yuriy N. Kozlov
7,8,
Nikolay S. Ikonnikov
1,
Dmitri Gelman
9 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
Synchrotron Radiation Facility SKIF, Boreskov Institute of Catalysis SB RAS, Nikolskii Prosp. 1, 630559 Koltsovo, Russia
5
National Research Center “Kurchatov Institute”, Akademika Kurchatova pl. 1, 123182 Moscow, Russia
6
Pirogov Russian National Research Medical University, Ostrovitianov Str. 1, 117997 Moscow, Russia
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
9
Institute of Chemistry, Edmond J. Safra Campus, The Hebrew University of Jerusalem, Jerusalem 91904, Israel
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(19), 6205; https://doi.org/10.3390/molecules27196205
Submission received: 29 August 2022 / Revised: 9 September 2022 / Accepted: 13 September 2022 / Published: 21 September 2022
(This article belongs to the Section Organometallic Chemistry)

Abstract

:
The first examples of metallasilsesquioxane complexes, including ligands of the 8-hydroxyquinoline family 19, were synthesized, and their structures were established by single crystal X-ray diffraction using synchrotron radiation. Compounds 19 tend to form a type of sandwich-like cage of Cu4M2 nuclearity (M = Li, Na, K). Each complex includes two cisoid pentameric silsesquioxane ligands and two 8-hydroxyquinoline ligands. The latter coordinates the copper ions and corresponding alkaline metal ions (via the deprotonated oxygen site). A characteristic (size) of the alkaline metal ion and a variation of characteristics of nitrogen ligands (8-hydroxyquinoline vs. 5-chloro-8-hydroxyquinoline vs. 5,7-dibromo-8-hydroxyquinoline vs. 5,7-diiodo-8-hydroxyquinoline) are highly influential for the formation of the supramolecular structure of the complexes 3a, 5, and 79. The Cu6Na2-based compound 2 exhibits high catalytic activity towards the oxidation of (i) hydrocarbons by H2O2 activated with HNO3, and (ii) alcohols by tert-butyl hydroperoxide. Studies of kinetics and their selectivity has led us to conclude that it is the hydroxyl radicals that play a crucial role in this process.

1. Introduction

Nanosized silsesquioxane [RSiO1.5]n units bridge a gap between organics and inorganics with their pure inorganic Si-O-Si main chain, surrounded by various organic substituents [1,2,3,4,5]. Silsesquioxanes are often regarded as unique molecular models of silica, both for investigations into surface phenomena [6] and heterogeneous catalysis [7]. These structural features of silsesquioxanes are widely used to design hybrid (organic–inorganic) materials [8,9,10,11]. Another important opportunity provided by silsesquioxanes is their applicability as ligands for a large number of metallacomplexes [12]. In this context, incompletely condensed silsesquioxanes [13] should be mentioned as an especially efficient synthon for the design of different cage-like metallasilisesquioxanes (CLMSs) [14,15,16,17,18,19,20,21,22,23,24]. As a result, CLMSs offer an impressively wide scope of applications. These include the investigation of flame-retardant properties [25,26,27,28,29,30] and the development of approaches to anode [31,32] and ceramic [33] materials. Recent results revealed intriguing magnetic effects [34,35] (including the observation of spin glass [36] and single-molecule magnet [37] behaviors) and photophysical properties [38,39,40,41,42,43,44] of CLMSs. CLMSs are widely discussed as homo- and heterogeneous catalysts [45,46,47,48,49], with the most recent papers reporting the impressive activity of metallasilsesquioxanes in (i) Chan–Evans–Lam coupling [50], (ii) the hydroboration of ketones [51], (iii) CO2 valorization [52], (iv) the synthesis of bio-derived ethers [53], and (v) oxidative amidation [54].
Additionally, it is extremely well known that transition metal complexes efficiently catalyze the oxidations of hydrocarbons and alcohols involving the functionalization of C–H bonds [55,56,57,58]. Copper derivatives bearing various nitrogen-containing ligands are among the most active catalysts in the oxidation of organic compounds with peroxides [59,60,61,62,63,64,65,66,67,68,69], and this applies to the full extent of several Cu-based sesquioxane complexes [70,71,72,73,74]. Interested in further studying functional cage-like metallasilsesquioxanes, we have synthesized a series of novel copper silsesquioxanes via complexation with ligands of the 8-hydroxyquinoline family. Recent publications by us [50,75,76,77,78,79] and others [37,40,51,80,81] showed that the self-assembly of CLMSs in the presence of various (N,N-, P,P-, O,O-) organic ligands could be regarded as an efficient and facile approach to new types of mixed-ligand metallasilsesquioxane complexes.
To the best of our knowledge, such a popular bidentate ligand as 8-hydroxyquinoline has never been applied for the design of CLMS derivatives, despite an enormous number of other 8-hydroxyquinoline complexes and their multiple applications [82,83,84,85,86]. Here, we present our results on the synthesis of the very first examples of heteroligand silsesquioxane/8-hydroxyquinoline metallacomplexes and their evaluation towards the oxidative functionalization of hydrocarbons.

2. Results and Discussion

2.1. Synthesis and Structure

For a synthesis method, a convenient “siloxanolate route” was chosen. This approach implies the in situ formation of active lithium siloxanolate [PhSi(O)OLi]n species via the interaction of PhSi(OMe)3 with LiOH in an ethanol solution. Later on, these species interact with CuCl2, and the copper–lithium silsesquioxane formed in situ is treated (Scheme 1) with 8-hydroxyquinoline (hq). The use of lithium hydroxide for the CLMS design is in accordance with recent publications [30,87]. At the same time, classical approaches rely on more complicated reactions with either BuLi/LiN(SiMe3)2 (for metalation synthesis [88,89,90,91]) or LiCl (for transmetalation synthesis [92]).
As a result, the very first example of a CLMS/8-hydroxyquinoline complex, compound 1 of [(Ph5Si5O10)2(OH)2Cu6Li2hq2(EtOH)2]·EtOH composition, was isolated in a 33% yield.
Compound 1 represents the skewed sandwich-like cage structure, as established by an X-ray diffraction study (Figure 1). The central core of complex 1 includes two almost linear Cu…Cu…Cu…Li fragments (Figure 2, top), coordinated by two cyclic silsesquioxane [Ph5Si5O10] ligands (Figure 2, bottom). Obviously, the CuII6 nuclearity (giving 12 positive charges) could not be compensated by two pentameric silsesquioxane ligands (giving 10 negative charges). The electroneutrality of complex 1 is reached by the presence of two hydroxyl groups. Additionally, two opposite copper ions are coordinated by 8-hydroxyquinolinolate ligands. In turn, the hydroxyl groups of the pristine 8-hydroxyquinoline molecules are metallated by lithium ions, forming OLi units. It is worth nothing that, despite the enormous popularity of 8-hydroxyquinoline ligands, cage 1, to the best of our knowledge, is only the second example of dual M-O-M’ (copper/alkali metal) complexation (M = Cu and M’ = an alkali metal), with the first one being reported for a CuIK 8-hydroxyquinoline compound [93].
The molecular structure of “complex 1 type” is easily reproduced. Additional runs of this reaction produced complex 1a of the [(Ph5Si5O10)2(OH)2Cu6Li2hq2(EtOH)1.6(H2O)0.4]·EtOHsq composition in a 39% yield. Complex 1a is a close analog of compound 1 (the only exception being a partial replacement of ethanol molecules coordinating lithium centers by water molecules). This difference causes negligible changes in cage structures (e.g., the intramolecular Li…Li distance is equal to 10.3512(2) Å and 10.3347(1) Å for 1 and 1a, respectively).
Further exploration of this reaction, namely a shift to sodium-based complexes (via the replacement of LiOH reagent by NaOH in Scheme 2), allowed us to isolate a series of CuNa cage compounds 24. These compounds include complex 2 of the [(Ph5Si5O10)2(OH)2Cu6Na2hq2(EtOH)2(H2O)]·EtOH composition (32% yield, Figure 3, top). Additionally, implementation of 5-chloro-8-hydroxyquinoline (Cl-hq) allowed us to isolate complex 3 of the [(Ph5Si5O10)2(OH)2Cu6Na2(Cl-hq)2(n-BuOH)4] composition (20% yield, Figure 3, middle). Finally, the use of 5,7-dibromo-8-hydroxyquinoline (Br2-hq) allowed us to isolate complex 4 of the [(Ph5Si5O10)2(OH)2Cu6Na2(Br2-hq)2(EtOH)2(Me2CO)(H2O)]·2Me2CO EtOH composition (44% yield, Figure 3 bottom). Compounds 24 represent the same type of molecular architecture as 11a. Differences in solvate systems and the nature of 8-hydroxyquinolinolate ligands cause negligible changes in the structural features of their cages (e.g., the intramolecular Na…Na distance is equal to 11.4133(4) Å, 11.4487(2) Å, and 11.4005(6) Å for 2, 3, and 4, respectively). Nevertheless, we would like to mention, once again, the acetone-solvated compound 4. We have recently reported the use of acetone as the ligand of choice as a very efficient method of isolating crystalline cage metallasilsesquioxanes (12 complexes) [94]. The smooth formation of complex 4 (in a higher yield than for compounds 13) supports the conclusion concerning the acetone potential in the CLMS design. In turn, a change in the solvate system used for the synthesis of complex 3, namely the application of an EtOH solution (Supplementary Scheme S1), results in the isolation of compound 3a of the [(Ph5Si5O10)2(OH)2Cu6Na2(Cl-hq)2(EtOH)4] composition in a 26% yield. In terms of molecular structure, compounds 3 and 3a are very close analogs (the intramolecular Na…Na distances are equal to 11.4487(2) Å and 11.4747(4) Å for 3 and 3a, respectively). At the same time, unlike the discrete structure of 3, compound 3a exhibits T-stacking interactions between the phenyl groups of neighboring cage units (Supplementary Figure S1). This points to a strong influence of solvate ligands on the supramolecular assembly of CLMSs, as will be discussed in more detail below.
A further investigation of such self-assembly reaction conditions (with KOH as a starting reagent, Scheme 3) allowed us to isolate the 8-hydroxyquinolinolate-containing complex 5. Complex 5 of the [(Ph5Si5O10)2(OH)2Cu6K2(hq)2(EtOH)2] composition (37% yield, Figure 4, top) represents the same fashion of molecular architecture as discussed for complexes 14. At the same time, the strong individuality of complex 5 is revealed by its unusual supramolecular structure (Figure 4, bottom). The inter-cage connections are realized via the metallocene K·π-Ph joints between the potassium center of one cage and the phenyl group of the silicon atom, which belongs to a neighboring cage (Figure 5, top). Only two similar observations could be cited in this context:(i) Na·π-Ph joints in discrete Mn,Na CLMS [95] and (ii) Cs(Rb)·π-Ph joints in CuCs(Rb) CLMS coordination polymers [96]. Interestingly, each cage complex in the supramolecular structure of 5 interacts with neighbors via both two potassium centers and two phenyl groups, thus forming an unusual 2D coordination polymer structure (Figure 5, bottom).
Nevertheless, a general conclusion that could be derived from the comparison of complexes 15 agrees with the results of aforementioned publications [97,98,99] that a large alkali metal ion could be easily involved in inter-cage connections. Indeed, the only coordination polymer among complexes 15, compound 5, includes potassium ions (unlike lithium- or sodium-based cages 14). To make a correct conclusion regarding the role played by the alkali metal ion, a series of additional experiments with one type of ligand, namely 5,7-dibromo-8-hydroxyquinoline, was performed (Scheme 4). These experiments let us isolate three compounds, including the reference alkali metal ions: lithium-complex [(Ph5Si5O10)2(OH)2Cu6Li2(Br2-hq)2(EtOH)2]·4 EtOH 6, sodium-complex [(Ph5Si5O10)2(OH)2Cu6Na2(Br2-hq)2(EtOH)2]·4 EtOH 7, and potassium-complex [(Ph5Si5O10)2(OH)2Cu6K2(Br2-hq)2(Me2CO)4]·2Me2CO 8 in 18%, 27%, and 39% yields, respectively.
All these complexes (Figure 6) belong to a general type of molecular architecture discussed for 15. What is important is that compounds 6-8 exhibit profound differences in the context of their supramolecular geometry. According to formal expectations, complex 6, containing smaller (lithium) ions, exhibits no supramolecular structure. In turn, both compounds 78, containing larger sodium or potassium ions, respectively, pack in the crystal in a way that forms a ladder-like motif (Figure 7, top and bottom, respectively). Unlike the “metallocene-connected” 2D coordination polymer 6, 1D coordination polymers 78 are assembled via almost rectangular units, formed by two additional contacts between the bromide center of Br2-hq ligand of one cage and sodium (or potassium) center of the neighboring cage. Distances Na…Br (in 7) and K…Br (in 8) are equal to 3.317(2) and 3.3816(10) Å, respectively, which points to their coordination/ionic character. Here, it is especially interesting to compare the “supramolecular” behaviors of close analogs, namely compounds 4 and 7,Cu6Na2-based complexes with 5,7-dibromo-8-hydroxyquinoline ligands. While compound 4 crystallizes in an island-like structure, compound 7 forms the aforementioned 1D coordination polymer. As the only difference between these two complexes is a set of solvate ligands (acetone/ethanol/water in the case of 4, and ethanol only in the case of 7), the role played by the solvate molecules in the formation/cleavage of the inter-cage association could be concluded. This observation is in perfect accordance with several earlier reports [72,97,98,100,101], and will definitely yield more confirmations in future research.
Considering the high efficiency of large-sized alkali metal ions, it was interesting to study the same effect for modified 8-hydroxyquinoline ligands bearing extra substituents. In this line of thought, the self-assembly synthesis of CuK-phenylsilsesquioxane with 5,7-diiodo-8-hydroxyquinoline (I2-hq) ligands has been performed (Scheme 5). This reaction smoothly produced CLMS 9 of the [(Ph5Si5O10)2(OH)2Cu6K2(I2-hq)2]·⅙MeOH composition in a 33% yield. The type of molecular structure of complex 9 (Figure 8) is similar to those of 18, but its supramolecular architecture is quite different compared to compounds 5, 7, and 8. To our surprise, no iodide centers get involved in the supramolecular interactions (K…I interactions are normal non-valence intramolecular ones). Nevertheless, the molecules in 9 pack into an unprecedented 3D structure via the crown ether-like contacts of the potassium ion of one cage and two oxygen atoms of the silsesquioxane ligand of the neighboring cage (Figure 9). Thus, in the crystal, compound 9 forms the unique “solvent-free” 3D framework via potassium ions, binding no solvate methanol molecules.
In the context of functional properties, a representative compound 2 (Cu6Na2-based complex) has been tested as a homogeneous catalyst towards the oxidative functionalization of hydrocarbons (including cyclohexane and n-heptane) and alcohols.

2.2. Catalyzed Oxidation of Cyclohexane

Complex 2 was studied more carefully in the oxidation of hydrocarbons with cyclohexane as a case example. Typical oxidation kinetics curves of cyclohexane are presented in Figure 10. It should be noted that under the studied conditions, nitric acid additives practically do not affect the rate of formation of cyclohexyl hydroperoxide within an hour. However, these additives have a significant effect on the subsequent course of the process. The catalyst deactivation observed during the reaction is apparently due to the destruction of organic ligands bound to copper ions and its hydrolysis in the presence of water traces. The striking similarity of the initial oxidation rates with hydrogen peroxide and tert-butyl hydroperoxide is also worth noting. It indicates a similar reactivity of these peroxides in their catalytic conversion under the action of 2. Initial reaction rates were found at various initial cyclohexane concentrations, presented below in Figure 10 and Figure 11.
We found that the rate of cyclohexyl hydroperoxide formation at a concentration of cyclohexane of 0.05 M decreases by a factor of 4.3 with the addition of linear alkane n-heptane at a concentration of 0.4 M. At the same time, the total yield of alkyl hydroperoxides generated from cyclohexane and n-heptane practically coincides with the yield of cyclohexyl hydroperoxide at a concentration of 0.46 M, in the absence of n-heptane additives. These data, as well as the mode of dependence of the initial rate of cyclohexyl hydroperoxide formation on the cyclohexane concentration (Figure 11), indicate that the stimulation of alkane oxidation is due to the same intermediate species, which decomposes hydrogen peroxide.
Taking into account the data presented above, as well as data on the selectivity of the oxidation of linear alkanes (the ratios of relative reactivities are C-1:C-2:C-3:C-4 = 1.0:2.0:2.4:2.5 for the n-heptane oxidation and 1:2:3 = 1.0:5.6:14.5 for the methylcyclohexane oxidation), indicates that it is hydroxyl radicals that act as the active species. The kinetic scheme, which describes the obtained results, is shown below:
2→ Z
Z → Y
H2O2 + Z →…→ HO
HO + CH3CN →…→ oxidation products
HO + C6H12 →…→ ROOH
HO +n-C7H16 →…→ rOOH
Here, the initial stage (Equation (1)) includes the transformation of the pristine complex 2 in the presence of water and 0.05 M HNO3 in the intermediate state Z, which genuinely manifests catalytic activity in the decomposition of H2O2 to generate hydroxyl radicals at a rate of W3 (for reaction 3).
The next step (Equation (2)) is the transformation of Z into a catalytically inactive form, Y.
Reactions (Equation (3)) are the sequence of transformations leading to the formation of HO radicals. The limiting stage of these reactions is the interaction of H2O2 with species Z.
Equations (4)–(6) describe the sequence of the transformations leading to the formation of oxidation products of acetonitrile (Equation (4)), cyclohexyl hydroperoxide (Equation (5)), and n-heptane hydroperoxides (Equation (6)), with the limiting stages being the HO﮲ interactions with CH3CN, C6H12, and n-C7H16, characterized by the rate constants k4, k5, and k6, respectively.
From the proposed kinetic scheme, within the framework of the quasi-stationary approximation, it is easy to obtain expressions for the rate of formation of ROOH (Equation (7)) and the rate of formation of the sum of peroxides ROOH and rOOH (Equation (8)).
d [ROOH]/dt = W3k5 [C6H12]/(k4 [CH3CN] + k5 [C6H12] + k6 [n-C7H16])
d ([ROOH] + [rOOH])/dt = W3 • (k5 [C6H12] + k6 [n-C7H16])/(k4 [CH3CN] + k5 [C6H12] + k6 [n-C7H16])
Let us employ Equations (7) and (8) to analyze the observed effect of the n-heptane additive. The proximity of the formation rates of ROOH at [C6H12] = 0.46 M and [n-C7H16] = 0 and the sums of ROOH and rOOH at [C6H12] = 0.05 M and [n-C7H16] = 0.4 M allows us to obtain the ratio 0.46k5/(k4 [CH3CN] + 0.46k5) ≈ (0.05k5 + 0.4k6)/(k4 [CH3CN] + 0.05k5 + 0.4k6), which is valid at k5k6. The coincidence of rate constants of the interaction of HO radicals with cyclohexane and n-heptane is consistent with published data [102,103,104]. The decrease in the rate of ROOH formation at [C6H12] = 0.05 M from the addition of 0.4 M n-C7H16 by a factor of 4.3 allows us to make another estimate of the ratio of rate constants. Using (Equation (7)) for these conditions and the relation k5k6 obtained above, we have:
(k4[CH3CN]/0,05 k5 + 1 + 8)/(k4[CH3CN]/0.05k5 + 1) = 4.3.
k4[CH3CN]/k5 = 0.086 M.
From the last relation, it follows that k4 [CH3CN]/k5 = 0.071 M.
An analysis of the experimental dependence of the initial rate of ROOH formation on the concentration of C6H12 (Figure 11A) in the coordinates corresponding to the linear anamorphosis of Equation (7), [C6H12]/(d [ROOH]/dt)0—ordinate, [C6H12]—abscissa (its results are presented in Figure 11B), showed the correspondence of the experimental data of the proposed kinetic model; a satisfactory linear relationship is observed in the proposed coordinates. Further, from the ratio of the magnitude of the segment of this dependence to the tangent of the angle of inclination, it follows:
k4 [CH3CN]/k5 = 0.086 M.
This value of the ratio of rate constants, as well as the value obtained above from the data on the effect of n-heptane additives on the rate of ROOH formation at [C6H12] = 0.05 M, are close to the value characteristic of reactions involving hydroxyl radicals. Thus, in addition to the regioselectivity data for the oxidation of linear alkanes, which indicate the induction of the oxidation reaction by hydroxyl radicals, we also have kinetic evidence for this.

2.3. Catalyzed Oxidation of Alcohols

Figure 12A–C represent kinetic curves of ketone accumulation during the oxidation of alcohols (including three substrates, 2-heptanol, cyclohexanol, and 1-phenylethanol) with tert-butylhydroperoxide in acetonitrile. 1-Phenylethanol is the most active substrate. It should be noted that the oxidation with hydrogen peroxide of this alcohol proceeds much less efficiently (see curve 2 in Figure 12C). Attention should be drawn to the fact that the oxidation of phenylethanol with hydrogen peroxide is low compared to tert-butyl hydroperoxide (Figure 12C). In contrast, the alkane oxidation with hydrogen peroxide in the presence of the same catalyst is extremely efficient under the same conditions (Figure 7). Perhaps, the mechanism of decomposition of hydrogen peroxide and the rate of the generation of particles induces the oxidation of alcohol change in the presence of alcohol. To establish the reason for this puzzling observation, more studies are needed in the future.

3. Conclusions

The very first examples of metallacomplexes, including simultaneously silsesquioxane and 8-hydroxyquinoline-derived ligands, were prepared via a convenient self-assembly approach. The exchange reaction of alkali-metal-based (M = Li, Na, or K) phenylsiloxanolates [PhSi(O)OM]x and copper (II) chloride, followed by the complexation with ligands of the 8-hydroxyquinoline family, smoothly lead to a series of similar sandwich-like octanuclear (Cu6M2) cage-like complexes. Each cage structure includes two pairs of ligands: two pentameric (Ph5Si5O10) cycles and two N,O-ligands of the 8-hydroxyquinolinolate type (derived from 8-hydroxyquinoline, 5-chloro-8-hydroxyquinoline, 5,7-dibromo-8-hydroxyquinoline, or 5,7-diiodo-8-hydroxyquinoline). Hydroxyl groups of such N,O-ligands are metallated by the corresponding metal ions (lithium, sodium, or potassium). The size of these metal ions is responsible for the supramolecular packing of cage complexes: (i) lithium-based compounds are island-like structures, (ii) potassium-based compounds are coordination polymers, and (iii) “intermediate” sodium-based compounds could realize both options, depending on additional factors. In sum, the use of a single type of organic ligand of the 8-hydroxyquinoline family allowed us to develop an impressively powerful approach to similar (in terms of the molecular cage structure) complexes, exhibiting principally different features of supramolecular packing (from discrete 0D to extended 3D geometry). The representative complex Cu6Na2-silsesquioxane/8-hydroxyquinoline 2 exhibits high catalytic activity in the oxidation of hydrocarbons and alcohols with H2O2 or tert-butyl hydroperoxide. Studies of kinetics and selectivity led us to the conclusion that the oxidation proceeds mainly with the participation of free hydroxyl radicals, as evidenced by the specific regioselectivity for individual chemical bonds in the oxidation of n-heptane and methylcyclohexane, as well as the dependence of the reaction rate on the initial concentration of cyclohexane. Of particular importance is the addition of nitric acid to hydrogen peroxide as a co-catalyst. Both research lines, (i) the design of heteroligand metallasilsesquioxane complexes and (ii) their evaluation in catalytic applications, are currently in further progress in our groups.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27196205/s1: Details of syntheses (including isolation and T-stacking structure of complex 3a), catalyzed reactions, and X-ray diffraction studies (CCDC 2203816-2203826) [105,106,107,108,109,110,111,112,113,114].

Author Contributions

Conceptualization, A.N.B. and G.B.S.; investigation, V.N.K., A.Y.Z., E.M.T., G.S.A., Y.V.Z., P.V.D., A.A.K., L.S.S., E.S.S., N.S.I., and D.G.; data curation, Y.N.K.; writing—original draft preparation, 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 22-13-00252.

Acknowledgments

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 the 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. acknowledges RUDN University Scientific Projects Grant System, project No. 025239-2-174, the ISF (Israel Science Foundation) Grant No. 370/20 and Esther K. and M. Mark Watkins Chair for Synthetic Organic Chemistry.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Kickelbick, G. Silsesquioxanes. In Functional Molecular Silicon Compounds I; Scheschkewitz, D., Ed.; Springer: Berlin/Heidelberg, Germany, 2014; pp. 1–28. [Google Scholar] [CrossRef]
  2. Loman-Cortes, P.; Huq, T.B.; Vivero-Escoto, J.L. Use of Polyhedral Oligomeric Silsesquioxane (POSS) in Drug Delivery, Photodynamic Therapy and Bioimaging. Molecules 2021, 26, 6453. [Google Scholar] [CrossRef] [PubMed]
  3. Mohamed, M.G.; Kuo, S.-W. Progress in the self-assembly of organic/inorganic polyhedral oligomeric silsesquioxane (POSS) hybrids. Soft Matter 2022, 18, 5535–5561. [Google Scholar] [CrossRef] [PubMed]
  4. Ding, S.; Zhao, S.; Gan, X.; Sun, A.; Xia, Y.; Liu, Y. Design of Fluorescent Hybrid Materials Based on POSS for Sensing Applications. Molecules 2022, 27, 3137. [Google Scholar] [CrossRef]
  5. Nowacka, M.; Kowalewska, A. Self-Healing Silsesquioxane-Based Materials. Polymers 2022, 14, 1869. [Google Scholar] [CrossRef]
  6. Feher, F.J.; Newman, D.A.; Walzer, J.F. Silsesquioxanes as models for silica surfaces. J. Am. Chem. Soc. 1989, 111, 1741–1748. [Google Scholar] [CrossRef]
  7. 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]
  8. Zhou, H.; Chua, M.H.; Xu, J. Functionalized POSS-Based Hybrid Composites. In Polymer Composites with Functionalized Nanoparticle, Synthesis, Properties, and Applications; Pielichowski, K., Majka, T.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2019; pp. 179–210. [Google Scholar] [CrossRef]
  9. Dong, F.; Lu, L.; Ha, C. Silsesquioxane-Containing Hybrid Nanomaterials: Fascinating Platforms for Advanced Applications. Macromol. Chem. Phys. 2019, 220, 1800324. [Google Scholar] [CrossRef]
  10. Du, Y.; Liu, H. Cage-like silsesquioxanes-based hybrid materials. Dalton Trans. 2020, 49, 5396–5405. [Google Scholar] [CrossRef] [PubMed]
  11. Ahmed, N.; Fan, H.; Dubois, P.; Zhang, X.; Fahad, S.; Aziz, T.; Wan, J. Nano-engineering and micromolecular science of polysilsesquioxane materials and their emerging applications. J. Mater. Chem. A 2019, 7, 21577–21604. [Google Scholar] [CrossRef]
  12. Feher, F.J.; Budzichowski, T.A. Silasesquioxanes as ligands in inorganic and organometallic chemistry. Polyhedron 1995, 14, 3239–3253. [Google Scholar] [CrossRef]
  13. Lee, D.W.; Kawakami, Y. Incompletely Condensed Silsesquioxanes: Formation and Reactivity. Polym. J. 2007, 39, 230–238. [Google Scholar] [CrossRef]
  14. Murugavel, R.; Voigt, A.; Walawalkar, M.G.; Roesky, H.W. Hetero- and Metallasiloxanes Derived from Silanediols, Disilanols, Silanetriols, and Trisilanols. Chem. Rev. 1996, 96, 2205–2236. [Google Scholar] [CrossRef]
  15. 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]
  16. 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]
  17. 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.–A Eur. J. 2004, 10, 4106–4114. [Google Scholar] [CrossRef] [PubMed]
  18. Lorenz, V.; Edelmann, F.T. Metallasilsesquioxanes. Adv. Organomet. Chem. 2005, 53, 101–153. [Google Scholar]
  19. 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]
  20. 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] [CrossRef]
  21. 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]
  22. Zheng, Y.; Gao, Z.; Han, J. Current Chemistry of Cyclic Oligomeric Silsesquioxanes. Curr. Org. Chem. 2018, 21, 2814–2828. [Google Scholar] [CrossRef]
  23. 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]
  24. 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]
  25. Wu, H.; Zeng, B.; Chen, J.; Wu, T.; Li, Y.; Liu, Y.; Dai, L. An intramolecular hybrid of metal polyhedral oligomeric silsesquioxanes with special titanium-embedded cage structure and flame retardant functionality. Chem. Eng. J. 2019, 374, 1304–1316. [Google Scholar] [CrossRef]
  26. Ye, X.; Li, J.; Zhang, W.; Yang, R.; Li, J. Fabrication of eco-friendly and multifunctional sodium-containing polyhedral oligomeric silsesquioxane and its flame retardancy on epoxy resin. Compos. Part B Eng. 2020, 191, 107961. [Google Scholar] [CrossRef]
  27. Zeng, B.; Hu, R.; Zhou, R.; Shen, H.; Liu, X.; Chen, G.; Xu, Y.; Yuan, C.; Luo, W.; Dai, L. Co-flame retarding effect of ethanolamine modified titanium-containing polyhedral oligomeric silsesquioxanes in epoxy resin. Appl. Organomet. Chem. 2020, 34, e5266. [Google Scholar] [CrossRef]
  28. Ye, X.; Li, J.; Zhang, W.; Pan, Y.-T.; Yang, R.; Li, J. Enhanced fire safety and mechanical properties of epoxy resin composites based on submicrometer-sized rod-structured methyl macrocyclic silsesquioxane sodium salt. Chem. Eng. J. 2021, 425, 130566. [Google Scholar] [CrossRef]
  29. 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]
  30. Ye, X.; Zhang, X.; Jiang, Y.; Qiao, L.; Zhang, W.; Pan, Y.-T.; Yang, R.; Li, J.; Li, Y. Controllable dimensions and regular geometric architectures from self-assembly of lithium-containing polyhedral oligomeric silsesquioxane: Build for enhancing the fire safety of epoxy resin. Compos. Part B Eng. 2021, 229, 109483. [Google Scholar] [CrossRef]
  31. 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]
  32. 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]
  33. Loganathan, P.; Pillai, R.S.; Jeevananthan, V.; David, E.; Palanisami, N.; Bhuvanesh, N.S.P.; Shanmugan, S. Assembly of discrete and oligomeric structures of organotin double-decker silsesquioxanes: Inherent stability studies. New J. Chem. 2021, 45, 20144–20154. [Google Scholar] [CrossRef]
  34. Levitsky, M.M.; Bilyachenko, A.N.; Shubina, E.S.; Long, J.; Guari, Y.; Larionova, J. Magnetic cage-like metallasilsesquioxanes. Coord. Chem. Rev. 2019, 398, 213015. [Google Scholar] [CrossRef]
  35. 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. [Google Scholar] [CrossRef]
  36. Bilyachenko, A.N.; Yalymov, A.I.; Korlyukov, A.A.; Long, J.; Larionova, J.; Guari, Y.; Zubavichus, Y.V.; Trigub, A.L.; Shubina, E.S.; Eremenko, I.L.; et al. Heterometallic Na6 Co3 Phenylsilsesquioxane Exhibiting Slow Dynamic Behavior in its Magnetization. Chem.–A Eur. J. 2015, 21, 18563–18565. [Google Scholar] [CrossRef]
  37. 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]
  38. Marchesi, S.; Carniato, F.; Boccaleri, E. Synthesis and characterisation of a novel europium(iii)-containing heptaisobutyl-POSS. New J. Chem. 2014, 38, 2480–2485. [Google Scholar] [CrossRef]
  39. 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.–A Eur. J. 2020, 26, 16594–16598. [Google Scholar] [CrossRef]
  40. 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]
  41. 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] [PubMed]
  42. 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]
  43. 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]
  44. 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]
  45. Abbenhuis, H.C.L. Advances in Homogeneous and Heterogeneous Catalysis with Metal-Containing Silsesquioxanes. Chem. Eur. J. 2000, 6, 25–32. [Google Scholar] [CrossRef]
  46. 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]
  47. 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]
  48. 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. A Chem. 2017, 426, 297–304. [Google Scholar] [CrossRef]
  49. 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]
  50. 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]
  51. Liu, Y.-N.; Su, H.-F.; Li, Y.-W.; Liu, Q.-Y.; Jagličić, Z.; Wang, W.-G.; Tung, C.-H.; Sun, D. Space Craft-like Octanuclear Co(II)-Silsesquioxane Nanocages: Synthesis, Structure, Magnetic Properties, Solution Behavior, and Catalytic Activity for Hydroboration of Ketones. Inorg. Chem. 2019, 58, 4574–4582. [Google Scholar] [CrossRef]
  52. 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]
  53. 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]
  54. 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. [Google Scholar] [CrossRef]
  55. Shilov, A.E.; Shul’Pin, G.B. Activation of C−H Bonds by Metal Complexes. Chem. Rev. 1997, 97, 2879–2932. [Google Scholar] [CrossRef] [PubMed]
  56. Shul’Pin, G.B. New Trends in Oxidative Functionalization of Carbon–Hydrogen Bonds: A Review. Catalysts 2016, 6, 50. [Google Scholar] [CrossRef]
  57. 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]
  58. Wójtowicz-Młochowska, 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 2016, 2017, 12–58. [Google Scholar] [CrossRef]
  59. Punniyamurthy, T.; Rout, L. Recent advances in copper-catalyzed oxidation of organic compounds. Coord. Chem. Rev. 2008, 252, 134–154. [Google Scholar] [CrossRef]
  60. Kirillov, A.M.; Kirillova, M.V.; Pombeiro, A.J. Multicopper complexes and coordination polymers for mild oxidative functionalization of alkanes. Coord. Chem. Rev. 2012, 256, 2741–2759. [Google Scholar] [CrossRef]
  61. 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]
  62. Conde, A.; Vilella, L.; Balcells, D.; Díaz-Requejo, M.M.; Lledós, A.; Pérez, P.J. Introducing Copper as Catalyst for Oxidative Alkane Dehydrogenation. J. Am. Chem. Soc. 2013, 135, 3887–3896. [Google Scholar] [CrossRef]
  63. 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]
  64. Kirillova, M.V.; de Paiva, P.T.; Carvalho, W.A.; Mandelli, D.; Kirillov, A.M. Mixed-ligand aminoalcohol-dicarboxylate copper(II) coordination polymers as catalysts for the oxidative functionalization of cyclic alkanes and alkenes. Pure Appl. Chem. 2016, 89, 61–73. [Google Scholar] [CrossRef]
  65. Armakola, E.; Colodrero, R.M.P.; Bazaga-García, M.; Salcedo, I.R.; Choquesillo-Lazarte, D.; Cabeza, A.; Kirillova, M.V.; Kirillov, A.M.; Demadis, K.D. Three-Component Copper-Phosphonate-Auxiliary Ligand Systems: Proton Conductors and Efficient Catalysts in Mild Oxidative Functionalization of Cycloalkanes. Inorg. Chem. 2018, 57, 10656–10666. [Google Scholar] [CrossRef] [PubMed]
  66. 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 the oxidation of alkanes and alcohols with peroxides. Dalton Trans. 2019, 48, 12656–12673. [Google Scholar] [CrossRef] [PubMed]
  67. Costa, I.F.M.; Kirillova, M.V.; André, V.; Fernandes, T.A.; Kirillov, A.M. Tetracopper(II) Cores Driven by an Unexplored Trifunctional Aminoalcohol Sulfonic Acid for Mild Catalytic C–H Functionalization of Alkanes. Catalysts 2019, 9, 321. [Google Scholar] [CrossRef]
  68. 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, 512, 119889. [Google Scholar] [CrossRef]
  69. Shul’Pin, G.B.; Shul’Pina, L.S. Oxidation of Organic Compounds with Peroxides Catalyzed by Polynuclear Metal Compounds. Catalysts 2021, 11, 186. [Google Scholar] [CrossRef]
  70. 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]
  71. 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]
  72. Kulakova, A.N.; Bilyachenko, A.N.; Levitsky, M.M.; Khrustalev, V.N.; Korlyukov, A.A.; Zubavichus, Y.V.; Dorovatovskii, P.V.; Lamaty, F.; Bantreil, X.; Villemejeanne, B.; et al. Si10Cu6N4 Cage Hexacoppersilsesquioxanes Containing N Ligands: Synthesis, Structure, and High Catalytic Activity in Peroxide Oxidations. Inorg. Chem. 2017, 56, 15026–15040. [Google Scholar] [CrossRef]
  73. 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]
  74. 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]
  75. 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]
  76. 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]
  77. 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]
  78. 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]
  79. Astakhov, G.S.; Bilyachenko, A.N.; Levitsky, M.M.; Shul’Pina, L.S.; Korlyukov, A.A.; Zubavichus, Y.V.; Khrustalev, V.N.; Vologzhanina, A.V.; Shubina, E.S.; Dorovatovskii, P.V.; et al. Coordination Affinity of Cu(II)-Based Silsesquioxanes toward N,N-Ligands and Associated Skeletal Rearrangements: Cage and Ionic Products Exhibiting a High Catalytic Activity in Oxidation Reactions. Inorg. Chem. 2020, 59, 4536–4545. [Google Scholar] [CrossRef]
  80. Anisimov, A.A.; Vysochinskaya, Y.S.; Kononevich, Y.N.; Dolgushin, F.M.; Muzafarov, A.M.; Shchegolikhina, O.I. Polyhedral phenylnickelsodiumsiloxanolate transformation in the presence of aromatic nitrogen-containing ligands. Inorg. Chim. Acta 2020, 517, 120160. [Google Scholar] [CrossRef]
  81. Sergienko, N.V.; Trankina, E.S.; Polshchikova, N.V.; Korlyukov, A.A. Metallosiloxanes with Acetylacetonate Ligands. Russ. J. Coord. Chem. 2021, 47, 382–392. [Google Scholar] [CrossRef]
  82. Singh, D.; Nishal, V.; Bhagwan, S.; Saini, R.K.; Singh, I. Electroluminescent materials: Metal complexes of 8-hydroxyquinoline—A review. Mater. Des. 2018, 156, 215–228. [Google Scholar] [CrossRef]
  83. Prachayasittikul, V.; Prachayasittikul, V.; Prachayasittikul, S.; Ruchirawat, S. 8-Hydroxyquinolines: A review of their metal chelating properties and medicinal applications. Drug Des. Dev. Ther. 2013, 7, 1157–1178. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  84. Savić-Gajić, I.M.; Savic, I. Drug design strategies with metal-hydroxyquinoline complexes. Expert Opin. Drug Discov. 2019, 15, 383–390. [Google Scholar] [CrossRef] [PubMed]
  85. Albrecht, M.; Fiege, M.; Osetska, O. 8-Hydroxyquinolines in metallosupramolecular chemistry. Coord. Chem. Rev. 2008, 252, 812–824. [Google Scholar] [CrossRef]
  86. Lipunova, G.N.; Nosova, E.V.; Charushin, V.N.; Chupakhin, O. Structural, Optical Properties, and Biological Activity of Complexes Based on Derivatives of Quinoline, Quinoxaline, and Quinazoline with Metal Centers from Across the Periodic Table. Comments Inorg. Chem. 2014, 34, 142–177. [Google Scholar] [CrossRef]
  87. 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] [PubMed]
  88. Annand, J.; Aspinall, H.C.; Steiner, A. Novel Heterometallic Lanthanide Silsesquioxane. Inorg. Chem. 1999, 38, 3941–3943. [Google Scholar] [CrossRef]
  89. Lorenz, V.; Gießmann, S.; Gun’Ko, Y.K.; Fischer, A.K.; Gilje, J.W.; Edelmann, F.T. Fully Metalated Silsesquioxanes: Building Blocks for the Construction of Catalyst Models. Angew. Chem. Int. Ed. 2004, 43, 4603–4606. [Google Scholar] [CrossRef]
  90. García, C.; Gómez, M.; Gómez-Sal, P.; Hernández, J.M. Monocyclopentadienyl(niobium) Compounds with Imido and Silsesquioxane Ligands: Synthetic, Structural and Reactivity Studies. Eur. J. Inorg. Chem. 2009, 2009, 4401–4415. [Google Scholar] [CrossRef]
  91. Gau, M.R.; Zdilla, M.J. Multinuclear Clusters of Manganese and Lithium with Silsesquioxane-Derived Ligands: Synthesis and Ligand Rearrangement by Dioxygen- and Base-Mediated Si–O Bond Cleavage. Inorg. Chem. 2021, 60, 2866–2871. [Google Scholar] [CrossRef] [PubMed]
  92. Sergienko, N.V.; Korlyukov, A.A.; Myakushev, V.D.; Antipin, M.Y.; Zavin, B.G. Ion exchange in bimetallic cage copper organosiloxanes and the synthesis of polynuclear metal complexes containing Li and CuII atoms. Bull. Acad. Sci. USSR Div. Chem. Sci. 2009, 58, 2258–2265. [Google Scholar] [CrossRef]
  93. Giri, R.; Brusoe, A.; Troshin, K.; Wang, J.Y.; Font, M.; Hartwig, J.F. Mechanism of the Ullmann Biaryl Ether Synthesis Catalyzed by Complexes of Anionic Ligands: Evidence for the Reaction of Iodoarenes with Ligated Anionic CuI Intermediates. J. Am. Chem. Soc. 2017, 140, 793–806. [Google Scholar] [CrossRef]
  94. 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]
  95. Bilyachenko, A.; Dronova, M.S.; Korlyukov, A.A.; Levitsky, M.M.; Antipin, M.Y.; Zavin, B.G. Cage-like manganesephenylsiloxane with an unusual structure. Bull. Acad. Sci. USSR Div. Chem. Sci. 2011, 60, 1762–1765. [Google Scholar] [CrossRef]
  96. 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]
  97. 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]
  98. 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]
  99. 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 Copper(II) Phenylsilsesquioxanes: Synthesis and High Activity in Oxidative Catalysis. ChemCatChem 2017, 9, 4437–4447. [Google Scholar] [CrossRef]
  100. Dronova, M.S.; Bilyachenko, A.; Korlyukov, A.A.; Arkhipov, D.E.; Kirilin, A.D.; Shubina, E.; Babakhina, G.M.; Levitskii, M.M. A new type of supramolecular organization in the cage-like metallasiloxanes. Bull. Acad. Sci. USSR Div. Chem. Sci. 2013, 62, 1941–1943. [Google Scholar] [CrossRef]
  101. Dronova, M.S.; Bilyachenko, A.N.; Yalymov, A.I.; Kozlov, Y.N.; Shul’Pina, L.S.; Korlyukov, A.A.; Arkhipov, D.E.; Levitsky, M.M.; Shubina, E.S.; Shul’Pin, G.B. Solvent-controlled synthesis of tetranuclear cage-like copper(ii) silsesquioxanes. Remarkable features of the cage structures and their high catalytic activity in oxidation with peroxides. Dalton Trans. 2013, 43, 872–882. [Google Scholar] [CrossRef]
  102. 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] [CrossRef]
  103. 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]
  104. Nesterov, D.S.; Chygorin, E.N.; Kokozay, V.N.; Bon, V.V.; Boča, R.; Kozlov, Y.N.; Shul’Pina, L.S.; Jezierska, J.; Ozarowski, A.; Pombeiro, A.J.L.; et al. Heterometallic CoIII4FeIII2Schiff Base Complex: Structure, Electron Paramagnetic Resonance, and Alkane Oxidation Catalytic Activity. Inorg. Chem. 2012, 51, 9110–9122. [Google Scholar] [CrossRef]
  105. 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]
  106. Olivo, G.; Lanzalunga, O.; Di Stefano, S. Non-Heme Imine-Based Iron Complexes as Catalysts for Oxidative Processes (Review). Adv. Synth. Catal. 2016, 358, 843–863. [Google Scholar] [CrossRef]
  107. Czerwińska, K.; Machura, B.; Kula, S.; Krompiec, S.; Erfurt, K.; Roma-Rodrigues, C.; Fernandes, A.R.; Shul’pina, L.S.; Ikonnikov, N.S.; Shul’pin, G.B. Copper(II) complexes of functionalized 2,2′:6′,2″-terpyridines and 2,6-di(thiazol-2-yl)pyridine: Structure, spectroscopy, cyto-toxicity and catalytic activity. Dalton Trans. 2017, 46, 9591–9604. [Google Scholar] [CrossRef] [Green Version]
  108. Bruker. SAINT; Bruker AXS Inc.: Madison, WI, USA, 2013. [Google Scholar]
  109. 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]
  110. CrysAlisPro, Version 1.171.41.106a; Rigaku Oxford Diffraction: Tokyo, Japan, 2021.
  111. 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 Cryst. 2011, D67, 271–281. [Google Scholar]
  112. Evans, P.R. Scaling and assessment of data quality. Acta Cryst. 2006, D62, 72–82. [Google Scholar]
  113. Spek, A.L. PLATON, A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2015. [Google Scholar]
  114. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar]
Scheme 1. General scheme of the synthesis of CuLi-silsesquioxane/8-hydroxyquinoline complex 1. Solvated molecules are omitted for clarity.
Scheme 1. General scheme of the synthesis of CuLi-silsesquioxane/8-hydroxyquinoline complex 1. Solvated molecules are omitted for clarity.
Molecules 27 06205 sch001
Figure 1. Two projections of the molecular structure of 1.
Figure 1. Two projections of the molecular structure of 1.
Molecules 27 06205 g001aMolecules 27 06205 g001b
Figure 2. Top. The metal-containing central core of 1. Bottom. The silsesquioxane ligand of 1. Phenyl groups at the silicon atoms are not shown. The color code is the same as in Figure 1.
Figure 2. Top. The metal-containing central core of 1. Bottom. The silsesquioxane ligand of 1. Phenyl groups at the silicon atoms are not shown. The color code is the same as in Figure 1.
Molecules 27 06205 g002
Scheme 2. General scheme of the synthesis of CuNa-silsesquioxane complexes with ligands of the 8-hydroxyquinoline family 24. Solvated molecules are omitted for clarity.
Scheme 2. General scheme of the synthesis of CuNa-silsesquioxane complexes with ligands of the 8-hydroxyquinoline family 24. Solvated molecules are omitted for clarity.
Molecules 27 06205 sch002
Figure 3. Top. The molecular structure of 2. Middle. The molecular structure of 3. Bottom. The molecular structure of 4.
Figure 3. Top. The molecular structure of 2. Middle. The molecular structure of 3. Bottom. The molecular structure of 4.
Molecules 27 06205 g003aMolecules 27 06205 g003b
Scheme 3. General scheme of the synthesis of CuK-silsesquioxane/8-hydroxyquinoline complex 5. Solvated molecules are omitted for clarity.
Scheme 3. General scheme of the synthesis of CuK-silsesquioxane/8-hydroxyquinoline complex 5. Solvated molecules are omitted for clarity.
Molecules 27 06205 sch003
Figure 4. The molecular structure (top) and crystal packing (bottom) of 5.
Figure 4. The molecular structure (top) and crystal packing (bottom) of 5.
Molecules 27 06205 g004
Figure 5. Top. Principle of the metallocene (potassium–phenyl) linkage observed in 5. Bottom. A fragment of the 2D coordination polymer structure of 5.
Figure 5. Top. Principle of the metallocene (potassium–phenyl) linkage observed in 5. Bottom. A fragment of the 2D coordination polymer structure of 5.
Molecules 27 06205 g005
Scheme 4. General scheme of the synthesis of CuLi- (6), CuNa- (7), and CuK-based (8) silsesquioxane complexes with 5,7-dibromo-8-hydroxyquinoline ligands. Solvated molecules are omitted for clarity.
Scheme 4. General scheme of the synthesis of CuLi- (6), CuNa- (7), and CuK-based (8) silsesquioxane complexes with 5,7-dibromo-8-hydroxyquinoline ligands. Solvated molecules are omitted for clarity.
Molecules 27 06205 sch004
Figure 6. Top. The molecular structure of 6. Middle. The molecular structure of 7. Bottom. The molecular structure of 8.
Figure 6. Top. The molecular structure of 6. Middle. The molecular structure of 7. Bottom. The molecular structure of 8.
Molecules 27 06205 g006aMolecules 27 06205 g006b
Figure 7. Top. A fragment of the 1D coordination polymer structure of 7. Bottom. A fragment of the 1D coordination polymer structure of 8.
Figure 7. Top. A fragment of the 1D coordination polymer structure of 7. Bottom. A fragment of the 1D coordination polymer structure of 8.
Molecules 27 06205 g007
Scheme 5. General scheme of the synthesis of CuK-silsesquioxane/5,7-diiodo-8-hydroxyquinoline ligand complex 9. Solvated molecules are omitted for clarity.
Scheme 5. General scheme of the synthesis of CuK-silsesquioxane/5,7-diiodo-8-hydroxyquinoline ligand complex 9. Solvated molecules are omitted for clarity.
Molecules 27 06205 sch005
Figure 8. The molecular structure of 9.
Figure 8. The molecular structure of 9.
Molecules 27 06205 g008
Figure 9. The 3D coordination polymer structure of 9.
Figure 9. The 3D coordination polymer structure of 9.
Molecules 27 06205 g009
Figure 10. Oxidation of cyclohexane (0.46 M) with hydrogen peroxide (50% aqueous; 2.0 M) catalyzed by complex 2 (concentration 5 × 10−4 M) in the presence of nitric acid (0.05 M) in acetonitrile at 50 °C. Concentrations of cyclohexanol (curve 1), cyclohexanone (curve 2), and sum cyclohexanol + cyclohexanone (curve 3) were determined after the reduction of the aliquots with solid PPh3. Concentrations of cyclohexanol (curve 1a) and cyclohexanone (curve 2a) in the absence of nitric acid in the analogous experiment. For the comparison, the accumulation of cyclohexanol + cyclohexanone in the oxidation of cyclohexane with tert-BuOOH under the same conditions is represented by curve 4.
Figure 10. Oxidation of cyclohexane (0.46 M) with hydrogen peroxide (50% aqueous; 2.0 M) catalyzed by complex 2 (concentration 5 × 10−4 M) in the presence of nitric acid (0.05 M) in acetonitrile at 50 °C. Concentrations of cyclohexanol (curve 1), cyclohexanone (curve 2), and sum cyclohexanol + cyclohexanone (curve 3) were determined after the reduction of the aliquots with solid PPh3. Concentrations of cyclohexanol (curve 1a) and cyclohexanone (curve 2a) in the absence of nitric acid in the analogous experiment. For the comparison, the accumulation of cyclohexanol + cyclohexanone in the oxidation of cyclohexane with tert-BuOOH under the same conditions is represented by curve 4.
Molecules 27 06205 g010
Figure 11. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 2 on the initial concentration of cyclohexane (graph A). Linearization of the curve from graph A in coordinates W0−1 − 1/[C6H12] is shown in graph B. Conditions. Concentrations: [1] = 5 × 10−4 M, [H2O2]0 = 2.0 M, [HNO3] = 0.05 M, in acetonitrile at 50 °C. Concentrations of cyclohexanone and cyclohexanol were determined after the reduction of the aliquots with solid PPh3.
Figure 11. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 2 on the initial concentration of cyclohexane (graph A). Linearization of the curve from graph A in coordinates W0−1 − 1/[C6H12] is shown in graph B. Conditions. Concentrations: [1] = 5 × 10−4 M, [H2O2]0 = 2.0 M, [HNO3] = 0.05 M, in acetonitrile at 50 °C. Concentrations of cyclohexanone and cyclohexanol were determined after the reduction of the aliquots with solid PPh3.
Molecules 27 06205 g011
Figure 12. Accumulation of 2-heptanone (graph A), cyclohexanone (graph B), and acetophenone (graph C), (curve 1), in the oxidation of 2-heptanol (0.42 M), cyclohexanol (0.5 M), and 1-phenylethanol (0.5 M), respectively, with tert-butyl hydroperoxide (1.5 M) catalyzed by complex 2 (5 × 10−4 M) at 50 °C in acetonitrile. Accumulation of acetophenone (graph C, curve 2) in the oxidation of 1-phenylethanol (0.5 M) with H2O2 (2,0 M) catalyzed by complex 2 (5 × 10−4 M) at 50 °C in acetonitrile. To quench the oxidation process, concentrations of products were measured by GC only after reducing the reaction sample with solid PPh3.
Figure 12. Accumulation of 2-heptanone (graph A), cyclohexanone (graph B), and acetophenone (graph C), (curve 1), in the oxidation of 2-heptanol (0.42 M), cyclohexanol (0.5 M), and 1-phenylethanol (0.5 M), respectively, with tert-butyl hydroperoxide (1.5 M) catalyzed by complex 2 (5 × 10−4 M) at 50 °C in acetonitrile. Accumulation of acetophenone (graph C, curve 2) in the oxidation of 1-phenylethanol (0.5 M) with H2O2 (2,0 M) catalyzed by complex 2 (5 × 10−4 M) at 50 °C in acetonitrile. To quench the oxidation process, concentrations of products were measured by GC only after reducing the reaction sample with solid PPh3.
Molecules 27 06205 g012
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.; 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. https://doi.org/10.3390/molecules27196205

AMA Style

Bilyachenko AN, Khrustalev VN, Zueva AY, Titova EM, Astakhov GS, Zubavichus YV, Dorovatovskii PV, Korlyukov AA, Shul’pina LS, Shubina ES, 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(19):6205. https://doi.org/10.3390/molecules27196205

Chicago/Turabian Style

Bilyachenko, Alexey N., Victor N. Khrustalev, Anna Y. Zueva, Ekaterina M. Titova, Grigorii S. Astakhov, Yan V. Zubavichus, Pavel V. Dorovatovskii, Alexander A. Korlyukov, Lidia S. Shul’pina, Elena S. Shubina, and et al. 2022. "A Novel Family of Cage-like (CuLi, CuNa, CuK)-phenylsilsesquioxane Complexes with 8-Hydroxyquinoline Ligands: Synthesis, Structure, and Catalytic Activity" Molecules 27, no. 19: 6205. https://doi.org/10.3390/molecules27196205

Article Metrics

Back to TopTop