Next Article in Journal
Per- and Poly-Fluoroalkyl Substances in Portuguese Rivers: Spatial-Temporal Monitoring
Next Article in Special Issue
Recognition Site Modifiable Macrocycle: Synthesis, Functional Group Variation and Structural Inspection
Previous Article in Journal
Alvespimycin Inhibits Heat Shock Protein 90 and Overcomes Imatinib Resistance in Chronic Myeloid Leukemia Cell Lines
Previous Article in Special Issue
Comparison of Photophysical Properties of Lanthanide(III) Complexes of DTTA- or DO3A-Appended Aryl-2,2′-Bipyridines
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Cagelike Octacopper Methylsilsesquioxanes: Self-Assembly in the Focus of Alkaline Metal Ion Influence—Synthesis, Structure, and Catalytic Activity

by
Alexey N. Bilyachenko
1,*,
Ivan S. Arteev
1,2,
Victor N. Khrustalev
3,4,
Anna Y. Zueva
1,3,
Lidia S. Shul’pina
1,
Elena S. Shubina
1,
Nikolay S. Ikonnikov
1 and
Georgiy B. Shul’pin
5,6,*
1
A. N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, Vavilov Str. 28, 119991 Moscow, Russia
2
Higher Chemical College, Mendeleev University of Chemical Technology of Russia, Miusskaya Sq. 9, 125047 Moscow, Russia
3
Research Institute of Chemistry, Peoples’ Friendship University of Russia (RUDN University), Miklukho-Maklay Str. 6, 117198 Moscow, Russia
4
Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences (RAS), Leninsky Prospect 47, 119991 Moscow, Russia
5
Semenov Federal Research Center for Chemical Physics, Russian Academy of Sciences, Ulitsa Kosygina 4, 119991 Moscow, Russia
6
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 2023, 28(3), 1211; https://doi.org/10.3390/molecules28031211
Submission received: 27 December 2022 / Revised: 21 January 2023 / Accepted: 24 January 2023 / Published: 26 January 2023
(This article belongs to the Special Issue Design and Synthesis of Macrocyclic Compounds)

Abstract

:
A family of unusual octacopper cage methylsilsesquioxanes 14 were prepared and characterized. Features of their cagelike (prismatic) structure were established using X-ray diffraction studies. Effects of distortion of prismatic cages 14 due to variation of (i) additional alkaline metal ions (K, Rb, or Cs), (ii) combination of solvating ligands, and (iii) nature of encapsulating species were found. Opportunities for the design of supramolecular 1D extended structures were found. These opportunities are based on (i) formate linkers between copper centers (in the case of Cu8K2-based compound 2) or (ii) crown ether-like contacts between cesium ions and siloxane cycles (in the case of Cu8Cs2-based compound 4). Cu8Cs2-complex 4 was evaluated in the catalysis of alkanes and alcohols. Complex 4 exhibits high catalytic activity. The yield of cyclohexane oxidation products is 35%. The presence of nitric acid is necessary as a co-catalyst. The oxidation of alcohols with the participation of complex 4 as a catalyst and tert-butyl hydroperoxide as an oxidizer also proceeds in high yields of up to 98%.

1. Introduction

Cagelike metallasilsesquioxanes (CLMSs) are a wide family of polynuclear metallacomplexes, providing a kaleidoscopic variety of molecular architectures [1,2,3,4,5,6,7,8,9,10,11] as well as impressive diversity of further applications. The latter comprises approaches to molecular magnets (including behaviors of single-molecule magnets or spin glass types) [12,13,14,15,16,17,18] and objects with photophysical properties (including the first example of a lanthanide-based luminescent thermometer) [19,20,21,22]. In the context of material science, CLMSs were successfully applied for the versatile design of functional MOFs and coordination polymers [23,24,25,26], anodes [27,28], biomedical agents [29], and fire retardants [30,31,32,33,34,35]. Numerous reviews summarize the effectiveness of CLMS in various catalyzed reactions, including oxidative C–H bond activations [36,37,38,39,40,41,42]. Among recent results we could mention activity in CO2 cycloaddition [43], Chan-Evans-Lam coupling [44], biomass transformations [45], and oxidative amidation [46]. Considering the significant potential of copper-based CLMSs as catalysts [47,48,49,50], we were interested in designing new types of these compounds. Numerous results pointed out the significant influence of the nature of the substituent at the silicon center on molecular geometry of arising CLMSs, e.g., for nickel-based (phenyl- [51,52,53,54] vs. 4-vinylbenzyl-substituted [55] CLMSs), titanium-based (t-Bu vs. (2,6-iPr2C6H3)N(SiMe3)-substituted [56,57] CLMSs), aluminum-based (cyclopentyl [58] vs. cyclohexyl-substituted [59] CLMSs) compounds. In turn, it is well known that the archetype of phenyl-substituted copper CLMSs is a hexanuclear prismatic structure [1,6,8,11,48,49]. Our newest results describing methyl-substituted Cu-CLMSs’ synthesis, structure determination, and evaluation in catalysis are presented below.

2. Results

2.1. Synthesis and Structure

It is known that variations of the nature of alkaline metal ions strongly influence the structural features of phenylsurrounded CLMSs, especially in the context of supramolecular aggregation observed for large-sized ions (K, Rb, Cs) [16,23,25,47,60]. For the method of synthesis of methylsurrounded Cu-CLMSs, a universal approach that included alkaline hydrolysis [61,62,63] of MeSi(OMe)3 assisted by the action of corresponding hydroxide MOH (M = Na, K, Rb, or Cs), was chosen. Then in situ formed alkaline metal siloxanolate [PhSi(O)OM]x species were brought into the exchange interaction with copper(II) chloride (Scheme 1). Characteristically in all cases arising CLMSs 14 adopted rare for metallasilsesquioxanes octacopper sandwich geometry (CCDC search revealed only two publications reported on the structure of Cu8 silsesquioxanes [64,65]). Noteworthy is that the presence of a large inner void in the sandwich structure of complexes reported in [64,65] favors its filling by different encapsulated species, i.e., pyrazine [64], and acetonitrile or sodium acetate [65]. According to these regulations, compounds [(MeSiO2)8]2Cu8(DMF)8 (EtOH)] • 2.75 (DMF) • 0.25 (Me2CO) 1 (isolated in 37% yield), {[(MeSiO2)8]2Cu8 (HCOO) (EtOH)4 K2 (MeCOO)] • 2 (EtOH)}n 2 (31% yield), [(MeSiO2)8]2Cu8 (DMF)8 Rb (MeO) (EtOH)] 3 (26% yield), {[(MeSiO2)8]2Cu8 Br2 (DMF)6 Cs2] • 2 (DMF)sq} 4 (42% yield) also feature the encapsulation phenomenon. More precisely, sodium-assisted synthesis (Scheme 1, compound 1), followed by the addition of dimethylformamide/acetone mixture as a crystallization facilitator, provided mixed-ligated complex 1. One could see (Figure 1) that while eight DMF molecules coordinate the copper centers of the compound, the ethanol molecule occupies the void of the cage structure. The top view of complex 1 clearly shows the non-regular octagon structure of a prism-like cage. This deviation is also obvious from the comparison of the Cu…Cu distances between opposite copper centers. All four possible Cu…Cu distances are different from each other and they vary in the range from 7.3781(7) Å to 7.6773(7) Å. Two factors potentially responsible for the distortion of the cage could be discussed: (i) the presence of the before mentioned encapsulated ethanol specie (its oxygen center coordinates only two of eight copper sites), and (ii) the variation of the location style of DMF molecules coordinating eight copper centers. Namely, while seven DMF ligands are located in the positions parallel to the basal plane of the copper ions, the eighth DMF molecule is arranged perpendicular to it. Recently, we reported on a similar effect (rotation of acetone solvates [48]) that causes a “shape switch effect” for the hexanuclear copperphenylsilsesquioxanes—from a regular form to an ellipsoid-like one. In the case of complex 1, due to the presence of the encapsulated component (which was not the case for Ref [48]), we could not state precisely if the distortion of the prismatic geometry of 1 is a consequence of DMFs rotation exclusively. Most probably, both factors are influential and act in parallel.
The self-assembly and formation of the Cu8-based geometry of CLMS 1 are easily reproduced. Additional runs of synthesis (in an alcohol-only reaction system) smoothly gave complex 1a [(MeSiO2)8]2Cu8(EtOH)7(MeOH)0.25] • EtOH (Figure 2) in 59% yield. A discussion of the structural features of 1a is given in Section 3.3.
Next, a similar synthesis as complex 1 has been performed for KOH-involved reaction conditions (Scheme 1). This reaction provided octacopper methylsilsesquioxane complex 2 (Figure 3), whose cage geometry at the first sight is a close analog of compound 1 (Figure 3 top). In point of fact, several deep differences between the two of these compounds could be drawn. First, the shape of the cage of 2 is much ellipsoid in comparison to 1—Cu…Cu distances in minor and major axes in ellipse fragment are equal to 6.8940(9) Å and 7.9608(7) Å, respectively. The other two opposite Cu…Cu distances are almost equal to each other (7.4115(9) Å and 7.4388(7) Å). Second, compound 2 features the presence of two potassium ions located in crown ether-like positions to siloxane cycles of silsesquioxane ligands. Each potassium center coordinates four oxygen atoms of the corresponding siloxane cycle. It is intriguing that the nature of the anions for potassium cations is different, some of them are carboxylate fragments that appeared as a consequence of the partial oxidation of alcohols present in the synthesis/crystallization system. This is in perfect accord with several earlier reports [49,65,66,67]. Precisely, the first type of anion in 2 is an encapsulated (in axial position) acetate fragment, with acetate oxygen atoms additionally coordinating four of eight copper centers. The next anion type in 2 is an external formate laying in the basal plane of the copper belt of the cage. Additionally, the formate ligand coordinates one of the copper centers. Symptomatically that formate species plays the role of a linker between neighboring cages causing an assembly of 1D coordination polymer (Figure 4). To the best of our knowledge, formate linkers have never been reported as the bridging components for CLMS’ coordination polymers, thus compound 2 shows new possibilities for the supramolecular chemistry of metallasilsesquioxanes. Finally, the last two copper ions in 2 are bonded to ethanol solvates.
Then, the shift to RbOHaq as a starting reagent (Scheme 1) allowed isolation of octacopper methylsilsesquioxane complex 3 (Figure 5). Compound 3, despite the same nuclearity, exhibits significant individuality in comparison to predecessors 1 and 2. Due to the larger size of the ion, rubidium center in 3 coordinates all eight oxygen atoms in the siloxane cycle of the silsesquioxane ligand. This causes the formation of a more regular prismatic type of cage in comparison to 2. The range between the shortest (7.3674(1) Å) and the longest (7.6497(9) Å) opposite Cu…Cu distance is just ~0.28 Å (for the ellipsoid cage of 2—~1.07 Å). The role of counterion for rubidium cation in complex 3 is played by the OMe group (located in the axial position). The oxygen atom of the methoxy group is laying in a basal plane of the copper belt and does not coordinate with any of the copper centers. In turn, each copper ion is ligated by an external DMF molecule. The screening effect of DMF solvates prevents the assembly of any supramolecular derivatives of 3 via its copper centers. The same could be said regarding screening ethanol solvate and coordinating rubidium ions. This also prevents the formation of the coordination polymer structure despite the significant coordinative potential of rubidium centers towards siloxanes, which could be fruitful in supramolecular assembly [25,68,69,70].
Finally, the shift to CsOHaq as a starting reagent (Scheme 1) allowed isolation of octacopper methylsilsesquioxane complex 4 (Figure 6). It is worthwhile to notice that we failed to crystallize the compound from general reaction conditions. Keeping in mind the recently reported high efficiency of exchange reactions with Et4NCl (or Ph4PCl) for the isolation of Ln-based CLMSs [14,20,22], we decided to use a similar reaction with Me4NBr here. To our surprise, instead of the formation of “cation-involved CLMS” (as in the case of Et4N+/Ph4P+ based compounds in Refs [14,20,22]), compound 4 is “anion-involved” CLMS bearing two bromide groups. It is intriguing that bromides are completely different in their role in the composition of 4. Namely, when the first of them is located in an external position to the cage and connected to copper ion, the second bromide is encapsulated into the inner void of the cage. The presence of additional bromide as carriers of two negative charges requires charge-balancers. These latter (cesium ions) are found in crown ether positions by siloxane cycles of silsesquioxane ligands and bring a similar effect as the rubidium ion in complex 3. Namely, cesium centers, owing to their large ion size, coordinate all eight oxygen centers in the siloxane cycle. Keeping this fact in mind, as well as the presence of DMF molecules as the only solvates in the composition of 4, one could expect a regular shape type of cage fragment of 4. In fact, the difference between the shortest (7.2469(2) Å) and the longest (7.7244(2) Å) distances between opposite copper ions is significantly larger than the before mentioned value for 3—~0.48 Å. This points to the irregularity of structure 4. Indeed, the presence of one Cu-Br fragment and one copper center without any solvate surrounding shows the inequality of copper sites. It is important to note that the copper atom free from an external ligand is linked to the encapsulated bromide anion at a distance of 3.069(3) Å. Each of the remaining six copper ions is ligated by a DMF molecule. This is especially intriguing because an absence of any screening solvates at cesium centers allows the assembly of an unusual supramolecular structure of 4. 1D coordination polymer 4 (Figure 7) is realized due to mutual Cs…O contacts between the cesium ion of one cage and the oxygen atom from the siloxane cycle of the neighboring cage. Thus, the rectangular linking unit Cs2O2 is connected to two cages. Symptomatically, this unit is repeated in a diagonal manner when the next cage fragment joins, giving non-linear supramolecular geometry.
Remarkably, the detailed structural analysis of the complexes studied here allows revealing the characteristic criterion of the template effect of encapsulated anions/molecules, namely, the copper atoms bound to the template anions/molecules remain externally “naked” (see complexes 2 and 4). The “naked” environment of the copper(II) cations is absolutely unusual taking into account the favorable [4+2]-coordination geometry for it due to the Jahn–Teller effect. Consequently, in contrast to the acetate anion in 2 and bromide anion in 4, the encapsulated ethanol molecule in 1 and the methoxy anion in 3 are not template agents. This statement is in perfect agreement with the Cu8- and Cu9-cage-like complexes studied by us previously (with the template agents of the acetate anion and DMSO molecule, respectively) [65,66].

2.2. Catalytic Studies

We discovered that complex 4 exhibits high catalytic activity in the oxidation of saturated hydrocarbons. Accumulations of the products are presented in Figure 8, Figure 9, Figure 10 and Figure 11. Oxidation of cyclohexane with hydrogen peroxide according to gas chromatography results in the formation of a mixture of cyclohexanol and cyclohexanone. Adding triphenylphosphine to the reaction mixture according to the method proposed previously by one of us [71,72], leads to a sharp increase in the concentration of cyclohexanol compared to the concentration of cyclohexanone, as shown in Figure 8B. (The ratio of cyclohexanone to cyclohexanol after adding triphenylphosphine, for example, is about 0.04 (0.007 M/0.142 M after 60 min). At the same time in the absence of phosphine, the ratio of cyclohexanone to cyclohexanol for complex 4 is about 0.7 (0.042 M/0.059 M after 60 min) Figure 8A. The effect of adding triphenylphosphine to a sample of the reaction mixture can be explained by the fact that the main product of the oxidation reaction is alkyl hydroperoxide. In the absence of phosphine, it decomposes to form a ketone and alcohol in comparable amounts in an injector of GC. Triphenylphosphine reduces an alkyl hydroperoxide to alcohol, so we see a sharp increase in the amount of alcohol on the chromatogram. The total yield of alcohol and ketone in the oxidation of cyclohexane catalyzed by complex 4 after 1.5 h is 35%, and TON is 320. To carry out the oxidation reaction of alkanes, it is necessary to add nitric acid. When carrying out reactions in the absence of nitric acid, only the decomposition of hydrogen peroxide (catalase activity) is observed and the yield of oxidation products is not more than 0.5% in two hours. In this case, rapid precipitation of a brown precipitate is observed, apparently, the decomposition of the complex occurs.
We investigated the oxidation of n-heptane with hydrogen peroxide catalyzed by complex 4. The results are given below (initial concentrations: n-heptane 0.5 M, H2O2 1.5 M, catalyst 5 × 10−4 M, HNO3 0.05 M, CH3CN up to 2.5 mL). After two hours, the following concentrations of (M) alcohol isomers were obtained (after reduction with triphenylphosphine maximum yield of products in the oxidation of n-heptane 10% was obtained): C (1) 0.0003M; C (2) 0.006; C (3) 0.006M: C (4) 0.003M. These values give a series of selectivity: C (1):C (2):C (3):C (4) = 1.0:5.6:5.6:5.6. The selectivity parameters for the oxidation of methylcyclohexane were obtained also: 1°:2°:3° = 1.0:4.8:14.2 (after reduction with triphenylphosphine maximum yield of products in the oxidation of methylcyclohexane 12% was obtained). However, the selectivity parameters obtained for n-heptane are slightly higher than the typical selectivity parameters for the oxidation of n-heptane with the participation of hydroxyl radicals [73,74,75]. We see two possible reasons for these differences. First, it is possible that, along with the hydroxyl radical, another oxidizing species is generated in the system, which exhibits greater regioselectivity than OH. This can be, in particular, the Cu(III) ion [76]. Secondly, it is possible that the differences in regioselectivity and reactivity are associated with steric hindrances in the access of an alkane to the reaction center near which hydroxyl radicals are generated. Consideration of the regio-and bond-selectivity in the oxidation of n-heptane and methylcyclohexane and the dependences of initial rates of oxygenates formation (Figure 9 and Figure 10) indicates that the oxidation proceeds in the main with the participation of free hydroxyl radicals.

2.3. Oxidation of Alcohols

Aliphatic and aromatic alcohols are oxidized with hydrogen peroxide with very low efficiency, however to the use of tert-butyl hydroperoxide, they give ketones as oxidation products in good to high yields. It should be noted that these reactions do not require the addition of nitric acid. The accumulation of products in such reactions is shown in Figure 11. We would like to mention that the principle of catalytic activity of complex 4 is a precatalytic one. In homogeneous conditions (and in the presence of peroxide and acid cocatalyst) complex 4 modifies into catalytic active species, which further provide efficient oxidations. The advantage of metallasilsesquioxane precatalysts is a high activity of catalysis; these cage catalysts show superior activity in comparison with the simple copper salts and complexes [77].

3. General Experimental Considerations

All reagents were purchased from the usual suppliers (Sigma, Fluka) and used without further purification. RbOH and CsOH were used as 50% w/w water solutions. Elemental analyses were carried out with an XRF spectrometer VRA-30 and Eurovector EA-3000 analyzer. IR spectra of the compounds (KBr pellets, Figure 12, see below) 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. Set of signals: 1650–1550 cm−1 (νC=N), 1420–1380 cm−1stretchC–Si), 1120 cm−1asSi–O–Si), 1050 cm−1sSi–O–Si), 960 cm−1asSi–O in Si–O–M fragment), 820 cm−1 and 760 cm−1wagC–Si). UV-Vis spectra (10 mm optical path length, acetone solutions) were recorded on a Cary 50 spectrophotometer. Figure 13 is to exemplify a general view of spectra (for complex 4, see below).

3.1. Syntheses of 14

In a typical procedure, 1.5 g (11 mmol) of MeSi(OMe3)3 and 11.7 mmol of corresponding MOH (M = Na, K, Rb, or Cs, see below for details) were heated at reflux in 40 mL of ethanol for 1.5 h. Then, 0.73 g (5.43 mmol) of CuCl2 was added and the resulting mixture was heated at reflux for 3 h. Then the solution was stirred without heating for 24 h, followed by centrifugation of the precipitate.
For compound [(MeSiO2)8]2Cu8(DMF)8 (EtOH)] • 2.75 (DMF) • 0.25 (Me2CO) 1: Mass of MOH (NaOH) = 0.47 g. Crystallization of filtrate (after the addition of 20 mL of DMF/acetone (1:1, v:v) gave a crystalline material in a week, including single crystals that were used for X-ray diffraction and elemental analyses. Anal. Calcd for [(CH3SiO2)8]2Cu8(C3H7NO)10.25(C2H5OH)(C3H6O)0.25: C, 23.59; H, 5.09; Cu, 20.17; N, 5.70; Si, 17.83. Found: C, 22.14; H, 4.91; Cu, 20.29; N, 5.61; Si, 17.91. The rest of crystalline material was dried in vacuum. Yield: 0.28 g (37%).
For compound [(MeSiO2)8]2Cu8(EtOH)7(MeOH)0.25] • EtOH 1a: Mass of MOH (NaOH) = 0.47 g. Crystallization of filtrate (after the addition of 10 mL of methanol) gave a crystalline material in 3–4 days, including single crystals that were used for X-ray diffraction and elemental analyses. Anal. Calcd for [(CH3SiO2)8]2Cu8(CH3OH)0.25(C2H5OH)8: C, 18.56; H, 4.69; Cu, 24.36; Si, 21.53. Found: C, 18.41; H, 4.58; Cu, 24.50; O, 30.97; Si, 21.60. The rest of crystalline material was dried in vacuum. Yield: 0.45 g (59%).
For compound {[(MeSiO2)8]2Cu8(HCOO)(EtOH)4K2(MeCOO)] • 2 (EtOH)}n 2: Mass of MOH (KOH) = 0.66 g. Crystallization of filtrate gave a crystalline material in a week, including single crystals that were used for X-ray diffraction and elemental analyses. Anal. Calcd for [(CH3SiO2)8]2Cu8(HCOO)(C2H5OH)6K2(CH3COO): C, 17.17; H, 4.09; Cu, 23.44; K, 3.61; Si, 20.72. Found: C, 17.01; H, 3.95; Cu, 23.53; K, 3.69; Si, 20.81. The rest of crystalline material was dried in vacuum. Yield: 0.35 g (31%).
For compound [(MeSiO2)8]2Cu8 (DMF)8 Rb (MeO) (EtOH)] 3: Mass of MOH (RbOH) = 1.20 g. Crystallization of filtrate (after the addition of 15 mL of DMF) gave a crystalline material in a week, including single crystals that were used for X-ray diffraction and elemental analyses. Anal. Calcd for [(CH3SiO2)8]2Cu8(C3H7NO)8(C2H5OH)Rb(CH3O): C, 21.01; H, 4.63; Cu, 20.69; N, 4.56; Rb, 3.48; Si, 18.28. Found: C, 20.69; H, 4.50; Cu, 20.77; N, 4.41; Rb, 3.59; Si, 18.37. The rest of crystalline material was dried in vacuum. Yield: 0.28 g (26%).
For compound {[(MeSiO2)8]2Cu8 Br2 (DMF)6 Cs2] • 2 (DMF)sq} 4: Mass of MOH (CsOH) = 1.75 g. Crystallization of filtrate was found unsuccessful and additional exchange reaction with Me4NBr (0.42 g, 2.75 mmol) was performed. Mixture was stirred without heating for 24 h, followed by centrifugation of precipitate. Resulting filtrate (after the addition of 15 mL of DMF) gave a crystalline material in a week, including single crystals that were used for X-ray diffraction analysis. Anal. Calcd for [(CH3SiO2)8]2Cu8(C3H7NO)8Cs2Br2: C, 17.66; H, 3.85; Br, 5.87; Cs, 9.77; Cu, 18.69; N, 4.12; Si, 16.52. Found: C, 17.52; H, 3.79; Br, 5.92; Cs, 9.82; Cu, 18.77; N, 4.05; Si, 16.60. The rest of crystalline material was dried in vacuum. Yield: 0.54 g (42%).

3.2. X-ray Crystal Structure Determination

The single-crystal X-ray diffraction study of 1–4 was carried out on a four-circle Rigaku Synergy S diffractometer equipped with a HyPix6000HE area-detector (T = 100 K, graphite monochromator, shutterless ω-scanning mode). The data were integrated and corrected for absorption by the CrysAlisPro program [78]. The single-crystal X-ray diffraction study of 1a was carried out on the ‘RSA’ beamline (T = 100 K, λ = 0.75270 Å) of the Kurchatov Synchrotron Radiation Source using a Rayonix SX165 detector. In total, 720 frames were collected with an oscillation range of 1.0° in the φ scanning mode using two different orientations for the crystal. The semi-empirical correction for absorption was applied using the Scala program [79]. The data were indexed and integrated using the utility iMOSFLM from the CCP4 software suite [80,81]. For details, see Table 1.
The structures were solved by intrinsic phasing modification of direct methods [82], and refined by a full-matrix least-squares technique on F2 with anisotropic displacement parameters for all non-hydrogen atoms. In the case of 4, all attempts to model and refine positions of the solvate DMF molecules were unsuccessful. Therefore, their contribution to the total scattering pattern was removed by use of the utility SQUEEZE in PLATON15 [83]. The hydrogen atoms of the hydroxyl groups were objectively localized in the difference-Fourier maps and included in the refinement within the riding model with fixed isotropic displacement parameters [Uiso(H) = 1.5Ueq(O)]. 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 [84,85].
Crystallographic data have been deposited with the Cambridge Crystallographic Data Center, CCDC 2232957 (1), CCDC 2235882 (1a), 2232958 (2), 2232959 (3), and 2232960 (4). 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: deposit@ccdc.cam.ac.uk or www.ccdc.cam.ac.uk).

3.3. Structural Features of Complex 1a

Similarly to complex 1, the inner void of cage structure 1a is occupied by ethanol molecules. The rest of the alcohol-solvating ligands (seven ethanol and one methanol) occupy external positions at the basal plane of complex 1a and coordinate copper centers. These differences in the nature of solvating ligands cause some transformations in non-regular octagon cages 1 and 1a. For example, distances between opposite copper centers in 1 vary between 7.3781(7) Å and 7.6773(7) Å. In turn, for compound 1a this deviation is larger (between 7.3789(3) Å and 7.8144(3) Å).

3.4. Catalytic Studies

All reactions were carried out in air in thermostatically controlled cylindrical glass vessels with vigorous stirring. The total volume of the reaction solution was 5 mL (WARNING: the combination of air or molecular oxygen and H2O2 with organic compounds can be explosive at elevated temperatures!). Initially, a portion of a 50% aqueous solution of hydrogen peroxide or tert-butyl hydroperoxide (70%) (for alcohols) was added to a solution of the catalyst (added in dry form), HNO3 (from the stock solution) and the substrate in acetonitrile). Attribution of peaks was made by comparison with chromatograms of authentic samples). Usually, samples were analyzed twice, i.e., before and after the addition portion by portion of the excess of solid PPh3. Aliquots of the reaction solution were analyzed by GC (CH3NO2 (250 μL) was used as an internal standard) (the instrument chromatograph—Cromas, FFAP/OV-101 20/80 w/w fused silica capillary column, 30 m × 0.2 mm × 0.3μm, argon as a carrier gas). The reactions of alkanes and alcohols were stopped by cooling and were usually analyzed twice, i.e., before and after addition of excess solid PPh3.

4. Conclusions

A series of octacopper cage methylsilsesquioxanes 14 were prepared and characterized. The strong influence of steric factors (nature of the substituent at silicon atom) on the self-assembly of coppersilsesquioxanes was found, keeping in mind that the archetype structure of phenylsilsesquioxane derivatives is a hexacopper cage. A change of geometry of prismatic cages 14 caused by the variation of their composition (different alkaline metal ions, solvates, or encapsulated species) was detected. It revealed the characteristic criterion of the template effect of encapsulated anions/molecules for the cage-like copper complexes. Noteworthy is that the large size of the inner void of octacopper silsesquioxanes points to the significant potential of further studies of high nuclear metallasilsesquioxanes in the context of potentially intriguing encapsulation effects. The same concerns observed in different types of intercage connectivity provided 1D coordination polymers for compound 2 (formate linkers) and 4 (crown ether contacts). Noteworthy is that carboxylate bridges in structures of cage metallasilsesquioxane coordination polymers were never reported before this work. This points to many prospects for supramolecular chemistry involving metallasilsesquioxanes. Complex 4 exhibits high catalytic activity in the oxidation of alkanes (cyclohexane, n-heptane, and methylcyclohexane) with hydrogen peroxide and alcohols (phenylethanol, cyclohexanol, 2-heptanol, and 2-hexanol) with tert-butyl hydroperoxide. Measurement of n-heptane regioselectivity and bond selectivity in the reactions of methylcyclohexane with H2O2 allows us to conclude that hydroxyl radicals play an important role in these oxidation reactions. It is possible that, along with the hydroxyl radical, another oxidizing species (it may be Cu(III)-ion) is generated in the system. The starting products in reactions with alkanes are alkyl hydroperoxides, which are reduced to the corresponding alcohols with an excess of triphenylphosphine (PPh3).

Author Contributions

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

Funding

This work (synthetic and catalytic studies) was supported by the Ministry of Science and Higher Education of the Russian Federation (Contract/agreement No. 075-00697-22-00) and was performed employing the equipment of Center for molecular composition studies of INEOS RAS. Funding for X-ray diffraction studies was provided by the Ministry of Education and Science of the Russian Federation (award no. 075-03-2020-223 (FSSF-2020-0017)). This work was also performed within the framework of the Program of Fundamental Research of the Russian Federation. Reg. No. 122040500068-0.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in the article.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

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]
  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]
  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]
  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]
  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. 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]
  31. 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]
  32. 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]
  33. 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]
  34. 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]
  35. 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]
  36. Abbenhuis, H.C.L. Advances in Homogeneous and Heterogeneous Catalysis with Metal-Containing Silsesquioxanes. Chem. Eur. J. 2000, 6, 25–32. [Google Scholar] [CrossRef]
  37. 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]
  38. 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]
  39. 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]
  40. 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]
  41. 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]
  42. 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]
  43. 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]
  44. 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]
  45. 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]
  46. 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]
  47. 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] [PubMed]
  48. 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]
  49. 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]
  50. 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]
  51. Levitsky, M.M.; Schegolikhina, O.I.; Zhdanov, A.A.; Igonin, V.A.; Ovchinnikov, Y.E.; Shklover, V.E.; Struchkov, Y.T. An unusual sandwich-type nickel oxide-siloxanolate complex {[PhSiO]64-O)23-O)4[Ni83-O)2](μ3-O)44-O)2[PhSiO]6} • 14n-BuOH • 10H2O • 2Me2CO. Synthesis and crystal structure. J. Organomet. Chem. 1991, 401, 199–210. [Google Scholar] [CrossRef]
  52. Bilyachenko, A.N.; Yalymov, A.I.; Shul’pina, L.S.; Mandelli, D.; Korlyukov, A.A.; Vologzhanina, A.V.; Es’kova, M.A.; Shubina, E.S.; Levitsky, M.M.; Shul’pin, G.B. Novel Cage-Like Hexanuclear Nickel(II) Silsesquioxane. Synthesis, Structure, and Catalytic Activity in Oxidations with Peroxides. Molecules 2016, 21, 665. [Google Scholar] [CrossRef] [Green Version]
  53. Bilyachenko, A.N.; Yalymov, A.I.; Korlyukov, A.A.; Long, J.; Larionova, J.; Guari, Y.; Vologzhanina, A.V.; Es’kova, 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]
  54. 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 2021, 517, 120160. [Google Scholar] [CrossRef]
  55. Korlyukov, A.A.; Eskova, M.A.; Tkachenko, I.M.; Kononevich, Y.N.; Shchegolikhina, O.I.; Muzafarov, A.M. Heteroligand nickel siloxane with 4-vinylbenzyl substituents. Mendeleev Commun. 2015, 25, 226–228. [Google Scholar] [CrossRef]
  56. Murugavel, R.; Bhattacharjee, M.; Roesky, H.W. Review Organosilanetriols: Model Compounds and Potential Precursors for Metal-containing Silicate Assemblies. Appl. Organometal. Chem. 1999, 13, 227–243. [Google Scholar] [CrossRef]
  57. Jutzi, P.; Lindemann, H.M.; Nolte, J.-O.; Schneider, M. Metallasilsesquioxanes–Synthetic and Structural Studies. Synthesis, Structure, and Reactivity of Novel Oligomeric Titanasiloxanes In Silicon Chemistry: From the Atom to Extended Systems; Jutzi, P., Schubert, U., Eds.; Wiley: Darmstadt, Germany, 2003; pp. 372–382. [Google Scholar]
  58. Skowronska-Ptasinska, M.; Duchateau, R.; van Santen, R.; Yap, G. Steric Factors Affecting the Brønsted Acidity of Aluminosilsesquioxanes. Eur. J. Inorg. Chem. 2001, 2001, 133–137. [Google Scholar] [CrossRef]
  59. Edelmann, F.T.; Gun’ko, Y.K.; Giessmann, S.; Olbrich, F.; Jacob, K. Silsesquioxane Chemistry:  Synthesis and Structure of the Novel Anionic Aluminosilsesquioxane [HNEt3][{Cy7Si7O9(OSiMe3)O2}2Al]·C6H14 (Cy = c-C6H11). Inorg. Chem. 1999, 38, 210–2114. [Google Scholar] [CrossRef]
  60. 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]
  61. 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]
  62. 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]
  63. Siripanich, P.; Bureerug, T.; Chanmungkalakul, S.; Sukwattanasinitt, M.; Ervithayasuporn, V. Mono and Dumbbell Silsesquioxane Cages as Dual-Response Fluorescent Chemosensors for Fluoride and Polycyclic Aromatic Hydrocarbons. Organometallics 2022, 41, 201–210. [Google Scholar] [CrossRef]
  64. Zherlitsyna, L.; Auner, N.; Bolte, M. Bis(μ6-cis2,4,6,8,10,12,14,16-octamethylcyclooctasiloxane-2,4,6,8,10,12,14,16- octolato)octa-kis[(dimethylformamide)copper(II)] dimethylformamide solvate enclosing a pyrazine molecule. Acta Crystallogr. Sect. C Cryst. Struct. Commun. 2006, 62, m199–m200. [Google Scholar] [CrossRef] [PubMed]
  65. 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]
  66. 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] [PubMed]
  67. 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]
  68. Dankert, F.; von Hänisch, C. Siloxane coordination revisited: Si−O bond character, reactivity and magnificent molecular shapes. Eur. J. Inorg. Chem. 2021, 2021, 2907–2927. [Google Scholar] [CrossRef]
  69. Jost, M.; Richter, R.-M.; Balmer, M.; Peters, B.; Dankert, F.; von Hänisch, C. Coordination polymers of alkali metal cyclosiloxazanides with one- and two-dimensional structures. Dalton Trans 2020, 49, 5787–5790. [Google Scholar] [CrossRef]
  70. Dankert, F.; Erlemeier, L.; Ritter, C.; von Hänisch, C. On the molecular architectures of siloxane coordination compounds: (re-)investigating the coordination of the cyclodimethylsiloxanes Dn (n = 5−8) towards alkali metal ions. Inorg. Chem. Front 2020, 7, 2138–2153. [Google Scholar] [CrossRef]
  71. 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]
  72. 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]
  73. 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]
  74. Nesterov, D.S.; Chygorin, E.N.; Kokozay, V.N.; Bon, V.V.; Boca, R.; Kozlov, Y.N.; Shul’pina, L.S.; Jezierska, J.; Ozarowski, A.; Pombeiro, A.J.L.; et al. Heterometallic CoIII4FeIII2 Schiff Base Complex: Structure, Electron Paramagnetic Resonance, and Alkane Oxidation Catalytic Activity. Inorg. Chem. 2012, 51, 9110–9122. [Google Scholar] [CrossRef]
  75. 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]
  76. 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]
  77. 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. [Google Scholar] [CrossRef]
  78. Rigaku. CrysAlisPro Software System, v. 1.171.41.106a; Rigaku Oxford Diffraction: Oxford, UK, 2021.
  79. Evans, P. Scaling and Assessment of Data Quality. Acta Crystallogr. 2006, D62, 72–82. [Google Scholar] [CrossRef]
  80. 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]
  81. 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. iMOSFLM: A new graphical interface for diffraction-image processing with MOSFLM. Acta Crystallogr. 2011, D67, 235–242. [Google Scholar]
  82. Sheldrick, G.M. SHELXT-Integrated Space-Group and Crystal-Structure Determination. Acta Crystallogr. 2015, A71, 3–8. [Google Scholar] [CrossRef] [Green Version]
  83. Spek, A.L. PLATON: A Multipurpose Crystallographic Tool; Utrecht University: Utrecht, The Netherlands, 2015. [Google Scholar]
  84. Sheldrick, G.M. A Short History of SHELX. Acta Crystallogr. 2008, A64, 112–122. [Google Scholar] [CrossRef] [Green Version]
  85. Sheldrick, G.M. Crystal Structure Refinement with SHELXL. Acta Crystallogr. 2015, C71, 3–8. [Google Scholar]
Scheme 1. General scheme of synthesis of cagelike octacopper methylsilsesquioxanes 14.
Scheme 1. General scheme of synthesis of cagelike octacopper methylsilsesquioxanes 14.
Molecules 28 01211 sch001
Figure 1. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8(DMF)8 (EtOH)] • 2.75 (DMF) • 0.25 (Me2CO) 1; the solvate dimethylformamide and acetone molecules are not shown. Hydrogen atoms are presented as green dots (the same concerns Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6).
Figure 1. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8(DMF)8 (EtOH)] • 2.75 (DMF) • 0.25 (Me2CO) 1; the solvate dimethylformamide and acetone molecules are not shown. Hydrogen atoms are presented as green dots (the same concerns Figure 2, Figure 3, Figure 4, Figure 5 and Figure 6).
Molecules 28 01211 g001aMolecules 28 01211 g001b
Figure 2. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8(EtOH)7(MeOH)0.25] • EtOH 1a.
Figure 2. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8(EtOH)7(MeOH)0.25] • EtOH 1a.
Molecules 28 01211 g002aMolecules 28 01211 g002b
Figure 3. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8 (HCOO) (EtOH)4 K2 (MeCOO)] • 2 (EtOH)}n 2; the solvate ethanol molecules are not shown.
Figure 3. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8 (HCOO) (EtOH)4 K2 (MeCOO)] • 2 (EtOH)}n 2; the solvate ethanol molecules are not shown.
Molecules 28 01211 g003
Figure 4. 1D coordination polymer structure in structure of 2; linked via formate bridges.
Figure 4. 1D coordination polymer structure in structure of 2; linked via formate bridges.
Molecules 28 01211 g004
Figure 5. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8 (DMF)8 Rb (MeO) (EtOH)] 3.
Figure 5. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8 (DMF)8 Rb (MeO) (EtOH)] 3.
Molecules 28 01211 g005
Figure 6. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8 Br2 (DMF)6 Cs2] • 2 (DMF)sq 4.; the solvate DMF molecules are not shown.
Figure 6. Two projections of molecular structure of complex [(MeSiO2)8]2Cu8 Br2 (DMF)6 Cs2] • 2 (DMF)sq 4.; the solvate DMF molecules are not shown.
Molecules 28 01211 g006aMolecules 28 01211 g006b
Figure 7. 1D coordination polymer structure of 4 via crown ether contacts of cesium and oxygen atoms of neighboring cages in crystal packing.
Figure 7. 1D coordination polymer structure of 4 via crown ether contacts of cesium and oxygen atoms of neighboring cages in crystal packing.
Molecules 28 01211 g007
Figure 8. Accumulation of cyclohexanol and cyclohexanone in oxidation of cyclohexane (0.46 M) with hydrogen peroxide (2.0 M, 50% aqueous) catalyzed by compound 4 (5 × 10−4 M) in the presence of HNO3 (0.05 M) in MeCN at 50 °C. Concentrations of cyclohexanone and cyclohexanol were determined before reduction of the PPh3 (Graph A), and the same reaction after reduction of the aliquots with solid PPh3 (Graph B).
Figure 8. Accumulation of cyclohexanol and cyclohexanone in oxidation of cyclohexane (0.46 M) with hydrogen peroxide (2.0 M, 50% aqueous) catalyzed by compound 4 (5 × 10−4 M) in the presence of HNO3 (0.05 M) in MeCN at 50 °C. Concentrations of cyclohexanone and cyclohexanol were determined before reduction of the PPh3 (Graph A), and the same reaction after reduction of the aliquots with solid PPh3 (Graph B).
Molecules 28 01211 g008
Figure 9. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 4 on the initial concentration of cyclohexane (Initial concentrations: Catalyst 5 × 10−4 M, H2O2 1.5 M, HNO3 0.05 M, CH3CN up to 2.5 mL, T = 50 °C).
Figure 9. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 4 on the initial concentration of cyclohexane (Initial concentrations: Catalyst 5 × 10−4 M, H2O2 1.5 M, HNO3 0.05 M, CH3CN up to 2.5 mL, T = 50 °C).
Molecules 28 01211 g009
Figure 10. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 4 on the initial concentration of catalyst. (Initial concentrations: cyclohexane 0.46 M, H2O2 1.5 M, catalyst 5 × 10−4 M, HNO3 0.05 M, CH3CN up to 2.5 mL, T = 50 °C).
Figure 10. Dependence of the initial rate of oxygenate (sum cyclohexanol + cyclohexanone) formation W0 for complex 4 on the initial concentration of catalyst. (Initial concentrations: cyclohexane 0.46 M, H2O2 1.5 M, catalyst 5 × 10−4 M, HNO3 0.05 M, CH3CN up to 2.5 mL, T = 50 °C).
Molecules 28 01211 g010
Figure 11. Accumulation of cyclohexanone (yield = 60%, TON = 600) in oxidation of cyclohexanol (0.5 M, green—curve 1), acetophenone (yield = 98%, TON = 980) in oxidation of phenylethanol (0.5 M, red—curve 2), 2-heptanone (yield = 48%, TON = 480) in oxidation of 2-heptanol (0.5 M, blue—curve 3) and 2-hexanone (yield = 50%, TON = 500) in oxidation of 2-hexanol (0.5 M, black—curve 4) with tert-butyl hydroperoxide (1.5 M, 70% aqueous) catalyzed by compound 4 (5 × 10−4 M) in MeCN at 50 °C.
Figure 11. Accumulation of cyclohexanone (yield = 60%, TON = 600) in oxidation of cyclohexanol (0.5 M, green—curve 1), acetophenone (yield = 98%, TON = 980) in oxidation of phenylethanol (0.5 M, red—curve 2), 2-heptanone (yield = 48%, TON = 480) in oxidation of 2-heptanol (0.5 M, blue—curve 3) and 2-hexanone (yield = 50%, TON = 500) in oxidation of 2-hexanol (0.5 M, black—curve 4) with tert-butyl hydroperoxide (1.5 M, 70% aqueous) catalyzed by compound 4 (5 × 10−4 M) in MeCN at 50 °C.
Molecules 28 01211 g011
Figure 12. From top to bottom. IR spectra for 14.
Figure 12. From top to bottom. IR spectra for 14.
Molecules 28 01211 g012aMolecules 28 01211 g012b
Figure 13. UV-vis spectrum for 4.
Figure 13. UV-vis spectrum for 4.
Molecules 28 01211 g013
Table 1. Crystal Data and Structure Refinement for Compounds 1–4.
Table 1. Crystal Data and Structure Refinement for Compounds 1–4.
Identification Code1 • 2¾C3H7NO • ¼C3H6O1a2 • 2C2H6O34 • 2C3H7NO
Empirical formulaC51H130.75Cu8N10.75O44Si16C129H388Cu32O161Si64C31H88Cu8K2O42Si16C43H113Cu8N8O42RbSi16C34H90Br2Cs2Cu8N6O38Si16
Formula weight2556.768347.752169.062457.732574.58
Temperature, K100(2)100(2)100(2)100(2)100(2)
Crystal size, mm0.20 × 0.18 × 0.150.12 × 0.06 × 0.030.23 × 0.13 × 0.120.15 × 0.13 × 0.120.35 × 0.05 × 0.03
Wavelength, Å0.710730.752701.541841.541841.54184
Crystal systemMonoclinicOrthorhombicMonoclinicMonoclinicMonoclinic
Space groupP21/cP21212C2/cP21/nP21/c
a, Å18.7364(2)28.231(3)14.60230(14)17.15942(11)10.68626(9)
b, Å19.9150(3)16.8711(14)22.9138(2)14.19682(8)25.0419(2)
c, Å29.0275(3)17.3129(15)24.0571(2)21.78734(10)35.8364(3)
α, deg.9090909090
β deg.94.2936(11)9098.2831(9)90.7073(5)95.0906(9)
γ, deg.9090909090
V, Å310,800.8(2)8245.9(12)7965.40(12)5307.19(5)9552.15(14)
Z41424
Density (calc.), g/cm31.5721.6811.8091.5381.790
μ, mm−11.8042.7296.3124.67911.331
F(000)52804274441625125104
Theta range, deg.1.996–35.4581.992–30.9923.713–77.6563.716–79.5993.741–78.402
Index ranges−28 ≤ h ≤ 29,
−31 ≤ k ≤ 26,
−44 ≤ l ≤ 46
−38 ≤ h ≤ 38,
−22 ≤ k ≤ 22,
−23 ≤ l ≤ 23
−18 ≤ h ≤ 18,
−26 ≤ k ≤ 28,
−30 ≤ l ≤ 30
−21 ≤ h ≤ 21,
−18 ≤ k ≤ 18,
−21 ≤ l ≤ 27
−13 ≤ h ≤ 13,
−31 ≤ k ≤ 31,
−45 ≤ l ≤ 45
Reflections collected215,76485,33247,38067,623346,984
Independent reflections44,011 (Rint = 0.0482)21,823 (Rint = 0.0701)8327 (Rint = 0.0436)11,268 (Rint = 0.0408)20,299 (Rint = 0.1134)
Reflections observed29,70913,993736510,61718,431
Restraints/parameters111/117071/75615/46267/4960/840
R1/wR2 (I > 2σ(I))0.0510/0.12030.0721/0.16490.0556/0.15200.0715/0.14630.1054/0.2197
R1/wR2 (all data)0.0927/0.14310.1124/0.18990.0610/0.15940.0740/0.14760.1116/0.2232
Goodness-of-fit on F21.0371.0221.0241.0051.028
Extinction coefficient0.00067(6)0.000250(16)0.000418(18)
Tmin/Tmax0.663/0.7070.716/0.9030.234/0.5000.447/0.5600.090/0.699
Δρmax/Δρmin, e.Å−31.653/−1.3632.043/−1.2661.297/−0.8142.604/−1.8982.120/−2.670
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Bilyachenko, A.N.; Arteev, I.S.; Khrustalev, V.N.; Zueva, A.Y.; Shul’pina, L.S.; Shubina, E.S.; Ikonnikov, N.S.; Shul’pin, G.B. Cagelike Octacopper Methylsilsesquioxanes: Self-Assembly in the Focus of Alkaline Metal Ion Influence—Synthesis, Structure, and Catalytic Activity. Molecules 2023, 28, 1211. https://doi.org/10.3390/molecules28031211

AMA Style

Bilyachenko AN, Arteev IS, Khrustalev VN, Zueva AY, Shul’pina LS, Shubina ES, Ikonnikov NS, Shul’pin GB. Cagelike Octacopper Methylsilsesquioxanes: Self-Assembly in the Focus of Alkaline Metal Ion Influence—Synthesis, Structure, and Catalytic Activity. Molecules. 2023; 28(3):1211. https://doi.org/10.3390/molecules28031211

Chicago/Turabian Style

Bilyachenko, Alexey N., Ivan S. Arteev, Victor N. Khrustalev, Anna Y. Zueva, Lidia S. Shul’pina, Elena S. Shubina, Nikolay S. Ikonnikov, and Georgiy B. Shul’pin. 2023. "Cagelike Octacopper Methylsilsesquioxanes: Self-Assembly in the Focus of Alkaline Metal Ion Influence—Synthesis, Structure, and Catalytic Activity" Molecules 28, no. 3: 1211. https://doi.org/10.3390/molecules28031211

Article Metrics

Back to TopTop