Family of Polynuclear Nickel Cagelike ... - ACS Publications

Oct 19, 2017 - Peoples' Friendship University of Russia, Miklukho-Maklay Str., 6, 117198 Moscow, Russia. §. Pirogov Russian National Research Medical...
4 downloads 0 Views 7MB Size
Article pubs.acs.org/IC

Cite This: Inorg. Chem. 2017, 56, 12751-12763

Family of Polynuclear Nickel Cagelike Phenylsilsesquioxanes; Features of Periodic Networks and Magnetic Properties Alexey N. Bilyachenko,*,†,‡ Alexey Yalymov,† Marina Dronova,† Alexander A. Korlyukov,†,§ Anna V. Vologzhanina,† Marina A. Es’kova,† Jérôme Long,∥ Joulia Larionova,*,∥ Yannick Guari,∥ Pavel V. Dorovatovskii,⊥ Elena S. Shubina,† and Mikhail M. Levitsky† †

A.N.Nesmeyanov Institute of Organoelement Compounds of Russian Academy of Sciences, Vavilov str., 28, 119991 Moscow, Russia Peoples’ Friendship University of Russia, Miklukho-Maklay Str., 6, 117198 Moscow, Russia § Pirogov Russian National Research Medical University, Ostrovitianov str., 1, 117997 Moscow, Russia ∥ Institut Charles Gerhardt de Montpellier (ICGM), UMR5253, Equipe IMNO, Université de Montpellier, Site Triolet, Place Eugène Bataillon, 34095 Montpellier cedex 5, France ⊥ National Research Center “Kurchatov Institute” Akademika Kurchatova pl., 1, 123098 Moscow, Russia ‡

S Supporting Information *

ABSTRACT: A new family of bi-, tetra-, penta-, and hexanickel cagelike phenylsilsesquioxanes 1−6 was obtained by selfassembly and transmetalation procedures. Their crystal structures were established by single-crystal X-ray analysis, and features of crystal packing relevant to the network formation were studied by a topological analysis. Compounds 1, 2, and 4 are isolated architectures, while 3, 5, and 6 present extended 1D and 3D networks. The investigation of magnetic properties revealed the presence of ferro- (1 and 3−5) or antiferromagnetic (2 and 6) interactions between Ni(II) ions, giving rise in the most cases (1, 2, and 4−6) to the presence of a slow relaxation of the magnetization, which can originate from the spin frustration.



INTRODUCTION Polynuclear metal-based architectures with a cagelike topology have attracted a great deal of attention over several decades due to their enormous structural diversity and their striking potential applications, in many fields including molecular recognition, catalysis, electronics, or performance of reactions within the cavities.1 Among those, particular interest has been devoted to the synthesis and investigations of cagelike polycyclic compounds based on silsesquioxane RSiO1.5 building blocks,2 including different types of metal-containing derivatives.3 Different approaches have been employed to design such isolated architectures of various nuclearities, containing transition-metal ions or lanthanides linked through silicon atoms, including self-assembling reactions of tri-3a,d,4 or tetrasilanols5 as well as noncondensed cubane silsesquioxane3b,g,6 building blocks. Indeed, polynuclear architectures with various structural organizations, such as Ti-based adamantine,7 Cu-based Cooling Tower8 and Knight Helmet,9 Fe-based drum10 or In-based bird cage,11 Sn-based tricubanes,12 and Ndbased tetracubanes13 are available using these synthetic strategies. Cagelike metallasilsesquioxanes (CLMSs) often exhibit high catalytic activity, which makes them interesting as molecular models for catalysis,3i,14 and sometimes they present unusual magnetic properties.15 On the other hand, © 2017 American Chemical Society

CLMSs may also be used as building blocks to design various extended molecule-based networks, also called coordination polymers, as has been demonstrated for the first time by some of us in the example of a Cu(II)-based silsesquioxane 1D−3D network.3i,16 Recently, we focused on the design of Ni(II)-containing polyhedral silsesquioxanes with penta- and hexanuclear cagelike architectures.14b Interestingly, the nuclearity in these compounds has been controlled by employment of specific solvents during the self-assembly reaction between precursors, which gave the formation of the pentanuclear architecture in the case of dimethylformamide (DMF) and the hexanuclear architecture in the case of dimethyl sulfoxide (DMSO). In both compounds, Ni(II) ions linked through oxygen atoms are arranged in slightly distorted pentagonal and hexagonal polyhedra with peripheral phenylsilsesquioxane moieties, permitting a good isolation of each cluster from others in the crystal packing. These architectures present a slow relaxation of the magnetization arising from the presence of the spin frustration of Ni(II) ions in each cluster rather than from the occurrence of weak dipolar interactions. Note that a few cagelike Received: June 13, 2017 Published: October 19, 2017 12751

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry

sided [PhSi(O)O]5 in the case of 2 (Figures 1 and 2). Between the bases of these molecular prisms one could observe Ni-

phosphonate-based Ni(II) architectures containing 8 and 12 nickel ions have previously been reported, but the presence of a slow relaxation of the magnetization has never been reported up to now.17 Encouraged by these results, we pursued our effort in the design of Ni(II)-based cage silsesquioxanes with different numbers of Ni(II) ions and focused for the first time on the design of extended networks including Ni(II)-based CLMSs as repeating units. In this article we report the synthesis, crystal structures, and investigations of the magnetic properties of six new bi-, tetra-, penta-, and hexanuclear Ni(II)-based CLMSs presenting discrete or extended crystal structures. Two different approaches consisting of the substitution of the solvating ligands or the transmetalation reaction have been used. A particular emphasis is given here to the influence of the synthetic conditions on the nuclearity and structural features. In addition, a comparison of various types of structural organizations and the magnetic properties of these cage silsesquioxanes are also discussed.



RESULTS AND DISCUSSION Synthesis and Crystal Structures of 1−6. It is wellknown3i that one of the features of CLMS chemistry consists of the remarkable opportunity to obtain a complex cage geometry from very simple educts. To take advantage of this, we performed the parallel syntheses of Ni(II) silsesquioxanes starting from the trivial silane building block PhSi(OEt)3. Depending on the molar ratio between two precursors, sodium siloxanolate (obtained in situ from PhSi(OEt)3) and NiCl2· 6NH3, this reaction (Scheme 1) leads to the formation of Scheme 1. Representation of the Self-Assembly or Transmetalation Reactions Conducting to Formation of CLMSs 1−6

Figure 1. Molecular structure of CLMS 1. Color code: red for O, beige for Si, green for Ni, yellow for Na, dark gray for C, light gray for H atoms. Solvating ligands are not shown for clarity.

CLMS (PhSiO2)12Ni4Na8(O)2 (1) or (PhSiO2)10Ni5(NaOH) (2). Single crystals of both compounds appropriate for X-ray diffraction analysis were obtained by slow evaporation of the solvents. Note that two earlier publications18 have previously been reported on the alternative synthesis of distorted prismatic Ni4 CLMSs, but 1 differs from Ni4Na6 of ref 18 in its composition. The structural analysis shows that compounds 1 and 2 are discrete architectures, which can be represented by a prismatic type of polyhedron (distorted for 1 and regular for 2), with sixsided polygonal bases [PhSi(O)O]6 in the case of 1 and five-

Figure 2. Molecular structure of CLMS 2. Color code: red for O, beige for Si, green for Ni, yellow for Na atoms. Solvating ligands are not shown for clarity.

containing cages (Ni4 for 1 (Figure S1 in the Supporting Information), Ni5 for 2 (Figure S2 in the Supporting Information)). The most important difference between these two compounds consists of the presence of silanolate fragments 12752

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry Table 1. Main Structural Parameters of Nickel-Containing Fragments in 1−6 and in Compound from Ref 23 dimensionality

Ni−O(Ni), Å

1 2 3 4 ref 23 5

0D 0D 3D 0D 0D 1D

2.049(4)−2.130(4) 2.016(3)−2.109(3)

6

1D

2.004(4)−2.027(4)

a

1.939(2)−1.978(2) 1.998(3)−2.046(3) 2.059(3)−2.135(3)

∠NiONi, deg

Ni−Ni, Å 3.143(1)−3.147(1), 11.574(1)a 2.719(1)−2.767(1), 9.248(1)a 6.009(1), 12.408a 2.808(1)−2.848(1), 12.201a 2.823(1)−2.848(1), 12.841(1)a 3.156(1)−3.183(1) 14.424(1),a 11.485(2) 2.832(1)−2.833(1), 11.015(1)a

96.2(2)−99.1(2) 84.0(1)−85.7(1), 71.05(8)−74.18(8),b 141.92(12)−146.36(13)b 90.90(1) − 94.41(1) 88.1(1)−90.8(1) 95.9(1)−100.2(1) 88.62(19)−89.80(18), 60c 120,c 180c

b

The shortest intermolecular Ni−Ni distance. Angles through the central oxygen atom. cAngles through the central chloride atom.

Si−O−Na in 1 (compound 2 includes only a Na+ cation as a fragment of the (Na)+(OH)− unit). It is noteworthy that sodium ions of silanolate fragments could be involved in CLMS coordination polymer formation.3i,16 The distances and angles for both structures are given in Tables 1 and 2. Despite our expectations, X-ray investigations of compounds 1 and 2 revealed no formation of extended structures with isolated clusters. Table 2. Main Structural Parameters of Siloxane Cycles in 1−6 and Compound from Ref 23 CLMS

Si−O(Si), Å

∠SiOSi, deg

1 2 3 4 ref 23 5 6

1.630(4)−1.652(4) 1.624(3)−1.658(3) 1.621(3)−1.653(3) 1.617(2)−1.629(2) 1.626(4)−1.641(4) 1.596(3)−1.649(3) 1.629(5)−1.647(5)

130.9(3)−141.7(3) 130.7(2)−138.0(2) 129.8(2)−154.0(2) 136.0(2)−137.8(6) 135.9(2)−137.0(2) 131.2(2)−139.3(2) 133.3(3)−134.6(3)

Being interested in the study of controlled formation of CLMS-based coordination polymers, we focus on several synthetic transformations of silanolate-containing compound 1. First of all, a procedure of substitution of solvating ligands in 1 was carried out. Recently, Hasell and Cooper19 have established that 1,4-dioxane is an unique solvent, which can induce the packing of organic cagelike compounds in extended networks. This observation is in agreement with the results obtained by some of us in the case of Cu(II)-based CLMS coordination polymers.3i,16,20 Along this line of thought, we reacted 1 with 1,4-dioxane (Scheme 1). This procedure induces not only a solvent replacement (from ethanol/water to 1,4dioxane) but also a deep structural reorganization of the cagelike molecular architecture conducted to the formation of the three-dimensional compound (PhSiO2)14Ni2Na12(CO3) (3). Figure 4A shows the molecular fragment of this CLMS structure, which is quite unusual due to the presence of two heptameric [PhSi(O)O]7 ligands (bringing 14 negative charges) and two Ni(II) ions, providing only four positive charges. Only two transition-metal ions provide the formation of a “non-cubane” cage silsesquioxane framework (Figure S3a,b in the Supporting Information). As rare examples of such binuclear CLMSs, we could mention Cr2,4d,21 Fe2,22 Cu2,4c,8 and Zn2 silsesquioxanes.4c The geometry of the molecular fragment could be assigned as distorted prismatic with a prominent shift of prism bases (silsesquioxane ligands) from the vertical axis. Note that 3 includes the 14-membered siloxane cycle (Figure 3C), which to the best of our knowledge is the first observation for CLMSs. Note also that such a nontrivially

Figure 3. Three types of siloxanolate ligands in 1−6: (A) 10membered cycle for 2; (B) 12-membered cycle for 1 and 4−6; (C) 14membered cycle for 3. Color code: O in red, Si in beige.

high number of alkaline-metal ions (for CLMSs) is required to balance the charge neutrality. For comparison, we could cite Co-based15a and Cu-based16b CLMSs, containing only six sodium and cesium ions, respectively. One could suggest that such a huge amount of sodium centers in 3 could be a positive feature in the context of coordination polymer structure formation. The three-dimensional structure is then formed by connection of molecular cages through 1,4-dioxane leading to the three-periodic primitive cubic net with the distances between connected nodes being between 15.4 and 16.5 Å (Figure 4B). Furthermore, complex 3 is uncommon due to its unusual encapsulation features: the inner void of 3 comprises a sodium bicarbonate molecule (Figure S3b). In principle, the encapsulation of different molecular entities (water, pyrazine, chloride anion, etc.) by CLMSs is well-known feature.3i However, to our knowledge, the encapsulation of a carbonate 12753

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry

Figure 5. Crystal structure of CLMS 4. Color code: O in red, Si in beige, Ni in green, Cl in light green, N in blue.

In the next stage, we explored the influence of alkali-metal ions on CLMS coordination polymer formation. Considering that these ions are usually located in external positions of a cage skeleton of metallasilsesquioxanes, their capacity to connect discrete CLMS architectures in order to form extended structures depends on their size. For this reason, different alkali-metal ions have been used in reactions with complex 1. Along this line of thought, the partial replacement of sodium by potassium ions in 1 led to the isolated unusual 1D compound [(PhSiO2)6]2Ni4Na4K2(OH)2 (5) (Scheme 1 and Figure 6). The cluster unit structure of CLMS 5 is quite close to that of the pristine complex 1, with two six-sided polygonal bases, [PhSi(O)O]6, coordinating the Ni4 central layer (Figure 6A and Figure S5 in the Supporting Information for the Ni(II) ion arrangement) and 12-membered siloxane cycles (Figure 3). Importantly, unlike complex 1, CLMS 5 forms a 1D coordination polymer structure through the connection of molecular units via the contacts of solvating ethanol molecule oxygen atoms with potassium centers of the cages (Figure 6B). Symptomatic of this, repeating contacts between cage fragments of 5 are implemented exclusively through potassium (not sodium) centers. This observation is in total agreement with the aforementioned suggestion concerning the role of the alkali metal ion size. Along the same line of thought, the interaction of potassium silanolate [(PhSiO1.5)(KO0.5)]n, obtained in situ from PhSi(OEt)3, with CsF and then with NiCl2·6NH3 gave the Cs+containing CLMS [(PhSiO2)6]2Ni6K1.5Cs0.5Cl(OH)] (6) having a 1D structural organization (Scheme 1 and Figure 7). The complex contains six nickel ions per cage organized in a hexagonal polygon with the chloride anion situated in the center and 12-membered siloxane cycles (Figure S6 in the Supporting Information and Figure 3B). The geometry of the molecular fragment of 6 is regular prismatic one. Positively charged alkaline ions balance the negative charges of OH− and chloride anions, encapsulated by inner voids of the metallasilsesquioxane framework. The cage fragments are connected through the linkages K−O−K in order to form a onedimensional chain structure along the c axis (Figure 7B and Figure S7 in the Supporting Information). The distances and angles are summarized in Tables 1 and 2. Note that the

Figure 4. (A) Molecular structure of CLMS 3. Color code: red for O, beige for Si, green for Ni, yellow for Na. Solvating ligands (except of 1,4-dioxane) are not shown for clarity. (b) View of the 3D network of 3.

is an unprecedented case. The distances and angles for 3 are summarized in Tables 1 and 2. The next step of our approach consists of the replacement of sodium by Ni(II) ions in 1 through the reaction with NiCl2· 6NH3 in the presence of 1,4-dioxane (transmetalation reaction) (Scheme 1). This results in the formation of the isolated cagelike Ni(II)-based phenylsilsesquioxane (PhSiO2)12Ni6 (4), having a crystal structure where six Ni(II) ions are organized into hexanuclear polygons (Figure 5, Tables 1 and 2, and Figure S4 in the Supporting Information) by means of the oxygen atoms and 12-membered siloxane cycles (Figure 3B). Its structural organization is close to that of the compound (PhSiO2)12Ni6(NaCl), previously reported by some of us.23 It should be noted that such transmetalation reactions are wellknown in the chemistry of CLMSs,3i,24 but no example involving a Ni(II)-based CLMS has been reported to date. Notwithstanding, despite the presence of 1,4-dioxane solvate, compound 4 does not form an extended structure. Thus, it has been established that the reactions with 1,4-dioxane cannot be considered as the usual way for the formation of CLMS-based coordination polymers. 12754

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry

Figure 6. (A) Molecular structure of CLMS 5. Color code: O in red, Si in beige, Ni in green, Na in yellow, K in purple. (B) View of the 1D network of 5. Figure 7. (A) Molecular structure of CLMS 6. Color code: O in red, Si in beige, Ni in green, Cl in light green, K in purple, Cs in violet. (B) View of the 1D network of 6 along the c crystallographic axis.

distances between the centroids of nickelphenylsilsesquioxane cages vary from 14.3 to 17.2 Å. As a short conclusion, two approaches consisting of the change of the solvent molecules by 1,4-dioxane as well as the transmetalation reaction dedicated to the formation of the extended structures from compound 1 have been explored. The strategy based on an involvement of an alkali metal with a greater ion size (K+ in CLMS 5 or Cs+ in CLMS 6 instead of Na+ in CLMS 1) has been found to be more productive than the method of 1,4-dioxane use as a linker (effective in the case of complex 3 but not for the product 4). Moreover, the

structural rearrangement is frequently concomitant with a modification of the cluster niclearity with respect to the pristine compound 1. Magnetic Properties. dc Magnetic Properties. The magnetic properties of all CLMSs were studied by using SQUID magnetometry in both static (direct current, dc) and dynamic (alternating current, ac) modes. With the exception of 12755

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry 6, the χT values at 300 K are in rather good accordance with the presence of the expected number of isolated octahedral Ni2+ ions (S = 1; g = 2.1, Table 3). The thermal dependence of χT Table 3. Main Magnetic Parameters for 1−6

1 2 3 4 5 6

χT (300 K)/ cm3 K mol−1

theor χT (300 K, g = 2.1)/cm3 K mol−1

M (70 kOe, 1.8 K)/μB

dominant interactiona

4.73 5.99 1.97 6.82 4.04 3.19

4.40 5.51 2.20 6.61 4.40 6.61

6.02 0.82 2.20 5.09 2.19 0.40

F AF F F F/AF AF

a

F denotes a ferromagnetic interaction and AF an antiferromagnetic interaction.

for 1, 3, and 4 reveals the presence of a plateau for all compounds down to 150 K before diverging from the Curie law upon further cooling. The observed positive deviation is indicative of dominant ferromagnetic interactions between the Ni2+ ions, and the maximum χT values are 9.23 (6 K), 3.68 (8 K), and 13.91 cm3 K mol−1 (8.5 K) for 1, 3, and 4, respectively. For 1 and 4, the maximum values of χT are lower than the theoretical value of 11.02 cm3 K mol−1 (S = 4 species, g = 2.1) and 23.1 cm3 K mol−1 (S = 6 species, g = 2.1) calculated for these isolated molecules assuming ferromagnetic interactions between all the nickel ions. For lower temperature, a steep decrease in χT is observed for 1, 3, and 4, which may originate from zero-field splitting and/or intermolecular interactions (Figure 8a). The field dependence of the magnetization, M, at 1.8 K confirms the presence of rather ferromagnetic interactions at low temperature for 1 and 4 without reaching a clear saturation, indicating the presence of a magnetic anisotropy (Figure 8b). Note that compound 3 presents a very weak value of the magnetization at 70 kOe, which may be explained by a strong zero-field splitting occurring due to an unusual geometry of Ni(II) ions in this structure. The room-temperature χT value for 5 is equal to 4.04 cm3 K mol−1, which is slightly lower than the expected value of 4.40 cm3 K mol−1 for four noninteracting octahedral Ni2+ ions with g = 2.1. The thermal dependence of χT shows a progressive decrease down to a minimum value of 3.25 cm3 K mol−1 at 60 K and then a dramatic increase to a value of 5.87 cm3 K mol−1 at 8 K (Figure 8a). This value remains lower than the expected value of 11.00 cm3 K mol−1 in the case of ferromagnetic coupling between all the spin carriers. Such a shape of the χT vs T curve reflects the occurrence of both competitive antiferromagnetic and ferromagnetic interactions within the tetranuclear unit mediated by different magnetic pathways. This is confirmed by an M vs H curve showing a low value of 2.19 μB at 70 kOe. In contrast, compounds 2 and 6 exhibit a strikingly different thermal dependence of χT. At room temperature, the experimental χT values are lower than the expected values for the isolated Ni2+ ions, indicating that antiferromagnetic interactions are still operative. Upon cooling, χT continuously decreases to reach a plateau at 14 K for both compounds before dramatically decreasing at very low temperature (Figure 8a). This curve profile confirms the presence of a predominant antiferromagnetic interaction. Though the steep decrease may originate from a zero-field splitting or intermolecular interactions, the plateau may be indicative of the presence of

Figure 8. Temperature dependence of χT under a 1000 Oe dc field.

short-range magnetic ordering as we previously evidenced in pentagonal M5 clusters (M = Co, Ni).15b,c The weak values of the magnetization at 70 kOe observed for these compounds corroborate the presence of dominant antiferromagnetic interactions (Table 3 and Figure 8b). Magnetic Irreversibility and ac Properties. The occurrence of a slow relaxation of the magnetization was investigated first by using the zero-field-cooled/field-cooled (ZFC/FC) curves and second by dynamic measurements by using alternating current (ac) mode. The ZFC curve was obtained when the sample was cooled in zero magnetic field down to 2 K, and then a magnetic field of 100 Oe was applied and the magnetization was measured as a function of temperature. The FC was recorded after the ZFC curve by decreasing the temperature under the same magnetic field. Among the six presented compounds only compound 3 shows no evidence of a slow relaxation of the magnetization. This fact is not surprising, considering the dinuclear structure of the cage fragment with a relatively long Ni(II)−Ni(II) distance (6.009 Å). Under a zero-dc field, only a weak frequency dependence without a maximum in the available temperature range can be observed for 1 (Figure S10 in the Supporting Information). Such slow relaxation of the magnetization may arise from either the presence of a single-molecule magnet (SMM) or spin glass behavior. In the case of SMM, it is known that applying a dc field can cut short the quantum tunneling of the magnetization 12756

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry

time leading to the parameters summarized in Table 4. Note that the τ0 values found are lower in comparison with those usually observed for superparamagnetic systems (10−8−10−12 s). This is further confirmed when the Mydosh parameter φ (φ = (Tmax − Tmin)/(Tmax × log νmax − log νmin)) is calculated, which is often used to discriminate spin-glass/superparamagnetic systems on the basis of the amplitude of the out-of-phase peak’s shift with frequency.25 For 2, this parameter is found to be equal to 0.093, which is slightly lower than the threshold value fixed for superparamagnetic systems (φ > 0.1). Another evidence of a spin-glass behavior is related to the field dependence of the ac susceptibility at fixed frequency. Such an analysis for 2 performed on χ′ shows that the maximum temperature decreases with field and follows the Almeida− Thouless law26 (for fields larger than 50 Oe). This fact is in agreement with the presence of a spin-glass behavior. All these data suggest that the slow relaxation observed in 2 arises from a spin-glass behavior induced by spin frustration due to the odd number of spin carriers forming the pentagonal cluster with the presence of different magnetic interaction pathways (between adjacent cis and trans Ni2+ ions). However, the presence of dipolar interactions between the cages could not be totally excluded. The ZFC curve for 4 reveals a maximum located at 3.9 K (Figure S17 in the Supporting Information). However, the increase in χ″ for temperatures lower than 2.5 K suggests the presence of a second relaxation process. The FC curve continuously increases and diverges from the ZFC curve at Tirr = 4.6 K. 4 exhibits an hysteresis loop at 1.8 K with a coercive field of 200 Oe, confirming a magnetic irreversibility (Figure S18 in the Supporting Information). The temperature dependence of the ac susceptibilities reveals a frequencydependent behavior, indicating also the presence of a slow relaxation of the magnetization (Figure 10). The temperature dependence of χ′ shows the presence of two different maxima, while χ″ exhibits only a single maximum accompanied by a tail at low temperature. The temperature dependence of the relaxation time for the maximum of χ″ gives Ueff = 84 K and τ0 = 1.9 × 10−14 s (Figure S19 in the Supporting Information, Table 4). As in the previous case of compound 2, the value of τ0 is lower than that expected for superparamagnetic systems and may suggest the occurrence of a spin-glass behavior resulting from the different exchange interactions between the paramagnetic centers within the hexanuclear unit. The Mydosh parameter, φ = 0.80, also suggests such a relaxation feature and the scenario is ultimately confirmed by the temperature dependence of the ac susceptibilities for various dc fields. It can be seen that, upon applying dc fields, the maxima of both χ′ and χ″ shift to lower temperature with magnetic fields (Figure S20 in the Supporting Information). The temperature dependence of the χ″ maximum follows the Almeida−Thouless (AT) equation, H ∝ (1 − Tmax/ Tf)3/2,26 Tf being the freezing temperature (Figure S21 in the

(QTM) and eventually shift the maximum to higher temperatures. ac measurements performed under a 900 Oe dc field reveal a more pronounced frequency dependence, but the maximum still remains located below 1.8 K (Figure S11 in the Supporting Information), precluding further analysis of the relaxation dynamics. The ZFC curve for 2 shows a maximum at Tmax = 3.3 K (Figure S12 in the Supporting Information), while the FC curve increases. Both curves overlap at high temperatures and start to separate at 5.0 K (irreversible temperature, Tirr). A clear opening in the hysteresis loop can be observed at 1.8 K with a coercive field of 350 Oe, confirming the magnetic irreversibility (Figure S13 in the Supporting Information). ac magnetic properties (Figure 9) reveal a frequency-dependent behavior

Figure 9. Temperature dependence of the in-phase (χ′) and out-of phase (χ″) susceptibilities performed under a zero dc field for 2.

with a single peak, for which the maximum shifts to higher temperatures as the frequency increases, indicating the occurrence of a slow relaxation of the magnetization. The Arrhenius equation τ = τ0 exp(Ueff/kBT) (Figure S14 in the Supporting Information), in which Ueff is the average energy barrier for the reversal of the magnetization, τ0 is the attempt time, and kB is the Boltzmann constant, was used for a description of the temperature dependence of the relaxation Table 4. Magnetic Irreversibility and ac Parameters for 1−6 CLMS

Tmax (ZFC)/K

Hc (1.8 K)/Oe

Ueff/K

1 2 4 5 6

2σ(I)) Rw,b % GOFc residual electron density, e Å−3 (dmin/dmax)

2

1

Table 5. Main Crystallographic Data and Experimental Details for 1−6 C98.5H141.75K2Na4Ni4O41.5Si12 2731.94 120.0(2) triclinic P1̅ 2 17.1030(5) 18.1176(5) 22.0120(6) 92.673(2) 94.282(2) 109.005(2) 6412.7(3) 1.415 3.106 2854 132 80058 21435 (0.068) 14218/113/1397 0.085 0.246 0.969 1.69/−1.78

5

C90H114ClCs0.5K1.5Ni6O33.75S9Si12 2759.78 120.0(2) trigonal R3̅ 3 26.086(5) 26.086(5) 17.175(3) 90 90 120 10121(4) 1.415 1.330 4431 60.4 24506 6620 (0.105) 3524/18/258 0.088 0.298 1.041 2.47/−1.75

6

Inorganic Chemistry Article

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry thermometer, and condenser. The mixture was heated at reflux for 1 h, and then nickel hexaamine chloride (Ni(NH3)6Cl2; 0.97 g, 4.17 mmol) was added at once. The reaction mixture was heated under reflux for 3.5 h and then cooled to 25 °C and filtered into a round-bottom flask for crystallization. After 5 days the formation of a yellow crystal fraction was observed. A few selected single crystals were used for the X-ray study (1·(EtOH)12(H2O)4; for details, see below). Anal. Calcd for (PhSiO2)12(Ni)4(Na)8(O)2: Ni, 11.20; Si, 16.07; Na, 8.77. Found (in a vacuum-dried sample): Ni, 11.11; Si, 16.25; Na, 8.69. Yield: 0.56 g, 45%. [(PhSiO2)5]2Ni5(NaOH) (2). Phenyltriethoxysilane (PhSi(OEt)3; 3 g, 12.48 mmol), water (H2O; 0.45 g, 25 mmol), and sodium hydroxide (NaOH; 0.50 g, 12,50 mmol) in 75 mL of EtOH were placed into a three-necked flask, equipped with a magnetic stirrer, thermometer, and condenser. The mixture was heated at reflux for 1 h and then cooled to room temperature. Then, after 1.5 h of stirring at room temperature, nickel hexaamine chloride (Ni(NH3)6Cl2; 1.38 g, 5.94 mmol) was added at once. The reaction mixture was heated under reflux for 3.5 h and then cooled to 25 °C and was filtered into a round-bottom flask for crystallization. After 5 days the formation of s yellow crystal fraction was observed. A few selected single crystals were used for the X-ray study (2·(EtOH)3(H2O)5; for details, see below). Anal. Calcd for (PhSiO2)10(Ni)5(NaOH): Ni, 17.21; Si, 16.47. Found (in a vacuum-dried sample): Ni, 17.08, Si, 16.36. Yield: 0.25 g, 22%. [(PhSiO2)7]2Ni2Na12(CO3) (3). A 0.25 g portion (0.09 mmol) of 1 was placed into a three-necked flask, equipped with a magnetic stirrer, thermometer, and condenser. Then 1.4-dioxane and methanol (60 mL, in 6:1 ratio) were added, and the resulting mixture was heated at reflux for 4 h and filtered. The filtrate gave (after 2 weeks of storing in a cool place) a batch of yellow crystals. After crystal fraction growth had ceased (∼1 week), a few selected single crystals were used for the Xray study (3·(C4H8O2)5(MeOH)7(EtOH); for details, see below). Anal. Calcd for [(PhSiO2)7]2Ni2Na12(CO3): Ni, 4.94; Si, 16.56; Na, 11.62. Found (in a vacuum-dried sample): Ni, 4.88; Si, 16.41; Na, 11.40. Yield: 0.08 g, 18%. [(PhSiO2)6]2Ni6 (4). A 0.25 g portion (0.09 mmol) of 1 was placed into a three-necked flask, equipped with a magnetic stirrer, thermometer, and condenser. Then 1,4-dioxane and methanol (60 mL, in 6:1 ratio) were added and the mixture was heated at reflux for 1 h. Afterward Ni(NH3)6Cl2 (0.043 g, 0.186 mmol) was added, and the reaction mixture was heated under reflux for 2 h and then cooled to 25 °C and filtered into a round-bottom flask (with 7 mL of benzonitrile) for crystallization. After 7 days the formation of a yellow crystal fraction was observed. A few selected single crystals were used for the X-ray study (4·(C4H8O2)7(PhCN)2; for details see below). Anal. Calcd for (PhSiO2)12Ni6: Ni, 17.62; Si, 16.86. Found (in a vacuum-dried sample): Ni, 17.00; Si, 16.28. Yield: 0.07 g, 38%. [(PhSiO2)6]2Ni4Na4K2(OH)2 (5). A 0.25 g portion (0.09 mmol) of 1 was placed into three-necked flask, equipped with a magnetic stirrer, thermometer, and condenser. Then ethanol and toluene (60 mL, in 3:1 ratio) were added and the mixture was heated at reflux for 1 h. Afterward KCl (0.02 g, 0.275 mmol) was added, and the reaction mixture was heated under reflux for 4 h and then cooled to 25 °C and filtered into a round-bottom flask for crystallization. After 7 days the formation of a yellow crystal fraction was observed. A few selected single crystals were used for the X-ray study (5·(EtOH)13(H2O)2.5; for details, see below). Anal. Calcd for (PhSiO2)12(Ni)4Na4K2(OH)2: Ni, 17.12; Si, 16.39. Found (in a vacuum-dried sample): Ni, 17.00; Si, 16.28. Yield: 0.07 g, 36%. [(PhSiO2)6]2Ni6K1.5Cs0.5Cl(OH)] (6). Phenyltriethoxysilane (PhSi(OEt)3; 3 g, 12.48 mmol), water (H2O; 0.45 g, 25 mmol), potassium hydroxide (KOH; 0.70 g, 12.50 mmol) in 75 mL of EtOH were placed into a three-necked flask, equipped with a magnetic stirrer, thermometer, and condenser. The mixture was heated at reflux for 1 h, and then cesium fluoride (CsF; 0.95 g, 6.24 mmol) and nickel hexaamine chloride (Ni(NH3)6Cl2; 0.96 g, 4.16 mmol) were added at once. The reaction mixture was heated under reflux for 3.5 h and then cooled to 25 °C and was filtered into an evaporation flask. The volatiles were removed under vacuum, and the solid product was recrystallized from DMSO (25 mL). Crystallization gave in 2 weeks

yellow crystals; a few selected single crystals were used for the X-ray study (6·DMSO 9 ; for details, see below). Anal. Calcd for (PhSiO2)12Ni6K1.5Cs0.5Cl(OH): Ni, 16.18; Si, 15.49; K, 2.70, Cs; 3.05. Found (in a vacuum-dried sample): Ni, 16.01; Si, 14.30; K, 2.61; Cs, 2.97. Yield: 0.22 g, 15%. X-ray Crystallography. X-ray diffraction studies were carried out on a bruker Apex DUO diffractometer using ω scans with Cu Kα radiation (5, λ = 1.54178 Å, multilayer optics) and Mo Kα radiation (the other compounds, λ = 0.71073 Å, graphite monochromator). The structures were solved by the direct method and refined by full-matrix least squares against F2. Non-hydrogen atoms were refined in anisotropic approximation except those for the disordered fragments. The crystal of 1 was twinned; twinning was resolved with Platon software29 and refined using BASF/HKLF 5 instructions. The H(C) atoms were included in the refinement by the riding model with Uiso(H) = nUeq(X), where n = 1.5 for methyl groups and water molecules and 1.2 for the other atoms. All calculations were made using the SHELXL-201430 and OLEX231 program packages. The main crystallographic data and experimental details are collected in Table 5. CCDC file numbers 1471547−1471550, 1556622, and 920380 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via http://www.ccdc.cam. ac.uk/conts/retrieving.html, or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, U.K. (fax, (+44) 1223-336-033; e-mail, [email protected]).



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acs.inorgchem.7b01436. Additional structures and analytical data as described in the text (PDF) Accession Codes

CCDC 1471547−1471550, 1556622, and 920380 contain the supplementary crystallographic data for this paper. These data can be obtained free of charge via www.ccdc.cam.ac.uk/ data_request/cif, or by emailing [email protected]. uk, or by contacting The Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033.



AUTHOR INFORMATION

Corresponding Authors

*E-mail for A.N.B.: [email protected]. *E-mail for J.L.: [email protected]. ORCID

Alexey N. Bilyachenko: 0000-0003-3136-3675 Anna V. Vologzhanina: 0000-0002-6228-303X Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS This work wassupported by the RSF (project 17-73-30036). J.L. acknowledges the support of the “RUDN University Program 5-100”, the University of Montpellier, CNRS, and the Plateforme d’ Analyse et de Caracterí sation ICGM for the support of X-ray studies.



REFERENCES

(1) For some recent reviews see: (a) Inokuma, Y.; Kawano, M.; Fujita, M. Crystalline molecular flasks. Nat. Chem. 2011, 3, 349−358. (b) Miras, H. N.; Yan, J.; Long, D.-L.; Cronin, L. Engineering polyoxometalates with emergent properties. Chem. Soc. Rev. 2012, 41, 12761

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry

generated polysiloxanes stabilizing μ3-oxo bridged Sb3 triangles. Dalton Trans. 2016, 45, 6963−6967. (5) Hirotsu, M.; Taruno, S.; Yoshimura, T.; Ueno, K.; Unno, M.; Matsumoto, H. Synthesis and Structures of the First Titanium(IV) Complexes with Cyclic Tetrasiloxide Ligands: Incomplete and Complete Cage Titanosiloxanes. Chem. Lett. 2005, 34, 1542−1543. (6) (a) Ehle, S.; Lorenz, V.; Liebing, P.; Hilfert, L.; Edelmann, F. T. Synthesis and structural characterization of two complex tantalum(v) siloxides. Inorg. Chem. Commun. 2016, 74, 82−85. (b) Gießmann, S.; Lorenz, V.; Liebing, P.; Hilfert, L.; Fischer, A.; Edelmann, F. T. Synthesis and structural study of new metallasilsesquioxanes of potassium and uranium. Dalton Trans. 2017, 46, 2415−2419. (7) Voigt, A.; Murugavel, R.; Roesky, H. W.; Schmidt, H.-G. Syntheses, spectroscopy and crystal structures of new group 4 metallasiloxanes. J. Mol. Struct. 1997, 436−437, 49−57. (8) Bilyachenko, A. N.; Dronova, M. S.; Yalymov, A. I.; Korlyukov, A. A.; Shul’pina, L. S.; Arkhipov, D. E.; Shubina, E. S.; Levitsky, M. M.; Kirilin, A. D.; Shul’pin, G. B. Binuclear Cage-Like Copper(II) Silsesquioxane (“Cooling Tower”) − Its High Catalytic Activity in the Oxidation of Benzene and Alcohols. Eur. J. Inorg. Chem. 2013, 2013, 5240−5246. (9) 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.; Golub, I. E.; Gulyaeva, E. S.; Shubina, E. S.; Shul’pin, G. B. Unusual tri-, hexa- and nonanuclear Cu(II) cage methylsilsesquioxanes: Synthesis, Structures and Catalytic Activity in Oxidations with Peroxides. Inorg. Chem. 2017, 56, 4093− 4103. (10) Nehete, U. N.; Anantharaman, G.; Chandrasekhar, V.; Murugavel, R.; Walawalkar, M. G.; Roesky, H. W.; Vidovic, D.; Magull, J.; Samwer, K.; Sass, B. Polyhedral Ferrous and Ferric Siloxanes. Angew. Chem., Int. Ed. 2004, 43, 3832−3835. (11) Voigt, A.; Walawalkar, M. G.; Murugavel, R.; Roesky, H. W.; Parisini, E.; Lubini, P. Organic-Soluble Neutral and Ionic Indium Siloxane Cages: Potential Precursors for Indium-Containing Silicates. Angew. Chem., Int. Ed. Engl. 1997, 36, 2203−2205. (12) Duchateau, R.; Dijkstra, T. W.; Severn, J. I.V. R.; van Santen, R. A.; Korobkov, I. V. Synthesis and characterization of tin containing polyhedral oligometallasilsesquioxanes (POMSS). Dalton Trans. 2004, 2677−2682. (13) Wu, G.; Chen, Y.; Xu, D.-J.; Liu, J.-C.; Sun, W.; Shen, Z. Synthesis and molecular structure of a tetrameric neodymiumsilsesquioxane cage complex: {[(i-C4H9)7(Si7O12)Nd]4NaCl}. J. Organomet. Chem. 2009, 694, 1571−1574. (14) (a) 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. (b) Levitsky, M. M.; Yalymov, A. I.; Kulakova, A. N.; Petrov, A. A.; Bilyachenko, A. N. Cage-like metallsilsesquioxanes in catalysis: A review. J. Mol. Catal. A: Chem. 2017, 426, 297−304. (c) 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. J. Organomet. Chem. 2017, DOI: 10.1016/j.jorganchem.2017.05.007. (d) Yalymov, A. I.; Bilyachenko, A. N.; Levitsky, M. M.; Korlyukov, A. A.; Khrustalev, V. N.; Shul’pina, L. S.; Dorovatovskii, P. V.; Es’kova, M. A.; Lamaty, F.; Bantreil, X.; Villemejeanne, B.; Martinez, J.; Shubina, E. S.; Kozlov, Y. N.; Shul’pin, G. B. High Catalytic Activity of Heterometallic (Fe6Na7 and Fe6Na6) Cage Silsesquioxanes in Oxidations with Peroxides. Catalysts 2017, 7 (4), 101. (e) 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.; Bantreil, X.; Villemejeanne, B.; Ruiz, C.; Martinez, J.; Shubina, E. S.; Shul’pin, G. B. Ionic Complexes of Tetra- and Nonanuclear Cage Cu(II)-phenylsilsesquioxanes: Synthesis and High Activity in Oxidative Catalysis. ChemCatChem 2017, DOI: 10.1002/ cctc.201701063. (15) (a) Bilyachenko, A. N.; Yalymov, A. I.; Korlyukov, A. A.; Long, J.; Larionova, J.; Guari, Y.; Zubavichus, Y. V.; Trigub, A. L.; Shubina, E.

7403−7430. (c) Cook, T. R.; Zheng, Y.-R.; Stang, P. J. Metal−Organic Frameworks and Self-Assembled Supramolecular Coordination Complexes: Comparing and Contrasting the Design, Synthesis, and Functionality of Metal−Organic Materials. Chem. Rev. 2013, 113, 734−777. (d) Zhang, G.; Mastalerz, M. Organic cage compounds − from shape-persistency to function. Chem. Soc. Rev. 2014, 43, 1934− 1947. (e) Ahmad, N.; Chughtai, A. H.; Younus, H. A.; Verpoort, F. Discrete metal-carboxylate self-assembled cages: Design, synthesis and applications. Coord. Chem. Rev. 2014, 280, 1−27. (f) Huang, S.-L.; Jin, G.-X.; Luo, H.-K.; Hor, T. S. A. Engineering Organic Macrocycles and Cages: Versatile Bonding Approaches. Chem. - Asian J. 2015, 10, 24− 42. (g) Xu, L.; Wang, Y.-X.; Yang, H.-B. Recent advances in the construction of fluorescent metallocycles and metallocages via coordination-driven self-assembly. Dalton Trans. 2015, 44, 867−890. (h) Brown, C. J.; Toste, F. D.; Bergman, R. G.; Raymond, K. N. Supramolecular Catalysis in Metal−Ligand Cluster Hosts. Chem. Rev. 2015, 115, 3012−3035. (2) (a) Markovic, E.; Constantopolous, K.; Matisons, J. G.; Hartmann-Thompson, C. Polyhedral Oligomeric Silsesquioxanes: From Early and Strategic Development through to Materials Application, in Applications of Polyhedral Oligomeric Silsesquioxanes. Adv. Silicon Sci. 2011, 3, 1−46. (b) Cordes, D. B.; Lickiss, P. D.; Hartmann-Thompson, C. Preparation and Characterization of Polyhedral Oligosilsesquioxanes, in Applications of Polyhedral Oligomeric Silsesquioxanes. Adv. Silicon Sci. 2011, 3, 47−133. (3) (a) 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. (b) 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. (c) 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. (d) 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. (e) Lorenz, V.; Edelmann, F. T. Metallasilsesquioxanes. Adv. Organomet. Chem. 2005, 53, 101−153. (f) Jutzi, P.; Lindemann, H. M.; Nolte, J.-O.; Schneider, M. Synthesis, Structure, and Reactivity of Novel Oligomeric Titana-siloxanes. In Silicon Chemistry: From the Atom to Extended Systems; Jutzi, P., Schubert, U., Eds.; Wiley: Hoboken, NJ, 2007, pp 372−382. (g) Edelmann, F. T. Metallasilsesquioxanes. Synthetic and Structural Studies. In Silicon Chemistry: From the Atom to Extended Systems; Jutzi, P., Schubert, U., Eds.; Wiley: Hoboken, NJ, 2007; pp 383−394. (h) Ward, A. J.; Masters, A. F.; Maschmeyer, T. Adv. Silicon Sci. 2011, 3, 135−166. (i) Levitsky, M. M.; Bilyachenko, A. N. Modern concepts and methods in the chemistry of polyhedral metallasiloxanes. Coord. Chem. Rev. 2016, 306, 235−269. (4) For some recent papers see: (a) Tan, G.; Yang, Y.; Chu, C.; Zhu, H.; Roesky, H. W. Cu24O24Si8R8: Organic Soluble 56-Membered Copper(I) Siloxane Cage and Its Use in Homogeneous Catalysis. J. Am. Chem. Soc. 2010, 132, 12231−12233. (b) Li, Y.; Wang, J.; Wu, Y.; Zhu, H.; Samuel, P. P.; Roesky, H. W. Synthesis of metallasiloxanes of group 13−15 and their application in catalysis. Dalton Trans. 2013, 42, 13715−13722. (c) Schax, F.; Braun, B.; Limberg, C. A Tripodal Trisilanol Ligand and Its Complexation Behavior towards CuI, CuII, and ZnII. Eur. J. Inorg. Chem. 2014, 2014, 2124−2130. (d) Schax, F.; Bill, E.; Herwig, C.; Limberg, C. Dioxygen Activation by Siloxide Complexes of Chromium(II) and Chromium(IV). Angew. Chem., Int. Ed. 2014, 53, 12741−12745. (e) Lokare, K. S.; Frank, N.; Braun-Cula, B.; Goikoetxea, I.; Sauer, J.; Limberg, C. Trapping Aluminum Hydroxide Clusters with Trisilanols during Speciation in Aluminum(III)−Water Systems: Reproducible, Large Scale Access to Molecular Aluminate Models. Angew. Chem., Int. Ed. 2016, 55, 12325−12329. (f) Santhana Raj Prabhu, M.; Ugandhar, U.; Baskar, V. In situ 12762

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763

Article

Inorganic Chemistry S.; Eremenko, I. L.; Efimov, N. N.; Levitsky, M. M. Heterometallic Na6Co3 Phenylsilsesquioxane Exhibiting Slow Dynamic Behavior in its Magnetization. Chem. - Eur. J. 2015, 21, 18563−18565. (b) 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. (c) Bilyachenko, A. N.; Yalymov, A. I.; Levitsky, M. M.; Korlyukov, A. A.; Eskova, M. A.; Long, J.; Larionova, J.; Guari, Y.; Shul’pina, L. S.; Ikonnikov, N. S.; Trigub, A. L.; Zubavichus, Y. V.; Golub, I. E.; Shubina, E. S.; Shul’pin, G. B. First Cage-Like Pentanuclear Co(II)-Silsesquioxane. Dalton Trans. 2016, 45, 13663− 13666. (d) Bilyachenko, A. N.; Levitsky, M. M.; Yalymov, A. I.; Korlyukov, A. A.; Vologzhanina, A. V.; Kozlov, Yu. N.; Shul’pina, L. S.; Nesterov, D. N.; Pombeiro, A. J. L.; Lamaty, F.; Bantreil, X.; Fetre, A.; Liu, D.; Martinez; Long, J.; Larionova, J.; Guari, Y.; Trigub, A. L.; Zubavichus, Y. V.; Golub, I. E.; Filippov, O. A.; Shubina, E. S.; Shul’pin, G. B. A Heterometallic (Fe6Na8) Cage-like Silsesquioxane: Synthesis, Structure, Spin Glass Behavior and High Catalytic Activity. RSC Adv. 2016, 6, 48165−48180. (16) (a) Dronova, M. S. Ph.D. thesis, 2013, Moscow. (b) Korlyukov, A. A.; Vologzhanina, A. V.; Buzin, M. I.; Sergienko, N. V.; Zavin, B. G.; Muzafarov, A. M. Cu(II)-Silsesquioxanes as Secondary Building Units for Construction of Coordination Polymers: A Case Study of CesiumContaining Compounds. Cryst. Growth Des. 2016, 16, 1968−1977. (17) (a) Breeze, B. A.; Shanmugam, M.; Tuna, F.; Winpenny, R. E. P. A series of nickel phosphonate−carboxylate cages. Chem. Commun. 2007, 5185−5187. (b) Sheikh, J. A.; Adhikary, A.; Jena, H. S.; Biswas, S.; Konar, S. High Nuclearity (Octa-, Dodeca-, and Pentadecanuclear) Metal (M = CoII, NiII) Phosphonate Cages: Synthesis, Structure, and Magnetic Behavior. Inorg. Chem. 2014, 53, 1606−1613. (c) Sheikh, J. A.; Jena, H. S.; Clearfield, A.; Konar, S. Phosphonate Based High Nuclearity Magnetic Cages. Acc. Chem. Res. 2016, 49, 1093−1103. (18) (a) Igonin, V. A.; Lindeman, S. V.; Potekhin, K. A.; Shklover, V. E.; Struchkov, Yu. T.; Shchegolikhina, O. I.; Zhdanov, A. A.; Razumovskaya, I. V. Organomet. Chem. USSR 1991, 4, 383. (Metalloorg. Khim. 1991, 4, 790) (b) Anisimov, A. A.; Kononevich, Y. N.; Zhemchugov, P. V.; Milenin, S. A.; Korlyukov, A. A.; Tsareva, U. S.; Peregudov, A. S.; Dorovatovskii, P. V.; Molodtsova, Y. A.; Takazova, R. U.; Shchegolikhina, O. I.; Muzafarov, A. M. Synthesis and structure of new polyhedral Ni, Na- and Cu, Na-metallasiloxanes with tolyl substituent at the silicon atom. RSC Adv. 2016, 6, 22052− 22060. (19) Hasell, T.; Culshaw, J. L.; Chong, S. Y.; Schmidtmann, M.; Little, M. A.; Jelfs, K. E.; Pyzer-Knapp, E. O.; Shepherd, H.; Adams, D. J.; Day, G. M.; Cooper, A. I. Controlling the Crystallization of Porous Organic Cages: Molecular Analogs of Isoreticular Frameworks Using Shape-Specific Directing Solvents. J. Am. Chem. Soc. 2014, 136, 1438− 1448. (20) 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 of benzene and alcohols with peroxides. Dalton Trans. 2014, 43, 872− 882. (21) Bilyachenko, A. N.; Korlyukov, A. A.; Levitskii, M. M.; Zavin, B. G.; Antipin, M.; Yu; Filippov, O. A.; Tsupreva, V. I. Dinuclear cage-like metalloorganosiloxane containing CrIII ions. Russ. Chem. Bull. 2008, 57, 2204. (22) Pinkert, D.; Demeshko, S.; Schax, F.; Braun, B.; Meyer, F.; Limberg, C. A Dinuclear Molecular Iron(II) Silicate with Two HighSpin Square-Planar FeO4 Units. Angew. Chem., Int. Ed. 2013, 52, 5155−5158. (23) 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.

(24) Bilyachenko, A. N.; Dronova, M. S.; Yalymov, A. I.; Lamaty, F.; Bantreil, X.; Martinez, J.; Bizet, C.; Shul’pina, L. S.; Korlyukov, A. A.; Arkhipov, D. E.; Levitsky, M. M.; Shubina, E. S.; Kirillov, A. M.; Shul’pin, G. B. Cage-like Copper(II) Silsesquioxanes: Transmetalation Reactions and Structural, Quantum Chemical, and Catalytic Studies. Chem. - Eur. J. 2015, 21, 8758−8770. (25) Mydosh, J. A. Spin glasses: an experimental introduction; Taylor & Francis: London, Washington, DC, 1993. (26) Almeida, J. R. L. D.; Thouless, D. J. Stability of the SherringtonKirkpatrick solution of a spin glass model. J. Phys. A: Math. Gen. 1978, 11 (5), 983. (27) Lartigue, L.; Oh, S.; Prouzet, E.; Guari, Y.; Larionova, J. Superspin-glass behavior of Co3[Fe(CN)6]2 Prussian blue nanoparticles confined in mesoporous silica. Mater. Chem. Phys. 2012, 132 (2−3), 438−445. (28) Banerjee, S.; Nandy, M.; Sen, S.; Mandal, S.; Rosair, G. M.; Slawin, A. M. Z.; García, C. J.; Gómez; Clemente-Juan, J. M.; Zangrando, E.; Guidoling, N.; Mitra, S. Isolation of four new CoII/ CoIII and NiII complexes with a pentadentate Schiff base ligand: syntheses, structural descriptions and magnetic studies. Dalton Trans. 2011, 40, 1652−1661. (29) Spek, A. L. Structure validation in chemical crystallography. Acta Crystallogr., Sect. D: Biol. Crystallogr. 2009, 65, 148−155. (30) Sheldrick, G. M. Crystal structure refinement with SHELXL. Acta Crystallogr., Sect. C: Struct. Chem. 2015, 71, 3−8. (31) Dolomanov, O. V.; Bourhis, L. J.; Gildea, R. J.; Howard, J. A. K.; Puschmann, H. OLEX2: a complete structure solution, refinement and analysis program. J. Appl. Crystallogr. 2009, 42, 339−341.

12763

DOI: 10.1021/acs.inorgchem.7b01436 Inorg. Chem. 2017, 56, 12751−12763