Next Article in Journal
Advances in Nanogels for Topical Drug Delivery in Ocular Diseases
Next Article in Special Issue
Editorial on the Special Issue: “Advances in Xerogels: From Design to Applications”
Previous Article in Journal
Synthesis and Plugging Performance of Nano-Micron Polymeric Gel Microsphere Plugging Agents for Oil-Based Drilling Fluids
Previous Article in Special Issue
Green Chemistry for the Transformation of Chlorinated Wastes: Catalytic Hydrodechlorination on Pd-Ni and Pd-Fe Bimetallic Catalysts Supported on SiO2
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Non-Hydrolytic Sol–Gel Route to Organic-Inorganic Hybrid Polymers: Linearly Expanded Silica and Silsesquioxanes

1
Institute of Inorganic Chemistry, Department of Chemistry and Physics, Technische Universität Bergakademie Freiberg (TUBAF), Leipziger Strasse 29, 09596 Freiberg, Saxony, Germany
2
Center of Efficient High Temperature Processes and Material Conversion (ZeHS), Technische Universität Bergakademie Freiberg (TUBAF), Winklerstr. 5, 09599 Freiberg, Saxony, Germany
3
Institute of Analytical Chemistry, Department of Chemistry and Physics, Technische Universität Bergakademie Freiberg (TUBAF), Leipziger Strasse 29, 09596 Freiberg, Saxony, Germany
*
Author to whom correspondence should be addressed.
Gels 2023, 9(4), 291; https://doi.org/10.3390/gels9040291
Submission received: 10 March 2023 / Revised: 27 March 2023 / Accepted: 30 March 2023 / Published: 2 April 2023
(This article belongs to the Special Issue Advances in Xerogels: From Design to Applications)

Abstract

:
Condensation reactions of chlorosilanes (SiCl4 and CH3SiCl3) and bis(trimethylsilyl)ethers of rigid, quasi-linear diols (CH3)3SiO–AR–OSi(CH3)3 (AR = 4,4′-biphenylene (1) and 2,6-naphthylene (2)), with release of (CH3)3SiCl as a volatile byproduct, afforded novel hybrid materials that feature Si–O–C bridges. The precursors 1 and 2 were characterized using FTIR and multinuclear (1H, 13C, 29Si) NMR spectroscopy as well as single-crystal X-ray diffraction analysis in case of 2. Pyridine-catalyzed and non-catalyzed transformations were performed in THF at room temperature and at 60 °C. In most cases, soluble oligomers were obtained. The progress of these transsilylations was monitored in solution with 29Si NMR spectroscopy. Pyridine-catalyzed reactions with CH3SiCl3 proceeded until complete substitution of all chlorine atoms; however, no gelation or precipitation was found. In case of pyridine-catalyzed reactions of 1 and 2 with SiCl4, a Sol–Gel transition was observed. Ageing and syneresis yielded xerogels 1A and 2A, which exhibited large linear shrinkage of 57–59% and consequently low BET surface area of 10 m2⋅g−1. The xerogels were analyzed using powder-XRD, solid state 29Si NMR and FTIR spectroscopy, SEM/EDX, elemental analysis, and thermal gravimetric analysis. The SiCl4-derived amorphous xerogels consist of hydrolytically sensitive three-dimensional networks of SiO4-units linked by the arylene groups. The non-hydrolytic approach to hybrid materials may be applied to other silylated precursors, if the reactivity of the corresponding chlorine compound is sufficient.

1. Introduction

Organic-inorganic hybrid materials usually have properties intermediate between organic polymers and inorganic substances, such as oxide glasses or ceramics [1,2,3]. A well-known approach to organic-inorganic hybrid materials uses the classic Sol–Gel process based on hydrolysis reactions of (semi)metal alkoxides [4,5,6,7]. Non-aqueous or non-hydrolytic [8,9,10,11] and non-oxide [12,13] Sol–Gel systems have also been reported. The so-called non-hydrolytic routes frequently refer to oxidic products, which may be used for similar applications as oxide products derived from hydrolytic routes; these include catalyst supports [14], optical materials [15], or aerogels [16,17,18], among many others [4,5,6,7,8,9].
Non-hydrolytic Sol–Gel systems are frequently based on metal halides and their reactions with oxygen containing organic or organometallic compounds such as alkoxides or acetoxides. They yield polymeric networks that are comprised of M–O–M’ backbone units. Thus, these non-aqueous systems are basically alternatives to the classic aqueous Sol–Gel routes as they afford oxides or metal oxide based organic-inorganic hybrid materials. The M–O–M’ bridges are generated via a nucleophilic attack of the oxygen atoms at the metal atom of the halide. Therefore, the common principle of most non-hydrolytic Sol–Gel processes relies on a scission of carbon-oxygen-bonds instead of hydrolysis involving a nucleophilic attack of OH-groups (Scheme 1a). The (semi)metals M and M’ may be identical. For example, the reaction of SiCl4 with tetra-n-butoxysilane (Scheme 1b) has been used to generate silica gels upon formation of the liquid side product n-butylchloride [19]. The underlying reaction may be considered as an ether cleavage. This route can also be applied for other alkyl groups and may be simplified by using alcohols like methanol, ethanol, or isopropanol in order to form the alkoxide in situ via reactions with an element halide. In analogy to silica gels [20], other single component oxide gels (M = M’), such as alumina [21], titania [22] gels, or tungsten oxide [23], were prepared. Different mixed oxide ceramics such as TiO2/SiO2 [24], TiO2/Al2O3 [25], SiO2/Al2O3, or SiO2/ZrO2 [24,26] were synthesized using the same approach. This route has also been applied to prepare multinary solids such as YAG (yttrium aluminum garnet) powders [27], cordierites [28], or β-SiAlON:Eu powders [29]. In many cases, precipitation of the products, rather than Sol–Gel transition and monolith formation, has been reported. A number of related approaches, such as nonhydrolytic solvothermal synthesis of oxide nanoparticles under anhydrous conditions, e.g., [30], have been reported in recent years. In addition, numerous reports on the controlled aggregation and gelation of (non-hydrolytically) formed oxide colloids, i.e., nano-particle dispersions or sols, have been published [31].
Pronounced Lewis acidic metal halides, such as TiCl4, react with organic ethers, e.g., diisopropyl ether or THF (tetrahydrofuran), to give mesoporous TiO2-carbon nanocomposites [32]. In this process, which involves no additional solvent, the ether acts as both an oxygen donor and source of oxide and as the sole carbon source.
Acetoxysilanes and metal alkoxides (the latter in analogy to metal halides) react with formations of carboxylic acid esters and of non-aqueous oxide gels (Scheme 1c). The alkoxide may be a compound of the main group elements (silicon [33] and aluminum [21,33]) or of a transition metal (zirconium [26,33] or titanium [33,34]).
Further non-hydrolytic Sol–Gel systems are based on reactive organometallic compounds or alkoxides and further organic oxygen-containing compounds, e.g., acetone and zinc alkoxides react with formation of gels under enolization of the ketone and elimination of the corresponding alcohol [35]. This reaction behavior is observed only for non-acidic alkoxides.
Only few non-oxide Sol–Gel systems are known, i.e., Sol–Gel routes to sulfides, carbides, or nitrides, which lead to monolithic xerogels [12,13]. Some are related to the synthesis described here. In analogy to the hydrolysis and condensation reactions of the classic oxide Sol–Gel systems, compounds of the heavier chalcogenides may be considered as precursors [12,36].
Many papers on Sol–Gel synthesis routes to sulfides were published, but only a few Sol–Gel transitions were reported. For ZnS it was described that treatment of Et2Zn or [Zn(SR)2]x with H2S generates precipitates, while reactions of dibenzyl trisulfide (BnS)2S and Et2Zn in pentane (Scheme 2a) yielded transparent sols, gels, and xerogels after removal of the solvent [37]. The successful preparation of colloidal CdS and CdSe [38] gels, as well as CdSe sols obtained from Cd(ethoxyacetate)2 and Se(Si(CH3)3)2 [39], represents another non-oxide Sol–Gel strategy. CdS/CdSe aerogels can also be obtained from nanocrstyalline starting materials [40]. In another example, Ge(OEt)4 was treated with H2S to form sulfide gels [41]. Interesting nanostructured sulfide materials have been obtained from molybdenum chloride and S(Si(CH3)3)2 in chloroform [42]. A similar approach, that is also based on the formation of trimethylchlorosilane, has been used to generate polydimethylsiloxane (PDMS) analogous silicon-sulfur polymers [43].
Even more work has been published on the synthesis of nitride materials via Sol–Gel routes. It may be assumed that simple replacement of water by ammonia or amines leads to nitride gels. However, it turns out that, in most cases, this approach is not successful, probably due to the much higher basicity of NH3 and RNH2 when compared to H2O. The first example of a Sol–Gel process based on an ammonolysis reaction yielding Si/(C)/N gels starts from a dialkylamino substituted cyclic trisilazane (Scheme 2b); it can be described as a transamination reaction [44]. These so-called silicon diimide gels have subsequently been used for stationary phases in thin layer chromatography for organic acids [45].
Hexagonal boron nitride (h-BN) and ternary B/C/N ceramics have been synthesized using “silazanolysis” reactions of B-trichloroborazene and its alkyl substituted derivatives with hexa- and heptamethyl disilazane [46]. A three-dimensional network infiltrated with liquid (CH3)3SiCl is formed that is analogous to the oxide polymer networks infiltrated by solvents and liquid reaction products. The dried xerogels showed low specific surface areas (<35 m2⋅g−1) when compared to oxide xerogels. Pyrolysis at 1200 °C produced hexagonal boron nitride.
Another approach to synthesize gel precursors for carbide and nitride materials is based on the pseudo-chalcogen concept [47]. In principle, the oxygen atoms of polysiloxanes are replaced by carbodiimide (–N=C=N–) groups in order to obtain (poly)silylcarbodiimide compounds [48]. Reactions of bis(trimethylsilyl)carbodiimide (CH3)3Si–NCN–Si(CH3)3 and dichlorosilanes (such as (CH3)2SiCl2) give linear and cyclic oligomers [49], while reactions of this carbodiimide with trichlorosilanes (RSiCl3 with R = alkyl, aryl or H) or with tetrachlorosilane may result in the formation of transparent gels (Scheme 3a) [50,51]. Several attempts to extend this Sol–Gel system through replacing the chlorosilanes with other element halides were not successful. For example, analogous reactions with titanium chlorides, such as TiCl4 or CpTiCl3, led to the formation of precipitates; the use of RAlCl2 and gallium chlorides as starting materials gave soluble oligomers [52]. However, B-trichloroborazene (B3N3H3Cl3) reacts similarly to the chlorosilanes with bis(trimethylsilyl)carbodiimide and forms B/C/N gels (Scheme 3b) that can be transformed into B/C/N and B4C ceramics [53].
We modified the above-described Sol–Gel approach based on (CH3)3SiCl formation by replacing bis(trimethylsilyl)carbodiimide with trimethylsilyl esters of cyanuric and cyameluric acid (Scheme 3c). Their reactions with SiCl4 and CH3SiCl3 yielded networks consisting of three-fold bridged SiO4 and CH3SiO3 units [54]. Similar three-fold bridged networks were formed upon acid catalyzed reactions of trihydroxybenzene and tetraethoxysilane (TEOS) [55]. The same approach was used with TEOS and dihydroxybenzenes as bifunctional precursors [56].
Here we present the extension of the concept of reaction of element chlorides and silylated precursors to obtain “bridged” or “linearly expanded” silicas and silsesquioxanes. Reactions of rigid, quasi-linear bis(trimethylsilyl)ethers of diols of the type (CH3)3SiO–AR–OSi(CH3)3 (AR = quasi-linear arylene spacer) and chlorosilanes (SiCl4 and CH3SiCl3) afforded novel hybrid oligomers, polymers, and xerogels; the latter were investigated with X-ray diffraction (XRD), 29Si nuclear magnetic resonance (NMR), and Fourier-transform infrared(FTIR) spectroscopy, as well as with scanning electron microcopy coupled with energy-dispersive X-ray spectroscopy (SEM/EDX), elemental analysis, thermal gravimetry with differential thermal analysis (TG/DTA), and TG-FTIR.

2. Results and Discussion

2.1. Selection, Synthesis and Characterisation of the Starting Materials

As mentioned in the introduction, the goal of the present work was to synthesize “linearly expanded silica and silsesquioxanes”. The latter term defines hybrid materials that differ from the well-known “organically bridged” silsesquioxanes O1.5Si–[linker]–SiO1.5 [57,58,59,60] (Figure 1a), which feature various organic groups as linkers, and related systems such as polysilsesquioxanes [R–SiO1.5]n, with R being a terminal group (such as hydrocarbyl) [61,62]. These materials are based on the replacement of Si–O bonds by Si–C bonds. The present approach retains original Si–O motifs but alters the –O– bridge into an extended –O–[linker]–O– quasi-linear bidentate connector. Thus, these systems may be compared to metal organic frameworks (MOFs) [63], which may be considered “expanded” coordination centers (such as metal oxide centers) bridged by rigid bidentate ligands (such as terephthalate [O2C–C6H4–CO2]2−) symbolized as LyM–L″–[linker]–L″–MLy (Figure 1b). This route to usually crystalline network structures has been extended to covalent organic frameworks (COFs) [64].
Silylcarbodiimides may also by viewed as “linearly expanded Si–O” compounds, since the oxygen atoms are replaced by NCN-units as alternative difunctional linkers of similar electronegativity [48,49,50,51,52,53]. This approach differs from the above oxides since Si–O bonds are replaced by Si–N bonds (Figure 1c). In the present work, we aimed to “expand” SiOx-units with rigid spacers without changing the Si–O bonds, i.e., incorporating Si–O–C linkages. In a former study, we used three-fold functional units as organic bridges (Figure 1d) [54]. The obtained hybrid materials are highly cross-linked gels. In order to synthesize polymeric gels with a lower degree of cross-linking, we decided to use two-fold functionalized rigid bridges (Figure 1e).
All attempts to generate the target materials using reactions of commercially available diols, including 4,4′-dihydroxybiphenyl with alkoxysilanes such as TEOS or tetramethoxysilane (TMOS), were not successful. The poor solubility of these diols in polar and non-polar organic solvents may be one of the reasons. In order to start from homogeneous solutions, we decided to react silylated diols with chlorosilanes.
4,4′-Bis(trimethylsiloxy)biphenyl 1 was synthesized according to a previously reported protocol [65]; the silylated product was additionally purified by crystallization from n-hexane. 2,6-Bis(trimethylsiloxy)naphthalene 2 was prepared analogously by silylation of the diol with hexamethyldisilazane (HMDS) using KOtBu as a catalyst. Upon vacuum distillation, compound 2 solidified to afford a colorless crystalline solid. Both starting materials (1 and 2) showed sharp melting points and delivered satisfying elemental analyses. In the FTIR spectra, the expected absorption bands for Calkyl–H, Caryl–H, C=C, Si–O, and Si(CH3)3 groups were present. For both compounds, the 1H, 13C, and 29Si NMR spectra exclusively exhibited the respective set of signals to be expected for the symmetric two-fold silylated products. No indications for monosilylated diols or for any other impurities were found.
In case of compound 2, a single-crystal X-ray diffraction analysis was performed, providing complete structural information for this molecule (Figure 2). Selected bond lengths and angles are summarized in Table 1. All remaining structural data can be found in Table S1 (bond lengths), Table S2 (bond angles), and Table S3 (torsion angles).
The bonds between C4 and C5 (1.364 Å) and between C3 and C2 (1.364 Å) are significantly shorter than the remaining C–C bonds of the naphthalene group. Thus, four of the eleven bonds have an enhanced double bonding character. Nevertheless, all bond angles of the naphthalene unit are very close to 120° in accordance with a planar six-membered arene motif.

2.2. Oligomer and Polymer Formation in Solution

Solutions of the precursors (i.e., silylated diol 1 or 2, chlorosilane CH3SiCl3, or SiCl4—with and without addition of pyridine, in THF as the solvent) were allowed to react at room temperature and at 60 °C. Only the reactions with SiCl4 and pyridine (1aPy and 2aPy) afforded gels. All other reaction mixtures (1a = 1/SiCl4/THF; 2a = 2/SiCl4/THF; 1b = 1/CH3SiCl3/THF; 2b = 2/CH3SiCl3/THF; 1bPy = 1/CH3SiCl3/pyridine/THF; 2bPy= 2/CH3SiCl3/pyridine/THF) were sealed in NMR tubes and investigated with 29Si NMR spectroscopy (using external D2O). The 29Si NMR spectrum of reaction mixture 1b after one day at room temperature indicates a rather slow reaction; the spectrum is dominated by the signals of the starting materials, but the emergence of new signals at 31 ppm ((CH3)3SiCl) and −11 ppm (CH3SiCl2(O-aryl) confirm the transsilylation reaction even at room temperature and in absence of pyridine catalyst (cf. Figure S1 in the supporting information).
Upon further storage of reaction mixture 1b (biphenyl precursor 1 and CH3SiCl3 in THF without pyridine) at room temperature, a new signal at −33 ppm (MeSiCl(O-aryl)2 moieties) emerges in the 29Si NMR spectrum; this is the sole product signal after 30 days (cf. Figure S2 in the supporting information). An analogous reaction mixture 2b (using the naphthalene precursor 2 and CH3SiCl3 in THF without pyridine) exhibits similar behavior, but the emergence of the signal of CH3SiCl(O-aryl)2 groups (at −32 ppm) within one day and the disappearance of the signal of CH3SiCl2(O-aryl) groups (at −11 ppm) within 8 days indicates more rapid reaction of precursor 2 vs. 1 (cf. Figure S3 in the supporting information). Both systems have in common that transsilylation essentially stopped at the CH3SiCl(O-aryl)2 stage.
However, supplementing a reaction mixture such as 2b with pyridine as a nucleophilic catalyst (2bPy = 2 plus CH3SiCl3 and pyridine in THF) enhances the already known transsilylation steps and gives rise to the emergence of a new 29Si NMR signal at −52 ppm (CH3Si(O-aryl)3 moieties) within 9 hours; this is the exclusive high field signal after 6 days (cf. Figure S4 in the supporting information). Thus, with pyridine, a complete substitution of all three Cl atoms of CH3SiCl3 by O-atoms occurs, while the reaction seems to cease at the stage of a two-fold substitution without addition of pyridine. This rate-increasing effect of pyridine is observed in a similar way for the biphenyl derivative (1bPy).
For tetrachlorosilane, the reaction progress of non-gel-forming reaction mixtures can also be followed in solution by 29Si NMR spectroscopy. The corresponding spectra are depicted in the supplementary material (Figures S5 and S6). They also indicate that the naphthalene derivative 2 reacts faster than the biphenyl compound 1.
In the reaction mixtures of SiCl4 with 1 and 2 plus pyridine (1aPy and 2aPy), a complete chlorine substitution is observed as in the reactions of CH3SiCl3 with 1 and 2 plus pyridine (1bPy und 2bPy), i.e., formation of “SiO4“ and “CH3SiO3” and acceleration of the reactions. In addition, gelation also occurred. This indicates that the degree of cross-linking in the case of starting material CH3SiCl3 is not sufficient to form a rigid network. The silsesquioxane units obviously lead to fewer cross-linked and THF soluble oligomeric structures.

2.3. Gel Synthesis and Characterisation

The reaction mixtures of both silylated precursors 1 and 2 with SiCl4 and pyridine (1aPy and 2aPy) formed gels within 24 hours. Over several days, strong syneresis was observed. After three weeks, the gels were dried in vacuum. The obtained xerogels 1A and 2A were brittle and glass-like; 1A was white and 2A was slightly brownish. Both xerogels were analyzed by elemental analysis, powder X-ray diffraction, and IR and solid state 29Si NMR spectroscopy. The latter spectroscopic investigations were also used to characterize the hydrolytic sensitivity of xerogels 1A and 2A.
EDX studies (performed during SEM analyses, vide infra) of both xerogels showed the presence of the elements Si, O, and C, as expected. A determination of the chlorine content indicated minor contents of 1.15 wt% in case of 1A and 0.89 wt% in case of 2A. This corresponds well with the 29Si NMR studies in solution (see Section 2.2), where a complete substitution of all three (in case of CH3SiCl3) and all four (in case of SiCl4) chlorine atoms was observed for the pyridine catalyzed reactions.
The powder X-ray diffraction data indicated completely amorphous structures for xerogels 1A and 2A. In the vibrational spectra, all typical absorption bands for the characteristic C–H, C–C, C–O, Si–O, and Si–C valence and deformation vibrations were present (see materials and methods section).
In Figure 3 and Figure 4, the solid state 29Si NMR spectra of both xerogels 1A and 2A are shown. Despite the fact that stoichiometric amounts of the starting materials SiCl4 and 1 or 2 were used, i.e., a molar ratio of 1:2, relatively strong signals for unreacted or partly reacted O–Si(CH3)3-groups in 1 (around 20 ppm), and some signals for potential side products with OSi(CH3)3-groups (signals in the range 10–16 ppm)), were found. The signals for the expected SiO4-moieties in the xerogels are clearly detectable (around −100 ppm). As the spectrum essentially consists of these two groups of signals, we conclude that moieties of partial substitution (ClSi(O-aryl)3, expected in the 29Si NMR shift range (−80 to −85 ppm) are almost absent. In Figure 3, it can be seen that after storing the powdered xerogel 1A for two days in ambient air, new upfield shifted signals emerge for both groups of signals; we attribute them to (aryl-O)3Si–O–Si-groups. The spectrum of xerogel 1A (in contrast to xerogel 2A) already shows a sharp signal at 16.1 ppm before contact with air, which is attributed to adsorbed (CH3)3SiOH. After the contact with humidity from air, the unreacted Si(CH3)3)-groups in both xerogels were partially hydrolysed and bonded to the Si–O network of the xerogel. The backbone of xerogel 1A is obviously more prone to hydrolysis when compared to that of xerogel 2A, because after two days of storage with air contact, an upfield signal around −110 ppm is present in the spectrum of sample 1A (Figure 3), but an analogous signal is less pronounced in the spectrum of 2A upon the same treatment (Figure 4, bottom).
This higher sensitivity toward hydrolysis of 1A may be caused by the fact that the biphenyl units provide more flexibility and more accessibility for water molecules due to the torsional freedom between both phenylene groups, while the naphthalene fragment is essentially rigid.

2.4. Morphological and Thermal Characterization of the Xerogels

The phenomenology of the Sol–Gel transition observed for the hybrid Sol–Gel system presented here is very similar to typical oxide Sol–Gel processes: a homogenous and transparent liquid reaction mixture solidifies—after a certain reaction time of typically hours to days—with formation of a shape retaining monolithic gel body that is usually translucent for visible light. In Figure 5, the appearance of gels of the system 1aPy is depicted.
The morphology of the xerogels was investigated using scanning electron microscopy (SEM). The image in Figure 6 shows that xerogel 1A is constituted of nanostructured primary particles with diameters of less than 50 nm. These primary particles are very strongly agglomerated, fused, or even coalesced.
Xerogel 2A is characterized by a very similar microstructure, with primary particle sizes in the range below 50 nm, as can be seen in Figure 7. The degree of agglomeration is very similar to that of xerogel 1A. The particles are almost fully coalesced to a dense homogeneous body, with only very little inter-particle spaces and corresponding porosity.
Nitrogen adsorption measurements, analyzed according to the Brunauer–Emmett–Teller (BET) model, indicated that xerogel 2A has a surface area of approximately 10 m2⋅g−1. This relatively low value corresponds very well with the SEM images, which show that most of the porosity and surface area (which was most likely present in the wet gels) strongly decreases during ageing and drying of the gels. This is also supported by the large linear shrinkage of 57% for xerogel 1A and 59% for xerogel 2A.
The thermal behavior of the xerogels was investigated under argon atmosphere applying a heating rate of 5 K⋅min−1 up to 800 °C. Further TG-FTIR studies were performed in the temperature range up to 750 °C, again with a heating rate of 5 K⋅min−1 under argon. The results of the measurements are shown in Figure 8 and Figure 9.
The TG/DTG curves of xerogels 1A and 2A are similar, showing three peaks in the comparable temperature ranges (1A: ca. 220 °C, 520 °C, and 600 °C; 2A: ca. 150 °C, 520 °C, and 600 °C). For both xerogels, the mass loss below 250 °C can be attributed at least in part to the desorption of residual THF and pyridine. This was supported by FTIR spectra of the evolved gases. Desorption of volatile species containing –O–Si(CH3)3 groups and also condensation reactions of –O–Si(CH3)3 terminal groups also occurs. Hexamethyldisiloxane and trimethylsilanol were clearly detectable in the FTIR spectra. At temperatures around 500 °C and 600 °C, decomposition reactions of aromatic C–H and remaining trimethylsilyl units led to the formation of methane for both xerogels.
In the case of xerogel 1A, the total mass loss in the temperature range from 20 °C to 800 °C is 68.2%; xerogel 2A displays a lower mass loss of 57.3% in the same temperature range. Although the pyrolytic residues were not analyzed, it is expected that Si/C/O/(H) materials are formed at 800 °C. Thus, a slightly higher thermal stability and a higher ceramic yield is found for xerogel 2A; this may be attributed to a higher degree of “graphitization” and, in turn, a higher thermal stability of the naphthalene moieties when compared to the biphenylene units.

3. Conclusions

Here we reported on a non-hydrolytic Sol–Gel system for “organically bridged” or “linearly expanded” silica and silsesquisiloxane hybrid materials. The synthesis is based on reactions of SiCl4 and CH3SiCl3 with rigid aromatic diols, which are two-fold trimethysilylated, namely compounds 1 and 2. The driving force for these reactions is the formation of (CH3)3SiCl.
The reactions proceed smoothly at room temperature and are accelerated by the addition of catalytic amounts of pyridine. Most importantly, pyridine addition also causes a complete substitution of all three or four chlorine atoms in the stoichiometric reaction mixtures with CH3SiCl3 and SiCl4, respectively, even at room temperature.
A Sol–Gel transition is observed only in case of reactions of SiCl4 in the presence of pyridine. The gels exhibit strong shrinkage, which leads to rather dense, nearly non-porous xerogels 1A and 2A, both of which are hydrolytically sensitive.
The route to organic-inorganic hybrid polymers and gels presented here may be applied to many related systems. We expect that other chlorosilanes with at least three chlorine atoms as well as other element halides with similar reactivity may be combined in suitable inert solvents with trimethylsilylated di-, tri-, or polyols to form (CH3)3SiCl or (CH3)3SiX and more-or-less branched and cross-linked network structures. The products may be calcined to form amorphous organic-inorganic hybrids with tailorable elemental composition and properties.
The hydrolytic sensitivity of as prepared xerogels may not necessarily be a disadvantage. This property might be exploitable for the development of environmentally degradable materials or for a targeted release of incorporated components triggered by contact with water.

4. Materials and Methods

General Techniques and Chemicals. All manipulations were performed under an inert (Ar or N2) atmosphere using standard Schlenk techniques or in a glovebox. The solvents, tetrahydrofuran (THF), and n-hexane were dried by distillation from a sodium/benzophenone mixture. The remaining starting materials, i.e., 4,4′-dihydroxybiphenyl, 2,6-dihydroxynaphthalene, hexamethyldisilazane (HMDS), potassium tert-butanolate (KOtBu), as well as the chlorosilanes SiCl4 and CH3SiCl3, were used as received without further purification.
Synthesis of the starting materials. 4,4′-Bis(trimethylsiloxy)biphenyl (1): The compound was synthesized according to a procedure reported in the literature [65]. However, the brownish product was impure and had to be purified. This was performed by extracting 45.0 g (136 mmol) of the raw product with 50 mL boiling n-hexane. In the extract solution, white crystals of the product formed at room temperature. After filtration and drying under vacuum, 40.9 g (124 mmol) analytically pure 1 were obtained. Mp 64 °C. NMR: 1H (400 MHz, CDCl3) δ [ppm]: 7.42 (m, 4H, 3JHH = 8.4 Hz, H2,2′), 6.88 (m, 4H, 3JHH = 8.4 Hz, H3,3′), 0.31 (s, 18H, CH3); 13C (100 MHz, CDCl3) δ [ppm]: 154.3 (C4, 4′), 134.2 (C1,1′), 127.7 (C2, 2′), 120.2 C3,3′), 0.2 (CH3); 29Si (79.5 MHz, CDCl3) δ [ppm]: 19.7 (s). IR (KBr pellet), ν [cm−1]: 3035–2960 (w) [aromatic C–H valence vibration]; 2900 (w) [C–H valence vibration of CH3-groups]; 1500 (m), 1600 (m) [aromatic C=C valence vibration]; 1263–1248 (s) [C–O valence vibration]; 1260–1250 (s), 843–817 (s), 760–752 (m) [Si(CH3)3 valence vibration]; 1174 (w), 1102 (w), 936–919 (s) [Si–O–C(aryl) valence vibration]. Elemental analysis calcd (%) for C18H26O2Si2 (M = 330.56 g/mol): C 65.40, H 7.93, found: C 65.35, H 8.09.
2,6-Bis(trimethylsiloxy)naphthalene (2): In a round bottomed flask, 5.0 g (31 mmol) 2,6-dihydroxynaphthalene was mixed with 18 mL (86 mmol) of hexamethyldisilazane (HMDS) and a catalytic amount (~50 mg) of potassium tert-butanolate. The reaction mixture was heated under reflux with vigorous stirring for 1 h. The end of the reaction was indicated when the ammonia formation has ceased. Isolation of the product was accomplished using fractioned distillation. After removal of excessive HMDS, pure 2 was collected at 130–134 °C/1.3–2.6 × 10−2 Pa as a colorless liquid. Upon cooling to room temperature, the distillate formed a white crystalline solid. Yield: 9.2 g (30 mmol, 97%). Mp 109 °C. NMR: 1H (400 MHz, CDCl3) δ [ppm]: 7.58 (d, 2H, 3JHH = 8.8 Hz, H4,8), 7.15 (d, 2H, 4JHH = 1.8 Hz, H1,5), 7.03 (dd, 2H, 3JHH = 8.8 Hz, 4JHH = 1.8 Hz, H3,7), 0.3 (s, 18H, CH3); 13C (100 MHz, CDCl3) δ [ppm]: 151.5 (C2), 130.3 (Cq), 128.0 (C4), 122.3 (C3), 114.9 (C1), 0.3 (CH3); 29Si (79.5 MHz, CDCl3) δ [ppm]: 19.9 (s). IR (KBr pellet), ν [cm−1]: 3050 (w), 2955 (w) [aromatic C-H valence vibration]; 2900 (w) [C–H valence vibration of CH3-groups]; 1600 (m), 1500 (m) [aromatic C=C valence vibration]; 1390–1375 (m) [CH3-deformation vibration]; 1270–1250 (s), 1121 (m), 1160 (m) [C–O valence vibration]; 1261–1244 (s), 874–847 (vs), 811 (w), 761 (m) [Si(CH3)3 valence vibration]; 966–958 (m) [Si–O–C(aryl) valence vibration]. Elemental analysis calcd (%) for C16H24O2Si2 (M = 304.53 g/mol): C 63.11, H 7.94, found: C 62.85, H 8.18. A single-crystal X-ray structure analysis of 2 was reported earlier [66] and provided further detailed information on this compound.
Polymer and Gel Synthesis. The polymer synthesis was performed analogous to the preparation of s-triazine- and s-heptazine-based hybrid materials, as described in [54]. In a typical experiment, 1 mmol of 1 was dissolved in 5 mL of dry THF and 0.5 mmol of SiCl4 was added at room temperature in a Schlenk tube. For comparison, the same experiment was conducted with 0.1 mmol of pyridine added as a catalyst. The liquid homogeneous reaction mixtures were stored at room temperature or heated to 60 °C. For the synthesis of less cross-linked polymers, 0.33 mmol of methyltrichlorosilane and 0.5 mmol of 1 were dissolved in 4 ml of dry THF, in absence and presence of 0.05 mmol of pyridine, and stored at room temperature or heated to 60 °C. Syntheses of polymers derived from the naphthalene derivative 2 were conducted analogously. Further details are summarized in Table 2 and Table 3.
Gelation was observed only in the cases of the pyridine catalyzed experiments with SiCl4. After an aging period of 21 days at 60 °C, the gels were carefully dried in vacuum. The obtained xerogels 1A and 2A consisted of glass-like bulk xerogel bodies. Yield and linear shrinkage were determined followed by characterization with powder XRD, gas adsorption (BET), elemental analysis, simultaneous thermal analysis (TG/DTA), and IR and 29Si CP/MAS NMR spectroscopy:
1A: Yield = 173 mg of a white, solid xerogel. NMR (CP/MAS at 5 kHz) δ [ppm]: 29Si (79.5 MHz): −100.2 (broad, SiO4-units), further sharp signals at 19.4, 17.4 and 16.1 (O–Si(CH3)3 groups). IR (KBr-pellet), ν [cm−1]: 2960(vw) [aromatic C–H valence vibration]; 1600 (w), 1500 (w) [aromatic C=C valence vibration]; 1250 (m) [C–O valence vibration]; 1251 (m), 845 (w), 821 (w) [Si(CH3)3 valence vibration]; 1169 (w), 1107 (w), 936 (w), 913 (w) [Si-O-Ar valence vibration]. Elemental analysis calcd (%) for [C24H16O4Si]n (M = 396.45 g/mol): Cl 0.0., Si 7.08, found: Cl 1.15, Si 8.79.
2A: Yield = 153 mg of a white and very rigid xerogel. NMR (CP/MAS at 5 kHz) δ [ppm]: 29Si (79.5 MHz): −100.1 (broad, SiO4-units), 19.0 (sharp, O–Si(CH3)3 groups). IR (KBr-pellet), ν [cm−1]: 3065 (vw), 2958 (vw) [aromatic C–H valence vibration]; 1600 (m), 1505 (w) [aromatic C=C valence vibration]; 1375 (m) [CH3-deformation vibration]; 1230 (m), 1119 (w), 1150 (w) [C–O valence vibration]; 870 (w), 802 (w), 750 (vw) [Si(CH3)3 valence vibration]; 983 (m) [Si–O–Ar valence vibration]. Elemental analysis calcd (%) for [C20H12O4Si]n (M = 344.38 g/mol): Cl 0.0., Si 8.2, found: Cl 0.89, Si 10.8.
Methods of Characterization. All manipulations of gels, xerogels, and pyrolysis products were performed under inert gas using glove box and/or Schlenk techniques.
IR Spectra were recorded using the standard KBr pellet method. For the starting materials, a Nicolet 510 FTIR spectrometer was used. The IR investigation of the xerogels was performed on a Varian FTIR spectrometer.
Solution 1H, 13C and 29Si NMR spectra were recorded on a BRUKER DPX 400 instrument at room temperature (TMS as internal standard). The 29Si NMR spectra were measured using inverse gated decoupling (IGATED).
Solid State 29Si NMR data of the dried gels 1A and 2A were acquired on a BRUKER AvanceTM WB 400 MHz spectrometer with a resonance frequency of 79.51 MHz using magic-angle spinning (MAS) at 5 kHz in 7 mm ZrO2 rotors and cross polarization (CP) with 5 ms contact time. The chemical shift is referenced externally to octakis(trimethylsiloxy)octasilsesquioxane Q8M8 (the most upfield signal of its Q4 groups at δiso = −109 ppm) and reported relative to tetramethylsilane (TMS).
Powder X-ray Diffraction (XRD): The patterns of the xerogels were obtained with a Siemens D-5000 diffractometer with Cu Kα radiation (λ = 1.5418 Å).
Single-crystal X-ray diffraction analysis of 2: When the freshly vacuum-distilled 2,6-bis(trimethylsiloxy)naphthalene had solidified, a tiny piece of the solid was found to be suitable for X-ray structure analysis. X-ray diffraction data were collected on a BRUKER-NONIUS X8 APEXII-CCD diffractometer with Mo-Kα-radiation (λ = 0.71073 Å). The structure was solved with direct methods and refined with full-matrix least-squares methods. All non-hydrogen atoms were refined anisotropically. Hydrogen atoms were placed in idealized positions and refined isotropically (riding model). Structure solution and refinement of F2 against all reflections were performed with SHELXS-97 and SHELXL-97 (G.M. Sheldrick, Universität Göttingen, Göttingen, Germany, 1986–1997).
2,6-bis(trimethylsiloxy)naphthalene: C16H24O2Si2, Mr = 304.53, colourless piece, 0.45 × 0.40 × 0.06 mm3, trigonal space group R-3, a = b = 21.0302(4), c = 10.5261(4) Å, α = β = 90°, γ = 120°, V = 4031.67(19) Å3, Z = 9, ρcalcd. = 1.129 g⋅cm−3, 2θmax = 60°, F(000) = 1476, μ = 0.197 mm−1, no absorption correction, T = 203(2) K, 24,421 recorded reflections (−29 ≤ h ≤ 29, −29 ≤ k ≤ 27, −14 ≤ l ≤ 14), 2608 unique and 2026 observed reflections with Fo > 2 σ(Fo), 91 parameters, R1 = 0.0401, wR2 = 0.1083 [I > 2σ(I)], R1 = 0.0578, wR2 = 0.1179 (all data), residual electron density (highest peak and deepest hole) 0.410 and −0.341 eÅ−3.
Crystallographic data (excluding structure factors) for the structure reported in this paper have been deposited earlier with the Cambridge Crystallographic Centre as private communication CCDC 622212. These data can be obtained free of charge from the Cambridge Crystallographic Data Centre via https://www.ccdc.cam.ac.uk/structures/ (accessed on 27 February 2023).
High Resolution Scanning Electron Microscopy (HR-SEM) was performed with an LEO1530 microscope equipped with an energy dispersive X-ray (EDX) system.
Nitrogen Adsorption Measurements (Micromeritics ASAP 2000) were carried out after desorption at 150 °C/~0.5 mbar for 12 h of samples (~0.5 g) which were ground in a mortar under inert conditions.
Thermal Analysis was first performed using a Seiko Instruments TG/DTA 22 under argon atmosphere applying a heating rate of 5K⋅min−1. Further TG-FTIR studies were performed on a Setaram Sensys TG-DSC coupled with a Varian FTIR instrument in the temperature range up to 750 °C with a heating rate of 5 K⋅min−1 under argon. Alumina crucibles were used.
Elemental Analyses of the starting materials 1 and 2 were carried out on a “Foss Heraeus CHN-O-Rapid” analyzer while the determination of chlorine and silicon contents in 1A and 2A was performed by the “Mikroanalytisches Labor Pascher” (Remagen-Bandorf, Germany).

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/gels9040291/s1, Figure S1: 29Si solution NMR spectrum of 1b (1 with CH3SiCl3) after 1 day at room temperature; Figure S2: 29Si solution NMR spectra of reaction mixture 1b (1/CH3SiCl3/THF), top: after 1 day; center: after 8 days; bottom: after 30 days at room temperature; Figure S3: 29Si solution NMR spectra of reaction mixture 2b (2/CH3SiCl3/THF), top: after 1 day; center: after 8 days; bottom: after 30 days at room temperature; Figure S4: 29Si solution NMR spectra of reaction mixture 2bPy (2/CH3SiCl3/pyridine/THF), top: after 9 h; center: after 6 days; bottom: after 27 days at room temperature; Figure S5: 29Si solution NMR spectra of reaction mixtures 1a (1/SiCl4/pyridine/THF) [top: after 9 h (with broad absorption band for the glas of the NMR tube); bottom: after 7 days]; Figure S6: 29Si solution NMR spectra of reaction mixtures 2a (2/SiCl4/pyridine/THF) [top: after 18 h (with broad absorption band for the glas of the NMR tube); bottom: after 7 days]; Figure S7: Molecular structure of 2 in the crystal (thermal displacement ellipsoids plotted at the 50% probability level, H-atoms are omitted for clarity). The bond C1–C1* of the molecule is located on a crystallographic center of inversion, the atoms of the asymmetric unit are labeled, the asterisk * indicates a symmetry equivalent position. Figure S8: Xerogel 1A after drying at 60 °C in vacuum for several hours. Table S1: Bond lengths [Å] of compound 2 (in its crystal structure); Table S2: Bond angles [deg.] of compound 2 (in its crystal structure); Table S3: Torsion angles [deg.] of compound 2 (in its crystal structure).

Author Contributions

Conceptualization, E.K. and J.W.; methodology, K.K., J.W., E.B. and E.K.; validation, E.K. and J.W.; investigation, K.K. (synthesis, analysis), J.W. (X-ray structures) and E.B. (solid state NMR studies); resources, E.K.; data curation, J.W.; writing, E.K.; writing—review and editing, J.W. and E.B.; visualization, K.K. and J.W.; supervision, E.K. and J.W.; All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by TU Bergakademie Freiberg (Saxony, Germany).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors gratefully acknowledge financial support by Deutsche Forschungsgemeinschaft (DFG, Bonn). Beate Kutzner, Conny Wiltzsch, and Florian Hoffmann are acknowledged for performing the solution NMR measurements. Marcus Schwarz is acknowledged for performing the REM/EDX measurements.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Judeinstein, P.; Sanchez, C. Hybrid organic–inorganic materials: A land of multidisciplinarity. J. Mater. Chem. 1996, 6, 511–525. [Google Scholar] [CrossRef]
  2. Kickelbick, G. Hybrid Materials: Synthesis, Characterization, and Applications; Wiley-VCH: Weinheim, Germany, 2006. [Google Scholar]
  3. Faustini, M.; Nicole, L.; Ruiz-Hitzky, E.; Sanchez, C. History of Organic–Inorganic Hybrid Materials: Prehistory, Art, Science, and Advanced Applications. Adv. Funct. Mater. 2018, 28, 1704158. [Google Scholar] [CrossRef]
  4. Brinker, C.J.; Scherer, G.W. Sol–Gel Science; Academic Press: San Diego, CA, USA, 1990. [Google Scholar]
  5. Hench, L.L.; West, J.K. The Sol–Gel Process. Chem. Rev. 1990, 90, 33–72. [Google Scholar] [CrossRef]
  6. Corriu, R.J.P.; Leclerq, D. Recent developments of molecular chemistry for Sol–Gel processes. Angew. Chem. Int. Ed. 1996, 35, 1420–1436. [Google Scholar] [CrossRef]
  7. Sakka, S. (Ed.) Sol–Gel Science and Technology; 4 volume set; Springer: New York, NY, USA, 2003. [Google Scholar]
  8. Vioux, A. Nonhydrolytic Sol−Gel Routes to Oxides. Chem. Mater. 1997, 9, 2292–2299. [Google Scholar] [CrossRef]
  9. Hay, J.N.; Raval, H.M. Synthesis of Organic−Inorganic Hybrids via the Non-hydrolytic Sol−Gel Process. Chem. Mater. 2001, 13, 3396–3403. [Google Scholar] [CrossRef]
  10. Debecker, D.P.; Mutin, P.H. Non-hydrolytic sol–gel routes to heterogeneous catalysts. Chem. Soc. Rev. 2012, 41, 3624–3650. [Google Scholar] [CrossRef]
  11. Smeets, V.; Styskalik, A.; Debecker, D.P. Non-hydrolytic sol–gel as a versatile route for the preparation of hybrid heterogeneous catalysts. J. Sol–Gel Sci. Technol. 2021, 97, 505–522. [Google Scholar] [CrossRef]
  12. Kroke, E. Novel Sol–Gel Routes to Non-Oxide Ceramics. In 9th CIMTEC—World Ceramics Congress, Ceramics: Getting into the 2000‘s—Part C; Vincenzini, P., Ed.; Techna: Faenza, Italy, 1999; pp. 123–134. [Google Scholar]
  13. Hector, A.L. Synthesis and processing of silicon nitride and related materials using preceramic polymer and non-oxide Sol–Gel approaches. Coord. Chem. Rev. 2016, 323, 120–137. [Google Scholar] [CrossRef] [Green Version]
  14. Rex, A.; dos Santos, J.H.Z. The use of Sol–Gel processes in the development of supported catalysts. J. Sol–Gel Sci. Technol. 2023, 105, 30–49. [Google Scholar] [CrossRef]
  15. Almeida, R.M.; Martucci, A.; Santos, L.; Rojas-Hernandez, R.E. (Eds.) Sol–Gel Derived Optical and Photonic Materials; Elsevier Ltd.: London, UK, 2021. [Google Scholar]
  16. Aegerter, M.A.; Leventis, N.; Koebel, M.M. (Eds.) Aerogels Handbook; Springer: New York, NY, USA, 2011. [Google Scholar]
  17. Hu, P.; Liu, L.; Zhao, M.; Wang, J.; Ma, X.; Wang, J. Review on design, synthesis, and use of high temperature resistant aerogels exceeding 800 °C. ES Mater. Manuf. 2022, 15, 14–33. [Google Scholar]
  18. Eychmueller, A. Nanoparticle-Based Aerogels and Their Prospective Future Applications. J. Phys. Chem. C 2022, 126, 19011–19023. [Google Scholar] [CrossRef]
  19. Gerrard, W.; Jones, J.V. Stability of isomeric butoxysilanes with respect to silicon tetrachloride and hydrogen chloride. J. Chem. Soc. 1952, 1690–1693. [Google Scholar] [CrossRef]
  20. Corriu, R.J.P.; Leclercq, D.; Lefèvre, P.; Mutin, P.H.; Vioux, A. Preparation of monolithic gels from silicon halides by a non-hydrolytic Sol–Gel process. J. Non-Cryst. Solids 1992, 146, 301–303. [Google Scholar] [CrossRef]
  21. Acosta, S.; Corriu, R.J.P.; Leclerq, D.; Mutin, P.H.; Vioux, A. Novel non-hydrolytic Sol–Gel route to metal oxides. J. Sol–Gel Sci. Technol. 1994, 2, 25–28. [Google Scholar] [CrossRef]
  22. Arnal, P.; Corriu, R.J.P.; Leclerq, D.; Mutin, P.H.; Vioux, A. A Solution Chemistry Study of Nonhydrolytic Sol−Gel Routes to Titania. Chem. Mater. 1997, 9, 694–698. [Google Scholar] [CrossRef]
  23. Oakton, E.; Siddiqi, G.; Fedorov, A.; Coperet, C. Tungsten oxide by non-hydrolytic sol–gel: Effect of molecular precursor on morphology, phase and photocatalytic performance. New J. Chem. 2016, 40, 217–222. [Google Scholar] [CrossRef]
  24. Andriamananarivelo, M.; Corriu, R.J.P.; Leclerq, D.; Mutin, P.H.; Vioux, A. Mixed oxides SiO2–ZrO2 and SiO2–TiO2 by a non-hydrolytic sol–gel route. J. Mater. Chem. 1996, 6, 1665–1671. [Google Scholar] [CrossRef]
  25. Andriamananarivelo, M.; Corriu, R.J.P.; Leclerq, D.; Mutin, P.H.; Vioux, A. Nonhydrolytic Sol−Gel Process:  Aluminum Titanate Gels. Chem. Mater. 1997, 9, 1098–1102. [Google Scholar] [CrossRef]
  26. Jansen, M.; Guenther, E. Oxide gels and ceramics prepared by a nonhydrolytic sol- gel process. Chem. Mater. 1995, 7, 2110–2114. [Google Scholar] [CrossRef]
  27. Nassar, E.J.; dos Santos Pereira, P.F.; de Oliveira Nassor, E.C.; Ávila, L.R.; Ciuffi, K.J.; Calefi, P.S. Nonhydrolytic Sol–Gel synthesis and characterization of YAG. J. Mater. Sci. 2007, 42, 2244–2249. [Google Scholar] [CrossRef]
  28. Janković-Častvan, I.; Lazarević, S.; Jordović, B.; Petrović, R.; Tanasković, D.; Janaćković, D. Electrical properties of cordierite obtained by non-hydrolytic sol–gel method. J. Eur. Ceram. Soc. 2007, 27, 3659–3661. [Google Scholar] [CrossRef]
  29. Gao, Y.; Hamana, D.; Iwasaki, R.; Iihama, J.; Honda, S.; Kumari, M.; Hayakawa, T.; Bernard, S.; Thomas, P.; Iwamoto, Y. Chemical route for synthesis of β-SiAlON:Eu2+ phosphors combining polymer-derived ceramics route with non-hydrolytic Sol–Gel chemistry. J. Sol–Gel Sci. Technol. 2022, 104, 711–723. [Google Scholar] [CrossRef]
  30. Koshevaya, E.; Mikhaylov, V.; Sitnikov, P.; Krivoshapkina, E.; Krivoshapkin, P. Electrosurface properties and acid-base equilibria of Ta2O5 and Ta2O5:Eu nanoparticles in NaCl solutions. Surf. Interfaces 2022, 29, 101713. [Google Scholar] [CrossRef]
  31. Shrestha, S.; Wang, B.; Dutta, P. Nanoparticle processing: Understanding and controlling aggregation. Adv. Colloid Interface Sci. 2020, 279, 102162. [Google Scholar] [CrossRef] [PubMed]
  32. Escamilla-Perez, A.M.; Louvain, N.; Boury, B.; Brun, N.; Mutin, P.H. Ethers as Oxygen Donor and Carbon Source in Non-hydrolytic Sol–Gel: One-Pot, Atom-Economic Synthesis of Mesoporous TiO2-Carbon Nanocomposites. Chem. Eur. J. 2018, 24, 4982–4990. [Google Scholar] [CrossRef]
  33. Corriu, R.J.P.; Leclerq, D.; Lefevre, P.; Mutin, P.H.; Vioux, A. Preparation of monolithic binary oxide gels by a nonhydrolytic Sol–Gel process. Chem. Mater. 1992, 4, 961–963. [Google Scholar] [CrossRef]
  34. Iwasaki, M.; Yasumori, A.; Shibata, S.; Yamane, M. Preparation of high homogeneity BaO-TiO2-SiO2 gel. J. Sol–Gel Sci. Technol. 1994, 2, 387–391. [Google Scholar] [CrossRef]
  35. Goel, S.C.; Chiang, M.Y.; Gibbons, P.C.; Buhro, W.F. New Chemistry for the Sol–Gel Process: Acetone as a New Condensation Reagent. Mater. Res. Soc. Symp. Proc. 1992, 271, 3–13. [Google Scholar] [CrossRef]
  36. Mehrotra, R.C. Present status and future potential of the Sol–Gel process. In Chemistry, Spectroscopy and Applications of Sol–Gel Glasses; Reisfeld, R., Jorgensen, C.K., Eds.; Springer: Berlin, Germany, 1992; p. 20. [Google Scholar]
  37. Guiton, T.A.; Czekaj, C.L.; Pantano, C.G. Organometallic Sol/Gel Chemistry of Metal Sulfides. J. Non-Cryst. Solids 1990, 121, 7–15. [Google Scholar] [CrossRef]
  38. Gacion, T.; Malier, L.; Boilot, J.-P. New Transparent Chalcogenide Materials Using a Sol–Gel Process. Chem. Mater. 1997, 9, 1502–1504. [Google Scholar]
  39. Ptatschek, V.; Schreder, B.; Herz, K.; Hilbert, U.; Ossau, W.; Schottner, G.; Rahäuser, O.; Bischof, T.; Lermann, G.; Materny, A.; et al. Sol–Gel Synthesis and Spectroscopic Properties of Thick Nanocrystalline CdSe Films. J. Phys. Chem. B 1997, 101, 8898–8906. [Google Scholar] [CrossRef]
  40. Rusch, P.; Luebkemann, F.; Borg, H.; Eckert, J.G.; Dorfs, D.; Bigall, N.C. Influencing the coupling between network building blocks in CdSe/CdS dot/rod aerogels by partial cation exchange. J. Chem. Phys. 2022, 156, 234701. [Google Scholar] [CrossRef]
  41. Seddon, A.B.; Hodgson, S.N.B.; Scott, M.G. Sol–Gel approach to preparing germanium disulphide. J. Mater. Sci. 1991, 26, 2599–2602. [Google Scholar] [CrossRef]
  42. Leidich, S.; Buechele, D.; Lauenstein, R.; Kluenker, M.; Lind, C. “Non-hydrolytic” sol–gel synthesis of molybdenum sulfides. J. Solid State Chem. 2016, 242, 175–181. [Google Scholar] [CrossRef] [Green Version]
  43. Gerwig, M.; Schwarz, M.; Brendler, E.; Kroke, E. [SiSxHy]n—Perhydridopolysilathianes—Cross-linked thio-analogs of polysiloxanes. Eur. J. Inorg. Chem. 2016, 2016, 5028–5035. [Google Scholar] [CrossRef]
  44. Rovai, R.; Lehmann, C.W.; Bradley, J.S. Non-Oxide Sol–Gel Chemistry: Preparation from Tris(dialkylamino)silazanes of a Carbon-Free, Porous, Silicon Diimide Gel. Angew. Chem. Int. Ed. 1999, 38, 2036–2038. [Google Scholar] [CrossRef]
  45. Kitney, S.P.; Sajedin, S.M.; Rocher, V.; Cheng, F.; Kelly, S.M. Silicon diimide gel as an efficient stationary phase in thin layer chromatography for acid-sensitive organic compounds. Chem. Commun. 2017, 53, 11080–11082. [Google Scholar] [CrossRef]
  46. Paine, R.T. Processing of boron-nitrogen preceramic polymers. J. Inorg. Organomet. Polym. 1992, 2, 183–195. [Google Scholar] [CrossRef] [Green Version]
  47. Jäger, L.; Köhler, H. Pseudochalkogene. Sulfur Reports 1992, 12, 159–203. [Google Scholar] [CrossRef]
  48. Riedel, R.; Kroke, E.; Greiner, A.; Gabriel, A.O.; Ruwisch, L.; Nicolich, J.; Kroll, P. Inorganic Solid State Chemistry with Main Group Element Carbodiimides. Chem. Mater. 1998, 10, 2964–2979. [Google Scholar] [CrossRef]
  49. Lippe, K.; Wagler, J.; Kroke, E.; Herkenhoff, S.; Ischenko, V.; Woltersdorf, J. Cyclic Silylcarbodiimides as Precursors for Porous Si/C/N Materials: Formation, Structures, and Stabilities. Chem. Mater. 2009, 21, 3941–3949. [Google Scholar] [CrossRef]
  50. Cheng, H.; Li, Y.; Kroke, E.; Herkenhoff, S. In situ synthesis of Si2N2O/Si3N4 composite ceramics using polysilyloxycarbodiimide precursors. J. Eur. Ceram. Soc. 2013, 33, 2181–2189. [Google Scholar] [CrossRef]
  51. Voelger, K.W.; Hauser, R.; Kroke, E.; Riedel, R.; Ikuhara, Y.H.; Iwamoto, Y. Synthesis and characterization of novel non-oxide Sol–Gel derived mesoporous amorphous Si-CN membranes. J. Ceram. Soc. Japan 2006, 114, 567–570. [Google Scholar] [CrossRef] [Green Version]
  52. Krupinski, K.; Brendler, E.; Gericke, R.; Wagler, J.; Kroke, E. A new aspect of the “pseudo water” concept of bis(trimethylsilyl)carbodiimide—“pseudohydrates” of aluminum. Z. Naturforsch. 2018, B73, 911–918. [Google Scholar] [CrossRef]
  53. Kroke, E.; Völger, K.W.; Klonczynski, A.; Riedel, R. A Sol–Gel Route to B4C. Angew. Chem. Int. Ed. 2001, 40, 1698–1700. [Google Scholar] [CrossRef]
  54. El-Gamel, N.E.A.; Schwarz, M.; Brendler, E.; Kroke, E. s-Triazine and tri-s-triazine based organic-inorganic hybrid gels prepared from chlorosilanes by exchange reactions. Chem. Commun. 2006, 4741–4743. [Google Scholar] [CrossRef]
  55. Mahadik, D.B.; Wang, Q.; Meti, P.; Lee, K.-Y.; Gong, Y.-D.; Park, H.-H. Synthesis of multi-functional porous superhydrophobic trioxybenzene cross-linked silica aerogels with improved textural properties. Ceram. Internat. 2020, 46, 17969–17977. [Google Scholar] [CrossRef]
  56. Wang, Q.; Mahadik, D.B.; Meti, P.; Gong, Y.-D.; Lee, K.-Y.; Park, H.-H. Dioxybenzene-bridged hydrophobic silica aerogels with enhanced textural and mechanical properties. Micropor. Mesopor. Mater. 2020, 294, 109863. [Google Scholar] [CrossRef]
  57. Small, J.H.; Shea, K.J.; Loy, D.A. Arylene- and alkylene-bridged polysilsesquioxanes. J. Non-Cryst. Solids 1993, 160, 234–246. [Google Scholar] [CrossRef]
  58. Shea, K.J.; Loy, D.A. Bridged Polysilsesquioxanes: Molecular Engineering of Hybrid Organic–Inorganic Materials. MRS Bulletin 2001, 26, 368–375. [Google Scholar] [CrossRef]
  59. Pfeifer, S.; Schwarzer, A.; Schmidt, D.; Brendler, E.; Veith, M.; Kroke, E. Precursors for pyromellit-bridged silica sol–gel hybrid materials. New J. Chem. 2013, 37, 169–180. [Google Scholar]
  60. Pfeifer, S.; Brendler, E.; Veith, M.; Kroke, E. Hybrid-Coatings Derived from Pyromellitic Acid Bridged Alkoxy-silylalkyl Precursors. J. Sol–Gel Sci. Technol. 2014, 70, 191–202. [Google Scholar] [CrossRef]
  61. Haas, K.-H. Hybrid Inorganic–Organic Polymers Based on Organically Modified Si-Alkoxides. Adv. Eng. Mater. 2000, 2, 571–582. [Google Scholar] [CrossRef]
  62. Croissant, J.G.; Cattoen, X.; Durand, J.-O.; Wong Chi Man, M.; Khashab, N.M. Organosilica hybrid nanomaterials with a high organic content: Syntheses and applications of silsesquioxanes. Nanoscale 2016, 8, 19945–19972. [Google Scholar] [CrossRef] [PubMed]
  63. Furukawa, H.; Cordova, K.E.; O’Keeffe, M.; Yaghi, O.M. The Chemistry and Applications of Metal-Organic Frameworks. Science 2013, 341, 974. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  64. Ding, S.-Y.; Wang, W. Covalent organic frameworks (COFs). From design to applications. Chem. Soc. Rev. 2013, 42, 548–568. [Google Scholar] [CrossRef] [PubMed]
  65. Ismail, R.M. Silylierung mit Hexamethyldisilazan. Z. Naturforsch. 1963, 18b, 582. [Google Scholar] [CrossRef]
  66. Lippe, K.; Kroke, E.; Wagler, J. CSD Private Communication, 2016; CCDC 622212.
Scheme 1. Examples of non-hydrolytic oxide-based Sol–Gel strategies: (a) Generic reaction of metal alkoxide and metal halide moiety; (b) Reaction of silicon tetrachloride and tetra-n-butoxysilane; (c) Generic reaction of a metal alkoxide and an acetoxysilane.
Scheme 1. Examples of non-hydrolytic oxide-based Sol–Gel strategies: (a) Generic reaction of metal alkoxide and metal halide moiety; (b) Reaction of silicon tetrachloride and tetra-n-butoxysilane; (c) Generic reaction of a metal alkoxide and an acetoxysilane.
Gels 09 00291 sch001
Scheme 2. Examples of non-hydrolytic non-oxide-based Sol–Gel strategies: (a) Reaction of dibenzyltrisulfide and diethylzinc; (b) Reaction of a dimethylamino substituted cyclotrisilazane and ammonia.
Scheme 2. Examples of non-hydrolytic non-oxide-based Sol–Gel strategies: (a) Reaction of dibenzyltrisulfide and diethylzinc; (b) Reaction of a dimethylamino substituted cyclotrisilazane and ammonia.
Gels 09 00291 sch002
Scheme 3. Examples of non-hydrolytic Sol–Gel strategies with release of (CH3)3SiCl as condensation product: (a) Generic reaction of trichlorosilanes and bis(trimethylsilyl)carbodiimide (R =, e.g., alkyl, aryl); (b) Reaction of B-trichloroborazene and bis(trimethylsilyl)carbodiimide; (c) Generic reaction of methylchlorosilanes and three-fold silylated cyameluric or cyanuric acid.
Scheme 3. Examples of non-hydrolytic Sol–Gel strategies with release of (CH3)3SiCl as condensation product: (a) Generic reaction of trichlorosilanes and bis(trimethylsilyl)carbodiimide (R =, e.g., alkyl, aryl); (b) Reaction of B-trichloroborazene and bis(trimethylsilyl)carbodiimide; (c) Generic reaction of methylchlorosilanes and three-fold silylated cyameluric or cyanuric acid.
Gels 09 00291 sch003
Figure 1. Approaches to “expanded oxides”: (a) The most frequently prepared hybrid materials which may be described as organically bridged silsesquioxanes; (b) MOFs (metal organic frameworks) can be considered as “organically expanded” coordination centers; (c) Replacing the oxygen atoms in Si–O compounds by carbodiimide units results in non-oxides, which are “expanded” compared to the oxide analogues; (d) “Trigonally expanded silicas” as prepared earlier; (e) “Linearly expanded silicas” as presented here.
Figure 1. Approaches to “expanded oxides”: (a) The most frequently prepared hybrid materials which may be described as organically bridged silsesquioxanes; (b) MOFs (metal organic frameworks) can be considered as “organically expanded” coordination centers; (c) Replacing the oxygen atoms in Si–O compounds by carbodiimide units results in non-oxides, which are “expanded” compared to the oxide analogues; (d) “Trigonally expanded silicas” as prepared earlier; (e) “Linearly expanded silicas” as presented here.
Gels 09 00291 g001
Figure 2. Molecular structure of 2 in the crystal (thermal displacement ellipsoids plotted at the 50% probability level). The bond C1–C1* of the molecule is located on a crystallographic center of inversion, the non-hydrogen atoms of the asymmetric unit are labeled, the asterisk * indicates a symmetry equivalent position. For clarity reasons hydrogen atoms (white balls) are not labeled.
Figure 2. Molecular structure of 2 in the crystal (thermal displacement ellipsoids plotted at the 50% probability level). The bond C1–C1* of the molecule is located on a crystallographic center of inversion, the non-hydrogen atoms of the asymmetric unit are labeled, the asterisk * indicates a symmetry equivalent position. For clarity reasons hydrogen atoms (white balls) are not labeled.
Gels 09 00291 g002
Figure 3. 29Si cross-polarization magic angle spinning (CP MAS) NMR spectra of xerogel 1A (obtained from 1/SiCl4/pyridine/THF), top (red trace): xerogel prepared under argon; bottom (blue trace): same sample of xerogel after two days in contact with ambient air. (The vertical lines at δ = +20 ppm and −100 ppm aid enhanced visualization of the chemical shift differences).
Figure 3. 29Si cross-polarization magic angle spinning (CP MAS) NMR spectra of xerogel 1A (obtained from 1/SiCl4/pyridine/THF), top (red trace): xerogel prepared under argon; bottom (blue trace): same sample of xerogel after two days in contact with ambient air. (The vertical lines at δ = +20 ppm and −100 ppm aid enhanced visualization of the chemical shift differences).
Gels 09 00291 g003
Figure 4. 29Si CP MAS NMR spectra of xerogel 2A (obtained from 2/SiCl4/pyridine/THF), top (red trace): xerogel prepared under argon; bottom (blue trace): same sample of xerogel after two days in contact with ambient air. (The vertical lines at δ = +20 ppm and −100 ppm aid enhanced visualization of the chemical shift differences).
Figure 4. 29Si CP MAS NMR spectra of xerogel 2A (obtained from 2/SiCl4/pyridine/THF), top (red trace): xerogel prepared under argon; bottom (blue trace): same sample of xerogel after two days in contact with ambient air. (The vertical lines at δ = +20 ppm and −100 ppm aid enhanced visualization of the chemical shift differences).
Gels 09 00291 g004
Figure 5. Reaction mixture 1aPy and formation of gel 1A: (a) at t = 0 h, (b) after one day gelation and partial syneresis already occurred, (c) aged for 21 days.
Figure 5. Reaction mixture 1aPy and formation of gel 1A: (a) at t = 0 h, (b) after one day gelation and partial syneresis already occurred, (c) aged for 21 days.
Gels 09 00291 g005
Figure 6. SEM image of xerogel 1A (obtained from 1/SiCl4/pyridine/THF).
Figure 6. SEM image of xerogel 1A (obtained from 1/SiCl4/pyridine/THF).
Gels 09 00291 g006
Figure 7. SEM image of xerogel 2A (obtained from 2/SiCl4/pyridine/THF). The composition and identity of the larger crystalline-appearing particle in the upper right corner of the image could not be resolved.
Figure 7. SEM image of xerogel 2A (obtained from 2/SiCl4/pyridine/THF). The composition and identity of the larger crystalline-appearing particle in the upper right corner of the image could not be resolved.
Gels 09 00291 g007
Figure 8. TG/DTG measurement of xerogel 1A (obtained from 1/SiCl4/pyridine/THF; TG trace dark blue, DTG trace light blue).
Figure 8. TG/DTG measurement of xerogel 1A (obtained from 1/SiCl4/pyridine/THF; TG trace dark blue, DTG trace light blue).
Gels 09 00291 g008
Figure 9. TG/DTG measurement of xerogel 2A (obtained from 2/SiCl4/pyridine/THF; TG trace dark blue, DTG trace light blue).
Figure 9. TG/DTG measurement of xerogel 2A (obtained from 2/SiCl4/pyridine/THF; TG trace dark blue, DTG trace light blue).
Gels 09 00291 g009
Table 1. Selected bond lengths and angles for compound 2. Symmetry equivalent atoms are marked with an asterisk *.
Table 1. Selected bond lengths and angles for compound 2. Symmetry equivalent atoms are marked with an asterisk *.
Bond Lengths [Å] Bond Angles [°]
C1–C2 1.424(2) C1*–C1–C2 119.2(2)
C2–C3 1.364(2) C1–C2–C3 120.5(1)
C3–C4 1.415(2) C2–C3–C4 120.0(1)
C4–C5 1.364(2) C3–C4–C5 120.6(1)
C5–C1* 1.413(2) C4–C5–C1* 120.7(1)
C1–C1* 1.413(2) C5*–C1–C1* 119.0(2)
C5*–C1–C2 121.9(1)
Table 2. Overview about the reactions of precursor 1 with SiCl4 and CH3SiCl3 in THF.
Table 2. Overview about the reactions of precursor 1 with SiCl4 and CH3SiCl3 in THF.
Reaction Mixture n(1) [mmol] THF [mL]n(SiCl4) [mmol] n(CH3SiCl3) [mmol] nPyridine [mmol] Observation
1a1 5 0.5 --no gelation after three weeks at 60 °C
1aPy1 5 0.5 -0.1 gelation after 1 day; strong syneresis in following weeks, linear shrinkage of about 57% for xerogel 1A
1b0.54 -0.33 -no gelation after three weeks at 60 °C
1bPy0.5 4 -0.33 0.05
Table 3. Overview about the reactions of precursor 2 with SiCl4 and CH3SiCl3 in THF.
Table 3. Overview about the reactions of precursor 2 with SiCl4 and CH3SiCl3 in THF.
Reaction Mixture n(2) [mmol] THF [mL] n(SiCl4) [mmol] n(CH3SiCl3) [mmol] nPyridine [mmol] Observations
2a1 5 0.5 --no gelation after three weeks at 60 °C
2aPy1 5 0.5 -0.1 gelation after 1 day; strong syneresis in following weeks, linear shrinkage of 59% for xerogel 2A
2b0.5 4 -0.33 -no gelation after three weeks at 60 °C
2bPy0.5 4 -0.33 0.05
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

Krupinski, K.; Wagler, J.; Brendler, E.; Kroke, E. A Non-Hydrolytic Sol–Gel Route to Organic-Inorganic Hybrid Polymers: Linearly Expanded Silica and Silsesquioxanes. Gels 2023, 9, 291. https://doi.org/10.3390/gels9040291

AMA Style

Krupinski K, Wagler J, Brendler E, Kroke E. A Non-Hydrolytic Sol–Gel Route to Organic-Inorganic Hybrid Polymers: Linearly Expanded Silica and Silsesquioxanes. Gels. 2023; 9(4):291. https://doi.org/10.3390/gels9040291

Chicago/Turabian Style

Krupinski, Katrin, Jörg Wagler, Erica Brendler, and Edwin Kroke. 2023. "A Non-Hydrolytic Sol–Gel Route to Organic-Inorganic Hybrid Polymers: Linearly Expanded Silica and Silsesquioxanes" Gels 9, no. 4: 291. https://doi.org/10.3390/gels9040291

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

Article Metrics

Back to TopTop