Abstract
This study is focused on structural characterization of hybrid glasses obtained by consolidation of melting gels. The melting gels were prepared in molar ratios of methyltriethoxysilane (MTES) and dimethyldiethoxysilane (DMDES) of 75%MTES-25%DMDES and 65%MTES-35%DMDES. Following consolidation, the hybrid glasses were characterized using Raman, 29Si and 13C Nuclear Magnetic Resonance (NMR) spectroscopies, synchrotron Small Angle X-Ray Scattering (SAXS) and scanning electron microscopy (SEM). Raman spectroscopy revealed the presence of Si-C bonds in the hybrid glasses and 8-membered ring structures in the Si-O-Si network. Qualitative NMR spectroscopy identified the main molecular species, while quantitative NMR data showed that the ratio of trimers (T) to dimers (D) varied between 4.6 and 3.8. Two-dimensional 29Si NMR data were used to identify two distinct types of T3 environments. SAXS data showed that the glasses are homogeneous across the nm to micrometer length scales. The scattering cross section was one thousand times lower than what is expected when phase separation occurs. The SEM images show a uniform surface without defects, in agreement with the SAXS results, which further supports that the hybrid glasses are nonporous.
Keywords: Melting gels, Raman, 29Si NMR and 13C NMR spectroscopy, synchrotron small angle X-ray scattering (SAXS), ultra-small angle X-ray scattering (USAXS), Scanning Electron Microscopy (SEM)
1. Introduction
The concept of melting gels was first introduced in 2001 by Matsuda et al.1. They studied the formation of poly(benzylsilsesquioxane) particles, which were then deposited on an ITO substrate by electrophoretic deposition. After heat treatment, continuous and uniform transparent thick coatings were obtained. In reality the process called “melting” is a softening process. Melting gels are hybrid organic inorganic materials in which the presence of the covalent bonds between the organic components and inorganic network places the melting gels in the Class II of hybrids2. The precursors used for preparation of hybrid melting gels are a mixture between a mono-substituted organo-modified alkoxide such as phenyltriethoxysilane3,4 or methyltriethoxysilane5 and a di-substituted organo-modified alkoxide such as diphenyldiethoxysilane3 or dimethyldiethoxysilane5. Hybrid melting gels differ from classical hybrids through their properties. The defining property of these hybrid melting gels is that they are solid at room temperature, become fluid at a softening temperature T1 (~ 110°C), return to rigidity at room temperature, and can be re-softened many times. All the hybrid melting gels also present a consolidation temperature which varies with the composition3,5. However, after consolidation at a temperature T2 (T2>T1), the gels no longer soften. The consolidation temperature T2 corresponds to cross-linking of the silica chains5,6. At the consolidation temperature the melting gels are transformed into hybrid glasses4.
The most studied system is the phenyl based melting gel3,6–8. For this system Kakiuchida et al.6,7 found a correlation between the structure of a melting gel and its viscoelastic properties. Using 29Si MAS NMR and Gel Permeation Chromatography (GPC) they identified the complex nature of the structure of the melting gels. Furthermore, they demonstrated that the softening ability is controlled by the number of oxygen bridges and the intramolecular structure, and the viscosity follows a free volume model. Despite a growing number of studies on melting gel properties, there are only a few studies that characterize the hybrid glasses obtained from the melting gel consolidation. Masai et al.8 prepared melting gels and the corresponding low temperature hybrid glasses from phenyltriethoxysilane and diphenyldiethoxysilane. Using 29Si MAS NMR and FT-IR spectroscopy, they identified the phenomena that take place during the thermal transition from melting gels to hybrid glasses. Using the same system3 and phenyltriethoxysilane along with dimethyldiethoxysilane4 we characterized the structure of the hybrid glasses by Raman and FT-IR spectroscopy. The structure and the orientation of the organic groups was found to promote hydrophobicity to the surface of the hybrid glasses.
In our earlier works, hybrid glasses were also obtained using methyl-substituted melting gels5,9–12. Mainly the focus was on preparation5,9 using methyltriethoxysilane (MTES) and dimethyldiethoxysilane (DMDES), on their thermal stability10,11, and on their viscoelastic properties11. The hybrid glasses obtained from these melting gels were characterized by FT-IR and Raman spectroscopy5,12. These glasses, which are obtained at relatively low temperatures (135–160 °C)5, perform well as anticorrosive coatings12 and as hermetic barriers13,14 and also have low dielectric constants15. While several applications have been pursued for these hybrid glasses, an understanding of their molecular structure is still incomplete.
NMR spectroscopy is among the most powerful techniques for structural studies of the silica-based sol-gel materials. For the sol-gel method, NMR spectroscopy can be used either in solutions to elucidate the early stages of formation of the molecular species of the inorganic or hybrid materials16 or on the final material when solid state NMR can be used to elucidate the structure17.
Historically, “horizontal and vertical polymerization in the silane phase” of MTES and DMDES was first studied by Sindorf et al.18 in comparison with their chlorosilane partners using 29Si and 13C NMR by using cross polarization (CP) and magic angle spinning (MAS). It was found that while that the ethoxysilanes and chlorosilanes have similar chemistries in the silylation reactions of silica surfaces, ethoxysilanes are less reactive than their chlorosilane partners. Fyfe et al.19 studied the copolymerization of the tetraethoxysilane (TEOS) and MTES using two dimensional 1H/29Si NMR correlation, which revealed these two components mixed well in the final gel and no phase separation was induced. Another interesting study on the copolymerization of TEOS along with MTES or ethyltriethoxysilane (ETES) or octyltriethoxysilane (OTES) was carried out by Peeters et al.20. Using the CP and single-pulse excitation (SPE) MAS NMR, the effect of the length of the organic tail on the degree of condensation was studied. They observed that CP MAS NMR cannot be used to quantitatively obtain reliable data for the system with a long organic tail due to the large distance between the proton and a part of Q4 or T3 silicon atoms. For this reason, SPE MAS-NMR was required in quantitative measurements20. It was also concluded that the total number of network bonds decreases with the increasing level of substitution.
The formations of silicon-oxycarbide glasses obtained by cohydrolysis and polycondensation of the TEOS along with DMDES21,22 or triethoxysilane (TES) along with methyldiethoxysilane (MDES)23 or TEOS and MTES24 or MTES along with methyltrimethoxysilane (MTMS)25 were investigated by using 29Si, 13C and 1H MAS NMR. 29Si and 13C MAS NMR23 was used for identification of the distribution of the organic groups in the gel phases. In addition, these techniques were used to characterize the evolution between gel and silicon oxycarbide glasses23,24. Moreover, Suyal et al.24 showed that the addition of colloidal silica delayed the decomposition of the methyl groups from MTES.
Brus et al.26–28, published a series of studies upon co-hydrolysis polycondensation of TEOS with DMDES or TEOS with MTES using 1H MAS NMR26 1H and 29Si CP/MAS NMR27 and 2D 1H-29Si heteronuclear experiments. They observed that the dynamic behavior and the nature of the silanol protons in the presence of water are influenced by the amount of methyl groups in the hybrid network26. In addition, it was observed that the formation of clusters of OH protons with different strength of hydrogen bond is related to the arrangements of the siloxane units in the final network28. Based on the interatomic distances a cage-like model was proposed for the structure of the final copolymer27.
Small angle X-ray scattering (SAXS), being a premier, nondestructive technique that probes mesoscopic structure inhomogeneities, provides statistically significant characterization of the network structure within hybrid gels and glasses. SAXS is based on the elastic scattering of X-rays at scattering angles very close to the direction of the incident beam. Depending on the detailed instrumental setup, SAXS can access structural information in the reciprocal space from sub-nanometer to tens of nanometers (conventional pinhole setup) and even tens of micrometers (Bonse-Hart type setup).
As a general technique, SAXS has been used to investigate the microstructures of polymeric gels and glasses29,30. In particular, SAXS was deployed to study the static structure and formation kinetics of a broad range of silicon-based materials, ranging from nanocomposites, nanoparticles, and mesoporous silica films31–42. Notably, SAXS was combined with 29Si to reveal the condensation kinetics of hydrolyzed alkoxides where progressive assembly of small organized units were found to be followed by growth of fractal clusters in tetravalent TEOS31. This study illustrated the potency of combined NMR and SAXS characterizations, where detailed local conformations can be derived from NMR and SAXS reveal global morphological characteristics.
This work will focus on the structure characterization of the hybrid organic-inorganic glasses which contain direct bond between silicon and methyl groups, obtained after consolidation of melting gels. The consolidation of the melting gels represent a new pathway to obtain hybrid organic-inorganic glasses. We seek to provide a comprehensive and across-length-scale view of the structures by making use of Raman spectroscopy, 13C and 29Si NMR spectroscopy along with synchrotron-based SAXS and scanning electron microscopy (SEM).
2. Experimental
2.1. Melting gel / hybrid glasses preparation
The preparation of the melting gels, reported previously5, is briefly discussed below. We used two types of alkoxides, mono-substituted methyltriethoxysilane (MTES) (Sigma-Aldrich, Milwaukee, WI1) and di-substituted one dimethyldiethoxysilane (DMDES) (Sigma-Aldrich, Milwaukee, WI), without further purification. Hydrochloric acid (Fisher Scientific, Atlanta, GA) and ammonia (Sigma-Aldrich, Milwaukee, WI) were used as catalysts while anhydrous ethanol (Sigma-Aldrich, Milwaukee, WI) was used as solvent. Two melting gels presented in Table 1, with compositions of 75% MTES-25% DMDES and 65% MTES-35% DMDES (in mol %) have been investigated in this study. The synthesis was performed in three different steps with predetermined amount of reactants. The molar ratios of MTES:EtOH:H2O:HCl were 1:4:3:0.01. First, water was mixed with hydrochloric acid and half of the ethanol. Separately, MTES was mixed with the other half of ethanol. Then, the mixture between ethanol and MTES was added dropwise to the water solution under continuous stirring. The container was sealed with Parafilm® while the mixture was continuously stirred at room temperature for three hours. In the second step, the di-substituted alkoxide DMDES was diluted with ethanol in a molar ratio of DMDES:EtOH = 1:4. The DMDES-EtOH mixture was added dropwise to the first mixture. This resulting solution was kept under continuous stirring in the sealed container at room temperature for two additional hours. In the third step, ammonia was added to the mixture, which was stirred for one more hour in the sealed container. The molar ratio of (MTES+DMDES):NH3=1:0.01. The final solution was continuously stirred for 48 hours at room temperature in an open container until gelation occurred. Ammonium chloride was formed as a byproduct during gelation. To remove this, 10 ml of dry acetone (Spectranal, Sigma-Aldrich) was added to the samples. The ammonium chloride was removed by vacuum filtration. Then the clear solution was stirred until all the acetone evaporated. The obtained gel was treated at 70 °C for 24 hours to remove any remaining acetone and ethanol, followed by another heat treatment at 110 °C for 24 hours to remove of unreacted water. The hybrid glasses were obtained by consolidation of the melting gels on 15 mm × 15 mm mica substrates (Grade V-4 muscovite, SPI Co., West Chester, PA). The consolidation temperatures5,10 for each composition are a function of the molar % between MTES and DMDES and are listed in Table 1. At these temperatures the melting gels are transformed into hybrid glasses within a 17-hour consolidation heat treatment.
Table 1.
The starting compositions of the melting gels, their temperatures of consolidation and the thicknesses and densities of the hybrid glasses obtained by consolidation of the melting gels
| Sample | Composition (mol%) |
Temperature of consolidation (°C) |
Thickness of coatings (mm) |
Density (g/cm3) |
BET surface area m2/g |
|
|---|---|---|---|---|---|---|
| MTES | DMDES | |||||
| 1 | 75 | 25 | 135 | 1.284 | 0.86 | 0.0255 |
| 2 | 65 | 35 | 150 | 1.252 | 1.06 | 0.0141 |
For the NMR and BET surface area measurements the hybrid glasses were removed from the mica substrates by thermally shocking the samples. The samples were immersed in the liquid nitrogen (77K) and then very quickly exposed to the room temperature. Repeating this process a few times allowed hybrid glasses to be detached from the substrates. Subsequently, the removed hybrid glasses underwent cryogenic grinding in liquid nitrogen until fine powders were achieved.
For the SAXS and Raman measurements, the hybrid glasses samples as deposited on the mica supports were used. The thickness and density of the coatings are listed in Table 1.
2.2. Sample characterization
The consolidated hybrid glasses were characterized using Raman, 13C and 29Si NMR spectroscopy, SAXS and SEM.
Raman spectra were recorded using Renishaw® in Via Raman Microscope (Renishaw, Gloucestershire, UK) equipped with 765 nm laser between 4000 – 200 cm−1. The spectra were acquired under a Leica optical microscope at 20× magnification.
NMR 13C and 29Si quantifications and T1 measurements were conducted at 7.0 T on a Bruker spectrometer equipped with a 4 mm double resonance probehead. The MAS frequency was set at 12 kHz. T1 longitudinal relaxation time measurements were conducted with a saturation-recovery approach, using 10 saturation pulses for both 13C and 29Si experiments with a saturation train of 10 pulses spaced by delays decreasing step by step from 50 to 5 ms. Quantitative measurements were collected with an echo-MAS sequence and a short echo duration of two rotor periods. Full magnetization recovery was ensured by recycling delays at least quintupling the longest T1 measured under the same conditions. The number of scans was adjusted for each recovery delay to optimize the experimental time while maintaining good signal-to-noise. The peak intensities were extracted from sets of peak-profile parameters (positions, widths and Gaussian to Lorentzian ratios), allowing only the amplitudes to vary. These sets of parameters were obtained, for each sample and experimental condition, from a simultaneous fit of multiple spectra, including the echo-MAS spectra with the highest signal to noise and CP-MAS spectra collected with different contact times, which ensures the robustness of the spectral decomposition into individual components. Uncertainty evaluation was performed with a Monte-Carlo approach using an in-house MATLAB program. Random noise within the experimental standard deviation was introduced to the experimental data to produce a large number of datasets, which were subsequently fitted. Statistical analysis was conducted on the fitting parameters to yield the uncertainty of the measured T1 value.
Two-dimensional 29Si[29Si] dipolar double quantum (DQ) experiments were conducted on sample 2 using a 9.4 T Bruker spectrometer with a 7 mm double-resonance probe, using symmetry-based43 SR26411 recoupling44, at the MAS frequency of 5.5 kHz and a 29Si nutation frequency of 35.7 kHz (optimized for maximum recoupling efficiency). Heteronuclear 1H decoupling was achieved using CW decoupling at a 1H nutation frequency of 60 kHz during the recoupling, and SPINAL64 at 60 kHz during acquisition in both dimensions. The DQ creation and reconversion blocks used 5 recoupling supercycles each, corresponding to 7.1 ms blocks. A two-rotor-period echo was added at the end of the pulse sequence to ensure a flat baseline. The indirect dimension was collected with 48 t1 increments using the States procedure45, with 7424 scans and a recycle delay of 1.5 s (total duration: 6 days). The spectral width in the indirect dimension was 5500 Hz (to ensure synchronization with the rotor rotation). This is too small to observe the full range of the DQ dimension, where cross-peak frequencies correspond to the sum of the individual sites that are coupled (such that the width is typically twice the width of the standard spectrum) and lead to a folding of the peaks. An unfolded spectrum can nevertheless be obtained by a “shearing” transformation that yields a SQ-SQ spectrum, instead of the double-quantum – single-quantum (DQ-SQ) 2D spectrum that is obtained after standard Fourier transformation. This is done by first performing Fourier transform on the direct dimension of the 2D time-domain signal, and then applying a first-order phase correction of the indirect-time domain signal with a coefficient that varies linearly with the frequency in the direct dimension, and finally applying the Fourier transform in the indirect dimension to obtain the sheared SQ-SQ spectrum46,47. All chemical shifts are given relative to (neat) tetramethylsilane (TMS).
The densities of the samples were measured on the ground hybrid glasses using a pycnometer with helium AccuPyc II 1340 (Micromeritics, Norcross, GA).
The BET surface area of the samples was determined by absorption/desorption of nitrogen at 77K using a Tristar II 3020 BET surface analyzer. Samples were previously outgassed at 110°C in a nitrogen flow overnight. Surface areas were determined using the Brunauer-Emmet-Teller (BET) method.
SAXS scattering experiments were conducted at the ultra-small angle X-ray scattering (USAXS) instrument at the Advanced Photon Source, Argonne National Laboratory48. This instrument makes use of Bonse-Hart type of crystal optics to access a scattering q range that is normally unavailable to a conventional pinhole based small angle scattering camera. Here q is the magnitude of the scattering vector, and is defined as q = 4π/λ sin(θ), where λ is the X-ray wavelength and θ is one half of the scattering angle. The USAXS instrument has a q resolution of ≈ 1 × 10−4 Å−1. Coupled with an add-on Pilatus 100K detector, the USAXS instrument can access a q range from 1 × 10−4 Å−1 to 1 Å−1 49. Particularly, it is worth noting that the USAXS instrument is primary-calibrated, i.e., the measured intensity is directly related to the differential scattering cross section, a physical property of the material being studied50.
1D-collimated USAXS experiments were conducted in fly-scan mode using monochromatic 17.5 keV X-rays (λ = 0.708481Å). The beam size was 0.8 mm × 0.8 mm. The X-ray flux density was ≈ 1013 photon s−1mm−2. The high X-ray flux ensures that weak scattering signals, commonly expected from a melting gel system, be captured. Total scan time was 120 s. The sample was placed on a standard sample holder. SAXS experiments were conducted twice on some of the samples as a check of damage to the samples caused by the X-ray beam. It was found that in these hybrid glasses, the exposed X-ray dosage does not alter the materials microstructures.
To visually examine the materials microstructures, the surface and the cross-sections of the fractures of the hybrid glasses were evaluated using a HITACHI S-2700 SEM. The SEM was operated at 25 keV acceleration voltage. The SEM was equipped with and AMT digital camera system.
3. Results and discussion
In this study we focused on the two compositions listed in Table 1 for two reasons. First, the melting gel with 75%MTES-25%DMES composition is the gel with the highest content of MTES which retains the melting properties5. On the other hand, the melting gel with 65%MTES-35%DMES composition is the borderline gel between solid and liquid state11 at room temperature. Second, the hybrid organic inorganic glasses obtained using those compositions proved to be the best anticorrosive coatings12.
The Raman spectra of the investigated hybrid glass samples are presented in Figure 1a while the assignments of the peaks are listed in Table 2.
Figure 1.
The Raman spectra of the hybrid glasses. (a) Complete Raman spectra; (b) partial spectra between 600 cm−1 and 900 cm−1 which illustrate an increasing of Si-C bonds in the hybrid glass with a higher amount of DMDES (sample 2).
Table 2.
The Raman peak assignments of the hybrid glasses obtained by consolidation of the melting gels
| Raman | Assignements |
|---|---|
| Raman Shift / cm−1 | |
| 2971 (w) | ν asym CH3 |
| 2901 (s) | ν sym CH3 |
| 2808 (vw) | overtone |
| 1413 (vs) | δ asym CH3(Si-CH3) |
| 1223 (sh) | δsym CH3 (Si-CH3) |
| 1097 (sh) | ν asym Si-O-Si |
| 860 (m) | δ asym Si-O-C (alkoxi) |
| 797 (w) | ν sym Si-O-Si |
| 743 (w sh) | ρ CH3 |
| 704 (m) | ν Si-C (Si-CH3) |
| 589 (w) | Si-O-Si ring breathing vibration for 6 member |
| 461 (s) | Si-O-Si ring breathing vibration for a 8 member ring |
| 202 (s) | σ Si-O-Si |
The peak intensities are given in parentheses as (s) for strong, (m) for medium, (w) for weak, (vw) for very weak and (sh) shoulder
Raman Characterization
Although more than 60% of the studied hybrid glasses contain SiO210, the characteristic peaks for the presence of the Si-O-Si bonds, identified at 790 cm−1 (assigned to νsym Si-O-Si) and at 1090 cm−1 (assigned to νasym Si-O-Si), have very low intensity51,52. On the other hand, the characteristic peaks for the Si-O-Si ring breathing mode, identified at 461 cm−1 (attributed to the 8-member ring structure) and at 202 cm−1 (characteristic σ Si-O-Si), are high in intensity53–55. These results demonstrate that the main structure of the matrix is formed by an 8-member ring structure. Additionally, we identified a very weak peak at 589 cm−1 that is characteristic of a 6-member ring structure53, which suggests that small amount of this structure is also present. Raman spectroscopy also reveals the presence of Si-C bonds. These was identified at 705 cm−1 and was assigned to the ν Si-C at52,55.
Besides the peaks mentioned above, the most intense peaks belong to the methyl groups bonded to the silica backbone. These were identified at 2971 cm−1 and 2901 cm−1, which were assigned to νasym CH3 and νsym CH3, respectively. In addition, Raman bands placed at 1413 cm−1, 1223 cm−1 and 743 cm−1 were assigned to δasymCH3, δsym CH3 and ρ CH3, respectively55. The high intensity of the peaks assigned to the organic groups can be credited to the presence of the methyl groups to the surface of the hybrid glasses investigated4,5,12. Moreover, in the Raman spectra of the both samples, we also identified a very weak peak at 864 cm−1 which was assigned to the δ Si-O-C. The existence of this band demonstrates that few ethoxy groups from the original alkoxides remain unreacted and are entrapped in the hybrid glasses.
Figure 1b details the normalized Raman spectra of the two hybrid glasses investigated in this study. It can be observed that the Raman intensity of the band assigned to the ν Si-C is increasing for sample 2, which has a higher amount of DMDES (35 mol %) than sample 1 (25 mol %). Here, the higher amount of the DMDES leads to a higher number of Si-C bonds, which is reflected by the higher intensity of the ν Si-C peak in the sample 2.
NMR peak assignments
29Si NMR measurements are used to distinguish between the different Si environments formed in the hybrid glasses on the basis of the number of attached carbons and the degree of condensation. The notation Tn and Dn are used to designate CSi(OSi)n(OR)3-n and C2Si(OSi)n(OR)2-n units originating from MTMES and DMDES monomers, respectively, with R = H corresponding to silanol groups or R = –CH2–CH3 corresponding to unreacted Si–O–CH2–CH3 functions from the alkoxide. Quantitative 29Si NMR spectra recorded at room temperature for the hybrid glasses samples are shown in Figure 2a. Assignments of the different signals to D1, D2, T2, and T3 signals can be made based on literature data56–59. The variations of relative intensities between Tn and Dn sites in the materials of the sample 1 and 2 reflect the different compositions of their synthesis mixtures, with larger proportions of DMDES in the sample 2 leading to higher relative intensities of the Dn sites.
Figure 2.
(a) Quantitative 29Si NMR spectra collected of the hybrid glasses. (b) Quantifications with estimated uncertainties based on 29Si NMR peak areas. Samples 1 and 2 are depicted in blue and red color, respectively.
At room temperature, different conformations of the polymer chains are dynamically averaged on the NMR timescale (fast exchange regime) since they are frozen. The linewidths in the slow motional regime are typically of the order of 5 ppm (full width at half maximum, FWHM) corresponding to 300 Hz for 29Si at 7.0 T. The characteristic time corresponding to the intermediate exchange regime for a pair of peaks with a frequency difference Δν between 2 sites is given by 1/(πΔν)60. Thus, in order for chemical shift dispersions on the order of 5 ppm to be dynamically averaged out, it typically requires the chain reorientation dynamics to be significantly faster than 10−3 s. The fast motional regime is clearly reached at room temperature. Some chemical shift dispersion nevertheless remains for each type of Tn or Dn environment, which points to chemically distinct local environments. This is particularly clear for D2 sites, which show two clearly-resolved peaks, but also for the T2 and T3 peaks which show more or less pronounced shoulders (more clearly visible on 29Si[1H] cross polarization MAS spectra that have better signal to noise, data not shown). The deconvolution of spectra in individual components are shown in Supplementary Information (Figure S1). These different components may be due to differences in the nature of the neighboring Si sites (D1, D2, T2 or T3) or, for T2 species, to the nature of the R group of uncondensed O atoms (O-H or unreacted O-CH2-CH3 moieties). Attempts to assign these different contributions are discussed further below.
The 29Si NMR spectra of the hybrid glasses, samples 1 and 2, shown in Figure 2a, are very similar and their assignments are presented in Table 3. This is more clearly reflected in a quantitative analysis of the 29Si NMR data shown in Figure 2b, which are based on the spectral decompositions shown in Supplementary Information (Fig. S1).
Table 3.
The assignments of the 29Si MAS NMR Spectra
| Chemical Shift / ppm
|
||||||||
|---|---|---|---|---|---|---|---|---|
| Sample | D° | D1 | D2 | Ta2 | Tb2 | Tc2 | Ta3 | Tb3 |
| 1 | −17.2 | −19.0 | −21.1 | −54.5 | −56.8 | −59.1 | −64.5 | −66.4 |
| 2 | −16.4 | −19.0 | −21.3 | −54.0 | −57.0 | −59.3 | −64.5 | −66.3 |
Another important observation from these quantitative analyses is the discrepancy between the overall populations of Dn and Tn sites and the relative amounts of DMDES and MTES used for the synthesis. The stars in Figure 2b indicate the populations expected based on the composition of the synthesis mixtures. They reveal that a significant proportion of the D sites have been lost in the course of the synthesis procedure or during the thermal treatments necessary to remove the solvent (70 °C/24 h), to remove the water (110 °C/24 h) or to consolidate the hybrid glasses (temperature of consolidation is listed in Table 1). This effect is significantly more pronounced in the case of the sample 2, which has the highest DMDES content (35%) in the synthesis mixture. As shown in Table 4, this different behavior appears to be due to a very subtle difference of Dn and Tn relative proportions, with Tn / Dn ratios varying between 4.6 and 3.8 in hybrid glass samples 1 and 2, respectively, compared to MTES/DMDES ratios of 3.0 and 1.9 in the synthesis mixtures.
Table 4.
Summary of the quantitative 13C and 29Si NMR analyses.
| Synthesis mixture composition |
NMR quantifications | |||||
|---|---|---|---|---|---|---|
| Sample | MTES (%) |
DMDES (%) |
MTES / DMDES |
ΣI(Tn)/ΣI(Dn) |
I(CH3-SiO3)/ ½I[(CH3)2-SiO2] |
n(O-CH2-CH3) / (n(D1)+n(T2)) a |
| 1 | 75 | 25 | 3.0 | 4.6 | 4.8 | 0.24 |
| 2 | 65 | 35 | 1.9 | 3.8 | 3.7 | 0.27 |
Proportion of incompletely condensed Si (D1 or T2 sites) forming Si-O-CH2-CH3 rather than Si-OH moieties, calculated from the combination of 13C and 29Si quantitative NMR data (see details in Supplementary information).
These unexpected results are corroborated by quantitative 13C NMR data, which also shed some light on the nature of uncondensed O atoms. Figure 3a shows quantitative 13C NMR spectra recorded at room temperature for hybrid glasses 1 (in blue) and 2 (in red). The signatures of (CH3)2–SiO2 species originating from DMDES monomers (at −3 ppm) and CH3–SiO3 species originating from MTES monomers (at -1 ppm) are well resolved in all 13C spectra. Changes in relative amplitudes reflect again to some extent the relative ratios of monomers in the synthesis mixtures and confirm the assignments. Two other peaks are detected in the 13C NMR spectra, at 18 and 58 ppm, which may be assigned to CH3–CH2–O and –CH2–O– environments (respectively) corresponding to partially unreacted ethoxy groups, Si–O–CH2–CH3 functions carried by D1 or T2 Si atoms. Those data are in good agreement with the Raman results which also show the presence of some unreacted ethoxy groups.
Figure 3.
(a) Quantitative 13C NMR spectra for the hybrid glasses after the consolidation treatment (b) Quantifications based on 13C NMR peak areas with hybrid glasses prepared using 75%MTES-25%DMDES and 65%MTES-35%DMDES depicted in blue and red, respectively. Uncertainties are estimations.
Quantitative analyses of 13C NMR data, shown in Figure 3b and Table 4, confirm the loss of a significant proportion of the DMDES monomers (or of their products) in the course of the materials synthesis. These data provide very reliable results since solid nature of the hybrid glasses allowed the use of faster MAS spinning rates (typically 12 kHz as compared to 5.5 kHz for gel samples). The ratios between the area of the peak assigned to CH3–SiO3 species (at -3 ppm) and half the area of the peak assigned to (CH3)2–SiO2 species (at -1 ppm) are 4.8 and 3.7 for gel samples 1 and 2, respectively. These results are in quantitative agreement with the ratios between the cumulated areas of Tn and Dn 29Si NMR peaks (4.6 and 3.8, respectively), and considerably higher than the MTES/DMDES ratios of the synthesis mixture (3.0 and 1.9, respectively). As already observed from 29Si NMR data, the loss of DMDES moieties is more severe in the sample 2 containing the higher amount of these monomers in the synthesis mixture, which tends to even out the contrast of compositions between the two samples.
Quantifications of the areas of the two resonances assigned to the –O–CH2–CH3 groups reveal the nature of the incompletely condensed D1 and T2 Si environments detected in 29Si NMR data. Combining both quantitative 29Si and 13C NMR data (see details in Supplementary Information), we find that 24% of non-polymerized Si–O sites are in the form of Si–O–CH2–CH3 groups in the sample of hybrid glass 1. (We note that we cannot tell whether these are supported by D1 and T2 sites.) This fraction is as high as 27% for the sample of hybrid glass 2, which has the higher DMDES content in the synthesis mixture. And yet the amount of D1 in the sample is too small to explain such a difference by a concentration of unreacted Si–O–CH2–CH3 on D1 sites. It seems instead that the more hydrophobic character of the sample of hybrid glass 2 stabilizes these Si–O–CH2–CH3 moieties (both on D1 and T2 sites), which may well contribute to their inherent flexibility and hydrophobicity. The amount of uncondensed Si sites (whether in the form of Si-OH or Si-OEt) D1 and T2 was used to calculate the probability for a Si-O bond to remain uncondensed. We find that, in both hybrid glasses, this probability is between 11 and 12% for Si-O bonds in T sites as compared to 3 to 5% for D sites. This means that DMDES precursors tend to condense more easily than MTES.
The network analysis of the hybrid glasses
One-dimensional 29Si NMR spectra resolve a certain number of local Si environments, or “molecular motifs”61 characterized by the number of Si-C bonds and the number of Si neighbors connected via bridging O atoms as showed in Figure 2. More extended motifs can however be identified by recording two-dimensional correlation spectra that probe the connectivities between Si atoms, in the hope to better understand in particular the fine structures of the 29Si spectra. Examples of these fine structures are highlighted on the top Figure 4 by the spectral decomposition of a 29Si[1H] CP-MAS NMR spectrum collected for the sample hybrid glass 2, with individual components shown as grey lines. Between 2 to 3 different types of D2, T2, and T3 environments are thus distinguished, which are designated thereafter as D2a, D2b, D2c, T2a, T2b, T2c, T3a, and T3b.
Figure 4.
Dipolar -mediated 29Si-29Si correlation NMR spectra for the sample hybrid glass 2 65%MTES-35%DMDES, revealing close proximities between Si atoms characteristic of Si-O-Si framework connectivity. The blue contours correspond to the original spectrum (sheared from SQ-DQ to DQ-DQ representation, see experimental section for details) whereas the red contours correspond to a spectrum obtained by symmetrization with respect to the diagonal (shown as a black dashed line) of the former, to increase signal to noise. Only the upper-left half of this symmetric spectrum is shown to better see the original spectrum underneath. 1D spectra shown in black on top and right side of the 2D spectrum are 29Si [1H] CP-MAS spectra recorded with identical contact time as the 2D experiment. The 1D model obtained by decomposition of this 29Si [1H] CP-MAS spectrum in individual contributions (shown as grey lines) is shown in red.
The two-dimensional spectra shown in Figure 4 were collected to try to better understand the origins of these different features. They were obtained with a sequence of radio-frequency pulses designed to reintroduce the homonuclear dipole-dipole couplings between nearby 29Si nuclei to probe their spatial proximities44. This experiment was conducted in conditions specifically optimized to yield intense cross peaks for the short Si-Si distances of ca. 3 Å between connected 29Si-O-29Si pairs and weak or negligible cross peaks for the longer distances associated with non-connected pairs62. The result of such an experiment (after the “shearing” transformation, see details in Experimental Section) is a 2D map in which pairs of correlation peaks appear on both parts of the spectrum diagonal at horizontal and vertical 29Si frequencies corresponding to the nuclei that are connected. Because of severe signal to noise limitations in the original 2D spectrum (shown in blue), a symmetrization procedure was applied to gain signal to noise (in principle by the square root of 2) by adding up the regions on both sides of diagonal. The resulting 2D spectrum is shown in red in Figure 4, overlaid with the original spectrum. Because of the symmetry of this spectrum with respect to the diagonal, only its upper-left half is shown to better illustrate the original (blue) spectrum underneath. (The symmetric representation can sometimes be misleading by making high noise level points look like correlation peaks).
The main features of the correlation spectrum reveal extended molecular motifs centered on Si-O-Si pairs that are depicted schematically in Figure 4 (using the same color code as in Figure 2), and which point to different levels of network ramification, with 4-fold-branched T3-T3 pairs, 3-fold-branched T2-T3 and D2-T3 pairs, linear D2-T2, D2-T2 and D2-D2 pairs. The 2D spectrum shows a higher intensity of D2-D2 correlations as compared to D2-T2 correlations, which barely point out of the noise level after symmetrization (and only at the D2c -T2 position). The quantitative 1D spectra of all samples show D2a intensities comparable to the D1 intensities, suggesting that the D2a position (-17 ppm) could be a signature of a D2-D1 end chain.
Semi-quantitative estimations of the conditional probability for each site to be correlated with another may be calculated assuming that (i) Si-Si distances between connected Si-O-Si sites are all identical (resulting in equal efficiencies of the double-quantum excitation during the NMR sequence) and (ii) that signal losses during the sequence are the same for all sites. The results indicate that site T3b environments are primarily connected to T3 environments (primarily to other T3b), with only ca. 12% probability to be connected to T2 sites and ca. 11% probability to be connected to D2 sites. They are hence indicative of strongly reticulated regions. In contrast, T3a environments have higher probabilities to be connected to T2 sites (17%) and 10% to be connected to D2 sites. They are nevertheless connected primarily to other T3a sites and to T3b environments (49% and 23%, respectively), indicating that they may well correspond to edges of these strongly interconnected domains.
The quantitative analyzes of 1D 29Si NMR spectra of samples may be re-examined in the light of this difference of reticulation degrees around T3a and T3b Si environments. In both samples the more reticulated T3b sites appear to account for ca. 46% of the total amount of T3 sites, in hybrid glass sample 1, 75%MTES-25%DMDES and 51% in hybrid glass sample 2, 65%MTES-35%DMDES. Other 2D peak intensities are less intense and thereby less reliable, making it difficult to understand the differences between different types of D2 sites and different types of T2 sites. We observe however that D2 sites in general tend to have a stronger probability to be connected to other D2 sites rather than to T2 sites, whereas T2 sites are connected to other T2 sites rather than to D2 sites. Both have comparable (and rather high) probabilities to be connected to T3 sites.
While NMR provides deep insights regarding the structure at the molecular level, the microstructural morphologies of the sol-gel materials are also known to affect materials properties63. SAXS, due to its nature of being a nondestructive and statistically significant technique, is applied to probe the possible segregation, separation, and presence of micro/nanopores within the consolidated hybrid glasses. In particular, we made use of the high-brilliance synchrotron X-ray beam to ensure that weak scattering intensities can be captured.
The SAXS profiles of both hybrid glass samples are shown in Figure 5. It is clear that these profiles bear much resemblance. First, the reduced (slit-smeared) differential scattering cross section d Σ(q)/VdΩ is very low in intensity and noisy, which indicates that the scattering contrast that gives rise to the scattering signal is weak. This high degree of noise originates from data reduction, whereas the sample scattering, although weak, is unmistakably evident from the raw scattering data (not shown). Second, a broad scattering feature at qs below 0.01 Å−1 exists, which indicates certain degree of microphase separation. This scattering feature also does not appear to have a strong dependence on the molar ratio between MTES and DMDES.
Figure 5.
Synchrotron SAXS data and fit using Debye-Bueche model of Sample 1 (75% MTES + 25% DMDES) and Sample 2 (65% MTES + 35% DMDES).
NMR results of these hybrid glass materials strongly indicate that they possess a network structure. To fully account for the scattering profiles, we made use of the Debye-Bueche model64, which describes scattering from a randomly distributed two-phase system, a situation very similar to the hybrid glasses. This model assumes that the pair-correlation function γ(r) follows a simple exponential decay,
| (1) |
where r is the interatomic distance, and ξ, the correlation length, is a measure of the average separation distance between the separated phases.
The scattering cross section, following Eqn. (1), is
| (2) |
Here, Δρ is the scattering contrast; ϕ1 and ϕ2 are the volume fraction of the two phases, respectively.
Figure 5 shows that despite the high degree of noise, the Debye-Bueche model describes the scattering intensity profiles very well. The correlation lengths were found to be 10.3 ± 0.8 nm for sample 1 and 7.5 ± 0.1 nm for sample 2, respectively. This result indicates that higher MTES-to- DMDES ratio leads to a larger phase-separation distance. We also note that our observed scattering profiles are similar to those acquired by Hagiwara et al65 in a similar hybrid organoalkoxide material prepared using sol-gel synthesis. Instead of characterizing possible phase separation, these authors opted to view the gels as fractal structures, perhaps due to the limited q range of their SAXS camera. In our case, the broad q range of the USAXS instrument leaves little double about the bend of scattering curve near 0.01 Å−1, which prompted our analysis of a phase-separated network structure that is consistent with NMR findings. Such type of separation was also visually confirmed by atomic force microscopy in a study of nanohybrids containing 1,8-bis(triethoxysilyl)octane66.
We also evaluated the degree of phase separation following Eqn. (2). For this purpose, we calculated the scattering length densities of both MTES and DMDES, which are tabulated in Table 5. Assuming complete phase separation, the theoretical X-ray scattering contrast between MTES and DMDES is 2.31 × 1020 cm−4. Empirically, the ratio between the experimental and theoretical scattering contrasts could serve to gauge the phase separation. Here, following the fitting results from Figure 5, we found that the experimental scattering contrasts for samples 1 and 2 are 8.35 × 1016 cm−4 and 1.97 × 1017 cm−4, respectively. In other words, the experimental contrasts are on the order of 10−3 of the theoretical contrast. This result shows that MTES and DMDES are fairly homogenously mixed and do not phase separate to form well defined phase structures. In other words, from an electron density point of view, the hybrid glass materials possess a high degree of uniformity across nm to micrometer length scales.
Table 5.
Scattering length density and contrast
| Mass density* (g/cm3) |
Scattering length density (1010cm−2) |
Scattering contrast (1020cm−4) |
|
|---|---|---|---|
| MTES | 1.332 | 11.66 | 2.31 |
| DMDES | 1.122 | 10.14 |
Mass densities of MTES and DMDES are estimated from a series of mass densities of hybrid glasses containing different molar ratios of MTES and DMDES.
Last but not least, we note that the low scattering intensities also confirm that micro- and nano-pores are completely absent in these hybrid glasses, as the pores have the highest possible X-ray scattering contrast and their presence would have been eminently captured by the scattering methods. From a mechanical property point of view, the lack of pore-related defects is beneficial in improving fracture toughness and reducing fatigue cracks.
The BET Surface area analysis
The N2 physisorption data are presented in the Table 1. These shows very low BET surface areas, <<1 m2/g. For the studied samples the BET surface area decreasing with an increase of the amount of disubstituted alkoxide. Those data are in good agreement with SAXS data which shows a lack of porosity for those samples.
The SEM analysis
The SEM images of the surface fractures and the cross section of the fracture of both studied hybrid glass samples are presented in Figure 6. The images of the surface Figure 6a and Figure 6b show a smooth surface without defects or glass domains. For a better focus and due to the fact that the surface are defects free the images were collected near to the fracture and defects which appeared during the fracture process. The images of the cross sections of the fracture (Figure 6c-6f) show homogenous structures without formation of any domains or without any visible phase separation which is in agreement with the previous data. No micro- or macro-pores were detected in the hybrid glasses. These surface-sensitive SEM data are in very good agreement with the SAXS data, which are connected to the bulk microstructure. The cross section of the fracture of the hybrid glass with composition 65%MTES-35%DMES presented in Figure 6d and 6f display a flaky area which shows also a certain orientation of the material. The presence of the flakes can be correlated with the higher concentration of the organic content in this sample. On the other hand, the images of the cross section of the hybrid glass with composition 75%MTES-25%DMES presented in Figure 6c and 6e show a clear fracture with longitudinal lines. This aspect of the fracture can be correlated with higher content of Si-O-Si bonds in sample 1 as it was revealed by the NMR analysis. The highest content of Si-O-Si bonds is increasing the rigidity of the sample which can determine a clearer fracture.
Figure 6.
The SEM images of the surfaces of (a) 75%MTES-25%DMDES; and (b) 65%MTES-35%DMES hybrid glass samples (c) and (e) show the cross-sections of fracture for the 75%MTES-25%DMDES sample; (d) and (f) show the cross-sections of fracture for the 65%MTES-35%DMDES sample.
4. Conclusions
The structure and morphology of hybrid glasses prepared with different ratios of MTES and DMDES have been investigated in this study using Raman, one- and two-dimensional 29Si and one dimensional 13C NMR spectroscopy, SAXS and SEM microscopy.
The Raman spectroscopy showed that the structure of the matrix is formed mainly by 8-member ring structures. In addition, the Raman measurements revealed that organic groups are placed at the surface of the hybrid glasses.
The 13C and 29Si NMR spectroscopy offered information about the molecular structures of the hybrid glasses. The main molecular species were quantitatively identified. Quantitative 29Si NMR of the hybrid glasses samples shown characteristic signals to D1, D2, T2, and T3. The quantitative measurements identify the final ratio between the mono substituted species CH3-SiO3 and the di-substituted species (CH3)2-SiO2 which are present in each studied hybrid glass. This points out a loss of DMDES precursors during the synthesis. Our data reveal the presence of unhydrolyzed Si-OEt groups which was also identified in the Raman spectra. These Si-OEt groups, seem to be stabilized when higher amounts of DMDES are present, and may increase the inter-chain mobility by acting as lubricant owing to their inherent flexibility and hydrophobicity, in contrast with Si-OH groups that form hydrogen bonds. This can explain the higher rigidity of sample 1.
Two-dimensional 29Si NMR data show more extended structural motifs within the framework, which seem to reveal that two distinct types of T3 Si environments (the dominant type of Si sites in both materials) exist, that correspond to Si atoms located in regions with different extents of framework condensation.
The SAXS data revealed that the mono substituted species (CH3-SiO3) and the di-substituted species ((CH3)2-SiO2) that are coming from hydrolysis and polycondensation of the MTES and DMDES are homogeneously mixed and do not separate to form a well-defined phase structure. In addition, the SAXS data confirmed the absence of any nano- or micro-pores.
The absence of the pores was confirmed also by the BET and SEM analyses. The SEM images of the fractures confirmed that the sample 1 (75%MTES-25%DMDES) has higher rigidity than sample 2(75%MTES-25%DMDES) which contains a higher amount of DMDES. Based on the NMR data it was concluded that the samples with higher amount of DMDES retain a higher amount of unhydrolyzed Si-OEt groups which can increase the flexibility of the hybrid glass 2.
We have performed a thorough investigation of the structure of hybrid glasses at both a molecular level and a microstructure level. We envisage that our findings in this study would serve as a bridge to connect the molecular arrangements of the hybrid glasses to their physical characteristics, and potentially impact the synthesis and design of a broader range of hybrid glass materials by consolidation of melting gels with improved physical, mechanical, chemical and electrochemical performances.
Supplementary Material
Acknowledgments
This work is supported by NSF –DMR Award #1313544, Materials World Network, SusChEM: Hybrid Sol-Gel Route to Chromate-free Anticorrosive Coating and by Ministerio de Economía y Competitividad, Spain (PCIN-2013-030). Use of the Advanced Photon Source, an Office of Science User Facility operated for the U.S. Department of Energy (DOE) Office of Science by Argonne National Laboratory, was supported by the U.S. DOE under Contract No. DE-AC02-06CH11357.
Footnotes
Certain commercial equipment, instruments, software or materials are identified in this paper to foster understanding. Such identification does not imply recommendation or endorsement by the Department of Commerce or the National Institute of Standards and Technology, nor does it imply that the materials or equipment identified are necessarily the best available for the purpose.
References
- 1.Matsuda A, Sasaki T, Hasegawa K, Tatsumisago M, Minami T. Thermal softening behavior and application to transparent thick films of poly(benzylsilsesquioxane) particles prepared by the sol-gel process. J Am. Ceram Soc. 2001;84:775–780. [Google Scholar]
- 2.Sanchez C. Chemical design of hybrid organic-inorganic materials synthesized via sol-gel. New J. Chem. 1994;10:1007–1040. [Google Scholar]
- 3.Jitianu A, Amatucci G, Klein LC. Phenyl-substituted siloxane hybrid gels that soften below 140˚C. J. Am. Ceram. Soc. 2008;92:36–40. [Google Scholar]
- 4.Jitianu A, Gonzalez G, Klein LC. Hybrid Sol-Gel Glasses with Glass Transition Temperatures below Room Temperature. J. Am. Cer. Soc. 2015;98:3673–3679. [Google Scholar]
- 5.Jitianu A, Doyle J, Amatucci G, Klein LC. Methyl modified siloxane melting gels for hydrophobic films. J. Sol-Gel. Sci. Technol. 2010;53:272–279. [Google Scholar]
- 6.Kakiuchida H, Takahashi M, Tokuda Y, Masai H, Kuniyoshi M, Yoko T. Viscoelastic and Structural properties of the Phenyl-Modified Polysiloxane System with a three-dimensional Structure. J. Phys. Chem. B. 2006;110:7321–7327. doi: 10.1021/jp0573543. [DOI] [PubMed] [Google Scholar]
- 7.Kakiuchida H, Takahashi M, Tokuda Y, Masai H, Kuniyoshi M, Yoko T. Effects of Organic Groups on Structure and Viscoelastic properties of Organic-inorganic polysiloxane. J. Phys. Chem. B. 2007;111:982–988. doi: 10.1021/jp064252j. [DOI] [PubMed] [Google Scholar]
- 8.Masai H, Tokuda Y, Yoko T. Gel-melting method for preparation of organically modified siloxane low-melting glasses. J. Mater. Res. 2005;20:1234–1241. [Google Scholar]
- 9.Klein LC, Jitianu A. Organic-inorganic hybrid melting gels. J. Sol-Gel Sci. &Tehnol. 2011;59:424–431. [Google Scholar]
- 10.Jitianu A, Lammers K, Arbuckle-Keil GA, Klein LC. Thermal analysis of organically modified siloxane melting gels. J. Therm. Anal. Calorim. 2012;107:1039–1045. [Google Scholar]
- 11.Jitianu M, Jitianu A, Stamper M, Aboagye D, Klein LC. Melting Gel Films for Low Temperature Seals. In: Jain M, Jia Q, Puig T, Kozuka H, editors. Solution Synthesis of Inorganic Functional Materials – Films, Nanoparticles, and Nanocomposites. Vol. 1547. Materials Research Society; Warrendale, PA: 2013. pp. 81–86. MRS. [Google Scholar]
- 12.Aparicio M, Jitanu A, Rodriguez G, Degnah A, Al-Marzoki K, Mosa J, Klein LC. Corrosion Protection of AISI 304 Stainless Steel with Melting Gels Coating Electrochim. Acta. 2016;202:325–332. [Google Scholar]
- 13.Jitianu A, Doyle J, Amatucci G, Klein LC. Methyl-modified melting gels for hermetic barrier coatings. Proceedings MS&T 2008 Enabling Surface Coating Systems: Multifunctional Coatings; Pittsburgh, PA: 2008. pp. 2171–2182. [Google Scholar]
- 14.Jitianu A, Klein LC. In: “Encapsulating Battery Components with Melting Gels” Ceramic Transactions, “Advances in Materials Science for Environmental and Energy Technologies III. Ohji T, Matyas J, Manjooran NJ, Pickrell G, Jitianu A, editors. Vol. 250. American Ceramic Soc; Westerville, OH: 2014. pp. 279–286. [Google Scholar]
- 15.Gambino L, Jitianu A, Klein LC. Dielectric behavior of organically modified siloxane melting gels. J. Non-Cryst. Solids. 2012;358:3501–3504. [Google Scholar]
- 16.Jitianu A, Britchi A, Deleanu C, Badescu V, Zaharescu M. Comparative study of the sol-gel processes starting with different substituted Si-alkoxides. J Non Cryst. Solids. 2003;319:263–279. [Google Scholar]
- 17.Smith ME, Holand D. In: Atomic Scale Structure of Gel Materials by Solid State NMR. Sakka S, editor. II. Handbook of Sol-Gel Science and Technology, Processing Characterization and Application; 2005. pp. 35–63. Characterization of the Sol-Gel materials and products. [Google Scholar]
- 18.Sindorf DW, Maciel GE. Solid-State NMR Studies of Reactions of Silica Surfaces with polyfunctional Chloromethylsilanes and Ethoxymethylsilanes. J. Am. Chem. Soc. 1983;105:3767–3776. [Google Scholar]
- 19.Fyfe CA, Zhang Z, Aroca P. An alternative preparation of organofunctionalized silica gels and their characterization by two-dimensional high-resolution solid state heteronuclear NMR correlation spectroscopy. J. Am. Chem. Soc. 1992;114:3252–3255. [Google Scholar]
- 20.Petters MPJ, Wakelkamp WJJ, Kentgens APM. A 29Si solid-state magic angle spinning nuclear magnetic resonance study of TEOS-based hybrid materials. J. Non-Cryst. Solids. 1995;189:77–89. [Google Scholar]
- 21.Bois L, Maquet J, Babonneau F, Mutin H, Bahloul D. Structural Characterization of sol-gel derived oxycarbide glasses. 1 Study of the pyrolysis process. Chem. Mater. 1994;6:796–802. [Google Scholar]
- 22.Bois L, Maquet J, Babonneau F, Bahloul D. Structural Characterization of sol-gel derived oxycarbide glasses. 2 Study of the thermal stability of silicon oxicarbide phase. Chem. Mater. 1995;7:975–981. [Google Scholar]
- 23.Trimmel G, Badheka R, Babonneau F, Latournerie J, Dempsey P, Baholul-Houlier D, Paramentier J, Soraru GD. Solid-datate NMR and TG/MS study of transformation of methyl groups during the pyrolysis of preceramic precursors to SOC glasses. J. Sol-gel Sci & Technol. 2003;26:279–283. [Google Scholar]
- 24.Syal N, Hoebbel D, Mennig M, Schmidt H. A solid state 29Si and 13C NMR study on synthesis of thin silicon-oxicarbide glass sheets by a sol-gel route. J. Mat. Chem. 1999;9:3061–3067. [Google Scholar]
- 25.Sitarz M, Czosnek C, Jelen P, Odziomek M, Olejniczak Z, Kozanecki M, Janik JF. SiOC glasses produced from silsequioxanes by aerosol-assisted vapor synthesis method. Spectrochem Acta part A, Molec and Biomolec Spec. 2013;112:440–445. doi: 10.1016/j.saa.2013.05.007. [DOI] [PubMed] [Google Scholar]
- 26.Brus J, Skrdlantova M. 1H MAS NMR study of structure of hybrid siloxane-based networks and the interaction with quartz filler. J. Non-Cryst Solids. 2001;281:61–71. [Google Scholar]
- 27.Brus J. Solid state NMR study on Phase separation and order of water molecules and silanol groups in polysiloxane networks. J. Sol-gel Sci & Technol. 2002;25:17–28. [Google Scholar]
- 28.Brus J, Dybal J. Hydrogen Bond Interactions in Organically-Modified Polysiloxane networks studied by 1D and 2d CRAMPS and double-quantum 1H MAS NMR. Macromolecules. 2002;35:10038–10047. [Google Scholar]
- 29.Chu B, Hsiao BS. "Small-angle X-ray scattering of polymers". Chemical Reviews. 2001;101(6):1727–1762. doi: 10.1021/cr9900376. [DOI] [PubMed] [Google Scholar]
- 30.Zhang F, Ilavsky J. "Ultra-small-angle X-ray scattering of polymers". Journal of Macromolecular Science, Part C: Polymer Reviews. 2010;50(1):59–90. [Google Scholar]
- 31.Devreux F, Boilot JP, Chaput F, Lecomte A. Sol-gel condensation of rapidly silicon alkoxides: A joint 29Si NMR and small-angle x-ray scattering study. Phys. Rev A. 1990;41:6901–6909. doi: 10.1103/physreva.41.6901. [DOI] [PubMed] [Google Scholar]
- 32.Xu Y, Zhang L, Wu D, Sun YH, Huang ZX, Jiang XD, Wei XF, Li ZH, Dong ZH, Wu BZ. Durable sol-gel antireflective films with high laser-induced damage thresholds for inertial confinement fusion. J. Opt. Soc. Am. B. 2005;22:905–912. [Google Scholar]
- 33.Li ZH, Gong YJ, Wu D, Sun YH, Wang J, Liu Y, Dong BZ. SAXS analysis of interface in organo modified mesoporous silica. Surf. Interface Anal. 2001;31:897–900. [Google Scholar]
- 34.Craievich AF. Small Angle X-ray scattering by nanostructured materials. In: Sakka S, editor. Handbook of Sol-Gel Science and Technology, Processing Characterization and Application. II. Characterization of the Sol-Gel materials and products; 2005. pp. 161–190. [Google Scholar]
- 35.Dahmouche K, Santilli C, Lafontaine E, Judeinstein P, Craievich AF. Effect of addition of MTES on the structure of Siloxane-PPG nanocomposites. J. Sol-gel Sci. & Technol. 2000;19:429–433. [Google Scholar]
- 36.Dahmouche K, Santilli CV, Pulcinelli SH, Craievich AF. Samall-angle X-Ray scattering study of sol-gel derived Siloxane-PEG and Siloxane-PPG hybrid materials. J. Phys. Chem. B. 1999;103:4937–4942. [Google Scholar]
- 37.Li ZH, Gong YJ, Pu M, Wu D, Sun YH, Dong BZ, Wu ZH, Zhao H. Determination of SiO2 colloid core size by SAXS. J. Mat. Sci. Let. 2003;22:33–35. [Google Scholar]
- 38.Porod G. In: Samll-Angle X-Ray Scattering. Glatter O, Kratky O, editors. Academic Press; London: 1982. Chapter 2. [Google Scholar]
- 39.Gong YJ, Li ZH, Dong BZ. Evaluation of average wall thickness of Oranically modified mesoporous silica. Chinese Chem. Let. 2005;16:139–142. [Google Scholar]
- 40.Boukari H, Lin JS, Harris MT. Small-angle X-ray scattering study of the formation of colloidal silica particles from alkoxides: primary particles or not? J. Colloid Interface Sci. 1997;194:311–318. doi: 10.1006/jcis.1997.5112. [DOI] [PubMed] [Google Scholar]
- 41.Boukari H, Lin JS, Harris MT. Probing the Dynamics of the silica nanostructure formation and Growth by SAXS. Chem. Mater. 1997;9:2376–2384. [Google Scholar]
- 42.Boukari H, Loung GG, Harris MT. Polydispersity during the formation and Growth of Stober Silica particles from Samall-Angle X-ray Scatterin measurements J. Colloid Interface Sci. 2000;229:129–139. doi: 10.1006/jcis.2000.7007. [DOI] [PubMed] [Google Scholar]
- 43.Carravetta M, Eden M, Zhao X, Brinkmann A, Levitt MH. Symmetry principles for the design of radiofrequency pulse sequences in the nuclear magnetic resonance of rotating solids. Chem. Phys. Lett. 2000;321:205–215. [Google Scholar]
- 44.Kristiansen PE, Carravetta M, Lai WC, Levitt MH. A robust pulse sequence for the determination of small homonuclear dipolar couplings in magic-angle spinning NMR. Chem. Phys. Lett. 2004;390:1–7. [Google Scholar]
- 45.States DJ, Haberkorn RA, Ruben DJ. A two-dimensional nuclear overhauser experiment with pure absorption phase in four quadrants. J. Magn. Reson. 1982;48:286–292. [Google Scholar]
- 46.Ernst RR, Bodenhausen G, Wokaun A. Principles of Nuclear Magnetic Resonance in One and Two Dimensions. Oxford Science Publications; New York: 1987. [Google Scholar]
- 47.Fayon F, Le Saout G, Emsley L, Massiot D. Through-bond phosphorus–phosphorus connectivities in crystalline and disordered phosphates by solid-state NMR. Chem. Commun. 2002:1702–1703. doi: 10.1039/b205037b. [DOI] [PubMed] [Google Scholar]
- 48.Ilavsky J, Jemian PR, Allen AJ, Zhang F, Levine LE, Long GG. "Ultra-small-angle X-ray scattering at the Advanced Photon Source". Journal of Applied Crystallography. 2009;42(3):469–479. [Google Scholar]
- 49.Ilavsky J, Zhang F, Allen A, Levine L, Jemian P, Long G. "Ultra-small-angle X-ray scattering instrument at the advanced photon source: history, recent development, and current status". Metallurgical and Materials Transactions A. 2013;44(1):68–76. [Google Scholar]
- 50.Zhang F, Ilavsky J, Long GG, Quintana JP, Allen AJ, Jemian PR. "Glassy carbon as an absolute intensity calibration standard for small-angle scattering". Metallurgical and Materials Transactions A. 2010;41(5):1151–1158. [Google Scholar]
- 51.Bertoluzza A, Fagnano C, Morelli MA, Gottardi V, Guglielmi M. Raman and infrared spectra on silica gel evolving towards glass. J. Non-Cryst. Solids. 1982;48:117–128. [Google Scholar]
- 52.Riegel B, Plittersdorf S, Kiefer W, Hüsing N, Schubert U. Raman spectroscopic analysis of the sol-gel processing of RSi(OMe)3/Si(OMe)4 mixtures. J. Molec. Struct. 1997;410–411:157–160. [Google Scholar]
- 53.Pakjamsai C, Kobayashi N, Koyano M, Sasaki S, Kawakami Y. Characterization of the benzene-insoluble fraction of the hydrolyzate of phenyltrimethoxysilane in the presence of benzyltrimethylamonium hydroxide. J. Polym. Sci: Part A: Polym. Chem. 2004;42:4587–4597. [Google Scholar]
- 54.Li YS. Vibrational spectroscopic studies on triethoxy (-(trifluoromethyl)-phenyl) silane and its sol-gel coatings, Spectrochim. ActaA: Molec. & Biomolec Spec. 2012;96:868–873. doi: 10.1016/j.saa.2012.07.019. [DOI] [PubMed] [Google Scholar]
- 55.Chomel AD, Dempsey P, Latournerie J, Hourlie-Bahloul D, Jayasooriya UA. Gel to glass transformation of methytriethoxysilane: a silicone oxycarbide glass precursor investigated using vibrational spectroscopy. Chem. Mater. 2005;17:4468–4473. [Google Scholar]
- 56.Glaser RH, Wilkes GL, Bronnimann Solid-state 29Si NMR of TEOS-based multifunctional sol-gel materials. J. Non-Cryst. Solids. 1989;113:73–87. [Google Scholar]
- 57.Zhang Y, Wu D, Sun Y, Peng S. Sol–Gel Synthesis of Methyl Modified Optical Silica Coatings and Gels from DDS and TEOS. J. Sol-Gel Sci Technol. 2005;33:19–24. [Google Scholar]
- 58.Engelhardt G, Jancke H, Lippmaa E, Samoson A. Structure investigations of solid organosilicon polymers by high resolution solid state 29Si NMR. J. Organomet. Chem. 1981;210:295–301. [Google Scholar]
- 59.Xu Y, Liu R, Wu D, Sun Y, Gao H, Yuan H, Deng F. Ammonia-catalyzed hydrolysis kinetics of mixture of tetraethoxysilane with methyltriethoxysilane by 29Si NMR. J. Non-Cryst. Solids. 2005;351:2403–2413. [Google Scholar]
- 60.Harris RK. Nuclear Magnetic Resonance Spectroscopy. Longman: Harlow: 1986. [Google Scholar]
- 61.Deschamps M, Fayon F, Hiet J, Ferru G, Derieppe M, Pellerin N, Massiot D. Spin-counting NMR experiments for the spectral editing of structural motifs in solids. Phys. Chem. Chem. Phys. 2008;10:1298–1303. doi: 10.1039/b716319c. [DOI] [PubMed] [Google Scholar]
- 62.Brouwer DH, Kristiansen PE, Fyfe CA, Levitt MH. Symmetry-Based 29Si Dipolar Recoupling Magic Angle Spinning NMR Spectroscopy: A New Method for Investigating Three-Dimensional Structures of Zeolite Frameworks. J. Am. Chem. Soc. 2005;127:542–543. doi: 10.1021/ja043228l. [DOI] [PubMed] [Google Scholar]
- 63.Brinker CJ, Scherer GW. Sol-gel science: the physics and chemistry of sol-gel processing. Academic press; 2013. [Google Scholar]
- 64.Debye P, Bueche A. "Scattering by an inhomogeneous solid". Journal of Applied Physics. 1949;20(6):518–525. [Google Scholar]
- 65.Hagiwara Y, Shimojima A, Kuroda K. "Alkoxysilylated-Derivatives of Double-Four-Ring Silicate as Novel Building Blocks of Silica-Based Materials†. "Chemistry of Materials. 2007;20(3):1147–1153. [Google Scholar]
- 66.Nakanishi T, Norisuye T, Sato H, Takemori T, Tran-Cong-Miyata Q, Sugimoto T, Nomura S. "Studies on microscopic structure of sol-gel derived nanohybrids containing heteropolyacid". Macromolecules. 2007;40(12):4165–4172. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.














