Next Article in Journal
Arginine Is a Novel Drug Target for Arginine Decarboxylase in Human Colorectal Cancer Cells
Previous Article in Journal
The Role of Extracellular Vesicles in Aging and Disease
Previous Article in Special Issue
General Capacitance Upper Limit and Its Manifestation for Aqueous Graphene Interfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Study of Organic Nanostructures Fabricated by Inclusion of 2-Methylbenzimidazole Molecules in Nanotubes of Chrysotile Asbestos, Mesoporous Silica, and Nanopores of Borate Glasses

Ioffe Institute, Russian Academy of Sciences, Politechnicheskaya 26, 194021 Saint Petersburg, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(18), 13740; https://doi.org/10.3390/ijms241813740
Submission received: 15 August 2023 / Revised: 27 August 2023 / Accepted: 28 August 2023 / Published: 6 September 2023

Abstract

:
New organic nanostructures were synthesized by introducing 2-methylbenzimidazole (MBI) molecules from a melt, gas phase, or alcoholic solution into nanosized voids of borate porous glasses (PG), nanotubes of chrysotile asbestos (ChA), and mesoporous silica (MS). The incorporation of MBI into borate glasses with different pore sizes is accompanied by the appearance of several phases formed by nanocrystallites which have a MBI crystal structure, but somewhat differ in lattice parameters. The size of some crystallites significantly exceeds the size of nanopores, which indicates the presence of long-scale correlations of the crystal structure. The size of MBI nanocrystallites in ChA was close to the diameter of nanotubes (D ~10 nm), which shows the absence of crystal structure correlations. The XRD pattern of mesoporous silica filled by MBI does not exhibit reflections caused by MBI and a presence of MBI was confirmed only by the analysis of correlation function. The incorporation of MBI molecules into matrices is observed through optical IR absorption spectroscopy (FTIR) and photoluminescence. Introducing MBI in ChA and MS is followed by the appearance of bright green photoluminescence, the spectral structure of which is analogous to MBI crystals but slightly shifted in the blue region, probably due to a quantum-size effect. The influence of MBI inclusion in PG and ChA on the permittivity, dielectric losses, conductivity, and parameters of their hopping conductivity is analyzed.

1. Introduction

Modern materials science is a multifaceted field of knowledge in which fundamentally new ideas and directions are being developed, at the present stage, which are related to the creation of nanomaterials of various natures and nanosystems based on them [1]. The design of new materials, the improvement of their structure and properties, and the creation of organic and bioorganic nanomaterials and nanosystems require the use of physical diagnostic tools associated with the use of electromagnetic radiation, including from visible light to hard X-rays, methods based on the scattering of various particles such as electrons, neutrons, and ions, and as atomic resolution methods such as atomic force microscopy and atomic resolution electron microscopy, along with dielectric spectroscopy methods that allow for revealing local correlated interactions.
The main types of nanomaterials that are currently developed are thin films [2,3,4], including two-dimensional and multilayer systems obtained using the Langmuir–Blodgett technique [5], liquid suspensions [6,7], porous materials [8], and filler/matrix nanocomposites [9,10,11]. The best-known matrices used to create nanostructures of the filler/matrix type are carbon nanotubes [12,13], zeolites [14], porous borate glasses [15,16], chrysotile asbestos [17], and mesoporous silica [18,19].
In connection with the successes in the field of molecular biology and bioengineering is the prospect of creating new devices and systems for medicine, electronics, and energy, based on the possibility of embedding organic, semi-organic, or bioorganic molecules into various structures with nanosized cavities, such as nanopores, nanotubes, and voids in the crystal structure of organic and semi-organic crystals [20,21,22,23,24].
Borate mesoporous bioactive glasses have been extensively studied for biomedical applications, in particular, as drug delivery instruments and as micro-fibrous dressings to treat chronic wounds for blood vessel formation for wound healing applications and improving bone healing [25]. Mesoporous silica can also be used for drug delivery, and its particles can penetrate certain cells depending on the type of chemical coating of the particles. When the particles are filled with a fluorescent dye, they can be used as a biosensor that transports the dye through the cell membrane. Note that zeolitic imidazole ZIF-67 nanoparticles show excellent adsorption performances for organic dyes [26].
Of particular interest in this area are nanostructures based on ferroics, i.e., on ferroelectric, piezoelectric (ferroelastics), magnetic, or superconducting fillers of nanocavities [27], as well as materials with plastic phases, which make it possible to create ordered nanostructures from multifunctional materials. Scientific interest in such nanostructures may be associated with the study of size effects, in particular, changes in the physical characteristics of ferroelectrics or magnets, such as the Curie temperature, the type of phase transition, the magnitude of the magnetization or polarization of nanoparticles, and, with the emergence of new properties of multiferroics, the appearance of luminescence, as well as others. From the point of view of practical application, it seems important to study the macroscopic (bulk) magnetic, dielectric, and optical properties of such nanostructures, their dependence on the type of filler, and the possibility of their control.
This work is devoted to the synthesis and study of new filler/matrix nanocomposites fabricated by filling voids of borate porous glasses, chrysotile asbestos, and mesoporous silica with molecules of organic 2-methylbenzimidazole C8H8N2 (MBI). The choice of MBI as a filler was due to several reasons.
First, the relatively small size of planar heterocyclic MBI molecules (the maximum size in the plane of molecule dmol ~0.6 nm [28]) makes it possible for them to penetrate the nanopores of matrices with a diameter Dpore > dmol. MBI crystallizes in a non-centrosymmetric lattice with monoclinic polar space group Pn, which is supported by ferroelectric domain switching [29] and the AFM visualization of domain wall moving under the action of an electric field [30]. Because noncentrosymmetric deformations of crystal lattices are very small and are indetectable by XRD, the MBI crystal structure is described by tetragonal group P42/n (86) with the unit cell parameters a = 13.950(9) Å and c = 7.192(3) Å and a unit cell volume Vcell = 1399.6(1.4) Å3 [28,29]. As was expected, MBI can form nanocrystals in voids for which the volume size and diameter are higher than the Vcell and dmol, correspondingly.
Secondly, pore filling is made possible by placing the matrices in the MBI melt (Tmelt ~174 °C), as well as by vacuum sublimation from the gas phase, since the MBI begins to sublimate before reaching the melting temperature. In addition, as shown by this study, filling the pores of matrices is possible using a solution of MBI in alcohol at room temperature and following alcohol evaporation. Previous investigations of micron-thick MBI textured films prepared by evaporation from a solution at room temperature or by vacuum sublimation on different substrates have shown that MBI crystallizes in elongated crystallites with a needle shape, forming on substrate in-plane blocks made of splitted crystals like spherulites [31,32,33]. It is important to note that MBI has a relatively small chemical activity. The inclusion of MBI molecules cannot destroy the material of matrices or the void structure. The penetration of MBI molecules into voids can be followed only by the appearance of hydrogen bonds between ions of the matrix and MBI that cannot affect the matrix material [24]. Nevertheless, the appearance of such bonds, most probably, can influence the properties of MBI nanocrystals.
Third, bulk MBI samples demonstrate ferroelectric properties above room temperature [29,31], almost up to the melting point, because the Curie temperature of MBI at normal pressure is close to or above the melting point [29]. Spontaneous polarization Ps in MBI occurs in directions of the [110]tetra type, perpendicular to the pseudotetragonal axis [001], and amounts to Ps ~5 μC/cm2. Dielectric hysteresis loops at room temperature show coercive fields Ec ~2–3 V/μm. An important property of MBI is the ability to switch polarization in different directions, perpendicular to the pseudotetragonal axis, by weak fields [29,31,32]. Therefore, it was interesting to see how the inclusion of MBI molecules affected the dielectric properties of the matrices.
It should also be noted that MBI, like many other organic substances, can be used in printing techniques [34]. In addition, the noncentrosymmetric crystal structure of MBI and the presence of spontaneous polarization in it makes it possible to use it to generate terahertz radiation [35].
The aim of this work was the synthesis of nanocomposites based on various matrices (borate porous glasses (PG), chrysotile asbestos (ChA), mesoporous silica (MS)) filled with organic 2-methylbenzimidazole molecules. Since the pores of borate glass, asbestos, and mesoporous silica tubes have a diameter of several nanometers, it was of interest to study the possibility of the formation of MBI nanocrystals in such systems, to study their crystal structure as well as the effect of the introduction of a filler on the infrared (FTIR) spectra, photoluminescence (PL), and dielectric properties of matrices. The description of the study samples is presented in the Section 3.

2. Results and Discussion

2.1. XRD Measurements

2.1.1. MBI in Porous Borate Glasses

X-ray phase analysis showed that all reflections observed in PG with a pore diameter Dpore of ~2.5 nm, ~7 nm, or ~30 nm (PG2.5, PG7, and PG30), filled by MBI from melt (samples MBI-PG2.5, MBI-PG7, and MBI-PG30, respectively), can be attributed to the MBI crystal structure (Figure 1).
A careful inspection of the reflection profiles shows that their broadening in the samples is different. Some are very narrow, others are very wide, and in some reflections, for example, in a reflection with Miller indices hkl = 002, split is clearly visible. These facts indicate the presence in the samples of nanocrystallites belonging to the same crystalline phase (MBI) but having different sizes and slightly different values of the unit cell parameters and, correspondingly, unit cell volume.
  • Sample MBI-PG2.5
In the MBI-PG2.5 sample, the reflections are characterized by the ratio of the full-width at half maximum (FWHM) FWHM to the integral width Bint, which lies within 0.637 ≈ 2/π < FWHM/Bint < (4∙ln(2)/π)1/2 ≈ 0.939 (Figure 2), i.e., they are pseudo-Voigt (pV)-type reflections [36]. Figure 2 shows the distribution of the crystallite sizes, D0hkl, of the MBI-PG2.5L and MBI-PG2.5M phases depending on the Bragg angles, 2θB, of the observed individual XRD reflections (see Section 3.2.2). The horizontal lines in Figure 2 are the mean values, D0, of the phase crystallite sizes obtained by root mean square (rms) averaging of the corresponding D0hklvalues.
As can be seen from Figure 2, there are two sets of reflections with narrow and wide FWHMs, which refer to crystallites with sizes of D0 ~80 nm and D0 ~15 nm. Taking into account that, according to the X-ray phase analysis, all the observed reflections belong to the MBI crystal structure, and also that the FWHM is inversely proportional to the crystallite size, these phases will be referred as MBI-PG2.5L (large size) and MBI-PG2.5M (mean size). The designations “mean” and “large” are used since the estimated crystallite sizes are larger than the average pore size in glass, which is ~2.5 nm.
Note that, for the MBI-PGI2.5M phase, the experimental D0hkl points lie well enough on the horizontal line corresponding to the average value of D0 = 15(5) nm. In contrast to that, for the MBI-PG2.5L phase, there is a systematic deviation of the experimental points from the horizontal line. The experimental values of D0hklfall better on a straight line with a negative slope. Taking into account that the instrumental broadening has been corrected, it is obvious that such a distribution of individual values of D0hkl over the Bragg angles 2θB of the corresponding reflections is due to the absence of microstrains in the MBI-PG2.5M, in contrast to the MBI-PG2.5L phase, in which the presence of microstrains is expected.
Indeed, the Williamson-Hall plot (WHP) and size-strain plot (SSP) [37] graphs, constructed using the individual reflections of MBI-PG2.5 (Figure 3), show that, instead of being grouped around a single line graph, the WHP and SSP experimental points fall into two separate regions that can be approximated by line graphs and correspond to the MBI-PG2.5M and MBI-PG2.5L phases.
As expected, the MBI-PG2.5M phase is characterized by the absence of microstrains (εs = 0), whereas a non-zero microstrain level was observed in MBI-PG2.5L. The average crystallite size in the MBI-PG2.5M phase, obtained by rms averaging of the sizes of the crystallites corresponding to individual MBI-PG2.5M reflections, is D = D0 = 15(5) nm. The values of crystallite size and microstrain obtained from the WHP and SSP plots for the MBI-PG2.5L phase are close: (D = 100(36)–119(20) nm, εs = 0.13(6)–0.16(4)%). Numerical results of the WHP and SSP calculations are shown in Table 1.
Therefore, both methods, WHP and SSP, give the same result: no contribution of microstrains to the broadening of reflections of the MBI-PG2.5M phase and the presence of microstrains in the MBI-PG2.5L phase.
The SSP graphs are characterized by a relatively smaller spread of experimental points around the approximating lines Y = A + BX compared to WHP, which is expressed by significantly higher values of the coefficient of determination Rcod = 91.52–96.00% in comparison to Rcod = 25.52–44.97% for WHP (see definition of Rcod in [37]). As a result, the A and B coefficients and, calculated from them, the D and εs values (see [37]) are characterized by smaller estimated standard deviations (e.s.d.s) in the case of SSP calculations. Therefore, the more precise values obtained during the SSP estimates will be discussed further.
An attempt at Rietveld fitting within the framework of the model of two phases of MBI with larger and smaller crystallite sizes (MBI-PG2.5L and MBI-PG2.5M with crystallite size ~100 nm and ~14 nm, respectively) led to a rather small value of the weighted profile factor, Rwp = 1.64% (atomic coordinates of the starting MBI model (CCDC code 1199885) were fixed for all phases during the refinement). However, the value of the weighted profile factor corrected for the background contribution, which better characterizes the quality of fitting reflection profiles, remained high, cRwp = 27.92%. The assumption of the presence of a third phase with a small crystallite size of the order of 2.5 nm pore diameter (MBI-PG2.5S) led to a noticeable decrease in all the factors of agreement (Rwp = 1.36%, cRwp = 18.86%). Despite the fact that the reflections of this MBI-PG2.5S phase cannot be visualized in the XRD pattern due to the fact that they are too broad, a significant improvement in the quality of the fit indicates its presence in the MBI-PG2.5 sample.
The results of the Rietveld fitting for the three-phase model (MBI-PG2.5L + MBI-PG2.5M + MBI-PG2.5S) are graphically illustrated in Figure 4 and summarized in Table 1.
As can be seen from Table 1, all MBI phases are characterized by close unit cell parameters. However, the unit cell volume, Vcell, of the MBI-PG2.5M phase is close to the Vcell of the MBI compound described in the CCDC databank (CCDC code 1199885). At the same time, the MBI-PG2.5L and MBI-PG2.5S phases are characterized by ~1–2% larger unit cell volumes. The sizes of the crystallites of the MBI-PG2.5L and MBI-PG2.5M phases (respectively, DTOPAS = 107(1) nm and 12(1) nm, as well as microstrain εsTOPAS = 0.082(1)% for the MBI-PG2.5L phase) obtained in the Rietveld refinement agree well with the values obtained in the SSP and WHP calculations, but are characterized by a smaller e.s.d.s. The refinement results in the size of the crystallites of the MBI-PG2.5S phase, DTOPAS = 2.4(2) nm, being almost equal to the pore size of the sample MBI-PG 2.5. The weight content of the MBI-PG2.5S phase with the smallest crystallite size is the largest, being 73.7(7) wt.%. At the same time, the weight content of the MBI-PG2.5L phase with the largest crystallite size is the smallest, only 0.80(2) wt.%. The remaining 25.5(7) wt.% of the weight content of the sample corresponds to the MBI-PG2.5M phase with an intermediate crystallite size.
Thus, various methods lead to the fact that the sizes of crystallites are not uniformly distributed around a mean value (Figure 2 and Table 1). The crystallite sizes of two phases (MBI-PG2.5L and MBI-PG2.5M) significantly exceed the diameter of the Dpore pores, ~2.5 nm, and the size of the crystallites of the MBI-PG2.5S phase is approximately the value of Dpore. The presence of crystallites with a size that is considerably higher than the size of the pores may be explained by taking into account two factors. The first one is that, for the θ–2θ scan mode used in our measurements, the XRD measures the size of the crystallite in the direction that is perpendicular to the diffracting interatomic planes, i.e., perpendicular to the surface of the glass sample. The second factor is that the pores in PG2.5 are curved hollow cylinder-like fibers that are distributed arbitrarily, relative to the surface of the sample.
MBI in the parts of the pores lying parallel to the surface of the glass sample will form MBI-PG2.5S crystallites, and MBI in the parts of the pores perpendicular to the sample surface or inclined to it will lead to the formation of MBI-PG2.5L and MBI-PG2.5M crystallites with larger sizes. Obviously, if the porous glass was not prepared in a special way, there cannot be many areas which result in the formation of large crystallites. As a result, the larger the crystallite size, the lower weight content of the corresponding MBI phase, and, accordingly, the smaller the relative volume occupied by the phase.
Since the MBI-PG2.5 XRD patterns (Figure 1 and Figure 4) contain reflections characterized by different Miller indices hkl (110, 011, 111, etc.), the crystallites of different MBI phases in the pores do not have exactly the same orientation. Nevertheless, as the Rietveld fit (Table 1) and comparison with the theoretical XRD pattern of MBI (cf. Figure 4 and Figure S1 of Section S1.1 of Supplementary Materials, simulated using structure data of CCDC 1199885 card by means of program PowderCell v.2.4 [39]) show, most of the crystallites of the MBI-PG2.5L, MBI-PG2.5M, and MBI-PG2.5S phases are oriented along the crystallographic directions [110], [001], and [211], respectively (i.e., the crystallographic planes (110), (001), and (211) in the crystallites of the corresponding phases lie preferentially parallel to the sample surface).
It should be noted that the sizes of the crystallites of the phases MBI-PG2.5L and MBI-PG2.5M are much larger than the pore diameter in glass, which indicates that the crystal structure of MBI is correlated over a large number of pores. Correlation manifests itself not only in the parameters of the MBI unit cell, but also in the orientation of the MBI crystal axes, discussed above, since each phase has its preferential orientation, described by the March-Dollase parameter, rMD (see Table 1). The appearance of large areas with a correlated crystal structure and orientation of the filler (i.e., crystallites) has also been observed in different nanosystems [40,41,42].
  • Sample MBI-PG7
The analysis of the XRD reflection profiles using the WHP and SSP methods shows that, for the MBI-PG7 sample, it is possible to distinguish three characteristic average sizes of crystallites which correspond to three MBI phases, MBI-PG7L, MBI-PG7M1, and MBI-PG7M2 (see Section S1.2 and Figures S2 and S3a,b of Supplementary Materials and Table 2).
The crystallite sizes of these phases are ~60 nm (MPB-PG7L phase), ~28 nm (MPB-PG7M1 phase), and ~17 nm (MPB-PG7M2 phase). However, the Rietveld fitting indicates the presence of a fourth phase, MPB-PG7S, with an even smaller average crystallite size. The assumption of the presence of the MBI-PG7S phase with crystallite sizes equal to or smaller than the pore sizes in the MBI-PG7 sample led to a significant decrease in the agreement factors, to Rwp = 1.64% and cRwp = 13.02% from Rwp = 2.53% and cRwp = 30.40%.
The graphical results of the Rietveld fit carried out within the framework of the four-phase model (MBI-PG7L + MBI-PG7M1 + MBI-PG7M2 + MBI-PG7S) are shown in Figure 5. The quantitative results are summarized in Table 2.
The average sizes of the crystallites of the phases MBI-PG7L, MBI-PG7M1, and MBI-PG7M2 (DTOPAS = 61(3) nm, 28(1) nm, and 17(1) nm, respectively) obtained in the Rietveld refinement coincide well with the values of D obtained by the SSP and WHP methods (Table 2), and are characterized by smaller e.s.d.s. The value of the microstrain of the crystallites of the MBI-PG7L phase (εsTOPAS = 0.014(6)%) obtained in the refinement also agrees satisfactorily with the SSP and WHP data (Table 2), taking into account its e.s.d. Refinement by the Rietveld method led to a small size of the crystallites of the MBI-PG2.5S phase, DTOPAS = 2.7(1) nm. As in the case of the MBI-PG2.5S phase with approximately the same small size of crystallites in the MBI-PG 2.5 sample, the broadening of the reflections of this phase is large, and they are visually indistinguishable in the XRD pattern. Nevertheless, the significant improvement in the quality of the fit described above evidences the presence of this phase. According to our quantitative estimates, when refined by the Rietveld method, the weight contents of phases in the composition of the MBI-PG7 sample are greater the smaller their size is (71.33(14) wt.%, 9.71(17) wt.%, 5.70(9) wt.%, and 0.80(2) wt.% for phases MBI-PG7L, MBI-PG7M1, MBI-PG7M2, and MBI-PG7S, respectively). Thus, the mass content of the MBI-PG7L and MSI-PG7S phases with the largest and smallest crystallite sizes, respectively, and the total mass content of the MBI-PG7M1 and MBI-PG7M2 phases with intermediate crystallite sizes, are approximately the same as for the similar MBI-PG2.5L, MSI-PG2.5S, and MBI-PG2.5M phases in the MBI-PG2.5 sample. Like the phases of the MBI-PG2.5 sample, all of the crystalline phases of the MBI-PG7 sample exhibit close unit cell parameters. In comparison to CCDC code 1199885 MBI, the MBI-PG7S phase is characterized by a ~1% smaller unit cell volume, Vcell, whereas the Vcell of all of the other phases is ~1–1.5% larger.
Table 1 and Table 2 show that the relative volume, Vi, occupied by a certain phase, i, in the MBI-PG2.5 and MBI- PG7 samples is roughly inversely proportional to the crystallite size Di (Vi~Di−1). Since the volume occupied by phase i is Vi = Nvi, where Ni is the number of crystallites of phase i, and vi~Di3 is the volume of a crystallite of phase i, we can assume that the number of crystallites of phase i is Ni~Di−4.
Thus, the WHP and SSP analysis of the XRD reflection profiles and Rietveld fitting of the XRD pattern leads to a model of the MBI-PG7 sample that is similar to the model of the MBI-PG2.5 sample of the MBI phases with close unit cell parameters and large, intermediate, and small crystallite phases. The only difference is that there is not one crystalline phase with an intermediate crystallite size, as in MBI-PG2.5, but two phases (MBI-PG7M1 and MBI-PG7M2). In addition, for the MBI-PG7S phase, the smallest crystallites are characterized by sizes smaller than the pore size Dpore ~7 nm, in contrast to MBI-PG2.5, where the smallest crystallites are approximately equal to the pore size. As discussed above, the weight contents of the MBI-PG7 phases follow the same tendency, as in MBI-PG2.5, i.e., for the larger crystallite size, a lower weight content of the corresponding MBI phase is observed. Therefore, for PG2.5, we can assume the same model of formation of the MBI phases in the pores of the PG7, which are curved, hollow, cylinder-like fibers that are arbitrarily distributed relative to the sample surface, resulting in the appearance of large areas with a correlated crystal structure and orientation of the MBI filler. As in the MBI-PG2.5 sample, the microstrain is observed in the MBI-PG7L phase, apparently due to the large size of the crystallites of this phase.
In contrast to MBI-PG2.5, in the MBI-PG7 sample, most of the crystallites of the MBI-PG7L, MBI-PG7M1, MBI-PG7M2, and MBI-PG7S phases are oriented along the crystallographic directions [001], [012], [241], and [211], respectively. The preferential orientation of the phases with the smallest crystallite size is the same as in MBI-PG7 and MBI-PG2.5 samples.
It is interesting to note the complete similarity of the XRD reflection profiles in the samples MBI-PG7 and MBI-PG30 (Figure 1), which is probably due to the formation of MBI crystalline phases in these samples with similar cell parameters, sizes, and orientations of the crystallites.

2.1.2. MBI in Chrysotile Asbestos

Two types of chrysotile asbestos (ChA) samples filled by MBI from melt were investigated. The first sample, designated as the MBI-ChA sample, was a plate of ~10 × 10 × 1 mm3 filled with MBI, where ChA fibers in the form of hollow tubes with a diameter of ~9 nm passed in one direction along one of the long sides of the plate and parallel to the surface of the plate. Another one, referred to as MBI-ChA-mill, was the sample MBI-ChA milled into powder.
X-ray phase analysis shows that all observed XRD reflections of the MBI-ChA and MBI-ChA-mill samples are attributed to the MBI structure. Line profile analysis of the XRD reflections using WHP and SSP techniques (Figures S4 and S5a,b in Section S1.3 of Supplementary Materials) results in a MBI crystallite size of about the diameter of the ChA tube, DSSPDWHP = 8.0(2.5) nm, and microstrain εsSSP = εsWHP = 0.
Analysis of the XRD patterns of the MBI-ChA and MBI-ChA-mill samples was performed using the Le Bail (LB) fitting method [43], which allows for fitting without using a structural model and, correspondingly, neglecting the influence of the preferential orientation of crystallites.
However, the assumption of only one MBI-ChA1 phase did not allow us to obtain a sufficient LB fit of the XRD pattern of the MBI-ChA sample. For example, the reflection profile with Miller indices hkl = 130 (Bragg angle 2θB ≈ 21.6°) did not fit properly, and the weighted profile agreement factor was only Rwp = 12.46%.
The LB fitting showed that the best correspondence to the experiment (Rwp = 3.58%) was achieved when two phases, MBI-ChA1 and MBI-ChA2, were present in the MBI-ChA sample, which have slightly different values of unit cell parameters and crystallite sizes a (~14.06 and 13.63 Å) and c (7.16 and 7.13 Å). The same was obtained for the powder sample MBI-ChA-mill (Rwp = 3.87%).
The final graphical results of the LB fitting for MBI-ChA are shown in Figure 6, and for MBI-ChA-mill are presented in Figure S6 of Supplementary Materials.
Table 3 shows the characteristics of these phases that were obtained from LB fitting for both the bulk MBI-ChA and milled MBI-ChA-mill samples.
It should be noted that milling did not lead to the appearance of spherical particles, and the elongated MBI particles were still stacked mainly along the tubular pores of ChA, so the results of both measurements turned out to be very close (cf. results for MBI-ChA and MBI-ChA-mill in Table 3). The crystallite sizes of the MBI-ChA1 and MBI-ChA2 phases in both samples are close to ~9 nm, which coincides with the pore diameter of chrysotile asbestos. The MBI-ChA1 phase occupies about two times more volume of the MBI-ChA and MBI-ChA-mill samples than the MBI-ChA2 phase.
The values of the unit cell parameters in both phases of the samples are close to those observed in bulk crystal (CCDC code 1199885). However, the unit cell parameter a in both phases is smaller than the tabular value, whereas the value of parameter c in the MBI-ChA1 phase is slightly larger and, in the MBI-ChA2 phase, it is noticeably smaller than in the bulk crystal. This is also reflected in the unit cell volume Vcell (Table 3), resulting in a ~1% larger value of Vcell in the MBI-ChA1 phase and, vice versa, a ~5% smaller value in MBI-ChA2.
Thus, the incorporation of MBI into ChA is also accompanied by the appearance of MBI phases. However, unlike porous glasses, MBI-ChA contains only two phases. The size of the MBI crystallites of these phases is close to the diameter of ChA nanotubes (Dpore ~9 nm), evidently because the ChA nanotubes run parallel to each other and the surface of the ChA sample, and the XRD (θ–2θ scan mode) measures the size of the crystallite in the direction that is perpendicular to the surface of the ChA sample. Phases with larger or smaller crystallites were not found in MBI-ChA. This means that there are no correlations of the MBI crystal structure between different chrysotile asbestos nanotubes.
The close but nevertheless different crystallite sizes of the MBI-ChA1 and MBI-ChA2 crystalline phases could be explained by two possible models. In the first case, the diameter of the ChA nanotubes is the same throughout the sample and is approximately equal to the maximum size of the crystallite (phase MBI-ChA1, D ≈ 9.7 nm averaged by MBI-ChA and MBI-ChA-mill samples). The phase MBI-ChA2 is characterized by a smaller crystallite size, D ≈ 8.75 nm (value, averaged by MBI-ChA and MBI-ChA-mill samples), since it does not fill the entire volume of the hollow ChA nanotube. In the second case, probably closer to reality, there are two characteristic close diameters of the nanotube in the ChA sample which is completely filled with MBI, Dpore ≈ 9.7 nm and Dpore ≈ 8.75 nm.
Visual inspection of the measured and the simulated theoretical MBI XRD patterns (Figure S6 of Supplementary Materials and Figure 6 in comparison with Figure S1 of Supplementary Materials) leads to the conclusion that the samples are strongly influenced by the effects of preferential orientation. The preferential orientation effects in the ChA samples filled with MBI are much stronger than those in porous glasses (for example, reflection 110, 021, etc. are not seen in the XRD patterns of the ChA samples (cf. Figure S6 of Supplementary Materials and Figure 6 comparing with Figure 4 and Figure 5)), which is expected from the location of the ChA nanotubes in the ChA samples. Consideration of the results of the LB fitting indicates that the main preferential directions of the MBI-ChA1 and MBI-ChA2 phases are different. For the MBI-ChA1 phase, the directions are [100] and [001], and for the MBI-ChA2 phase, they are [001] and [013].

2.1.3. MBI in Mesoporous Silica (MBI-MS)

In contrast to the MBI-PG2.5, MBI-PG7, and MBI-ChA bulk-plate samples, the sample of mesoporous silica (MS), with a cylindrical pore diameter of Dpore ~3 nm and filled with MBI (referred as MBI-MS), was a powder. In contrast to the MBI-PG2.5, MBI-PG7, MBI-ChA, and MBI-ChA-mill samples, the XRD patterns of the MBI-MS sample do not show the presence of reflections attributed to MBI or any other phase. Only a halo due to the amorphous SiO2 matrix is observed (Figure S7 in Section S1.4 of Supplementary Materials).
Therefore, to detect MBI, total correlation functions c(r) were built for the experimental XRD patterns of the MBI-MS and pure MS according to the method from [44] (Section 3.2.5 and Section S2 of Supplementary Materials).
In Figure 7, the symbol f indicates a false peak, which arises due to the breakage of the Fourier series. The numbers indicate the correlation maxima, the positions of which correspond to the average distances between atoms in the amorphous material under study and can be correlated with the average interatomic distances in the crystalline material of the same composition.
The positions of the c(r) maxima, i.e., the observed correlation distances for MBI-MS and pure MS, coincide well with each other and with the positions of the sodium-borate mesoporous alumina oxide (MS5) glass with a pore size of Dpore ~5 nm and a composition of 0.2Na2O·3.8B2O3·96SiO2 (mol.%) from [44].
The peaks of the correlation function c(r) of MBI-MS are characterized by lower heights (intensities) and being somewhat wider than the c(r) of pure MS. In [44], the wider c(r) peaks of lower intensity were also observed for natrium-borate bulk glass 18.1Na2O∙16.9B203∙65SiO2 (mol.%) in comparison with MS5 glass. It was concluded that this is caused by the greater disorder of the local bulk glass structure compared to MS5. It is also possible that MBI-MS is characterized by a greater disorder of the local structure compared to pure MS due to the entry of MBI into the pores of the MBI-MS sample.
Thus, in contrast to MBI-ChA and MBI-PG, there are no reflections on the MBI-MS XRD pattern. Apparently, this is due to the relatively low filling of the MS nanotubes with MBI filler and the lack of correlation of the MBI structure in neighboring MS nanotubes. As a result, the MBI-MS nanocrystallites are very small and, consequently, the XRD reflections are very broad. Nevertheless, studies have shown that the correlation functions, c(r), in MBI-MS are characterized by lower intensities and are somewhat wider than in pure MS, which may indicate the presence of MBI nanocrystallites.

2.2. FTIR and Photoluminescence Measurements

2.2.1. IR Absorption in ChA and MBI-ChA Nanostructure

To obtain IR absorption spectra, FTIR measurements were performed on thin fibers split from a bulk MBI-ChA sample. The measurements were carried out in the transmission mode. The methods of reflection and frustrated total reflection did not give the best results. The spectrum of MBI-ChA, shown in Figure 8 (see also Table 4), is consistent with the literature data on IR absorption in MBI crystals [45,46] and in ChA [47].
The main absorption peaks of ChA are observed in the spectral region ν = 900–1200 cm−1. Three modes at ν = 986, 1008, and 1088 cm−1, usually attributed to the Si–O valence modes, were observed in our samples. The peak at 1088 cm−1 corresponds to the stretching vibrations of Si-O bonds in the Si-O-Si system [47]. The wider intense bands at 986 and 1008 cm−1 are associated with antisymmetric vibrations in the Si-O-Si complexes.
In the MBI-ChA sample, in addition to ChA peaks, two groups of lines in the ranges of 1200–1750 cm−1 and 2500–3200 cm−1 are observed in the spectrum, and are associated with vibrations in the MBI molecules [45,46]. Table 4 shows the positions of the main absorption peaks and their interpretations from [46].
A broad band at ν = 1635 cm−1 that was observed earlier in ChA in Reference [47] is also presented in our sample at 1639 cm−1 (Figure 8). Note that this band has also been observed in some crystals containing water and was attributed to bending vibrations of the water molecules [48,49,50]. The bands at 2846, 2918, and 2993 cm−1 also refer to stretching vibrations of water [48]. The position of water stretching vibrations indicates strong hydrogen bonds and the participation of water molecules in various hydrogen bond structures [48]. The appearance of water vibrations in the absorption spectrum of ChA indicates the presence of water molecules in the pores of the ChA.
Thus, the IR absorption spectrum clearly shows the presence of MBI molecules in the MBI-ChA sample. Since the MBI lines in the IR spectrum are associated mainly with the internal vibrations in the MBI molecule, which are only slightly affected by the environment, the resonance energy and line width are practically the same as in the MBI crystal or in the MBI solution. In the MBI-PG and MBI-MS samples, IR absorption from the MBI was not observed in the FTIR experiments because of very high contribution to absorption from the matrixes.

2.2.2. Photoluminescence of MBI Single Crystals

The bulk MBI crystals under study were prepared by the condensation of MBI vapor. The crystals were almost colorless, with a slight beige tinge, and with an average volume of about 1 mm3. The room temperature luminescence spectrum of the crystals upon excitation with 405 nm light (Ephoton = 3.061 eV) is shown in Figure 9.
The spectrum consists of a series of broad overlapping bands occupying almost the entire visible region. To our knowledge, the energy gap, Eg, in MBI crystals is unknown. On the other hand, a sharp increase in optical absorption in the MBI solution in ethanol, which can be related to the highest occupied molecular orbital-lowest unoccupied molecular orbital (HOMO-LUMO) transitions in MBI molecules, was found to start at a photon energy with ≥ 3.8 eV. Apparently, this value can be considered in the first approximation as an estimate of the Eg value in MBI crystals. Therefore, the emission of MBI crystals observed in the region < 3 eV (Figure 9a) should be ascribed to the recombination of free or localized excitons (hereafter referred to as emission centers). The observed structure of the emission spectrum can be due to electron transitions from excited emissive states in the MBI crystal to vibrationally excited substates of the ground electronic state (Figure 9b).
In such a case, an emission spectrum is expected to consist of the band of a purely electronic transition at energy E0−0 and its vibrational replicas at energies Ei = E0−0ħωi, where ħωi is the energy of the i-th vibration (phonon) involved in the transition (Figure 9b). The spectral contour of MBI emission shown in Figure 9a indicates the presence of at least seven overlapping bands (inset in Figure 9a). Figure 9a also shows the IR vibrational spectra of MBI (according to [45]). As can be seen, there is an obvious correlation of the positions of the vibronic emission bands with the vibrational spectrum of the crystal. The best agreement between the considered model of vibronic optical transitions and the vibrational spectrum of the crystal is achieved by assuming the existence of two emission centers (emissive states) in the crystal with energies of purely electronic transitions, E 0 0 ( 1 ) and E 0 0 ( 2 ) (Figure 9a).
The MBI molecule consists of imidazole and benzene rings (see inset in Figure 9c). In studying this connection, it is of undoubted interest to understand the role of these rings in the formation of the emission spectrum of the molecule. The emission spectra of imidazole (image of imidazole molecule is shown in Figure 9c) and MBI crystals at low temperatures turned out to be the most indicative in this respect. As can be seen from Figure 9c, the structures of the low-temperature spectra of these crystals are almost identical. This suggests that the emission spectrum of MBI crystals is most likely formed mainly by electronic transitions in the imidazole ring. The detailed interpretation of the MBI low-temperature emission spectrum will be given elsewhere.

2.2.3. Photoluminescence of MBI-ChA

“Pure” ChA is practically nonluminous at room temperature. The introduction of MBI into the pores of asbestos is accompanied by the appearance of quite bright photoluminescence, the spectrum of which is represented in Figure 10. It is known that, compared with bulk crystals, nanocrystals have a larger fraction of surface atoms and, accordingly, an increased role of surface states, which often serve as channels for the radiationless deactivation of electronic excitations and thereby contribute to a decrease in the photoluminescence intensity. The relatively bright luminescence of MBI nanoparticles in ChA shows that the effect of such states on the luminescence intensity of MBI nanocrystals is small.
As in the case of pure MBI crystals, the short-wavelength part of the MBI emission spectrum in MBI-ChA is well structured and similar to the structure of the spectrum of pure MBI. It is also noticeable that the bands in the spectrum of MBI in asbestos are broader than in the emission spectrum of the crystals and are shifted to the shorter wavelength side by ~16 meV.
A slight shift of the emission spectrum of MBI nanoparticles in asbestos to a shorter wavelengths relative to the spectrum of bulk MBI crystals can be attributed to the quantum-size effect in MBI nanoparticles. This effect has been theoretically investigated in nanoclusters of some polycyclic aromatic hydrocarbons [51]. It was shown that, in such nanoclusters, the HOMO-LUMO gap increases with a decreasing number of molecules in the nanocluster. In this case, the observed broadening of the bands in the spectrum of nanoparticles can be explained by the inhomogeneous broadening of the spectrum due to the MBI-nanoparticle size dispersion in ChA.

2.2.4. Photoluminescence in Porous Borate Glasses Filled with MBI

As in the case of asbestos, the introduction of MBI into porous borate glasses is accompanied by the appearance of bright photoluminescence. The room-temperature photoluminescence spectra of glasses with 2.5 and 7 nm pores filled with MBI are shown in Figure 10b. Comparison of the spectra depicted in Figure 10b shows that the spectrum of the MBI-PG2.5 glass composite is very close to that of the bulk crystal. The small “blue” shift in the composite spectrum with respect to the crystal spectrum can be attributed to the nanoparticle-size effect mentioned above. The position of the emission bands in the 7 nm glass, where the larger pore size allows for the formation of larger nanoparticles, already coincides with the position of the spectral bands in the bulk crystal. The large pore size in 7 nm glass also appears to contribute to the relatively large dispersion of MBI nanoparticle sizes, resulting in significant inhomogeneous broadening of its emission spectrum (Figure 10b).

2.2.5. Photoluminescence of Mesoporous Silica Filled with MBI

The photoluminescence spectrum of MBI introduced into mesoporous silica with ~3 nm pores is shown in Figure 10c. The spectrum consists of a broad, weakly structured band with a maximum at max ≈ 2.370 eV (λmax ≈ 507 nm). Note that the luminescence intensity of MBI in MS is weaker than that of MBI in asbestos and borate glasses. The absence of a pronounced structure makes it difficult to directly compare the emission spectrum of MBI nanoparticles in silica with the spectrum of bulk MBI crystal. The difference in the shape of the spectrum of MBI in porous silica from the above-described emission spectra of MBI in asbestos and porous glasses, as well as MBI crystals, can indicate the presence of a specific interaction between the MBI molecules and silica matrix. Due to the synthesis conditions of mesoporous silica, its surface contains a significant number of hydroxyl groups that are capable of forming hydrogen bonds with MBI molecules on the nanoparticle surfaces, which affects the electronic states of the MBI molecules and the vibrational spectrum of the associates.
The perturbations of the electronic and vibrational spectra are expected to be different for MBI molecules in different microenvironments in the pores of an MS matrix. This leads to an additional inhomogeneous broadening of the emission spectrum of the nanoparticle ensemble, which is also characteristic of solutions. In this connection, it can be noted that the shape of the spectrum of MBI in mesoporous silica appears to be close to the shape of the emission spectrum of MBI solution in ethanol (Figure 10c), which is also a proton donor. At the same time, the perturbation of the spectra is relatively small, so that the main emission still falls on the green region of the spectrum, as in the case of the emissions of MBI in ChA and in borate glasses.
The formation of hydrogen bonds between MBI molecules and the MS matrix can also account for the relatively weak luminescence of MBI in MS, which is caused by the effect of hydrogen bonds on the rate of the radiationless deactivation of the molecules. Another reason for the lower intensity of the emission of MBI in the MS matrix is the lower content of MBI per unit volume of the MS matrix compared to its content in ChA and borate glasses.

2.3. Dielectric Properties

2.3.1. MBI-ChA

Figure 11 shows the frequency dependencies of the real and imaginary parts of the effective permittivity, εeff and εeff, of the ChA and MBI-ChA samples, measured in the frequency range f = 25–106 Hz for an electric field applied perpendicularly to the nanotubes. In the low-frequency region (f ~102–104 Hz), we see considerable differences in the values of both the real and imaginary parts of the effective permittivity, εeff and εeff, of the samples. The introduction of MBI in nanotubes of ChA leads to a significant decrease in the effective dielectric constant, εeff, and tgδ. In particular, at f = 100 Hz, the εeff and tgδ = εeff/εeff in MBI-ChA are almost two times smaller than in the “empty” ChA.
Large values of εeff and tgδ in ChA samples at low frequencies are associated with the response of molecules, in particular water molecules, adsorbed in nanotubes from the air to a driving electric field [52]. With the introduction of MBI, the permittivity of ChA-MBI approaches the values observed in MBI crystals [31], which indicates the absence of water molecules in the nanotubes.
At high frequencies (f ~105–106 Hz), the values of εeff both in ChA and MBI-ChA are almost the same, εeff ≈ 10. This is due to the fact that the permittivities in MBI crystals (ε′ ~8) and ChA ′ ~10) are very close in this frequency range [53]. At f = 1 MHz, the values of the εeff and dielectric losses in MBI-ChA and ChA are also close. The value of tgδ in MBI-ChA (tgδ ≅ 0.08) is even lower than in ChA (tgδ ≅ 0.12) (Figure 11b). This is because, at a high frequency, the contribution of water molecules is considerably reduced.
Figure 11c shows the frequency dependencies of conductivity (σ = ωεε0tgδ) in the MBI-ChA and ChA samples in a double logarithmic scale. At low frequencies, the conductivity of MBI-ChA is close to that of MBI crystals and much lower than that of ChA. The large values of ChA conductivity are due to the adsorption of water molecules in the nanotubes. In both samples, strong frequency dependencies on conductivity are observed, including in the low frequency region. This indicates that the direct current (DC) conductivity is low and does not prevail over the alternating current (AC) conductivity even at low frequencies (f ~100 Hz). At high frequencies, the slopes of frequency dependence on conductivity in ChA and MBI-ChA have different values: 0.8 and 0.94, respectively.

2.3.2. MBI in PG

Figure 12 shows the frequency dependencies of the effective permittivity, εeff—the real part of the dielectric constant (a), imaginary part of dielectric constant, εeff, (b), and conductivity, σ, (c) in the porous glass samples PG7 and MBI-PG7. Similar to the case of the ChA and MBI-ChA samples, significant differences between the porous glass samples without MBI (PG7) and the samples with MBI included in the pores (MBI-PG7) are observed in the low-frequency region. The effective permittivity, εeff and εeff, are much higher in the porous glass without MBI. The permittivity, εeff and εeff, is significantly reduced when MBI is introduced into glass nanopores. The values of the real εeff and imaginary part εeff of the effective permittivity in MBI-PG7 are close and even somewhat lower than in MBI crystals, which indicates the absence of water molecules in the pores.
At high frequencies, the value of the effective permittivity of PG7 is lower than in the MBI-PG7 samples, which confirms the inclusion of MBI in the pores of the glass, since the value of the permittivity of glass without pores is ε′ ≈ 5, which is almost half the permittivity of MBI crystals.
At low frequencies (f = 10–104 Hz), the magnitude and frequency dependencies of the conductivity in the PG7 and MBI-PG7 samples show significant differences (Figure 12c). For f < 100 Hz, the conductivity in PG7 is practically frequency independent, which indicates the prevalence of DC conductivity with a value of σDC ≈ 8⋅10−8 S/m. In MBI-PG7, the conductivity at low frequencies is much lower and strongly depends on the frequency. This indicates that the regime of dominance of DC conductivity is not achieved. At a high frequency, the conductivity in both of the samples shows close frequency dependencies, with a slope of 0.90 ± 0.01 in PG7 and 0.94 ± 0.01 in MBI-PG7.
To show the effect of pore size on the dielectric characteristics of porous glasses filled with MBI, measurements were carried out on samples with pore sizes of 2.5 nm (MBI-PG2.5), 7 nm (MBI-PG7), and about 30 nm (MBI-PG30). Frequency dependencies of the real part of the effective permittivity, εeff, in MBI-PG7, MBI-PG2.5 (a), and conductivity, σ, (b) in the MBI-PG7, MBI-PG2.5, and MBI-PG30 samples are shown in Figure 13. The inset shows the frequency dependence of permittivity, εeff, in the MBI crystals along the [110] direction.
Figure 14 presents frequency dependences of the conductivity, σ, in MBI-PG2.5 at different temperatures (a) and temperature dependencies of conductivity, σ, in the MBI-PG7 and MBI-PG2.5 samples for f = 60 Hz (b). The inset shows the dependence of the conductivity, σ, on the inverse temperature in the MBI-PG7 and MBI-PG2.5 samples.

2.3.3. Calculations

For calculations of effective permittivity in the MBI-PG7 sample, we used the theoretical expressions of the dielectric response of ferroelectric-dielectric composites (Equation (1) [54] and Bruggerman equation (Equation (2) [55]) for dielectric composites obtained in the effective medium approximation (EMA). The first approach considers the composite formed by the dielectric matrix with permittivity εd, volume concentration q and ferroelectric inclusions with permittivity εf, and gives the following expression for effective permittivity, εeff:
εeff(q) = ¼[−εd + 3d + 2εf − 3f + (8εdεf + (−εd + 3d + 2εf − 3f)2)1/2].
The second approach gives the possibility of calculating the εeff of a composite formed by N components with a volume concentration qj, and with a permittivity εdj, by solving the equation:
j = 1 N q j · ε e f f ε d j ε d j + 2 · ε e f f = 0 ,
In calculation, we used the values of the dielectric constant and the volume of PG matrix, εd = 5 and q = 0.75–0.8, respectively.
The calculated effective permittivity, εeff for Equations (1) and (2) is shown in Figure 13a as red and green lines, correspondently. The calculations were carried out using the experimental dependencies of εeff for the MBI crystal shown in the inset of Figure 13a and the value of the permittivity for glass without pores εd = 5. It can be seen from Figure 13a that a good agreement between the experimental and calculated values of εeff in the MBI-PG7 and MBI-PG2.5 samples was achieved. In Equation (2), we also take into account, besides glass and MBI, empty pores with a volume of about 3% on the sample volume. This gives a somewhat better description of the experimental data. It should be noted that, during preparation, the entire pore volume is occupied by the MBI melt, but small empty volumes may appear due to a decrease in the volume of MBI after crystallization.
Equations (1) and (2) show that the effective permittivity is determined by the total volume of pores in the sample, and not by their linear size. Samples MBI-PG7 and MBI-PG2.5 are characterized by approximately the same values of permittivity, as follows from Figure 13a. This indicates that, at the same permittivities of the matrix and filler, the effective permittivity of the composite is determined only by the volume occupied by the components, which is practically the same in MBI-PG7 and MBI-PG2.5. The same pore volume for both MBI-PG7 and MBI-PG2.5 is determined only by the same composition of borate glass used to create PG7 and PG2.5 nanoglasses. The calculated dependencies of the effective permittivity obtained within the framework of the models used are consistent with the experimental data.
The AC conductivity of a material, σAC = ωεε0tgδ = ωεε0 (where ε0 is the permittivity of a vacuum), is usually described by the expression [56,57]:
σAC = σDC + s,
where σDC is the DC conductivity, ω = 2πf is the circular frequency, A is the constant, and s is the exponent of frequency dependence. In the case of hopping conductivity 0.5 < s ≤ 1 [56,57].
At room temperature, only in the PG7 sample, the conductivity has no frequency dependence at low frequencies (Figure 12c) and σDC ≈ 8⋅10−8 S/m. For MBI-ChA and MBI-PG7, the conductivity has a strong frequency dependence in the studied frequency range, which means that the DC conductivity mode is not achieved DC << σAC) (Figure 11c and Figure 12c).
At high frequencies, f ~ 105–106 Hz, in all samples, σ is frequency dependent. Linear approximation of σ(f) dependence in the range f = 0.1–1 MHz gives the value of the exponent as s = 0.80 in ChA, s = 0.94 in MBI-ChA, and s = 0.90 and 0.94 for PG7 and MBI-PG7, correspondingly (Figure 11c and Figure 12c). The non-unity in the exponent values may indicate specific conductivity mechanisms [58,59]. The most probable of these seems to be hopping charge transfer between localized states separated by an energy barrier. The correlated barrier hopping (CBH) model considers carrier hopping between barrier-separated states that are randomly distributed in the sample volume [58]. This model was successfully used to describe the low frequency conductivity in thin films of scandium oxide [58], as well as in betaine phosphate crystals with a 5% addition of BPI [59]. The value of the s parameter makes it possible to estimate the energy difference Wm = 6kT/(1 − s) between the ground state and the free state in which a carrier can travel over the lattice. The described calculations yield Wm ≅ 2.58 eV for the MBI-ChA nanostructure and Wm ≅ 0.77 eV for the ChA. The values of Wm and ε can now be used to determine the Bohr radius, a, of a localized carrier, a = e2/(2εWm) ≈ 4.4 Å for the MBI-ChA and 12 Å for ChA. Therefore, the incorporation of MBI molecules in a ChA matrix results in a considerable change in parameters of the charge carrier hopping mechanism. For PG7 and MBI-PG7 nanostructures, our calculations give closer values: Wm ≅ 1.55 eV and 2.58 eV, and a ≈ 5.1 Å and 7.3 Å, respectively. Since the glass used for the preparation of PG has a low dielectric constant and very small dielectric losses, the frequency dependencies of the conductivity and dielectric losses of MBI-PG7 are caused mainly by the MBI component.
To reveal the influence of the pore size on the frequency dependencies of the conductivity, samples of porous glass with the inclusion of MBI with pore sizes of 30 nm, 7 nm, and 2.5 nm were studied. The frequency dependencies of conductivity for MBI-PG2.5, MBI-PG7, and MBI-PG30 samples are shown in Figure 13c. The value of the exponent s in MBI-PG7 is s = 0.94 ± 0.01 and, in MBI-PG30, s = 0.91 ± 0.01. These values are close and are considerably higher than s = 0.79 ± 0.01, which was observed for MBI-PG2.5. In MBI-PG2.5, Wm ≅ 0.74 eV and a ≈ 18.1 Å. Note that, in this case, the carrier localization radius approximately corresponds to the radius of the pores filled with MBI.
The frequency dependencies of the conductivity in the MBI-PG7 and MBI-PG2.5 samples at different temperatures are shown in Figure 14. In MBI-PG2.5 (and MBI-PG7), for measurement temperature T > 247 K, the conductivity at low frequencies is frequency independent. This means that the conductivity at low frequencies is determined by σDC, which increases with the temperature (Figure 14a). The linear fitting of conductivity frequency dependences at high frequency reveals a decrease in exponent s with a temperature increase that is in agreement with the CBH model [58].
In both the MBI-PG7 and MBI-PG2.5 samples, an increase in temperature above 340 K results in a drastic increase in the DC conductivity, σDC, which make it possible to calculate the activation energies, Ea (Figure 14b). The inset in Figure 14b shows the dependencies of the conductivity, σ, on the inverse temperature in these samples. The calculated values of Ea in the MBI-PG7 and MBI-PG2.5 samples are equal to 1.22 eV and 1.32 eV, respectively. These activation energies for DC conduction are higher than the Ea ≈ 1.1 eV found in an MBI film [31], which is characteristic of proton conduction in crystals with chains of hydrogen bonds [60]. The parameters s, Wm, a, and Ea for the ChA, MBI-ChA, PG, and MBI-PG structures are shown in Table 5.
In all samples, the exponent s is less than one. In MBI-ChA, PG7, MBI-PG7, and MBI-PG30, in which the sizes of the pores are Dpore ≥ 7 nm, the values of s are close and vary from s = 0.90 to s = 0.94. In MBI-PG2.5, the s is considerably smaller: s = 0.79. This is also reflected in the smaller values of Wm and the larger carrier localization, a.

3. Materials and Methods

3.1. Sample Preparation

Organic nanostructures used in this work were synthesized by introducing MBI molecules into porous borate glasses, chrysotile asbestos, and mesoporous silica. MBI crystals used for nanostructure preparation were grown by evaporation, as well as by slow cooling from saturated solutions in ethanol or acetone (or deuterated acetone (d-acetone)) [61] of MBI powder obtained by chemical methods. The solution was purified with activated carbon. To obtain more perfect crystals, they were repeatedly recrystallized. Also, MBI crystals were obtained from the gas phase by vacuum sublimation. Examples of MBI crystals grown by different methods are shown in Figure 15 along with image of an MBI film with a thickness of ~5 µm obtained by evaporation on substrate.
Depending on the growth conditions, MBI crystals can have different shapes, such as split crystals of the spherulite type, formed in an ethanol solution by elongated crystallites growing radially along the [001]tetra pseudotetragonal axes from one crystallization center (Figure 15a); dendritic-type crystals after sublimation of MBI from the gas phase onto a Pt/glass substrate (Figure 15b); MBI films several microns thick obtained by evaporating an ethanol solution of MBI (Figure 15c), consisting of self-organizing spherulitic blocks [31,32]. Bulk MBI crystals can be prepared from acetone (or an alcohol) solution (Figure 15e) or from the gas phase (Figure 15d).
Description of preparation methods and properties of various types of porous glasses can be found in [62,63,64]. A typical example of porous silica glass is Vycor glass (Corning 7930) with standard chemical composition: 96% SiO2, 3% B2O3, 0.40/a Na2O, R2O3 ± RO2 < 1% (R = Al2O3 or ZrO2). It is made by acid leaching the boron-rich phase in a phase-separated borosilicate glass. The result is a high-content (96%) silica glass containing an interconnected network of pores typically less than 100 Å in diameter. In our experiments, we used Vycor glasses with average pore diameter of Dpore ~2.5 nm, ~7 nm, and ~30 nm which was evaluated by nitrogen adsorption and/or desorption measurements, and a standard porosimetry analysis.
For introducing MBI into pores of Vycor glasses, the templates of glasses with dimensions of ~5 × 5 × 0.5 mm3 were placed in MBI melt (Tmelt ~174 °C) for several days without any external pressure. After removing from the melt and cooling up to room temperature, the samples were mechanically cleaned from tailings of MBI remaining on the surface of the matrix. The analysis of sample weight before and after MBI introduction shows ~80% filling of pores by MBI.
Also, the introduction of MBI into borate glasses was carried out from the gas phase. For this purpose, MBI crystals and a Vycor glass template were placed in an evacuated and sealed quartz tube and heated to a temperature below Tmelt. A few days later, the matrix was filled with MBI at 80% filling of pores. Photos of porous glass before and after filling with MBI are shown in Figure 16a–d.
Chrysotile asbestos, Mg3Si2O5(OH)4—magnesium hydrosilicate, refers to layered silicates. In the case of chrysotile, the layers are twisted into tubules with an inner diameter of 30–60 Å and an outer one, on average, 300–400 Å. Chrysotile asbestos fibers form a hexagonal, dense packing. The walls of the tubes are formed by about 20 double-ribbon layers. One layer consists of silicon-oxygen tetrahedra with hydroxyl groups in the vertex plane. Three oxygen atoms of these tetrahedra are shared with neighboring tetrahedra. The second layer is located on the tops of these tetrahedra. It is composed of hydroxyl groups and magnesium ions and twists the double-ribbon layer into a hollow tube. There are up to 20 such double layers in the wall. Chrysotile asbestos with the inclusion of 2-methylbenzimidazole was prepared by placing and impregnating an asbestos template in an MBI melt. We used natural chrysotile asbestos with a tube diameter of Dpore ~9 nm (Figure 16e).
An important advantage of monodisperse spherical mesoporous silica particles (MSMPs) is the presence of an internal system of cylindrical nanochannels of the same diameter (controllably varied within 2–5 nm, on average ~2.5 nm) with a volume of up to 60% of the particle volume (Figure 16f). MSMPs were synthesized via basic hydrolysis of tetraethoxysilane in a water-ethanol-ammonia mixture containing cylindrical micelles of a surfactant pore-forming agent, cetyltrimethylammonium bromide. The synthesis procedure is described in detail in [65]. The molar ratio of the reagents tetraethoxysilane: NH3:H2O:C2H5OH was 1:60:370:230. The synthesis duration was 1 h. The particles obtained were centrifuged, dried in air at 80 °C for 24 h, and calcined at 550 °C for 5 h. The average diameter of the particles was 510 ± 30 nm, pore size was 3.1 ± 0.2 nm, pore volume was 0.48 cm3/g (~50% of particle volume), specific surface area (BET) was 800 m2/g.
To introduce MBI into the pores of silica particles, a capillary impregnation method was used. For this, a weighed portion of SiO2 particles was impregnated with alcohol solution of MBI under ambient conditions, and then dried at 60 °C. The amount of the solution and concentration of MBI were chosen so that, after embedding into the pores of silica particles, its weight content was 10%. After impregnation of MBI content (determined gravimetrically) into obtained MBI-MS (MS filled with MBI), powder sample was found to be 9.8 wt.%, which, taking the density of the filler to be close to 1 cm3/g, amounted to ~20% of the pore volume.

3.2. Details of X-ray Diffraction Experiment and Analysis

3.2.1. Measurement of XRD Patterns

XRD measurements were carried out on a D2 Phaser X-ray powder diffractometer (Bruker AXS, Karlsruhe, Germany) using an X-ray tube with a copper anode, monochromatized with a Ni filter (Cu-Kα radiation). A LYNXEYE (Bruker AXS, Karlsruhe, Germany) semiconductor linear X-ray detector was used to record XRD patterns.
The measurements were carried out in the symmetrical scanning mode θ-2θ (2θ is diffraction angle and ω = 2θ/2 = θ is the angle of incidence of X-rays on the surface of the sample) in the vertical Bragg-Brentano θ-θ geometry. To reduce the influence of the possible effect of the preferential orientation of crystallites, during measurements, the sample was rotated around the axis of the holder, which coincided with the axis of the diffractometer goniometer.
The XRD patterns of the MBI-PG2.5, MBI-PG7, and MBI-PG30 samples for phase analysis and Rietveld fitting were recorded in the range of 2θ = 6°–85° with a step of Δ2θ = 0.02°. The XRD patterns of the MBI-ChA and MBI-ChA-mill containing more intensive reflections were recorded in a wider range of 2θ = 6°–120°. The amorphous-like XRD pattern of the MBI-MS was measured in maximum possible range 2θ = 5.6°–142° to reduce possible Fourier series breakage errors when constructing a correlation function.
To correct the diffraction patterns for counter zero shift (Δ2θzero) and to correct (during calculations) the angular positions of the reflections for a shift due to a possible misalignment with the focal plane (Δ2θdispl), additional measurements were carried out. For these measurements, the plate samples were embedded in the Si640f X-ray powder standard (NIST, Gaithersburg, MD, USA) in such a way that the surface of the plates was at the same level with the surface of the Si powder, and both the sample surface and the surface of the Si powder fell into the X-ray beam during measurements. The Si reflections recorded on these XRD patterns were used as an internal standard to correct the angular positions of the reflections from the corresponding samples measured together with the standard. In turn, the obtained corrected angular positions of reflections with a strong intensity from the samples were used as an external standard for correcting the angular positions of the reflections of the samples in subsequent measurements without use of the Si640f standard.
X-ray phase analysis of the measured diffraction patterns was carried out using the EVA v.5.1.0.5 [66] program and the powder PDF-2 (Powder Diffraction File-2, ICDD, 2014) database [67].

3.2.2. Determination of Microstructure Parameters from XRD Patterns by Methods of Analyzing the Profiles of Observed XRD Reflections

One of the main reasons for the broadening of XRD reflections is the broadening due to the small size of the crystallites and due to the presence of microstrains in the crystallites. To show the absence or presence of the contribution of crystallite microstrains to the observed broadening of reflections, WHP and SSP graphs were constructed using SizeCr v.11.01 program [37], which takes into consideration the type of the observed XRD reflections (pV type).
For calculations using the SizeCr program, it is necessary to know the parameters of independent X-ray reflections; therefore, strongly superimposed reflections that cannot be confidently separated due to the coincidence or closeness of their angular positions were not taken into account. The parameters of XRD reflections (their observed Bragg angles 2θBobs, FWHMs FWHM, integral (Iint) and maximum (Imax) reflection intensities, and integral width Bint = Iint/Imax) were determined using a combination of TOPAS v.5.0 [68] and EVA [66] programs. To do this, first, the approximate positions of the observed reflections of the MBI phase were indicated in the TOPAS program, assuming a profile type adopted in TOPAS (for example, pV type). The reflection profiles were fitted by refining their angular positions 2θBobs, FWHM, the parameters of the reflection profile, and the background parameters (for modeling the background and emission spectrum, see Section 3.2.5). Further, the fitted profiles of individual MBI reflections were extracted from the simulated XRD pattern and processed in the EVA program, which, after taking into account the background contribution and correcting the contribution of Cu-Kα2 radiation, gave the desired reflection parameters.
The WHP and SSP graph points were calculated from the FWHMs of reflections corrected for the instrumental broadening (designated as FWHMcorr) for pV reflections [69] observed in the measured XRD patterns. The values of the angular positions of the reflections with corrections for Δ2θzero and Δ2θdispl,
2 θ B = 2 θ B o b s + 2 θ z e r o + 2 θ d i s p l · cos θ B o b s ,
were used for calculations. When constructing the WHP and SSP plots, we used the coefficient Kstrain = 4 [70] of the Stokes equation, relating the broadening FWHMstrain due to the presence of microstrain εs with tg(θB), and the Scherrer coefficient KScherrer = 0.94 [71] in the Scherrer equation, and relating the broadening FWHMsize due to the size D of crystallite with cos(θB).
If there is a contribution of microstrains to the broadening of reflections, the average size D of nanocrystallites and the average value of microstrains εs in them is determined from the slope of the linear approximating lines Y = A + B·X of WHP or SSP and the value of their intersection with the Y axis of the graphs [37] (expressions for X and Y of WHP and SSP graphs for pV reflections are shown in the captions of corresponding axes of Figure 3a,b and in [37]).
For WHP method (pV reflections) [37]:
D = K S c h e r r e r · λ A ,
ε s % = B 1 / 2 · 100 % .
In case of SSP technique (pV reflections) [37]:
D = 1 B   ,
ε s % = 2 · A K s t r a i n · 100 % .
Both of the A and B values must be greater than zero in this case. In the case of B < 0 for WHP (Equation (6) or A < 0 for SSP (Equation (8)), it is assumed that the microstrain is zero (εs = 0) [36,37]). Correspondingly, for WHP and SSP, if A ≤ 0 in Equation (5) or B ≤ 0 in Equation (7), then D = ∞ (ideal perfect single crystal) [37].
In the absence of the contribution of microstrains to reflection broadening (model εs = 0), the crystallite sizes D0hkl were estimated for every individual reflection hkl using the Scherrer equation and averaged by the least squares method over the entire set of observed reflections.

3.2.3. Rietveld Quantitative Analysis

Using the angular positions 2θB of the extracted individual reflections (after correction for Δ2θzero and Δ2θdispl according to (4)), the unit cell parameters of the crystalline phases recorded in the samples were calculated by the least squares method utilizing the program Celsiz v.1.1 [72]. The obtained values of the unit cell parameters were used as starting points in the Rietveld fitting [73] of the calculated XRD pattern to the experimental one using the TOPAS v.5.0 program [68]. The coordinates of the atoms of the structure of the MBI phases for the Rietveld analysis were taken from the Cambridge Crystallographic Data Center (CCDC) [74] (code 1199885). The same atomic coordinates were utilized for all MBI phases.
Quantitative Rietveld analysis was carried out. In the course of fitting the simulated XRD patterns to the experimental ones, the weight contents of crystalline phases in the sample were determined by the TOPAS program from refined scaling factors (scale-factors) according to the well-known formalism [75].
Of the structural parameters, only the parameters of the unit cell and the isotropic temperature factors Bisooverall of the atoms, which are common to all atoms in the structures of each phase, have been refined. The atomic coordinates were fixed and not refined for all phases during the quantitative Rietveld analysis.
When modeling calculated XRD patterns in the TOPAS program, the emission spectrum (Cu-Kα doublet) was described by the Berger model of 5 spectral lines with different FWHMs [76]. The standard weight scheme wi = 1/yi was used, where yi is the intensity of the XRD pattern at the point 2θi = 2θstart + (i − 1)∙Δ2θstep. Atomic factors of neutral atoms were used.
The background was modeled using the Chebyshev polynomial of the 7th order and the contribution of a hyperbolic additive for 2θ < 10° range. The amorphous halo (for MBI-PG samples) was considered as an addition to the background and was modeled using the split Pearson VII function (SPVII) [68,77], which allows modeling asymmetric peaks.
The typical course of fitting was as follows. In the start of the Rietveld fitting, after setting the unit cell parameters of the observed crystalline phases to the start values determined in the calculations by Celsiz, the angular corrections Δ2θzero and displ were first refined,
2 θ d i s p l = 2 · d i s p l R g o n · c o s ( θ ) ,
where Rgon is the known radius of the diffractometer goniometer. Since the angular correction Δ2θzero was already introduced into the experimental XRD patterns from measurements of the mixture of the samples with the certified powder, the angular corrections obtained by Rietveld refinement were close to zero.
After that, the unit cell parameters were included in the refinement, but they were refined in different cycles with angular corrections to avoid correlations. The XRD patterns were fitted using instrumentally broadened XRD reflection profiles calculated from the geometry and diffractometer slits used (FP (first principles) profiles in the TOPAS program [68] and FP (fundamental parameters) profiles in [78]), and assuming that the broadening of reflections occurs due to the size of the crystallites and the contribution of microstrains to the broadening. This made it possible to refine the Gaussian and Lorentzian components of the mean crystallite size and microstrain in accordance with the double-Voigt approach [79] adopted in TOPAS, and to calculate their values.
In TOPAS, the crystallite size DTOPAS is designated as Lvol-FWHM. To compare with the microstrain values εsWHP and εsSSP obtained, respectively, in WHP and SSP calculations done by program SizeCr, the microstrain parameter e0 obtained by TOPAS was recalculated to
ε T O P A S % = 2 · e 0 · 100 %
(see [61] for explanation).
In order to compare the obtained crystallite size DTOPAS values with those (DSSP and DWHP) estimated by the SizeCr program, the same Scherrer coefficient KScherrer = 0.94 was used in Rietveld refinement as for SizeCr. Initially, for all crystalline phases, microstrains were not included in the refinement. Based on the results obtained during the construction of WHP and SSP plots, at subsequent stages of refinement, microstrain was included in the refinement for the crystalline phase with largest DTOPAS for which WHP and SSP plots showed non-zero microstrain values (for the phase with the smallest crystallite size DTOPAS ~2.5 nm, for which no reflections were observed in the XRD patterns due to their wide FWHMs, the absence of microstrains was also assumed). As a rule, taking into account the contribution of the microstrain for the phase mentioned above gave an improvement in weighted profile factor Rwp of only ~0.05%.
The inclusion in the Rietveld refinement of the Gaussian and the Lorentzian component of the microstrain parameters for all phases, for which the absence of the contribution of microstrains to the broadening of the reflections was initially supposed, led to εsTOPAS ≈ 0 within e.s.d. and did not improve the quality of fitting of the XRD patterns, which confirmed the suppositions made about the absence of microstrains in these phases. An example of this inclusion of the microstrain in refinement for all phases is shown in Table 3 for MBI-ChA and MBI-ChA-mill samples.
The inclusion in the refinement of the parameters of the preferred orientation of crystallites according to the March-Dollase (MD) [80] model along the crystallographic directions, along which the reflections which showed the greatest intensity, gave a decrease in Rwp by ~1% for all Rietveld fittings done. The influence of other directions of predominant orientation was corrected using the 8th order spherical harmonics model [81], which led to a further decrease in Rwp by ~1.5%. Structural parameters and preferred orientation parameters were refined in different cycles to avoid correlations between them.
At the last stage, the structure parameter was refined, namely, the total isotropic temperature coefficient of atoms, which led to a further decrease in Rwp, as a rule, by ~0.05%. Refinement of the background parameters and the scale factor was carried out at each stage of fitting. The procedure for refining the structural and non-structural parameters was repeated until their change stopped.
The quality of the fitting of the XRD patterns was controlled visually (graphically), as well as quantitatively using a weighted profile agreement factor Rwp. Additionally, at each step, the decrease in other agreement factors, the profile Rp, the weighted profile cRwp, and the profile cRp corrected for the background contribution, as well as the Bragg factor RB, was checked (for the definition of agreement factors, see, for example [68,82]). Weighted profile factor cRwp and profile factor cRp are calculated using program RietEsd v.6.03 [83] due to wrong values given by TOPAS v.5.0, when using the hyperbolic additive to background (see explanations in [83]).
The factor me.s.d., which corrects the e.s.d.s of the refined parameters, obtained in the Rietveld program, for serial correlation according to the procedure [84] by multiplication, was estimated using the RietEsd v.6.03 [83] program.

3.2.4. Le Bail Fitting

The LB fitting [43], which does not require knowledge of atomic coordinates, but only approximate values of the unit cell parameters and the space group of symmetry of crystal phases, was carried out for the MBI-ChA and MBI-ChA-mill samples, characterized by the strongest influence of the effects of the preferential orientation. In the LB method, the intensity of reflections is not calculated on the basis of a structural model, but is extracted from the angular position and reflection profile calculated on the basis of refined unit cell parameters and refined profile parameters directly from the XRD pattern during fitting. As a result, the LB technique, even in the case of strong preferential orientation effects, allows you to get a good quality fitting of the simulated XRD pattern to the experimental one without using preferential orientation models.
The good quality of the LB fitting with a description of the profiles of all observed reflections indicate the presence of crystalline phases in the sample that were used for fitting. The LB method results in obtaining precise refined values of the parameters of the unit cells of the crystalline phases. When using the reflection profiles of the FP type, the LB method also gives, with good precision, the values of the parameters of the microstructure, mean crystallite size DTOPAS, and absolute value εTOPAS of the mean microstrain.
The course of the LB fitting and the list of refined structural and non-structural parameters are completely similar to those described in Section 3.2.3 for the Rietveld fitting, with the exception of parameters related to atoms (temperature factors of atoms, in particular) and with a preferential orientation model, which are not involved for the LB fitting.
Unlike the Rietveld method, the LB method does not allow for obtaining the weight content of the phases by refinement, since there is no information about the content of the unit cell. Nevertheless, taking into account that the molecular weights of different MBI-ChA phases are the same, the weight content of the phases was estimated as the ratio of the areas under the XRD reflections of the MBI-ChAi (i = 1, 2) phase extracted from the simulated XRD pattern of this phase and the total area of reflections of both MBI phases (after subtraction of the background contribution),
                    W t i = I i n t M B I C h A i I i n t c a l c ,
where I i n t M B I C h A i is the integral intensity of reflections of the MBI-ChAi phase, and I i n t c a l c is the total integral intensity of reflections of both phases, MBI-ChA1 and MBI-ChA2.

3.2.5. Calculation of Correlation Function

To analyze the amorphous-like XRD pattern of MBI-MS, the method of total correlation function c(r) was chosen [84]. The construction of this function from the measured XRD pattern is described in more detail in Section S2 of Supplementary Materials.
Briefly, in this method, the experimental XRD pattern I(2θ) is first corrected for the contributions of Cu-Kα2 radiation (for example, using the EVA program) and a small-angle background (utilizing a graphical program, for example, Origin v. 6.0 or later [85]). After that, I(2θ) is also corrected for Lorentz, polarization, and absorption factors (for formulas, see, for example, [68,86,87]) and normalized to the function < f c o r r 2 > at the high diffraction angles 2θ > 90°,
< f c o r r 2 > = ( c i · f i c o r r ) 2 ,
where ci is the atomic concentration of element i in the sample, and ficorr is the atomic scattering factor fi of element i, calculated analytically [88] and corrected for the coefficients Δfi′ and Δfi″ of anomalous dispersion.
By Fourier transformation of the function (S(Q) − 1), a correlation function c(r) is obtained,
c r = 2 · r π Q · S Q 1 · sin Q · r d Q ,
where r is a correlation distance, S(Q) is the total structure factor of the sample,
S Q = I c o r r n o r m < f c o r r 2 > + < f c o r r > 2 < f c o r r > 2 ,
I c o r r n o r m is the XRD pattern after all corrections and normalization,
< f c o r r > 2 = ( c i · f i c o r r ) 2 ,
and all the dependences on the diffraction angle 2θ are recalculated depending on the modulus of the scattering wave vector;
Q = 4 π · s i n ( θ ) λ
where λ is the wavelength of Cu-Kα1 radiation (after correction of the Cu-Kα2 contribution), θ is half the diffraction angle.
The correlation function c(r) does not require information about the average atomic density of the sample for its construction. The maxima of c(r) correspond to the mean interatomic distances in the sample (with the exception of “false” maxima arising from the breakage of the Fourier series, which usually appear at small values of the correlation distance r).

3.3. Optical and Dielectric Measurements

Measurements of the IR absorption spectra (Fourier-transform infrared (FTIR) spectra) were carried out using an IR-Fourier spectrophotometer IRPrestige-21 (Shimadzu Corporation, Kyoto, Japan) with an IR microscope AIM-8000 (Shimadzu Corporation, Kyoto, Japan), both in the specular reflection mode and in the transmission mode, followed by the Kramers–Kronig transformation. The results were then converted to absorbance. The measured spectral range was from 650 to 5000 cm−1. Measurements were carried out on split asbestos fibers shown in Figure 16f.
The photoluminescence spectrum was measured using diffraction spectrometers DFS-24 (LOMO, Saint Petersburg, Russia) and FSD-8 (Optofiber LLC, Moscow, Russia) with 1200 lines/mm and 300 lines/mm gratings, respectively. PMT and CCD matrix were used for radiation detection. Luminescence was excited by CW radiation of a laser operating at a wavelength of 405 nm.
Measurements of capacity and dielectric losses in samples of porous borate glasses and asbestos with and without MBI were performed in the frequency range of 25 Hz−1 MHz with LCR-meters MIT 9216A (Protek, Seattle, WA, USA) and E7-20 (MNIPI, Minsk, Belarus), using the LabView software package (Version 2011, NIST, Gaithersburg, MD, USA). Thin metallic foil was used as electrodes for measurements of capacity and dielectric losses of samples.

4. Conclusions

This study showed that MBI molecules from a melt, gas phase, or solution of MBI easily penetrate into the nanopores of borate glasses, nanotubes of chrysotile asbestos, and mesoporous silica without the use of external pressure, which makes it possible to almost completely fill the PG and ChA pores from the melt or from the gas phase. A lower degree of filling is realized in MS when the pores are filled with an MBI solution in alcohol.
According to our XRD measurements, the incorporation of MBI into PG with pore sizes of Dpore ~ 2.5 and 7 nm is accompanied by the appearance in the matrix of various phases which have the structure of an MBI single crystal with slightly different crystal parameters, paramters a and c. The sizes of the crystallites of some phases can significantly exceed the sizes of PG pores. The appearance of such crystallites indicates the correlation of the crystal lattice in the pores, and the correlation diameter can significantly exceed Dpore. In MBI-PG2.5, there are two such phases with sizes of D ~ 100 nm and 15 nm, while, in MBI-PG7, there are three phases, with D ~ 60, 30, and 18 nm. In both MBI-PG samples, there is also a phase with crystallite sizes that are equal to or smaller than the pore sizes. The volumes occupied by the phase Vi are approximately inversely proportional to the sizes of crystallites Di, and the number of crystallites Ni ~Di−4. Thus, the maximum volume of MBI-PG is occupied by the phase with DDpore, and much smaller volumes are occupied by phases with D > Dpore. In MBI-ChA, two phases are observed with very similar crystal parameters, parameters a and c, which occupy approximately the same volumes. The characteristic size of the crystallites of these phases approximately coincides with the diameter of the nanotubes. This suggests that there is no correlation of crystallites in different ChA nanotubes.
In MBI-MS, the XRD patterns do not show reflections which correspond to the MBI lattice, since the degree of filling of the pores is much less than in the MBI-PG and MBI-ChA samples and there is no correlation of the crystal structure, even if it exists. As a result, the width of the reflections must be very large. The presence of MBI in MS, however, manifests itself in the features of the correlation functions, c(r).
The presence of MBI molecules in the studied structures was confirmed in optical experiments. In particular, the bands corresponding to the molecular vibrations of MBI are well manifested in the IR spectra of the MBI–ChA. In all of the studied structures, MBI-PG, MBI-ChA, and MBI-MS, the introduction of MBI is accompanied by the appearance of strong photoluminescence. The spectral dependence of PL in the MBI-PG2.5 and MBI-ChA samples is similar to the spectrum of the MBI single crystal; however, the observed bands are broadened and slightly shifted to the short wavelength region, which may be due to the manifestation of the nanoparticle size effect. The MBI-PG7 and MBI-PG30 “blue sifts” are much smaller, since the size of the nanoparticles approaches those that are typical for bulk crystals. The widest and most weakly structured PL spectrum is observed in MBI-MS. From a practical point of view, the bright luminescence of these nanostructures makes them promising materials for application as medical biosensors.
The introduction of MBI into PG and ChA is accompanied by a change in the dielectric properties of the matrices. A strong decrease in the effective dielectric permittivity εeff and dielectric loss tgδ at low frequencies is associated with the substitution of MBI molecules in pores and nanotube matrices, for example, water adsorbed from air. Our calculations show that the value of εeff in MBI-PG structures is mainly determined by the total pore volume filled with MBI and weakly depends on the pore size. The conductivity σ of PG, ChA, MBI- PG, and MBI-ChA at high frequencies ω is frequency dependent and is described by a power dependence σ ~ ω s, which indicates a hopping type of conduction. The main parameters of hopping conductivity—energy barrier, Wm, and the carrier localization radius a in structures with pore sizes Dpore ≥ 7 nm, are close in value. In contrast, in the MBI-PG2.5 structure, with a smaller pore size, Dpore ~2.5 nm, a significantly smaller value of Wm and a larger value of activation energy Ea for DC conductivity are observed, which indicates the manifestation of a nanoparticle size effect.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms241813740/s1.

Author Contributions

Formal analysis, E.B., A.A.L., B.K., A.S., S.P., A.F., D.K. and D.E.; Investigation, E.B., A.A.L., A.S., S.P., A.F., D.K. and D.E.; Writing—original draft, E.B., A.A.L., A.S. and B.K.; Writing—review and editing, E.B., B.K. and A.A.L.; Conceptualization, E.B. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Acknowledgments

The XRD powder characterizations were performed using equipment and software of the Joint Research Center “Material science and characterization in advanced technology” (Ioffe Institute, St.-Petersburg, Russia).

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

MBI2-methylbenzimidazole or C8H8N2
Dporecharacteristic pore or tube diameter
dmolmaximum molecular size of MBI
Dcharacteristic size of crystallites
aMBI lattice parameter
cMBI lattice parameter
Vcellunit cell volume
PGborate porous glasses
ChAchrysotile asbestos
MSmesoporous silica
PG2.5PG with a pore diameter of 2.5 nm
PG7PG with a pore diameter of 2.5 nm and 7 nm
MBI-PG2.5PG2.5 filled by MBI from melt
MBI-PG7PG7 filled by MBI from melt
MBI-PG7gPG7 filled by MBI from gas
MBI-MSmesoporous silica filled by MBI
PLphotoluminescence
Phases in MBI PG2.5
MBI-PG2.5LMBI phase Large
MBI-PG2.5MMBI phase Mean
MBI-PG2.5SMBI phase Small
Phases in MBI-PG7
MBI-PG7LMBI phase Large
MBI-PG7M1MBI phase Mean1
MBI-PG27M2MBI phase Mean2
MBI-PG7SMBI phase Small
Phases in MBI-ChA
MBI-ChA1MBI phase 1
MBI-ChA2MBI phase 2
Ephphoton energy
f, ωfrequency, ω = 2πf
εeffeffective permittivity
tgδdielectric loss
σconductivity

References

  1. Alshammari, B.H.; Lashin, M.M.A.; Mahmood, M.A.; Al-Mubaddel, F.S.; de Nasir Ilyas, R.N.; Sohail, M.; Khan, A.; Abdullaev, S.S.; Khan, R. Organic and inorganic nanomaterials: Fabrication, properties and applications. RSC Adv. 2023, 13, 13735. [Google Scholar] [CrossRef] [PubMed]
  2. Zeijl, H.W. Thin Film Technologies for Micro/Nano Systems; A Review. ECS Trans. 2014, 61, 191–206. [Google Scholar] [CrossRef]
  3. Plawsky, J.L.; Fedorov, A.G.; Garimella, S.V.; Ma, H.B.; Maroo, S.C.; Chen, L.; Nam, Y. Nano-and Microstructures for Thin-Film Evaporation—A Review. Nanoscale Microscale Thermophys. Eng. 2014, 18, 251–269. [Google Scholar] [CrossRef]
  4. Krishnamoorthy, G.; Chidambaram, R. Nanostructured thin films and nanocoatings. In Micro and Nano Technologies, Emerging Applications of Nanoparticles and Architecture Nanostructures; Barhoum, A., Makhlouf, A.A.S., Eds.; Elsevier: Amsterdam, The Netherlands, 2018; pp. 533–552. ISBN 9780323512541. [Google Scholar] [CrossRef]
  5. Bune, A.V.; Fridkin, V.M.; Ducharme, S.; Blinov, L.M.; Palto, S.P.; Sorokin, A.V.; Yudin, S.G.; Zlatkin, A. Two-dimensional ferroelectric films. Nature 1998, 391, 874–877. [Google Scholar] [CrossRef]
  6. Agarwal, P.; Qi, H.; Archer, L.A. The Ages in a Self-Suspended Nanoparticle Liquid. Nano Lett. 2010, 10, 111–115. [Google Scholar] [CrossRef]
  7. Wells, J.; Kazakova, J.; Posth, O.; Steinhoff, U.; Petronis, S.; Bogart, L.K.; Southern, P.; Pankhurst, Q.; Johansson, C. Standardisation of magnetic nanoparticles in liquid suspension. J. Phys. D Appl. Phys. 2017, 50, 383003. [Google Scholar] [CrossRef]
  8. Pergher, S.; Rodríguez-Castellón, E. Nanoporous Materials and Their Applications. Appl. Sci. 2019, 9, 1314. [Google Scholar] [CrossRef]
  9. Sivasankaran, S. (Ed.) Nanocomposites Recent Evolutions. In Open Access Peer-Reviewed Edited Volume; IntechOpen: London, UK, 2019. [Google Scholar] [CrossRef]
  10. Sharma, A.K.; Priya; Kaith, B.S. Polymer Nanocomposite Matrices: Classification, Synthesis Methods, and Applications. In Handbook of Polymer and Ceramic Nanotechnology; Hussain, C.M., Thomas, S., Eds.; Springer: Cham, Switzerland, 2021. [Google Scholar] [CrossRef]
  11. Bréchet, Y.; Cavaillé, J.Y.Y.; Chabert, E.; Chazeau, L.; Dendievel, R.; Flandin, L.; Gauthier, C. Polymer based nanocomposites: Effect of filler-filler and filler-matrix interactions. Adv. Eng. Mater. 2001, 3, 571–577. [Google Scholar] [CrossRef]
  12. Popov, V.N. Carbon Nanotubes: Properties and Application. Mater. Sci. Eng. R Rep. 2004, 43, 61–102. [Google Scholar] [CrossRef]
  13. Lin, Y.; Taylor, S.; Li, H.; Fernando, K.A.S.; Qu, L.; Wang, W.; Gu, L.; Zhou, B.; Sun, Y.P. Advances toward Bioapplications of Carbon Nanotubes. J. Mater. Chem. 2004, 14, 527–541. [Google Scholar] [CrossRef]
  14. Vasconcelos, A.A.; Len, T.; de Nazaré de Oliveira, A.; da Costa, A.A.F.; da Costa, C.E.F.; Luque, R.; da Rocha Filho, G.N.; Noronha, R.C.R.; do Nascimento, L.A.S. Zeolites: A Theoretical and Practical Approach with Uses in (Bio) Chemical Processes. Appl. Sci. 2023, 13, 1897. [Google Scholar] [CrossRef]
  15. Rahaman, M.N.; Liang, W.; Day, D.E. Preparation and Bioactive Characteristics of Porous Borate Glass Substrates. In Advances in Bioceramics and Biocomposites: Ceramic Engineering and Science Proceedings; Mizuno, M., Ed.; The American Ceramic Society: Westerville, OH, USA, 2005. [Google Scholar] [CrossRef]
  16. Ege, D.; Zheng, K.; Boccaccini, A.R. Borate Bioactive Glasses (BBG): Bone Regeneration, Wound Healing Applications, and Future Directions. ACS Appl. Bio Mater. 2022, 5, 3608–3622. [Google Scholar] [CrossRef] [PubMed]
  17. Pye, A.M. A review of asbestos substitute materials in industrial applications. J. Hazard. Mater. 1979, 3, 125–147. [Google Scholar] [CrossRef]
  18. Porrang, S.; Davaran, S.; Rahemi, N.; Allahyari, S.; Mostafavi, E. How Advancing are Mesoporous Silica Nanoparticles? A Comprehensive Review of the Literature. Int. J. Nanomed. 2022, 17, 1803–1827. [Google Scholar] [CrossRef]
  19. Niculescu, V.-C. Mesoporous Silica Nanoparticles for Bio-Applications. Front. Mater. 2020, 7, 36. [Google Scholar] [CrossRef]
  20. Atwood, J.L.; Steed, J.W. (Eds.) Organic Nanostructures; John Wiley & Sons: Hoboken, NJ, USA, 2008; 370p, ISBN 978-3-527-31836-0. [Google Scholar]
  21. O’Carrol, D.M. Organic photonic nanostructures. In Handbook of Organic Materials for Electronic and Photonic Devices, 2nd ed.; Woodhead Publishing Series in Electronic and Optical Materials; Woodhead Publishing: Cambridge, UK, 2019; pp. 111–138. [Google Scholar]
  22. Al-Shamery, K.; Rubahn, H.-G.; Sitter, H. Organic Nanostructures for Next Generation Devices, 2nd ed.; Woodhead Publishing Series in Electronic and Optical Materials; Woodhead Publishing: Cambridge, UK, 2019; pp. 111–138. [Google Scholar]
  23. Lombardo, D.; Pasqua, L.; Magazù, S. Self-Assembly of Organic Nanomaterials and Biomaterials: The Bottom-Up Approach for Functional Nanostructures Formation and Advanced Applcations. Materials 2020, 13, 1048. [Google Scholar] [CrossRef]
  24. Balashova, E.; Levin, A.A.; Davydov, V.; Smirnov, A.; Starukhin, A.; Pavlov, S.; Krichevtsov, B.; Zolotarev, A.; Zhang, H.; Li, F.; et al. Croconic Acid Doped Glycine Single Crystals: Growth, Crystal Structure, UV-Vis, FTIR, Raman and Photoluminescence Spectroscopy. Crystals 2022, 12, 1342. [Google Scholar] [CrossRef]
  25. Kermani, F.; Nazarnezhad, S.; Mollaei, Z.; Mollazadeh, S.; Ebrahimzadeh-Bideskan, A.; Askari, V.R.; Oskuee, R.K.; Moradi, A.; Hosseini, S.A.; Azari, Z.; et al. Zinc-and Copper-Doped Mesoporous Borate Bioactive Glasses: Promising Additives for Potential Use in Skin Wound Healing Applications. Int. J. Mol. Sci. 2023, 24, 1304. [Google Scholar] [CrossRef]
  26. Wang, H.; Cheng, Z.; Zhang, P.; Ye, J.; Ding, L.; Jia, W. Ultra-high adsorption behavior of zeolitic imidazole framework-67 nanoparticles for removing brilliant green dye. AIP Adv. 2021, 11, 095304. [Google Scholar] [CrossRef]
  27. Kumzerov, Y.; Vakhrushev, S. Nanostructures within porous materials. In Encyclopedia of Nanoscience and Nanotechnology; Nalwa, H.S., Ed.; American Scientific Publishers: Los Angeles, CA, USA, 2004; Volume 3, pp. 811–849. [Google Scholar]
  28. Obodovskaya, A.E.; Starikova, Z.A.; Belous, S.N.; Pokrovskaya, I.E. Crystal and molecular structure of 2-methylbenzimidazole. J. Struct. Chem. 1991, 32, 421–422. [Google Scholar] [CrossRef]
  29. Horiuchi, S.; Kagawa, F.; Hatahara, K.; Kobayashi, K.; Kumai, R.; Murakami, Y.; Tokura, Y. Above-room-temperature ferroelectricity and antiferroelectricity in benzimidazoles. Nat. Commun. 2012, 3, 1308. [Google Scholar] [CrossRef]
  30. Balashova, E.V.; Krichevtsov, B.B.; Kunkel, T.S.; Ankudinov, A.V. AFM Visualization of Ferroelastic and Ferroelectric Domains in 2-Methylbenzimidazole C8H8N2 Crystals. J. Surf. Investig. X-ray Synchrotron Neutron Tech. 2021, 15, 1165–1167. [Google Scholar] [CrossRef]
  31. Balashova, E.V.; Svinarev, F.B.; Ankudinov, A.V.; Pankova, G.A.; Lityagin, G.A.; Kunkel, T.S.; Krichevtsov, B.B. Polarization switching, dielectric, structural and elastic properties of 2-Methylbenzimidazole crystals and films. Ferroelectrics. 2019, 538, 74–82. [Google Scholar] [CrossRef]
  32. Svinarev, F.B.; Balashova, E.V.; Krichevtsov, B.B. Dielectric properties of self-assembled spherulite films of organic ferroelectric 2-methylbenzimidazole. Ferroelectrics 2019, 543, 167–174. [Google Scholar] [CrossRef]
  33. Yuan, Y.; Ni, Y.; Jiang, X.; Yun, Y.; Li, J.; Xu, X. Highly Oriented Organic Ferroelectric Films with Single-Crystal-Level Properties from Restrained Crystallization. Cryst. Growth Des. 2022, 22, 2124–2131. [Google Scholar] [CrossRef]
  34. Noda, Y.; Yamada, T.; Kobayashi, K.; Kumai, R.; Horiuchi, S.; Kagawa, F.; Hasegawa, T. Few-Volt Operation of Printed Organic Ferroelectric Capacitor. Adv. Mater. 2015, 27, 6475. [Google Scholar] [CrossRef]
  35. Kinoshita, Y.; Sotome, M.; Miyamoto, T.; Uemura, Y.; Arai, S.; Horiuchi, S.; Hasegawa, T.; Okamoto, H.; Kida, N. Observation of the Three-Dimensional Polarization Vector in Films of Organic Molecular Ferroelectrics Using Terahertz Radiation Emission. Phys. Rev. Appl. 2020, 14, 054002. [Google Scholar] [CrossRef]
  36. Langford, J.I.; Cernik, R.J.; Louer, D. The Breadth and Shape of Instrumental Line Profiles in High-Resolution Powder Diffraction. J. Appl. Phys. 1991, 24, 913–919. [Google Scholar] [CrossRef]
  37. Levin, A.A. Program SizeCr for Calculation of the Microstructure Parameters from X-ray Diffraction Data. Preprint. 2022. Available online: https://www.iucr.org/resources/other-directories/software/sizecr (accessed on 1 October 2022).
  38. Hahn, T. International Tables for Crystallography. Volume A. Space Group Symmetry; D. Reidel Publishing Company: Dordrecht, The Netherlands; Boston, MA, USA, 1983; pp. 344–345. [Google Scholar]
  39. Kraus, W.; Nolze, G. POWDER CELL—A program for the representation and manipulation of crystal structures and calculation of the resulting X-ray powder patterns. J. Appl. Crystallogr. 1996, 29, 301–303. [Google Scholar] [CrossRef]
  40. Fokin, A.V.; Kumzerov, Y.A.; Okuneva, N.M.; Naberezhnov, A.A.; Vakhrushev, S.B.; Golosovsky, I.V.; Kurbakov, A.I. Temperature Evolution of Sodium Nitrite Structure in a Restricted Geometry. Phys. Rev. Lett. 2002, 89, 175503. [Google Scholar] [CrossRef]
  41. Golosovsky, I.V.; Delaplane, R.G.; Naberezhnov, A.A.; Kumzerov, Y.A. Thermal motions in lead confined within porous glass. Phys. Rev. B 2004, 69, 132301. [Google Scholar] [CrossRef]
  42. Golosovsky, I.V.; Smirnov, O.P.; Delaplane, R.G.; Wannberg, A.; Kibalin, Y.A.; Naberezhnov, A.A.; Vakhrushev, S.B. Atomic motion in Se nanoparticles embedded into a porous glass matrix. Eur. Phys. J. 2006, B54, 211–216. [Google Scholar] [CrossRef]
  43. Le Bail, A.; Duroy, H.; Fourquet, J.L. Ab-initio structure determination of LiSbWO6 by X-ray powder diffraction. Mater. Res. Bull. 1988, 23, 447–452. [Google Scholar] [CrossRef]
  44. Paufler, P.; Filatov, S.K.; Shakhverdova, I.P.; Bubnova, R.S.; Reibold, M.; Müller, B.; Levin, A.A.; Meyer, D.C. Mechanical properties and structure of a nanoporous sodium borosilicate glass. Glass Phys. Chem. 2007, 33, 187–198. [Google Scholar] [CrossRef]
  45. Spectral Database for Organic Compounds, SDBS. National Institute of Advanced Industrial Science and Technology (AIST) Japan. Available online: https://sdbs.db.aist.go.jp (accessed on 1 October 2022).
  46. Güllüoglu, M.T.; Özduran, M.; Kurt, M.; Kalaichelvan, S.; Sundaraganesan, N. Molecular structure and vibrational spectra of 2- and 5-methylbenzimidazole molecules by density functional theory. Spectrochim. Acta Part A Mol. Biomol. Spectrosc. 2010, 76, 107–114. [Google Scholar] [CrossRef] [PubMed]
  47. Anbalagan, G.; SakthiMurugesan, K.; Balakrishnan, M.; Gunasekaran, S. Structural analysis, optical absorption and EPR spectroscopic studies on chrysotile. Appl. Clay Sci. 2008, 42, 175–179. [Google Scholar] [CrossRef]
  48. Frost, R.L.; Scholz, R.; Belotti, F.M.; López, A.; Theiss, F.L. A vibrational spectroscopic study of the phosphate mineral vantasselite Al4(PO4)3(OH)39H2O. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2015, 147, 185–192. [Google Scholar] [CrossRef] [PubMed]
  49. Balashova, E.V.; Svinarev, F.B.; Zolotarev, A.A.; Levin, A.A.; Brunkov, P.N.; Davydov, V.Y.; Smirnov, A.N.; Redkov, A.V.; Pankova, G.A.; Krichevtsov, B.B. Crystal structure, Raman spectroscopy and dielectric properties of new semi-organic crystals based on 2-methylbenzimidazole. Crystals 2019, 9, 573. [Google Scholar] [CrossRef]
  50. Bartoš, J.; Arrese-Igor, S.; Švajdlenková, H.; Kleinová, A.; Alegría, A. Dynamics of Confined Short-Chain alkanol in MCM-41 by Dielectric Spectroscopy: Effects of matrix and system Treatments and Filling Factor. Polymers 2020, 12, 610. [Google Scholar] [CrossRef]
  51. Chen, D.; Wang, H.J. HOMO-LUMO Gaps of Homogeneous Polycyclic Aromatic Hydrocarbon Clusters. Phys. Chem. C 2019, 123, 27785–27793. [Google Scholar] [CrossRef]
  52. Gutina, A.; Axelrod, E.; Puzenko, A.; Rysiakiewicz-Pasek, E.; Kozlovich, N.; Feldman, Y. Dielectric relaxation of porous glasses. J. Non-Cryst. Solids 1998, 235–237, 302–307. [Google Scholar] [CrossRef]
  53. Datta, A.K.; Bhattacherjee, S. An electrical study of chrysotile asbestos. J. Mater. Sci. 1986, 21, 1041–1045. [Google Scholar] [CrossRef]
  54. Sherman, V.O.; Tagantsev, A.K.; Setter, N. Ferroelectric-dielectric tunable composites. J. Appl. Phys. 2006, 99, 074104. [Google Scholar] [CrossRef]
  55. Bruggeman, D.A.G. Dielectric constant and conductivity of mixtures of isotropic materials. Ann. Phys. 1935, 24, 636–679. [Google Scholar] [CrossRef]
  56. Jonscher, A.K. The ‘universal’ dielectric response. Nature 1977, 267, 673–679. [Google Scholar] [CrossRef]
  57. Jonscher, A.K. A new understanding of the dielectric relaxation of solids. J. Mater. Sci. 1981, 16, 2037–2060. [Google Scholar] [CrossRef]
  58. Pike, G.E. ac Conductivity of Scandium Oxide and a New Hopping Model for Conductivity. Phys. Rev. B 1972, 6, 1572. [Google Scholar] [CrossRef]
  59. Hutton, S.L.; Fehst, I.; Böhmer, R.; Braune, M.; Mertz, B.; Lunkenheimer, P.; Loidl, A. Proton glass behavior and hopping conductivity in solid solutions of antiferroelectric betaine phosphate and ferroelectric betaine phosphite. Phys. Rev. Lett. 1991, 66, 1990. [Google Scholar] [CrossRef]
  60. Totz, J.; Michel, D.; Banys, J.; Klopperpieper, A. Conductivity processes in deuterated betaine phosphate1−xbetaine phosphitex mixed crystals. J. Phys. Condens. Matter 1998, 10, 9281–9292. [Google Scholar] [CrossRef]
  61. Balashova, E.; Levin, A.A.; Fokin, A.; Redkov, A.; Krichevtsov, B. Structural Properties and Dielectric Hysteresis of Molecular Organic Ferroelectric Grown from Different Solvents. Crystals 2021, 11, 1278. [Google Scholar] [CrossRef]
  62. Morimoto, S. Porous Glass: Preparation and Properties. Key Engineering Materials; Trans Tech Publications, Ltd.: Stafa-Zurich, Switzerland, 1995; Volume 115, pp. 147–158. [Google Scholar] [CrossRef]
  63. Zhu, B.; Zhang, Z.; Zhang, W.; Wu, Y.; Zhang, J.; Imran, Z.; Zhang, D. Synthesis and Applications of Porous Glass. J. Shanghai Jiaotong Univ. Sci. 2019, 24, 681–698. [Google Scholar] [CrossRef]
  64. Inayat, A.; Reinhardt, B.; Jan Herwig, J.; Küster, C.; Uhlig, H.; Krenkel, S.; Edda Raedlein, E.; Enke, D. Recent advances in the synthesis of hierarchically porous silica materials on the basis of porous glasses. New J. Chem. 2016, 40, 4095–4114. [Google Scholar] [CrossRef]
  65. Trofimova, E.Y.; Kurdyukov, D.A.; Yakovlev, S.A.; Kirilenko, D.A.; Kukushkina, Y.A.; Nashchekin, A.V.; Sitnikova, A.A.; Yagovkina, M.A.; Golubev, V.G. Monodisperse spherical mesoporous silica particles: Fast synthesis procedure and fabrication of photonic-crystal films. Nanotechnology 2013, 24, 155601. [Google Scholar] [CrossRef]
  66. Diffrac. Suite Eva, Version 5.1.0.5; Bruker AXS: Karlsruhe, Germany, 2019.
  67. International Centre for Diffraction Data (ICDD). Powder Diffraction File-2 Release 2014; ICDD: Newton Square, PA, USA, 2014. [Google Scholar]
  68. Bruker AXS. Technical reference. In TOPAS Version 5; Bruker AXS: Karlsruhe, Germany, 2014. [Google Scholar]
  69. Rehani, B.R.; Joshi, P.B.; Lad, K.N.; Pratap, A. Crystallite size estimation of elemental and composite silver nano-powders using XRD principles. Indian J. Pure Appl. Phys. 2006, 44, 157–161. [Google Scholar]
  70. Stokes, A.R.; Wilson, A.J.C. The diffraction of X-rays by distorted crystal aggregates. Proc. Phys. Soc. Lond. 1944, 56, 174–181. [Google Scholar] [CrossRef]
  71. Scherrer, P. Bestimmung der Grösse und der inneren Struktur von Kolloidteilchen mittels Röntgenstrahlen. Nachr. Königl. Ges. Wiss. Göttingen. 1918, 26, 98–100. [Google Scholar]
  72. Maunders, C.; Etheridge, J.; Wright, N.; Whitfield, H.J. Structure and microstructure of hexagonal Ba3Ti2RuO9 by electron diffraction and microscopy. Acta Cryst. B 2005, 61, 154–159. [Google Scholar] [CrossRef]
  73. Rietveld, H.M. Line profiles of neutron powder-diffraction peaks for structure Refinement. Acta Crystallogr. 1967, 22, 151–152. [Google Scholar] [CrossRef]
  74. Cambridge Crystallographic Data Center (CCDC). Available online: https://www.ccdc.cam.ac.uk/about-us/ (accessed on 8 October 2022).
  75. Rietveld, H.M. The Rietveld Method: A retrospection. Z. Kristallogr. 2010, 225, 545–547. [Google Scholar] [CrossRef]
  76. Berger, H. Study of the K alpha emission spectrum of copper. X-ray Spectrom. 1986, 15, 241–243. [Google Scholar] [CrossRef]
  77. Hall Jnr, M.M.; Veeraraghavan, V.G.; Rubin, H.; Winchell, P.G. The approximation of symmetric X-ray peaks by Pearson type VII distributions. J. Appl. Crystallogr. 1977, 10, 66–68. [Google Scholar] [CrossRef]
  78. Cheary, R.W.; Coelho, A.A. A fundamental parameters approach to X-ray line-profile fitting. J. Appl. Crystallogr. 1992, 25, 109–121. [Google Scholar] [CrossRef]
  79. Balzar, D. Voigt-function model in diffraction line-broadening analysis. In Defect and Microstructure Analysis by Diffraction; Snyder, R.L., Fiala, J., Bunge, H.J., Eds.; IUCr, Oxford University Press: Oxford, UK, 1999; pp. 94–126. [Google Scholar]
  80. Dollase, W.A. Correction of Intensities of Preferred Orientation in Powder Diffractometry: Application of the March Model. J. Appl. Crystallogr. 1986, 19, 267–272. [Google Scholar] [CrossRef]
  81. Jarvinen, M. Application of symmetrized harmonics expansion to correction of the preferred orientation effect. J. Appl. Crystallogr. 1993, 26, 525–531. [Google Scholar] [CrossRef]
  82. Hill, R.J.; Fischer, R.X. Profile Agreement Indices in Rietveld and Pattern-Fitting Analysis. J. Appl. Crystallogr. 1990, 23, 462–468. [Google Scholar] [CrossRef]
  83. Levin, A.A. Program RietESD for Correction of Estimated Standard Deviations Obtained in Rietveld-Refinement Program. Preprint. 2022. Available online: https://www.iucr.org/resources/other-directories/software/rietesd (accessed on 1 October 2022).
  84. Bérar, J.-F.; Lelann, P. ESD’s and Estimated Probable Error Obtained in Rietveld Refinements with Local Correlations. J. Appl. Crystallogr. 1991, 24, 1–5. [Google Scholar] [CrossRef]
  85. Moberly, J.G.; Bernards, M.T.; Waynant, K.V. Key features and updates for Origin. J. Cheminform. 2018, 10, 5. [Google Scholar] [CrossRef]
  86. Lipson, H.; Langford, J.I.; Hu, H.-C. Trigonometric intensity factors. In International Tables for Crystallography, Vol. C, Mathematical, Physical and Chemical Tables, 3rd ed.; Prince, E., Ed.; International Union of Crystallography, Kluwer Academic Publishers: Dordrecht, The Netherlands; Boston, MA, USA, 2004; pp. 596–598. [Google Scholar]
  87. Maslen, E.N. X-ray absorption. In International Tables for Crystallography, Vol. C, Mathematical, Physical and Chemical Tables, 3rd ed.; Prince, E., Ed.; International Union of Crystallography, Kluwer Academic Publishers: Dordrecht, The Netherlands; Boston, MA, USA, 2004; pp. 599–608. [Google Scholar]
  88. Brown, A.G.; Fox, A.G.; Maslen, E.N.; O’Keefe, M.A.; Willis, B.T.M. Intensity of diffracted intensities. In International Tables for Crystallography, Vol. C, Mathematical, Physical and Chemical Tables, 3rd ed.; Prince, E., Ed.; International Union of Crystallography, Kluwer Academic Publishers: Dordrecht, The Netherlands; Boston, MA, USA, 2004; pp. 554–595. [Google Scholar]
Figure 1. Parts of the XRD patterns of the borate glasses filled by MBI, MBI-PG2.5, MBI-PG7, and MBI-PG30. The Miller indices hkl of MBI XRD reflections which make the greatest contribution to the observed reflections are shown. The inset illustrates the difference in the broadening of the observed XRD with hkl = 221 and 002.
Figure 1. Parts of the XRD patterns of the borate glasses filled by MBI, MBI-PG2.5, MBI-PG7, and MBI-PG30. The Miller indices hkl of MBI XRD reflections which make the greatest contribution to the observed reflections are shown. The inset illustrates the difference in the broadening of the observed XRD with hkl = 221 and 002.
Ijms 24 13740 g001
Figure 2. MBI-PG2.5 sample. angle distribution of crystallite sizes, D0hkl, of MBI-PG2.5L and MBI-PG2.5M phases calculated in the model of zero microstrain (εs = 0) for observed individual reflections of the phases.
Figure 2. MBI-PG2.5 sample. angle distribution of crystallite sizes, D0hkl, of MBI-PG2.5L and MBI-PG2.5M phases calculated in the model of zero microstrain (εs = 0) for observed individual reflections of the phases.
Ijms 24 13740 g002
Figure 3. (a) WHP and (b) SSP graphs, constructed for the reflections of MBI-PG2.5L and MBI-PG2.5M crystalline phases of the MBI-PG2.5 sample. In (a,b), FWHM and FWHMcorr are FWHM values without and with correction to instrumental broadening, respectively; Kstrain = 4 and KScherrer = 0.94 (see Section 3.2.2); θB is half of Bragg angle 2θB corrected to zero shift and displacement (Section 3.2.2); dhkl is interplane distance corresponding to XRD reflection with Miller indices hkl and Bragg angle 2θB; λ = 0.15406 nm is wavelength of Cu-Kα1 radiation used (after correction to Cu-Kα2 contribution). The straight approximating lines Y = A + BX of the WHP and SSP are shown in (a,b). Expressions for X and Y are given in axes captions of WHP and SSP graphs.
Figure 3. (a) WHP and (b) SSP graphs, constructed for the reflections of MBI-PG2.5L and MBI-PG2.5M crystalline phases of the MBI-PG2.5 sample. In (a,b), FWHM and FWHMcorr are FWHM values without and with correction to instrumental broadening, respectively; Kstrain = 4 and KScherrer = 0.94 (see Section 3.2.2); θB is half of Bragg angle 2θB corrected to zero shift and displacement (Section 3.2.2); dhkl is interplane distance corresponding to XRD reflection with Miller indices hkl and Bragg angle 2θB; λ = 0.15406 nm is wavelength of Cu-Kα1 radiation used (after correction to Cu-Kα2 contribution). The straight approximating lines Y = A + BX of the WHP and SSP are shown in (a,b). Expressions for X and Y are given in axes captions of WHP and SSP graphs.
Ijms 24 13740 g003
Figure 4. Sample MBI-PG2.5. The results of the Rietveld fit within the three-phase model MBI-PG2.5L + MBI-PG2.5M + MBI-PG2.5S. The MBI-G2.5L, MBI-PG2.5M, and MBI-PG2.5S phases are designated as L, M, and S, respectively. The positions of the phase reflections calculated from the unit cell parameters refined by the Rietveld method are shown as vertical bars below. In addition to the experimental (Iexp), calculated (Icalc), and difference (IexpIcalc) diagrams, the calculated XRD patterns (together with the background contribution) corresponding to individual crystalline phases are shown in the region 2θ = 6–37°. The Miller indices hkl of some selected observed reflections are indicated. The inset shows, on an enlarged scale, part of the diagrams in the region of reflections 221 and 002 of the MBI phases.
Figure 4. Sample MBI-PG2.5. The results of the Rietveld fit within the three-phase model MBI-PG2.5L + MBI-PG2.5M + MBI-PG2.5S. The MBI-G2.5L, MBI-PG2.5M, and MBI-PG2.5S phases are designated as L, M, and S, respectively. The positions of the phase reflections calculated from the unit cell parameters refined by the Rietveld method are shown as vertical bars below. In addition to the experimental (Iexp), calculated (Icalc), and difference (IexpIcalc) diagrams, the calculated XRD patterns (together with the background contribution) corresponding to individual crystalline phases are shown in the region 2θ = 6–37°. The Miller indices hkl of some selected observed reflections are indicated. The inset shows, on an enlarged scale, part of the diagrams in the region of reflections 221 and 002 of the MBI phases.
Ijms 24 13740 g004
Figure 5. Sample MBI-PG7. Results of the Rietveld fit within the four-phase MBI-PG7L + MBI-PG7M1 + MBI-PG7M2 + MBI-PG7S model. The MBI-PG7L, MBI-PG7M1, MBI-PG7M2, and MBI-PG7S phases are designated as L, M1, M2, and S, respectively. Other details are the same as in caption for Figure 4.
Figure 5. Sample MBI-PG7. Results of the Rietveld fit within the four-phase MBI-PG7L + MBI-PG7M1 + MBI-PG7M2 + MBI-PG7S model. The MBI-PG7L, MBI-PG7M1, MBI-PG7M2, and MBI-PG7S phases are designated as L, M1, M2, and S, respectively. Other details are the same as in caption for Figure 4.
Ijms 24 13740 g005
Figure 6. Sample MBI-ChA. Results of the LB fit in frames of the two-phase MBI-ChA1 + MBI-ChA2 model. The MBI-ChA1 and MBI-ChA2 phases are designated as M1 and M2, respectively. Other details are the same as in the caption to Figure 4. The positions of the phase reflections calculated from the unit cell parameters refined by the LB method are shown as vertical bars below. The experimental, calculated, and difference diagrams are marked as Iexp, Icalc, and IexpIcalc, respectively. The Miller indices, hkl, of some selected observed reflections are indicated. The inset shows the experimental and calculated XRD patterns and calculated diagrams (together with the background contribution) corresponding to individual crystalline phases in the region 2θ = 17–39°.
Figure 6. Sample MBI-ChA. Results of the LB fit in frames of the two-phase MBI-ChA1 + MBI-ChA2 model. The MBI-ChA1 and MBI-ChA2 phases are designated as M1 and M2, respectively. Other details are the same as in the caption to Figure 4. The positions of the phase reflections calculated from the unit cell parameters refined by the LB method are shown as vertical bars below. The experimental, calculated, and difference diagrams are marked as Iexp, Icalc, and IexpIcalc, respectively. The Miller indices, hkl, of some selected observed reflections are indicated. The inset shows the experimental and calculated XRD patterns and calculated diagrams (together with the background contribution) corresponding to individual crystalline phases in the region 2θ = 17–39°.
Ijms 24 13740 g006
Figure 7. Total correlation functions c(r) built from the measured XRD patterns of MBI-MS and pure MS. For comparison, the c(r) function for mesoporous silica MS5 glass with composition 0.2Na2O·3.8B2O3·96SiO2 is shown. Designations of the figure are explained in text. Numbers indicate the correlation maxima and correspond to mean distances in silica structure. Symbol f indicates the false peak arising due to the breakage of the Fourier series during calculations of c(r).
Figure 7. Total correlation functions c(r) built from the measured XRD patterns of MBI-MS and pure MS. For comparison, the c(r) function for mesoporous silica MS5 glass with composition 0.2Na2O·3.8B2O3·96SiO2 is shown. Designations of the figure are explained in text. Numbers indicate the correlation maxima and correspond to mean distances in silica structure. Symbol f indicates the false peak arising due to the breakage of the Fourier series during calculations of c(r).
Ijms 24 13740 g007
Figure 8. (a) IR absorption spectra of samples of “pure” ChA and MBI-ChA. (b) Enlarged image of the IR absorption spectra of the samples in the wavenumber region ν = 1150–3300 cm−1. Spectra are shifted vertically for clarity.
Figure 8. (a) IR absorption spectra of samples of “pure” ChA and MBI-ChA. (b) Enlarged image of the IR absorption spectra of the samples in the wavenumber region ν = 1150–3300 cm−1. Spectra are shifted vertically for clarity.
Ijms 24 13740 g008
Figure 9. (a) The emission spectrum of a bulk MBI crystal (blue line) upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 293 K. The insets show the decomposition of the MBI emission contour into Gaussian bands (olive lines) and fragments of the IR absorption spectra of MBI (red lines). (b) A scheme of vibronic transitions in an emission center. (c) Luminescence spectra of imidazole and MBI crystals upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 80 K.
Figure 9. (a) The emission spectrum of a bulk MBI crystal (blue line) upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 293 K. The insets show the decomposition of the MBI emission contour into Gaussian bands (olive lines) and fragments of the IR absorption spectra of MBI (red lines). (b) A scheme of vibronic transitions in an emission center. (c) Luminescence spectra of imidazole and MBI crystals upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 80 K.
Ijms 24 13740 g009
Figure 10. (a) Emission spectra of MBI nanoparticles in MBI-CA (red line) and MBI bulk crystal (blue line) upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 300 K. (b) Emission spectra of MBI nanoparticles in porous borate glasses with pores of 2.5 nm (1) and 7 nm (2) and MBI bulk crystal (3) upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 300 K. (c) PL spectra of MBI in mesoporous silica (1) and in ethanol solution (2) under 405 nm light excitation. Measurement temperature T = 300 K.
Figure 10. (a) Emission spectra of MBI nanoparticles in MBI-CA (red line) and MBI bulk crystal (blue line) upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 300 K. (b) Emission spectra of MBI nanoparticles in porous borate glasses with pores of 2.5 nm (1) and 7 nm (2) and MBI bulk crystal (3) upon excitation by light with a wavelength of 405 nm (Ephoton = 3.061 eV). Measurement temperature T = 300 K. (c) PL spectra of MBI in mesoporous silica (1) and in ethanol solution (2) under 405 nm light excitation. Measurement temperature T = 300 K.
Ijms 24 13740 g010aIjms 24 13740 g010b
Figure 11. Frequency dependencies of effective permittivity. (a) εeff, (b) εeff, and (c) conductivity σ in MBI-ChA and ChA.
Figure 11. Frequency dependencies of effective permittivity. (a) εeff, (b) εeff, and (c) conductivity σ in MBI-ChA and ChA.
Ijms 24 13740 g011
Figure 12. Frequency dependencies of the real and imaginary parts of effective permittivity. (a) εeff, (b) εeff, and (c) conductivity σ in PG7, MBI-PG7 samples.
Figure 12. Frequency dependencies of the real and imaginary parts of effective permittivity. (a) εeff, (b) εeff, and (c) conductivity σ in PG7, MBI-PG7 samples.
Ijms 24 13740 g012
Figure 13. Frequency dependencies of the (a) real part of effective permittivity, εeff, in MBI-PG7 (blue) and MBI-PG2.5 (purple), and (b) conductivity σ in MBI-PG7 (blue), MBI-PG2.5 (purple), and MBI-PG30 (green) samples. Inset in (a) shows frequency dependence of permittivity εeff in MBI crystals along [110] direction. Red and green lines in (a) show results of calculations.
Figure 13. Frequency dependencies of the (a) real part of effective permittivity, εeff, in MBI-PG7 (blue) and MBI-PG2.5 (purple), and (b) conductivity σ in MBI-PG7 (blue), MBI-PG2.5 (purple), and MBI-PG30 (green) samples. Inset in (a) shows frequency dependence of permittivity εeff in MBI crystals along [110] direction. Red and green lines in (a) show results of calculations.
Ijms 24 13740 g013
Figure 14. Frequency dependencies of the conductivity, σ, in MBI-PG2.5 at different temperatures (a) and temperature dependencies of conductivity, σ, in MBI-PG7 and MBI-PG2.5 samples for f = 60 Hz (b). Inset in (b) shows dependence of conductivity, σ, on inverse temperature in MBI-PG7 and MBI-PG2.5 samples.
Figure 14. Frequency dependencies of the conductivity, σ, in MBI-PG2.5 at different temperatures (a) and temperature dependencies of conductivity, σ, in MBI-PG7 and MBI-PG2.5 samples for f = 60 Hz (b). Inset in (b) shows dependence of conductivity, σ, on inverse temperature in MBI-PG7 and MBI-PG2.5 samples.
Ijms 24 13740 g014
Figure 15. Images of MBI crystals prepared by different methods. (a) Split crystals of the spherulite type, consisting of elongated crystallites growing radially in an ethanol solution along pseudotetragonal axes [001]tetr from one crystallization center. (b) Dendrite MBI crystals prepared by sublimation of gas phase on glass surface. (c) Image of textured MBI film grown by evaporation of MBI ethanol solution on NdGaO3 substrate in crossed polarizers. (d) MBI crystals grown by vacuum sublimation from gas phase. (e) MBI crystal from acetone solution.
Figure 15. Images of MBI crystals prepared by different methods. (a) Split crystals of the spherulite type, consisting of elongated crystallites growing radially in an ethanol solution along pseudotetragonal axes [001]tetr from one crystallization center. (b) Dendrite MBI crystals prepared by sublimation of gas phase on glass surface. (c) Image of textured MBI film grown by evaporation of MBI ethanol solution on NdGaO3 substrate in crossed polarizers. (d) MBI crystals grown by vacuum sublimation from gas phase. (e) MBI crystal from acetone solution.
Ijms 24 13740 g015aIjms 24 13740 g015b
Figure 16. Photos of porous glasses and asbestos filled by MBI. (a) Porous glass with Dpore ~7 nm pores size without MBI. Porous glass with diameter of pores (b) Dpore ~30 nm, (c) ~7 nm, and (d) ~2.5 nm filled by MBI from melt. (e) Images of chrysotile asbestos matrix (tube diameter of ChA, Dpore ~9 nm) filled by MBI from melt. Split asbestos fibers were used for IR absorption (FTIR) measurements. Size of samples ~(8–12) × (6–10) × 1 mm3. (f) Schematic representation of the location of nanopores in a silica microparticle.
Figure 16. Photos of porous glasses and asbestos filled by MBI. (a) Porous glass with Dpore ~7 nm pores size without MBI. Porous glass with diameter of pores (b) Dpore ~30 nm, (c) ~7 nm, and (d) ~2.5 nm filled by MBI from melt. (e) Images of chrysotile asbestos matrix (tube diameter of ChA, Dpore ~9 nm) filled by MBI from melt. Split asbestos fibers were used for IR absorption (FTIR) measurements. Size of samples ~(8–12) × (6–10) × 1 mm3. (f) Schematic representation of the location of nanopores in a silica microparticle.
Ijms 24 13740 g016
Table 1. Results of Rietveld quantitative analysis of the sample MBI-PG2.5 obtained by means of Rietveld program TOPAS (unit cell parameters a and c and volume Vcell, size DTOPAS of nanocrystallites and microstrain εsTOPAS, weight content of the crystalline phases Wt, overall isotropic temperature factor of atoms Bisooverall, March-Dollase parameter rMD of the preferential orientation), and fitting quality characteristics (agreement Bragg factor RB, weighted profile factor Rwp and profile factor Rp, weighted profile factor cRwp and profile factor cRp corrected to background contribution). Additionally, microstructure parameters DSSP and εsSSP estimated using SSP method and factors me.s.d. for correction of e.s.d.s, obtained in Rietveld refinement, are shown.
Table 1. Results of Rietveld quantitative analysis of the sample MBI-PG2.5 obtained by means of Rietveld program TOPAS (unit cell parameters a and c and volume Vcell, size DTOPAS of nanocrystallites and microstrain εsTOPAS, weight content of the crystalline phases Wt, overall isotropic temperature factor of atoms Bisooverall, March-Dollase parameter rMD of the preferential orientation), and fitting quality characteristics (agreement Bragg factor RB, weighted profile factor Rwp and profile factor Rp, weighted profile factor cRwp and profile factor cRp corrected to background contribution). Additionally, microstructure parameters DSSP and εsSSP estimated using SSP method and factors me.s.d. for correction of e.s.d.s, obtained in Rietveld refinement, are shown.
Crystalline
Phase a
a, Å
c, Å
Vcell, Å3DTOPAS, nm
εsTOPAS, %
DSSP, nm
εsSSP, %
Wt, wt.%
Bisooverall, Å−2
RB, %
rMD b
Rwp, %
Rp, %
cRwp, %
cRp, %
MBI-PG2.5L + MBI-PG2.5M + MBI-PG2.5S, me.s.d. = 7.529 c
MBI-PG2.5L13.979(8)1220.3 (2.1)107(1)119(20) d0.80(2)2.681.36
1.00
7.529
18.86
19.86
7.211(11)0.082(1)0.16(4) d0.5(5)0.34(1)
MBI-PG2.5M13.873(60)1209.4 (7.5)12(1)15(5) d25.5(7)0.39
7.256(9)00 d4.3(4)0.53(1)
MBI-PG72.5S14.060(45)1240.0 (14.0)2.4(2)73.7(7)1.07
7.243(75)0BisooverallMBI-M5.8(6)
a space group of the MBI phases is P42/n (86) (Choice 2 according to [38]). Table unit cell parameters according to Cambridge Crystallographic Data Center (CCDC) databank, a = 13.950(9) Å, c = 7.192(3) Å, and unit cell volume Vcell = 1212.1(1.2) Å3 (CCDC code 1199885). b consistent with [110], [001], and [211] for MBI-PG2.5L, MBI-PG2.5M, and MBI-PG2.5S, respectively. c all e.s.d.s of the parameters shown in the table and obtained by refinement during Rietveld fitting are corrected on underestimation due to serial correlations by multiplication on the coefficient me.s.d. d WHP calculations result in close values, DWHP = 100(36) nm and εsWHP = 0.13(6)% for MBI-PG2.5L, DWHP = 15(5) nm, and εsWHP = 0 for MBI-PG2.5M phase.
Table 2. Results of Rietveld quantitative analysis of the sample MBI-PG7 obtained by means of Rietveld program TOPAS (see head of Table 1 for description of parameters shown).
Table 2. Results of Rietveld quantitative analysis of the sample MBI-PG7 obtained by means of Rietveld program TOPAS (see head of Table 1 for description of parameters shown).
Crystalline
Phase a
a, Å
c, Å
Vcell, Å3DTOPAS, nm
εsTOPAS, %
DSSP, nm
εsSSP, %
Wt, wt.%
Bisooverall, Å−2
RB, %
rMD b
Rwp, %
Rp, %
cRwp, %
cRp, %
MBI-PG7L + MBI-PG7M1 + MBI-PG7M2 + MBI-PG7S, me.s.d. = 5.832 c
MBI-PG7L13.943(2)1224.3 (6)61(3)57(5) d0.80(2)0.471.63
1.27
13.02
16.15
7.272(3)0.014(6)0.06(12) d1.0(4)1.24(3)
MBI-PG7M114.025(11)1230.1 (1.4)28(1)31(3) d5.70(9)0.40
7.221(2)00 d4.3(6)0.62(1)
MBI-PG7M214.014(17)1222.9 (2.9)17(1)18(4) d9.71(17)1.26
7.190(12)00 d3.9(5)0.21(1)
MBI-PG7S13.910(58)1202.5 (9.1)2.7(1)71.33(14)0.27
7.176(34)0BisooverallMBI-M20.10(1)
a see corresponding note to Table 1. b corresponds to [001], [012], [241], and [211] for MBI-PG7L, MBI-PG7M1, MBI-PG7M2, and MBI-PG7S, respectively. c see corresponding note to Table 1. d WHP calculations result in close values, DWHP = 56(7) nm and εsWHP = 0.10(9)% for MBI-PG7L, DWHP = 31(3) nm and εsWHP = 0 for MBI-PG7M1, and DWHP = 18(4) nm and εsWHP = 0 for MBI-PG7M2.
Table 3. Results of LB fitting of the XRD patterns of MBI-ChA and MBI-ChA-mill samples obtained by means of Rietveld program TOPAS (see head of Table 1 for description of parameters shown).
Table 3. Results of LB fitting of the XRD patterns of MBI-ChA and MBI-ChA-mill samples obtained by means of Rietveld program TOPAS (see head of Table 1 for description of parameters shown).
Crystalline
Phase a
a, Å
c, Å
Vcell, Å3DTOPAS, nm b
εsTOPAS, % b
Wt, wt.% cRwp, %
Rp, %
cRwp, %
cRp, %
MBI-ChA, me.s.d. = 4.715 d
MBI-ChA114.062(1)
7.156(1)
1225.4(2)9.98(6)
0.000(3)
62.43.58
2.49
4.01
2.91
MBI-ChA213.630(3)
7.129(2)
1147.0(5)9.13(7)
0.006(7)
35.6
MBI-ChA-mill, me.s.d. = 7.492 d
MBI-ChA114.063(3)
7.160(4)
1226.3(8)9.41(7)
0.000(70)
77.33.87
2.33
4.71
3.12
MBI-ChA213.620(6)
7.131(2)
1145.6(8)8.43(11)
0.002(90)
22.7
a Space group of the MBI phases is P42/n (86) (Choice 2 according to [38]). Table unit cell parameters according to CCDC databank, a = 13.950(9) Å, c = 7.192(3) Å, and unit cell volume Vcell = 1212.1(1.2) Å3 (CCDC code 1199885). b Both WHP and SSP calculations result in the same values of D = 8.0(2.5) nm and εs = 0 for MBI-ChA. c The Wt values are estimated by the relative areas under the XRD reflections of the MBI-ChA1 and MBI-ChA2 phases on the XRD pattern, taking into account the same molecular weights of the MBI1 and MBI2 phases (see Section 3.2.4). The mean weight contents of the MBI-ChA1 and MBI-ChA2 phases averaged over MBI-ChA and MBI-ChA-mill samples are Wt = 70(7) wt.% and 30(7) wt.%, respectively. d All e.s.d.s obtained during LB refinement are corrected for underestimation due to serial correlations by multiplying by a coefficient me.s.d.
Table 4. Wavenumber (ν) positions of the absorption peaks in the IR spectra of an asbestos sample with the addition of MBI. The last column shows the interpretation of lines for MBI from [46]. Abbreviations: Γ, out-of-plane bending; δ, planar bending; ν, stretching (stretching vibrations); M, methyl group.
Table 4. Wavenumber (ν) positions of the absorption peaks in the IR spectra of an asbestos sample with the addition of MBI. The last column shows the interpretation of lines for MBI from [46]. Abbreviations: Γ, out-of-plane bending; δ, planar bending; ν, stretching (stretching vibrations); M, methyl group.
ν, cm−1Assignment from [46]
1219δCCH + νCN + νCC
1271νCN + νCC + δCCH
1360νCCH + νCN
1389νCC + νCN + MδCH2
1418CH2 + δCCH
1448δCCH + νCC
1485CH2 + ΓCCCN + δCCH
1555νCN + νCC
1624νCC + νCN
2544
2583
2679
2725
2756
2791
2847
2876
2918
2995
3061CH
3096CH
3111CH
3177νCH
Table 5. Exponent s of conductivity frequency dependence and parameters characterizing hopping conductivity: energy barrier, Wm; localization radius a of carriers; and activation energy Ea in ChA, MBI-ChA, PG, and MBI-PG structures.
Table 5. Exponent s of conductivity frequency dependence and parameters characterizing hopping conductivity: energy barrier, Wm; localization radius a of carriers; and activation energy Ea in ChA, MBI-ChA, PG, and MBI-PG structures.
sWm, eVa, Å
ChA0.800.7712
MBI-ChA0.942.584.4
MBI-PG300.911.726.6
PG70.901.555.1
MBI-PG70.942.587.3
MBI-PG2.50.790.7418.1
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

Balashova, E.; Levin, A.A.; Pavlov, S.; Starukhin, A.; Fokin, A.; Kurdyukov, D.; Eurov, D.; Krichevtsov, B. Synthesis and Study of Organic Nanostructures Fabricated by Inclusion of 2-Methylbenzimidazole Molecules in Nanotubes of Chrysotile Asbestos, Mesoporous Silica, and Nanopores of Borate Glasses. Int. J. Mol. Sci. 2023, 24, 13740. https://doi.org/10.3390/ijms241813740

AMA Style

Balashova E, Levin AA, Pavlov S, Starukhin A, Fokin A, Kurdyukov D, Eurov D, Krichevtsov B. Synthesis and Study of Organic Nanostructures Fabricated by Inclusion of 2-Methylbenzimidazole Molecules in Nanotubes of Chrysotile Asbestos, Mesoporous Silica, and Nanopores of Borate Glasses. International Journal of Molecular Sciences. 2023; 24(18):13740. https://doi.org/10.3390/ijms241813740

Chicago/Turabian Style

Balashova, Elena, Aleksandr A. Levin, Sergey Pavlov, Anatoly Starukhin, Alexander Fokin, Dmitry Kurdyukov, Daniil Eurov, and Boris Krichevtsov. 2023. "Synthesis and Study of Organic Nanostructures Fabricated by Inclusion of 2-Methylbenzimidazole Molecules in Nanotubes of Chrysotile Asbestos, Mesoporous Silica, and Nanopores of Borate Glasses" International Journal of Molecular Sciences 24, no. 18: 13740. https://doi.org/10.3390/ijms241813740

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