Next Article in Journal
Separation of Drugs by Commercial Nanofiltration Membranes and Their Modelling
Previous Article in Journal
Molecular Characterization of Membrane Gas Separation under Very High Temperatures and Pressure: Single- and Mixed-Gas CO2/CH4 and CO2/N2 Permselectivities in Hybrid Networks
Previous Article in Special Issue
MOF-Derived Nanoporous Carbon Incorporated in the Cation Exchange Membrane for Gradient Power Generation
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Hydrothermal Stability of Microporous Niobia-Silica Membranes: Effect of Niobium Doping Contents

1
School of Urban Planning and Municipal Engineering, Xi’an Polytechnic University, Xi’an 710048, China
2
Xi’an Thermal Power Research Institute Co., Ltd., Xi’an 710032, China
*
Author to whom correspondence should be addressed.
Submission received: 14 April 2022 / Revised: 8 May 2022 / Accepted: 13 May 2022 / Published: 17 May 2022

Abstract

:
Methyl-modified niobium-doped silica (Nb/SiO2) materials with various Nb/Si molar ratios (nNb) were fabricated using tetraethoxysilane and methyltriethoxysilane as the silica source and niobium pentachloride as the niobium source by the sol–gel method, and the Nb/SiO2 membranes were prepared thereof by the dip-coating process under an N2 calcining atmosphere. Their microstructures were characterized and gas permeances tested. The results showed that the niobium element existed in the formation of the Nb-O groups in the Nb/SiO2 materials. When the niobium doping content and the calcining temperature were large enough, the Nb2O5 crystals could be formed in the SiO2 frameworks. With the increase of nNb and calcination temperature, the formed particle sizes increased. The doping of Nb could enhance the H2/CO2 and H2/N2 permselectivities of SiO2 membranes. When nNb was equal to 0.08, the Nb/SiO2 membrane achieved a maximal H2 permeance of 4.83 × 10−6 mol·m−2·Pa−1·s−1 and H2/CO2 permselectivity of 15.49 at 200 °C and 0.1 MPa, which also exhibited great hydrothermal stability and thermal reproducibility.

1. Introduction

Environmental protection and resource shortage are nowadays issues that need to be faced in the process of world development. The continuous growth of the world’s population has led to the increasing consumption of the earth’s resources. Some experts have suggested hydrogen as an alternative fuel because of its zero pollution. Nowadays, hydrogen is primarily produced from fossil fuels, such as natural gas and coal through steam reforming/gasification and water gas shift reactions [1]. In order to obtain high purity hydrogen from either syngas or the products of the water-gas shift reaction, separation of H2 from other gases such as CO2, CO, or CH4 is necessary. Consequently, hydrogen purification from the above CO2-containing reaction gas mixture is becoming an important issue. After long-term research and efforts by scientists, a large number of experimental results have shown that the membrane separation technology has shown great potential in gas separation. An inorganic membrane has the advantages of good resistance to high temperature and pressure, high mechanical strength, good chemical stability, long service life, and resistance to halite, which make it attractive in the field of gas separation [2]. In the last two decades, H2-separation membranes have been developed using various materials, such as palladium and its alloys, silica, alumina, etc. In the research to date, inorganic silica membranes, especially those derived from the sol-gel technique, are some of the best among the various inorganic materials for the separation of hydrogen-containing gas mixtures.
However, in high temperature and humid air, pure SiO2 membranes showed poor hydrothermal stability. The Si-O-Si bonds were broken and Si-OH bonds were formed when inter-played with water, resulting in the densification of the silica structure [3,4]. Many scientists have carried out a lot of research work to improve the hydrothermal stability of silica membrane materials. In recent years, the two primary methods have been the introduction of groups such as F, Cl, -CnH2n, -CnH2n+1, phenyl groups, etc. [5,6], and the doping of transition metals such as nickel [7,8,9], palladium [10,11], zirconium [12], niobium [13], magnesium [14], aluminum [15], cobalt [16,17], etc. Wei et al. [6] prepared a perfluorodecyl-modified silica membrane by the sol-gel method using tetraethylorthosilicate (TEOS) and 1H, 1H, 2H, and 2H-perflouorodecyltriethoxysilane (PFDTES) as precursors. The H2 permeance of the as-prepared membrane was 9.71 × 10−7 mol·m−2·s−1·Pa−1, and the H2/CO2 permselectivity and binary gas separation factor were 7.19 and 12.11, respectively. Under humid conditions with a temperature of 250 °C and a water vapor molar ratio of 5%, the single H2 permeance and H2/CO2 permselectivity remained almost constant for at least 200 h.
Among the transition metals, niobium doping has caught researchers’ eyes. Boffa et al. [18] prepared niobia-silica membranes using tetraethyl orthosilicate (TEOS) as the Si source and niobium penta (n-butoxide) as the Nb source. Their research results showed that the hydrothermal stability of the microporous niobia-silica membranes was better than that of the pure SiO2 membrane because the incorporation of Nb ions into the silica matrix. Qi et al. [13] prepared a novel microporous hybrid silica membrane using 1,2-bis (triethoxysilyl) ethane (BTESE) and niobium penta (n-butoxide) as precursors for the permselectivity of CO2. The result showed that the permselectivity of H2/CO2 for the Nb-BTESE membrane could be tuned by altering the calcination temperature. Lin et al. [19] investigated the influence of the sol particle size on the gas permselectivity of the niobium-doped hybrid silica membrane. The prepared Nb/SiO2 membrane had an H2 permeability of 8.36 × 10−8 mol·m−2·s−1·Pa−1 with a mean particle size of 5 nm. However, there have been few studies focusing on the effects of niobium doping content on the microstructures and gas permeances of niobium-doped SiO2 membranes.
In this work, we prepared a kind of new niobium-doped SiO2 membrane using tetraethoxysilane (TEOS) and methyltriethoxysilane (MTES) as the Si sources and niobium pentachloride (NbCl5) as the niobium source. Nb/SiO2 membranes with different Nb/Si molar ratios (nNb) were prepared. The effects of nNb and calcination temperature on the microstructures of Nb/SiO2 materials were studied in detail. Characterization and results were attained by Fourier transform infrared (FTIR) spectroscopy, X-ray diffraction (XRD), X-ray photo electron spectroscopy (XPS), N2 adsorption/desorption measurements, and scanning electron microscopy (SEM). The gas permeation tests and hydrothermal stability of the Nb/SiO2 membranes were performed and are discussed.

2. Experimental

2.1. Fabrication of Nb/SiO2 Sols

The Nb/SiO2 sols were prepared using tetraethoxysilane (TEOS) and methyltriethoxysilane (MTES) as the silica sources, niobium pentachloride (NbCl5) as the niobium source, absolute ethanol (EtOH) as the solvent, and HCl as the catalyst. High purity solid NbCl5 powder was dissolved in absolute ethanol to obtain a 0.43 M NbCl5 solution. This process was carried out in a fume hood, stirring and dissolving with a glass rod until the HCl gas was released. The reaction equation is as follows:
NbCl5 (s) + nC2H5OH (l) → NbCl5-n(OC2H5)n (l) + nHCl (g)
According to the mol ratio of TEOS:MTES:EtOH:H2O:HCl:NbCl5 = 1:0.8:16:7:0.085:nNb, the EtOH, TEOS, and MTES were first mixed with the solution of NbCl5 solution in an ice bath and stirred magnetically for 40 min. After that, a mixture of HCl and water was carefully added drop-wise and then refluxed in a water bath at 60 °C for 3 h. In this way, the Nb-doped SiO2 sols were obtained. The nNb is the molar ratio of Nb/TEOS, which was 0, 0.08, 0.2, and 1, respectively. The Nb-doped SiO2 sol with nNb = 0 is also referred to as the SiO2 sol. The Nb-doped SiO2 sols were diluted three times using absolute ethanol to obtain the final Nb/SiO2 sols.

2.2. Fabrication of Nb/SiO2 Materials

The as-prepared Nb/SiO2 sols were dried in a vacuum oven to prepare the dry gels. The obtained dry gels were then ground into fine powders and calcined under N2 atmosphere in a temperature-controlled tubular furnace at various temperatures (200 °C, 400 °C, 600 °C, 800 °C) for 2 h with a ramping rate of 0.5 °C·min−1. The final niobium-doped silica (Nd/SiO2) materials were produced. The Nb/SiO2 materials with nNb = 0 were also referred to as the SiO2 materials.

2.3. Fabrication of Nb/SiO2 Membranes

To obtain the Nb/SiO2 membranes, part of the above Nb/SiO2 sols were applied to the surface of porous α-Al2O3 composite discs (Hefei Shijie Membrane Engineering Co., Ltd., Hefei, China) by the dip-coating method. The discs each had a thickness of 4 mm, a diameter of 30 mm, a mean pore diameter of 100 nm, and a porosity of 40%. The dipping time was 6 s. Four-layer Nb/SiO2 membranes were prepared, and each Nb/SiO2 layer was individually calcined under N2 atmospheres in a temperature-controlled furnace to 400 °C at a ramping rate of 0.5 °C·min−1 and with a dwell time of 2 h. The prepared Nb/SiO2 membranes were used to test the permeances of H2, CO2, and N2; the preparation process is shown in Figure 1.

2.4. Steam-Treatment and Regeneration of Nb/SiO2 Membranes

The steam stability of the membranes was tested by exposure to saturated steam at 25 °C for 10 d. The thermal regeneration of Nb/SiO2 membranes after steam treatment was carried out at a calcination temperature of 350 °C by the same calcining procedure as described above. After the steam-treatment and regeneration, the gas permeances of Nb/SiO2 membranes were tested, respectively.

2.5. Characterization

The sol densities were determined using a pycnometer. The solid contents of sols were determined by the weighing method. The particle size distributions of Nb/SiO2 sols were measured using a Malvern Nano ZS size analysis instrument (Malvern Instruments Ltd., Malvern, UK) The functional groups of samples were characterized by Fourier transform infrared spectroscopy (FTIR, Nicolet 5700, Thermo Nicolet Corporation, Fitchburg, WI, USA), and the wavelength measurement range was 400–4000 cm−1 as determined by the KBr compression method. The material phase structure was determined by a Rigaku D/max-2550 pc X-ray diffractometer (XRD, Rigaku D/max-2550, Hitachi, Tokyo, Japan) with CuKα radiation under the conditions of 40 kV and 40 mA. The chemical components of Nb/SiO2 samples were performed by an X-ray photoelectron spectrometer (XPS, ESCALAB250xi, Thermo Scientific, Waltham, MA, USA) with AlKα excitation. The morphologies of surfaces and cross-sections for the membranes were observed by scanning electron microscopy (SEM, JEOL JSM-6300, Hitachi, Tokyo, Japan) under 5 kV acceleration voltage. Before the SEM tests, the samples were treated with gold spraying. The BET surface area and pore volume of the samples were measured by N2 sorption/desorption isotherms with a specific surface area and pore and pore-size analyzer (ASAP 2020, Micromeritics, Norcross, GA, USA).
Single gas permeances of Nb/SiO2 membranes were measured under a transmembrane pressure of 0.1-0.4 MPa using the schematic diagram of the experimental setup shown in Figure 2. Prior to the gas permeation measurement, the membranes were mounted in a stainless-steel module with a cylindrical geometry and placed inside the furnace in the gas permeation measurement rig at a temperature range of 25 to 200 °C. The membranes were tested using single gases with different kinetic diameters (H2, CO2, and N2), and permeance was measured by using a soap film flow meter. The gas permselectivities, also known as ideal selectivities, were calculated by the permeance ratio between two gases. The steam stability of the membranes was tested by exposure to saturated steam at 25 °C for 10 d. The thermal regeneration of Nb/SiO2 membranes after steam stability testing was carried out at a calcination temperature of 350 °C by the same calcining procedure as described above. It should also be noted that permeate gas flow was only recorded after a steady state was attained.

3. Results and Discussion

3.1. Analysis of Nb/SiO2 Sol Performance

The influence of nNb on the pH value, density, and solid content of Nb/SiO2 sol is shown in Table 1. The particle size distributions of Nb/SiO2 sols with various nNb are depicted in Figure 3. As seen in Table 1, with increasing nNb, the sol pH value decreased while the density and solid content increased. As seen in Figure 3, with the increase of nNb, the mean particle sizes of Nb/SiO2 sols increased, and their particle size distributions became wider. The particle size distributions of Nb/SiO2 sols with nNb = 0, 0.08, and 0.2 were narrow, and their mean particle sizes were small. However, when nNb was increased to 1.0, the mean particle size of Nb/SiO2 sol increased greatly. In the experiment, the prepared NbCl5 solution was acidic. The increase of nNb in Nb/SiO2 sol made the acidity of the sol increase, which sped up the hydrolysis–polycondensation reaction. As a result, the sol density, solid content, and particle size increased. On the other hand, with the occurrence of the hydrolysis-polycondensation reaction, Nb-O-Si bonds were formed gradually. The radius of the niobium atom (2.08 Å) is larger than that of the Si atom (1.46 Å), and the bond length of Nb-O is longer than that of Si-O (the bond lengths of Nb-O and Si-O are about 2.13 and 1.64 Å, respectively), which contributed to the increase of sol particles.

3.2. Chemical Structure Analysis

The phase-chemical structure of Nb/SiO2 materials may be influenced by the introduced Nb element, so that the effect may have been more evident in the samples containing higher Nb contents. In order to study the effect of calcination temperature on Nb/SiO2 materials, the Nb/SiO2 samples with nNb = 1 were used for the measurement analysis. Figure 4 shows the FTIR spectra of Nb/SiO2 materials with nNb = 1 at different calcination temperatures. In Figure 4, the absorption peaks located at 1055 and 788 cm−1 were associated with the Si-O-Si bonds. The peaks at 2985 cm−1 and 1630 cm−1 were assigned to the mode of -CH3 groups and Si-OH bonds, respectively [11]. The band at 1278 cm−1 was designated as Si-CH3 groups. The intensity of the absorption peaks at 2985 cm−1 and 1278 cm−1 decreased obviously with increases in the calcination temperatures and disappeared as the calcination temperature reached 600 °C, indicating that the -CH3 groups had been broken down at this temperature. The absorption peak observed at 619 cm−1 corresponded to the Nb-O bonds, which enhanced in intensity with the increase of calcination temperatures.
The FTIR spectra of Nb/SiO2 materials with various nNb calcined at 400 °C are provided in Figure 5. It can be seen from Figure 5 that the hydrophobic Si-CH3 groups located at 1278 cm−1 were all involved in the Nb/SiO2 materials with various nNb. The absorption peak locations of Nb/SiO2 materials with nNb = 0.08, 0.2, and 1 were roughly the same, but the intensities of individual peaks were different. The peak located at 619 cm−1 corresponding to the Nb-O groups did not exist in the pure SiO2 materials. This certified that the Nb had been successfully incorporated into the silica frameworks. At the same time, the peak intensity due to the Nb-O groups became strong continuously with the increase of niobium doping.

3.3. Phase Structure Analysis

The XRD patterns of Nb/SiO2 materials with nNb = 1 calcined at various temperatures are given in Figure 6. The broad diffraction peak at the range of about 2θ = 20–30° was assigned to the amorphous SiO2. An obvious crystallization peak of Nb2O5 was detected when the calcination temperature was increased to 600 °C. According to the conclusion from Kosutova, there will be two new crystalline phases formed sequentially when the sample is heated up to 450 °C [20]. The first crystalline phase was identified as hexagonal TT-Nb2O5 (JCPDS No-400-028-0317) [21]. The TT notation comes from the German Tief-Tief (low–low), referring to the temperature at which the structure was observed in the sequence of niobium oxides obtained at elevated temperatures, first used in Ref [22]. In Figure 6, the diffraction peaks at 2θ = 22.6°, 28.6°, 36.8°, 46.2°, 50.8°, and 55.2° were assigned to the (001), (180), (201), (002), (380), and (182) crystal planes, respectively, of hexagonal TT-Nb2O5, which indicated the formation of hexagonal TT-Nb2O5 (JCPDS No.00-030-0873). In Figure 6, another phase transition was observed at 800 °C, which was associated with the transition from hexagonal TT-Nb2O5 to orthorhombic T-Nb2O5 (JCPDS No. 404-007-0752) [23]. The transformation caused the splitting of the diffraction peaks [24]. It is evident that a high calcining temperature can improve the crystallinity of Nb2O5 species.
The XRD patterns of the Nb/SiO2 materials with various nNb calcined at 400 °C are provided in Figure 7. From Figure 7, it can be seen that as the nNb ≤ 0.2, the XRD curves of the Nb/SiO2 materials were similar, and only the diffraction peak between 20° and 30° corresponding to the amorphous SiO2 could be observed. When the nNb was increased to 1, a new diffraction peak assigned to the hexagonal TT-Nb2O5 appeared besides for the amorphous SiO2. This means there was no crystalline form of hexagonal TT-Nb2O5 when the Nb-doped content was low. When the niobium doping content and the calcining temperature were large enough, the Nb2O5 crystals could be formed in the SiO2 frameworks.
In order to explore the crystal form of niobium, the Nb/SiO2 materials with nNb = 1 calcined at various temperatures were characterized by XPS, which are shown in Figure 8. In Figure 8, the NbTT 3d and NbT 3d mean the Nb-O of the hexagonal crystal system TT-Nb2O5 and orthorhombic crystal system T-Nb2O5, respectively. It can be seen that the NbTT 3d5/2 peak and the NbTT 3d3/2 peak appeared at 206.4 eV and 209.1 eV, respectively, in the sample calcined at 400 °C. The data matched with the species of niobium oxide, which proved the formation of Nb2O5 crystals. The temperature related to the formation of the TT-Nb2O5 phase was reported to be approximately 500 °C or higher, according to the studies conducted on nanostructures or thin films [24,25]. Furthermore, the calcination temperature reached 600 °C, and the NbT 3d5/2 and NbT 3d3/2 peaks also appeared at 209.0 eV and 211.8 eV, respectively, which were the Nb 3d peaks of orthorhombic phase T-Nb2O5. The conclusion was as same as that of XRD analysis.

3.4. SEM Analysis

Figure 9 shows the SEM images of Nb/SiO2 materials with various nNb calcined at various temperatures. From Figure 9a–d, it can be seen that when calcined at 400 °C, the samples had the morphology of nanoparticles with dispersed amorphous structures, and the particle sizes increased with increasing nNb.
Comparing Figure 9d with Figure 9e,f, it can be clearly observed that the morphology and particle size of Nb/SiO2 material with nNb = 1 underwent a large change after being calcined at 600 °C, and the materials had a tendency to develop from an amorphous state to a spherical shape. It can be seen from Figure 9f that when the calcination temperature reached 800 °C, the materials exhibited a spherical structure, the particle size increased, and there were more small-sized spherical particles formed around the large particles. In summary, with the increase of calcination temperature and Nb-doping, the particle sizes increased.
Moreover, the SEM images of the pure SiO2 membrane and Nb/SiO2 with nNb = 0.08 calcined at 400 °C are depicted in Figure 10. From Figure 10 it can be observed that there were no visible cracks or pinholes on the membrane surfaces, indicating that the membranes were well coated. It could be seen that niobium doping could make particle sizes increase.

3.5. Pore Structure Analysis

The pore properties of the as-prepared Nb/SiO2 materials were investigated by N2 adsorption/desorption to characterize the surface area, pore volume, and porosity. Figure 11 shows the N2 adsorption/desorption isotherms of Nb/SiO2 materials with various nNb calcined at 400 °C under an N2 atmosphere. After calcining, all isotherms displayed similar tendencies, which were similar to a tape I isotherm. It could be proved that the Nb/SiO2 materials all exhibited the adsorption isotherm characteristics of microporous materials. This material was subjected to a strong force due to the gas in the micropores, so it could quickly reach the absorption saturation state.
Figure 12 manifests the pore size distribution of Nb/SiO2 materials with various nNb calcined at 400 °C. As shown in Figure 12, as nNb ≤ 0.2, with the Nb doping, the pore size of Nb/SiO2 materials were widened, and the microporous structure was maintained. When nNb = 1, the Nb/SiO2 material became dense.
The pore structure parameters of Nb/SiO2 materials with various nNb calcined at 400 °C are shown in Table 2. It can be observed that with the increases of nNb, the mean pore size, BET surface area, and total pore volume increased until nNb = 0.08, and then they began to decrease. This is because the bond length of Nb-O is longer than that of Si-O (the bond lengths of Nb-O and Si-O are about 2.13 and 1.64 Å, respectively). The formation of the Nb-O bond helped the formation of larger particles. Thus, with the increase of the Nb doping amount, the particle size increased, and the formed pore grew gradually. However, when nNb was >0.08, that changed. Some small hexagonal TT-Nb2O5 particles were formed and distributed in the SiO2 network, which led to a decrease in the mean pore size, BET surface area, and total pore volume

3.6. Gas Permeation and Separation Property Analysis

Based on all the results of the above, the Nb-doping content showed a significant impact on the microstructures of Nb/SiO2 materials. Compared with the Nb/SiO2 membrane with nNb = 0.08 and nNb = 0.2, the pore volume of the Nb/SiO2 membrane with nNb = 1 was too small, which is bad for gas separation. Therefore, the Nb/SiO2 membrane with nNb = 1 was not considered here.
A transient test was conducted on the Nb/SiO2 membranes with various nNb at 25 °C and a differential pressure of around 0.1 MPa, and the gas permeances and H2 permselectivities are shown in Figure 13. In Figure 13a, the H2 and N2 permeances of Nb/SiO2 membranes with nNb = 0.08 and 0.2 were greater than those of the pure SiO2 membrane, while the change trends of CO2 permeances were the contrary. Compared with the pure SiO2 membrane, the H2 permeances of Nb/SiO2 membranes with nNb = 0.08 and 0.2 increased by 39.0% and 8.9%, respectively. In Figure 13b, the permselectivities of H2/CO2 and H2/N2 for the pure SiO2 membrane were greater than the values based on Knudsen diffusion (4.69 and 3.74, respectively), which means the transport in pure SiO2 membrane is controlled by molecular sieving plus Knudsen diffusion. The H2/CO2 and H2/N2 permselectivities for the Nb/SiO2 membranes with nNb = 0.08 and 0.2, respectively, were obviously higher than those of the pure SiO2 membrane. Compared with the pure SiO2 membrane, the H2/CO2 permselectivity for the Nb/SiO2 membranes with nNb = 0.08 and 0.2 increased by 59.5% and 6.3%, respectively. The larger mean pore size and higher pore volume of Nb/SiO2 membranes with nNb = 0.08 and 0.2 than those of the SiO2 membrane can explain the increase of gas permeance. The larger mean pore size of Nb/SiO2 membrane indicated that the increase of H2/CO2 permselectivity was not due to molecular sieving. This suggests that the doping of Nb introduced another transport mechanism. Some researchers have proposed that doping transition metals into the microporous SiO2 network will generate Lewis acids on the surface of the membrane materials and ultimately endow the Nb/SiO2-derived microporous membrane with sufficiently high H2/CO2 permselectivity. The exceptionally low permeance of CO2 is explained as a consequence of strong chemical interactions between CO2 and the materials of the membrane pore surface, presumably Nb-bound hydroxy groups. The results indicate that the H2/CO2 separation was based on sorption rather than on the differences in molecular sizes [26]. In other words, the existence of acid sites on the surface of membranes may play a key role in reducing CO2 permeance. The H2/CO2 and H2/N2 permselectivities of the Nb/SiO2 membrane with nNb = 0.08 obtained the maximum. Furthermore, with the further increase of nNb, although there are still acid sites in the membrane materials, the formed Nb2O5 in the high-content Nb/SiO2 materials agglomerates to form a non-selective interfacial gap, which will lead to the densification of the membrane materials [13]. Thereby the average pore size will become smaller, resulting in the decrease of gas permeances and H2 permselectivities of the Nb/SiO2 membrane.
Figure 14 displays the influence of temperature differences on the gas permeance of the Nb/SiO2 membrane with various nNb at a pressure difference of 0.1 MPa. As shown in Figure 14, the H2 permeance and H2/CO2 permselectivities in the Nb/SiO2 membranes with different nNb revealed an upward trend with increasing temperature when maintaining a constant pressure difference of 0.1 MPa, which indicated that the transport of H2 molecules through the membranes was activated.
The permeance and permselectivity of the Nb/SiO2 membrane with nNb = 0.08 at a pressure difference of 0.1 MPa and temperature change from 25 °C to 200 °C are manifested in Figure 15. It can be clearly seen that H2 permeance was significantly enhanced with increasing temperature, and the other permeances of CO2 and N2 were reduced slightly. The results were the same as the temperature dependency of permeance of several gases in the NS (Nb/SiO2) membrane from Boffa [27]. Thus, this gives rise to the permselectivity of H2/CO2 and H2/N2 elevating with increasing temperature. With increasing temperatures, the H2 permeance of the Nb/SiO2 membrane increased gradually, which shows that the permeation behavior of H2 in the membranes mainly follows an activation–diffusion mechanism. In the case of activated diffusion, molecules penetrate through the micropore while being subjected to a repulsive force from the pore walls, and the molecules with sufficient kinetic energy to overcome the repulsive force can penetrate the pores [16]. Conversely, the permeance of CO2 and N2 decreased slightly, similar to the trend of Knudsen diffusion, in which molecules collide with the pore walls more regularly than permeating molecules.
After the previous discussion, we believe that the permselectivity of Nb/SiO2 membranes has a sorption separation mechanism. High temperature is conducive to the permselectivity of H2, indicating that the permselectivity mechanism of H2 to other gases is mainly dominated by activation diffusion, which follows the Arrhenius equation.
F = A 0 exp ( E a RT )
In the formula, F means permeance, A0 means former factor, Eα means apparent activation energy, R means the ideal gas constant, T means the temperature, and Equation (1) can be described in another form:
lnF = lnA 0 E a RT = A E a RT
The 1/T is used as the abscissa and lnF as the ordinate to draw the graph. It can be seen from the above formula that it is a straight line in theory. The apparent activation energy can be calculated from the slope of the straight line. Then the Arrhenius curves of the three gases are shown in Figure 16. Moreover, from the slope of the straight-line fitting in Figure 16, the apparent activation energy of the three gases can be obtained: Em = Ea + Qst [28]. The results are shown in Table 3.
From Table 4, Qst means isosteric heat of adsorption, CO2 exhibits a greater adsorption heat, and the gas with the lowest adsorption heat is H2. The apparent activation energy (Ea) can be positive or negative, depending on their relative magnitudes. A negative value of Ea is generally interpreted as being caused by strong sorption of the molecule on the pore surface. Such a negative value suggests a high enthalpy of sorption [26]. The mobility energy (Em) of gas molecules moving on the surface of the Nb/SiO2 membrane was Em (CO2) > Em (N2) > Em (H2). The permselectivity of the Nb/SiO2 membrane towards H2/CO2 increased rapidly as a function of temperature. This was probably a result of the high activation energy of the mobility of hydrogen and the high heat of sorption of carbon dioxide. This result has never been reported for pure silica [29,30,31,32]. Thus, the strong heat of adsorption should be related to the presence of Nb ions in the microporous framework. Table 3 shows the Ea value of H2, H2 permeance, H2 permselectivities, and mean pore diameter of silica membranes by the sol–gel method from other researchers. From Table 3, it can be observed that it is difficult to improve the H2 permeance and permselectivity at the same time. In addition, a higher Ea always corresponds to a smaller mean pore diameter and lower H2 permeance. This means that the Ea value maybe have a link with the mean pore diameter and the interplay between the molecules of H2 and the pore walls of the membrane. Therefore, compared with other research groups, the Nb-doping in this work may be the reason for the lower Ea value of H2.
Table 3. Ea of H2, H2 permeances, H2 permselectivities, and mean pore diameter for various SiO2 membranes prepared by other researchers using the sol–gel process.
Table 3. Ea of H2, H2 permeances, H2 permselectivities, and mean pore diameter for various SiO2 membranes prepared by other researchers using the sol–gel process.
Membrane TypeTemperature and PressureEa of H2 (kJ·mol−1)H2 Permeance (mol·m−2·Pa−1·s−1)H2 Permselectivities Mean Pore Diameter (nm)Ref.
H2/CO2H2/N2
SiO2200 °C, 2 bar-4.62 × 10−73.710.50.30–0.54[33]
SiO2(400)200 °C, 1 bar817.4 × 10−77.5640.38–0.55[28]
SiO2(600)200 °C, 2 bar7.64.03 × 10−766-0.36–0.38[28]
Pd/SiO2200 °C, 0.3 MPa-7.26 × 10−74.3140.57[34]
Co/SiO2200 °C, 0.2 MPa1.981.97 × 10−510.4813.082.34[35]
Nb/SiO2 *200 °C, 0.1 MPa2.534.83 × 10−615.499.542.4549
* In this work.
The pressure dependence of various gas permeances and H2/CO2 permselectivities in the pure SiO2 membrane and Nb/SiO2 membrane with various nNb at 200 °C were further investigated in the pressure difference range from 0.10 MPa to 0.40 MPa, which is shown in Figure 17. It could be observed that H2 permeance with various nNb increased with the pressure-dependence increase. This is because the pressure difference elevated, and the gas force increased, resulting in an elevation in the gas concentration in the membrane, and then leading to higher H2 permeance. As seen in Figure 17b, the permselectivity of H2/CO2 changed slightly with increases in the pressure difference.
Figure 18 demonstrates the effect of pressure differences on the permeance and permselectivity of different gases for the Nb/SiO2 membrane with nNb = 0.08 and a temperature at 200 °C. As seen in Figure 18a, as the pressure difference increased from 0.10 to 0.40 MPa, all of the gas permeances increased. The reason is that the increases in intake pressure resulted in an elevation of the gas concentration, thus yielding a higher permeance. Furthermore, the relationship between permselectivity and pressure differences is shown in Figure 18b. For example, with the pressure increasing from 0.10 MPa to 0.40 MPa, the permeance of N2 and CO2 slightly increased. This was due to the small influence of pressure on Knudsen diffusion, as previously reported in the literature [30]. Hence, no matter how high the pressure was, the change in permselectivity would not increase significantly.
Traditional SiO2 membranes have poor hydrothermal stability due to a large amount of Si-OH groups on their surfaces, which easily absorb water vapor in the air. In the Nb/SiO2 membrane, not only the Nb-doping but also the introduced hydrophobic groups can improve the vapor stability, which can reduce the hydroxyl groups on the pore surface and enhance the hydrophobicity. In order to test the hydrothermal stability of membranes, the Nb/SiO2 membranes with nNb = 0 and 0.08 were chosen to investigate the gas permeance before and after steam treatment.
Figure 19 shows the permeances of various gases (H2, CO2, and N2) and H2 permselectivities of Nb/SiO2 membranes with nNb = 0 and 0.08 at 25 °C and a pressure difference of 0.1 MPa before and after steam treatment and regeneration. As shown in Figure 19, compared with the fresh membranes, the H2 permeances of SiO2 and Nb/SiO2 membranes after steam treatment decreased by 17.90% and 6.68%, respectively, while their H2/CO2 permselectivities decreased by 3.2% and increased by 1.6%, respectively. After regeneration by calcination at 350 °C, the gas permeances and the permselectivities of H2/CO2 and H2/N2 for the two membranes showed an upward trend. Compared with the fresh membranes, the H2 permeances of SiO2 and Nb/SiO2 membranes after regeneration decreased by 10.95% and 3.21%, respectively, while their H2/CO2 permselectivities increased by 2.8% and 2.1%, respectively. The above results indicate that niobium doping improves the hydrothermal stability of SiO2 membranes.

4. Conclusions

To sum up, using the sol-gel technique, Nb/SiO2 materials and membranes with various nNb were successfully synthesized. Their microstructures and gas permeances were investigated. The results showed that the niobium element existed in the formation of the Nb-O groups in the Nb/SiO2 materials. As the niobium doping content and the calcining temperature were high enough, the Nb2O5 crystals could be formed in the SiO2 frameworks. With the increase of nNb, the formed particle sizes increased, and the mean pore size, BET surface area, and total pore volume also increased until nNb = 0.08, and then they began to decrease. The doping of Nb could enhance the H2/CO2 and H2/N2 permselectivities of the SiO2 membrane. When nNb was equal to 0.08, the Nb/SiO2 membrane achieved a maximal H2 permeance of 4.83 × 10−6 mol·m−2·Pa−1·s−1 and H2/CO2 permselectivity of 15.49 at 200 °C and 0.1MPa, which increased by 36.7% and 155.47%, respectively, compared with that of the pure SiO2 membrane. Compared with the fresh membranes, the H2 permeances of SiO2 and Nb/SiO2 membranes after steam treatment decreased by 17.90% and 6.68%, respectively, while their H2/CO2 permselectivities decreased by 3.2% and increased by 1.6%, respectively. After regeneration, the gas permeances and the permselectivities of H2/CO2 and H2/N2 for the two membranes showed an upward trend. Compared with the fresh membrane, the H2 permeances of SiO2 and Nb/SiO2 membranes after regeneration decreased by 10.95% and 3.21%, respectively, while their H2/CO2 permselectivities increased by 2.8% and 2.1%, respectively. Niobium doping improved the hydrothermal stability of the SiO2 membrane. The Nb/SiO2 membranes also exhibited great thermal reproducibility.

Author Contributions

Conceptualization, J.X. and J.Y.; methodology, H.Z.; formal analysis, Y.G. and R.Z.; writing-original draft preparation, J.X. and J.Y.; project administration, J.Y.; funding acquisition, J.Y. and Y.G. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Scientific Research Project of Shaanxi province of China [2022SF-287 and 2021GY-147] and the Scientific Research Project of Shaanxi Education Department, China [19JC017 and 21JK0650].

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Johnston, B.; Mayo, M.C.; Anshuman, K. Hydrogen: The energy source for the 21st century. Technovation 2005, 25, 569–585. [Google Scholar] [CrossRef]
  2. Ali, A.; Pothu, R.; Siyal, S.; Phulpoto, D.; Sajjad, M.; Thebo, K. Graphene-based membranes for CO2 separation. Mater. Sci. Energy Technol. 2019, 2, 83–88. [Google Scholar] [CrossRef]
  3. Ameh, A.E.; Eze, C.P.; Antunes, E.; Cornelius, M.-L.U.; Musyoka, N.; Petrik, L.F. Stability of fly ash-based BEA-zeolite in hot liquid phase. Catal. Today 2019, 357, 416–424. [Google Scholar] [CrossRef]
  4. Prodinger, S.; Derewinski, M.A. Recent Progress to Understand and Improve Zeolite Stability in the Aqueous Medium. Pet. Chem. 2020, 60, 420–436. [Google Scholar] [CrossRef]
  5. Yang, X.; Du, H.; Lin, Y.; Song, L.; Zhang, Y.; Gao, X.; Kong, C.; Chen, L. Hybrid organosilica membrane with high CO2 permselectivity fabricated by a two-step hot coating method. J. Membr. Sci. 2016, 506, 31–37. [Google Scholar] [CrossRef]
  6. Wei, Q.; Ding, Y.-L.; Nie, Z.-R.; Liu, X.-G.; Li, Q.-Y. Wettability, pore structure and performance of perfluorodecyl-modified silica membranes. J. Membr. Sci. 2014, 466, 114–122. [Google Scholar] [CrossRef]
  7. Li, W.; Li, D.; Zheng, L.; Fan, J.; Zhang, Z. Preparation of Nickel-Doped Silica Films and Gas Separation Performance for CH4/CO2. J. Chin. Ceram. Soc. 2014, 42, 416–422. [Google Scholar]
  8. Yang, J.; Xing, X.; Zhao, Y.; Mu, R.; Guo, Y.; Hou, H. Microstructures of nickel-doped methylated silica membrane materials calcined in air: Influence of Ni content. Ferroelectrics 2020, 562, 85–95. [Google Scholar] [CrossRef]
  9. Yang, J.; Tian, L. Preparation, Characterization and Surface Free Energy of Nickel-Doped Silica Organic Inorganic Hybrid Membrane for H2/CO2 Separation. J. Nanosci. Nanotechnol. 2019, 19, 3180–3186. [Google Scholar] [CrossRef]
  10. Lei, J.; Song, H.; Wei, Y.; Zhao, S.; Qi, H. A novel strategy to enhance hydrothermal stability of Pd-doped organosilica membrane for hydrogen separation. Microporous Mesoporous Mater. 2017, 253, 55–63. [Google Scholar] [CrossRef]
  11. Gu, Y.; Zhong, S. Preparation and Application of Supported Palladium-Modified Polyimide-Silica Hybrid Membrane for Selective Separation of H2. Chin. J. Catal. 2006, 3, 250–254. [Google Scholar]
  12. Hove, M.; Nijmeijer, A.; Winnubst, L. Facile synthesis of zirconia doped hybrid organic inorganic silica membranes. Sep. Purif. Technol. 2015, 147, 372–378. [Google Scholar] [CrossRef]
  13. Qi, H.; Han, J.; Xu, N. Effect of calcination temperature on carbon dioxide separation properties of a novel microporous hybrid silica membrane. J. Membr. Sci. 2011, 382, 231–237. [Google Scholar] [CrossRef]
  14. Karakiliç, P.; Huiskes, C.; Luiten-Olieman, M.W.J.; Nijmeijer, A.; Winnubst, L. Sol-gel processed magnesium-doped silica membranes with improved H2/CO2 separation. J. Membr. Sci. 2017, 543, 195–201. [Google Scholar] [CrossRef]
  15. Kanezashi, M.; Miyauchi, S.; Nagasawa, H.; Yoshioka, T.; Tsuru, T. Gas permeation properties through Al-doped organosilica membranes with controlled network size. J. Membr. Sci. 2014, 466, 246–252. [Google Scholar] [CrossRef]
  16. Igi, R.; Yoshioka, T.; Ikuhara, Y.H.; Iwamoto, Y.; Tsuru, T. Characterization of Co-doped silica for improved hydrothermal stability and application to hydrogen separation membranes at high temperatures. J. Am. Ceram. Soc. 2008, 91, 2975–2981. [Google Scholar] [CrossRef]
  17. Yang, J.; Ai, Y.; Fan, W.; Mu, R.; Chen, X.; Hou, H. Cobalt-Doped Silica Organic-Inorganic Materials by Sol-Gel Method: Preparation and Thermal Stability Calcined under N2 Atmosphere. Integr. Ferroelectr. 2021, 219, 247–259. [Google Scholar] [CrossRef]
  18. Boffa, V.; Blank, D.; Elshof, J.E.T.; Tenelshof, J. Hydrothermal stability of microporous silica and niobia–Silica membranes. J. Membr. Sci. 2008, 319, 256–263. [Google Scholar] [CrossRef]
  19. Lin, D.; Song, H.; Zhu, H.; Chen, J.; Qi, H. The influence of the particle size of polymer sol on the gas separation performance of niobium-doped organic-inorganic hybrid SiO2 membrane. Membr. Sci. Technol. 2016, 36, 23–29. [Google Scholar]
  20. Košutová, T.; Horák, L.; Pleskunov, P.; Hanuš, J.; Nikitin, D.; Kúš, P.; Cieslar, M.; Gordeev, I.; Burazer, S.; Choukourov, A.; et al. Thermally-driven morphogenesis of niobium nanoparticles as witnessed by in-situ x-ray scattering. Mater. Chem. Phys. 2021, 277, 125466. [Google Scholar] [CrossRef]
  21. Frevel, L.K.; Rinn, H.W. Powder Diffraction Standards for Niobium Pentoxide and Tantalum Pentoxide. Anal. Chem. 1955, 27, 1329–1330. [Google Scholar] [CrossRef]
  22. Brauer, G. Die oxyde des niobs. Z. Anorg. Allg. Chem. 1941, 248, 1–31. [Google Scholar] [CrossRef]
  23. Kato, K.; Tamura, S. Die kristallstruktur von T-Nb2O5. Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem. 1975, 31, 673–677. [Google Scholar] [CrossRef]
  24. Angerer, P.; van Karsbergen, V.; Weinberger, N.; Strauss, G.; Neubauer, E.; Friessnegger, B.; Marsoner, S. In-situ high-temperature X-ray diffraction investigations of magnetron sputtered niobium oxide layers up to 900 °C. Thin Solid Films 2019, 674, 33–38. [Google Scholar] [CrossRef]
  25. Francisco, M.; Gushikem, Y. Synthesis and characterization of SiO2–Nb2O5 systems prepared by the sol–gel method: Structural stability studies. J. Mater. Chem. 2002, 12, 2552–2558. [Google Scholar] [CrossRef] [Green Version]
  26. Boffa, V.; Elshof, J.E.T.; Petukhov, A.V.; Blank, D.H.A. Microporous Niobia–Silica Membrane with Very Low CO2 Permeability. ChemSusChem 2008, 1, 437–443. [Google Scholar] [CrossRef] [Green Version]
  27. Boffa, V.; Elshof, J.T.; Garcia, R.; Blank, D. Microporous niobia-silica membranes: Influence of sol composition and structure on gas transport properties. Microporous Mesoporous Mater. 2009, 118, 202–209. [Google Scholar] [CrossRef]
  28. de Vos, R.M.; Verweij, H. Improved performance of silica membranes for gas separation. J. Membr. Sci. 1998, 143, 37–51. [Google Scholar] [CrossRef]
  29. Nair, B.N.; Keizer, K.; Suematsu, H.; Suma, Y.; Kaneko, N.; Ono, S.; Okubo, A.T.; Nakao, S.-I. Synthesis of Gas and Vapor Molecular Sieving Silica Membranes and Analysis of Pore Size and Connectivity. Langmuir 2000, 16, 4558–4562. [Google Scholar] [CrossRef]
  30. Uhlhorn, R.; Keizer, K.; Burggraaf, A. Gas transport and separation with ceramic membranes. Part II. Synthesis and separation properties of microporous membranes. J. Membr. Sci. 1992, 66, 271–287. [Google Scholar] [CrossRef] [Green Version]
  31. Kanezashi, M.; Fujita, T.; Asaeda, M. Nickel-Doped Silica Membranes for Separation of Helium from Organic Gas Mixtures. Sep. Sci. Technol. 2005, 40, 225–238. [Google Scholar] [CrossRef]
  32. Yoshida, K.; Hirano, Y.; Fujii, H.; Tsuru, T.; Asaeda, M. Hydrothermal Stability and Performance of Silica-Zirconia Membranes for Hydrogen Separation in Hydrothermal Conditions. J. Chem. Eng. Jpn. 2001, 34, 523–530. [Google Scholar] [CrossRef] [Green Version]
  33. Qureshi, H.F.; Nijmeijer, A.; Winnubst, L. Influence of sol–gel process parameters on the micro-structure and performance of hybrid silica membranes. J. Membr. Sci. 2013, 446, 19–25. [Google Scholar] [CrossRef]
  34. Song, H.; Zhao, S.; Lei, J.; Wang, C.; Qi, H. Pd-doped organosilica membrane with enhanced gas permeability and hydrothermal stability for gas separation. J. Mater. Sci. 2016, 51, 6275–6286. [Google Scholar] [CrossRef]
  35. Wang, L.; Yang, J.; Mu, R.; Guo, Y.; Hou, H. Sol-Gel Processed Cobalt-Doped Methylated Silica Membranes Calcined under N2 Atmosphere: Microstructure and Hydrogen Perm-Selectivity. J. Mater. 2021, 14, 4188. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of the fabrication of Nb/SiO2 membranes by the sol–gel process.
Figure 1. Schematic diagram of the fabrication of Nb/SiO2 membranes by the sol–gel process.
Membranes 12 00527 g001
Figure 2. Schematic diagram of the experimental setup.
Figure 2. Schematic diagram of the experimental setup.
Membranes 12 00527 g002
Figure 3. The particle size distributions of Nb/SiO2 sols with various nNb.
Figure 3. The particle size distributions of Nb/SiO2 sols with various nNb.
Membranes 12 00527 g003
Figure 4. FTIR spectra of Nb/SiO2 materials with nNb = 1 calcined at various temperatures.
Figure 4. FTIR spectra of Nb/SiO2 materials with nNb = 1 calcined at various temperatures.
Membranes 12 00527 g004
Figure 5. FTIR spectra of Nb/SiO2 materials with various nNb calcined at 400 °C.
Figure 5. FTIR spectra of Nb/SiO2 materials with various nNb calcined at 400 °C.
Membranes 12 00527 g005
Figure 6. XRD patterns of Nb/SiO2 materials with nNb = 1 calcination at various temperatures.
Figure 6. XRD patterns of Nb/SiO2 materials with nNb = 1 calcination at various temperatures.
Membranes 12 00527 g006
Figure 7. XRD patterns of Nb/SiO2 materials doped with various nNb calcined at 400 °C.
Figure 7. XRD patterns of Nb/SiO2 materials doped with various nNb calcined at 400 °C.
Membranes 12 00527 g007
Figure 8. The Nb 3d XPS patterns of Nb/SiO2 materials with nNb = 1 calcined at various temperatures: (a) 400 °C, (b) 600 °C, and (c) 800 °C.
Figure 8. The Nb 3d XPS patterns of Nb/SiO2 materials with nNb = 1 calcined at various temperatures: (a) 400 °C, (b) 600 °C, and (c) 800 °C.
Membranes 12 00527 g008
Figure 9. SEM image of Nb/SiO2 materials with various nNb calcinations at various temperatures: (a) nNb = 0, 400 °C; (b) nNb = 0.08, 400 °C; (c) nNb = 0.2, 400 °C; (d) nNb = 1, 400 °C; (e) nNb = 1, 600 °C; (f) nNb = 1, 800 °C.
Figure 9. SEM image of Nb/SiO2 materials with various nNb calcinations at various temperatures: (a) nNb = 0, 400 °C; (b) nNb = 0.08, 400 °C; (c) nNb = 0.2, 400 °C; (d) nNb = 1, 400 °C; (e) nNb = 1, 600 °C; (f) nNb = 1, 800 °C.
Membranes 12 00527 g009
Figure 10. SEM image of Nb/SiO2 membranes with nNb = (a) 0 and (b) 0.08 calcined at 400 °C.
Figure 10. SEM image of Nb/SiO2 membranes with nNb = (a) 0 and (b) 0.08 calcined at 400 °C.
Membranes 12 00527 g010
Figure 11. N2 adsorption–desorption isotherms of Nb/SiO2 materials doped with various nNb calcined at 400 °C.
Figure 11. N2 adsorption–desorption isotherms of Nb/SiO2 materials doped with various nNb calcined at 400 °C.
Membranes 12 00527 g011
Figure 12. Pore size distribution of Nb/SiO2 materials doped with various nNb calcined at 400 °C.
Figure 12. Pore size distribution of Nb/SiO2 materials doped with various nNb calcined at 400 °C.
Membranes 12 00527 g012
Figure 13. Effect of nNb on (a) gas permeance and (b) H2 permselectivity of Nb/SiO2 membrane at a pressure difference of 0.1 MPa and 25 °C.
Figure 13. Effect of nNb on (a) gas permeance and (b) H2 permselectivity of Nb/SiO2 membrane at a pressure difference of 0.1 MPa and 25 °C.
Membranes 12 00527 g013
Figure 14. Effect of temperature on (a) H2 permeance and (b) H2/CO2 permselectivity of Nb/SiO2 membranes with different nNb at a pressure difference of 0.1 MPa.
Figure 14. Effect of temperature on (a) H2 permeance and (b) H2/CO2 permselectivity of Nb/SiO2 membranes with different nNb at a pressure difference of 0.1 MPa.
Membranes 12 00527 g014
Figure 15. Effect of temperature on (a) gas permeance and (b) H2 permselectivity of Nb/SiO2 membrane with nNb = 0.08 at a pressure difference of 0.1 MPa.
Figure 15. Effect of temperature on (a) gas permeance and (b) H2 permselectivity of Nb/SiO2 membrane with nNb = 0.08 at a pressure difference of 0.1 MPa.
Membranes 12 00527 g015
Figure 16. Arrhenius plots of gas (H2, CO2, and N2) permeances in Nb/SiO2 membrane with nNb = 0.08 at a pressure difference of 0.1 MPa.
Figure 16. Arrhenius plots of gas (H2, CO2, and N2) permeances in Nb/SiO2 membrane with nNb = 0.08 at a pressure difference of 0.1 MPa.
Membranes 12 00527 g016
Figure 17. Effect of pressure difference on (a) H2 permeance and (b) H2/CO2 permselectivity of Nb/SiO2 membranes with different nNb at 200 °C.
Figure 17. Effect of pressure difference on (a) H2 permeance and (b) H2/CO2 permselectivity of Nb/SiO2 membranes with different nNb at 200 °C.
Membranes 12 00527 g017
Figure 18. Effect of pressure difference on (a) gas permeances and (b) H2 permselectivities of Nb/SiO2 membrane at 200 °C.
Figure 18. Effect of pressure difference on (a) gas permeances and (b) H2 permselectivities of Nb/SiO2 membrane at 200 °C.
Membranes 12 00527 g018
Figure 19. (a) Gas permeance and H2 permselectivities and (b) H2 permselectivities of Nb/SiO2 membranes with nNb = 0 and 0.08 at a pressure difference of 0.1 MPa before and after steam treatment and regeneration.
Figure 19. (a) Gas permeance and H2 permselectivities and (b) H2 permselectivities of Nb/SiO2 membranes with nNb = 0 and 0.08 at a pressure difference of 0.1 MPa before and after steam treatment and regeneration.
Membranes 12 00527 g019
Table 1. The influence of nNb on the pH value, density, and solid content of Nb/SiO2 sol.
Table 1. The influence of nNb on the pH value, density, and solid content of Nb/SiO2 sol.
nNbpHDensity/g·cm−3Solid Content/%
03.41 ± 0.040.8418 ± 0.000722.31 ± 0.04
0.082.93 ± 0.030.8529 ± 0.000622.58 ± 0.05
0.22.64 ± 0.020.8710 ± 0.000822.80 ± 0.06
11.02 ± 0.020.9130 ± 0.000624.46 ± 0.07
Table 2. Pore structure parameters of Nb/SiO2 materials with various nNb calcined at 400 °C.
Table 2. Pore structure parameters of Nb/SiO2 materials with various nNb calcined at 400 °C.
nNbBET/ m2·g−1Vt/ cm3·g−1Vmic/ cm3·g−1Mean Pore Width/nm
0386.45450.17160.11151.8759
0.08778.71210.49010.08672.4549
0.2535.40720.46320.07482.2591
186.15990.07620.02661.2176
Table 4. Ea, Qst, and Em values of gases (H2, CO2, and N2) calculated from the Arrhenius formula for the Nb/SiO2 membrane with nNb = 0.08 at a pressure difference of 0.1 MPa.
Table 4. Ea, Qst, and Em values of gases (H2, CO2, and N2) calculated from the Arrhenius formula for the Nb/SiO2 membrane with nNb = 0.08 at a pressure difference of 0.1 MPa.
GasesEa/ kJ·mol−1Qst/ kJ·mol−1 [28] Em/ kJ·mol−1
H22.5368.53
CO2−4.282419.72
N2−4.071813.93
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Xia, J.; Yang, J.; Zhang, H.; Guo, Y.; Zhang, R. Microstructure and Hydrothermal Stability of Microporous Niobia-Silica Membranes: Effect of Niobium Doping Contents. Membranes 2022, 12, 527. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes12050527

AMA Style

Xia J, Yang J, Zhang H, Guo Y, Zhang R. Microstructure and Hydrothermal Stability of Microporous Niobia-Silica Membranes: Effect of Niobium Doping Contents. Membranes. 2022; 12(5):527. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes12050527

Chicago/Turabian Style

Xia, Jiachen, Jing Yang, Hao Zhang, Yingming Guo, and Ruifeng Zhang. 2022. "Microstructure and Hydrothermal Stability of Microporous Niobia-Silica Membranes: Effect of Niobium Doping Contents" Membranes 12, no. 5: 527. https://0-doi-org.brum.beds.ac.uk/10.3390/membranes12050527

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