Next Article in Journal
Heat Treatments for Stress Relieving AlSi9Cu3 Alloy Produced by Laser Powder Bed Fusion
Next Article in Special Issue
Can Mesoporous Silica Speed Up Degradation of Benzodiazepines? Hints from Quantum Mechanical Investigations
Previous Article in Journal
Fluorescent Materials for Monitoring Mitochondrial Biology
Previous Article in Special Issue
Effect of Modification of Amorphous Silica with Ammonium Agents on the Physicochemical Properties and Hydrogenation Activity of Ir/SiO2 Catalysts
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Sol-Gel Processed Cobalt-Doped Methylated Silica Membranes Calcined under N2 Atmosphere: Microstructure and Hydrogen Perm-Selectivity

1
School of Urban Planning and Municipal Engineering, Xi’an Polytechnic University, Xi’an 710048, China
2
School of Environment & Chemical Engineering, Xi’an Polytechnic University, Xi’an 710048, China
*
Author to whom correspondence should be addressed.
Materials 2021, 14(15), 4188; https://doi.org/10.3390/ma14154188
Submission received: 1 July 2021 / Revised: 16 July 2021 / Accepted: 24 July 2021 / Published: 27 July 2021
(This article belongs to the Special Issue Mesoporous Silica and Their Applications)

Abstract

:
Methyl-modified, cobalt-doped silica (Co/MSiO2) materials were synthesized by sol-gel technique calcined in N2 atmospheres, and membranes were made thereof by coating method. The effects of Co/Si molar ratio (nCo) on the physical-chemical constructions of Co/MSiO2 materials and microstructures of Co/MSiO2 membranes were systematically investigated. The gas permeance performance and hydrothermal stability of Co/MSiO2 membranes were also tested. The results show that the cobalt element in Co/MSiO2 material calcined at 400 °C exists not only as Si–O–Co bond but also as Co3O4 and CoO crystals. The introduction of metallic cobalt and methyl can enlarge the total pore volume and average pore size of the SiO2 membrane. The activation energy (Ea) values of H2, CO2, and N2 for Co/MSiO2 membranes are less than those for MSiO2 membranes. When operating at a pressure difference of 0.2 MPa and 200 °C compared with MSiO2 membrane, the permeances of H2, CO2, and N2 for Co/MSiO2 membrane with nCo = 0.08 increased by 1.17, 0.70, and 0.83 times, respectively, and the perm-selectivities of H2/CO2 and H2/N2 increased by 27.66% and 18.53%, respectively. After being steamed and thermally regenerated, the change of H2 permeance and H2 perm-selectivities for Co/MSiO2 membrane is much smaller than those for MSiO2 membrane.

1. Introduction

Hydrogen has been recognized as an ideal energy carrier because of its clean, renewable, and high-calorific value features [1]. However, the industrially produced hydrogen from water-gas conversion process or steam reforming process [2,3] contains some other contaminants (impurities), such as CO2, N2, CH4, CO, H2O, etc. Additionally, from an environmental point of view, refining and extracting hydrogen from industrial waste gas is necessary [4]. Therefore, in order to obtain high-purity hydrogen, it is necessary to separate H2 from the gas mixture. Consequently, the upgrading of hydrogen is of significant attention due to the versatile requirement for hydrogen with good purity in transportation, distributed heat, power generation, and other advanced applications [5]. Membrane separation technology has become an interesting alternative for separation and purification of hydrogen in the industry due to its energy-saving and excessive efficiency. Currently, the membranes for separating hydrogen from gas mixture mainly include inorganic membranes, such as palladium metal membranes, carbon molecular sieve membranes, and microporous ceramic membranes [6,7,8]. Among of them, microporous silica membranes are the most widely investigated because of their good chemical stability, large gas-permeation flux, and high selectivity.
However, pure silica membranes exhibit poor hydrothermal stability in high temperature and humid air. Because Si–O–Si linkages are damaged upon interplay with water, Si–OH hydroxyl agencies are formed, which creates destruction and reconstruction of Si–O–Si bonds in the silica structure, resulting in densification of the silica structure [9,10]. This has been identified as a drawback in the development of silica membranes for gas separation in moist environments for practical industrial applications. An extremely good quantity of work has been done to develop and improve the stability of silica membranes under hydrothermal conditions. For one thing, the incorporation of hydrophobic groups can improve the stability of silica membranes under hydrothermal conditions, such as octyl [11], phenyl [12], alkylamine [13], methyl [14,15], perfluorodecyl [16], etc., which effectively reduce the Si–OH concentration on the surface of silica membrane materials, thereby reducing the physical adsorption of water molecules and enhancing the hydrophobicity of silica membranes. For another, further improvement of hydrothermal stability of silica membranes have been researched by introducing different types of inorganic metal/metal oxides, such as aluminum [17], zirconium [18], nickel [19,20], titanium [21], cobalt [22,23], palladium [24,25], magnesium [26], niobium [27], PdCo [28], FeCo [29], etc. They were all added in the process of sol synthesis, which produced great beneficial effects. A large number of studies have found that by introducing metal, the structure of the membrane appears to be denser, and the structural stability of the membrane material is particularly improved. This is because the mixed oxide network structure formed by incorporating transition metals is more stable than amorphous silica materials [27,30]. Among them, cobalt (Co) is an excellent dopant, and the doping of Co to SiO2 matrix can reduce water adsorption, providing greater resistance by way of reducing the hydrophilicity of silica [31].
Numerous investigations have been conducted on silica materials/membranes modified by cobalt. For example, Smart et al. [32] synthesized methyl-modified Co/SiO2 membranes calcined at 630 °C under air atmosphere using methyltriethoxysilane (MTES) and tetraethylorthosilicate (TEOS) as the silica source and Co(NO3)2·6H2O as the cobalt source. When testing the permeability of single gas at 600 °C and a feed pressure of 600 KPa, it was observed that the H2 permeation reached 1.9 × 10−7 mol·m−2·Pa−1·s−1, and a H2/CO2 perm-selectivity exceeded 1500. Liu et al. [33] investigated the hydrothermal stability of the Co/SiO2 xerogels calcined at 630 °C in an air atmosphere under various hydrothermal treatment conditions. For unstable xerogels (cobalt/silicon < 0.25), their stability was significantly reduced due to steam content and exposure time, leading to a surface-area reduction of nearly 90%. However, it is found that the xerogels with high cobalt content (cobalt/silicon ≥ 0.25) contained Co3O4 and were more stable, with a surface-area reduction of less than 25%. Esposito et al. [34] prepared cobalt-doped silica nanocomposites with various cobalt contents (cobalt/silicon = 0.111, 0.250, and 0.428) by the sol-gel process. After treatment at 400 °C under air atmosphere, the lowest cobalt-loading, cobalt-doped silica nanocomposites appeared amorphous and contained solely tetrahedral complexes of Co2+, whilst Co3O4 was current as the solely crystalline section at greater cobalt content, besides the strong interaction of Co2+ ions with the siloxane matrix. Many research groups found that Co3O4 was the main existing form of cobalt in the Co/SiO2 material calcined under air atmosphere. However, Co3O4 is unstable in a hydrogen atmosphere and is easily reduced. According to Uhlmann et al. [35], after hydrogen reduction, the cobalt-doped xerogels calcined at 500 °C lost their crystal structure and showed no Co3O4 or CoO peaks but only a wide peak similar to that of amorphous silica. There have been a number of works revealing the effects of preparation conditions on the properties of cobalt-doped silica materials/membranes. Unfortunately, as far as we are aware, the influence of calcination atmospheres on the microstructures and characteristics of permeability for Co/SiO2 membrane is crucial, but it has rarely been reported before, especially under a non-oxidizing atmosphere, such as N2 atmosphere. Besides, there are few papers elaborating the influence of Co/Si molar ratio and methyl modification on the microstructures and characteristics of permeability for Co/SiO2 membrane.
In this work, methyl-modified Co/SiO2 (Co/MSiO2) materials and membranes with different Co/Si molar ratio (nCo) were prepared. The effect of nCo on the physical-chemical structures and microstructures of Co/MSiO2 membrane calcined under N2 atmosphere was studied in detail. Characterization and results were attained by X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), Fourier transform infrared (FTIR) spectroscopy, N2 sorption/desorption measurements, transmission electron microscope (TEM), and scanning electron microscope (SEM). Some gas-permeation measurements of Co/MSiO2 membranes were performed and compared with each other. By observing the changes of gas-permeation characteristics before and after exposure to saturated steam, the hydrothermal stability of the Co/MSiO2 membranes was additionally investigated. Subsequently, regeneration performance of the Co/MSiO2 membranes for single gas was discussed.

2. Experimental Part

2.1. Fabrication of Methyl-Modified Co/SiO2 Sols

The methyl-modified Co/SiO2 (Co/MSiO2) sols were prepared via sol-gel process, using methyltriethoxysilane (MTES, purity 99%) and tetraethylorthosilicate (TEOS, p.a. grade) as silica sources, the cobaltous nitrate hexahydrate (Co(NO3)2·6H2O, purity 98%) as a cobalt source, and the nitric acid (HNO3, p.a. grade) as a catalyst in conjunction with absolute ethanol (EtOH, purity 99.7%) as a solvent. The specific synthesis process is as follows: in accordance to the molar ratio of MTES:TEOS:Co(NO3)2·6H2O:HNO3:EtOH:H2O = 0.8:1:nCo:0.085:8.5:6.8, the required quantity of MTES, TEOS, and Co(NO3)2·6H2O solution was completely dissolved in ethanol. Then, the mixed solution was placed in an ice-water bath and stirred intensely to make it fully mixed into a homogeneous solution. After the mixture solution of H2O and HNO3 was poured dropwise, the reaction mixture was then stirred continuously and refluxed in a water bath at 60 °C for 180 min. Therefore, a final Co/MSiO2 sol was obtained. The nCo is Co/Si molar ratio, which is 0, 0.08, 0.15, and 0.5, respectively.

2.2. Fabrication of Unsupported Co/MSiO2 Materials

The prepared Co/MSiO2 sols were dried at 40 °C in a vacuum oven to prepare the dry gels. The obtained dry gels were then ground into fine powders and calcined under N2 atmosphere at 400 °C in a temperature-controlled tubular furnace with the temperature rising rate of 1 °C·min−1 for a resident time of 2 h to prepare the unsupported Co/MSiO2 materials.

2.3. Fabrication of Supported Co/MSiO2 Membranes

To obtain the supported Co/MSiO2 membranes, part of the above Co/MSiO2 sols was applied to the surface of porous α-Al2O3 composite discs (Hefei Shijie Membrane Engineering Co. Ltd., Hefei, China) by dip-coating method. The discs had a thickness of 4 mm, a diameter of 20 mm, a mean pore diameter of 100 nm, and a porosity of 40%. The dipping time was 6 s. After dipping, they were dried for 3 h at 40 °C in an electric heating blast drying oven and then calcined at 400 °C under N2 atmosphere in a tubular furnace for a resident time of 2 h. The process for dipping, drying, and calcination was performed once more in order to minimize any defects that might be occurred in the Co/MSiO2 membrane layer. The Co/MSiO2 membrane with nCo = 0 is also referred to as MSiO2 membrane. The prepared supported MSiO2 and Co/MSiO2 membranes were used to test the permeances of H2, CO2, and N2.

2.4. Characterization

The material phase structure was determined by a Rigaku D/max-2550pc X-ray diffractometer (XRD, Rigaku D/max-2550pc, Hitachi, Tokyo, Japan) with CuKα radiation under the conditions of 40 kV and 40 mA. The functional groups of samples were characterized by Fourier transform infrared spectroscopy (FTIR, Nicolet 5700, Thermo Nicolet Corporation, WI, USA), and the wavelength measurement range was 400~4000 cm−1 by KBr compression method. The chemical components of Co/MSiO2 samples were performed by an X-ray photoelectron spectrometer (XPS, ESCALAB250xi, Thermo Scientific, MA, USA) with AlKα excitation. The transmission electron microscopy (TEM, JEM 2100F, JEOL, Tokyo, Japan) was used to analyse the crystallization of the Co/MSiO2 powders. The morphologies of surface and cross-sections for the membranes were observed by a scanning electron microscopy (SEM, JEOL JSM-6300, Hitachi, Tokyo, Japan) under 5 kV acceleration voltage. The BET surface area and pore volume of the samples was measured by N2 sorption/desorption isotherm with a specific surface area and pore-size analyzer (ASAP 2020, Micromeritics, GA, USA).
The schematic diagram of experimental devices used to test single gas-permeation measurement is shown in Figure 1. Before testing the experiments, pressure and temperature were kept at 0.35 MPa and 200 °C for 0.5 h until gas permeation achieved a stable state. The MSiO2 and Co/MSiO2 membranes were tested using high-purity H2, CO2, and N2, respectively. Additionally, the steam stability of MSiO2 and Co/MSiO2 membranes was examined by exposure to saturated steam at 200 °C for 10 days. After steam-stability test, the thermal regeneration of MSiO2 and Co/MSiO2 membranes was carried out at 350 °C with the aid of the equal calcination process as described above. The gas permeance was calculated as the usage of the outlet gas flow rate. The values of gas perm-selectivity were calculated from the ratio of individual gas-permeance values at the same transmembrane pressure difference and temperature.

3. Results

3.1. Phase Structure Analysis

The XRD patterns of the unsupported Co/MSiO2 materials with various nCo calcined at 400 °C under N2 atmosphere are provided in Figure 2. A distinct diffraction peak at about 2θ = 23° is assigned to the feature of amorphous SiO2 for all samples, and the peak intensity decreases with increasing nCo. It is probably because the introduction of cobalt atoms replaces the original silicon atoms and forms the Si–O–Co bonds, resulting ultimately in an increase in Si–O–Co bonds and a decrease in SiO2. When nCo is equal to 0.5, significant absorption peaks appeared at 2θ = 36.57, 42.49, 61.64, 73.86, and 77.74°, which corresponds to the plane reflections of (111), (200), (220), (311), and (222) of CoO crystalline phase (PDF No. 70-2855), respectively, which indicates part of the cobalt is dispersed on the surface of the material in the shape of CoO. There are no peaks of CoO and Co3O4 in unsupported Co/MSiO2 material with nCo = 0. However, the characteristic peak of CoO in the samples with nCo = 0.08 and 0.15 is not obviously observed, which may be due to the fact that the CoO amount is too low or the size of the formed CoO is too small to be detected [36,37]. According to the literature, the XRD patterns of Co-doped silica powders with a cobalt mole fraction of 33% sintered at 550 °C with an existing Co3O4 peak [38]. In this paper, when nCo is equal to 0.08–0.5, no characteristic peak of Co3O4 is observed in the samples calcined at 400 °C, which does not mean that Co3O4 is not present in the sample. This may be owing to the fact that the content of Co3O4 is small and cannot be detected by XRD [36].
The full-width at half maxima of the characteristic reflection with the highest intensity (200) was used to calculate the mean crystallite size with the aid of making use of the Scherrer equation [39]:
D = k λ β cos θ
where D is the size of CoO crystallites, k is the constant value of Scherrer (0.89), λ is the wavelength of X-ray source (0.154 nm), β is the full width at half maximum intensity, and θ is the Bragg angle. Hence, the mean size of CoO crystals in the samples with nCo = 0.5 is calculated as 3.8 nm.
In order to further determine the presence of Co species in the unsupported Co/MSiO2 materials with nCo = 0.08, 0.15, and 0.5 calcined at 400 °C under N2 atmosphere. We characterized the unsupported Co/MSiO2 material with nCo = 0.08 by XPS. The XPS spectra curves for samples are displayed in Figure 3. The peaks at 786.2 eV and 801.3 eV are assigned to the 2p3/2 and 2p1/2 peaks of Co3O4, respectively, and the peaks of 789.1 eV and 803.3 eV correspond to the 2p3/2 and 2p1/2 peaks of CoO, respectively [40]. Additionally, two shake-up satellite peaks can be seen at 792.8 eV and 808.3 eV, which are because of the multi-electron excitation of Co2+ [28]. However, the binding energies of Co3O4 and CoO in the SiO2 material are higher than that of the pure component, as previously reported [41]. This is resulted from the formation of the Si–O–Co bond between Co oxides and Si atoms due to the interaction of electrons. Hence, the XPS investigation clearly indicated that the doped Co component is present in the oxide rather than the metal. This result was confirmed by XRD analysis. For the samples with even higher Co content, such as unsupported Co/MSiO2 materials with nCo = 0.15 and 0. 5, the Co signal becomes very strong (especially the latter one), and it is obvious that the peaks of Co3O4 and CoO are found in the samples by spectrum analysis. Therefore, the results are not included for comparison.

3.2. FTIR Analysis

To further consider the influence of nCo on chemical structure for the unsupported Co/MSiO2 materials, the unsupported Co/MSiO2 materials with various nCo calcined in N2 atmosphere at 400 °C are characterized by FTIR spectra, which are shown in Figure 4. In Figure 4, the peak appearing at about 2978 cm−1 is attributed to the –CH3 groups for TEOS and MTES. The absorption peak located at 1640 cm−1 is associated with the stretching and bending vibration of –OH groups [42] from the absorbed water and ethanol as well as Si–OH. Undoubtedly, the existence of 1276 cm−1 band indicates the stretching vibration of the Si–CH3 groups. The band located at 770–800 cm−1 is accompanied by a shoulder, which is attributed to the asymmetric tensile vibration of the Si–O–Si bonds [43]. For Co/MSiO2 material with nCo = 0, the absorption peak at about 1055 cm−1 is also assigned to the vibration of Si–O–Si bonds. With the increasing nCo, the Si–O–Si bonds centered at 1055 cm−1 gradually shift to a lower value. The movement of Si–O–Si bonds indicates that Co enters the SiO2 lattice and that Si–O–Co bonds exist in the materials, which destroys the symmetry of SiO2 and causes the move of peak position. A similar phenomenon has been reported in other literature [44,45]. In addition, when nCo = 0.5, an additional peak is found located at 960 cm−1 and corresponding to the Si–O–Co vibration, suggesting that cobalt enters into the silica framework and forms the Si–O–Co bonds. However, Si–O–Co bonds in the samples with nCo = 0.08 and 0.15 are not obvious, which may be owing to the fact that the Si–O–Co bonds cannot be detected when the content of doped cobalt is small. Generally, the FTIR bands assigned to Co3O4 are located at 571 cm−1 and 664 cm−1 [46], but there is no obvious peak of Co3O4 in this figure, which does not imply that Co3O4 is not present in the samples. The reason may be that the content of Co3O4 is too small, and the peak is not revealed.

3.3. Pore-Structure Analysis

The physical characteristics of the prepared samples can be greatly influenced by their specific surface area and porous structure. The N2 adsorption-desorption isotherm curves of unsupported Co/MSiO2 materials with various nCo at 400 °C are displayed in Figure 5a. As shown in Figure 5a, the isotherms for the four samples all show a similar trend, which can be categorized as type I isotherm. In the range of low relative pressure P/P0 < 0.1, a substantial amount of adsorption indicates that there is a giant quantity of micropores in the materials. As the relative pressure increases, the isotherm gradually increases, which confirms the existence of a small amount of mesopores. The N2 adsorption of unsupported Co/MSiO2 materials increases initially with the increasing nCo, then begins to decrease as nCo > 0.08, which indicates the change of pore-volume variation trend. The distributions of pore size for all samples are depicted in Figure 5b. It is found that the unsupported Co/MSiO2 materials with nCo = 0.08–0.5 have a wider pore size distribution and a bigger mean pore size than the sample with nCo = 0. In addition, the pore diameters of all samples are mainly concentrated around 1.3 nm. The detailed information about the pore size and surface area for these four samples is provided in Table 1. It can be observed that, with the increases of nCo, the mean pore size increases, and the micropore volume decreases; the total pore volume and BET surface area increase until nCo = 0.08, after which they begin to decrease. This is because the added cobalt atoms exist in the form of Si–O–Co bonds in the SiO2 skeleton, and the atomic radius of the cobalt atoms is larger than that of the silicon atoms, which plays a role in expanding the pores [45]. So, with the increase of nCo, the particle size, mean pore size, total pore volume, and surface area, the distribution of pore size becomes wider and shifts gradually to the direction of the mesopores. However, as nCo > 0.08, in addition to the existence of cobalt in the skeleton in the amorphous form, there are also some cobalt oxides interspersed in the pores to block part of the pores, which leads to a decreasing in the pore volume and BET surface area. When nCo = 0.08, the unsupported Co/MSiO2 material is more favorable for achieving a higher total pore volume (0.424 cm3·g−1) and BET surface area (775.344 m2·g−1), with the minimum mean pore diameter (2.34 nm). Therefore, the unsupported Co/MSiO2 material with nCo = 0.08 is more favorable for gas separation.

3.4. TEM Analysis

To gain an insight into the Co species, the TEM analysis was performed to determine the presence of Co species in unsupported Co/MSiO2 materials with various nCo calcined in N2 atmosphere at 400 °C. Figure 6 demonstrates the transmission electron microscope images for unsupported Co/MSiO2 materials with nCo = 0, 0.08, and 0.15 calcined in N2 atmosphere at 400 °C. For these three samples, because of differences in electronic density, the darker-contrast particles can be attributed to cobalt oxide, while the lighter-contrast particles are attributed to silica carrier. In Figure 6a, amorphous silica can be observed on Co/MSiO2 material with nCo = 0. In Figure 6b,c, a small amount of the CoO crystals are uniformly dispersed on the surface of Co/MSiO2 materials with nCo = 0.08 and 0.15, and the particle size of the CoO crystals increases with the increasing nCo. Furthermore, the crystal size of CoO in Co/MSiO2 material with nCo = 0.08 is in the range of 1.3–2.2 nm, and that in the Co/MSiO2 material with nCo = 0.15 is in the range of 1.9–2.7 nm. Nevertheless, no Co3O4 particles were found on the outside surfaces of two samples, which does not imply that Co3O4 does not exist in the samples. This may be due to the tiny amount. For the unsupported Co/MSiO2 material with nCo = 0.5, it is clear that particles of Co3O4 and CoO are scattered on the surface of the sample by TEM analysis because of a higher Co content. Thus, the sample is not introduced here. It can be seen that the conclusions obtained above are consistent with the results from the XRD and XPS analysis.

3.5. Gas-Permeance Analysis

Based on the XRD, XPS, FTIR, N2 adsorption-desorption, and TEM results, the nCo shows obvious impact on the physical-chemical structures of unsupported Co/MSiO2 materials. However, compared with the unsupported Co/MSiO2 materials with nCo = 0.08 and 0.15, the pore volume and surface area for unsupported Co/MSiO2 materials with nCo = 0.5 are smaller, whereas the mean pore size is larger. As far as we know, the permeation and selectivity of gas-separation membrane are dependent on the pore structure and surface area. It could suggest that the supported Co/MSiO2 membrane with nCo = 0.5 is not suitable for gas-permeation experiments. Therefore, the supported Co/MSiO2 membrane with nCo = 0.5 is not considered here. The permeances for gases (H2, CO2, and N2) to the gas molecules’ kinetic diameters (dk) and H2 perm-selectivities of Co/MSiO2 membrane with nCo = 0, 0.08, and 0.15 at a pressure difference of 0.2 MPa and 200 °C are shown in Figure 7. The gas molecules’ kinetic diameters can be acquired from the report in [47]. The Co/MSiO2 membrane with nCo = 0 is also referred to as MSiO2 membrane. From Figure 7a, as nCo > 0.08, the permeances of Co/MSiO2 membrane to H2, CO2, and N2 appears to decrease. The H2 permeance of Co/MSiO2 membranes with nCo = 0, 0.08, and 0.15 are 9.07 × 10−6, 1.97 × 10−5, and 1.41 × 10−5 mol·m−2·Pa−1·s−1, respectively. Compared with the Co/MSiO2 membrane with nCo = 0, the H2, CO2, and N2 permeances of Co/MSiO2 membrane with nCo = 0.08 increase by 1.17, 0.70, and 0.83 times, respectively. It can be seen from the pore structure analysis that the total pore volume and average pore diameters for the silica membranes enlarge slightly with increasing nCo, giving an explanation for the increase in gas permeance. Furthermore, for the same membrane, the order for permeance of gas molecules is N2 < CO2 < H2. The gas permeation decreases as the dk increases, suggesting that all membranes show molecular sieve characteristics. These above results indicate that the porosity of the membrane shows a mean pore size of approximately 0.3 nm. It can be observed from Figure 7b that the perm-selectivities of H2/CO2 and H2/N2 for Co/MSiO2 membranes with various nCo are all significantly higher than the ideal perm-selectivities of Knudsen diffusion, which are 4.69 and 3.74, respectively. Compared with MSiO2 membrane, the H2/CO2 and H2/N2 perm-selectivities of Co/MSiO2 membrane with nCo = 0.08 increase by 27.66% and 18.53%, respectively. Therefore, the amplification of H2 perm-selectivities is not entirely ruled by the aid of molecular sieving; however, it may additionally be partly attributed to the improved adsorption of hydrogen by means of the Co/MSiO2 membrane matrix. The consequences of improved H2 perm-selectivities have additionally been mentioned for the Ni/SiO2 [20] and Pd/SiO2 [24] membranes, which have been ascribed to the enlarge affinity of H2 with the aid of the metallic particles. Moreover, when nCo = 0.08, the H2 perm-selectivities of the Co/MSiO2 membrane reach the maximum value. However, as the nCo continues to increase, the H2 perm-selectivities show a gradual decrease. The above consequences point out that the nCo performs an advantageous function in the impact of gas permeation for the membrane, but it does not mean that the higher the nCo, the better the gas-permeation effect. Thus, the Co/MSiO2 membrane with nCo = 0.08 has good gas permeability and selectivity, which is more appropriate for gas-permeation experiments.
Figure 8 displays the influence for pressure difference on the gases’ (H2, CO2, and N2) permeances of MSiO2 membrane at 200 °C; their pressure difference is generally unbiased, whilst that for H2 in Co/MSiO2 membrane with nCo = 0.08 increases significantly with growing pressure difference. It suggests that, due to the impact of the incorporated metal cobalt, the mechanism of H2 diffusion for Co/MSiO2 membrane is different from that for MSiO2 membrane. There are small mesopores on the membrane surface, which leads to the result that the obtained Co/MSiO2 membranes are accompanied by Knudsen diffusion. However, the doped metallic cobalt can improve the surface diffusion of H2 molecules in SiO2 membranes, and the growth of pressure is conducive to the adsorption of hydrogen. In addition, with the gradual increase of pressure, slight increases in the permeances of CO2 and N2 for Co/MSiO2 membrane are observed, which is due to the small influence of pressure on Knudsen diffusion, as previously reported in other literature [48,49]. Accordingly, the permeances of CO2 and N2 in Co/MSiO2 membrane are slightly dependent on pressure, whereas the H2 permeance increases within the pressure range due to the enhanced surface diffusion of hydrogen molecules by the cobalt particles.
The temperature dependence of the various gases’ (H2, CO2, and N2) permeances and H2/CO2 perm-selectivities in the MSiO2 membrane and Co/MSiO2 membrane with nCo = 0.08 at a pressure difference of 0.2 MPa are further investigated in the temperature range of 25–200 °C, which is depicted in Figure 9. In Figure 9a, with the continuous increase of temperature, the H2 permeance in MSiO2 and Co/MSiO2 membranes gradually increases, indicating that the permeation behavior of H2 in the two membranes mainly follows the activation-diffusion mechanism. In contrast, the permeances of CO2 and N2 are slightly decreased in a similar tendency to Knudsen diffusion in which molecules collide with pore walls more regularly than permeating molecules. In the case of activated diffusion, molecules permeate via micropores whilst being uncovered to repelling forces from the pore walls, and molecules with sufficient kinetic energy to conquer the repulsive force can permeate into the pores [38]. The decreasing permeances of CO2 and N2 are attributed to the violent movement of molecules and the increase of the mean free path when the temperature increases. As shown in Figure 9b, as temperature continues to increase in this range, the H2/CO2 and H2/N2 perm-selectivities for MSiO2 and Co/MSiO2 membranes all show a gradual, increasing trend. Compared with MSiO2 membrane, when operated at 25 °C, the permeance of H2, and the perm-selectivities of H2/CO2 and H2/N2 for Co/MSiO2 membrane increase by 121.53%, 22.76%, and 16.50%, respectively; on the other hand, when operated at 200 °C, those increase by 116.73%, 27.66%, and 18.53%, respectively. In addition, it can be found that, in the temperature range of 25–200 °C, the perm-selectivities of H2/CO2 and H2/N2 in both membranes are greater than the ideal perm-selectivities of Knudsen diffusion (4.69 and 3.74). The above results show that the Co/MSiO2 membrane has better perm-selectivity and permeance of H2 than those of MSiO2 membrane under same conditions.
The apparent activation energy (Ea) is an index of the probability of molecules passing through shrinkage, so lower activation energy is related to higher permeability. According to the Arrhenius equation [16,38], permeability F is a temperature-related parameter, which can be expressed by modified Fick’s law:
F = exp ( E a R T )
where F is the gas permeability, Ea is the apparent permeation-activation energy, F0 is a temperature-independent parameter, R is the constant of gas and T is the temperature of gas, and the unit is K. Equation (2) can be described in another form:
ln F = ln F 0 E a R T
In order to further study the diffusion phenomenon of gas molecules through MSiO2 and Co/MSiO2 membrane, the Arrhenius diagram is established, and the Ea values of gases’ (H2, CO2, and N2) permeations in MSiO2 and Co/MSiO2 film are calculated by using Arrhenius relationship between natural logarithm of permeation and reciprocal of temperature; the corresponding results are plotted in Figure 10. The Ea values of gases (H2, CO2, and N2) can be calculated from the Arrhenius formula for MSiO2 membrane and Co/MSiO2 membrane with nCo = 0.08 at 0.2 MPa, which are listed in Table 2. It can be observed from Table 2 that the Ea value of H2 is positive, while the Ea values of CO2 and N2 are negative, which are conclusions that resemble those of previous reports [50]. The positive or negative value of Ea is related to the activated transportation behavior. The activation energy indicates the repulsive energy of osmotic substance passing through the pore structure of membrane [51]. Gas transport in the microporous state is carried out by heat-activated surface-diffusion mechanism [27]. The negative values Ea of CO2 and N2 indicate that there is a percolation path in the membrane, which is ample enough to enable the diffusion of larger molecules. The Ea values of gases (H2, CO2, and N2) in the Co/MSiO2 membrane are less than those in the MSiO2 membrane, which indicates that the structure of the Co/MSiO2 membrane is more open than that of the MSiO2 membrane, and the above observations are very consistent with the results of N2 adsorption-desorption. This result also suggests that the doping of cobalt successfully reduces the densification of SiO2 network. The larger porosity of Co/MSiO2 membrane leads to the kinetic energy of gas molecules overcoming the membrane pore barrier since it is less than that of MSiO2 membrane. Therefore, the gases’ (H2, CO2, and N2) permeances of Co/MSiO2 membrane are greater than those of MSiO2 membrane.
Table 3 shows perm-selectivities of H2, permeances of H2, Ea values of H2, and pore diameters for various SiO2 membranes prepared by other researchers using sol-gel process. As can be viewed from Table 3, it is hard to enhance the perm-selectivity and permeance of gas for the SiO2 membranes at the same time. Generally, larger average pore diameter leads to higher permeance of H2, lower perm-selectivity of H2, and smaller Ea value of H2. Besides, the Ea value of H2 has a link with the interplay between the molecules of H2 and the pore walls of the membrane. Therefore, probably due to the fact that the average pore diameters of the prepared Co/MSiO2 membranes are larger than that of the membrane obtained by different researchers listed in Table 3, it leads to smaller Ea values of H2 and higher permeance of H2. Due to the drawbacks for the conditions of experiment and technology, the perm-selectivities of H2 for the prepared Co/MSiO2 membrane is insufficient to reach significant value.
The transport of H2, CO2, and N2 in MSiO2 membrane is controlled by molecular sieving, but Knudsen diffusion still exists due to the presence of small mesopores. However, thanks to the absorption for molecules of H2 by cobalt, the introduction of cobalt particles improves the surface diffusion of molecules of H2 in SiO2 membranes, which promotes the transmission of H2 and leads to the growth for permeance of H2 in the Co/MSiO2 membrane. Hence, we can find that, compared with CO2 and N2, the Co/MSiO2 membrane has the higher permeation rate to H2. Consequently, compared with MSiO2 membrane, the permeance of H2 and the perm-selectivities of H2/CO2 and H2/N2 in Co/MSiO2 membrane increase simultaneously. The possible change mechanism of MSiO2 and Co/MSiO2 membranes for separating H2/CO2 is schematically illustrated in Figure 11.
In order to look into the stability for MSiO2 membrane and Co/MSiO2 membrane with nCo = 0.08 under hydrothermal conditions, they were subjected to saturated steam at 200 °C for 10 days and then regenerated by calcination at 350 °C. Figure 12 compares the effects of hydrothermal conditions on the both membrane samples. The experimental data were obtained at a pressure difference of 0.2 MPa and 200 °C. After steam treatment, the permeances of H2, CO2, and N2 for MSiO2 and Co/MSiO2 membranes appear to decrease. Compared with the untreated fresh samples, the permeance of H2 for MSiO2 and Co/MSiO2 membranes after steam aging for 10 days decrease by 21.06% and 7.48%, respectively, and the perm-selectivities of H2/CO2 and H2/N2 for MSiO2 membrane decrease by 4.13% and 3.54%, respectively, whereas those of Co/MSiO2 membrane increase by 3.37% and 2.55%, respectively. After regeneration by calcination at 350 °C, the permeances of gases (H2, CO2, and N2), the perm-selectivities H2/CO2 and H2/N2 for two membranes show an upward trend. However, compared with those of the untreated fresh samples, the permeances of H2 for MSiO2 and Co/MSiO2 membranes after regeneration decrease by 11.25% and 4.15%, respectively, whereas the perm-selectivities of H2/CO2 and H2/N2 for MSiO2 membrane increase by 5.80% and 4.64%, respectively, and those for Co/MSiO2 membrane increase by 5.03% and 3.49%, respectively. These results mightily indicate that regeneration causes structural changes of SiO2 membrane. The reduction of permeation for two membranes indicates that pore shrinkage of the membrane occurs after regeneration by calcination at 350 °C. The obtained smaller pores result in a lower permeance and higher perm-selectivity. However, a smaller decrease in permeation of H2 for Co/MSiO2 membrane indicates that the diffusion of hydrogen through the surface of cobalt particles could dominate permeation of H2. Therefore, the above results indicate that the Co/MSiO2 membrane has better hydrothermal stability and reproducibility than MSiO2 membrane.

3.6. SEM Analysis

The SEM images of membrane surface and cross-sections for Co/MSiO2 membranes with nCo = 0 and 0.08 calcined at 400 °C under N2 atmosphere are shown in Figure 13. In Figure 13a,b, it can be observed that there are no visible cracks and pinholes on Co/MSiO2 membranes surface, indicating that all membranes are well coated. Moreover, the particles on the surfaces of Co/MSiO2 membranes with nCo = 0 are relatively uniform with the particle diameters in the range of 1.2–5.0 nm, while particle diameters of Co/MSiO2 membrane with nCo = 0.08 are in the range of 1.6–6.3 nm. The cross-section of membrane indicates a classic, uneven configuration, which is related to the morphology of the supported SiO2 membrane. In the cross-section of the SEM image, there is a clear boundary between the support layer and selective layer. The selective layer is found to be smaller for Co/MSiO2 membrane with nCo = 0, with a total thickness about 2.3 μm, whereas the wider selective layer with a total thickness of approximately 2.5 μm can be observed for Co/MSiO2 membrane with nCo = 0.08. In addition, the consequences of gas-permeation measurements show that a complete selective layer has been successfully loaded on the support.

4. Conclusions

In summary, Co/MSiO2 materials and membranes with various nCo were successfully synthesized under N2 atmosphere by sol-gel technique. The effect of nCo on the microstructures and perm-selectivities of H2 for Co/MSiO2 membranes were investigated extensively. The results indicate that the cobalt element in Co/MSiO2 material calcined at 400 °C mainly exists in the form of Si–O–Co bond, Co3O4, and CoO crystals. The nCo has little influence on the thermal stability of Si–CH3 groups of the methyl-modified silica materials. In addition, the introduction of metallic cobalt can enlarge the total pore volume and average pore diameter of the MSiO2 membranes. However, the nCo has a large impact on the gas separation of Co/MSiO2 membrane. When operated at a pressure difference of 0.2 MPa and 200 °C, the Co/MSiO2 membrane with nCo = 0.08 has better gas permeability and selectivity. Compared with MSiO2 membrane, the H2, CO2, and N2 permeances of Co/MSiO2 membrane with nCo = 0.08 increased by 1.17, 0.70, and 0.83 times, respectively, and the perm-selectivities of H2/CO2 and H2/N2 increased by 27.66% and 18.53%, respectively. The Ea values of gases (H2, CO2, and N2) in the Co/MSiO2 membrane are less than those in the MSiO2 membrane. After steam treatment, the H2 permeance for MSiO2 and Co/MSiO2 membranes decreased by 21.06% and 7.48%, respectively, and the perm-selectivities of H2/CO2 and H2/N2 for MSiO2 membrane decreased by 4.13% and 3.54%, respectively, whereas those of Co/MSiO2 membrane increased by 3.37% and 2.55%, respectively. It is observed that, after regeneration, the permeances of gases (H2, CO2, and N2), the perm-selectivities H2/CO2 and H2/N2 for two membranes show an upward trend. However, compared with the untreated fresh samples, the permeances of H2 for MSiO2 and Co/MSiO2 membranes decreased by 11.25% and 4.15%, respectively, whereas the perm-selectivities of H2/CO2 and H2/N2 for MSiO2 membrane increased by 5.80% and 4.64%, respectively, and those for Co/MSiO2 membrane increased by 5.03% and 3.49%, respectively. The Co/MSiO2 membrane has better hydrothermal stability and reproducibility than MSiO2 membrane. In the future, we will further study these separation properties of Co/MSiO2 membranes for mixed gas with water vapor and compare the similarities and differences between the separation of single gas and mixed gas.

Author Contributions

Conceptualization, L.W. and J.Y.; methodology, R.M.; formal analysis, Y.G.; writing—original draft preparation, L.W. and J.Y.; project administration, R.M. and H.H.; funding acquisition, L.W., Y.G., and J.Y. 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 [2021GY-147] and [2021JQ-688], the Scientific Research Project of Shaanxi Education Department, China [No. 19JC017], the Xi’an Municipal Science and Technology Project, China [No. 2020KJRC0025], and Graduate Scientific Innovation Fund for Xi’an Polytechnic University, China [chx2021043].

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. Wang, Y.; Wang, H.; Fu, W.; Yang, X.; Zhang, H.; Huang, Z.; Li, J. Recent Advancement in Heterostructured Catalysts for Hydrogen Evolution Reaction. J. Mater. Chem. A 2020, 8, 6926–6956. [Google Scholar] [CrossRef]
  2. Chen, W.-H.; Chen, C.-Y. Water gas shift reaction for hydrogen production and carbon dioxide capture: A review. Appl. Energy 2019, 258, 114078. [Google Scholar] [CrossRef]
  3. Chen, S.; Pei, C.; Gong, J. Insights into Interfacial Catalysis of Steam Reforming Reactions for Hydrogen Production. Energy Environ. Sci. 2019, 12, 3473–3495. [Google Scholar] [CrossRef]
  4. Ishaq, H.; Dincer, I. Development and multi-objective optimization of a newly proposed industrial heat recovery based cascaded hydrogen and ammonia synthesis system. Sci. Total Environ. 2020, 743, 140671. [Google Scholar] [CrossRef]
  5. Edwards, P.; Kuznetsov, V.; David, W.I. Hydrogen energy. Philos. Trans. R. Soc. A Math. Phys. Eng. Sci. 2007, 365, 1043–1056. [Google Scholar] [CrossRef]
  6. Wei, Y.; Zhang, H.; Lei, J.; Song, H.; Qi, H. Controlling pore structures of Pd-doped organosilica membranes by calcination atmosphere for gas separation. Chin. J. Chem. Eng. 2019, 27, 3036–3042. [Google Scholar] [CrossRef]
  7. Cao, M.; Zhao, L.; Xu, D.; Ciora, R.; Liu PK, T.; Manousiouthakis, V.I.; Tsotsis, T.T. A carbon molecular sieve membrane-based reactive separation process for pre-combustion CO2 capture. J. Membr. Sci. 2020, 605, 118028. [Google Scholar] [CrossRef]
  8. Tahmasbi, D.; Hossainpour, S.; Babaluo, A.A.; Rezakazemi, M.; Mousavi Nejad Souq, S.S.; Younas, M. Hydrogen separation from synthesis gas using silica membrane: CFD simulation. Int. J. Hydrogen Energy 2020, 45, 19381–19390. [Google Scholar] [CrossRef]
  9. Ameh, A.E.; Eze, C.P.; Antunes, E.; Cornelius, M.-L.U.; Musyoka, N.M.; Petrik, L.F. Stability of fly ash-based BEA-zeolite in hot liquid phase. Catal. Today 2019, 357, 416–424. [Google Scholar] [CrossRef]
  10. 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]
  11. Kim, S.-H.; Park, J.-H.; Hong, Y.; Lee, C.-Y. Removal of BTX using granular octyl-functionalized mesoporous silica nanoparticle. Int. Biodeterior. Biodegrad. 2014, 95, 219–224. [Google Scholar] [CrossRef]
  12. Araki, S.; Imasaka, S.; Tanaka, S.; Miyake, Y. Pervaporation of organic/water mixtures with hydrophobic silica membranes functionalized by phenyl groups. J. Membr. Sci. 2011, 380, 41–47. [Google Scholar] [CrossRef]
  13. Messaoud, S.B.; Takagaki, A.; Sugawara, T.; Kikuchi, R.; Oyama, S.T. Alkylamine-silica hybrid membranes for carbon dioxide/methane separation. J. Membr. Sci. 2015, 477, 161–171. [Google Scholar] [CrossRef]
  14. Kato, H.; Lundin, S.-T.B.; Ahn, S.-J.; Takagaki, A.; Kikuchi, R.; Oyama, S.T. Gas Separation Silica Membranes Prepared by Chemical Vapor Deposition of Methyl-Substituted Silanes. Membranes 2019, 9, 144. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  15. Han, Y.-J.; Ko, K.-J.; Choi, H.-K.; Moon, J.-H.; Lee, C.-H. Kinetic effects of methane on binary mixture separation on methyltriethoxysilane templated silica membranes. Sep. Purif. Technol. 2017, 182, 151–159. [Google Scholar] [CrossRef]
  16. 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]
  17. Ma, X.; Janowska, K.; Boffa, V.; Fabbri, D.; Magnacca, G.; Calza, P.; Yue, Y. Surfactant-Assisted Fabrication of Alumina-Doped Amorphous Silica Nanofiltration Membranes with Enhanced Water Purification Performances. Nanomaterials 2019, 9, 1368. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Lawal, S.; Kanezashi, M.; Nagasawa, H.; Tsuru, T. Development of an acetylacetonate-modified silica-zirconia composite membrane applicable to gas separation. J. Membr. Sci. 2020, 599, 117844. [Google Scholar] [CrossRef]
  19. Darmawan, A.; Karlina, L.; Astuti, Y.; Sriatun; Motuzas, J.; Wang, D.; da Costa, J.C.D. Structural evolution of nickel oxide silica sol-gel for the preparation of interlayer-free membranes. J. Non-Cryst. Solids 2016, 447, 9–15. [Google Scholar] [CrossRef]
  20. 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]
  21. Lee, J.; Ha, J.-H.; Song, I.-H.; Park, J.-W. Facile surface modification of ceramic membranes using binary TiO2/SiO2 for achieving fouling resistance and photocatalytic degradation. J. Sol. Gel. Sci. Technol. 2019, 91, 198–207. [Google Scholar] [CrossRef]
  22. Yang, X.; Sheridan, S.; Ding, L.; Wang, D.K.; Smart, S.; Diniz da Costa, J.C.; Liubinas, A.; Duke, M. Inter-layer free cobalt-doped silica membranes for pervaporation of ammonia solutions. J. Membr. Sci. 2018, 553, 111–116. [Google Scholar] [CrossRef]
  23. Pakdehi, S.G.; Rahimi, E.; Shafiei, K. Effect of Co-SiO2 mesoporous layer coating step on liquid fuel dimethyl amino ethyl azide (DMAZ) dehydration performance. Chem. Eng. Technol. 2019, 42, 996–1001. [Google Scholar] [CrossRef]
  24. Yang, J.; Fan, W.; Bell, C.-M. Effect of calcination atmosphere on microstructure and H2/CO2 separation of palladium-doped silica membranes. Sep. Purif. Technol. 2019, 210, 659–669. [Google Scholar] [CrossRef]
  25. 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]
  26. 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]
  27. Boffa, V.; Blank, D.H.A.; Elshof, J.E. Hydrothermal stability of microporous silica and niobia-silica membranes. J. Membr. Sci. 2008, 319, 256–263. [Google Scholar] [CrossRef]
  28. Ballinger, B.; Motuzas, J.; Smart, S.; da Costa, J.C.D. Palladium cobalt binary doping of molecular sieving silica membranes. J. Membr. Sci. 2014, 451, 185–191. [Google Scholar] [CrossRef]
  29. Darmawan, A.; Motuzas, J.; Smart, S.; Julbe, A.; da Costa, J.C.D. Binary iron cobalt oxide silica membrane for gas separation. J. Membr. Sci. 2015, 474, 32–38. [Google Scholar] [CrossRef] [Green Version]
  30. Díez, B.; Roldán, N.; Martín, A.; Sotto, A.; Perdigón-Melón, J.A.; Arsuaga, J.; Rosal, R. Fouling and biofouling resistance of metal-doped mesostructured silica/polyethersulfone ultrafiltration membranes. J. Membr. Sci. 2017, 526, 252–263. [Google Scholar] [CrossRef]
  31. Battersby, S.; Smart, S.; Ladewig, B.; Liu, S.; Duke, M.C.; Rudolph, V.; da Costa, J.C.D. Hydrothermal stability of cobalt silica membranes in a water gas shift membrane reactor. Sep. Purif. Technol. 2009, 66, 299–305. [Google Scholar] [CrossRef]
  32. Smart, S.; Vente, J.F.; da Costa, J.C.D. High temperature H2/CO2 separation using cobalt oxide silica membranes. Int. J. Hydrog. Energy 2012, 37, 12700–12707. [Google Scholar] [CrossRef] [Green Version]
  33. Liu, L.; Wang, D.K.; Martens, D.L.; Smart, S.; Strounina, E.; da Costa, J.C.D. Physicochemical characterisation and hydrothermal stability investigation of cobalt-incorporated silica xerogels. RSC Adv. 2014, 4, 18862–18870. [Google Scholar] [CrossRef]
  34. Esposito, S.; Turco, M.; Ramis, G.; Bagnasco, G.; Pernice, P.; Pagliuca, C.; Bevilacqua, M.; Aronne, A. Cobalt-silicon mixed oxide nanocomposites by modified sol-gel method. J. Solid State Chem. 2007, 180, 3341–3350. [Google Scholar] [CrossRef]
  35. Uhlmann, D.; Liu, S.; Ladewig, B.P.; da Costa, J.C.D. Cobalt-doped silica membranes for gas separation. J. Membr. Sci. 2009, 326, 316–321. [Google Scholar] [CrossRef]
  36. Ma, H.; Xu, J.; Chen, C.; Zhang, Q.; Ning, J.; Miao, H.; Zhou, L.; Li, X. Catalytic aerobic oxidation of ethylbenzene over Co/SBA-15. Catal. Lett. 2007, 113, 104–108. [Google Scholar] [CrossRef]
  37. Zhang, W.; Wang, B.; Luo, H.; Jin, F.; Ruan, T.; Wang, D. MoO2 nanobelts modified with an MOF-derived carbon layer for high performance lithium-ion battery anodes. J. Alloys Compd. 2019, 803, 664–670. [Google Scholar] [CrossRef]
  38. 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]
  39. Morad, I.; Liu, X.; Qiu, J. Crystallization-induced valence state change of Mn2+→Mn4+ in LiNaGe4O9 glass-ceramics. J. Am. Ceram. Soc. 2020, 103, 3051–3059. [Google Scholar] [CrossRef]
  40. Riva, R.; Miessner, H.; Vitali, R.; Del Piero, G. Metal–support interaction in Co/SiO2 and Co/TiO2. Appl. Catal. A Gen. 2000, 196, 111–123. [Google Scholar] [CrossRef]
  41. Lukashuk, L.; Yigit, N.; Li, H.; Bernardi, J.; Föttinger, K.; Rupprechter, G. Operando XAS and NAP-XPS investigation of CO oxidation on meso- and nanoscale CoO catalysts. Catal. Today 2019, 336, 139–147. [Google Scholar] [CrossRef]
  42. Okoye-Chine, C.G.; Mbuya, C.O.L.; Ntelane, T.S.; Moyo, M.; Hildebrandt, D. The effect of silanol groups on the metal-support interactions in silica-supported cobalt Fischer-Tropsch catalysts. A temperature programmed surface reaction. J. Catal. 2020, 381, 121–129. [Google Scholar] [CrossRef]
  43. Azmiyawati, C.; Niami, S.S.; Darmawan, A. Synthesis of silica gel from glass waste for adsorption of Mg2+, Cu2+, and Ag+ metal ions. IOP Conf. Ser. Mater. Sci. Eng. 2019, 509, 1–6. [Google Scholar] [CrossRef]
  44. Zhang, S.; Liu, X.; Shao, Z.; Wang, H.; Sun, Y. Direct CO2 hydrogenation to ethanol over supported Co2C catalysts: Studies on support effects and mechanism. J. Catal. 2020, 382, 86–96. [Google Scholar] [CrossRef]
  45. Li, S.; Wang, J.; Ye, Y.; Tang, Y.; Li, X.; Gu, F.; Li, L. Composite Si-O-Metal network catalysts with uneven electron distribution: Enhanced activity and electron transfer for catalytic ozonation of carbamazepine. Appl. Catal. B 2020, 263, 118311. [Google Scholar] [CrossRef]
  46. Khalil, A.; Ali, N.; Khan, A.; Asiri, A.M.; Kamal, T. Catalytic potential of cobalt oxide and agar nanocomposite hydrogel for the chemical reduction of organic pollutants. Int. J. Biol. Macromol. 2020, 164, 2922–2930. [Google Scholar] [CrossRef] [PubMed]
  47. Mukherjee, S.; Chen, S.; Bezrukov, A.A.; Mostrom, M.; Terskikh, V.V.; Franz, D.; Wang, S.-Q.; Kumar, A.; Chen, M.; Space, B.; et al. Ultramicropore engineering by dehydration to enable molecular sieving of H2 by calcium trimesate. Angew. Chem. Int. Ed. 2020, 59, 16188–16194. [Google Scholar] [CrossRef] [PubMed]
  48. He, X.; Hagg, M.-B. Optimization of Carbonization Process for Preparation of High Performance Hollow Fiber Carbon Membranes. Ind. Eng. Chem. Res. 2011, 50, 8065–8072. [Google Scholar] [CrossRef]
  49. De Vos, R.M.; Verweij, H. Improved performance of silica membranes for gas separation. J. Membr. Sci. 1998, 143, 37–51. [Google Scholar] [CrossRef]
  50. Liu, L.; Wang, D.K.; Martens, D.L.; Smart, S.; da Costa, J.C.D. Binary gas mixture and hydrothermal stability investigation of cobalt silica membranes. J. Membr. Sci. 2015, 493, 470–477. [Google Scholar] [CrossRef] [Green Version]
  51. Hacarlioglu, P.; Lee, D.; Gibbs, G.V.; Oyama, S.T. Activation energies for permeation of He and H2 through silica membranes: An ab initio calculation study. J. Membr. Sci. 2008, 313, 277–283. [Google Scholar] [CrossRef]
  52. 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]
  53. 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]
  54. Qureshi, H.F.; Besselink, R.; ten Elshof, J.E.; Nijmeijer, A.; Winnubst, L. Doped microporous hybrid silica membranes for gas separation. J. Sol Gel Sci. Technol. 2015, 75, 180–188. [Google Scholar] [CrossRef]
Figure 1. Schematic diagram of experimental devices for permeation measurement of single gas.
Figure 1. Schematic diagram of experimental devices for permeation measurement of single gas.
Materials 14 04188 g001
Figure 2. XRD patterns of unsupported Co/MSiO2 materials with various nCo calcined at 400 °C under N2 atmosphere.
Figure 2. XRD patterns of unsupported Co/MSiO2 materials with various nCo calcined at 400 °C under N2 atmosphere.
Materials 14 04188 g002
Figure 3. Co2p XPS spectra of unsupported Co/MSiO2 material with nCo = 0.08 calcined at 400 °C under N2 atmosphere.
Figure 3. Co2p XPS spectra of unsupported Co/MSiO2 material with nCo = 0.08 calcined at 400 °C under N2 atmosphere.
Materials 14 04188 g003
Figure 4. FTIR spectra curves of unsupported Co/MSiO2 materials with various nCo calcined in N2 atmosphere at 400 °C (the green arrow represents the direction in which the Si−O−Si bonds centered at 1055 cm−1 moves; the dotted rectangle emphasizes the peaks at 571 and 664 cm−1).
Figure 4. FTIR spectra curves of unsupported Co/MSiO2 materials with various nCo calcined in N2 atmosphere at 400 °C (the green arrow represents the direction in which the Si−O−Si bonds centered at 1055 cm−1 moves; the dotted rectangle emphasizes the peaks at 571 and 664 cm−1).
Materials 14 04188 g004
Figure 5. (a) N2 adsorption-desorption isotherms and (b) pore size distributions of unsupported Co/MSiO2 materials with various nCo at 400 °C.
Figure 5. (a) N2 adsorption-desorption isotherms and (b) pore size distributions of unsupported Co/MSiO2 materials with various nCo at 400 °C.
Materials 14 04188 g005
Figure 6. TEM images of unsupported Co/MSiO2 materials with nCo = (a) 0, (b) 0.08, and (c) 0.15 calcined in N2 atmosphere at 400 °C.
Figure 6. TEM images of unsupported Co/MSiO2 materials with nCo = (a) 0, (b) 0.08, and (c) 0.15 calcined in N2 atmosphere at 400 °C.
Materials 14 04188 g006
Figure 7. (a) Gases’ (H2, CO2, and N2) permeances versus gas molecules’ kinetic diameters (dk) and (b) H2 perm-selectivities of supported Co/MSiO2 membranes with various nCo at a pressure difference of 0.2 MPa and 200 °C.
Figure 7. (a) Gases’ (H2, CO2, and N2) permeances versus gas molecules’ kinetic diameters (dk) and (b) H2 perm-selectivities of supported Co/MSiO2 membranes with various nCo at a pressure difference of 0.2 MPa and 200 °C.
Materials 14 04188 g007
Figure 8. Influence of pressure difference on the gases’ (H2, CO2, and N2) permeances for supported (a) MSiO2 membrane and (b) supported Co/MSiO2 membrane with nCo = 0.08 at 200 °C.
Figure 8. Influence of pressure difference on the gases’ (H2, CO2, and N2) permeances for supported (a) MSiO2 membrane and (b) supported Co/MSiO2 membrane with nCo = 0.08 at 200 °C.
Materials 14 04188 g008
Figure 9. Influence of temperature on the (a) gases’ (H2, CO2, and N2) permeances and (b) H2 perm-selectivities for supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at 0.2 MPa.
Figure 9. Influence of temperature on the (a) gases’ (H2, CO2, and N2) permeances and (b) H2 perm-selectivities for supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at 0.2 MPa.
Materials 14 04188 g009
Figure 10. Arrhenius plots of gases’ (H2, CO2, and N2) permeances in supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at 0.2 MPa.
Figure 10. Arrhenius plots of gases’ (H2, CO2, and N2) permeances in supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at 0.2 MPa.
Materials 14 04188 g010
Figure 11. A possible schematic diagram for the change mechanism of supported MSiO2 and Co/MSiO2 membranes for separating H2/CO2.
Figure 11. A possible schematic diagram for the change mechanism of supported MSiO2 and Co/MSiO2 membranes for separating H2/CO2.
Materials 14 04188 g011
Figure 12. Effect of hydrothermal conditions on the (a) permeance of H2 and (b) perm-selectivities of H2 for supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at a pressure difference of 0.2 MPa and 200 °C.
Figure 12. Effect of hydrothermal conditions on the (a) permeance of H2 and (b) perm-selectivities of H2 for supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at a pressure difference of 0.2 MPa and 200 °C.
Materials 14 04188 g012
Figure 13. SEM images of membrane surface (top images) and cross-sections (bottom images) for supported Co/MSiO2 membranes with nCo = (a) 0 and (b) 0.08 calcined at 400 °C under N2 atmosphere.
Figure 13. SEM images of membrane surface (top images) and cross-sections (bottom images) for supported Co/MSiO2 membranes with nCo = (a) 0 and (b) 0.08 calcined at 400 °C under N2 atmosphere.
Materials 14 04188 g013
Table 1. Parameters of pore structure for unsupported Co/MSiO2 materials with various nCo after calcination at 400 °C.
Table 1. Parameters of pore structure for unsupported Co/MSiO2 materials with various nCo after calcination at 400 °C.
Membrane MaterialsBET Surface
Area (m2/g)
Total Pore
Volume (cm3/g)
Micropore
Volume (cm3/g)
VM/VtMean Pore
Size (nm)
nCo = 0389.380.2300.1500.6521.75
nCo = 0.08775.340.4240.0450.1062.34
nCo = 0.15644.100.4020.0350.0872.73
nCo = 0.5494.540.3990.0340.0853.14
Table 2. Ea values of gases (H2, CO2, and N2) calculated from Arrhenius formula for supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at 0.2 MPa.
Table 2. Ea values of gases (H2, CO2, and N2) calculated from Arrhenius formula for supported MSiO2 membrane and supported Co/MSiO2 membrane with nCo = 0.08 at 0.2 MPa.
GasesEa/KJ·mol−1
MSiO2Co/MSiO2
H22.141.98
CO2−1.34−1.72
N2−1.21−1.37
Table 3. Perm-selectivities of H2, permeances of H2, Ea value of H2, and pore diameter for various SiO2 membranes prepared by other researchers using sol-gel process.
Table 3. Perm-selectivities of H2, permeances of H2, Ea value of H2, and pore diameter for various SiO2 membranes prepared by other researchers using sol-gel process.
Membrane TypeTemperature and PressurePermeance of H2 (mol·m−2·Pa−1·s−1)Perm-Selectivities of H2Calcination AtmosphereEa Value of H2 (KJ·mol−1)Pore Diameter
(nm)
H2/CO2H2/N2
Si(400) [49]200 °C, 1 bar1.74 × 10−67.564.4Air80.38–0.55
SiO2 [52]200 °C, 2 bar4.62 × 10−73.710.5N20.30–0.54
SiO2–ZrO2 [53]500 °C, 100 KPa2 × 10−615190Air13
Pd/SiO2 [25]200 °C, 0.3 MPa7.26 × 10−74.314H2, N2~0.57
Nb/SiO2 [54]200 °C, 2 bar5.03 × 10−73.56.5N2~0.55
Co/SiO2 [50]200 °C, 500 KPa5 × 10−831.6Air13.8
Co/SiO2 *200 °C, 0.2 MPa1.97 × 10−510.4813.08N21.980.3–2.3
* In this work.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

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. Materials 2021, 14, 4188. https://doi.org/10.3390/ma14154188

AMA Style

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. Materials. 2021; 14(15):4188. https://doi.org/10.3390/ma14154188

Chicago/Turabian Style

Wang, Lunwei, Jing Yang, Ruihua Mu, Yingming Guo, and Haiyun Hou. 2021. "Sol-Gel Processed Cobalt-Doped Methylated Silica Membranes Calcined under N2 Atmosphere: Microstructure and Hydrogen Perm-Selectivity" Materials 14, no. 15: 4188. https://doi.org/10.3390/ma14154188

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