Next Article in Journal
Printable Conductive Hydrogels and Elastomers for Biomedical Application
Previous Article in Journal
Biological Models for Evaluating Hydrogel-Based Formulations in Wound Healing
Previous Article in Special Issue
The Behaviour of Contaflex Soft Contact Lens Material During Hydration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Microstructure and Photocatalytic Performance of BaTi5O11 Nanocrystals Synthesized via Sol-Gel Method Mediated by Organic Solvents

1
School of Materials Science and Engineering, Wuhan University of Technology, Wuhan 430070, China
2
Ji Hua Laboratory, Foshan 528200, China
*
Author to whom correspondence should be addressed.
Gels 2025, 11(9), 706; https://doi.org/10.3390/gels11090706
Submission received: 25 July 2025 / Revised: 29 August 2025 / Accepted: 1 September 2025 / Published: 3 September 2025
(This article belongs to the Special Issue Innovative Gels: Structure, Properties, and Emerging Applications)

Abstract

BaTi5O11 nanocrystals were synthesized via a sol–gel method employing different organic solvents. The influence of solvent choice on microstructure and photocatalytic performance was investigated through methylene blue (MB) degradation under UV light irradiation. The monoclinic BaTi5O11 nanocrystals were successfully synthesized, where solvent selection significantly affected their grain size and Brunauer–Emmett–Teller (BET) surface area. The BaTi5O11 nanocrystals synthesized using polyethylene glycol-200 (PEG-200) exhibited the highest BET surface area (9.78 m2/g) and smallest average pore size (17.8 nm). The BaTi5O11 nanocrystals also displayed a larger optical bandgap (3.61 eV), attributed to pronounced quantum confinement and surface effects. Consequently, the PEG-200-derived BaTi5O11 photocatalyst achieved complete MB degradation within 30 min under UV light irradiation. This enhanced performance was attributed to the high BET surface area providing abundant active sites. Furthermore, the BaTi5O11 nanocrystal photocatalyst maintained excellent reusability and stability over four consecutive cycles.

1. Introduction

With the acceleration of industrialization, industrial wastewater containing a large number of organic pollutants, such as dyes, fertilizers, pesticide or surfactant molecules, is continuously discharged into water bodies [1,2,3,4]. Those organic matters are difficult for microorganisms to directly degrade in nature. Methylene blue (MB), which can be found in industrial wastewater, is a carcinogenic pollutant due to its hazardous impacts on humans, even in low concentration [5]. It is recognized that perovskite nanocrystals are efficient in the photocatalytic degradation of MB. Photocatalysis is considered as a prospective technology for the degradation of MB due to high efficiency, no secondary pollution, and environmental compatibility [6]. However, most outstanding perovskite photocatalysts, such as PrZrxTi1−xO3, are hazardous to both humans and environment due to the Pb element. As typical lead-free perovskite photocatalysts, the compounds in the TiO2-rich BaO-TiO2 system, such as BaTi4O9 and BaTi5O11, have been extensively studied [7,8,9].
The BaTi5O11 compound has attracted extensive research interest due to its exceptional permittivity and low dielectric loss, which is widely applied in microwave communications [10,11]. Many researchers have attempted to synthesize the BaTi5O11 compound in a variety of ways. Recent advancements in nanofabrication techniques have also expanded its application frontiers, with hydrothermal synthesis, co-precipitation, and sol–gel methods enabling precise control over crystallinity and morphology of BaTi5O11 nanostructures [12,13,14,15,16,17]. Notably, emerging studies revealed that nanostructured BaTi5O11 exhibited remarkable photocatalytic activity, attributed to its unique crystal architecture [16]. The distorted TiO6 octahedra in its lattice generated spontaneous polarization that established built-in electric fields, while nanoscale dimensions shortened charge migration paths—a synergistic effect that significantly enhanced photogenerated carrier separation. Zhang et al. demonstrated 82% MB degradation using hydrothermal-synthesized BaTi5O11 nanocrystals within 30 min of UV light irradiation [15], while Yang et al. achieved 96.5% (60 min UV light irradiation) degradation efficiency through sol–gel-derived BaTi5O11 nanocrystals by tailoring their microstructures [17].
Appropriate synthetic approaches are important to the BaTi5O11 nanomaterials with designed microstructure and properties. Among various synthesis approaches, the sol–gel method presents distinct advantages for photocatalytic material engineering, including stoichiometric precision, low-temperature processing, and cost-effectiveness [17,18,19,20,21]. During sol–gel processing, parameters, such as sintering temperature, sintering time and solvent choice, decisively govern the crystallite size and morphology of BaTi5O11 nanocrystals. Solvent polarity and coordination ability fundamentally determine hydrolysis–condensation rates, thereby governing pore structure development and surface reactivity [22]. Although previous studies have established sintering temperature and time as key factors affecting crystallite size of BaTi5O11 nanocrystals [13], systematic investigations into solvent-mediated morphological control of BaTi5O11 nanocrystals are conspicuously lacking. Mehdi R. et al. [23] prepared ZnO nanocrystals via sole-gel methods with ethanol, ethylene glycol, 1,4-butanediol and PEG-600. The obtained results showed that higher viscosities and lower vapor pressures of solvents allow for more rapid growth of the ZnO nuclei, resulting in larger particles. Wang Y. et al. [24] prepared sol–gel-derived pure BaTiO3 particles using acetylacetone. The results showed that the product maintained stable degradation performance in environments containing inorganic anions, with a high k value of 5.541 × 10−2 min−1 degrading RhB. Mageste F. et al. [25] prepared SiO2 via the sol–gel method using methanol, ethanol, isopropanol and tert-butanol as solvents. The results showed that higher molecular weight led to larger pore size and higher specific surface area of the product, which was due to the different rates of hydrolysis and condensation reactions. In addition, the lower polarity of the solvent and the hydrogen bond combination led to larger particle size of the product, which resulted from slower nucleation. The dielectric constants of solvents also have a remarkable influence on the crystalline phase [26] solvent with high boiling point suppresses the hydrolysis of (CH3CH2O)4Ti, minimizes impurity formation and enhances the crystallinity of the material, resulting in a more homogeneous structure [27].
At present, there is no systematic study on the mechanisms of solvents to the synthesis of BaTi5O11 nanomaterials. Solvents not only dissolve reactants, but also have important effects on nucleation, nuclear growth, the grain production process, morphology control, etc. Choosing the appropriate solvent can optimize the reaction conditions and improve the quality and performance of the product.
In this study, BaTi5O11 nanocrystals were synthesized by the sol–gel method, and the effects of different solvents, including methanol, ethanol, 2-methoxyethanol, acetylacetone, ethylene glycol (EG) and polyethylene glycol-200 (PEG-200), on microstructure and photocatalytic MB degradation were systematically investigated. Also, a full understanding of the types of solvents and various parameters on the mechanism of regulating the synthesis process of nanomaterials, micromorphology and photocatalytic performance will provide guidance for the design of high-performance photocatalytic materials, even for other nanomaterials.

2. Results and Discussions

Figure 1 shows the XRD results of BaTi5O11 samples synthesized using different solvents. All XRD patterns display pure crystalline phases without distinct impurity phases, and were indexed to the standard monoclinic BaTi5O11 phase (JCPDS No. 35-0805, Joint Committee on Powder Diffraction Standards). The diffraction peaks of all BaTi5O11 samples closely matched the monoclinic BaTi5O11 phase. The XRD patterns of BaTi5O11 samples prepared with different solvents were remarkably similar, indicating that all chose solvents facilitated the formation of the monoclinic BaTi5O11 phase.
The morphologies of BaTi5O11 samples synthesized using different solvents are shown in Figure 2. The solvents with low boiling points and high volatility (methanol and ethanol) evaporated rapidly, resulting in larger BaTi5O11 grain sizes and significant agglomeration. In contrast, the solvents with higher boiling points, lower volatility and slower evaporation rates (2-methoxyethanol and acetylacetone) helped preserve the porous microstructure of the Ba–Ti gels, yielding smaller BaTi5O11 grains. Meanwhile, the solvents exhibiting even higher boiling points and substantially greater viscosity (EG and PEG-200) further slowed evaporation, which promoted nucleation over crystal growth, effectively minimizing particle agglomeration. These results demonstrated that solvents’ properties (including boiling point, viscosity, surface tension, polarity and volatility) fundamentally governed the morphology, grain size, agglomeration and specific surface area of sol–gel-derived BaTi5O11 nanocrystals [28].
To further characterize the microstructural difference in the BaTi5O11 nanocrystals synthesized with different solvents, their pore structure and specific surface area were investigated. The BaTi5O11 nanocrystals adsorption behavior by BET method is physical adsorption, which means the adsorbed gas molecules bind to the nanocrystal with a weak van der Waals force. Figure 3a presents their N2 gas adsorption–desorption isotherms. Following the IUPAC (International Union of Pure and Applied Chemistry) classification, all physical absorption–desorption isotherms exhibited type IV behavior [29], with adsorption capacity increasing relative to pressure. The presence of hysteresis loops and absence of distinct adsorption plateaus indicated disordered mesoporous microstructures in all BaTi5O11 nanocrystal samples. Figure 3b displays their Brunauer–Emmett–Teller (BET) surface areas and average pore sizes. As the solvents changed from methanol, to ethanol, 2-methoxyethanol, acetylacetone, EG and PEG-200, the BET surface area increased while the average pore size decreased. The methanol-synthesized BaTi5O11 nanocrystals showed the smallest BET surface area (3.69 m2/g) and largest average pore size (43.8 nm), whereas PEG-200 yielded the highest BET surface area (9.78 m2/g) and smallest average pore size (17.8 nm). Table 1 summarizes the relevant solvent physicochemical properties. The boiling point, viscosity, volatility and surface tension of the solvents might be the primary factors influencing the microstructure of BaTi5O11 nanocrystals. This correlation was further evidenced by pore-size distributions (Figure 3c). Methanol, ethanol, 2-methoxyethanol and acetylacetone produced pores mainly between 10 and 50 nm, while PEG-200 generated significantly smaller pores of about 10 nm. However, BaTi5O11 nanocrystals prepared using EG solvent exhibited a broad pore size distribution, ranging from several nanometers to nearly 70 nm. The sol–gel method provided a relatively stable environment at a middle temperature. The rate and degree of hydrolysis and polycondensation reactions are easily affected by polarity, viscosity, boiling temperature and density. PEG-200 solvent, exhibiting even higher boiling points and substantially greater viscosity, further slowed evaporation, promoted nucleation over crystal growth, effectively minimizing particle agglomeration, which results in minimizing particle agglomeration.
Figure 4 presents the XPS spectra of BaTi5O11 nanocrystals synthesized using PEG-200 as the solvent. This spectroscopic tool is used for further understanding the chemical state of elements at the surface of BaTi5O11. The peak positions were calibrated by referencing the C 1s peak at 284.8 eV. The survey spectrum confirmed the presence of Ba, Ti and O elements, as shown in Figure 4a. The high-resolution spectra and their Gaussian fittings for Ba 3d, Ti 2p and O 1s are displayed in Figure 4b–d. The Ba 3d peaks resolved into Ba 3d5/2 (779.6 eV) and Ba 3d3/2 (794.6 eV) components. The Ti 2p peaks exhibited Ti 2p3/2 (458.3 eV) and Ti 2p1/2 (464.1 eV) peaks, consistent with Ti4+. Deconvolution of the O 1s spectrum revealed lattice oxygen of Ba–O (529.7 eV), lattice oxygen of Ti–O (531.0 eV), and a peak at 532.5 eV, attributed to surface OH groups from adsorbed water molecules on the surface of the prepared BaTi5O11 nanocrystals [24].
Figure 5 presents the typical TEM images of BaTi5O11 nanocrystals synthesized using PEG-200 as the solvent. As shown in Figure 5a, the granular microstructure was consistent with the SEM results (Figure 2e). To assess the crystallinity of BaTi5O11 nanocrystals, the high-resolution TEM image of the nanocrystals was characterized, as shown in Figure 5b. The measured lattice spacings of 0.3341 nm and 0.3679 nm corresponded well to the (220) (0.3338 nm) and (002) (0.3727 nm) planes of the monoclinic BaTi5O11 phase, respectively. Furthermore, the measured interplanar angle of 81.96o between the (220) and (002) planes was in close agreement with the theoretical value of 82.67° for the monoclinic BaTi5O11 phase (JCPDS No. 35-0805). These findings confirmed that highly crystalline BaTi5O11 nanocrystals with monoclinic crystal structure were synthesized by the sol–gel method.
The optical properties of BaTi5O11 nanocrystals synthesized using different solvents were characterized using the UV–vis diffuse reflectance spectroscopy in the wavelength range of 200–600 nm. As shown in Figure 6a, the BaTi5O11 nanocrystals showed strong UV absorption. The optical band gap (Eg) was calculated from the reflectance data using the Kubelka–Munk model of Equation (1).
(αhν)n = A (Eg)
where α is the absorption coefficient, is the photon energy and A is a constant. As the BaTi5O11 compound is a direct bandgap semiconductor [12], n is equal to two. Figure 6b shows the corresponding Tauc plots of (αhv)2 versus hv. The optical Eg values, determined from these plots, were 3.45 eV for methanol, 3.44 eV for ethanol, 3.44 eV for 2-methoxyethanol, 3.45 eV for acetylacetone, 3.59 eV for EG, and 3.61 eV for PEG-200. The optical Eg of semiconductor nanomaterials is tunable, primarily governed by quantum confinement effects and surface effects. Notably, BaTi5O11 nanocrystals synthesized using EG and PEG-200 exhibited significantly larger Eg values (3.59 eV and 3.61 eV, respectively) compared to those prepared with other solvents (ranging from 3.44 to 3.45 eV). When combined with the FESEM observations (Figure 2) and specific surface area data (Figure 3), the large Eg values observed for the BaTi5O11 nanocrystals using EG and PEG-200 were attributed to pronounced quantum confinement and surface effects.
The photocatalytic performance of BaTi5O11 nanocrystals was evaluated through the photodegradation of MB under UV light irradiation. The MB concentration was determined from the UV–vis absorption intensity at its characteristic wavelength of 666 nm, which exhibited a linear relationship with concentration. Figure 7a presents the time-dependent UV–vis absorption spectra (450–800 nm) of MB solutions during degradation catalyzed by BaTi5O11 nanocrystals synthesized using PEG-200. The characteristic absorption peak at 666 nm decreased rapidly with increasing irradiation time, indicating efficient MB degradation. Figure 7b plots the normalized MB concentration (C/C0, where C0 and C represent the initial concentration and concentration at time of t, respectively) versus irradiation time for BaTi5O11 nanocrystals synthesized with different solvents. Prior to irradiation, the MB solution containing the BaTi5O11 nanocrystal photocatalyst was stirred in darkness for 30 min to establish adsorption–desorption equilibrium. The control experiments showed that MB was stable under UV light irradiation, with only 5.8% degradation occurring after 30 min in the absence of the catalyst. In contrast, the addition of BaTi5O11 nanocrystal photocatalysts led to significant MB degradation. After 30 min of UV light irradiation, the degradation efficiencies for BaTi5O11 nanocrystal photocatalysts synthesized with methanol, ethanol, 2-methoxyethanol, acetylacetone, EG and PEG-200 were 75.9%, 82.2%, 75.1%, 88.5%, 69.7% and 100%, respectively. Notably, the PEG-200-synthesized BaTi5O11 nanocrystal photocatalyst achieved complete degradation within 30 min. In our previous reports [15,16,17], the hydrothermally synthesized BaTi5O11 nanocrystals achieved 82.27% MB degradation after 30 min UV light irradiation, while the sol–gel-derived BaTi5O11 nanocrystals reached 96.5% MB degradation after 60 min UV light irradiation. The superior efficiency of the PEG-200-synthesized BaTi5O11 nanocrystals highlighted the critical influence of the solvent. The photocatalytic efficiency depends on multiple factors (catalyst properties, reaction conditions and pollutant characteristics) [30]. This study focused on the intrinsic properties of BaTi5O11 nanocrystal photocatalysts under fixed reaction conditions and pollutant parameters. The previous results (Figure 3b and Figure 6b) indicated that the solvent significantly affected the specific surface area and Eg of the BaTi5O11 nanocrystals. The PEG-200-synthesized BaTi5O11 nanocrystals exhibited the highest BET surface area, providing more active sites. However, it also showed a widened Eg (Figure 6b) due to the quantum confinement effects, potentially hindering light absorption. The experimental results demonstrated that the beneficial effect of the increased BET surface area dominated, leading to enhanced photocatalytic efficiency. Interestingly, the EG-synthesized BaTi5O11 nanocrystals also possessed a relatively large BET surface area but exhibited the lowest photocatalytic degradation efficiency (69.7%). This discrepancy was primarily attributed to differences in the solvent templating effect [23]. PEG-200, with its strong steric hindrance and templating capability, promoted the formation of a uniform, open porous microstructure. This microstructure enhanced not only the BET surface area but also dye molecule adsorption and diffusion kinetics, facilitating photocatalysis. Conversely, EG provided a weaker templating effect, often resulting in denser or agglomerated microstructures. Consequently, despite a nominally high BET surface area, insufficient exposure of active sites and restricted mass transfer in the EG-synthesized BaTi5O11 nanocrystals led to diminished photocatalytic performance.
To further study the photocatalytic kinetics of BaTi5O11 nanocrystals, their degradation reaction constant was calculated using Equation (2)
ln(C0/C) = kt
where t represents the irradiation time, and k is the pseudo-first-order kinetic reaction rate constant. The k value serves as a direct indicator of photocatalytic performance, and the larger k value corresponds to better photocatalytic performance. Figure 7c presents degradation kinetics of the BaTi5O11 nanocrystal photocatalysts. The BaTi5O11 nanocrystal photocatalyst synthesized using PEG-200 solvent exhibited the highest k value of 0.28 min−1, while the BaTi5O11 nanocrystal photocatalyst prepared using EG solvent showed the lowest k value of 3.98 × 10−2 min−1. The coefficient of determination R2 of degradation reaction for samples were 0.994 (PEG-200), 0.976 (acetylacetone), 0.966 (ethanol), 0.958 (2-methoxyethanol), 0.958 (methanol), and 0.977 (EG), respectively. The photocatalytic cycling performance of the PEG-200-derived BaTi5O11 nanocrystal photocatalyst is shown in Figure 7d. It retained the high MB degradation efficiency of 99.7% even after four consecutive cycles, confirming the excellent reusability and stability of the BaTi5O11 nanocrystal photocatalyst synthesized using PEG-200 solvent. Based on the aforementioned results, a possible explanation for the superiority of the degradation performance of BaTi5O11 over MB synthesized using PEG-200 as solvent is proposed. The sol–gel method provided a relatively stable environment at a middle temperature. The rate and degree of hydrolysis and polycondensation reactions are easily affected by polarity, viscosity, boiling temperature and density. PEG-200 solvent, exhibiting even higher boiling points and substantially greater viscosity, further slowed evaporation, promoted nucleation over crystal growth, effectively minimizing particle agglomeration, which results in minimizing particle agglomeration. As a result, BaTi5O11 nanocrystals synthesized using PEG-200 display the highest BET surface area and smaller pores, providing more active sites to improve degradation performance. Finally, the synthesis conditions, physical characteristics, and degradation performance of BaTi5O11 synthesized using different solvents are summarized in Table 1.

3. Conclusions

The monoclinic BaTi5O11 nanocrystals were successfully synthesized by the sol–gel method employing different organic solvents, and their photocatalytic performance was investigated by the degradation of MB under UV light irradiation. The choice of solvent significantly influenced the grain size and BET surface area of BaTi5O11 nanocrystals. Specifically, BaTi5O11 nanocrystals synthesized using PEG-200 exhibited the high BET surface area (9.78 m2/g) and small average pore size (17.8 nm). The BaTi5O11 nanocrystals also demonstrated a larger optical Eg value (3.61 eV), attributed to pronounced quantum confinement and surface effects. Consequently, the PEG-200-synthesized BaTi5O11 nanocrystal photocatalyst achieved complete degradation of MB under 30 min UV light irradiation and exhibited the highest k value of 0.28 min−1. This enhanced photocatalytic performance was ascribed to the high BET surface area, which provided abundant active sites. Furthermore, the BaTi5O11 nanocrystal photocatalyst maintained excellent reusability and stability over four consecutive cycles. This study presents a simple sol–gel process for fabricating efficient BaTi5O11 nanocrystal photocatalysts, offering insights for wastewater treatment and environmental purification. Moreover, it provides fundamental guidance for understanding the role of solvents in constructing high-efficiency BaTi5O11 photocatalysts.

4. Materials and Methods

4.1. BaTi5O11 Nanocrystal Preparation

All reagents were of analytical grade and used without further purification. Barium acetate (Ba(CH3COO)2) and tetrabutyl titanate (Ti(OC4H9)4) (Ba:Ti molar ratio = 1:5) were employed as Ba2+ and Ti4+ sources, respectively. The desired amount of Ba(CH3COO)2 was dissolved in acetic acid (20 mL), while Ti(OC4H9)4 was dissolved in the chosen solvent (80 mL, methanol, ethanol, 2-methoxyethanol, acetylacetone, EG or PEG-200). The two solutions were mixed together to form transparent Ba–Ti precursors (0.1 mol/L). Figure 1 shows the results of differential scanning calorimetry and thermogravimetric analysis (DSC-TG) for the Ba–Ti precursors with different solvents. The temperature increased from room temperature to 750 °C in N2 atmosphere with a heat rate of 10 °C/min. The Ba–Ti precursor prepared with methanol as the solvent exhibited an endothermic peak at a relatively low heating temperature (about 50 °C), which was attributed to the lowest boiling point and high volatility. As the boiling points of the solvents increased, the temperatures corresponding to their endothermic peaks increased. Based on Figure 8a, the baking temperatures for Ba-Ti precursors could be identified as 150 °C for those prepared with methanol, ethanol, 2-methoxyethanol, and acetylacetone, 220 °C for those prepared with EG, and 250 °C for those prepared with PEG-200. As shown in Figure 8b, the pronounced weight loss could be observed when the heating temperature approached the boiling point of each respective solvent. There was a significant difference between the weight loss of material obtained using different solvents, which was consistent with the viscosity. This phenomenon was explained by the fact that the samples synthesized using higher viscosity solvent contain more solvent in the Ba–Ti precursors.
However, the Ba-Ti precursor prepared with PEG-200 exhibited nearly linear weight loss between 100 and 380 °C. It was mainly attributed to the high viscosity of PEG-200, which resulted in slow solvent evaporation. As the temperature continued to increase, the distinct endothermic peaks appeared around 350 °C, corresponding to the pyrolysis of organic components and the formation of amorphous Ba–Ti–O powder. The TG curve also indicated a weight loss in this temperature region; consequently, 350 °C was selected as the thermal decomposition temperature. A pronounced exothermic peak near 700 °C signified the onset of BaTi5O11 crystallization, and 750 °C was chosen as the sintering temperature for the BaTi5O11 phase.
The BaTi5O11 nanocrystals were synthesized by the sol–gel method under flowing O2 in a quartz-tube furnace (Hangzhou Zhuochi Co., Hangzhou, China) according to the three-step heat-treatment process. In the heat-treatment process, the increase rate of temperature was 10 °C/min. These precursors were firstly baked at different temperatures (150 °C for methanol, ethanol, 2-methoxyethanol and acetylacetone; 220 °C for EG; 250 °C for PEG-200) for 180 min to evaporate the solvents, then heated at 350 °C for 180 min to obtain amorphous Ba–Ti–O powder, and finally sintered at 750 °C for 180 min to form BaTi5O11 nanocrystals. The key physicochemical properties of the solvents are listed in Table 2.

4.2. Material Characterization

Thermal analysis of the Ba–Ti precursors prepared with different solvents was performed with a simultaneous thermal analyzer (STA 449C, NETZSCH Group, Bavaria, Germany) to establish their optimal heat-treatment parameters. Crystal phase of the BaTi5O11 nanocrystals was characterized by an X-ray diffractometer (XRD, D8 Advance, Bruker Co., Billerica, MA, USA, Cu Kα, λ = 1.5406 Å, 40 kV and 40 mA). Morphologies were examined by field emission scanning electron microscopy (FESEM, Zeiss Ultra Plus, Jena, Germany) and transmission electron microscope (TEM, JEM-2100UHR, JEOL Ltd., Tokyo, Japan, 200 kV). Specific surface area and pore-texture properties of the BaTi5O11 nanocrystals were determined from N2 adsorption–desorption isotherms were measured by a Micromeritics ASAP 2460 analyzer (Micromeritics Instrument Co., GA, USA). Surface chemistry and elemental oxidation states of the BaTi5O11 nanocrystals were probed by X-ray Photoelectron Spectroscopy (XPS, Thermo Scientifc ESCALAB 250Xi instrument, Thermo Fisher Scientific, Inc., Waltham, MA, USA, Al Kα radiation).

4.3. Photocatalytic MB Degradation

The photocatalytic abilities of the BaTi5O11 nanocrystals were evaluated by monitoring the UV-driven degradation of MB. A mercury lamp (150 W) served as the UV source. Prior to irradiation, the BaTi5O11 nanocrystals (50 mg) were dispersed in aqueous MB solution (50 mL, 10 mg/L) with continuous stirring for 30 min in darkness to reach adsorption–desorption equilibrium between MB and the BaTi5O11 nanocrystal surface. When the UV light illuminated, 2 mL aliquots were taken out every 10 min, followed by centrifugation (10,000 rpm, 10 min) to remove the BaTi5O11 nanocrystal photocatalysts. The residual MB concentration in the supernatant was measured using UV–vis spectrophotometry (UV1800PC, Jinghua, Shanghai, China).

Author Contributions

Conceptualization, D.G. and H.W.; methodology, H.W. and J.W.; validation, H.W., Y.H.; formal analysis, H.W. and J.W.; investigation, H.W. and T.G.; resources, Z.H. and D.G.; data curation, X.L.; writing—original draft preparation, H.W.; writing—review and editing, D.G.; visualization, H.W. and Y.H.; supervision, Z.H.; project administration, D.G.; funding acquisition, Z.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Natural Science Foundation of China (Grant No. 50902108).

Data Availability Statement

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

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Masekela, D.; Hintsho-Mbita, N.; Sam, S.; Yusuf, T.; Mabuba, N. Application of BaTiO3-based catalysts for piezocatalytic, photocatalytic and piezo-photocatalytic degradation of organic pollutants and bacterial disinfection in wastewater: A comprehensive review. Arab. J. Chem. 2023, 16, 104473. [Google Scholar] [CrossRef]
  2. Ray, S.; Cho, J.; Hur, J. A critical review on strategies for improving efficiency of BaTiO3-based photocatalysts for wastewater treatment. J. Environ. Manage. 2021, 290, 112679. [Google Scholar] [CrossRef]
  3. Taghreed, A.; Riham, S.; Muftah, H. Organic Contaminants in Industrial Wastewater: Prospects of Waste Management by Integrated Approaches. In Combined Application of Physico-Chemical & Microbiological Processes for Industrial Effluent Treatment Plant; Springer: Singapore, 2020; pp. 205–235. [Google Scholar]
  4. Yarbay, R.; Şimşek, V.; Bogdan, L.; Tomašić, V. Photocatalytic Degradation of Neonicotinoids—A Comparative Study of the Efficacy of Hybrid Photocatalysts. Processes 2024, 12, 489. [Google Scholar] [CrossRef]
  5. Bollinger, J.C.; Lima, E.C.; Mouni, L.; Salvestrini, S.; Tran, H.N. Molecular properties of methylene blue, a common probe in sorption and degradation studies: A review. Environ. Chem. Lett. 2025, 23, 1403–1424. [Google Scholar] [CrossRef]
  6. Liu, M.; You, X.; Li, Y.; Yang, Y. Critical review on development of methylene blue degradation by wet catalytic methods. Rev. Chem. Eng. 2025, 41, 179–195. [Google Scholar] [CrossRef]
  7. Yamashita, Y.; Tada, M.; Kakihana, M.; Osada, M.; Yoshida, K. Synthesis of RuO2-loaded BaTinO2n+1 (n = 1, 2 and 5) using a polymerizable complex method and its photocatalytic activity for the decomposition of water. J. Mater. Chem. 2002, 12, 1782–1786. [Google Scholar] [CrossRef]
  8. Kohno, M.; Kaneko, T.; Ogura, S.; Sato, K.; Inoue, Y. Dispersion of ruthenium oxide on barium titanates (Ba6Ti17O40, Ba4Ti13O30, BaTi4O9 and Ba2Ti9O20) and photocatalytic activity for water decomposition. J. Chem. Soc. Faraday Trans. 1998, 94, 89–94. [Google Scholar] [CrossRef]
  9. Manar, M.; Mohamed, F.; Mohamed, E.; Nabil, E.; Moritz, R.; Ghada, B. Studying the kinetic parameters of BaTi5O11 by using the thermoluminescence technique. Arab. J. Chem. 2023, 16, 105247. [Google Scholar]
  10. Álvarez, D.; Carmen, M.; Reinosa, J.; Canu, G.; Buscaglia, M.; Buscaglia, M.; Fernández, J. Revealing the Role of the Intermediates during the Synthesis of BaTi5O11. Inorg. Chem. 2019, 58, 8120–8129. [Google Scholar] [CrossRef]
  11. Chen, Y.; Li, E.; Duan, S.; Zhang, S. Low temperature sintering kinetics and microwave dielectric properties of BaTi5O11 ceramic. ACS Sustain. Chem. Eng. 2017, 5, 10606–10613. [Google Scholar] [CrossRef]
  12. Fukui, T.; Sakurai, C.; Okuyama, M. Effects of heating rate on sintering of alkoxide-derived BaTi5O11 powder. J. Mater. Res. 1992, 7, 192–196. [Google Scholar] [CrossRef]
  13. Tangjuank, S.; Tunkasiri, T. Effects of heat-treatment on structural evolution and morphology of BaTi5O11 powder synthesized by the sol-gel method. Mater. Sci. Eng. B 2004, 108, 223–226. [Google Scholar] [CrossRef]
  14. Sözen, A.P. Synthesis of BaTi5O11 by an aqueous co-precipitation method via a stable organic titanate precursor. J. Serb. Chem. Soc. 2021, 86, 415–427. [Google Scholar]
  15. Zhang, W.; Han, J.; Cheng, X.; Guo, D. Photocatalytic activity of BaTi5O11 nanocrystals synthesized by hydrothermal method. Funct. Mater. Lett. 2024, 249, 2450015. [Google Scholar] [CrossRef]
  16. Cheng, X.; Lu, G.; Huang, Y.; Guo, D. Photocatalytic mechanism of BaTi5O11 nanocrystals for methylene blue degradation. Mater. Lett. 2024, 376, 137302. [Google Scholar] [CrossRef]
  17. Yang, G.; Zhao, X.; Che, J.; Huang, K.; Peng, S.; Wang, J.; Guo, D. Photocatalytic performance of BaTi5O11 nanocrystals synthesized by sol-gel method for methylene blue degradation. Funct. Mater. Lett. 2025, 18, 2550013. [Google Scholar] [CrossRef]
  18. Gong, Y.; Liu, C.; Guo, D.; Wang, C.; Shen, Q.; Zhang, L. Effect of Zr content on microstructure of BaTi2O5 powders prepared by sol-gel method. Mater. Lett. 2012, 73, 72–74. [Google Scholar] [CrossRef]
  19. Panomsuwan, G.; Manuspiya, H. Correlation between size and phase structure of crystalline BaTiO3 particles synthesized by sol-gel method. Mater. Res. Exp. 2019, 6, 065062. [Google Scholar] [CrossRef]
  20. Bokov, D.; Jalil, A.T.; Chupradit, S.; Suksatan, W.; Ansari, M.J.; Shewael, I.H.; Valiev, G.H.; Kianfar, E. Nanomaterial by sol-gel method: Synthesis and application. Adv. Mater. Sci. Eng. 2021, 2021, 5102014. [Google Scholar] [CrossRef]
  21. Zarkov, A. Sol–Gel Technology Applied to Materials Science: Synthesis, Characterization and Applications. Materials 2024, 17, 462. [Google Scholar] [CrossRef]
  22. Niederberger, M. Nonaqueous sol-gel routes to metal oxide nanoparticles. Acc. Chem. Res. 2007, 9, 793–800. [Google Scholar] [CrossRef]
  23. Rezapour, M.; Talebian, N. Comparison of structural, optical properties and photocatalytic activity of ZnO with different morphologies: Effect of synthesis methods and reaction media. Mater. Chem. Phys. 2011, 129, 249–255. [Google Scholar] [CrossRef]
  24. Wang, Y.; Zhong, S.; Zeng, X.; Lin, M.; Lin, C.; Lin, T.; Gao, M.; Zhao, C.; Wu, X. Microstructural manipulation and enhanced piezo-photocatalytic performance of sol-gel-derived pure BaTiO3 particles. Ceram. Int. 2025, 51, 3960–3969. [Google Scholar] [CrossRef]
  25. Mageste, F.; Sampaio, D.; Mayrink, T.; Alcamand, H.; Palhares, H.; Nunes, E.; Houmard, M. Influence of the alcoholic solvent and gelation temperature on the structural and water vapor adsorption properties of silica adsorbents synthesized by sol-gel method without catalyst. Microporous Mesoporous Mater. 2022, 342, 112114. [Google Scholar] [CrossRef]
  26. Dutta, A.; Behera, R.K.; Pradhan, N. Solvent Polarity: How Does This Influence the Precursor Activation, Reaction Rate, Crystal Growth, and Doping in Perovskite Nanocrystals? ACS Energy Lett. 2019, 4, 926–932. [Google Scholar] [CrossRef]
  27. Xu, Z.; Wang, W.; Li, Y.; Meng, S.; Liu, Y.; Cao, Y.; You, J.; Zhang, B.; Zang, J. The influence of solvent-temperature parameters of sodium titanium phosphate anode materials prepared by sol-gel method. Mater. Chem. Phys. 2024, 315, 128946. [Google Scholar] [CrossRef]
  28. Gholami, T.; Seifi, H.; Dawi, E.A.; Pirsaheb, M.; Seifi, S.; Aljeboree, A.M.; Hamoody, A.M.; Altimari, U.S.; Abass, M.A.; Salavati-Niasari, M. A review on investigating the effect of solvent on the synthesis, morphology, shape and size of nanostructures. Mater. Sci. Eng. B 2024, 304, 117370. [Google Scholar] [CrossRef]
  29. Donohue, M.D.; Aranovich, G.L. Adsorption hysteresis in porous solids. J. Colloid Interface Sci. 1998, 205, 121–130. [Google Scholar] [CrossRef]
  30. Wan, S.; Wang, W.; Cheng, B.; Luo, G.; Shen, Q.; Yu, J.; Zhang, J.; Cao, S.; Zhang, L. A superlattice interface and S-scheme heterojunction for ultrafast charge separation and transfer in photocatalytic H2 evolution. Nat. Commun. 2024, 15, 9612. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of BaTi5O11 samples synthesized using different solvents: (a) methanol, (b) ethanol, (c) 2-methoxyethanol, (d) acetylacetone, (e) PEG-200 and (f) EG.
Figure 1. XRD patterns of BaTi5O11 samples synthesized using different solvents: (a) methanol, (b) ethanol, (c) 2-methoxyethanol, (d) acetylacetone, (e) PEG-200 and (f) EG.
Gels 11 00706 g001
Figure 2. SEM patterns of the BaTi5O11 samples synthesized using different solvents: (a) methanol, (b) ethanol, (c) 2-methoxyethanol, (d) acetylacetone, (e) PEG-200 and (f) EG.
Figure 2. SEM patterns of the BaTi5O11 samples synthesized using different solvents: (a) methanol, (b) ethanol, (c) 2-methoxyethanol, (d) acetylacetone, (e) PEG-200 and (f) EG.
Gels 11 00706 g002
Figure 3. (a) N2 gas adsorption–desorption isotherms, (b) specific surface area and average pore size and (c) pore-size distribution of the BaTi5O11 nanocrystals synthesized with different solvents.
Figure 3. (a) N2 gas adsorption–desorption isotherms, (b) specific surface area and average pore size and (c) pore-size distribution of the BaTi5O11 nanocrystals synthesized with different solvents.
Gels 11 00706 g003
Figure 4. XPS spectra of BaTi5O11 nanocrystals using PEG-200 as the solvent: (a) XPS survey spectrum, (bd) high-resolution XPS spectra Ba 3d, Ti 2p and O 1s, respectively.
Figure 4. XPS spectra of BaTi5O11 nanocrystals using PEG-200 as the solvent: (a) XPS survey spectrum, (bd) high-resolution XPS spectra Ba 3d, Ti 2p and O 1s, respectively.
Gels 11 00706 g004
Figure 5. (a) TEM images of BaTi5O11 nanocrystals using PEG-200 as the solvent, and (b) its high-resolution TEM image.
Figure 5. (a) TEM images of BaTi5O11 nanocrystals using PEG-200 as the solvent, and (b) its high-resolution TEM image.
Gels 11 00706 g005
Figure 6. (a) UV–vis diffuse reflectance spectra of BaTi5O11 nanocrystals synthesized with different solvents, and their (b) Tauc plots of (αhv)2 vs. hv.
Figure 6. (a) UV–vis diffuse reflectance spectra of BaTi5O11 nanocrystals synthesized with different solvents, and their (b) Tauc plots of (αhv)2 vs. hv.
Gels 11 00706 g006
Figure 7. (a) UV–vis absorption spectra of MB solutions photodegraded at different times by BaTi5O11 nanocrystals synthesized with PEG-200 as the solvent, (b) time-dependent photocatalytic degradation efficiencies, and (c) their pseudo-first-order kinetics for MB in the presence of BaTi5O11 nanocrystals synthesized with different solvents, (d) photocatalytic cycling performance of BaTi5O11 nanocrystals synthesized with PEG-200 for MB degradation.
Figure 7. (a) UV–vis absorption spectra of MB solutions photodegraded at different times by BaTi5O11 nanocrystals synthesized with PEG-200 as the solvent, (b) time-dependent photocatalytic degradation efficiencies, and (c) their pseudo-first-order kinetics for MB in the presence of BaTi5O11 nanocrystals synthesized with different solvents, (d) photocatalytic cycling performance of BaTi5O11 nanocrystals synthesized with PEG-200 for MB degradation.
Gels 11 00706 g007
Figure 8. (a) DSC and (b) TG curves of BaTi5O11 precursors with different solvents.
Figure 8. (a) DSC and (b) TG curves of BaTi5O11 precursors with different solvents.
Gels 11 00706 g008
Table 1. The synthesis conditions, physical characteristics and degradation performance of BaTi5O11 synthesized using different solvents.
Table 1. The synthesis conditions, physical characteristics and degradation performance of BaTi5O11 synthesized using different solvents.
SolventsMethanolEthanol2-MethoxyethanolAcetylacetoneEGPEG-200
Molecular formulaCH4OC2H6OC3H8O2C5H8O2(CH2OH)2(CH2OH)n
Baking temperature (℃)150150150150220250
Viscosity
(mPa·s, 25 °C)
0.5501.071.801.4817.322.0
Specific surface area (m2/g)3.693.783.974.698.699.78
Average pore size (nm)43.840.542.928.927.817.8
Eg (eV)3.453.613.443.453.593.61
K Value (min−1)4.98 × 10−29.12 × 10−25.13 × 10−20.133.98 × 10−20.28
Table 2. The physicochemical characteristics of solvents.
Table 2. The physicochemical characteristics of solvents.
SolventsAcetic AcidMethanolEthanol2-MethoxyethanolAcetylacetoneEGPEG-200
Molecular FormulaC2H4O2CH4OC2H6OC3H8O2C5H8O2(CH2OH)2(CH2OH)n
Molecular Weight6032467610062200
Boiling Point (℃)117.964.778.3124141197250
Permittivity6.1532.725.816.9323.13720.5
Viscosity (mPa·s, 25 °C)1.220.5501.071.801.4817.322.0
Density (g/cm3)1.050.7910.7900.9650.9751.111.27
Surface Tension31.920.122.327.627.548.449.1
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Wang, H.; Gao, T.; Li, X.; Huang, Y.; Wang, J.; Huang, Z.; Guo, D. Microstructure and Photocatalytic Performance of BaTi5O11 Nanocrystals Synthesized via Sol-Gel Method Mediated by Organic Solvents. Gels 2025, 11, 706. https://doi.org/10.3390/gels11090706

AMA Style

Wang H, Gao T, Li X, Huang Y, Wang J, Huang Z, Guo D. Microstructure and Photocatalytic Performance of BaTi5O11 Nanocrystals Synthesized via Sol-Gel Method Mediated by Organic Solvents. Gels. 2025; 11(9):706. https://doi.org/10.3390/gels11090706

Chicago/Turabian Style

Wang, Honghua, Tianchen Gao, Xinyi Li, Yuci Huang, Junjie Wang, Zhixiong Huang, and Dongyun Guo. 2025. "Microstructure and Photocatalytic Performance of BaTi5O11 Nanocrystals Synthesized via Sol-Gel Method Mediated by Organic Solvents" Gels 11, no. 9: 706. https://doi.org/10.3390/gels11090706

APA Style

Wang, H., Gao, T., Li, X., Huang, Y., Wang, J., Huang, Z., & Guo, D. (2025). Microstructure and Photocatalytic Performance of BaTi5O11 Nanocrystals Synthesized via Sol-Gel Method Mediated by Organic Solvents. Gels, 11(9), 706. https://doi.org/10.3390/gels11090706

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