Next Article in Journal
Sustainable Cellulose Nanofibril−Stabilized Pickering Emulsions for Fresh Egg Preservation
Previous Article in Journal
First-Principles Guided Design of Pd-Decorated V2O5/Porous Silicon Composites for High-Performance NO2 Sensing at Room Temperature
Previous Article in Special Issue
A Bifunctional Nanostructured RuPt/C Electrocatalyst for Energy Storage Based on the Chlor-Alkali Process
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

In Situ Anchored, Ultrasmall, Oxygen Vacancy-Rich TiO2−x on Carbonized Bacterial Cellulose for the Efficient Adsorption and Separation of Organic Pollutants

1
Jiangsu Province Key Laboratory of Fine Petrochemical Engineering, Changzhou University, Changzhou 213164, China
2
BGRIMM Technology Group, Daxing, Beijing 102600, China
3
Zhongxi Rare Earth New Materials Co., Ltd., Changzhou 213164, China
4
School of Materials Science and Engineering, Nanjing University of Posts and Telecommunications, Nanjing 210023, China
*
Authors to whom correspondence should be addressed.
Nanomaterials 2025, 15(7), 514; https://doi.org/10.3390/nano15070514
Submission received: 17 February 2025 / Revised: 16 March 2025 / Accepted: 26 March 2025 / Published: 28 March 2025
(This article belongs to the Special Issue Nano-Enabled Materials for Clean Water and Energy Generation)

Abstract

:
Superior selective adsorption of organic dye is still a big challenge in the process of dye wastewater treatment. Meanwhile, low-price and environmentally friendly biomass-based adsorbents show huge potential in the fields of separation and purification. In this study, we adopted the “hydrolysis–calcination method” to develop a novel in situ anchoring strategy for ultrasmall TiO2−x on carbonized bacterial cellulose (CBC), which was derived from natural bacterial cellulose. Notably, 3D networks of porous CBC played a dual role for both providing hydrolytic sites and controlling the oxygen vacancies (Vo) of TiO2−x. As for the single-dye adsorption, the TiO2−x/CBC had a strong adsorption ability (101.4 mg/g) for removing methylene blue (MB), which was much higher than that of methyl orange (MO), malachite green (MG), rhodamine B (RhB), and tetracyclines (TC). Moreover, under the optimized carbonization temperature (Tc) of 300 °C, the TiO2−x/CBC-300 exhibited an outstanding separation efficiency of 97.07% for the MB/MO solution. Detailed analysis confirmed that Tc was a key regulator for adjusting the Vo concentration, which directly influenced the surface charge density and, further, the separation efficiency of TiO2−x/CBC. Additionally, the used adsorbent could be easily regenerated from washing by ethanol. After 4 regenerations, the adsorption efficiency declined only by 6.9% after 20 min and 13.6% after 120 min adsorption, respectively. Ultimately, this oxygen vacancy-rich TiO2−x/BC system illuminated good prospects for mixed dye wastewater adsorption and separation.

1. Introduction

On a global scale, the textile, printing, and dyeing industry still discharges the majority of industrial wastewater [1]. In industrial production, such as printing and dyeing and textile manufacturing, organic dyes such as methyl orange (MO), malachite green (MG), and rhodamine B (RhB) are widely used as raw materials. The wastewater containing these dyes, which is discharged after production, seriously threatens the natural environment and the health of all living organisms [2]. Wastewater from printing and dyeing is characterized by its large volume, complex composition, as well as high chromaticity, which increases the difficulty of water purification [3]. According to the acceptable limits of many dyes (0.01–0.05 mg/L) [4] for industrial discharge, some new treatment processes were widely reported alongside conventional methods, such as photocatalysis methods, biosorption methods [5], electrochemical methods, and electron beam treatments [6]. However, some disadvantages (e.g., high costs, complex operation, and low selectivity) limit the further application of these wastewater treatment technologies, especially for particular purposes like rapidness, simplicity, reusability, and high selectivity during the purification process [7].
In recent years, physical adsorption methods have been deeply studied, with the development of various inorganic adsorbents for dye wastewater treatment [8]. As a class of typical adsorbents, porous carbon materials derived from natural biomass become an efficient solution in the field of adsorption due to their abundant pore structures, good stability, and reusability [9]. For example, porous N-doped carbons (NPCs) derived from the bottlebrush flower reached a large specific surface area (~2025 m2·g−1), which exhibited high efficiency in removing binary organic pollutants [10]. Similarly, natural biomass-derived, carbon-based adsorbents have been studied extensively, such as eggplant-based porous carbon materials (~560 m2·g−1) [11], coconut-based carbon materials (~900 m2·g−1) [12], bamboo-based carbon adsorbents (~933 m2·g−1) [13], corncob-based carbon adsorbents (~1722 m2·g−1), etc. [14]. Different from these natural plant, bacterial cellulose (BC), with its unique 3D cross-linked structure, is an ideal candidate for dye wastewater treatment [15]. The key advantage of BC is its special synthesis process of bacterial fermentation, which leads to a high purity and high hydroxyl density. Based on the methods of chemical modification, hydroxyl groups play an important role for the in situ growth of metal hydroxide and metallic oxide [16]. In the aqueous solution system, MB ionizes to release a cation (-N+(CH3)3), while MO exists in an anionic state due to the ionization of the sulfonate group (-SO3⁻). The OH groups on the surface of the nanomaterial can enhance the adsorption of MB cations. For example, several BC-derived composites, such as poly(vinyl amine)-grafted BC (PVAm-g-BC) [17], achieve a 96.5% MB adsorption efficiency. FeOOH/carbonized bacterial cellulose (FeOOH/CBC) [18] shows a 96.2% MO adsorption efficiency. These materials efficiently remove single-dye components. However, these physical adsorption processes can hardly achieve selective dye adsorption due to their lack of functional adsorption sites [19].
In practice, real textile and dye wastewaters always present as a mixture, containing different kinds of cationic, anionic, and neutral dyes [20]. Compared with non-selective adsorption, a highly selective adsorption process with a simple and effective regeneration ability is one of the ideal solutions but is still a challenging task. Till now, some adsorbents with selective adsorption properties have been reported by anchoring semiconductors as functional sites, such as MIL-101(Fe) [21], La2Mo2O9 [22], ZIF-67 [23], carbon microtubes [24], etc. However, these artificial nanostructures still have some hindering factors, including their high cost, high toxicity, and complex synthesis, which prevent their industrial applications. Different from the nanostructures mentioned above, TiO2 is an old but newly developed semiconductor covering many fields [25], such as heterogeneous catalysis [26], photocatalysis [27], electrocatalysis [28], etc. Over the past decades, TiO2 semiconductors, especially nano-TiO2, have been used as traditional adsorbents [29]; this combination provided two obvious advantages. Firstly, both TiO2 and carbon materials are eco-friendly materials. Secondly, owing to the low densities of both two components, TiO2-anchored carbon composites have great potential for the highly efficient adsorption and separation process [30,31]. For over a decade, TiO2-based materials were designed and constructed with unique morphologies (e.g., hollow microspheres [32], mesoporous structures [33,34], and nanotubes [35]). To overcome the shortcomings of low selective adsorption on the surface of pure TiO2, researchers focused on surface and interface chemistry, including amino modification [36], co-doping [37,38], Ag2O modification [39], bimetallic alloy deposition [40], etc. Overall, TiO2-based adsorbents with a low price, high selectivity, and easy regeneration have rarely been reported.
Herein, a new class of 3D carbonized bacterial cellulose (CBC)-supported TiO2-based composites (TiO2−x/CBC) was synthesized using the “hydrolysis–calcination method”. Natural BC fibers with a large surface area and rich hydroxyl groups could provide ideal anchoring sites during the in situ hydrolysis process. Benefiting from the highly dispersed precursor, the carbonization temperature (Tc) acted as a key regulator for adjusting both the degree of carbonization and the concentration of oxygen vacancy. In the field of materials science, chemical vacancies (acidic sites) and electronic vacancies (tuning d-band center) have a significant impact on the characteristics of oxygen vacancies. As acidic sites, chemical vacancies can form bonds with oxygen atoms by providing protons or electrons, facilitating the desorption of oxygen atoms and, thus, contributing to the formation of oxygen vacancies. Electronic vacancies, on the other hand, can influence the interaction between oxygen vacancies and surrounding atoms by adjusting the energy center of the d-band, thereby altering the catalytic performance related to oxygen vacancies. The relationship between the state of surface charge and the adsorption performance of TiO2−x/CBC was evaluated under four different conditions of Tc from 200 °C to 500 °C. The optimized TiO2−x/CBC adsorbent showed excellent selective performance and good cycle ability for separating the binary mixtures of MB/MO.

2. Materials and Methods

2.1. Materials

All chemicals utilized in the experiment were of analytical-grade purity and required no further purification. Tetrabutyl titanate (TBT, C16H36O4Ti), anhydrous ethanol (C2H6O), anhydrous methanol (CH4O), hydrochloric acid (HCl), sodium hydroxide (NaOH), methylene blue (MB, C16H18ClN3S), methyl orange (MO, C14H14N3NaO3S), malachite green (MG, C23H25ClN2), rhodamine B (RhB, C28H31ClN2O3), and tetracycline (TC, C22H24N2O8) were procured from Shanghai Aladdin Biochemical Technology Co., Ltd. (Shanghai, China). Biomass from BC membranes was sourced from Guilin Qihong Co., Ltd. (Guilin, China). Liquid nitrogen was purchased from Changzhou Huayang Gas Co., Ltd. (Changzhou, China). Distilled water was used throughout the experimentation. The freeze dryer (FD-2A) was purchased from Shanghai Bilang Instrument Manufacturing Co., Ltd. (Shanghai, China), and the tube furnace (GSL-1100X-S) was purchased from Hefei Kejing Materials Technology Co., Ltd. (Heifei, China)

2.2. Purification of BC

To remove residual impurities from the fibers of the purchased BC, a purification process was followed before the growth of TiO2. Initially, an NaOH solution with a concentration of 2% was prepared. Then, the BC membranes were immersed in the NaOH solution at 90 °C for 1 h. After cooling down to room temperature, the BC membranes were carefully washed with deionized water until it reached a neutral pH [41]. To maintain the 3D structure of the biomass, the purified BC membranes were post-treated with liquid nitrogen quick-freezing for several seconds. Finally, the purified and dried BC carriers were obtained after freeze-drying for 24 h. The freeze-drying temperature was fixed at −80 °C (cold trap temperature).

2.3. Preparation of TiO2/BC and TiO2−x/CBC Adsorbents

In a typical synthesis of TiO2/BC, 36 mL of deionized water was mixed with 4 mL of HCl, followed by adding different volumes of TBT (1, 2, 3, and 4 mL, separately). For the preparation of TiO2/BC samples, about 100 mg pure BC precursor was used after purification. Subsequently, the BC membranes were placed into a Teflon reactor and reacted at 90 °C for 9 h. The resultant solid materials were washed with ethanol and distilled water and then freeze-dried for 6 h to obtain TiO2/BC. According to the different amount of TBT, the intermediate products were denoted as TiO2/BC-1, TiO2/BC-2, TiO2/BC-3, and TiO2/BC-4, respectively.
The sample of TiO2/BC-4 was used for the study of the calcination temperature. In a typical synthesis of TiO2−x/CBC (Figure 1), TiO2/BC precursors were calcined under the same conditions, including a calcination atmosphere (flowing pure N2) and the same heating rate (5 °C/min) and calcination time (1 h). The only key variable of the calcination process was the calcination temperature (Tc) of four different points (200 °C, 300 °C, 400 °C, and 500 °C). Correspondingly, these partially carbonized composites were labeled as TiO2−x/CBC-200, TiO2−x/CBC-300, TiO2−x/CBC-400, and TiO2−x/CBC-500.

2.4. Characterization

The phase and crystallinity information of the composites were qualitatively analyzed using X-ray diffraction (XRD, Rigaku Co., Ltd., Tokyo, Japan). The analysis employed Cu Kα radiation (λ = 1.54056 Å), with the instrument operating at 40 kV and 100 mA. A JSM-6360LA scanning electron microscope (SEM, JEOL, Tokyo, Japan) and a JEM-2100 transmission electron microscope (TEM, JEM-2100, JEOL, Tokyo, Japan) were used to observe the microscopic morphology and surface structure of all the composites. The samples were tested for nitrogen adsorption–desorption using a fully automated specific surface area analyzer (Micromeritics ASAP 2000, Norcross, GA, USA), and the pore size and specific surface area of the materials were obtained by analyzing the isothermal adsorption–desorption curves using the BET method. The thermal stability was analyzed via thermogravimetric analysis (TGA, TG209F3 Tarsus, Netzsch, Selb, Germany), and X-ray photoelectron spectroscopy (XPS, Thermo Scientific KAlpha, Waltham, MA, USA) was performed to characterize the compositions of materials. In addition, a Fourier transform infrared spectrometer (FTIR, Nicolet iS50, Madison, WI, USA) and Raman spectroscopy were used to analyze the surface functional groups and chemical compositions of the samples. Solid UV-vis-NIR diffuse reflectance spectroscopy (DRS, UV-3600i Plus UV-vis-NIR, Shimadzu, Kyoto, Japan) was used to characterize the band gap widths of the samples. To assess the surface charge density and the intensity of electrostatic interactions on the synthesized nanoparticles, the Zeta potential of the samples at different calcination temperatures was measured using a ZEN3600 Laser Particle Size and Zeta Potential Analyzer (Zetasizer Software 7.11).

2.5. Evaluation of the Adsorption Activity

Firstly, our study verified the single-component adsorption activities of as-prepared TiO2−x/CBC composites using MO, MG, RhB, TC, and MB as typical probes. The concentrations of all the above solutions were kept at 20 mg/L. To identify the separation effect, a mixed-dye solution (50 mL) was formed from MB and MO with equal concentrations of 20 mg/L. The quality of each adsorbent was fixed at 10 mg for each test. After adsorption for a period of time, 2 mL of the reaction suspension samples was collected and centrifuged (8000 rpm, 3 min) before the absorption spectra were measured. The absorbance values near 664 nm and 465 nm were analyzed using a UV-vis spectrophotometer (Shanghai Precision & Scientific Instrument Co., Ltd., Shanghai, China) to calculate the concentrations of MB and MO, respectively. The removal efficiency (R) of MB or MO by different samples is expressed as Equation (1) [42]:
R = C 0 C t C 0 × 100 %
where C0 is the initial MB or MO concentration, and Ct is the MB or MO concentration after the reaction for time t. The separation efficiency for the mixture of organic dyes, which, in this context, represents selectivity, can be evaluated using Equation (2) [43]:
Separation efficiency   ( % ) = C ( M O ) t C ( M O ) t + C ( M B ) t × 100 %
where C(MO)t C(MO)t and C(MB)t C(MB)t are the concentrations of MO and MB dyes remaining in the solution after the adsorption.
The adsorption kinetics was studied using the pseudo-second-order model, as shown below [44]:
  t / Q t = 1 / k 2 Q e 2 + t / Q e
In Equation (3), k2 is the pseudo-second-order rate constant (mg/g·min), Qe is the adsorption capacity at equilibrium, and Qt is the adsorption capacity at time t. The values of Qe and k2 can be obtained from the linear relationship plot of Qt Qt/t versus t.

3. Results and Discussion

3.1. Material Characterization

Figure 2 shows the XRD patterns and SEM images of TiO2/BC with four different concentrations of TBT (Figure 2a–e), as well as four different calcination temperatures (Tc) (Figure 2f–j). Two typical peaks at 14.5° and 22.8° corresponding to the (110) and (200) planes of cellulose clearly existed in both BC and TiO2/BC. The XRD pattern of pure TiO2 exhibited a mixture of anatase and rutile phases. Moreover, from the red line to the green line in Figure 2a, the (110) peak intensities of TiO2 varies with increasing amounts of TBT, indicating an increase loading of TiO2. The intensity ratio of (200) peaks at 22.8° among pure BC (black line), TiO2/BC-1 (red line), and TiO2/BC-4 (green line) was 100:78:20, respectively. The detailed peak intensities can be found in Table S1 and Figure S1. The variation in TiO2 peak intensity between TiO2/BC-3 and TiO2/BC-4 is minimal, indicating that the available sites on the fibers had been occupied, preventing further loading. As shown in Figure 2b–e, BC carriers showed a typical three-dimensional structure with plenty of empty spaces, which helped TBT molecules quickly diffuse and anchor onto BC fibers. Under the condition of 1 mL (Figure 2b) and 2 mL (Figure 2c) of TBT, only a few TiO2 precursors were loaded onto BC fibers. When the amount of TBT grew up to 3 mL (Figure 2d) and 4 mL (Figure 2e), the gaps between BC fibers were partially shortened, while the diameter fibers became larger due to the adhesion effect in hydrolysis. Figure 2f shows the XRD patterns of TiO2−x/CBC under different Tc values. The characteristic peaks of CBC could be clearly seen in the samples of TiO2−x/CBC-300, TiO2−x/CBC-400, and TiO2−x/CBC-500. Interestingly, this kind of effect caused the lowest ratio of A/R (1.25) at 300 °C. But when using a high temperature of 500 °C, the ratio of A/R grew up to 2.50. The reason might be the interaction under a suitable Tc inhibited the growth of crystalline structures as well as the transformation from anatase to rutile phase. In our study, the calcination temperature was selected between 300 °C and 500 °C, which was much lower than the effective transition temperature. As shown in Table S2, carbon-based networks under a higher temperature (e.g., 400 °C and 500 °C), which derived from BC, significantly inhibit the phase transition process from the anatase phase to the rutile phase of TiO2 [45,46]. For comparison, Figure 2g–j present the morphologies of TiO2−x/CBC under different Tc values. Obviously, when the Tc was 200 °C (Figure 2g) or 300 °C (Figure 2h), ultrasmall TiO2−x particles were well distributed on the surface of the CBC fibers. However, as the Tc increased to 400 °C, partial aggregation of TiO2−x appeared in Figure 2i. Moreover, the condition of 500 °C caused serious aggregation of and structure damage to the CBC fibers (Figure 2j). The element mapping spectra and EDS energy spectrum of TiO2−x/CBC-300 can be seen in Figure S2. The atomic ratio of the C, Ti, and O elements, as determined via EDS, is presented in Table S3.
The detailed morphologies of metallic oxides on CBC fibers were analyzed via TEM (Figure 3a,b). Firstly, many “windows” with a small size (<50 nm) were observed between fibers, which were smaller than blank CBC (Figure S3). In addition, a large number of small particles were loaded on the surface of CBC. These ultrasmall particles occupied the initial windows within CBC membranes, resulting in a certain decrease in the specific surface area of TiO2−x/CBC-300. It is also demonstrated in subsequent sections through N2 adsorption–desorption isotherms. After measuring over 150 particles (Figure S4), the particle size distribution was determined, as shown in Figure 3c. The average size of TiO2−x was confirmed to be ~4.09 nm, which proved that the TiO2−x particles were well dispersed, with an ultrasmall particle size. To confirm the pore size and surface areas of adsorbents, the N2 adsorption–desorption isotherms of blank CBC, pure TiO2, and TiO2−x/CBC-300 and the relative results are shown in Figure 3d and Table S4. The adsorption–desorption hysteresis loop of TiO2 was relatively narrow, and the BET surface area (SBET) for pure TiO2 was 34.7 m2/g. In contrast, the desorption curve of CBC fibers became flatter with increasing pressure, suggesting that the pores of CBC were primarily composed of crack-like micropores. The SBET of CBC was 245.6 m2/g, much larger than BC fibers (77.82 m2/g) [47]. The adsorption isotherm of TiO2−x/CBC-300 exhibited a type IV isotherm with a high nitrogen adsorption rate at low pressures and a large hysteresis loop, indicating the presence of a large number of mesoporous structures. Compared with CBC, the SBET of TiO2−x/CBC-300 decreased to 182.6 m2/g due to the possibility that ultrasmall TiO2−x particles could easily fill the cracks and defects of CBC fibers. Considering the much smaller molecular sizes of the four organic dyes used in this work than the pore size of TiO2−x/CBC (Table S4), the pore size effect can be excluded.
XPS analysis was performed to investigate the surface chemistry and valence states (Figure 3e–f, Figures S5 and S6). Compared with pure TiO2, all the binding energies of the Ti 2p1/2 and Ti 2p3/2 peaks increased after combining CBC with TiO2−x (Figure 3e). Moreover, a higher calcination temperature induced the shift to higher binding energy, which indicated that a part of the Ti-O-Ti bonds were substituted by Ti-O-C and O=C-O-Ti bonds due to the higher electronegativity of the carbon element. As shown in Figure 3f, the binding energies of 529.3~530.1 eV, 530.3~530.7 eV, and 531.5~532.7 eV correspond to lattice oxygen (Ti-O), adsorbed oxygen on oxygen vacancies, and surface hydroxyl oxygen, respectively. As we know, oxygen vacancy (Vo) can directly determine the performance of selective adsorption [48,49]. It is worth to mention that the Tc of 300 °C led to the highest concentration of oxygen vacancy (Vo). Furthermore, a weak signal around the binding energy of 531.5 eV in TiO2−x/CBC-300 exhibited some kind of hydroxyl oxygen residue under the temperature of 300 °C. The electron paramagnetic resonance (EPR) spectra of all samples are shown in Figure S7. The unpaired electrons associated with the defects are identified by a single Lorentzian line at g = 2.003. The signal intensity of the EPR spectra corresponds to the concentration of oxygen vacancies. Based on the EPR, it can be inferred that the oxygen vacancy concentration is highest in TiO2−x/CBC-300, and this is consistent with the changes observed in the XPS. It means that either a too high or too low Tc could greatly influence the interaction between TiO2 and biomass fibers and further limit the oxygen vacancy concentration.
FTIR spectra were employed to analyze the surface functional groups on the pure TiO2, blank BC, CBC, and TiO2−x/CBC scaffolds, respectively. In Figure 4a, peaks of BC (black line) and CBC (red line) around 1059 cm−1, 2950 cm−1, and 3349 cm−1 were caused by symmetric stretching of C-O, deformation vibrations of the C-H bond, and stretching vibrations of O-H groups, respectively. For comparison, from TiO2−x/CBC-200 (green line) to TiO2−x/CBC-500 (pink line), the peak intensities of the C-O, C-H, and H-O-H groups were weakened, while the characteristic peak of TiO2, such as Ti-O-Ti, was gradually enhanced. The FTIR results indicate a strong interaction between titanium oxides and the CBC carriers. Furthermore, the Raman spectra of TiO2 and TiO2−x/CBC-300 are shown in Figure 4b. Compared with pure TiO2, the TiO2−x/CBC-300 had two extra peaks that belonged to the D band (1334 cm−1) and G band (1585 cm−1) of CBC. The value of Id/Ig is only 0.24, which indicated few defects and a high degree of graphitization. Additionally, the Eg mode of anatase TiO2 was observed at 141.4 cm−1 and the Bg, A1g, and Eg modes of rutile TiO2 were confirmed at 438.9 cm−1, 605.8 cm−1, and 237 cm−1, respectively. The results were consistent with the XRD analysis that both the anatase phase and rutile phase were formed. In Figure 4c, TG analysis shows the thermal stability of TiO2−x/CBC-300. The decomposition temperature range moved from 193 °C–348 °C to 286 °C–411 °C. We chose 300 °C as the target temperature. At this point, the BC is in a state between initial and complete decomposition. This moderate decomposition generates small molecular fragments and new active sites, which are crucial for anchoring and evenly dispersing TiO2−x nanoparticles, thus enhancing the material’s performance. The TG results confirmed that the anchoring of TiO2 enhanced the thermal stability of natural BC fibers. As presented in Figure 4d, the UV-vis light absorption properties of the TiO2 and TiO2−x/CBC-300 were studied. The band-gap energy of the TiO2 was determined to be 2.89 eV by using Kubelka–Munk models (Figure 4e). Compared with pure TiO2, the band-gap energy of the TiO2−x/CBC-300 significantly reduced and shifted towards the visible light region (red-shift), which could lead to a better optical response. In Figure S9, the highest photocurrent response achieved by TiO2−x/CBC-300 illustrates the best separation capability of electron–hole pairs. PL spectrum observation also shows that TiO2−x/CBC-300 has a much lower PL intensity (Figure S9) than that of others. That is to say, TiO2−x/CBC-300 obtains not only the optimal light absorption capacity but also excellent performance in electron–hole separation. The photocatalytic degradation effect of MB by TiO2−x/CBC adsorbents is listed in Figure S10. As shown in Figure S10, during the 180 min photocatalytic degradation process, the degradation capabilities of the various catalysts towards MB exhibited significant differences. Specifically, the TiO2−x/CBC-300 catalyst demonstrated the highest degradation efficiency, followed by TiO2−x/CBC-400 and TiO2−x/CBC-500, while pure TiO2 and TiO2−x/CBC-200 exhibited relatively weaker degradation capabilities. This result not only validates the photocatalytic activities of these composites but also provides a potential application value for regeneration via photocatalytic degradation.

3.2. Adsorption Performance

To investigate the adsorption performance of TiO2−x/CBC for organic dyes, we analyzed the adsorption curves for both individual and mixed-dye solutions. Figure 5a compared the single-component adsorption activities for five different organic pollutants, including MO, MG, RhB, TC, and MB, separately (Figure S11). Taking TiO2−x/CBC-300 as an example, MB exhibited the highest adsorption effect while achieving a removal efficiency (R) of 86.74% after 300 min. In contrast, the R values for MO, MG, RhB, and TC were relatively lower at 2.45%, 9.11%, 9.42%, and 12.16%, respectively. MB and MO, as two representative organic dye compounds, have been extensively utilized in various industrial applications and scientific research fields. Further analysis of mixed solutions of MO/MB revealed distinct adsorption characteristics under different calcination temperatures. When the Tc condition was 200 °C, the weak removal rate for both MB and MO stayed at a very low level (Figure 5b). Compared with CBC and pure TiO2 (Figure S12), however, TiO2−x/CBC-300 showed significantly enhanced selectivity for the mixed solution after 120 min, with an ultrahigh Ct/C0 value for MO (97.2%) and an ultralow value for MB (6.9%) in Figure 5c. Moreover, the selective adsorption performance of TiO2−x/CBC-400 and TiO2−x/CBC-500 gradually decreased with higher calcination temperatures of 400 °C (Figure 5d) and 500 °C (Figure 5e). Overall, when the calcination temperature (Tc) varied from 200 °C to 500 °C, the separation efficiencies of the TiO2−x/CBC adsorbents for the mixed dyes at 120 min were 56.27%, 97.07%, 87.97%, and 78.56%, respectively. As a result, TiO2−x/CBC-300 had a superior selective adsorption ability due to its 3D cross-linked structure equipped with oxygen vacancy-rich surfaces. Notably, it was found consistent with the conclusion in terms of their surface Zeta potentials, as shown in Figure 6c. Therefore, the excellent selective adsorption capacity of TiO2−x/CBC-300 towards cationic organic dyes can be attributed to the increased surface negative charge stemming from the synergism between vacancy-rich TiO2 and π–π interaction. As we know, the stability of adsorbent is critical for practical applications. Thus, the cycling performance of TiO2−x/CBC-300 was carefully studied in the mixed solution of MB and MO. The solids were collected after adsorption and washed three times with water and methanol alternately under the condition of ultrasonication for 5 min in a 10 mL centrifuge tube. After being gathered via centrifugation, the adsorbent was then dried in an oven at 60 °C. After 6 h, the composite was ready for reuse in the next cycle. After four cycles of the adsorption–desorption process under dark conditions, the value of the separation efficiency remained at 83.84%, representing a slightly decrease by only 13.23% compared to the initial value of separation efficiency (Figure 5f). The SEM (Figure S13) patterns of TiO2−x/CBC-300 after regeneration exhibited similarity to those before the reaction, with no significant differences observed. We confirmed that the system of TiO2−x/CBC has both excellent selective adsorption and good stability for separating mixed dyes.
Figure 6a shows the adsorption kinetic curves of MB on CBC, TiO2, and TiO2−x/CBC. Based on the pseudo-second-order fitting plots for MB, within the first 20 min, four TiO2−x/CBC samples reached over 70% of the final saturated adsorption capacity, which indicated the quick adsorption abilities of TiO2−x/CBC. The final saturated adsorption capacities of blank CBC, pure TiO2, TiO2−x/CBC-200, TiO2−x/CBC-300, TiO2−x/CBC-400, and TiO2−x/CBC-500 after 120 min adsorption were 9.6, 11.9, 24.4, 97.2, 91.4, and 87.5 mg/g. In Figure 6b and Table S5, it can be seen that the correlation coefficient (R2 = 0.998) from the pseudo-second-order model is was higher than that from the pseudo-first-order model (R2 = 0.878), indicating that this TiO2−x/CBC system was controlled by liquid film diffusion, as well as internal diffusion and chemical adsorption. Moreover, the Zeta potential of composite is also crucial, particularly for the separation of anionic and cationic dyes. Figure 6c presents the Zeta potential profiles for TiO2−x/CBC across different calcination temperatures. TiO2−x/CBC-200 showed a weak Zeta potential of −2.6 mV, indicating weak electrostatic attraction for cationic dyes, correlating with the poor dye adsorption capacity shown in Figure 5b. In contrast, TiO2−x/CBC-300 achieved the strongest Zeta potential of −22.3 mV, while higher calcination temperatures of 400 °C and 500 °C yielded the Zeta potentials of −17.6 mV and −12.3 mV, respectively. Electrostatic interaction and π–π interaction were determined as the main driving forces, and the 3D cross-linked structure also contributed to the adsorption process. In conclusion, there is a strong dependency relationship among the (i) Vo concentration, (ii) surface electronegativity, and (iii) separation efficiencies. Compared with related publications (Figure 6d, Table S6), our TiO2-BC based material is convenient for synthesis with a higher separation efficiency. A possible mechanism for the high selectivity of TiO2−x/CBC-300 is shown in Figure 6e. From a preparation perspective, the carbonization temperature (Tc) plays a key role in adjusting the selective adsorption effects of anionic/cationic dyes.

4. Conclusions

In this work, we proposed a novel route for the in situ fabrication of ultrasmall TiO2−x particles on CBC fibers using a facile hydrolysis–calcination process. The rich -OH groups of BC ensured a well-dispersed and size-controlled synthesis of TiO2 particles. Additionally, the optimized calcination temperature of 300 °C introduced the highest oxygen vacancy and the biggest Zeta potential (absolute value) to the 3D scaffold materials. Regarding the investigation of adsorption performance, our initial efforts were concentrated on single-dye adsorption experiments, involving MB, MO, MG, RhB, and TC. The TiO2−x/CBC material manifested a robust adsorption capacity of 101.4 mg/g for MB. This value was markedly higher compared to the adsorption capacities for the other single dyes examined in this study. Subsequently, a binary mixture of MB and MO was prepared and subjected to adsorption tests. TiO2−x/CBC-300 exhibited an outstanding adsorption performance, especially high selectivity for mixed pollutants. As for the mixture of MB/MO, the separation efficiency of TiO2−x/CBC-300 was 97.07%. Moreover, the adsorbents could be easily regenerated following methanol washing. After four cycles of adsorption–desorption in a dark atmosphere, the separation efficiencies of the TiO2−x/CBC-300 remained over 83.84% for the target dye. This work provides compelling evidence for the huge potential of TiO2-based materials in designing high-efficiency adsorbents with excellent selectivities.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/nano15070514/s1. See Refs [24,43,50,51,52,53,54,55,56,57].

Author Contributions

Conceptualization, J.Z. (Jian Zhou) and L.L.; Methodology, M.Z., H.M. and J.Z. (Jin Zhao); Software, Y.Z. (Yanli Zhou) and M.N.; Validation, S.X., J.Z. (Jian Zhou), L.L. and Z.L.; Formal analysis, Y.Z. (Yuzhe Zhang) and H.M.; Resources, M.N. and S.X.; Data curation, Y.Z. (Yanli Zhou); Writing—original draft, Y.Z. (Yanli Zhou); Writing—review & editing, M.Z., J.Z. (Jin Zhao) and Z.L.; Visualization, M.N., S.X., J.Z. (Jian Zhou) and L.L.; Supervision, M.Z., Y.Z. (Yuzhe Zhang), J.Z. (Jin Zhao) and Z.L.; Project administration, Y.Z. (Yuzhe Zhang). All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Jiangsu Province Key Laboratory of Fine Petrochemical Engineering (No. KF2204).

Data Availability Statement

The original contributions presented in this study are included in the article/Supplementary Material. Further inquiries can be directed to the corresponding authors.

Conflicts of Interest

Author Jian Zhou was employed by the company Zhongxi Rare Earth New Materials Co., Ltd. The remaining authors declare that the research was conducted in the absence of any commercial or financial relationships that could be construed as a potential conflict of interest.

References

  1. Aramesh, N.; Bagheri, A.R.; Zhang, Z.; Yadollahi, B.; Lee, H.K. Polyoxometalate-based materials against environmental pollutants: A review, Coordin. Chem. Rev. 2024, 507, 215767. [Google Scholar] [CrossRef]
  2. Fallah, S.; Mamaghani, H.R.; Yegani, R.; Hajinajaf, N.; Pourabbas, B. Use of graphene substrates for wastewater treatment of textile industries. Adv. Compos. Hybrid Mater. 2020, 3, 187–193. [Google Scholar] [CrossRef]
  3. Xin, Z.; Duan, K.; Zhuo, Q.; He, Q.; Zhang, X.; Zheng, C.; Han, X.; Han, T.; Fu, Z.; Xu, X.; et al. Novel nanozyme Ag/Fe3O4@h-BN with peroxidase-mimicking and oxidase-mimicking activities for dye degradation, As(V) removal and detection. Chem. Eng. J. 2023, 461, 141589. [Google Scholar] [CrossRef]
  4. Hessel, C.; Allegre, C.; Maisseu, M.; Charbit, F.; Moulin, P. Guidelines and legislation for dye house effluents. J. Environ. Manag. 2007, 83, 171–180. [Google Scholar] [CrossRef]
  5. Bashir, A.; Malik, L.A.; Ahad, S.; Manzoor, T.; Bhat, M.A.; Dar, G.N.; Pandith, A.H. Removal of heavy metal ions from aqueous system by ion-exchange and biosorption methods. Environ. Chem. Lett. 2018, 17, 729–754. [Google Scholar] [CrossRef]
  6. Das, P.P.; Sharma, M.; Purkait, M.K. Recent progress on electrocoagulation process for wastewater treatment: A review. Sep. Purif. Technol. 2022, 292, 121058. [Google Scholar] [CrossRef]
  7. Wang, T.; Wang, W.; Hu, H.; Khu, S.-T. Effect of coagulation on bio-treatment of textile wastewater: Quantitative evaluation and application. J. Clean. Prod. 2021, 312, 127798. [Google Scholar] [CrossRef]
  8. Omer, A.S.; El Naeem, G.A.; Abd-Elhamid, A.I.; Farahat, O.O.M.; El-Bardan, A.A.; Soliman, H.M.A.; Nayl, A.A. Adsorption of crystal violet and methylene blue dyes using a cellulose-based adsorbent from sugercane bagasse: Characterization, kinetic and isotherm studies. J. Mater. Res. Technol. 2022, 19, 3241–3254. [Google Scholar] [CrossRef]
  9. Yuan, X.; Wang, X.; Hu, S.; Sun, S.; Li, C. PDA@UiO-66-NH2-derived nitrogen and oxygen-doped hierarchical porous carbon for efficient adsorption of BPA and dyes. Sep. Purif. Technol. 2024, 343, 127169. [Google Scholar] [CrossRef]
  10. Tian, W.; Sun, H.; Duan, X.; Zhang, H.; Ren, Y.; Wang, S. Biomass-derived functional porous carbons for adsorption and catalytic degradation of binary micropollutants in water. J. Hazard. Mater. 2020, 389, 121881. [Google Scholar] [CrossRef]
  11. Liu, Z.; Huang, X.; Miao, Y.; Gao, B.; Shi, Y.; Zhao, J.; Tan, S.H. Eggplant biomass based porous carbon for fast and efficient dye adsorption from wastewater. Ind. Crop. Prod. 2022, 187, 115510. [Google Scholar] [CrossRef]
  12. Wu, L.; Zhang, X.; Si, Y. Polydopamine functionalized superhydrophilic coconut shells biomass carbon for selective cationic dye methylene blue adsorption. Mater. Chem. Phys. 2022, 279, 125767. [Google Scholar] [CrossRef]
  13. Zhao, Z.; Lin, L.; Zhang, J.; Xu, B.; Ma, Y.; Li, J. A green approach to enhance the adsorption capacity: Synthesis of bamboo-based adsorbent by biological pretreatment. Ind. Crop. Prod. 2024, 213, 118388. [Google Scholar] [CrossRef]
  14. Sun, Z.; Qu, K.; Li, J.; Yang, S.; Yuan, B.; Huang, Z.; Guo, Z. Self-template biomass-derived nitrogen and oxygen co-doped porous carbon for symmetrical supercapacitor and dye adsorption. Adv. Compos. Hybrid Mater. 2021, 4, 1413–1424. [Google Scholar] [CrossRef]
  15. Wu, W.; Ni, M.; Feng, Q.; Zhou, Y.; Cui, Y.; Zhang, Y.; Xu, S.; Lin, L.; Zhou, M.; Li, Z. A wet bacterial cellulose film self-anchored by phosphotungstic acid: Flexible, quick-response and stable cycling performance for photochromic application. Mater. Design 2024, 238, 112613. [Google Scholar] [CrossRef]
  16. Araujo, I.M.S.; Silva, R.R.; Pacheco, G.; Lustri, W.R.; Tercjak, A.; Gutierrez, J.; Junior, J.R.S.; Azevedo, F.H.C.; Figueredo, G.S.; Vega, M.L.; et al. Hydrothermal synthesis of bacterial cellulose-copper oxide nanocomposites and evaluation of their antimicrobial activity. Carbohyd. Polym. 2018, 179, 341–349. [Google Scholar] [CrossRef]
  17. Rattanawongwiboon, T.; Khongbunya, N.; Namvijit, K.; Lertsarawut, P.; Laksee, S.; Hemvichian, K.; Madrid, J.F.; Ummartyotin, S. Eco-friendly dye adsorbent from poly(vinyl amine) grafted onto bacterial cellulose sheet by using gamma radiation-induced simultaneous grafting and base hydrolysis. J. Polym. Environ. 2023, 32, 3048–3060. [Google Scholar] [CrossRef]
  18. Wang, Z.; Ma, Y.; He, H.; Pei, C.; He, P. A novel reusable nanocomposite: FeOOH/CBC and its adsorptive property for methyl orange. Appl. Surf. Sci. 2015, 332, 456–462. [Google Scholar] [CrossRef]
  19. Huo, M.; Jin, Y.; Sun, Z.; Ren, F.; Pei, L.; Ren, P. Facile synthesis of chitosan-based acid-resistant composite films for efficient selective adsorption properties towards anionic dyes. Carbohyd. Polym. 2021, 254, 117473. [Google Scholar] [CrossRef]
  20. Gnanasekaran, G.; Sudhakaran, M.S.P.; Kulmatova, D.; Han, J.; Arthanareeswaran, G.; Jwa, E.; Mok, Y.S. Efficient removal of anionic, cationic textile dyes and salt mixture using a novel CS/MIL-100 (Fe) based nanofiltration membrane. Chemosphere 2021, 284, 131244. [Google Scholar] [CrossRef]
  21. Vinothkumar, K.; Jyothi, M.S.; Lavanya, C.; Sakar, M.; Valiyaveettil, S.; Balakrishna, R.G. Strongly co-ordinated MOF-PSF matrix for selective adsorption, separation and photodegradation of dyes. Chem. Eng. J. 2022, 428, 132561. [Google Scholar] [CrossRef]
  22. Sun, M.; Ma, Y.; Tan, H.; Yan, J.; Zang, H.; Shi, H.; Wang, Y.; Li, Y. Lanthanum molybdenum oxide as a new platform for highly selective adsorption and fast separation of organic dyes. RSC Adv. 2016, 6, 90010–90017. [Google Scholar] [CrossRef]
  23. Du, X.; Wang, C.; Liu, J.; Zhao, X.; Zhong, J.; Li, Y.; Li, J.; Wang, P. Extensive and selective adsorption of ZIF-67 towards organic dyes: Performance and mechanism. J. Colloid Interface Sci. 2017, 506, 437–441. [Google Scholar] [CrossRef]
  24. Ji, L.; Luo, L.; Jin, D.; Qin, X. Preparation of aligned porous carbon microtubes by a reactant permeation template method and the highly selective adsorption of methylene blue dye from wastewater. Sep. Purif. Technol. 2024, 332, 125884. [Google Scholar] [CrossRef]
  25. Al-Amrani, W.A.; Onaizi, S.A. Adsorptive removal of heavy metals from wastewater using emerging nanostructured materials: A state-of-the-art review. Sep. Purif. Technol. 2024, 343, 127018. [Google Scholar] [CrossRef]
  26. Kumari, N.; Mahata, A.; Chakraborty, B. Immobilization of phosphamide on the TiO2 surface for heterogeneous phase catalytic appel reaction. Inorg. Chem. 2023, 62, 7728–7737. [Google Scholar] [CrossRef]
  27. Guo, Q.; Ma, Z.; Zhou, C.; Ren, Z.; Yang, X. Single molecule photocatalysis on TiO2 surfaces. Chem. Rev. 2019, 119, 11020–11041. [Google Scholar] [CrossRef]
  28. Xie, A.; Martínez-Vargas, D.R.; Yang, Z.; Zou, S. Efficient selenate removal from impaired waters with TiO2-assisted electrocatalysis. Water Res. 2024, 262, 122134. [Google Scholar] [CrossRef]
  29. Huo, H.; Su, H.; Tan, T. The influence of trace TiO2 on adsorption of Ag+-imprinted adsorbents made from chitosan and mycelium. Biotechnol. Bioproc. Eng. 2008, 13, 77–83. [Google Scholar] [CrossRef]
  30. Poudel, B.R.; Aryal, R.L.; Bhattarai, S.; Koirala, A.R.; Gautam, S.K.; Ghimire, K.N.; Pant, B.; Park, M.; Paudyal, H.; Pokhrel, M.R. Agro-waste derived biomass impregnated with TiO2 as a potential adsorbent for removal of As(III) from water. Catalysts 2020, 10, 1125. [Google Scholar] [CrossRef]
  31. Adesina, M.O.; Block, I.; Gunter, C.; Unuabonah, E.I.; Taubert, A. Efficient removal of tetracycline and bisphenol a from water with a new hybrid clay/TiO2 composite. ACS Omega 2023, 8, 21594–21604. [Google Scholar] [CrossRef]
  32. Wang, R.; Cai, X.; Shen, F. TiO2 hollow microspheres with mesoporous surface: Superior adsorption performance for dye removal. Appl. Surf. Sci. 2014, 305, 352–358. [Google Scholar] [CrossRef]
  33. Wang, Y.; He, J.; Shi, Y.; Zhang, Y. Structure-dependent adsorptive or photocatalytic performances of solid and hollow dendritic mesoporous silica & titania nanospheres. Micropor. Mesopor. Mater. 2020, 305, 110326. [Google Scholar] [CrossRef]
  34. Su, J.; Lu, H.; Wang, F.; Lu, Y.; Zhang, R.; Xiao, J.; Cui, Y.; Zhu, L. High-efficient, broad-spectrum and recyclable mesoporous TiO2 adsorbent for water treatment. Micropor. Mesopor. Mater. 2021, 325, 111345. [Google Scholar] [CrossRef]
  35. Sattarfard, R.; Behnajady, M.A.; Eskandarloo, H. Hydrothermal synthesis of mesoporous TiO2 nanotubes and their adsorption affinity toward Basic Violet 2. J. Porous Mater. 2017, 25, 359–371. [Google Scholar] [CrossRef]
  36. Xie, G.; Wang, L.; Zhu, Q.; Chu, Z.; Tian, S.; Gui, Z.; Song, K.; Yu, Z. Regulating alkyl chain length on quaternization TiO2 for boosting photocatalytic performance: Synergism of promoting photogenerated charge separation and improving reactant adsorption. ACS Appl. Mater. Interfaces 2022, 14, 57428–57439. [Google Scholar] [CrossRef]
  37. Shabir, M.; Shezad, N.; Shafiq, I.; Maafa, I.M.; Akhter, P.; Azam, K.; Ahmed, A.; Lee, S.H.; Park, Y.-K.; Hussain, M. Carbon nanotubes loaded N,S-codoped TiO2: Heterojunction assembly for enhanced integrated adsorptive-photocatalytic performance. J. Ind. Eng. Chem. 2022, 105, 539–548. [Google Scholar] [CrossRef]
  38. Xiao, F.; Guo, X.; Li, J.; Sun, H.; Zhang, H.; Wang, W. Electrospinning preparation and dye adsorption capacity of TiO2@Carbon flexible fiber. Ceram. Int. 2019, 45, 11856–11860. [Google Scholar] [CrossRef]
  39. Bibi, S.; Bibi, F.; Amir, A.; Mazhar, D.; Waseem, M. Selective uptake of cationic dyes by microemulsion driven Ag2O/TiO2 nanocomposites. Chem. Pap. 2023, 78, 1791–1804. [Google Scholar] [CrossRef]
  40. Chen, L.; Lei, J.; Tian, L.; Lv, K.; Jin, S. Schottky construction of bimetallic Au-Cu alloy@TiO2 hollow nanoboxes embedded optical switch for enhancing photocatalytic and selective adsorption activities via one-pot deposition-precipitation strategy. J. Alloys Compd. 2023, 943, 168978. [Google Scholar] [CrossRef]
  41. Zhou, M.; Chen, J.; Hou, C.; Liu, Y.; Xu, S.; Yao, C.; Li, Z. Organic-free synthesis of porous CdS sheets with controlled windows size on bacterial cellulose for photocatalytic degradation and H2 production. App. Surf. Sci. 2019, 470, 908–916. [Google Scholar] [CrossRef]
  42. Li, T.; Liu, L.; Zhang, Z.; Han, Z. Preparation of nanofibrous metal-organic framework filter for rapid adsorption and selective separation of cationic dye from aqueous solution. Sep. Purif. Technol. 2020, 237, 116360. [Google Scholar] [CrossRef]
  43. Kumari, S.; Khan, A.A.; Chowdhury, A.; Bhakta, A.K.; Mekhalif, Z.; Hussain, S. Efficient and highly selective adsorption of cationic dyes and removal of ciprofloxacin antibiotic by surface modified nickel sulfide nanomaterials: Kinetics, isotherm and adsorption mechanism. Colloids Surf. A 2020, 586, 124264. [Google Scholar] [CrossRef]
  44. Abd-Elhamid, A.I.; Ali, H.H.; Nayl, A.A. Modification of sugarcane bagasse as a novel lignocellulosic biomass adsorbent nanocomposite to improve adsorption of methylene blue. Cellulose 2023, 30, 5239–5258. [Google Scholar] [CrossRef]
  45. Watanabe, T.; Takawane, S.; Baba, Y.; Akaiwa, J.; Kondo, A.; Ohba, T. Superior thermal stability and high photocatalytic activity of titanium dioxide nanocatalysts in carbon nanotubes. J. Phys. Chem. C 2023, 127, 16861–16869. [Google Scholar]
  46. Ye, S.; Cheng, C.; Wang, S.; Xie, R.; Wang, D. Carbon-regulated titania crystallization to construct a comprehensive palette of anatase/rutile mixed phases for advanced photocatalysis. Mater. Chem. Front. 2023, 7, 3693–3705. [Google Scholar] [CrossRef]
  47. Feng, Q.; Wu, W.; Cui, Y.; Zhou, Y.; Zhang, Y.; Xu, S.; Lin, L.; Zhou, M.; Li, Z. Reversible wettability control of self-assembled TiO2 scaffolds on bacterial cellulose from superhydrophobicity to superhydrophilicity. Cellulose 2024, 31, 2907–2920. [Google Scholar] [CrossRef]
  48. Dong, Z.; Wang, D.; Liu, X.; Pei, X.; Chen, L.; Jin, J. Bio-inspired surface-functionalization of graphene oxide for the adsorption of organic dyes and heavy metal ions with a superhigh capacity. J. Mater. Chem. A 2014, 2, 5034–5040. [Google Scholar] [CrossRef]
  49. Cao, N.; Zhao, X.; Gao, M.; Li, Z.; Ding, X.; Li, C.; Liu, K.; Du, X.; Li, W.; Feng, J.; et al. Superior selective adsorption of MgO with abundant oxygen vacancies to removal and recycle reactive dyes. Sep. Purif. Technol. 2021, 275, 119236. [Google Scholar] [CrossRef]
  50. Wang, Q.; Ju, J.; Tan, Y.; Hao, L.; Ma, Y.; Wu, Y.; Zhang, H.; Xia, Y.; Sui, K. Controlled synthesis of sodium alginate electrospun nanofiber membranes for multi-occasion adsorption and separation of methylene blue. Carbohyd. Polym. 2019, 205, 125–134. [Google Scholar] [CrossRef]
  51. Yang, Q.; Ren, S.; Zhao, Q.; Lu, R.; Hang, C.; Chen, Z.; Zheng, H. Selective separation of methyl orange from water using magnetic ZIF-67 composites. Chem. Eng. J. 2018, 333, 49–57. [Google Scholar] [CrossRef]
  52. Mohammed, N.; Lian, H.; Islam, M.S.; Strong, M.; Shi, Z.; Berry, R.M.; Yu, H.; Tam, K.C. Selective adsorption and separation of organic dyes using functionalized cellulose nanocrystals. Chem. Eng. J. 2021, 417, 129237. [Google Scholar] [CrossRef]
  53. Obayomi, K.S.; Lau, S.Y.; Danquah, M.K.; Zhang, J.; Chiong, T.; Obayomi, O.V.; Meunier, L.; Rahman, M.M. A response surface methodology approach for the removal of methylene blue dye from wastewater using sustainable and cost-effective adsorbent. Process Saf. Environ. 2024, 184, 129–150. [Google Scholar] [CrossRef]
  54. Tu, M.; Yu, J.; Wang, J.; Shi, X.; Fu, Z.; Hu, S.; Zhong, M.; Fei, Z. Coral-like TiO2/organosilane hybrid particles with rapid adsorption of methyl orange. Sep. Purif. Technol. 2022, 309, 123000. [Google Scholar] [CrossRef]
  55. Jaseela, P.K.; Garvasis, J.; Joseph, A. Selective adsorption of methylene blue (MB) dye from aqueous mixture of MB and methyl orange (MO) using mesoporous titania (TiO2)—Poly vinyl alcohol (PVA) nanocomposite. J. Mol. Liq. 2019, 286, 110908. [Google Scholar] [CrossRef]
  56. Thuy, L.T.T.; Thuy, P.T.; Anh, P.T.G.; Tien, N.A.; Bich, N.T.H.; Khieu, D.Q. Simultaneous adsorption of malachite green, methyl orange, and rhodamine B with TiO2/macadamia nutshells-derived activated carbon composite. Mater. Res. Express 2023, 10, 125602. [Google Scholar] [CrossRef]
  57. Zeng, X.; Lu, M.; Xu, Z.; Lin, Z.; Fan, F.; Xia, Y.; Li, W.; Zhang, M.; Chen, T.; Zhou, C. Facile fabrication of superhydrophilic copper foam for rapid and efficient filtration of cationic dyes from water. Sep. Purif. Technol. 2024, 354, 128943. [Google Scholar] [CrossRef]
Figure 1. In situ synthesis of 3D structured TiO2−x/CBC containing four states of (a) natural BC membranes, (b) purified BC membranes, (c) TiO2/BC after hydrolysis, and (d) TiO2−x/CBC after drying and calcination.
Figure 1. In situ synthesis of 3D structured TiO2−x/CBC containing four states of (a) natural BC membranes, (b) purified BC membranes, (c) TiO2/BC after hydrolysis, and (d) TiO2−x/CBC after drying and calcination.
Nanomaterials 15 00514 g001
Figure 2. (a) XRD patterns and (be) SEM images of TiO2/BC with various TBT concentrations; (f) XRD patterns and (gj) SEM images of TiO2−x/CBC under different calcination temperatures (Tc).
Figure 2. (a) XRD patterns and (be) SEM images of TiO2/BC with various TBT concentrations; (f) XRD patterns and (gj) SEM images of TiO2−x/CBC under different calcination temperatures (Tc).
Nanomaterials 15 00514 g002
Figure 3. (a,b) TEM images of TiO2−x/CBC-300; (c) size distribution of TiO2−x nanoparticles; (d) N2 adsorption–desorption isotherms of TiO2−x/CBC-300; XPS spectra of (e) Ti 2p and (f) O 1s of TiO2−x/CBC-300 (dot line represents the original data).
Figure 3. (a,b) TEM images of TiO2−x/CBC-300; (c) size distribution of TiO2−x nanoparticles; (d) N2 adsorption–desorption isotherms of TiO2−x/CBC-300; XPS spectra of (e) Ti 2p and (f) O 1s of TiO2−x/CBC-300 (dot line represents the original data).
Nanomaterials 15 00514 g003
Figure 4. (a) FT-IR spectra of TiO2, BC, CBC, and TiO2−x/CBC under different calcination temperatures; (b) Raman spectra of TiO2, and TiO2−x/CBC-300; (c) TG profiles of BC and TiO2−x/CBC-300; (d) UV-vis DRS spectra and (e) Kubelka–Munk plots of TiO2 and TiO2−x/CBC-300.
Figure 4. (a) FT-IR spectra of TiO2, BC, CBC, and TiO2−x/CBC under different calcination temperatures; (b) Raman spectra of TiO2, and TiO2−x/CBC-300; (c) TG profiles of BC and TiO2−x/CBC-300; (d) UV-vis DRS spectra and (e) Kubelka–Munk plots of TiO2 and TiO2−x/CBC-300.
Nanomaterials 15 00514 g004
Figure 5. (a) Single-component adsorption activities of TiO2−x/CBC-300 for five pollutants. (b) Double-component selective adsorption activities of (b) TiO2−x/CBC-200, (c) TiO2−x/CBC-300, (d) TiO2−x/CBC-400, and (e) TiO2−x/CBC-500. (f) Cycling test of TiO2−x/CBC-300 in the dark.
Figure 5. (a) Single-component adsorption activities of TiO2−x/CBC-300 for five pollutants. (b) Double-component selective adsorption activities of (b) TiO2−x/CBC-200, (c) TiO2−x/CBC-300, (d) TiO2−x/CBC-400, and (e) TiO2−x/CBC-500. (f) Cycling test of TiO2−x/CBC-300 in the dark.
Nanomaterials 15 00514 g005
Figure 6. (a) Adsorption kinetics of blank CBC, pure TiO2, and TiO2−x/CBC under different calcination temperatures, (b) pseudo-second-order model of TiO2−x/CBC-300, (c) Zeta potentials of TiO2−x/CBC adsorbents under four different calcination temperatures, (d) the separation efficiencies for MB/MO in the reported adsorbents (MAPCTs [24], NS-4 [43], SA [50], MZIF-67 [51], CNCs [52], CaO-ES@GO and CaO-FB@GO [53]), and (e) possible mechanism for the high selectivity of TiO2−x/CBC-300.
Figure 6. (a) Adsorption kinetics of blank CBC, pure TiO2, and TiO2−x/CBC under different calcination temperatures, (b) pseudo-second-order model of TiO2−x/CBC-300, (c) Zeta potentials of TiO2−x/CBC adsorbents under four different calcination temperatures, (d) the separation efficiencies for MB/MO in the reported adsorbents (MAPCTs [24], NS-4 [43], SA [50], MZIF-67 [51], CNCs [52], CaO-ES@GO and CaO-FB@GO [53]), and (e) possible mechanism for the high selectivity of TiO2−x/CBC-300.
Nanomaterials 15 00514 g006
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

Zhou, M.; Zhou, Y.; Ni, M.; Zhang, Y.; Xu, S.; Ma, H.; Zhou, J.; Zhao, J.; Lin, L.; Li, Z. In Situ Anchored, Ultrasmall, Oxygen Vacancy-Rich TiO2−x on Carbonized Bacterial Cellulose for the Efficient Adsorption and Separation of Organic Pollutants. Nanomaterials 2025, 15, 514. https://doi.org/10.3390/nano15070514

AMA Style

Zhou M, Zhou Y, Ni M, Zhang Y, Xu S, Ma H, Zhou J, Zhao J, Lin L, Li Z. In Situ Anchored, Ultrasmall, Oxygen Vacancy-Rich TiO2−x on Carbonized Bacterial Cellulose for the Efficient Adsorption and Separation of Organic Pollutants. Nanomaterials. 2025; 15(7):514. https://doi.org/10.3390/nano15070514

Chicago/Turabian Style

Zhou, Man, Yanli Zhou, Minmin Ni, Yuzhe Zhang, Song Xu, Hao Ma, Jian Zhou, Jin Zhao, Liwei Lin, and Zhongyu Li. 2025. "In Situ Anchored, Ultrasmall, Oxygen Vacancy-Rich TiO2−x on Carbonized Bacterial Cellulose for the Efficient Adsorption and Separation of Organic Pollutants" Nanomaterials 15, no. 7: 514. https://doi.org/10.3390/nano15070514

APA Style

Zhou, M., Zhou, Y., Ni, M., Zhang, Y., Xu, S., Ma, H., Zhou, J., Zhao, J., Lin, L., & Li, Z. (2025). In Situ Anchored, Ultrasmall, Oxygen Vacancy-Rich TiO2−x on Carbonized Bacterial Cellulose for the Efficient Adsorption and Separation of Organic Pollutants. Nanomaterials, 15(7), 514. https://doi.org/10.3390/nano15070514

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