Home Low-cost and large-scale preparation of ultrafine TiO2@C hybrids for high-performance degradation of methyl orange and formaldehyde under visible light
Article Open Access

Low-cost and large-scale preparation of ultrafine TiO2@C hybrids for high-performance degradation of methyl orange and formaldehyde under visible light

  • Ruhumuriza Jonathan , Shafiq Ur Rehman , Feng Cao , Hui Xu , Xuejuan Ma EMAIL logo , Junwei Wang , Yifan Liu , Yinghua Niu , Xian Jian EMAIL logo and Nasir Mahmood EMAIL logo
Published/Copyright: July 10, 2023
Become an author with De Gruyter Brill

Abstract

Developing high-efficiency and low-cost visible light photocatalyst is a great challenge for degrading both air and liquid pollutants. Herein, we developed a large-scale preparation of ultrafine TiO2@C hybrid visible light photocatalyst for high-performance degradation of formaldehyde and methyl orange (MO) at low cost using the ultra-low temperature (<200°C) air calcination method. The as-designed TiO2@C hybrids are at the scale range of 2–5 nm and modified by ultrafine carbon layers enabling the strong physical adsorption and narrowing the corresponding band gap. Specifically, the photocatalytic performance of TiO2@C hybrids for formaldehyde and MO degradation was investigated both in the air and liquid pollutant. After optimization, the TiO2@C hybrid obtained at 175°C possesses relatively better photocatalytic degradation performance than other parallel control composites under visible light irradiation. The enhanced photodegradation ability of TiO2@C-175°C hybrid with visible light response attributes to novel hybrid structure with rich defect active sites and narrow band gap (2.51 eV), favoring dual functions of physical adsorption and chemical degradation. This ultra-low temperature air calcination approach can open a low-cost and scalable pathway to design TiO2@C hybrids for green environment.

1 Introduction

Photocatalytic technology is a sustainable green technology driven by sunlight as energy, and it is one of the most promising methods to effectively solve the current energy crisis and environmental pollution. Since 1972, Honda and Fujishima introduced a new renewable energy technology, i.e., hydrogen fuel production by photocatalytic water splitting using TiO2 photocatalyst [1]. As a widely used photocatalytic material, TiO2 has attracted much attention due to its high stability, excellent photocatalytic properties, and environmentally friendly characteristics [2,3]. To fully utilize solar energy, especially visible light, various methods have been developed to modify TiO2 to overcome its own defects of poor absorption for target reactant and the electron–hole recombination, including heteroatom doping, assembly of single atom co-catalysts, and defect engineering [4,5,6]. Thus, a series of highly active TiO2 photocatalysts with visible light response are obtained [7,8 9,10,11].

Generally, doping with other materials could improve the properties of TiO2 owing to the extended light absorption, enhanced contaminant adsorption, and promoted charge-carrier separation efficiency granted by various dopants. TiO2-modified materials can be roughly divided into doping with noble metal and non-noble metal [12,13,14], metal oxides [15,16], carbonaceous materials [17,18,19,20], and dual cocatalyst [21,22]. Carbon (C) materials possess a unique property of lower charge transfer resistance and, thus, can be combined with TiO2 to extend the applications of TiO2 [18,23,24]. The TiO2@C composites can decrease the band gap of TiO2 and bring a red shift in the absorption spectra of TiO2, and this can enhance the photocatalytic activity of TiO2. For example, Zhang et al. have fabricated a TiO2@C core–shell structure composed by the novel polymerization method to degrade organic pollutants [25]. Li et al. have constructed a ternary yolk–shell structure based on TiO2 by solvothermal method to achieve high catalytic performance [26]. Due to fascinating chemical and physical properties, the TiO2@C nanocomposite has shown tremendous applications in sensors, photocatalysis, solar cells, and heterogenous catalysis [27,28]. However, the synthetic methods of TiO2 composites are multi-step, time-consuming and often require high temperatures and pressures, or protective gas, which increase the cost of large-scale industrial preparation of TiO2. Therefore, we have demonstrated a promising strategy to synthesize in situ C-doped thin layers by air-calcined pristine TiO2 at low temperature. This will improve the visible light absorption and introduced more defect sites to fully absorb contaminants. Moreover, the synthesized catalysts were applied to photocatalytic degradation pollution both in air and solution. Formaldehyde (HCHO) was chosen as the model of air pollution due to its notorious indoor pollutant that imposes great threat to human health [29,30]. Methyl orange (MO), as the model of photocatalytic degradation in solution, is widely used in several industries including the paper, printing, and food industries [31].

Herein, we aim to provide an alternative solution for the low-cost and large-scale preparation of highly active TiO2. To achieve this goal, a series of ultrafine TiO2@C hybrid materials were effectively obtained by air-calcined pristine TiO2 powder at low temperature by a one-step method. The TiO2@C hybrid obtained by calcination at the relatively lower temperature of 175°C has the better photocatalytic degradation performance under visible light irradiation. The degradation efficiency of HCHO and MO is improved up to 1.95 and 2.44 times than that of pristine TiO2, respectively. The high photodegradation ability of TiO2@C-175 under irradiation of visible light is attributed to strong physical adsorption and the reduction of the band gap energy of TiO2@C-175 to 2.52 eV. This study proposes a scalable and commercially viable strategy for environmental purification.

2 Experimental

A detailed description of characterization, active species trapping experiments, and photoelectrochemical measurements is provided in the supporting information.

2.1 Preparation of TiO2@C

First, TiO2@C7H8O powder was prepared using the sol–gel method. Generally, 6.25 mL of titanium tetrachloride (TiCl4) was slowly dissolved in 1 L of benzyl alcohol (C7H8O) to adjust the hydrolysis rate of TiCl4. The above solution was stirred for 5 days at 50°C. Then, the solid product was centrifuged and washed with alcohol for 3–4 times to remove most of the adsorbed C7H8O. A few parts of benzyl alcohol were left on the surface of TiO2 powder. The slight yellow TiO2@C7H8O powder was finally obtained by drying at 60°C for 24 h. Second, the TiO2@C composite was prepared using the chemical vapor deposition (CVD) method. In this process, 0.2 g of the as-synthesized slight yellow TiO2@C7H8O was calcined in a CVD furnace for 2 h at different temperatures (100, 150, 175, 200, 225, 250, 300°C) under air atmosphere and atmospheric pressure to obtain TiO2@C-100, TiO2@C-150, TiO2@C-175, TiO2@C-200, TiO2@C-225, TiO2@C-250, and TiO2@C-300.

2.2 Photocatalytic degradation of MO and formaldehyde

First, MO was selected as a target contaminant to evaluate the photocatalytic activity of as-prepared catalysts. Typically, 2.5 mg of photocatalyst was dispersed into 50 mL of MO solution (0.2 mol L−1). After stirring and mixing for 10 min, the suspension was irradiated under a 200 W LED lamp (Philips, China). One milliliter of the solution was collected at a certain time interval. Subsequently, the obtained solution was centrifuged, and the absorbance value of MO at 446 nm in the supernatant was recorded with a UV–Vis spectrophotometer. To examine the photocatalysis mechanism, isopropanol (IPA), ascorbic acid (AA), and triethanolamine (TEOA) were used in MO solution (2 × 10−5 mol L−1) as scavengers of hydroxyl radicals (˙OH), superoxide radicals ( ˙ O 2 ), and holes (h+). IPA, AA, and TEOA have concentrations of 2 mmol L−1 as shown in the supplementary information in Figure S8. The introduction of IPA as a ˙OH scavenger has minimal influence on the photodegradation process, whereas the presence of AA (a quencher of ˙ O 2 ) dramatically reduces the efficiency of photodegradation of MO, demonstrating that ˙ O 2 is the principal active species involved in the photodegradation process. On the other hand, the photodegradation efficiency of TiO2@C is reduced to 77.4% with the addition of TEOA as the h+ scavenger, implying that the h+ is considered a minor active species in the process. Thus, ˙ O 2 plays a major role in MO photodegradation, while h+ has a minimal impact on MO degradation when exposed to visible light.

Second, the photocatalytic degradation experiments of formaldehyde (HCHO) were carried out using a designed device as shown in Figure S1. A 10 mL and 30 g L−1 catalyst suspension was sprayed on a piece of cotton cloth, and then 1 mL, 37 wt% HCHO solution (Aladdin, Shanghai) sprayed on the surface of the cloth. The chamber is equipped with a fan, which can evenly diffuse the HCHO gas. Then, the cloth was irradiated by four 5 W LEDs, the volatilization amount of HCHO in the transparent chamber was monitored and recorded by the formaldehyde sensors in real time. We have used the visible light LED to study the photocatalytic reaction both on MO and HCHO degradation and the temperature control system was adjusted to room temperature (25°C).

3 Results and discussion

The TiO2@C hybrids were synthesized by the CVD method under air atmosphere and ambient pressure. The electronic structure of TiO2@C was changed to fully absorb visible light for efficient photocatalytic degradation of organic pollutants. First, the TiO2@C7H8O was obtained by sol–gel approach with TiCl4 as the precursor and C7H8O as solvent and carbon source. As shown in Figure 1a, the TiO2@C is heated in a CVD furnace under air atmosphere and ambient pressure, and the residual C7H8O is decomposed into highly reactive carbon and in situ deposited on the surface of the TiO2.

Figure 1 
               (a) Schematic illustration for the preparation method and photocatalytic application of TiO2@C. (b–e) Optical photos of different TiO2@C. (f) Schematic diagram of the change of TiO2 during the calcination process.
Figure 1

(a) Schematic illustration for the preparation method and photocatalytic application of TiO2@C. (b–e) Optical photos of different TiO2@C. (f) Schematic diagram of the change of TiO2 during the calcination process.

The obtained TiO2@C hybrids are shown in Figure 1b–e, TiO2@C-100 is a light brown powder, while TiO2@C-150 becomes brown. The color of TiO2@C composite becomes darker with increasing calcination temperature; it becomes dark brown for TiO2@C-200 and dark for TiO2@C-250 composite. The color of the hybrid material gradually tends to white with further increasing temperature and at 500°C; the TiO2@C composite becomes white as shown in Figure S2. This can be explained by a series of changes in the TiO2 surface during high-temperature calcination, as shown in Figure 1f; the C7H8O on the surface of TiO2 was gradually carbonized and in situ deposited on the surface to obtain uniformly C-doped TiO2. As the calcination temperature increases the C7H8O decomposition increases until it is completely carbonized, and then, the carbon on the surface is decomposed and disappeared.

XRD characterization was conducted to determine the crystalline phases of TiO2@C hybrids, and the corresponding results are shown in Figure 2a. As displayed, all the diffraction peaks of as-prepared TiO2@C quite correspond to the structure of TiO2 (PDF 21-1272), which confirms that TiO2@C7H8O and TiO2@C hybrids belong to the anatase phase. Moreover, three distinct diffraction peaks at 25.11, 37.72, and 47.69° can be clearly observed both in the XRD pattern of TiO2@C7H8O and TiO2@C hybrids. These peaks are assigned to their respective lattice planes; (101), (004) and (200) are in agreement with the previous studies [32,33]. This suggests that the crystal structure of TiO2@C7H8O is not damaged after high-temperature calcination and surface loading of carbon (C).

Figure 2 
               (a) XRD patterns and (b) Raman spectra of TiO2@C7H8O, TiO2@C-175, TiO2@C-200, and TiO2@C-225; (c) Raman spectra of TiO2@C7H8O, TiO2@C-175, TiO2@C-200, and TiO2@C-225 at 1,000–2,000 cm−1 wavenumber.
Figure 2

(a) XRD patterns and (b) Raman spectra of TiO2@C7H8O, TiO2@C-175, TiO2@C-200, and TiO2@C-225; (c) Raman spectra of TiO2@C7H8O, TiO2@C-175, TiO2@C-200, and TiO2@C-225 at 1,000–2,000 cm−1 wavenumber.

Furthermore, Raman spectroscopic techniques were adopted to analyze the detailed crystal phases and molecular structure of TiO2@C hybrids. As shown in Figure 2b, the TiO2@C7H8O and TiO2@C composites show four peaks at 154, 398, 512, and 634 cm−1, assigned to the E g, B 1g, A 1g, and E g modes of the anatase TiO2 phase, respectively [34,35]. This shows that the in situ loading of carbon atoms in TiO2@C does not significantly change the crystalline phase of TiO2. Before calcination, the TiO2@C7H8O possesses the E g in TiO2 with relatively high intensity based on Raman testing as shown in Figure 2b. After calcination, organic molecules of C7H8O were transformed into carbon layers through temperatures (175C–225°C), resulting in the formation of carbon defects. So, these carbon defects lead to a decrease in the intensity of E g peaks as shown in Figure 2b, which agrees with the literature [36,37]. The Raman peaks of the hybrids were also observed in the wavenumber range of 1,000–2,000 cm−1 as shown in Figure 2c. The two peaks centered at 1,343 and 1,580 cm−1 belong to the D and G bands of carbon, respectively, which indicates the existence of an amorphous carbon layer [38,39]. The intensity ratio of the D and G bands (I D/I G) of TiO2@C-175, TiO2@C-200, and TiO2@C-225 were 0.8312, 0.8040, and 0.8046, respectively. The increase in the I D /I G values of TiO2@C-175 compared with that of TiO2@C-200 and TiO2@C-225 reflects the introduction of more disordered sp3 defects in TiO2 after calcination at 175°C. These results indicate that the TiO2@C composites are successfully constructed. The in situ formation of C atoms on the TiO2 surface construct unsaturated sites, which are highly active for photocatalytic processes.

The corresponding morphologies of TiO2@C7H8O and TiO2@C-175 were tested by high-resolution transmission electron microscopy (HRTEM). As can be seen from Figure S3, the TiO2@C7H8O is uniform in size and ultra-fine with a diameter of about 2–5 nm. Figure 3a and b shows the clear lattice fringes of the TiO2@C7H8O with an inter-planar spacing of 0.35 nm which well agrees with that for the (101) crystallographic planes in anatase TiO2 [40]. And the HRTEM images indicate that TiO2@C7H8O composites possess the ordered lattice fringes. In contrast, Figure 3c and d shows that the lattice fringes of TiO2@C-175 clearly contain disordered layers, which are mainly due to the presence of surface C and possible defects. Meanwhile, the particle TiO2@C-175 basically has not changed in size and dispersed uniformly, as shown in Figure S4. This not only confirms that the calcination process has not caused TiO2@C7H8O to agglomerate but also proves again that the carbon on the surface of the TiO2 hybrids originated from the in situ generation during the calcination process. Furthermore, the EDS mapping analysis of TiO2@C-175 in Figure 3e–h directly reflects the distribution of Ti, O, and C elements in the TiO2@C-175 composite, which is consistent with the above results. These characterization results powerfully certify the successful construction of TiO2@C hybrids.

Figure 3 
               (a and b) TEM images of TiO2@C7H8O (c and d) TiO2@C-175. Elemental mapping of Ti (e), O (f), C (g), and C (h) elements of TiO2@C-175.
Figure 3

(a and b) TEM images of TiO2@C7H8O (c and d) TiO2@C-175. Elemental mapping of Ti (e), O (f), C (g), and C (h) elements of TiO2@C-175.

We further investigated the chemical composition and valence states of TiO2@C photocatalysts through X-ray photoelectron spectroscopy (XPS), and all the binding energy values were calibrated with respect to C 1s peak at 284.8 eV. The XPS survey spectra of prepared TiO2, TiO2@C-175, and TiO2@C-200 are shown in Figure 4a. It proves the presence of Ti, O, and C elements in TiO2@C composites, which illustrates that calcining TiO2@C7H8O in the air has not introduced other components. Furthermore, a carbon layer and defects were formed on TiO2 surface after the calcination. As given in Figure 4a, the intensity of Ti–O–C peak in TiO2@C-175 is lower than that of TiO2@C-200, indicating more carbon defect formation at 175°C to consume carbon source. Therefore, the less carbon molecule was transformed into carbon layers, resulting in the relatively low intensity of C elements in TiO2@C-175 the surface disordered C layers with black lines and possible defects with red circle lines.

Figure 4 
               (a) Full spectra and high-resolution XPS spectra of (b) Ti 2p, (c) O 1s, and (d) C 1s regions for TiO2@C7H8O, TiO2@C-175, and TiO2@C-200, respectively.
Figure 4

(a) Full spectra and high-resolution XPS spectra of (b) Ti 2p, (c) O 1s, and (d) C 1s regions for TiO2@C7H8O, TiO2@C-175, and TiO2@C-200, respectively.

The high-resolution XPS spectra of Ti 2p, O 1s, and C 1s are shown in Figure 4b–d. In the Ti 2p spectra (Figure 4b), two peaks located at 458.7 and 464.4 eV of Ti4+ in TiO2@C7H8O are attributed to Ti-2p3/2 and Ti-2p1/2, respectively [41]. After carbon loading, the peaks obviously shifted towards lower binding energies for the TiO2@C-175 and the TiO2@C-200, which reveal the formation of a junction between activated C atoms and TiO2. A possible reason might be that the electronegativity of C atoms is less than that of Ti, which increases the electron cloud density around Ti4+ on the interface and results in the decrease in the binding energy of Ti 2p [42]. As seen in Figure 4c, the calcination process significantly impacted the chemical state of O element. The O 1s XPS spectrum of TiO2@C7H8O was deconvoluted into the main peaks located at 529.8 and 531.0 eV, which can be successively ascribed to Ti–O–Ti and surface OH bonds [43,44]. After doping with carbon, the original TiO2@C7H8O and as-designed TiO2@C hybrids showed a Ti–O–C peak at 532.1 eV [20,45]. Moreover, it is found that the Ti–O–Ti bond content in TiO2@C-175 and TiO2@C-200 is relatively low, while that of the Ti–O–C in TiO2@C hybrids has an obvious increase. Especially for the TiO2@C-175, the content of Ti–O–C peaks increased significantly, indicating that more oxygen and highly active carbon were successfully introduced into the TiO2 hybrids surface [46,47]. From the C 1s spectra of TiO2@C-175 and TiO2@C-200 in Figure 4d, three peaks located at 289.2, 286.4, and 284.4 eV are clearly discerned which are assigned to Ti–O–C, C–O, and C–C, respectively [48]. These results revealed that the C is doped on the surface of TiO2 by Ti–O–C species, and more oxygen is introduced to the surface of TiO2. The result in Figure 4d indicates that TiO2@C-175 has more oxygen formation than TiO2@C-200.

To evaluate the excellent adsorption and photocatalytic properties of TiO2 hybrids both in air and solution. First, as shown in Figure S1, the as-synthesized samples were directly used to degrade HCHO as a pollutant in the photocatalytic device. The photocatalyst was put on the surface of a cotton cloth containing HCHO in a closed chamber, then the LED light was turned on the volatilization amount of HCHO in the chamber was detected after different times. To confirm the photocatalytic performance of samples, the variation in organic pollutant (MO and HCHO) concentration (C/Co%) with the irradiation time over the TiO2@C hybrids under the visible light. Co is the initial concentration; C is the present concentration related to the irradiation time The results are displayed in Figure 5a; it can be found that only a slight decrease in HCHO concentration within 8 h in the dark and blank experiments, which may be due to the escape of HCHO along the wiring pores. As for TiO2@C7H8O, the HCHO degradation rates are only 50.3% for 8 h. Compared to TiO2@C7H8O, the HCHO degradation rates of TiO2@C-175 and TiO2@C-200 are greatly improved, which can reach 98.0 and 95.9% for 8 h, respectively. Meanwhile, the HCHO degradation rate of TiO2@C-175 composites (27.3% for 2 h) is higher than that of TiO2@C-200 composites (11.1% for 2 h). It was concluded that the construction of heterojunction of active C atoms and introduction of oxygen vacancies are an effective and valuable approach to improve the photocatalytic activity of TiO2 in the visible light range. In addition, the gas-phase formaldehyde adsorption on TiO2@C-175 in both the dark and light conditions is shown in Figure S7. The results of the adsorption experiments show that there is a rapid decrease in formaldehyde concentration in the light condition compared to the dark condition after 120 min, indicating the presence of photocatalytic activity. Furthermore, we also provided data on the degradation efficiency of formaldehyde on TiO2@C-175, which is 97.1% in the light and 70.1% in the dark. This indicates that the observed degradation of formaldehyde is mainly due to the photocatalytic activity of the TiO2@C-175 sample, as the degradation efficiency in the dark is significantly lower than in the light. The photocatalysts started to degrade formaldehyde after 120 min, which was mainly caused by the heteroatom doping and defect sites. Where doping develops regions that are likely responses to visible light causing to construct active surface and accelerate the separation of electron–hole pairs, this facilitates the absorption of pollutant molecules. In other words, the introduction of C doping in our catalyst creates large surface and surface oxygen vacancies showing a significant contribution to the high photocatalytic activity [49,50,51,52].

Figure 5 
               Photocatalytic performance of (a) HCHO and (b) MO degradation of TiO2@C hybrids. (c) UV–Vis absorbance spectra in the range of 200–800 nm and (d) Tauc plot by Kubelka–Munk equation of Pristine TiO2@C7H8O, TiO2@C-150, TiO2@C-175, and TiO2@C-200.
Figure 5

Photocatalytic performance of (a) HCHO and (b) MO degradation of TiO2@C hybrids. (c) UV–Vis absorbance spectra in the range of 200–800 nm and (d) Tauc plot by Kubelka–Munk equation of Pristine TiO2@C7H8O, TiO2@C-150, TiO2@C-175, and TiO2@C-200.

Second, the degradation of MO as organic pollutants in aqueous solutions by photocatalyst also has been investigated. Figure 5b shows the dark and light after TiO2 was doped with C atom, the adsorption capacity of MO was significantly enhanced, from 18.0% for TiO2@C7H8O to 61.2% for TiO2@C-175, which indicated that the introduction of C and oxygen vacancies were beneficial to the adsorption of organic pollutants. And along with the progress of the illumination reaction, the degradation rate of MO by the hybrid catalysts also increased significantly, and the color change in MO in Figure S5 can be observed that MO is significantly degraded. And the MO photodegradation efficiencies within 60 min under the visible light were 73.3, 77.7, 68.5, and 56.5% for TiO2@C-150, TiO2@C-175, TiO2@C-200, and TiO2@C-225, respectively. The raw TiO2@C7H8O has a removal efficiency of 31.9% in the same experimental condition. Interestingly, the TiO2@C-175 sample also exhibited the excellent photocatalytic degradation efficiency for MO compared to other TiO2 hybrids. The degradation efficiencies for other TiO2 hybrid photocatalysts such as Ag/N-TiO2, N-doped TiO2/C, Nd-C TiO2, and TiO2@C/Ag are 54, 92.5, 40, and 91%, respectively [1,2,53,54]. The high degradation ability of TiO2@C-175 for MO attributes to the defect sites and heterojunction structure of TiO2 and C atoms.

We further probed the band structures of TiO2@C7H8O, TiO2@C-150, TiO2@C-175, and TiO2@C-200 heterojunctions by UV–Vis absorbance spectra. Obviously, in comparison with pristine TiO2@C7H8O, the higher light absorption edge is observed over TiO2@C hybrids, enabling TiO2@C hybrids to get higher visible light trapping capacity. The absorption edge of the TiO2@C-150, TiO2@C-175, and TiO2@C-200 is at 419, 437, and 398 nm, respectively, which have obvious redshifts compared with 374 nm of TiO2@C7H8O. Among them, TiO2@C-175 has more redshifts. The Tauc plots were obtained according to the Kubelka–Munk formula [55].

A h ν 1 / 2 = A ( h ν E g ) ,

where A is absorption value, h is Planck’s constant, ѵ is light frequency, and E g is energy band gap as illustrated in Figure 5d, the bandgap energy (E g) values of TiO2@C7H8O, TiO2@C-150, TiO2@C-175, and TiO2@C-200 are estimated to be 3.07, 2.65, 2.51, and 2.78 eV, respectively.

Furthermore, the N2 adsorption/desorption isotherms of the TiO2@C7H8O and TiO2@C-175 in the pressure range of (P/P0 = 0–1) and their surface area with the porosity of samples are shown in Figure S6 in supplementary data.

In Figure 6a, LSV measurement was taken from 0.6 to 0.9 V (vs SCE) at a scan rate of 10 mV s−1. Where the weak photocurrent response in TiO2@C7H8O is due to the wide bandgap and low visible-light absorption ability. When TiO2@C-175 composites are used, the photocurrent density significantly increases compared to TiO2@C7H8O composites. The photocurrent density of TiO2@C-175 reaches a maximum (4.39 µA cm−2) at 0.6 V (vs SCE), which is approximately 3.4 times higher than that of bare TiO2@C7H8O (1.29 µA cm−2). As a result of the addition of carbon dopant, it is confirmed that TiO2@C-175 exhibits improved current density which might be the reason for the improved photoelectrochemical performance of TiO2@C-175. The transient photocurrent responses of TiO2@C-175 and TiO2@C7H8O were tested to further understand the photocatalyst activity and photo-induced charge separation ability as shown in Figure 6b. The light was turned on and off at intervals of 30 s while being biased at 0.6 V under visible light irradiation (>446 nm). Figure 6b demonstrates that the photocurrent density of TiO2@C-175 is greater than that of TiO2@C7H8O, indicating that the TiO2@C-175 is more efficient at separating photocarrier charges. Due to the defect sites and C doping, photogenerated electrons and holes are less likely to recombine, contributing to the increased photocurrent density. The effectiveness of the separation can also be analyzed using EIS measurement. Figure 6c, which is in line with the photocurrent density result, indicates that the defect site and doping carbon effectively improve the photogenerated charge separation efficiency and interfacial transfer. Furthermore, the electronic potential of TiO2@C7H8O and TiO2@C-175 was investigated using Mott-Schottky. TiO2@C7H8O and TiO2@C-175 exhibit a positive slope as shown in Figure 6d, which verifies that all samples are n-type semiconductors, where electrons are the dominant carriers. The flat band potential (E fb) of TiO2@C7H8O and TiO2@C-175 is found to be −0.19 and −0.30 V respectively. Mott-Schotty was conducted to study the electronic potential of the TiO2@C7H8O and TiO2@C-175. A positive slope corresponds to an n-type semiconductor, whereas a negative slope corresponds to a p-type semiconductor [56]. As shown in Figure 6d, both the TiO2@C7H8O and TiO2@C-175 show the positive slope. And it confirms that all samples are n-type semiconductors, which means that the electrons are considered the major carriers [57]. Through extrapolating the linear part of 1/C s 2 plots to y = 0, the flat band potential (E fb) of the TiO2@C7H8 and TiO2@C-175 is determined as −0.19 and −0.30 V vs SCE, respectively. For TiO2@C-175, the relatively negative shift of E fb demonstrates that it improves the separation and transfer process, and it has a stronger reduction reaction ability [58].

Figure 6 
               (a) LSV curves from 0.6 to 0.9 V (vs SCE) at a scan rate of 10 mV s−1 (b) transient photocurrent responses at 0.6 V bias (vs SCE). (c) EIS Nyquist plots at 0.6 v (vs SCE) bias with the sweeping frequency from 100 kHz to 0.1 Hz. (d) Mott-Schottky plots of the TiO2@C7H8O and TiO2@C-175 at a frequency of 1,000 Hz under the visible light illumination (λ > 446).
Figure 6

(a) LSV curves from 0.6 to 0.9 V (vs SCE) at a scan rate of 10 mV s−1 (b) transient photocurrent responses at 0.6 V bias (vs SCE). (c) EIS Nyquist plots at 0.6 v (vs SCE) bias with the sweeping frequency from 100 kHz to 0.1 Hz. (d) Mott-Schottky plots of the TiO2@C7H8O and TiO2@C-175 at a frequency of 1,000 Hz under the visible light illumination (λ > 446).

4 Conclusion

In conclusion, in situ carbon-doped TiO2 nanoparticles with uniform morphology were synthesized by one-step low-temperature calcination in air. TiO2 @C hybrids were synthesized on large scale by directly calcining TiO2@C7H8O in the air in the temperature range (100–500°C). Furthermore, the TiO2@C catalysts were applied to study the photocatalytic degradation of HCHO and MO in the air and solution, respectively. When TiO2@C7H8O was calcined at 175°C for 2 h, the adsorption capacity and degradation efficiency of TiO2@C-175 hybrid catalyst were enhanced, which exhibited a high photocatalytic performance for the degradation of HCHO and MO. The HCHO removal efficiency of the TiO2@C-175 reached 98.0% under visible light for 8 h, which was 1.95 times higher than TiO2@C7H8O powder (50.3%). The MO removal efficiency of the TiO2@C-175 reached 77.7% under visible light for 60 min, which was 2.44 times higher than the TiO2@C7H8O powder (31.9%). The high photodegradation ability of TiO2@C-175 under irradiation of visible light attributes to the more defect sites, heterojunction structure of TiO2 and C atoms, and the reduction of the band gap energy of TiO2@C-175 to 2.51 eV. This study provides a cost-effective and facile approach to preparing the TiO2 photocatalysts, which can be feasibly scaled up for industrial applications.

  1. Funding information: This work was financially supported by the Fundamental Research Funds for the Chinese Central Universities, China (No. ZYGX2019J025), Sichuan Science and Technology Program (No. 2022NSFSC1087, No. 2021YFG0373), Sichuan Science and Technology Program (2022YFSY0040), and Huzhou Sci-tech Special Commissioner project (2021KT54).

  2. Author contributions: All authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Conflict of interest: The authors state no conflict of interest.

References

[1] Fujishima A, Honda K. Electrochemical photolysis of water at a semiconductor electrode. Nature. 1972;238:37–8.10.1038/238037a0Search in Google Scholar PubMed

[2] Ruan X, Cui X, Cui Y, Fan X, Li Z, Xie T, et al. Favorable energy band alignment of TiO2 anatase/rutile heterophase homojunctions yields photocatalytic hydrogen evolution with quantum efficiency exceeding 45.6%. Adv Energy Mater. 2022;12:1–9.10.1002/aenm.202200298Search in Google Scholar

[3] Chen J, Kang Y, Zhang W, Zhang Z, Chen Y, Yang Y, et al. Lattice-confined single-atom Fe1Sx on mesoporous TiO2 for boosting ambient electrocatalytic N2 reduction reaction. Angew Chem Int Ed. 2022;51900:1–9.Search in Google Scholar

[4] Yuan W, Cheng L, An Y, Lv S, Wu H, Fan X, et al. Laminated hybrid junction of sulfur-doped TiO2 and a carbon substrate derived from Ti3C2 MXenes: Toward highly visible Light-driven photocatalytic hydrogen evolution. Adv Sci. 2018;5:1700870–80.10.1002/advs.201700870Search in Google Scholar PubMed PubMed Central

[5] Lee BH, Park S, Kim M, Sinha AK, Lee SC, Jung E, et al. Reversible and cooperative photoactivation of single-atom Cu/TiO2 photocatalysts. Nat Mater. 2019;18:620–6.10.1038/s41563-019-0344-1Search in Google Scholar PubMed

[6] Han Q, Wu C, Jiao H, Xu R, Wang Y, Xie J, et al. Rational design of high-concentration Ti3 + in porous carbon-doped TiO2 nanosheets for efficient photocatalytic ammonia synthesis. Adv Mater. 2021;33:1–9.10.1002/adma.202008180Search in Google Scholar PubMed

[7] Rengifo-Herrera JA, Osorio-Vargas P, Pulgarin C. A critical review on N-modified TiO2 limits to treat chemical and biological contaminants in water. Evidence that enhanced visible light absorption does not lead to higher degradation rates under whole solar light. J Hazard Mater. 2022;425:127979.10.1016/j.jhazmat.2021.127979Search in Google Scholar PubMed

[8] Qi MY, Lin Q, Tang ZR, Xu YJ. Photoredox coupling of benzyl alcohol oxidation with CO2 reduction over CdS/TiO2 heterostructure under visible light irradiation. Appl Catal B Env. 2022;307:121158–66.10.1016/j.apcatb.2022.121158Search in Google Scholar

[9] Huy TH, Phat BD, Kang F, Wang YF, Liu SH, Thi CM, et al. SnO2/TiO2 nanotube heterojunction: The first investigation of NO degradation by visible light-driven photocatalysis. Chemosphere. 2019;215:323–32.10.1016/j.chemosphere.2018.10.033Search in Google Scholar PubMed

[10] Liu SH, Lin WX. A simple method to prepare g-C3N4-TiO2/waste zeolites as visible-light-responsive photocatalytic coatings for degradation of indoor formaldehyde. J Hazard Mater. 2019;368:468–76.10.1016/j.jhazmat.2019.01.082Search in Google Scholar PubMed

[11] Wu Q, Ye J, Qiao W, Li Y, Niemantsverdriet JWH, Richards E, et al. Inhibit the formation of toxic methylphenolic by-products in photo-decomposition of formaldehyde–toluene/xylene mixtures by Pd cocatalyst on TiO2. Appl Catal B. 2021;291:120118–28.10.1016/j.apcatb.2021.120118Search in Google Scholar

[12] Zhang Z, Wu Q, Johnson G, Ye Y, Li X, Li N, et al. A Generalized synthetic strategy for transition metal doped brookite-phase TiO2 nanorods. J Am Chem Soc. 2019;141:16548–52.10.1021/jacs.9b06389Search in Google Scholar PubMed

[13] Wang H, Song L, Yu L, Xia X, Bao Y, Lourenco M, et al. Charge transfer between Ti4 +, Sn4 + and Pt in the tin doped TiO2 photocatalyst for elevating the hydrogen production efficiency. Appl Surf Sci. 2022;581:152202–13.10.1016/j.apsusc.2021.152202Search in Google Scholar

[14] Zhou J, Gao Z, Xiang G, Zhai T, Liu Z, Zhao W, et al. Interfacial compatibility critically controls Ru/TiO2 metal-support interaction modes in CO2 hydrogenation. Nat Commun. 2022;13:1–10.10.1038/s41467-021-27910-4Search in Google Scholar PubMed PubMed Central

[15] Li S, Cai J, Wu X, Zheng F. Sandwich-like TiO2@ZnO-based noble metal (Ag, Au, Pt, or Pd) for better photo-oxidation performance: Synergistic effect between noble metal and metal oxide phases. Appl Surf Sci. 2018;443:603–12.10.1016/j.apsusc.2018.03.017Search in Google Scholar

[16] Navarra W, Ritacco I, Sacco O, Caporaso L, Farnesi Camellone M, Venditto V, et al. Density functional theory study and photocatalytic activity of ZnO/N-doped TiO2 heterojunctions. J Phys Chem C. 2022;126:7000–11.10.1021/acs.jpcc.2c00152Search in Google Scholar

[17] Padmanabhan NT, Thomas N, Louis J, Mathew DT, Ganguly P, John H, et al. Graphene coupled TiO2 photocatalysts for environmental applications: A review. Chemosphere. 2021;271:129506–39.10.1016/j.chemosphere.2020.129506Search in Google Scholar PubMed

[18] Jiao W, Zhang L, Yang R, Ning J, Xiao L, Liu Y, et al. Synthesis of monolayer carbon-coated TiO2 as visible-light-responsive photocatalysts. Appl Mater Today. 2022;27:101498–508.10.1016/j.apmt.2022.101498Search in Google Scholar

[19] Xiong H, Fu J, Li J, Ali R, Wang H, Liu Y, et al. Strain-regulated sensing properties of α-Fe2O3 nano-cylinders with atomic carbon layers for ethanol detection. J Mater Sci Technol. 2021;68:132–9.10.1016/j.jmst.2020.08.009Search in Google Scholar

[20] Zhang Y, Hawboldt K, Zhang L, Lu J, Chang L, Dwyer A. Carbonaceous nanomaterial-TiO2 heterojunctions for visible-light-driven photocatalytic degradation of aqueous organic pollutants. Appl Catal A Gen. 2022;630:118460–75.10.1016/j.apcata.2021.118460Search in Google Scholar

[21] Meng A, Zhang L, Cheng B, Yu J. Dual cocatalysts in TiO2 photocatalysis. Adv Mater. 2019;31:1–31.10.1002/adma.201804710Search in Google Scholar PubMed

[22] Barba-Nieto I, Colón G, Fernández-García M, Kubacka A. Shepherding reaction intermediates to optimize H2 yield using composite-doped TiO2-based photocatalysts. Chem Eng J. 2022;442:136333–46.10.1016/j.cej.2022.136333Search in Google Scholar

[23] Ning J, Mu C, Guo X, Yang R, Jonathan R, Jiao W, et al. Efficient defect engineering and in-situ carbon doping in ultra-fine TiO2 with enhanced visible-light-response photocatalytic performance. J Alloy Compd. 2022;901:163490–500.10.1016/j.jallcom.2021.163490Search in Google Scholar

[24] Justh N, Mikula GJ, Bakos LP, Nagy B, László K, Parditka B, et al. Photocatalytic properties of TiO2@polymer and TiO2@carbon aerogel composites prepared by atomic layer deposition. Carbon. 2019;147:476–82.10.1016/j.carbon.2019.02.076Search in Google Scholar

[25] Zhang J, Vasei M, Sang Y, Liu H, Claverie JP. TiO2@carbon photocatalysts: The effect of carbon thickness on catalysis. ACS Appl Mater Interfaces. 2016;8:1903–12.10.1021/acsami.5b10025Search in Google Scholar PubMed

[26] Li Y, Xue J, Shen Q, Jia S, Li Q, Li Y, et al. Construction of a ternary spatial junction in yolk–shell nanoreactor for efficient photo-thermal catalytic hydrogen generation. Chem Eng J. 2021;423:16–8.10.1016/j.cej.2021.130188Search in Google Scholar

[27] Kholodnaya G, Sazonov R, Ponomarev D. TiO2@ C nanocomposites–from synthesis to application: A review. Fuller Nanotub. 2021;29:487–526.10.1080/1536383X.2020.1820994Search in Google Scholar

[28] Matos J, Miranda C, Poon PS, Mansilla HD. Nanostructured hybrid TiO2-C for the photocatalytic conversion of phenol. J Sol Energy Eng. 2016;134:64–71.10.1016/j.solener.2016.04.043Search in Google Scholar

[29] Li X, Fang G, Qian X, Tian Q. Z-scheme heterojunction of low conduction band potential MnO2 and biochar-based g-C3N4 for efficient formaldehyde degradation. Chem Eng J. 2022;428:131052–61.10.1016/j.cej.2021.131052Search in Google Scholar

[30] Li X, Li H, Huang Y, Cao J, Huang T, Li R, et al. Exploring the photocatalytic conversion mechanism of gaseous formaldehyde degradation on TiO2–x-OV surface. J Hazard Mater. 2022;424:127217–27.10.1016/j.jhazmat.2021.127217Search in Google Scholar PubMed

[31] Yao X, Zhang B, Cui S, Yang S, Tang X. Fabrication of SnSO4-modified TiO2 for enhance degradation performance of methyl orange (MO) and antibacterial activity. Appl Surf Sci. 2021;551:149419–38.10.1016/j.apsusc.2021.149419Search in Google Scholar

[32] Shao Z, Zhang D, Li H, Su C, Pu X, Geng Y. Fabrication of MIL-88A/g-C3N4 direct Z-scheme heterojunction with enhanced visible-light photocatalytic activity. Sep Purif Technol. 2019;220:16–24.10.1016/j.seppur.2019.03.040Search in Google Scholar

[33] Zhang Y, Zhou J, Chen X, Wang L, Cai W. Coupling of heterogeneous advanced oxidation processes and photocatalysis in efficient degradation of tetracycline hydrochloride by Fe-based MOFs: Synergistic effect and degradation pathway. Chem Eng J. 2019;369:745–57.10.1016/j.cej.2019.03.108Search in Google Scholar

[34] Naik KM, Higuchi E, Inoue H. Pt nanoparticle-decorated two-dimensional oxygen-deficient TiO2 nanosheets as an efficient and stable electrocatalyst for the hydrogen evolution reaction. Nanoscale. 2020;12:11055–62.10.1039/D0NR02092CSearch in Google Scholar PubMed

[35] Yao S, Ma Y, Xu T, Wang Z, Lv P, Zheng J, et al. Ti–C bonds reinforced TiO2@C nanocomposite Na-ion battery electrodes by fluidized-bed plasma-enhanced chemical vapor deposition. Carbon. 2021;171:524–31.10.1016/j.carbon.2020.09.023Search in Google Scholar

[36] Lv C, Qian Y, Yan C, Ding Y, Liu Y, Chen G, et al. Defect engineering metal-free polymeric carbon nitride electrocatalyst for effective nitrogen fixation under ambient conditions. Angew Chem Int Ed Engl. 2018;57:10403–7.10.1002/ange.201806386Search in Google Scholar

[37] Yu P, Tang W, Wu FF, Zhang C, Luo HY, Liu H, et al. Recent progress in plant-derived hard carbon anode materials for sodium-ion batteries a review. Rare Met. 2020;39:1019–33.10.1007/s12598-020-01443-zSearch in Google Scholar

[38] Presumido PH, do Santos LF, Neuparth T, Santos MM, Feliciano M, Primo A, et al. A Novel ceramic tubular membrane coated with a continuous graphene-TiO2 nanocomposite thin-film for CECs mitigation. Chem Eng J. 2022;430:132639–53.10.1016/j.cej.2021.132639Search in Google Scholar

[39] Kisielewska A, Spilarewicz-Stanek K, Cichomski M, Kozłowski W, Piwoński I. The role of graphene oxide and its reduced form in the in situ photocatalytic growth of silver nanoparticles on graphene-TiO2 nanocomposites. Appl Surf Sci. 2022;576:151759–72.10.1016/j.apsusc.2021.151759Search in Google Scholar

[40] Chen Y, Wang Y, Li W, Yang Q, Hou Q, Wei L, et al. Enhancement of photocatalytic performance with the use of noble-metal-decorated TiO2 nanocrystals as highly active catalysts for aerobic oxidation under visible-light irradiation. Appl Catal B Env. 2017;210:352–67.10.1016/j.apcatb.2017.03.077Search in Google Scholar

[41] Yuan L, Lu KQ, Zhang F, Fu X, Xu YJ. Unveiling the interplay between light-driven CO2 photocatalytic reduction and carbonaceous residues decomposition: A case study of Bi2WO6-TiO2 binanosheets. Appl Catal B Env. 2018;237:424–31.10.1016/j.apcatb.2018.06.019Search in Google Scholar

[42] Li Y, Shen J, Li J, Liu S, Yu D, Xu R, et al. Constructing a novel strategy for carbon-doped TiO2 multiple-phase nanocomposites toward superior electrochemical performance for lithium ion batteries and the hydrogen evolution reaction. J Mater Chem A. 2017;5:7055–63.10.1039/C7TA01184ASearch in Google Scholar

[43] Wang W, Wu M, Han P, Liu Y, He L, Huang Q, et al. Understanding the behavior and mechanism of oxygen-deficient anatase TiO2 toward sodium storage. ACS Appl Mater Interfaces. 2019;11:3061–9.10.1021/acsami.8b19288Search in Google Scholar PubMed

[44] Li Y, Zhang H, Liu P, Wang D, Li Y, Zhao H. Cross-linked g-C3N4/rGO nanocomposites with tunable band structure and enhanced visible light photocatalytic activity. Small. 2013;9:3336–44.10.1002/smll.201203135Search in Google Scholar PubMed

[45] Cong Y, Li X, Qin Y, Dong Z, Yuan G, Cui Z, et al. Carbon-doped TiO2 coating on multiwalled carbon nanotubes with higher visible light photocatalytic activity. Appl Catal B Env. 2011;107:128–34.10.1016/j.apcatb.2011.07.005Search in Google Scholar

[46] Wang J, Sun Y, Fu L, Sun Z, Ou M, Zhao S, et al. A defective g-C3N4/RGO/TiO2composite from hydrogen treatment for enhanced visible-light photocatalytic H2 production. Nanoscale. 2020;12:22030–5.10.1039/D0NR05141ASearch in Google Scholar

[47] Chen M, Wang W, Qiu Y, Wen H, Li G, Yang Z, et al. Identification of active sites in HCHO oxidation over TiO2‑Supported Pt catalysts. ACS Catal. 2022;12:5565–73.10.1021/acscatal.2c01150Search in Google Scholar

[48] Zou Y, Shi JW, Ma D, Fan Z, Lu L, Niu C. In situ synthesis of C-doped TiO2@g-C3N4 core-shell hollow nanospheres with enhanced visible-light photocatalytic activity for H2 evolution. Chem Eng J. 2017;322:435–44.10.1016/j.cej.2017.04.056Search in Google Scholar

[49] Li K, Zhou W, Li X, Li Q, Carabineiro SA, Zhang S, et al. Synergistic effect of cyano defects and CaCO3 in graphitic carbon nitride nanosheets for efficient visible-light-driven photocatalytic NO removal. J Hazard Mater. 2023;442:130040–53.10.1016/j.jhazmat.2022.130040Search in Google Scholar PubMed

[50] Kang N, Zhou W, Qi Z, Li Y, Wang Z, Li Q, et al. Recent progress of natural mineral materials in environmental remediation. Catalysts. 2022;12(9):996.10.3390/catal12090996Search in Google Scholar

[51] Gao J, Xue J, Jia S, Shen Q, Zhang X, Jia H, et al. Self-doping surface oxygen vacancy-induced lattice strains for enhancing visible light-driven photocatalytic H2 evolution over black TiO2. ACS Appl Mater Interface. 2021;13(16):18758–71.10.1021/acsami.1c01101Search in Google Scholar PubMed

[52] Lin T, Yang C, Wang Z, Yin H, Lü X, Huang F, et al. Effective nonmetal incorporation in black titania with enhanced solar energy utilization. Energy Env Sci. 2014;7(3):967–72.10.1039/c3ee42708kSearch in Google Scholar

[53] Lotfi S, Fischer K, Schulze A, Schäfer AI. Photocatalytic degradation of steroid hormone micropollutants by TiO2-coated polyethersulfone membranes in a continuous flow-through process. Nat Nanotechnol. 2022;17:417–23.10.1038/s41565-022-01074-8Search in Google Scholar PubMed

[54] Gao X, Zheng K, Zhang Q, Cao X, Wu S, Su J. Self-assembly TiO2-RGO/LDHs nanocomposite: Photocatalysis of VOCs degradation in simulation air. Appl Surf Sci. 2022;586:152882–92.10.1016/j.apsusc.2022.152882Search in Google Scholar

[55] Zhao K, Lu Y, Lu N, Zhao Y, Yuan X, Zhang H, et al. Design of H3PW12O40/TiO2 nano-photocatalyst for efficient photocatalysis under simulated sunlight irradiation. Appl Surf Sci. 2013;285:616–24.10.1016/j.apsusc.2013.08.101Search in Google Scholar

[56] Xiao L, Liu T, Zhang M, Li Q, Yang J. Interfacial construction of zero-dimensional/one-dimensional g-C3N4 nanoparticles/TiO2 nanotube arrays with Z-scheme heterostructure for improved photoelectrochemical water splitting. ACS Sustain Chem Eng. 2018;7(2):2483–91.10.1021/acssuschemeng.8b05392Search in Google Scholar

[57] Wang J, Xue C, Yao W, Liu J, Gao X, Zong R, et al. MOF-derived hollow TiO2@C/FeTiO3 nanoparticles as photoanodes with enhanced full spectrum light PEC activities. Appl Catal B. 2019;250:369–81.10.1016/j.apcatb.2019.03.002Search in Google Scholar

[58] Feng C, Tang L, Deng Y, Wang J, Liu Y, Ouyang X, et al. A novel sulfur-assisted annealing method of g-C3N4 nanosheet compensates for the loss of light absorption with further promoted charge transfer for photocatalytic production of H2 and H2O2. Appl Catal B. 2021;281:119539.10.1016/j.apcatb.2020.119539Search in Google Scholar

Received: 2022-10-31
Revised: 2023-04-08
Accepted: 2023-05-11
Published Online: 2023-07-10

© 2023 the author(s), published by De Gruyter

This work is licensed under the Creative Commons Attribution 4.0 International License.

Articles in the same Issue

  1. Research Articles
  2. Preparation of CdS–Ag2S nanocomposites by ultrasound-assisted UV photolysis treatment and its visible light photocatalysis activity
  3. Significance of nanoparticle radius and inter-particle spacing toward the radiative water-based alumina nanofluid flow over a rotating disk
  4. Aptamer-based detection of serotonin based on the rapid in situ synthesis of colorimetric gold nanoparticles
  5. Investigation of the nucleation and growth behavior of Ti2AlC and Ti3AlC nano-precipitates in TiAl alloys
  6. Dynamic recrystallization behavior and nucleation mechanism of dual-scale SiCp/A356 composites processed by P/M method
  7. High mechanical performance of 3-aminopropyl triethoxy silane/epoxy cured in a sandwich construction of 3D carbon felts foam and woven basalt fibers
  8. Applying solution of spray polyurea elastomer in asphalt binder: Feasibility analysis and DSR study based on the MSCR and LAS tests
  9. Study on the chronic toxicity and carcinogenicity of iron-based bioabsorbable stents
  10. Influence of microalloying with B on the microstructure and properties of brazed joints with Ag–Cu–Zn–Sn filler metal
  11. Thermohydraulic performance of thermal system integrated with twisted turbulator inserts using ternary hybrid nanofluids
  12. Study of mechanical properties of epoxy/graphene and epoxy/halloysite nanocomposites
  13. Effects of CaO addition on the CuW composite containing micro- and nano-sized tungsten particles synthesized via aluminothermic coupling with silicothermic reduction
  14. Cu and Al2O3-based hybrid nanofluid flow through a porous cavity
  15. Design of functional vancomycin-embedded bio-derived extracellular matrix hydrogels for repairing infectious bone defects
  16. Study on nanocrystalline coating prepared by electro-spraying 316L metal wire and its corrosion performance
  17. Axial compression performance of CFST columns reinforced by ultra-high-performance nano-concrete under long-term loading
  18. Tungsten trioxide nanocomposite for conventional soliton and noise-like pulse generation in anomalous dispersion laser cavity
  19. Microstructure and electrical contact behavior of the nano-yttria-modified Cu-Al2O3/30Mo/3SiC composite
  20. Melting rheology in thermally stratified graphene-mineral oil reservoir (third-grade nanofluid) with slip condition
  21. Re-examination of nonlinear vibration and nonlinear bending of porous sandwich cylindrical panels reinforced by graphene platelets
  22. Parametric simulation of hybrid nanofluid flow consisting of cobalt ferrite nanoparticles with second-order slip and variable viscosity over an extending surface
  23. Chitosan-capped silver nanoparticles with potent and selective intrinsic activity against the breast cancer cells
  24. Multi-core/shell SiO2@Al2O3 nanostructures deposited on Ti3AlC2 to enhance high-temperature stability and microwave absorption properties
  25. Solution-processed Bi2S3/BiVO4/TiO2 ternary heterojunction photoanode with enhanced photoelectrochemical performance
  26. Electroporation effect of ZnO nanoarrays under low voltage for water disinfection
  27. NIR-II window absorbing graphene oxide-coated gold nanorods and graphene quantum dot-coupled gold nanorods for photothermal cancer therapy
  28. Nonlinear three-dimensional stability characteristics of geometrically imperfect nanoshells under axial compression and surface residual stress
  29. Investigation of different nanoparticles properties on the thermal conductivity and viscosity of nanofluids by molecular dynamics simulation
  30. Optimized Cu2O-{100} facet for generation of different reactive oxidative species via peroxymonosulfate activation at specific pH values to efficient acetaminophen removal
  31. Brownian and thermal diffusivity impact due to the Maxwell nanofluid (graphene/engine oil) flow with motile microorganisms and Joule heating
  32. Appraising the dielectric properties and the effectiveness of electromagnetic shielding of graphene reinforced silicone rubber nanocomposite
  33. Synthesis of Ag and Cu nanoparticles by plasma discharge in inorganic salt solutions
  34. Low-cost and large-scale preparation of ultrafine TiO2@C hybrids for high-performance degradation of methyl orange and formaldehyde under visible light
  35. Utilization of waste glass with natural pozzolan in the production of self-glazed glass-ceramic materials
  36. Mechanical performance of date palm fiber-reinforced concrete modified with nano-activated carbon
  37. Melting point of dried gold nanoparticles prepared with ultrasonic spray pyrolysis and lyophilisation
  38. Graphene nanofibers: A modern approach towards tailored gypsum composites
  39. Role of localized magnetic field in vortex generation in tri-hybrid nanofluid flow: A numerical approach
  40. Intelligent computing for the double-diffusive peristaltic rheology of magneto couple stress nanomaterials
  41. Bioconvection transport of upper convected Maxwell nanoliquid with gyrotactic microorganism, nonlinear thermal radiation, and chemical reaction
  42. 3D printing of porous Ti6Al4V bone tissue engineering scaffold and surface anodization preparation of nanotubes to enhance its biological property
  43. Bioinspired ferromagnetic CoFe2O4 nanoparticles: Potential pharmaceutical and medical applications
  44. Significance of gyrotactic microorganisms on the MHD tangent hyperbolic nanofluid flow across an elastic slender surface: Numerical analysis
  45. Performance of polycarboxylate superplasticisers in seawater-blended cement: Effect from chemical structure and nano modification
  46. Entropy minimization of GO–Ag/KO cross-hybrid nanofluid over a convectively heated surface
  47. Oxygen plasma assisted room temperature bonding for manufacturing SU-8 polymer micro/nanoscale nozzle
  48. Performance and mechanism of CO2 reduction by DBD-coupled mesoporous SiO2
  49. Polyarylene ether nitrile dielectric films modified by HNTs@PDA hybrids for high-temperature resistant organic electronics field
  50. Exploration of generalized two-phase free convection magnetohydrodynamic flow of dusty tetra-hybrid Casson nanofluid between parallel microplates
  51. Hygrothermal bending analysis of sandwich nanoplates with FG porous core and piezomagnetic faces via nonlocal strain gradient theory
  52. Design and optimization of a TiO2/RGO-supported epoxy multilayer microwave absorber by the modified local best particle swarm optimization algorithm
  53. Mechanical properties and frost resistance of recycled brick aggregate concrete modified by nano-SiO2
  54. Self-template synthesis of hollow flower-like NiCo2O4 nanoparticles as an efficient bifunctional catalyst for oxygen reduction and oxygen evolution in alkaline media
  55. High-performance wearable flexible strain sensors based on an AgNWs/rGO/TPU electrospun nanofiber film for monitoring human activities
  56. High-performance lithium–selenium batteries enabled by nitrogen-doped porous carbon from peanut meal
  57. Investigating effects of Lorentz forces and convective heating on ternary hybrid nanofluid flow over a curved surface using homotopy analysis method
  58. Exploring the potential of biogenic magnesium oxide nanoparticles for cytotoxicity: In vitro and in silico studies on HCT116 and HT29 cells and DPPH radical scavenging
  59. Enhanced visible-light-driven photocatalytic degradation of azo dyes by heteroatom-doped nickel tungstate nanoparticles
  60. A facile method to synthesize nZVI-doped polypyrrole-based carbon nanotube for Ag(i) removal
  61. Improved osseointegration of dental titanium implants by TiO2 nanotube arrays with self-assembled recombinant IGF-1 in type 2 diabetes mellitus rat model
  62. Functionalized SWCNTs@Ag–TiO2 nanocomposites induce ROS-mediated apoptosis and autophagy in liver cancer cells
  63. Triboelectric nanogenerator based on a water droplet spring with a concave spherical surface for harvesting wave energy and detecting pressure
  64. A mathematical approach for modeling the blood flow containing nanoparticles by employing the Buongiorno’s model
  65. Molecular dynamics study on dynamic interlayer friction of graphene and its strain effect
  66. Induction of apoptosis and autophagy via regulation of AKT and JNK mitogen-activated protein kinase pathways in breast cancer cell lines exposed to gold nanoparticles loaded with TNF-α and combined with doxorubicin
  67. Effect of PVA fibers on durability of nano-SiO2-reinforced cement-based composites subjected to wet-thermal and chloride salt-coupled environment
  68. Effect of polyvinyl alcohol fibers on mechanical properties of nano-SiO2-reinforced geopolymer composites under a complex environment
  69. In vitro studies of titanium dioxide nanoparticles modified with glutathione as a potential drug delivery system
  70. Comparative investigations of Ag/H2O nanofluid and Ag-CuO/H2O hybrid nanofluid with Darcy-Forchheimer flow over a curved surface
  71. Study on deformation characteristics of multi-pass continuous drawing of micro copper wire based on crystal plasticity finite element method
  72. Properties of ultra-high-performance self-compacting fiber-reinforced concrete modified with nanomaterials
  73. Prediction of lap shear strength of GNP and TiO2/epoxy nanocomposite adhesives
  74. A novel exploration of how localized magnetic field affects vortex generation of trihybrid nanofluids
  75. Fabrication and physicochemical characterization of copper oxide–pyrrhotite nanocomposites for the cytotoxic effects on HepG2 cells and the mechanism
  76. Thermal radiative flow of cross nanofluid due to a stretched cylinder containing microorganisms
  77. In vitro study of the biphasic calcium phosphate/chitosan hybrid biomaterial scaffold fabricated via solvent casting and evaporation technique for bone regeneration
  78. Insights into the thermal characteristics and dynamics of stagnant blood conveying titanium oxide, alumina, and silver nanoparticles subject to Lorentz force and internal heating over a curved surface
  79. Effects of nano-SiO2 additives on carbon fiber-reinforced fly ash–slag geopolymer composites performance: Workability, mechanical properties, and microstructure
  80. Energy bandgap and thermal characteristics of non-Darcian MHD rotating hybridity nanofluid thin film flow: Nanotechnology application
  81. Green synthesis and characterization of ginger-extract-based oxali-palladium nanoparticles for colorectal cancer: Downregulation of REG4 and apoptosis induction
  82. Abnormal evolution of resistivity and microstructure of annealed Ag nanoparticles/Ag–Mo films
  83. Preparation of water-based dextran-coated Fe3O4 magnetic fluid for magnetic hyperthermia
  84. Statistical investigations and morphological aspects of cross-rheological material suspended in transportation of alumina, silica, titanium, and ethylene glycol via the Galerkin algorithm
  85. Effect of CNT film interleaves on the flexural properties and strength after impact of CFRP composites
  86. Self-assembled nanoscale entities: Preparative process optimization, payload release, and enhanced bioavailability of thymoquinone natural product
  87. Structure–mechanical property relationships of 3D-printed porous polydimethylsiloxane films
  88. Nonlinear thermal radiation and the slip effect on a 3D bioconvection flow of the Casson nanofluid in a rotating frame via a homotopy analysis mechanism
  89. Residual mechanical properties of concrete incorporated with nano supplementary cementitious materials exposed to elevated temperature
  90. Time-independent three-dimensional flow of a water-based hybrid nanofluid past a Riga plate with slips and convective conditions: A homotopic solution
  91. Lightweight and high-strength polyarylene ether nitrile-based composites for efficient electromagnetic interference shielding
  92. Review Articles
  93. Recycling waste sources into nanocomposites of graphene materials: Overview from an energy-focused perspective
  94. Hybrid nanofiller reinforcement in thermoset and biothermoset applications: A review
  95. Current state-of-the-art review of nanotechnology-based therapeutics for viral pandemics: Special attention to COVID-19
  96. Solid lipid nanoparticles for targeted natural and synthetic drugs delivery in high-incidence cancers, and other diseases: Roles of preparation methods, lipid composition, transitional stability, and release profiles in nanocarriers’ development
  97. Critical review on experimental and theoretical studies of elastic properties of wurtzite-structured ZnO nanowires
  98. Polyurea micro-/nano-capsule applications in construction industry: A review
  99. A comprehensive review and clinical guide to molecular and serological diagnostic tests and future development: In vitro diagnostic testing for COVID-19
  100. Recent advances in electrocatalytic oxidation of 5-hydroxymethylfurfural to 2,5-furandicarboxylic acid: Mechanism, catalyst, coupling system
  101. Research progress and prospect of silica-based polymer nanofluids in enhanced oil recovery
  102. Review of the pharmacokinetics of nanodrugs
  103. Engineered nanoflowers, nanotrees, nanostars, nanodendrites, and nanoleaves for biomedical applications
  104. Research progress of biopolymers combined with stem cells in the repair of intrauterine adhesions
  105. Progress in FEM modeling on mechanical and electromechanical properties of carbon nanotube cement-based composites
  106. Antifouling induced by surface wettability of poly(dimethyl siloxane) and its nanocomposites
  107. TiO2 aerogel composite high-efficiency photocatalysts for environmental treatment and hydrogen energy production
  108. Structural properties of alumina surfaces and their roles in the synthesis of environmentally persistent free radicals (EPFRs)
  109. Nanoparticles for the potential treatment of Alzheimer’s disease: A physiopathological approach
  110. Current status of synthesis and consolidation strategies for thermo-resistant nanoalloys and their general applications
  111. Recent research progress on the stimuli-responsive smart membrane: A review
  112. Dispersion of carbon nanotubes in aqueous cementitious materials: A review
  113. Applications of DNA tetrahedron nanostructure in cancer diagnosis and anticancer drugs delivery
  114. Magnetic nanoparticles in 3D-printed scaffolds for biomedical applications
  115. An overview of the synthesis of silicon carbide–boron carbide composite powders
  116. Organolead halide perovskites: Synthetic routes, structural features, and their potential in the development of photovoltaic
  117. Recent advancements in nanotechnology application on wood and bamboo materials: A review
  118. Application of aptamer-functionalized nanomaterials in molecular imaging of tumors
  119. Recent progress on corrosion mechanisms of graphene-reinforced metal matrix composites
  120. Research progress on preparation, modification, and application of phenolic aerogel
  121. Application of nanomaterials in early diagnosis of cancer
  122. Plant mediated-green synthesis of zinc oxide nanoparticles: An insight into biomedical applications
  123. Recent developments in terahertz quantum cascade lasers for practical applications
  124. Recent progress in dielectric/metal/dielectric electrodes for foldable light-emitting devices
  125. Nanocoatings for ballistic applications: A review
  126. A mini-review on MoS2 membrane for water desalination: Recent development and challenges
  127. Recent updates in nanotechnological advances for wound healing: A narrative review
  128. Recent advances in DNA nanomaterials for cancer diagnosis and treatment
  129. Electrochemical micro- and nanobiosensors for in vivo reactive oxygen/nitrogen species measurement in the brain
  130. Advances in organic–inorganic nanocomposites for cancer imaging and therapy
  131. Advancements in aluminum matrix composites reinforced with carbides and graphene: A comprehensive review
  132. Modification effects of nanosilica on asphalt binders: A review
  133. Decellularized extracellular matrix as a promising biomaterial for musculoskeletal tissue regeneration
  134. Review of the sol–gel method in preparing nano TiO2 for advanced oxidation process
  135. Micro/nano manufacturing aircraft surface with anti-icing and deicing performances: An overview
  136. Cell type-targeting nanoparticles in treating central nervous system diseases: Challenges and hopes
  137. An overview of hydrogen production from Al-based materials
  138. A review of application, modification, and prospect of melamine foam
  139. A review of the performance of fibre-reinforced composite laminates with carbon nanotubes
  140. Research on AFM tip-related nanofabrication of two-dimensional materials
  141. Advances in phase change building materials: An overview
  142. Development of graphene and graphene quantum dots toward biomedical engineering applications: A review
  143. Nanoremediation approaches for the mitigation of heavy metal contamination in vegetables: An overview
  144. Photodynamic therapy empowered by nanotechnology for oral and dental science: Progress and perspectives
  145. Biosynthesis of metal nanoparticles: Bioreduction and biomineralization
  146. Current diagnostic and therapeutic approaches for severe acute respiratory syndrome coronavirus-2 (SARS-COV-2) and the role of nanomaterial-based theragnosis in combating the pandemic
  147. Application of two-dimensional black phosphorus material in wound healing
  148. Special Issue on Advanced Nanomaterials and Composites for Energy Conversion and Storage - Part I
  149. Helical fluorinated carbon nanotubes/iron(iii) fluoride hybrid with multilevel transportation channels and rich active sites for lithium/fluorinated carbon primary battery
  150. The progress of cathode materials in aqueous zinc-ion batteries
  151. Special Issue on Advanced Nanomaterials for Carbon Capture, Environment and Utilization for Energy Sustainability - Part I
  152. Effect of polypropylene fiber and nano-silica on the compressive strength and frost resistance of recycled brick aggregate concrete
  153. Mechanochemical design of nanomaterials for catalytic applications with a benign-by-design focus
Downloaded on 9.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/ntrev-2022-0556/html
Scroll to top button