Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

A facile synthesis of Zn-doped TiO2 nanoparticles with highly exposed (001) facets for enhanced photocatalytic performance

Kun Jiangab, Jin Zhangb, Rui Luob, Yingfei Wanb, Zengjian Liub and Jinwei Chen*b
aDepartment of Integrated Traditional Chinese and Western Medicine, West China Hospital, Sichuan University, Chengdu 610041, PR China
bCollege of Materials Science and Engineering, Sichuan University, Chengdu 610065, PR China. E-mail: jwchen@scu.edu.cn; Tel: +86-28-8541-8786

Received 2nd November 2020 , Accepted 22nd January 2021

First published on 17th February 2021


Abstract

It is a great challenge to simultaneously improve the visible light absorption capacity and enhance photon-generated carrier separation efficiency of photocatalysts. Herein, Zn-doped TiO2 nanoparticles with high exposure of the (001) crystal face were prepared via a one-step hydrothermal decomposition method. A detailed analysis reveals that the electronic structures were modulated by Zn doping; thus, the responsive wavelength was extended to 600 nm, which effectively improved the visible light absorption of TiO2. More importantly, the surface heterojunction of TiO2 was created because of the co-existing specific facets of (101) and (001). Therefore, the surface separation efficiency of photogenerated electron and hole pairs was greatly enhanced. So, the optimal TiO2 photocatalyst exhibited excellent photocatalytic activity, in which the Rhodamine B (RhB) degradation efficiency was 98.7% in 60 min, under the irradiation of visible light. This study is expected to provide guidance for the rational design of TiO2 photocatalysts.


1. Introduction

Titanium dioxide (TiO2) has emerged as one of the most promising photocatalysts for the elimination of wastewater pollution due to its excellent properties such as high stability, non-toxicity, low-cost and outstanding photocatalytic activity.1,2 However, the inherent electronic structure (CB = −0.3 eV, VB = +2.9 eV, Eg = 3.2 eV) of TiO2 weakens its utilization efficiency of solar energy.3,4 More importantly, the short free time and severe surface recombination of photogenerated electrons greatly reduce the efficiency of the photogenerated carriers.5,6 Therefore, it is urgent to develop some effective strategies to widen the optical capture range and enhance the separation efficiency of photo-generated carriers of TiO2.

It is urgent to modify TiO2 with extended visible light absorption and increased visible light activities. Elemental doping is an effective method to modulate the electronic structure of TiO2, which decreases the bandgap to enhance the visible light absorption capacity.7 According to the principle of elemental doping and the structure of TiO2, doping of elements mainly exists in three ways: (1) O atom sites, such as N,8,9 C,10 and defects.11–13 (2) Ti atom sites, such as Fe,14,15 Cu,16,17 and Zn,18 (3) Interstitial doping.19 Elemental doping can greatly enhanced the visible light absorption of TiO2. However, doped atoms are usually used as carrier recombination centers, which further weaken the efficiency of photo-generated carrier separation.20 Hence, how to coordinate the relationship between the photo-generated carrier separation and the visible light absorption of TiO2 is more important.

In order to overcome the above-mentioned obstacles, it is highly desirable to develop a simple and effective strategy to enhance the separation of photo-electrons and hole pairs of TiO2. In general, the construction of heterojunctions or loading transition metal nanoparticles on TiO2 surfaces is an effective strategy to decrease the recombination of photo-electron and hole pairs, which can effectively improve the bulk phase separation efficiency of photogenerated electrons in TiO2.21,22 In addition, the surface separation of photogenerated electrons plays a crucial role in the photocatalytic reaction. Based on this consideration, researchers found that electrons exhibit anisotropic properties on specific facets of TiO2. For example, the concept of surface heterojunction was proposed by Yu et al.23–25 Under the function of the surface heterojunction, the oriented photogenerated holes migrate to the (001) crystal face and electrons migrate to the (101) crystal face, which greatly inhibited the surface recombination of photogenerated carriers, thus improving the utilization efficiency of photogenerated carriers. Similarly, Liu et al. prepared TiO2 with a high density of surface heterojunction and confirmed the existence of surface heterojunction on the TiO2 surface.26 However, to the best of our knowledge, TiO2 with high exposure of (001) specific facets usually has a poor capability to capture visible light and thus greatly decreased the visible-light photocatalytic activity.

Herein, Zn-doped TiO2 nanoparticles with highly exposed (001) facets are synthesized by a one-step hydrothermal method. The characteristics and photocatalytic performance of the obtained catalysts are investigated. The results show that Zn doping can effectively modulate the electronic structure of TiO2, which greatly enhanced the visible light absorption capacity of TiO2. More importantly, the co-existing specific facets of (101) and (001) on TiO2 surfaces greatly promote the separation efficiency of photogenerated carriers. Hence, under the synergistic effects of Zn doping and surface heterojunction, the photocatalytic activity of TiO2 for RhB degradation under visible light was greatly improved.

2. Experimental

2.1 Materials

Titanium(IV) sulfate Ti(SO4)2 (≥99% purity), zinc fluoride ZnF2 (≥99% purity) and zinc sulfate ZnSO4 (≥99% purity) were supplied by Aladdin, and were used as titanium and zinc precursors. All the chemical reagents, such as sodium fluoride (NaF) and urea (CH4N2O) were of analytic grade. All materials were used without further pretreatment. P25 (TiO2, Degussa) was purchased for comparison.

2.2 Sample preparation

Zn-doped TiO2 nanoparticles with high (001) facets were synthesized by a one-step hydrothermal method. A certain amount of Ti(SO4)2 was dissolved in a DI water solution to configure with 0.5 mol L−1 of Ti(SO4)2 for current work. Then, 9.09 g of CH4N2O and a certain amount of a mixture of NaF and ZnF2 were mixed into a 75 ml Ti(SO4)2 solution under vigorous stirring until complete dissolution. The obtained solution was placed into a 100 ml Teflon-lined autoclave at 140 °C for 12 h. After cooling to room temperature, the nanoparticles were washed with deionized water and alcohol three times and freeze-dried at −60 °C. The obtained photocatalysts were denoted as x%-Zn-TiO2(001), where x% represented the Zn/Ti molar ratio. Zn-doped TiO2 (Zn-TiO2) was synthesized by the same procedure by replacing the mixture of NaF and ZnF2 with ZnSO4. TiO2(001) and TiO2 were prepared by the same procedure with and without NaF, respectively. For comparison, P25 was used as the commercial reference material.

2.3 Characterization

XRD patterns were obtained on a Rigaku XRD Ultimate IV using Cu Kα radiation (λ = 0.154 nm) at a scanning range of 5–85°. The morphological properties were determined using a transmission electron microscope (TEM, Tecnai G2 F20, FEI Co., USA). The X-ray photoelectron spectrometer (XPS) was used to characterize the surface information of the composite (Escalab 250Xi, TMO, USA). A spectrophotometer (UV3600) was used to record the ultraviolet-visible diffusing reflection spectrometry patterns of the specimens. Electron spin resonance (ESR, JES FA200, JEOL) was employed to detect the transient radical intermediates.

2.4 Catalytic activity test

Rhodamine B (RhB) was employed to evaluate the photocatalytic performance of the as-prepared samples under visible irradiation. First, RhB solution and TiO2 photocatalysts were prepared. Second, 50 mg each of the TiO2 photocatalyst was dispersed in 100 ml RhB solution (20 mg L−1) individually at 20 °C in a jacketed glass reactor. Third, the obtained solution was fiercely stirred for 60 min in a dark environment for achieving adsorption–desorption equilibrium. Then, the solution was illuminated by a xenon lamp with a power of 300 W and an optical filter (λ > 420 nm, AM = 1.5) was used to cut off the ultraviolet light. The distance between the solution and the light source was 10 cm. At last, 4 ml solution was extracted during the process of degradation every 20 min centrifuged at 14[thin space (1/6-em)]000 rpm for 5 min and examined via UV-vis spectrophotometry. The individual absorption peak of rhodamine B at 554 nm was used to determine its concentration in the solution. The degradation efficiency of RhB was calculated using the following equation.
image file: d0ra09318a-t1.tif
where C0 and Ct are the initial and equilibrium concentrations of RhB, respectively.

3. Results and discussion

3.1 Characterization of the photocatalysts

Morphological information including the size, shape and crystalline interplanar spacing of the as-prepared photocatalysts was acquired via TEM and HRTEM. TEM and HRTEM images of Zn-TiO2(001) are shown in Fig. 1. The results of Fig. 1(a) and (b) show that Zn-TiO2(001) nanoparticles have a foursquare shape with an average size of 40 nm. It has been reported that the foursquare shape particle exhibits a large percentage of (001) facets. Fig. 1(c) and (d) are the HRTEM images of Zn-TiO2(001). On the surface of square-shaped particles, the inter-planar distance of 0.236 nm can be attributed to (001) atomic planes and the inter-planar distance of 0.35 nm can be attributed to (101) atomic planes.26,27 It can be seen from Fig. 1(c) and (d) that the fluorinated photocatalyst exposes a high percentage of (001) facets.
image file: d0ra09318a-f1.tif
Fig. 1 TEM image of (a and b) Zn-TiO2(001), (c and d) HR-TEM images of Zn-TiO2(001).

The crystalline structure of the as-synthesized photocatalysts is investigated by XRD. For all TiO2 samples, as shown in Fig. 2, it can be clearly observed that the characteristic peaks match perfectly with the standard peaks of the anatase phase (JCPDS no. 21-1272) without peaks of other crystalline forms. This result proves that the as-prepared TiO2 has a high purity of the anatase TiO2 structure and Zn2+ may have been uniformly doped into the lattice of TiO2 without forming other crystalline phases. To investigate the chemical state of each element in the TiO2(001) and Zn-TiO2(001) samples, X-ray photoelectron spectroscopy (XPS) was performed. As shown in Fig. 2(b), the presence of Ti, O, C, Zn and F can be clearly observed through the full spectrum scan of the Zn-TiO2(001) sample, revealing the formation of Zn-TiO2(001) nanoparticles. As compared to Zn-TiO2(001), Fig. 2(b) shows the full spectrum scan of the TiO2(001) sample, it is obvious that there was no presence of a Zn peak in the full spectrum scan. In addition, the carbon peak is attributed to the residual carbon from the sample and adventitious hydrocarbon from the XPS instrument itself. Furthermore, the doping concentration of Zn was 2.27% from XPS results. According to the results of XPS, we can confirm that Zn is successfully mixed into TiO2, and the doping concentration is 2.7%.


image file: d0ra09318a-f2.tif
Fig. 2 (a) XRD patterns of TiO2, Zn-TiO2, TiO2(001) and Zn-TiO2(001). (b) XPS spectra of Zn-TiO2(001) and TiO2(001).

The XPS peaks of Ti 2p and O 1s for Zn-TiO2(001) and TiO2(001) are shown in Fig. 3(a) and (b). Comparison of their high-resolution XPS spectra for Ti 2p (Fig. 3(a)) and O 1s (Fig. 3(b)) regions indicates that the introduction of Ov into Zn-TiO2(001) results in decreased binding energies for both Ti and O. This is because two additional electrons are left once an oxygen atom is removed from the surface of Zn-TiO2(001), causing increased electron cloud density around the Ti and O atoms near the Ov sites. Furthermore, in the Zn-2p photoelectron spectra of Zn-TiO2(001), Zn 2p3/2 and Zn 2p1/2 peaks appear at 1021.6 eV and 1044.7 eV, respectively, with the energy splitting of 23.1 eV being a signature of the +2-oxidation state of Zn.28,29 The above observation can be explained in the light of Zn2+ being doped substitutionally in the anatase TiO2 lattice. As the charge on the dopant ions is lower, the electron cloud of Zn2+ experiences a slightly stronger pull from the neighboring Ti4+ ions with a higher oxidation potential.30 Fig. 3(d) shows the high-resolution XPS spectrum of the F 1s region. The measured binding energy is 684.5 eV, which is a typical value for fluorinated TiO2 systems such as [triple bond, length as m-dash]Ti–F species on the TiO2 crystal surface. No signal for F in the lattice of TiO2 (BE = 688.5 eV) is found.31–33


image file: d0ra09318a-f3.tif
Fig. 3 High-resolution XPS spectra of (a) Ti 2p, (b) O 1s, (c) Zn 2p, and (d) F 1s of TiO2(001) and Zn-TiO2(001).

In order to confirm whether the ZnO was formed on the TiO2 surface during the hydrothermal process. The surface state of the Zn-TiO2(001) photocatalyst was determined by Raman spectroscopy, as shown in Fig. 4(a). The peaks at 151.7, 396.5, 516, and 639 cm−1 are very close to those reported for anatase. Therefore, the above-mentioned Raman data signify the absence of the ZnO phase and TiF phase in Zn-doped TiO2 materials. In addition, The N2 sorption isotherms of P25, TiO2(001), and Zn-TiO2(001) samples show a typical curve with the hysteresis loop of the type IV isotherms in Fig. 4(b). The BET surface areas of all samples were calculated. The BET surface area of P25, TiO2(001), and Zn-TiO2(001) are 56 m2 g−1, 94.17 m2 g−1, and 156.7 m2 g−1, respectively. According to the result of the N2 sorption isotherms, we can speculate that the Zn-TiO2(001) sample may show superior photocatalytic activity because its large surface area could expose more active sites and increase solid–liquid interfaces, which enhance the transportation of photocatalyst-relevant species.


image file: d0ra09318a-f4.tif
Fig. 4 (a) The Raman scattering spectra of Zn-TiO2(001), (b) N2 physisorption isotherms, (c) photocatalytic activity for RhB degradation using different photocatalysts, and (d) photocatalytic activity for RhB degradation using different percentages of Zn doped TiO2.

3.2 Photocatalytic performance

The photocatalytic activities of the as-prepared catalysts were evaluated via the photo-degradation of RhB under visible-light irradiation. Fig. 4(c) shows the photocatalytic activities of RhB photosensitization P25, TiO2, Zn-TiO2, TiO2(001) and Zn-TiO2(001) samples for RhB degradation under visible-light irradiation. After illumination for 60 min under visible light, around 3.3%, 19% and 20.1% of RhB were removed via RhB self-photo-degradation, P25 and TiO2 and TiO2(001) shows higher activity than TiO2 prepared without the F based additive. However, the degradation efficiency of RhB is still only around 30%. The doping of Zn seems to show a more pronounced promotion. Zn-TiO2 could remove 65.7% of RhB under 60 min visible light irradiation. Furthermore, 98.7% of RhB was degraded by Zn-TiO2(001), which demonstrates the synthetic effects of Zn doping and high (001) facet exposure for excellent photocatalytic activity under visible light. In order to investigate the effects of the Zn doping concentration on photocatalytic activity, TiO2(001) with different Zn doping concentrations were designed and prepared. As shown in Fig. 4(d), the TiO2-anatase has no obvious effect on RhB degradation under visible light. However, Zn-doped TiO2 samples presented higher photocatalytic activity than TiO2. By increasing the Zn doping amount, the photocatalytic activity of TiO2 was also promoted. When the ratio of Zn to Ti is 0.05, the 5%-Zn-TiO2(001) sample exhibited the highest photocatalytic activity that the degradation efficiency of RhB was 98.7% under 60 min visible light irradiation. Yet, beyond 5% zinc content, the photocatalytic performance of doped Zn-TiO2(001) obviously declined, which may have resulted from the formation of recombination center of photogenerated carriers by the excessive impurity atoms. Based on the literature reported in the last three years, RhB degradation efficiency on different TiO2-based materials is listed in Table 1. In the percent study, RhB degradation with 0.5 g L−1 Zn-TiO2(001) is 98.7% in 1 h. This result is higher than that reported for other TiO2-based catalysts.
Table 1 Comparison of RhB degradation efficiencies with different photocatalysts
Photocatalyst Reaction conditions Degradation efficiency Ref.
Zn-TiO2(001) RhB, 20 mg L−1; catalyst, 0.5 g L−1 98.7%, 1 h This work
Pd/TNTA RhB, 10 mg L−1; catalyst, 1 g L−1 99.7%, 2 h 34
Tm@Nd@TiO2 RhB, 20 mg L−1; catalyst, 0.5 g L−1 88%, 1.5 h 35
PCN/Fe0(1%)–TiO2 RhB, 10 mg L−1; catalyst, 0.5 g L−1 98%, 1.5 h 36
TiOF2/TiO2 RhB, 20 mg L−1; catalyst, 0.5 g L−1 97.7%, 1.5 h 37


In addition, the stability of the Zn-TiO2(001) photocatalyst is evidenced by the cycling test, as shown in Fig. 5(a), only a slight decrease was observed after 4th cycling. This result shows that the Zn-TiO2(001) catalyst has excellent photochemical stability.


image file: d0ra09318a-f5.tif
Fig. 5 (a) The reusability of the Zn-TiO2(001) degradation efficiency of RhB under visible light, (b) UV-Vis absorption spectra, band gap energies (inset), (c) ESR signals of O vacancy, and (d) photocatalytic activity of RhB degradation with and without different scavengers.

3.3 Mechanism of the enhancement

The light absorption abilities of the as-prepared samples, including P25, Zn-TiO2, TiO2(001) and Zn-TiO2(001), were studied via UV-Vis diffuse reflectance spectroscopy. As shown in Fig. 5(b), pure TiO2 (P25) shows a typical absorption band edge at around 375 nm owing to the large band gap of TiO2. Compared with P25, the absorption band edge of TiO2(001) and Zn-TiO2(001) clearly exhibit red-shift and extend to the visible light region. In particular, Zn-TiO2(001) shows the highest response in the visible light range in that the effective adsorption of visible light could extent to 600 nm. The band gaps of P25, Zn-TiO2, TiO2(001) and Zn-TiO2(001) were calculated using the Kubelk–Munk function, as shown in Fig. 5 (b, inset). According to Fig. 5 (b, inset), the band gap of P25 was 3.31 eV. It is consistent with previous research. However, when Zn was introduced into TiO2, the band gap decreased to 3.2 eV and 3.23 eV for Zn-TiO2(001) and Zn-TiO2, respectively. The band gap did not change with Zn doping, so what causes Zn-TiO2(001) sample shows a very strong band tail absorption in the visible-light region (420–600 nm)? According to our previous report, the visible light absorption capacity of a photocatalyst was enhanced by defects.38 Therefore, surface defects on Zn-TiO2(001) were confirmed via electron spin resonance (ESR) spectra, as shown in Fig. 5(c). According to previous reports, the characteristics g value was 1.998. However, the present ESR studies show almost no presence of Ti3+ in the Zn-TiO2(001) photocatalyst. The ESR spectra of Zn-TiO2(001) exhibit a distinct signal at g = 2.01 related to the O vacancies of TiO2. Therefore, the ESR data confirm the presence of O-vacancies in Zn-TiO2(001). Under the function of O vacancies, the visible light absorption capacity of Zn-TiO2(001) is greatly enhanced.

Compared to the photocatalytically active substance for degradation in a liquid solution, the existence of a dissolved scavenger is harmful for organic pollutant degradation. The photocatalytic activity is reduced by adding isopropanol (hydroxyl scavenger), ammonium oxalate (hole scavenger), and benzoquinone (superoxide radical scavenger) into the organic solution with visible light irradiation. As shown in Fig. 5(d), the photocatalytic degradation activity of Zn-TiO2(001) has varying degrees of reduction upon the addition of different amounts of scavengers. In particular, after adding ammonium oxalate, the photo-degradation efficiency is reduced to zero, indicating that the photo-generation of hole radicals is critical for the activity substance during the photo-degradation process. It demonstrates that Zn-TiO2(001) with Zn doping and high (001) facets is beneficial for the formation of hole radicals generated on the TiO2 surface.

The ESR spectra were contained to detect the transient radical intermediates in P25, Zn-TiO2 and Zn-TiO2(001). Fig. 6 shows the formation of e, h+, ˙O2− and ˙OH after 10 min of visible irradiation under N2 atmosphere. As show in Fig. 6(a) and (b), a strong peak is observed for all catalysts, which are corresponding to the photogenerated of e and h+. Compared with P25 and Zn-TiO2, the intensity of the characteristic peak increases with the existence of Zn doping and high (001) facets of Zn-TiO2(001), which implies that doped Zn could promote the photo-generated carrier and the special crystal facet structure could show more holes appearing on the surface of TiO2. Furthermore, as Fig. 6(c) and (d) show, the order of the ˙O2− and ˙OH radical intensity, both are following Zn-TiO2(001) > Zn-TiO2 > P25 order. It could be concluded that the as-prepared Zn-TiO2(001) photocatalysts could produce more radicals to degrade RhB under visible light. This conclusion is in accordance with the results of the RhB degradation efficiency (Fig. 7).


image file: d0ra09318a-f6.tif
Fig. 6 ESR spectra of P25, Zn-TiO2 and Zn-TiO2(001): (a) e under visible light conditions, (b) h+ under visible light conditions, (c) ˙O2 under visible light conditions and (d) ˙OH under visible light conditions.

image file: d0ra09318a-f7.tif
Fig. 7 The mechanism diagram of Zn-TiO2(001).

Based on the experimental data, the mechanism for the improved photocatalytic performance of Zn-TiO2(001) is proposed. The electrons and holes in TiO2 catalysts could be generated under visible light illumination. Then, the electrons in CB and the holes in VB should transfer to the different crystal surfaces of TiO2. During photocatalytic degradation, the main active substance consuming organic pollutants is superoxide radical (˙O2), hydroxyl radical (˙OH), and holes (h+). The mechanism of RhB photo-degradation is proposed as follows:

TiO2 → hVB+ + eCB

O2 + eCB → ˙O2

O2 + 2H+ + eCB → H2O2

4H2O + 2hVB+ → 4H2O2 + 2H+

H2O2 → 2˙OH

hVB+/˙OH + RhB → CO2 + H2O

4. Conclusion

In summary, Zn-doped TiO2 nanoparticles with highly exposed (001) facets were synthesized by a one-step hydrothermal method. The photocatalytic activity of RhB degradation over the Zn-TiO2(001) photocatalysts reaches 98.7% after 60 min of visible light irradiation. Zn doping effectively enhances the concentration of O vacancies and accelerates the formation of hole and hydroxyl radicals from the generated electron–hole pairs and special crystal facet structure exposes more holes on the surface of TiO2, which could contribute to the enhanced photocatalytic activity. The facile preparation method and outstanding photocatalytic performance provide a promising photocatalyst for organic pollutant degradation.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The work was supported by the National Key Research and Development Program of China (2018YFB1502700), the Provincial Nature Science Foundation of Sichuan (2017CC0017, 2018FZ0105, 2019YJ0025), the Research and Development Program of Chengdu (2019-YF05-01193-SN), and the Fundamental Research Funds for the Central Universities (No. YJ201746).

Notes and references

  1. J. Ji, Y. Xu, H. B. Huang, M. He, S. L. Liu, G. Y. Liu, R. J. Xie, Q. Y. Feng, Y. J. Shu, Y. J. Zhan, R. M. Fang, X. G. Ye and D. Y. C. Leung, Chem. Eng. J., 2017, 327, 490–499 CrossRef CAS.
  2. F. V. S. Lopes, R. A. R. Monteiro, A. M. T. Silva, G. V. Silva, J. L. Faria, A. M. Mendes, V. J. P. Vilar and R. A. R. Boaventura, Chem. Eng. J., 2012, 204, 244–257 CrossRef.
  3. V. Kurnaravel, S. Mathew, J. Bartlett and S. C. Pillai, Appl. Catal., B, 2019, 244, 1021–1064 CrossRef.
  4. K. Wetchakun, N. Wetchakun and S. Sakulsermsuk, J. Ind. Eng. Chem., 2019, 71, 19–49 CrossRef CAS.
  5. M. Z. Ge, J. S. Cai, J. Iocozzia, C. Y. Cao, J. Y. Huang, X. N. Zhang, J. L. Shen, S. C. Wang, S. N. Zhang, K. Q. Zhang, Y. K. Lai and Z. Q. Lin, Int. J. Hydrogen Energy, 2017, 42, 8418–8449 CrossRef CAS.
  6. S. Kohtani, A. Kawashima and H. Miyabe, Catalysts, 2017, 7(10), 303 CrossRef.
  7. C. Burda, Y. B. Lou, X. B. Chen, A. C. S. Samia, J. Stout and J. L. Gole, Nano Lett., 2003, 3, 1049–1051 CrossRef CAS.
  8. S. Sakthivel and H. Kisch, ChemPhysChem, 2003, 4, 487–490 CrossRef CAS.
  9. M. Batzill, E. H. Morales and U. Diebold, Phys. Rev. Lett., 2006, 96, 4 CrossRef.
  10. S. U. M. Khan, M. Al-Shahry and W. B. Ingler, Science, 2002, 297, 2243–2245 CrossRef CAS.
  11. K. Bhattacharyya, J. Majeed, K. K. Dey, P. Ayyub, A. K. Tyagi and S. R. Bharadwaj, J. Phys. Chem. C, 2014, 118, 15946–15962 CrossRef CAS.
  12. K. Bhattacharyya, B. Modak, C. Nayak, R. G. Nair, D. Bhattacharyya, S. N. Jha and A. K. Tripathi, New J. Chem., 2020, 44, 8559–8571 RSC.
  13. K. Bhattacharyya, S. Varma, A. K. Tripathi, S. R. Bharadwaj and A. K. Tyagi, J. Phys. Chem. C, 2008, 112, 19102–19112 CrossRef CAS.
  14. V. Moradi, M. B. G. Jun, A. Blackburn and R. A. Herring, Appl. Surf. Sci., 2018, 427, 791–799 CrossRef CAS.
  15. F. F. Wang, T. Shen, Z. P. Fu, Y. L. Lu and C. X. Chen, Nanotechnology, 2018, 29, 7 Search PubMed.
  16. R. Camarillo, D. Rizaldos, C. Jimenez, F. Martinez and J. Rincon, J. Supercrit. Fluids, 2019, 147, 70–80 CrossRef CAS.
  17. A. Heidarpour, Y. Mazaheri, M. Roknian and S. Ghasemi, J. Alloys Compd., 2019, 783, 886–897 CrossRef CAS.
  18. Y. Yu, J. Wang, W. Li, W. Zheng and Y. Cao, CrystEngComm, 2015, 17, 5074–5080 RSC.
  19. J. Li, J. Xu, W.-L. Dai, H. Li and K. Fan, Appl. Catal., B, 2009, 85, 162–170 CrossRef CAS.
  20. A. Kudo, H. Kato and I. Tsuji, Chem. Lett., 2004, 33, 1534–1539 CrossRef CAS.
  21. S. Sun, W. Wang, L. Zhang, M. Shang and L. Wang, Catal. Commun., 2009, 11, 290–293 CrossRef CAS.
  22. M. Jakob, H. Levanon and P. V. Kamat, Nano Lett., 2003, 3, 353–358 CrossRef CAS.
  23. A. Y. Meng, J. Zhang, D. F. Xu, B. Cheng and J. G. Yu, Appl. Catal., B, 2016, 198, 286–294 CrossRef CAS.
  24. M. Y. Xing, B. X. Yang, H. Yu, B. Z. Tian, S. Bagwasi, J. L. Zhang and X. Q. Gongs, J. Phys. Chem. Lett., 2013, 4, 3910–3917 CrossRef CAS.
  25. J. G. Yu, J. X. Low, W. Xiao, P. Zhou and M. Jaroniec, J. Am. Chem. Soc., 2014, 136, 8839–8842 CrossRef CAS.
  26. H. G. Yang, C. H. Sun, S. Z. Qiao, J. Zou, G. Liu, S. C. Smith, H. M. Cheng and G. Q. Lu, Nature, 2008, 453(7195), 638–641 CrossRef CAS.
  27. Q. J. Xiang, K. L. Lv and J. G. Yu, Appl. Catal., B, 2010, 96, 557–564 CrossRef CAS.
  28. Z. Wang, Y. Gu, J. Qi, S. Lu, P. Li, P. Lin and Y. Zhang, RSC Adv., 2015, 5, 42075–42080 RSC.
  29. K. Bhattacharyya, A. Danon, B. K. Vijayan, K. A. Gray, P. C. Stair and E. Weitz, J. Phys. Chem. C, 2013, 117, 12661–12678 CrossRef CAS.
  30. R. G. Nair, S. Mazumdar, B. Modak, R. Bapat, P. Ayyub and K. Bhattacharyya, J. Photochem. Photobiol., A, 2017, 345, 36–53 CrossRef CAS.
  31. J. C. Yu, J. G. Yu, W. K. Ho, Z. T. Jiang and L. Z. Zhang, Chem. Mater., 2002, 14, 3808–3816 CrossRef CAS.
  32. J. Yu, W. Wang, B. Cheng and B.-L. Su, J. Phys. Chem. C, 2009, 113, 6743–6750 CrossRef CAS.
  33. H. G. Yang, G. Liu, S. Z. Qiao, C. H. Sun, Y. G. Jin, S. C. Smith, J. Zou, H. M. Cheng and G. Q. Lu, J. Am. Chem. Soc., 2009, 131, 4078–4083 CrossRef CAS.
  34. L. F. Cui, T. T. Pu, Z. F. Shen, S. S. Li, S. F. Kang, Q. N. Xia, Y. G. Wang and X. Li, Res. Chem. Intermed., 2019, 45, 2167–2177 CrossRef CAS.
  35. H. N. Huang, H. L. Li, Z. Y. Wang, P. Wang, Z. K. Zheng, Y. Y. Liu, Y. Dai, Y. J. Li and B. B. Huang, Chem. Eng. J., 2019, 361, 1089–1097 CrossRef CAS.
  36. K. Rabe, L. F. Liu, N. A. Nahyoon, Y. Z. Zhang, A. M. Idris, J. Q. Sun and L. X. Yuan, J. Taiwan Inst. Chem. Eng., 2019, 96, 463–472 CrossRef CAS.
  37. Z. D. Liu, X. N. Liu, Q. F. Lu, Q. Y. Wang and Z. Ma, J. Taiwan Inst. Chem. Eng., 2019, 96, 214–222 CrossRef CAS.
  38. J. Zhang, J. Chen, Y. Wan, H. Liu, W. Chen, G. Wang and R. Wang, ACS Appl. Mater. Interfaces, 2020, 12, 13805–13812 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2021
Click here to see how this site uses Cookies. View our privacy policy here.