Next Article in Journal
Invasome: A Novel Nanocarrier for Transdermal Drug Delivery
Next Article in Special Issue
Pb-Free Cs3Bi2I9 Perovskite as a Visible-Light-Active Photocatalyst for Organic Pollutant Degradation
Previous Article in Journal
Structural Features and Photoelectric Properties of Si-Doped GaAs under Gamma Irradiation
Previous Article in Special Issue
Enhanced Degradation of Sulfamethoxazole (SMX) in Toilet Wastewater by Photo-Fenton Reactive Membrane Filtration
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Liquid-Plasma Hydrogenated Synthesis of Gray Titania with Engineered Surface Defects and Superior Photocatalytic Activity

1
State Key Laboratory of Precision Spectroscopy, East China Normal University, Shanghai 200062, China
2
Shanghai Key Laboratory of Modern Optical System, Engineering Research Center of Optical Instrument and System, Ministry of Education, School of Optical-Electrical and Computer Engineering, University of Shanghai for Science and Technology, Shanghai 200093, China
3
Chongqing Institute of East China Normal University, Chongqing 401120, China
4
CAS Center for Excellence in Ultra-Intense Laser Science, Shanghai 201800, China
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Submission received: 19 January 2020 / Revised: 10 February 2020 / Accepted: 13 February 2020 / Published: 17 February 2020

Abstract

:
Defect engineering in photocatalysts recently exhibits promising performances in solar-energy-driven reactions. However, defect engineering techniques developed so far rely on complicated synthesis processes and harsh experimental conditions, which seriously hinder its practical applications. In this work, we demonstrated a facile mass-production approach to synthesize gray titania with engineered surface defects. This technique just requires a simple liquid-plasma treatment under low temperature and atmospheric pressure. The in situ generation of hydrogen atoms caused by liquid plasma is responsible for hydrogenation of TiO2. Electron paramagnetic resonance (EPR) measurements confirm the existence of surface oxygen vacancies and Ti3+ species in gray TiO2−x. Both kinds of defects concentrations are well controllable and increase with the output plasma power. UV–Vis diffused reflectance spectra show that the bandgap of gray TiO2−x is 2.9 eV. Due to its extended visible-light absorption and engineered surface defects, gray TiO2−x exhibits superior visible-light photoactivity. Rhodamine B was used to evaluate the visible-light photodegradation performance, which shows that the removal rate constant of gray TiO2−x reaches 0.126 min−1 and is 6.5 times of P25 TiO2. The surface defects produced by liquid-plasma hydrogenation are proved stable in air and water and could be a candidate hydrogenation strategy for other photocatalysts.

Graphical Abstract

1. Introduction

Titanium dioxide (TiO2) due to its nontoxic, high stability, and high photocatalytic performance is considered as a promising photocatalyst, which has been widely used in environmental remedies and water splitting [1,2,3]. However, the photoactivity of TiO2 is largely limited by its wide bandgap (3.2 eV) and only responds to UV light, nearly 4% part of solar energy, which severely restricts its practical applications. In order to enhance solar light absorption, much effort has been devoted to bandgap engineering, by means of noble metal (Au, Pt, Ag, etc.) and nonmental (N, S, I, etc.) doping, artificial crystal structure modification, and carbonaceous nanomaterial compositing (carbon nanotubes, graphene, and fullerenes) [4,5,6,7,8,9,10]. Nevertheless, the visible-light energy conversion remains insufficient due to little solar light absorption and many carrier-recombination centers [11,12]. In recent years, black TiO2 has gained enormous attention owing to its extraordinary visible and infrared light absorption, prompting superior performances in photodegradation and water splitting [13,14,15]. The enhanced solar light absorption of black TiO2 is ascribed to additional intermediate electronic states (e.g., Vo or Ti3+ states) and disorder surface caused by hydrogenation [16,17,18]. More importantly, defect engineering, for instance defects concentration and distribution, also plays an important role in tuning photocatalytic activity of black TiO2. It was reported that the high defect ratio of surface to bulk can significantly enhance the photoactivity owing to preferential diffusion from bulk to surface of photoinduced charges [19,20]. However, surface defects are not stable enough as it can be spontaneously oxidized in air and water. So far, mainstream strategies to synthesize black/gray TiO2−x are relied on the reduction of pristine stoichiometric TiO2 [20,21,22,23,24], such as annealing under high pressures of H2 or NH3, high-energy particle bombardment (hydrogen plasma, laser plasma, or high-energy electrons), and chemical reduction with reducing agent in vacuum. Apparently, it is significant and desirable to develop a simple and feasible strategy for massive production of black/gray TiO2−x with engineered surface defects and related abundant solar absorption.
Hereafter, we conduct a one-pot synthesis of gray TiO2−x with large solar harvesting and engineered surface defects using a liquid-plasma technology at room temperature and atmospheric pressure [25]. There has recently been increasing interest in liquid-plasma discharges and their potential applications in various technologies, including environmental remediation, nanomaterial synthesis, and surface processing [26,27,28]. One of the most important advantages of liquid plasma is that various reactions instantaneously occur, including active species oxidation (e.g., OH, O, H2O2, etc.), ultraviolet radiation, electric fields, and shock waves. The synergy among those intense reactions is considered to possess a much higher efficiency than traditional chemical methods. In addition, liquid plasma provides an environmentally green and safe method, as fabrication itself consumes no extra chemicals and produces no chemical residuals.
We got mass production of black/gray TiO2−x via liquid plasma driven by a bipolar pulse power. The synthesis setup, as shown in Figure 1A, is quite simple and it requires no complicated vacuum chambers, no reducing gas, or reagents. The bright liquid plasma was generated on the surface of metallic titanium electrode by applying high voltage pulses with high frequency as shown in Figure 1B. The typical generation procedure of liquid plasma can be detailed as follows. When appropriate pulse voltages were applied, some vapors were formed to cover the cathodes surface, i.e., “gas sheath”, by the Joule heating effect [29]. The “gas sheath” breakdown took place with the increase of electrolyte temperature and input voltage, and concurrently, glow discharge was generated near the cathodes surface. The temperature of cathode plasma was estimated to be 2000–3000 K according to previous studies [30,31]. Afterwards, cathodic plasma dissociated water molecules to produce the hydrogen environment at/near the plasma region [32]. After 1 h of liquid-plasma treatments, the color of 300 g anatase titania was varied from white to dark gray, and the digital pictures of dry powder samples are shown in Figure 1C. Figure 2 shows the optical emission spectrum of liquid plasma and clearly exhibits various emission peaks attributed to Ti I (neutral), Ti II (single-charged ions), hydroxyl radicals, hydrogen, and atomic oxygen, which were originated from the interaction between plasma and liquid. Thus, we hypothesize that white TiO2 nanoparticles dissolved in electrolyte entered the hot plasma area, and thus underwent plasma thermal treatment and hydrogenation at/near the plasma region. This synergy played a significant role in synthesizing gray TiO2−x, and the synthesis mechanism of gray TiO2−x is depicted in Figure 2. From EPR measurements, a tunable surface-defect concentration of TiO2−x was achieved by applying a different plasma discharge power. Interestingly, even though surface defects are easily oxidized under strong oxidation environments, these surface defects generated through the liquid-plasma strategy were verified as quite stable.

2. Experimental Section

2.1. Preparation of Gray TiO2−x

Commercial white anatase TiO2 nanopowders were bought from Shanghai Xieqing industrial Co., Ltd.(Shanghai, China), with 30 nm average size. P25 TiO2 of 20–30 nm was bought from Degussa Corp, which consisted of 78% anatase and 28% rutile. Briefly, we mixed 300 mL of 100 mg/mL well-dispersed anatase TiO2 solution and 7 mL of 0.1 mol/L nitric acid to get the electrolyte, which was afterwards transferred to electrolytic cell. One anodic platinum sheet (20 × 20 × 1 mm3, 99.99%) and two cathodic titanium rods (4-mm diameter, 99.9%) were sealed into a corundum tube in 300-mL nitric acid electrolyte as shown in Figure 1A. Two cathodes were used to generate glow discharges and avoid unbalanced flow and temperature gradients in the electrolyte. The HNO3 electrolyte here acted as a conductive solution. Pulsed voltages were applied between anode and cathodes to produce intense plasma nearby the cathode surfaces. Liquid plasma was produced on the cathode surface when we applied an appropriate pulse voltage power (600 V). The produced plasma whose intensity is dependent on glow discharge power, induces the high temperature at plasma–electrolyte interface and hydrogen generation. To avoid the electrolyte evaporation, a water chiller was used to maintain the electrolyte temperature at 80 °C. The color of TiO2 electrolyte gradually turned from white to dark gray within 1 h. The reaction mechanism for the formation of gray TiO2−x can be ascribed to the hydrogenation effect of liquid plasma. Briefly, TiO2 nanopowders well dispersed in electrolyte took disordered movements due to intense liquid-plasma shockwaves. Some TiO2 nanoparticles moved to the plasma region and underwent hydrogenation under high temperature (2000–3000 K) owing to the hydrogen environment derived from water dissociation. Then, large numbers of nanoparticles repeated plasma hydrogenation treatments and hydrogenated gray TiO2−x from white TiO2 was finally obtained. Hereafter, we dubbed the sample according to the output plasma power, for example GT-360 refers to gray TiO2−x obtained by applying an output plasma power of 360 W.

2.2. Characterization

The phase and crystallinity for all samples were tested by the X-ray powder diffraction (XRD) using a Rigaku Smartlab (Rigaku, Tokyo, Japan) machine equipped with Cu Ka irradiation (λ = 1.54056 Å). The morphology was characterized by TEM using a JEM-2500SE (JEOL, Tokyo, Japan) instrument operating at an acceleration voltage of 200 kV. UV–Vis diffused reflectance spectra were measured by the Shimadzu UV-2700 spectrophotometer at a wavelength range of 200–800 nm at room temperature. The Raman spectra of samples were recorded by Thermo scientific DXR Raman microscope with 532-nm laser excitation at room temperature. The X-ray photoelectron spectra (XPS) were recorded with thermos Escalab 250Xi. The existence of defects doped in the TiO2−x nanoparticles was confirmed by the X-band EPR spectra recorded at room temperature. The plasma discharges were controlled by applying square-wave voltages between titanium cathodes and Pt anode, with variable pulsed voltage of 0~800 V, pulsed current of 0~5 A, and repetition rate of 0~5 kHz. The voltage was increased at a ratio of 20 V/s.

2.3. Visible-Light Photocatalytic Degradation

The evaluation of visible-light photoactivity was individually tested with three typical waste-water pollutants, including rhodamine B (RhB), methyl orange (MO), and phenol. We used a 300-W Xenon lamp with a 420-nm cutoff filter as the visible-light source. Degussa P25 TiO2 was used as a standard photocatalyst for photoactivity comparison with as-prepared samples. The concentrations of RhB, MO, and phenol were 20, 20, and 10 mg/L, respectively. Firstly, 50 mg as-prepared TiO2−x nanopowders and 50 mL of pollutant solution were put in a 500-mL beaker. Before illumination, the mixture was placed in a dark environment for 30 min with magnetic stirring for adsorption–desorption equilibrium. During photodegradation, we took out 1-mL of solution in the glass every 10 min and centrifuged it at 10,000 rpm for 11 min. Finally, the concentration of pollutant solution was analyzed at a specific wavelength by the UV-2700 spectrophotometer, where RhB, MO, and phenol are located at 554, 464, and 270 nm, respectively. Photodegradation experiments for whole tested samples were carried out under the same conditions.

3. Results and Discussion

The crystal structures of gray TiO2−x obtained at different plasma discharge powers were detected by XRD analysis as shown in Figure 3. The XRD patterns of the as-grown gray TiO2−x exhibit characteristic diffraction peaks matching the (101), (103), (004), (112), (200), (105), (211), (204), (116), (220), (215) facets of anatase crystal without any additional diffraction peaks, showing identical diffraction peaks with pristine TiO2. In addition, the full width at half maximum and height (FWMH), as well as the integrated peak area of the diffraction peak at 101 facets are nearly unchanged after plasma treatments as seen in Table S1 (see Support Information). This result indicates liquid plasma treatments neither destroy the long-range order lattice structure nor enlarge nanoparticle size of gray TiO2−x.
The morphology and lattice information of pristine TiO2 and gray TiO2−x nanoparticles are presented in Figure 4A–D. Figure 4A shows that pristine TiO2 nanoparticles possessed serious agglomeration. Conversely, as shown in Figure 4B, little agglomeration was observed in gray TiO2−x, which could be ascribed to increasement of surface energy in nanoparticles owing to plasma treatments. Both pristine TiO2 and gray TiO2−x nanocrystals are on average ~30 nm in diameter, which is well consistent with XRD measurements and suggests that hot temperature in liquid plasma does not prompt TiO2 crystal growth. Figure 4C,D present the high-resolution morphological structures of pristine TiO2 and TiO2−x nanoparticles. From Figure 4C, marginal defects of pristine TiO2 are observed and could be produced during the crystal growth procedure. As indicated in Figure 4D, it seems that a thin disorder shell, nearly 1–2 nm in thickness, appears on the gray TiO2−x surface, while the bulk maintains good crystallinity.
The UV–Vis diffuse reflectance spectra (DRS) are used to evaluate the light absorption properties of gray TiO2−x samples. As shown in Figure 5A, pristine TiO2 shows merely ultraviolet light absorption, while all as-prepared gray TiO2−x samples exhibit a broadband absorption from ultraviolet to visible even infrared regions. Moreover, from the inset of Figure 5A, the as-prepared TiO2−x gradually exhibited a deepened color with the change of output power from 360 to 480 W. Figure 5B plots (αhν)1/2 versus hν by using the Kubelka–Munk function [33]. From which we calculate the bandgap of pristine TiO2, GT-360, GT-420, and GT-480 as 3.14, 3.07, 3.00, and 2.90 eV, respectively. Apparent red-shifts were observed for the absorption edge of gray TiO2−x compared with pristine TiO2. The higher discharge power is, the narrower bandgap of gray TiO2−x is obtained. The decreased bandgap in our case implies some self-doped crystal defects, for instance oxygen vacancy and/or Ti3+ centers, has been produced after plasma hydrogenation. Since self-doped defects inevitably introduce doped states between the bandgap such as oxygen-vacancy and Ti3+ induced intermedium states, which accounts for the visible-light absorption of gray TiO2−x. The color depth of defective TiO2−x is reported and can be related to the amount of surface defects, which plays a role of “color centers”, and the more surface defects TiO2−x possesses, the more visible light TiO2−x captures [34,35]. To verify the existence of surface defects, gray TiO2−x was heated at 300 °C under atmosphere for 3 h. Unexpectedly, the color was changed from gray to white, which can be attributed to the fact that surface defects diffuse from surface to bulk or can be oxidized by air. Thus, we conclude that some surface defects most likely exist in gray TiO2−x and its quantity could be controlled directly by varying plasma discharge power.
The Raman spectra of gray TiO2−x and pristine TiO2 are shown in Figure 6. Both samples exhibit similar Raman peaks at 142, 194, 395, 514, and 637 cm−1, respectively, which is ascribed to anatase TiO2, well consistent with the XRD results. It is observed that the Eg mode of anatase associated with the symmetric stretching vibration of O-Ti-O is shifted to higher wavenumbers compared with pristine TiO2, which reveals that the Ti-O bands environment is changed. In principle, this phenomenon can result from lattice defects and/or finite size effect (<10 nm). Our TEM measurements exclude the size effect as pristine TiO2 and gray TiO2−x have almost the same nanoparticle size. Consequently, we infer that the difference between gray TiO2−x and pristine TiO2 is mainly due to the lattice defects effect.
XPS analyses are further performed to investigate the surface chemical states of Ti and O in gray TiO2−x. Figure 7A shows the full-scale XPS spectra of pristine TiO2 and gray TiO2−x. No obvious differences are observed between pristine TiO2 and gray TiO2−x, indicating no other dopant elements in gray TiO2−x. Figure 7B shows a high-resolution Ti 2p XPS spectra for both samples. There are two main peaks for pristine TiO2 centered at a binding energy of 458.6 and 464.4 eV, which are ascribed to Ti4+ 2p3/2 and Ti4+ 2p1/2, respectively [36]. In comparison with pristine TiO2, two similar peaks of gray TiO2−x are detected but with a slight shift toward lower binding energy at 458.3 and 464 eV, assigned to Ti3+ 2p3/2 and Ti3+ 2p1/2, respectively [37]. Figure 7C shows the XPS O 1s spectra for pristine TiO2 and gray TiO2−x. From the O 1s spectrum of pristine TiO2, two deconvoluted peaks at 529.6 and 531.4 eV are related to lattice oxygen (Ti-O) and surface absorption hydroxyl (Ti-OH), respectively [38]. Comparably, the fitting peaks of O 1s spectrum of gray TiO2−x show an obvious shift toward higher binding energy at 529.7 and 531.5 eV, which is mainly due to the existence of oxygen vacancies. Moreover, the area of fitting peak of Ti-OH in gray TiO2−x is little larger than that of pristine TiO2, indicating that many Ti dangling bonds induced by oxygen vacancies are produced.
The existence of surface defects was confirmed by EPR spectra at room temperature as shown in Figure 8. Unexpectedly, the pristine TiO2 shows an apparent paramagnetic signal centered at g = 2.001 ascribed to surface oxygen defects, which is consistent with HRTEM measurement. In comparison with TiO2, gray TiO2−x possesses two extinguished peaks at g = 2.001 and g = 1.974, corresponding to surface oxygen vacancies and surface Ti3+ species, respectively, which is in good accordance with XPS results [39,40]. As is well known, the formation of Ti3+ centers is always accompanied by the presence of oxygen vacancies due to the charge unbalance. It should be noticed that both of the EPR signals gradually increase with plasma discharge power. It is therefore proved that the high concentration of surface defects can be obtained when applying sufficient output energy.
Based on the obtained results, the generation of surface defects in gray TiO2−x could be understood that hydrogen atoms generated by liquid-plasma electrolysis reduced TiO2 to produce Ti3+ and oxygen vacancies. Furthermore, hydrogen atoms entered in the TiO2−x lattice and some atoms bonded with Ti atoms while others bonded with O atoms, and thus created the surface lattice disorders [41]. Massive hydrogen atoms produced by enhanced output power create an intensified plasma hydrogenation effect, which induces the increased contents of surface Ti3+ and oxygen vacancies [42]. In addition, there is no color fade for gray TiO2−x even if exposed in air and water for six months, indicating the existences of surface defects are quite stable. The robust surface defects could be attributed to the wrapped outer disorder surface with 1–2 nm thickness, hindering further strong oxidation by liquid plasma. On the other hand, considering the presence of surface Ti3+ and oxygen vacancies, the rather stable point defect structure of Ti3+-Vo-Ti3+ could be formed in order to maintain the electrostatic balance [43].
The photocatalytic performance for whole samples was tested with degradation of RhB under visible-light illumination (λ > 420 nm), as shown in Figure 9A. It is clearly seen that gray TiO2−x shows a much higher photoactivity than that of pristine TiO2 and P25. Figure 9B shows the calculated ln(C0/Ct) with the irradiation time (t), i.e., ln(C0/Ct) = kt, where k is the apparent rate constant, which follows first-order kinetic behavior and can directly give the efficiency of photodegradation. The best performance of gray TiO2−x is 0.126 min−1, which is far higher than pristine TiO2 and is 6.5 times of P25. In order to evaluate the stability of gray TiO2−x, the photocatalytic reaction was repeated four times under the same experimental conditions. As shown in Figure 9C, after four repeated cycles, the photocatalytic performance has no obvious differences, which indicates that the photoactivity of gray TiO2−x is rather stable. Additionally, the visible-light degradation of methyl orange and phenol are shown in Figure 9D. After 2 h of photodegradation, the degradation rate of MO using gray TiO2−x is 96.25%, while pristine TiO2 shows only 7.52%. The degradation rate of phenol for pristine TiO2 is 4.26%, while gray TiO2−x shows a much higher rate than pristine TiO2 and reaches 89.2%. Furthermore, some hydrogenated titania and doped TiO2 with high performance in RhB photodegradation are provided for comparison with gray TiO2−x as reflected in Table 1.
On basis of the above results, we can conclude that the superior photocatalytic activity of gray-TiO2−x is mainly due to the existence of surface defects. It is generally accepted that photocatalytic activity is largely dependent on the quantity of absorbed photons. The more photons TiO2 can absorb, the more photoinduced electrons and holes it can produce. Owing to the existence of oxygen vacancies and Ti3+ centers, dopants states such as Ti3+ states and oxygen states appear between the bandgap, which can narrow the bandgap and allow more long-wavelength light absorption. Even though, lots of photoinduced e-h pairs yet are easily recombined in the bulk of TiO2 crystal, leading to a serious waste of photoinduced charges. Surface defects serve as holes absorbers and thus cause the directional diffusion of photogenerated e-h pairs to the TiO2 surface. Therefore, benefits of surface defects in gray TiO2−x not only narrow its bandgap, but also reduce the recombination rate of e-h pairs. On the other hand, with the increase of output energy, the concentrations of surface defects are correspondingly improved, and lead to a narrower bandgap, as well as faster e-h pairs diffusion. It thus explains discharge power-dependent photoactivity of gray TiO2−x.
Considering the high temperature involved in liquid plasma treatments (2000–3000 K), the possible modification of black TiO2 deriving from the titanium cathodes should be discussed. Originally, we carried out the liquid plasma experiment to produce black TiO2, which was reported by others [16,17,18]. Typically, Ti cathodes (99.9%) and Pt anode (99.99%) were used and all other experimental conditions and setup are same with this work, except no TiO2 powder was mixed in the electrolyte. The black TiO2 nanoparticles were derived from Ti cathodes after plasma anodization and heat treatment, which shows abundant surface oxygen vacancies. However, the quantity of black TiO2 is quite few (about 10 mg), and it cannot meet our demands in the experiment and some practical applications. Furthermore, visible-light photodegradation of black TiO2 shows a heavy decrease in photocatalytic performance, which was ascribed to unstable existences of surface oxygen vacancies due to photocorrosion. To improve the quantity and photocatalytic performance, lots of anatase TiO2 powders were added and mixed in the electrolyte, which subsequently produced gray TiO2−x with mass production and stable surface defects. Overall, the contamination of black TiO2 generated from Ti cathodes existed. Nevertheless, the stabilities of surface oxygen defects, as well as photodegradation performance exhibited in gray TiO2−x, which confirms that the mechanism of plasma power-related hydrogenation with TiO2 plays the major role in whole synthesis procedure. Therefore, contamination from Ti cathodes does not affect the performance of gray TiO2−x, and the modification can be ignored. On the other hand, platinum sheet as inert metals is resisted by heavy corrosion due to intense plasma anodization, and there is no mass decay of Pt anode after plasma treatments. Transmission electron microscopy-energy dispersive X-ray spectroscopy (TEM-EDS) was carried out to exclude the influence of Pt element as seen in Figure S1.

4. Conclusions

In summary, a simple green method assisted by liquid plasma has been demonstrated for the synthesis of gray TiO2−x at a low temperature and atmospheric pressure. The synthesis mechanism can be explained that the in situ production of plasma-induced hydrogen atoms plays significant roles in the hydrogenation reactions of gray TiO2−x. The liquid-plasma technique can effectively manipulate the optical properties and band structures of gray TiO2−x, and herein enhances accompanying visible-light absorption. One interesting finding is that the engineered surface defects can be realized through directly controlling the plasma discharge power. EPR measurements confirm the existences of surface oxygen vacancies and Ti3+ species in gray TiO2−x. Both kinds of defects concentrations are well controllable and increase with the output plasma power. The synergistic effect with narrowed bandgap and directional e-h pairs diffusion results in superior visible-light photoactivity. Rhodamine B was used to evaluate the visible-light photodegradation performance. The results show that the removal rate constant of gray TiO2−x reaches 0.126 min−1, and is 6.5 times of P25. Moreover, photodegradations of methyl orange and phenol confirm that gray TiO2−x exhibits much higher photoactivity than pristine TiO2. We believe that this liquid plasma hydrogenation strategy can be extended to other photocatalysts with mass stable surface defects. Overall, our gray TiO2−x with improved solar energy absorption possesses potential for more efficient polluted water cleaning.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/2079-4991/10/2/342/s1, Figure S1: (A) The selected area for elements analysis in TEM of Ti420, and (B) the EDS resulte, Table S1: The integrated peak area of the diffraction peak at (101) facet.

Author Contributions

Data curation and writing—original draft preparation, F.Z.; conceptualization, methodology, supervision, and editing, G.F.; visualization and investigation, M.H.; writing—reviewing, Y.H.; writing—reviewing and funding acquisition, H.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the National Key Research and Development Program of China (2018YFB0407100, 2018YFB0504400), the National Natural Science Foundation of China (11727812 and 11621404), and Key Projects from Shanghai S&T Commission (18JC1412000) and Shanghai Educational Commission (2017-01-07-00-05-E000021).

Conflicts of Interest

The authors declare no competing financial interest.

References

  1. Fujishima, A.; Honda, K. Electrochemical photolysis of water at a semiconductor electrode. Nature 1972, 238, 37–38. [Google Scholar] [CrossRef]
  2. Grätzel, M. Photoelectrochemical cells. Nature 2001, 414, 338–344. [Google Scholar] [CrossRef]
  3. Linsebigler, A.L.; Lu, G.; Yates, J.T. Photocatalysis on TiO2 surfaces: Principles, mechanisms, and selected results. Chem. Rev. 1995, 95, 735–758. [Google Scholar] [CrossRef]
  4. Guo, Q.; Zhang, Z.; Ma, X.; Jing, K.; Shen, M.; Yu, N.; Tang, J.; Dionysiou, D.D. Preparation of N,F-codoped TiO2 nanoparticles by three different methods and comparison of visible-light photocatalytic performances. Sep. Purif. Technol. 2017, 175, 305–313. [Google Scholar] [CrossRef]
  5. Chen, Q.; Zhang, Y.; Zhang, D.; Yang, Y. Ag and N co-doped TiO2 nanostructured photocatalyst for printing and dyeing wastewater. J. Water Process Eng. 2017, 16, 14–20. [Google Scholar] [CrossRef]
  6. Daté, M.; Haruta, M. Moisture effect on CO oxidation over Au/TiO2 catalyst. J. Catal. 2001, 201, 221–224. [Google Scholar] [CrossRef]
  7. Vaiano, V.; Iervolino, G.; Sannino, D.; Murcia, J.J.; Hidalgo, M.C.; Ciambelli, P.; Navío, J.A. Photocatalytic removal of patent blue V dye on Au-TiO2 and Pt-TiO2 catalysts. Appl. Catal. B Environ. 2016, 188, 134–146. [Google Scholar] [CrossRef]
  8. Lin, Y.-H.; Hsueh, H.-T.; Chang, C.-W.; Chu, H. The visible light-driven photodegradation of dimethyl sulfide on S-doped TiO2: Characterization, kinetics, and reaction pathways. Appl. Catal. B Environ. 2016, 199, 1–10. [Google Scholar] [CrossRef]
  9. Xiang, P.; Lv, F.; Xiao, T.; Jiang, L.; Tan, X.; Shu, T. Improved performance of quasi-solid-state dye-sensitized solar cells based on iodine-doped TiO2 spheres photoanodes. J. Alloy. Compd. 2018, 741, 1142–1147. [Google Scholar] [CrossRef]
  10. Chen, X.; Burda, C. The Electronic origin of the visible-light absorption properties of C-, N- and S-doped TiO2 nanomaterials. J. Am. Chem. Soc. 2008, 130, 5018–5019. [Google Scholar] [CrossRef]
  11. Zhang, K.; Park, J.H. Surface localization of defects in black TiO2: Enhancing photoactivity or reactivity. J. Phys. Chem. Lett. 2017, 8, 199–207. [Google Scholar] [CrossRef] [PubMed]
  12. Wu, P.; Dai, S.; Chen, G.; Zhao, S.; Xu, Z.; Fu, M.; Chen, P.; Chen, Q.; Jin, X.; Qiu, Y.; et al. Interfacial effects in hierarchically porous α-MnO2/Mn3O4 heterostructures promote photocatalytic oxidation activity. Appl. Catal. B Environ. 2019, 118418. [Google Scholar] [CrossRef]
  13. Zuo, F.; Wang, L.; Wu, T.; Zhang, Z.; Borchardt, D.; Feng, P. Self-doped Ti3+ enhanced photocatalyst for hydrogen production under visible light. J. Am. Chem. Soc. 2010, 132, 11856–11857. [Google Scholar] [CrossRef] [PubMed]
  14. Xia, T.; Chen, X. Revealing the structural properties of hydrogenated black TiO2 nanocrystals. J. Mater. Chem. A 2013, 1, 2983–2989. [Google Scholar] [CrossRef]
  15. Chen, X.; Liu, L.; Huang, F. Black titanium dioxide (TiO2) nanomaterials. Chem. Soc. Rev. 2015, 44, 1861–1885. [Google Scholar] [CrossRef]
  16. Li, L.; Song, L.; Zhu, L.; Yan, Z.; Cao, X. Black TiO2−x with stable surface oxygen vacancies as the support of efficient gold catalysts for water-gas shift reaction. Catal. Sci. Technol. 2018, 8, 1277–1287. [Google Scholar] [CrossRef]
  17. Liu, N.; Schneider, C.; Freitag, D.; Hartmann, M.; Venkatesan, U.; Müller, J.; Spiecker, E.; Schmuki, P. Black TiO2 nanotubes: Cocatalyst-free open-circuit hydrogen generation. Nano Lett. 2014, 14, 3309–3313. [Google Scholar] [CrossRef]
  18. Panomsuwan, G.; Watthanaphanit, A.; Ishizaki, T.; Saito, N. Water-plasma-assisted synthesis of black titania spheres with efficient visible-light photocatalytic activity. Phys. Chem. Chem. Phys. 2015, 17, 13794–13799. [Google Scholar] [CrossRef]
  19. Zhang, H.; Cai, J.; Wang, Y.; Wu, M.; Meng, M.; Tian, Y.; Li, X.; Zhang, J.; Zheng, L.; Jiang, Z.; et al. Insights into the effects of surface/bulk defects on photocatalytic hydrogen evolution over TiO2 with exposed 001 facets. Appl. Catal. B Environ. 2018, 220, 126–136. [Google Scholar] [CrossRef]
  20. Kong, M.; Li, Y.; Chen, X.; Tian, T.; Fang, P.; Zheng, F.; Zhao, X. Tuning the relative concentration ratio of bulk defects to surface defects in TiO2 nanocrystals leads to high photocatalytic efficiency. J. Am. Chem. Soc. 2011, 133, 16414–16417. [Google Scholar] [CrossRef]
  21. Ullattil, S.G.; Periyat, P. Microwave-power induced green synthesis of randomly oriented mesoporous anatase TiO2 nanoparticles for efficient dye sensitized solar cells. Sol. Energy 2017, 147, 99–105. [Google Scholar] [CrossRef]
  22. Ullattil, S.G.; Narendranath, S.B.; Pillai, S.C.; Periyat, P. Black TiO2 nanomaterials: A review of recent advances. Chem. Eng. J. 2018, 343, 708–736. [Google Scholar] [CrossRef]
  23. Lepcha, A.; Maccato, C.; Mettenbörger, A.; Andreu, T.; Mayrhofer, L.; Walter, M.; Olthof, S.; Ruoko, T.P.; Klein, A.; Moseler, M.; et al. Electrospun black titania nanofibers: Influence of hydrogen plasma-induced disorder on the electronic structure and photoelectrochemical performance. J. Phys. Chem. C 2015, 119, 18835–18842. [Google Scholar] [CrossRef]
  24. Zimbone, M.; Buccheri, M.A.; Cacciato, G.; Sanz, R.; Rappazzo, G.; Boninelli, S.; Reitano, R.; Romano, L.; Privitera, V.; Grimaldi, M.G. Photocatalytical and antibacterial activity of TiO2 nanoparticles obtained by laser ablation in water. Appl. Catal. B Environ. 2015, 165, 487–494. [Google Scholar] [CrossRef]
  25. Feng, G.; Wu, B.; Khan, A.Q.; Zeng, H. In situ glow discharge plasma electrolytic synthesis of reduced TiO2 for enhanced visible light photocatalysis. Mater. Res. Express 2018, 5, 055022. [Google Scholar] [CrossRef]
  26. Jedsukontorn, T.; Ueno, T.; Saito, N.; Hunsom, M. Facile preparation of defective black TiO2 through the solution plasma process: Effect of parametric changes for plasma discharge on its structural and optical properties. J. Alloy. Compd. 2017, 726, 567–577. [Google Scholar] [CrossRef]
  27. Wang, K.; Xu, M.; Gu, Y.; Gu, Z.; Liu, J.; Fan, Q.H. Low-temperature plasma exfoliated n-doped graphene for symmetrical electrode supercapacitors. Nano Energy 2017, 31, 486–494. [Google Scholar] [CrossRef] [Green Version]
  28. Mauchauffé, R.; Bonot, S.; Moreno-Couranjou, M.; Detrembleur, C.; Boscher, N.D.; Van De Weerdt, C.; Duwez, A.-S.; Choquet, P. Fast Atmospheric plasma deposition of bio-inspired catechol/quinone-rich nanolayers to immobilize NDM-1 enzymes for water treatment. Adv. Mater. Interfaces 2016, 3, 1500520. [Google Scholar] [CrossRef]
  29. Gupta, P.; Tenhundfeld, G.; Daigle, E.O.; Ryabkov, D. Electrolytic plasma technology: Science and engineering—An overview. Surf. Coat. Technol. 2007, 201, 8746–8760. [Google Scholar] [CrossRef]
  30. Bruggeman, P.; Leys, C. Non-thermal plasmas in and in contact with liquids. J. Phys. D Appl. Phys. 2009, 42, 053001. [Google Scholar] [CrossRef]
  31. Chen, Q.; Li, J.; Li, Y. A review of plasma–liquid interactions for nanomaterial synthesis. J. Phys. D Appl. Phys. 2015, 48, 424005. [Google Scholar] [CrossRef] [Green Version]
  32. Bardos, L.; Baránková, H.; Bardos, A. Production of hydrogen-rich synthesis gas by pulsed atmospheric plasma submerged in mixture of water with ethanol. Plasma Chem. Plasma Process. 2017, 37, 115–123. [Google Scholar] [CrossRef] [Green Version]
  33. Zhu, Q.; Peng, Y.; Lin, L.; Fan, C.-M.; Gao, G.-Q.; Wang, R.-X.; Xu, A.-W. Stable blue TiO2−x nanoparticles for efficient visible light photocatalysts. J. Mater. Chem. A 2014, 2, 4429–4437. [Google Scholar] [CrossRef]
  34. Kuznetsov, V.N.; Serpone, N. On the origin of the spectral bands in the visible absorption spectra of visible-light-active TiO2 specimens analysis and assignments. J. Phys. Chem. C 2009, 113, 15110–15123. [Google Scholar] [CrossRef]
  35. Wang, Y.; Feng, C.; Zhang, M.; Yang, J.; Zhang, Z. Visible light active N-doped TiO2 prepared from different precursors: Origin of the visible light absorption and photoactivity. Appl. Catal. B Environ. 2011, 104, 268–274. [Google Scholar] [CrossRef]
  36. Moslah, C.; Aguilar, T.; Alcántara, R.; Ksibi, M.; Navas, J. Synthesis of W-doped TiO2 by low-temperature hydrolysis: Effects of annealing temperature and doping content on the surface microstructure and photocatalytic activity. J. Chin. Chem. Soc. 2019, 66, 99–109. [Google Scholar] [CrossRef] [Green Version]
  37. Hao, W.; Li, X.; Qin, L.; Han, S.; Kang, S.-Z. Facile preparation of Ti3+ self-doped TiO2 nanoparticles and their dramatic visible photocatalytic activity for the fast treatment of highly concentrated Cr(vi) effluent. Catal. Sci. Technol. 2019, 9, 2523–2531. [Google Scholar] [CrossRef]
  38. Zheng, Z.; Huang, B.; Lu, J.; Wang, Z.; Qin, X.; Zhang, X.; Dai, Y.; Whangbo, M.-H. Hydrogenated titania: Synergy of surface modification and morphology improvement for enhanced photocatalytic activity. Chem. Commun. 2012, 48, 5733–5735. [Google Scholar] [CrossRef]
  39. Li, L.; Yu, L.; Lin, Z.; Yang, G. Reduced TiO2-graphene oxide heterostructure as broad spectrum-driven efficient water-splitting photocatalysts. ACS Appl. Mater. Interfaces 2016, 8, 8536–8545. [Google Scholar] [CrossRef]
  40. Hamdy, M.S.; Saputera, W.H.; Groenen, E.J.; Mul, G. A novel TiO2 composite for photocatalytic wastewater treatment. J. Catal. 2014, 310, 75–83. [Google Scholar] [CrossRef]
  41. Chen, X.; Liu, L.; Yu, P.Y.; Mao, S.S. Increasing Solar absorption for photocatalysis with black hydrogenated titanium dioxide nanocrystals. Science 2011, 331, 746. [Google Scholar] [CrossRef] [PubMed]
  42. Deskins, N.A.; Rousseau, R.; Dupuis, M. Distribution of Ti3+ surface sites in reduced TiO2. J. Phys. Chem. C 2011, 115, 7562–7572. [Google Scholar] [CrossRef]
  43. Yu, X.; Kim, B.; Kim, Y.K. Highly enhanced photoactivity of anatase TiO2 nanocrystals by controlled hydrogenation-induced surface defects. ACS Catal. 2013, 3, 2479–2486. [Google Scholar] [CrossRef]
  44. Teng, F.; Li, M.; Gao, C.; Zhang, G.; Zhang, P.; Wang, Y.; Chen, L.; Xie, E. Preparation of black TiO2 by hydrogen plasma assisted chemical vapor deposition and its photocatalytic activity. Appl. Catal. B Environ. 2014, 148–149, 339–343. [Google Scholar] [CrossRef]
  45. Zhang, Y.; Xing, Z.; Liu, X.; Li, Z.; Wu, X.; Jiang, J.; Li, M.; Zhu, Q.; Zhou, W. Ti3+ self-doped blue TiO2(b) single-crystalline nanorods for efficient solar-driven photocatalytic performance. ACS Appl. Mater. Interfaces 2016, 8, 26851–26859. [Google Scholar] [CrossRef]
  46. Zhou, Y.; Chen, C.; Wang, N.; Li, Y.; Ding, H. Stable Ti3+ Self-doped anatase-rutile mixed TiO2 with enhanced visible light utilization and durability. J. Phys. Chem. C 2016, 120, 6116–6124. [Google Scholar] [CrossRef]
  47. Chen, S.; Wang, Y.; Li, J.; Hu, Z.; Zhao, H.; Xie, W.; Wei, Z. Synthesis of black TiO2 with efficient visible-light photocatalytic activity by ultraviolet light irradiation and low temperature annealing. Mater. Res. Bull. 2018, 98, 280–287. [Google Scholar] [CrossRef]
  48. Pei, Z.; Ding, L.; Feng, W.; Weng, S.; Liu, P. Defect self-doped TiO2 for visible light activity and direct noble metal anchoring. Phys. Chem. Chem. Phys. 2014, 16, 21876–21881. [Google Scholar] [CrossRef]
  49. An, H.-R.; Park, S.Y.; Kim, H.; Lee, C.Y.; Choi, S.; Lee, S.C.; Seo, S.; Park, E.C.; Oh, Y.-K.; Song, C.-G.; et al. Advanced nanoporous TiO2 photocatalysts by hydrogen plasma for efficient solar-light photocatalytic application. Sci. Rep. 2016, 6, 29683. [Google Scholar] [CrossRef]
  50. Ariyanti, D.; Mills, L.; Dong, J.; Yao, Y.; Gao, W. NaBH4 modified TiO2: Defect site enhancement related to its photocatalytic activity. Mater. Chem. Phys. 2017, 199, 571–576. [Google Scholar] [CrossRef]
  51. Li, Y.; Wang, W.; Wang, F.; Di, L.; Yang, S.; Zhu, S.; Yao, Y.; Ma, C.; Dai, B.; Yu, F. Enhanced photocatalytic degradation of organic dyes via defect-rich TiO2 prepared by dielectric barrier discharge plasma. Nanomaterials 2019, 9, 720. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Liang, Y.; Wang, H.; Sanchez Casalongue, H.; Chen, Z.; Dai, H. TiO2 nanocrystals grown on graphene as advanced photocatalytic hybrid materials. Nano Res. 2010, 3, 701–705. [Google Scholar] [CrossRef] [Green Version]
  53. Pol, R.; Guerrero, M.; García-Lecina, E.; Altube, A.; Rossinyol, E.; Garroni, S.; Baró, M.D.; Pons, J.; Sort, J.; Pellicer, E. Ni-, Pt- and (Ni/Pt)-doped TiO2 nanophotocatalysts: A smart approach for sustainable degradation of Rhodamine B dye. Appl. Catal. B Environ. 2016, 181, 270–278. [Google Scholar] [CrossRef] [Green Version]
Figure 1. (A) The setup for liquid plasma generation, (B) digital picture of as-produced liquid plasma, and (C) digital pictures for gray titanium dioxide (TiO2−x) and pristine white TiO2.
Figure 1. (A) The setup for liquid plasma generation, (B) digital picture of as-produced liquid plasma, and (C) digital pictures for gray titanium dioxide (TiO2−x) and pristine white TiO2.
Nanomaterials 10 00342 g001
Figure 2. The synthesis mechanism of gray TiO2−x nanosphere assisted by liquid plasma and the corresponding optical emission spectrum originated from cathodic Ti electrodes.
Figure 2. The synthesis mechanism of gray TiO2−x nanosphere assisted by liquid plasma and the corresponding optical emission spectrum originated from cathodic Ti electrodes.
Nanomaterials 10 00342 g002
Figure 3. XRD diffraction patterns of gray TiO2−x samples prepared with different glow discharge powers, and black line refers to anatase TiO2.
Figure 3. XRD diffraction patterns of gray TiO2−x samples prepared with different glow discharge powers, and black line refers to anatase TiO2.
Nanomaterials 10 00342 g003
Figure 4. TEM images of (A,C) pristine TiO2 and (B,D) gray TiO2−x.
Figure 4. TEM images of (A,C) pristine TiO2 and (B,D) gray TiO2−x.
Nanomaterials 10 00342 g004
Figure 5. (A) UV–Vis diffuse reflectance spectra of as-prepared samples and the digital photos shown in the inset. (B) The plot of (αhν)1/2 versus hν using the Kubelka–Munk function.
Figure 5. (A) UV–Vis diffuse reflectance spectra of as-prepared samples and the digital photos shown in the inset. (B) The plot of (αhν)1/2 versus hν using the Kubelka–Munk function.
Nanomaterials 10 00342 g005
Figure 6. Raman spectra of pristine TiO2 and gray TiO2−x. The inset shows an enlarged view of Eg mode of pristine TiO2 and gray TiO2−x.
Figure 6. Raman spectra of pristine TiO2 and gray TiO2−x. The inset shows an enlarged view of Eg mode of pristine TiO2 and gray TiO2−x.
Nanomaterials 10 00342 g006
Figure 7. (A) XPS survey spectra, (B) Ti 2p XPS spectra, and (C) O 1s XPS spectra of samples.
Figure 7. (A) XPS survey spectra, (B) Ti 2p XPS spectra, and (C) O 1s XPS spectra of samples.
Nanomaterials 10 00342 g007
Figure 8. EPR(Electron paramagnetic resonance) spectra of the as-synthesized gray TiO2−x samples obtained with different plasma discharge powers.
Figure 8. EPR(Electron paramagnetic resonance) spectra of the as-synthesized gray TiO2−x samples obtained with different plasma discharge powers.
Nanomaterials 10 00342 g008
Figure 9. (A) Photocatalytic degradation curves of RhB(Rhodamine B) over different photocatalysts under visible-light irradiation. (B) The pseudo-first-order reaction rate constant for all samples. (C) Recycling test results using GT-480 as the photocatalyst. (D) Photodegradation rates of methyl orange and phenol with TiO2 and GT-480 under visible-light irradiation.
Figure 9. (A) Photocatalytic degradation curves of RhB(Rhodamine B) over different photocatalysts under visible-light irradiation. (B) The pseudo-first-order reaction rate constant for all samples. (C) Recycling test results using GT-480 as the photocatalyst. (D) Photodegradation rates of methyl orange and phenol with TiO2 and GT-480 under visible-light irradiation.
Nanomaterials 10 00342 g009
Table 1. Comparison of rhodamine B photodegradation efficiency over gray TiO2−x with reported colored titania and highly efficient photocatalysts.
Table 1. Comparison of rhodamine B photodegradation efficiency over gray TiO2−x with reported colored titania and highly efficient photocatalysts.
PhotocatalystSynthesis MethodLight SourceCRhBWCatPerformanceRef.
Gray TiO2Liquid-plasma hydrogenation of anatase TiO2Visible light
300 W
50 mL RhB, 20 ppm50 mgT90% = 20 minThis work
Black TiO2Hydrogen plasma assisted chemical vapor depositionSolar light
50 W
50 mL RhB, 2 mg/L10 mgT90% = 20 min[44]
Blue TiO2Hydrothermal treatment and annealing in argon environmentVisible light
350 W
30 mL RhB, 10 mg/L20 mgT90% = 150 min[45]
Blue TiO2Solvothermal reaction with TiCl3 solutionVisible light
500 W
50 mL RhB, 10 mg/L100 mgT90% = 110 min[46]
Black TiO2Ultraviolet light irradiation and low temperature annealingSolar light
500 W
40 mL RhB, 4×10−5 M10 mgT90% = 130 min[47]
Gray TiO2Anodization and liquid plasma hydrogenationVisible light
300 W
50 mL RhB, 20 ppm50 mgT90% = 70 min[25]
Brown TiO2Noble metal deposited on defective TiO2−xVisible light
300 W
10 uM RhB, 80 mL80 mgT90% = 120 min[48]
Gray TiO2Hydrogen plasma-treated nanoporous TiO2Solar light
150 W
3 mg/mL RhB,0.5 g/LT90% = 90 min[49]
Black TiO2Reduced by NaBH4 under argon atmosphere.Solar light
300 W
50 mL RhB, 5 ppm50 mgT90% = 50 min[50]
Ar-TiO2Argon plasma treatments in a dielectric barrier dischargeVisible light
300 W
100 mL RhB 10 mg/L50 mgT90% = 90 min[51]
Graphene-P25Hydrothermal reaction of graphene and P25Solar light
100 W
100 mL RhB, 0.03 mmol/L6.3 mgT90% = 15 min[52]
Pt-doped nanoporous TiO2Sol-gel hydrothermal and annealing under H2 atmosphere320–500 nm
220 W
10 mL RhB, 10 ppm3 mgT90% > 180 min[53]
CRhB: Initial RhB concentration; Wcat: Catalyst usage; and T90%: Time 90% conversion of RhB; Ref.: Reference.

Share and Cite

MDPI and ACS Style

Zhang, F.; Feng, G.; Hu, M.; Huang, Y.; Zeng, H. Liquid-Plasma Hydrogenated Synthesis of Gray Titania with Engineered Surface Defects and Superior Photocatalytic Activity. Nanomaterials 2020, 10, 342. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10020342

AMA Style

Zhang F, Feng G, Hu M, Huang Y, Zeng H. Liquid-Plasma Hydrogenated Synthesis of Gray Titania with Engineered Surface Defects and Superior Photocatalytic Activity. Nanomaterials. 2020; 10(2):342. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10020342

Chicago/Turabian Style

Zhang, Feng, Guang Feng, Mengyun Hu, Yanwei Huang, and Heping Zeng. 2020. "Liquid-Plasma Hydrogenated Synthesis of Gray Titania with Engineered Surface Defects and Superior Photocatalytic Activity" Nanomaterials 10, no. 2: 342. https://0-doi-org.brum.beds.ac.uk/10.3390/nano10020342

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop