Next Article in Journal
Thermal Stability and Potential Cycling Durability of Nitrogen-Doped Graphene Modified by Metal-Organic Framework for Oxygen Reduction Reactions
Previous Article in Journal
Organocatalysis: A Brief Overview on Its Evolution and Applications
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Criticial Assessment of the Photocatalytic Reduction of Cr(VI) over Au/TiO2

1
Leibniz-Institute for Catalysis, Albert-Einstein-Str. 29a, D-18059 Rostock, Germany
2
School of Chemical Engineering, Hanoi University of Science and Technology, Dai Co Viet Street no.1, 100000 Ha Noi, Vietnam
3
Institute for Chemistry, University of Rostock, Albert-Einstein-Str. 3a, D-18059 Rostock, Germany
4
Department of Life, Light and Matter, Interdisciplinary Faculty, University of Rostock, Albert-Einstein-Str. 25, D-18059 Rostock, Germany
*
Author to whom correspondence should be addressed.
Catalysts 2018, 8(12), 606; https://doi.org/10.3390/catal8120606
Submission received: 3 November 2018 / Revised: 19 November 2018 / Accepted: 23 November 2018 / Published: 3 December 2018
(This article belongs to the Section Photocatalysis)

Abstract

:
The purification of drinking water is one of the most urgent challenges in developing countries, for which the efficient removal of traces of heavy metals, e.g., Cr(IV), represents a key technology. This can be achieved via photocatalysis. In this study, we compare the performance of Au/TiO2 to bare TiO2 P25 catalysts. Furthermore, the influence of the sacrificial reagent citric acid under UV-Vis and Vis excitation was investigated and a detailed investigation of the catalysts before and after reaction was performed. During the photocatalytic reduction of Cr(IV) under acidic conditions, both leaching of Au, as well as absorption of Cr, occur, resulting in new catalyst systems obtained in situ.

Graphical Abstract

1. Introduction

Nowadays, access to clean and accessible drinking water is one of the key challenges for humankind. Therefore, it was included among the 17 United Nations Sustainability Development Goals in 2017. Especially in developing countries, the rapid growth of industries led to an increase in wastewater without proper management and purification. Most of the water contaminants are harmful to both the environment and human health. Besides toxic organic compounds, heavy metals in high concentrations are also present in wastewater. Among these, Cr(VI) is considered to be one of the most toxic. Chromium compounds have been applied in many industries such as painting, furnace linings, tanning and dying processes, photography, and steel [1]. Chromium has two major stable oxidation states: Cr(VI) and Cr(III). While a moderate intake of Cr(III) is considered to be necessary for health, Cr(VI) is known to be very harmful even in small concentrations. Exposure (inhalation, ingestion, skin and eye contact) may cause cancer [2]. Therefore, Cr(VI) needs to be treated thoroughly before being released into the environment.
Many methods for Cr(VI) treatment have been investigated such as absorption by carbon materials with large surface area, ion exchange, membrane filtration, and chemical and electrochemical methods [3]. However, these methods have disadvantages: Absorption, ion exchange, and membrane filtration cannot thoroughly handle Cr(VI), since new toxic materials are produced. Chemical methods have the advantage of being simple but they imply the consumption of chemicals in a strongly acidic environment. Afterwards, the pH value must be raised to basic conditions to precipitate Cr(OH)3. Electrochemical methods just remove Cr(VI) at high concentrations, hence, the wastewater from electroplating still has a significant amount of Cr(VI). Thus, it is of high interest to purify contaminated water that has a very small concentration of Cr (<50 ppm).
Photocatalysis is one promising method to reduce Cr(VI) to Cr(III). The advantage of this process is the complementary degradation of organic compounds together with the detoxification of heavy metals. One of the most popular photocatalysts is TiO2 because of its high photoactivity, stability, non-toxicity, and low cost. However, with a large band gap energy (3.0–3.2 eV) TiO2 can only be excited in the ultraviolet light region, making it ineffective under sunlight. In the past, many researchers tried to narrow the band gap energy of TiO2 by doping with anions (e.g., N, C) [4] or cations (e.g., Sb, Cr, Fe, V, Mo) [5]. However, these methods had limited success, due to the fast recombination of conduction band electrons and holes [6]. Recently, a new generation of photocatalyst has been investigated by depositing noble metals e.g., Au on TiO2 surfaces. The excitation of the so-called surface plasmon resonance (SPR) effect of the Au nanoparticles (NPs) enables the injection of hot electrons to TiO2 in visible light [7].
During the photocatalytic reduction process, Cr(VI) is converted to Cr(V)/Cr(IV)/Cr(III) by conduction band electrons, see Figure 1 [8]. The photogenerated holes are able to oxidize H2O and also to re-oxidize Cr(III) to Cr(VI), Cr(V), and Cr(IV) [9]. This re-oxidation reduces the overall efficiency of the Cr(VI) reduction process. Therefore, the presence of a sacrificial agent (SR), which is easier to oxidize than H2O, prevents the re-oxidation of Cr(III) as well as decreases the occurrence of the recombination of electron and holes which can increase the photocatalytic performance.
Using citric acid (2-hydroxyl-propane-1,2,3-tricarboxylic acid) (CA), a common ingredient in cosmetics, pharmaceuticals, dietary supplements, and food industries, the photogenerated holes of TiO2 can be easily oxidized [10]. Furthermore, citric acid can generate a stable Cr(V)-Cit complex which enables reaction monitoring by EPR spectroscopy [9].
We became interested in the Cr(VI) reduction during our investigation of proton reduction of water with advanced Au/TiO2 [11]. Indeed, several publications have already described the improved catalytic Cr(VI) reduction on an Au/TiO2 catalyst, see Figure 1 [12,13,14,15], but these are limited to special conditions, e.g., selected irradiation, pH, or SR. In principle, most electronic processes are known. However, investigation regarding the processes on the surface of the catalyst e.g., adsorption/desorption during irradiation is still needed. Unexpected results raised some questions about the role of Au, Cr, TiO2(P25), and CA.
In this communication, we present a detailed investigation of the Cr(VI) reduction involving reactions with and without P25, Au/TiO2 catalysts prepared by Deposition–Precipitation (DP) [11] (Au/P25-DP) as well as calcined Au/P25 DP (Au/P25-Cal) catalysts, see Figure 2. The Au/P25 catalyst, prepared by the DP method, comprises of gold hydroxide [11]. Calcination of this catalyst results in the formation of pure Au NPs on P25 (Au/P25-Cal).
All catalysts were examined with and without citric acid as the electron donor under UV-Vis and Vis irradiation. The catalysts were characterized before and after reaction by ICP and XRD. Additionally, selected catalysts were characterized by UV-Vis DRS, XPS, and TEM.

2. Results

2.1. Photocatalytic Activity

In general, the concentration of Cr(VI) was determined by UV-Vis without an additional complexation agent (absorbance at 350 nm, see more information in the experimental section). The catalytic performance was evaluated with a standard solution of Cr(VI) (20 ppm, pH 2, adjusted by H2SO4 0.4 M). Three photocatalysts were investigated for Cr(VI) reduction: P25, Au/P25-DP, Au/P25-Cal. Photoreactions were carried out under UV-Vis (Xe-Lamp 300 W, no filter) or Vis excitation (cutoff filter 420 nm), in the absence of CA and the presence CA (molar ratio CA/Cr(VI) 25/1), see Table 1. After centrifugation, the conversion was detected after a 3-h reaction time by UV-Vis spectroscopy.
Under exclusion of light, no conversion was detected, indicating the negligible adsorption of Cr(VI) in the dark. Thus, the Cr(VI) conversion obtained occurs only by photoreaction. Under UV-Vis light (without CA) the photoactivity of P25 (75%) is much higher than with both Au/P25 catalysts (44% and 48%). This was a surprise since all other reports show a higher activity for Au/TiO2 [12,13,14,15]. Under visible irradiation, only Au/P25-Cal showed photoactivity, indicating the necessity of the Au NPs on the catalyst. However, the efficiency of Cr(VI) reduction was moderate; after reaction for 3 h, just 18% of Cr(VI) was reduced. In contrast, the non-existent activity of Au/P25-DP in visible light indicates that under these conditions no Au NPs are formed from Au(OH)3 under visible irradiation, which is in contrast to our recent investigations [11].
In the presence of citric acid (CA/Cr = 25/1), all catalysts completely reduced Cr(VI) to Cr(III) under UV light. It demonstrates the high necessity of citric acid as SA in photoreaction. Unexpectedly, citric acid alone can reduce Cr(VI) under UV-Vis. Here, 74% of Cr(VI) was reduced under UV-Vis light, and even 32% was reduced under visible light, raising the suspicion that a photoreaction between CA and Cr(VI) might take place. Indeed, with visible light and in the presence of CA, high conversions were still achieved even with pure P25 and Au/P25-DP. Although, Au/P25-DP and P25 have no absorption in the visible range. Even pure P25 showed a higher conversion than the Au-containing catalysts. Thus, the roles of these catalysts could be clarified by the results from the catalytic characterization methods, as described below.

2.2. Catalytic Characterization

2.2.1. UV-Vis Diffuse Reflectance Spectroscopy (UV-Vis DRS)

First, we were interested in the change of the absorbance of the SPR band in the Au/TiO2 catalysts by UV-Vis spectroscopy, see Figure 3. All photocatalysts showed strong absorbance in the region below 387 nm due to the P25 support. Au/TiO2-DP (before reaction) shows a broad absorption band in the visible region due to Au(OH)3. After reaction under UV light, a broad absorption at 553 nm was observed, indicating the formation of a low amount of Au NPs generated by UV light [16]. Au/P25-Cal showed a maximum absorbance at 600 nm, indicating the formation and agglomeration towards large Au NPs [11]. However, the absorbance in the visible region of Au/P25-Cal is much lower than in the UV region. This might be the reason for the low yield of Cr(VI) reduction under visible light (only 18%).

2.2.2. Chemical Composition of the Photocatalysts

Next, we analyzed the elemental composition of the catalysts before and after reaction by ICP-OES; the results of which are provided in Table 2. The main objective of this analysis was the detection of the Au that was still deposited on the surface after the reaction.
By the used deposition—precipitation method, 0.62 wt% of Au can be deposited. However, a significant decrease of the gold content was observed after the reaction. Overall, the gold content decreased more in Au/P25-DP than in the calcined sample. Without CA and visible light, nearly all gold was removed from the sample, which could explain the loss of activity. Indeed, in Au/P25-DP, Au(OH)3 or Au2O3 is present on the surface. This Au can be easily dissolved in an acidic environment or chelated and leached out by CA. This was not reported previously. However, if the Au NPs are already present in the starting material (e.g., Au/P25-Cal), then leaching and complexation are reduced.
Interestingly, we detected a significant amount of Cr in the catalyst sample after the reaction, see Table 2. Indeed, the results showed that, depending on the condition, chromium was chemically absorbed on the surface of the catalyst. In general, the chromium content on P25 was more than on the Au/P25 catalysts. With P25 and without CA under UV-VIS, nearly 100% of the converted Cr is present on the catalyst.
The valence state of the adsorbed chromium and gold on catalysts were measured by X-ray photoelectron spectroscopy, see Figure 4. Analysis from Au/P25-DP after reaction (no CA/with UV-VIS), showed the presence of Cr(III) (Cr 2p, 587 and 577.3 eV) and Au(0) (Au 4f, 87 and 83 eV) on the catalyst surface [11]. No Au(III) or Au(I) was detected which could be caused by the removal under acidic condition and/or CA.

2.2.3. Crystal Structure and TEM

However, the state and distribution of Cr(III) is still unclear. Thus, the crystal structure of the photocatalysts were analyzed by powder X-ray diffraction. The results are shown in Figure 5.
The results of the X-ray diffraction analysis showed that the crystal structure of Anatase and Rutile in P25 did not change after modification by deposited Au NPs, nor after reaction with Cr(VI). Especially in Au/P25-Cal, see Figure 5e, and in Au/P25-DP after the reaction, see Figure 5d, Au NPs were detected due to the presence of characteristic diffraction peaks at 38.1°, 44.5°, 64.5°, and 77.5°. However, no reflections from Cr2O3 were detected. Although, the presence of Cr(III) on the catalyst after the reaction has been proved by ICP-OES and XPS. These apparently contradictory results could be explained by the dispersion of the low content of Cr(III) on the TiO2 surface or the formation of amorphous Cr species.
To our disappointment, even transmission electron microscopy (TEM) cannot distinguish Cr and Ti from each other. The difference in the relative mass of Cr and Ti is too small to separate these elements. Furthermore, the amount of Cr is quite low. Only the distribution and size of the Au NPs (10–20 nm in Au/P25-DP under UV-Vis light) can be provided by TEM, see Figure 6, which is in accordance with the results published previously [11].

2.2.4. Discussion of the Mechanism

The combined results indicate very important points which should always be addressed together with the Cr(IV) conversion, but have not been reported together before:
In the presence of CA or another SA, intermediate Cr(V)-CA complexes are formed which are photocatalytically (redox) active. This can distort the conversion of different photocatalysts, especially, under visible light.
The formation of the Cr(V)-CA complex was previously reported and can be proven by Electron Paramagnetic Resonance (EPR) Spectroscopy, see Figure 7A. The conversion of Cr(V)-CA under visible and UV-Vis light is shown in Figure 7B. Here, the Cr(V)-CA complex showed fast conversion in UV-Vis as well as slow conversion with visible light, even without P25. It can be assumed that the colored Cr(V)-CA complex can act as a photosensitizer.
In bare P25, the reduction of Cr(V) occurs only at the active centers on the TiO2 surface. This is associated with adsorption of the formed Cr(III) on the surface of TiO2, see Figure 8. Depending on the condition, complete adsorption or partial dissolution can occur in the presence of CA.
The newly formed Cr(III)/P25 catalyst has a lower activity than bare P25. For example, the reuse of P25 (after full conversion with CA and UV-VIS) results only in 70% conversion, see Table 1 and also the literature mentioned by Litter [8]. It can be assumed that the active reduction sites are blocked by Cr(III) adsorbates.
Under acidic conditions, no Au NPs are formed from Au(OH)3. Instead, Au(OH)3 is dissolved out from the TiO2 surface, see Figure 9. This dissolution can be enhanced by the addition of CA. If Au NPs are present from the beginning, even partial dissolution of the Au NP can occur under acidic conditions. Under photocatalytic conditions, the reduction to Cr(III) occurs at the Au NP surface. Thus, the Au NP may be covered by Cr(III), see Figure 10. This is well known from the preparation of Cr/Rh core-shell catalysts [17,18,19]. However, the formation of an analog Au–Cr core-shell NPs still has to be proven.
In principle, the adsorption of Cr always results in a change of the photocatalysts independent from the substructure; therefore, a new Cr-containing photocatalyst is formed. Furthermore, assuming batch reactions, even this new catalyst will be continually changed during Cr(VI) reduction. This change of the catalyst systems should be considered quite critically and is not discussed in literature. In principle, under these conditions, TiO2 and Au/TiO2 catalysts are unsuitable for this type of photocatalytic reduction.
Very often, conversion of Cr(IV) is accompanied with an adsorption (mentioned as removal) [20]; although, these processes work in opposing directions. Thus, some strategies for the full recovery of the initial photocatalyst need to be developed in the future.

3. Materials and Methods

3.1. Materials

Au/TiO2 catalysts were synthesized by TiO2 P25 (anatase/rutile = 85:15, Evonik, S BET = 50 m2/g) and HAuCl4·3H2O (Aldrich, Germany). K2Cr2O7 (CAS 7778-50-9) was used for Cr(VI) solution. Citric acid C6H8O7 (CAS 5949-29-1) 99–102% was purchased from Alfa Aesar (Karlsruhe, Germany). H2SO4 and NaOH were used for adjusting the pH.

3.2. Catalytic Synthesis

Deposition of gold nanoparticles onto the TiO2 surface was carried out by the deposition-precipitation method [11]. A total of 50 mL HAuCl4 (5 mM) was heated to 70 °C then adjusted to pH 7 by 0.1 M NaOH. A total of 27 mL of Au-containing solution was added to 242 mL distilled water and stirred for 15 min at 70 °C prior to the addition 2.5 g of the P25 support. The resulting suspension was stirred for 1 h at 70 °C and for 1 h at 25 °C. The dry catalyst (Au/P25-DP) has Au(OH)3 on the surface. Au NPs can generate on P25 by reaction under UV light or by calcination. At high temperature, Au(OH)3 converts into Au2O3 and then to Au NP (Au/P25-Cal).

3.3. Catalytic Characterization

The elemental composition of catalysts was determined by ICP-OES using a Varian 715-ES ICP emission spectrometer (now Agilent, Germany) and ICP software. X-ray diffraction spectra were recorded using a VG Thermo ESCALAB 220iXL instrument (now Thermofisher Scientific, Germany) with monochromatic Al Kα radiation (E = 1486.6 eV). Peaks were fitted by Gaussian-Lorentzian curves after Shirley background subtraction with a mean error in the binding energies of ± (0.1–0.2) eV. Electron binding energy (XPS) was referenced to the adventitious carbon with a C 1s peak at 284.8 eV. For quantitative analysis, the peak areas were determined and divided by the element-specific Scofield factor and the analyzer-dependent transmission. Images of Au NPs and chromium on the TiO2 surface were investigated by transmission electron microscopy (TEM) at 200 kV using a JEM-ARM200F instrument (JOEL, Peabody, MA, USA).
In situ EPR experiments were recorded by a Bruker EMX CW microspectrometer (Bruker, Germany) using an ER 4119HS-WI high-sensitivity optical resonator with a grid on the front side. The temperature was adjusted by a Bruker Digital Temperature Control System, ER4131VT. For light irradiation, the beam of a Lot-Oriel 300 WXe lamp (LSB530) with a UV cutoff filter (GG42). The g values were calculated from the resonance field B0 and the resonance frequency ν using the resonance condition.

3.4. Photocatalysis

Photoreactions were carried out in a vessel with a water jacket to keep the temperature of the solution at room temperature (25 °C) during the reaction. The light source was supplied by a Xe-arc lamp 300 W (LOT Oriel). The distance between the light source and the reactor was 11 cm, intensity was 411 mW/cm2 in UV light and 365 mW/cm2 in visible light. A 50 mg powder of the catalyst was reacted with 40 mL Cr(VI), 20 ppm at pH 2 (adjusted by H2SO4 4 M). The mixed suspensions were first magnetically stirred in the dark for 0.5 h to reach the adsorption–desorption equilibrium. The photoreaction was carried out in 3 h, with stirring during reaction time. After the reaction, the solution was filtered to measure the conversion. Cr2O72−—a form of Cr(VI) in an acidic environment—has two maximum absorbances at 275 nm and 350 nm, see Figure 11A. Citric acid has a maximum absorbance at 275 nm. Compared to Cr(VI), Cr(III) (from Cr(NO3)3) has a maximum absorbance at 300 nm, 410 nm, and 577 nm. However, the absorbance intensity of Cr(III) can only be seen at high concentrations; >200 ppm, see Figure 11B. Low concentrations of Cr(III) (20 ppm or less) can almost not be detected. Therefore, the signal absorbance of the solution at 350 nm was used to measure the Cr(VI) conversion.

4. Conclusions

Within this study, the photocatalytic reduction ability towards Cr(VI) of Au/TiO2 catalysts was compared to standard P25 catalysts. Depending on the method of synthesis, less reactivity (compare to P25) was obtained under UV-Vis excitation. The efficiency of Cr(VI) reduction increases strongly in the presence of citric acid under UV light. The Cr(VI) reduction by bare P25 is much better than Au/P25. Photocatalyst Au/P25-Cal (with Au NP) can reduce Cr(VI) under visible light, but only achieves low conversion amounts. Interestingly, under irradiation even without a catalyst, a Cr(VI)-CA complex is formed which can act as a reductant itself.
These results can be explained by different parallel mechanisms occurring during catalysis. Partial dissolution of the gold hydroxide by citric acid (not reported before) as well as the reduction and desorption of Cr on the surface and/or Au nanoparticles need to be considered. Thus, under the reaction conditions (UV-Vis and Vis), complete new catalysts are formed containing Cr (and depleted of Au) on the surface. These mechanisms and processes resulting in the change of the catalyst system should be taken into account critically for recycling experiments as well as for the recovery of the original photocatalysts, but also for the use of other metallic co-catalysts such as Cu, Ag, and Pt.

Author Contributions

A.B.N. and D.H. designed the research; A.B.N. performed the research and analyzed the data; A.B.N. and D.H. wrote the paper; H.L.N. revised the paper. All authors read and approved the final manuscript.

Funding

This work has been supported by the RoHan Project funded by the German Academic Exchange Service (DAAD, No. 57315854) and the Federal Ministry for Economic Cooperation and Development (BMZ) inside the framework “SDG Bilateral Graduate School Program”.

Acknowledgments

D.H. thanks Carsten Robert Kreyenschulte for TEM measurements, Henrik Lund for XRD measurements, Anja Simmula for ICP-OES analysis, Angelika Brückner for the discussion during the project, and Esteban Mejia for correcting the manuscript.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jacobs, J.A.; Testa, S.M. Overview of Chromium(VI) in the Environment: Background and History. In Chromium(VI) Handbook; Guertin, J., Jacobs, J.A., Avakian, C.P., Eds.; CRC Press: Boca Raton, FL, USA, 2004; Chapter 1. [Google Scholar]
  2. Guertin, J. Toxicity and Health Effects of Chromium (All Oxidation States). In Chromium(VI) Handbook; Guertin, J., Jacobs, J.A., Avakian, C.P., Eds.; CRC Press: Boca Raton, FL, USA, 2004; Chapter 6. [Google Scholar]
  3. Hawley, E.L.; Deeb, R.A.; Kavanaugh, M.C.; Jacobs, J.A. Treatment Technologies for Chromium(VI). In Chromium(VI) Handbook; Guertin, J., Jacobs, J.A., Avakian, C.P., Eds.; CRC Press: Boca Raton, FL, USA, 2004; Chapter 8. [Google Scholar]
  4. Yuan, J.; Chen, M.; Shi, J.; Shangguan, W. Preparations and photocatalytic hydrogen evolution of N-doped TiO2 from urea and titanium tetrachloride. Int. J. Hydrog. Energy 2006, 31, 1326–1331. [Google Scholar] [CrossRef]
  5. Kato, H.; Kudo, A. Visible-Light-Response and Photocatalytic Activities of TiO2 and SrTiO3 Photocatalysts Codoped with Antimony and Chromium. J. Phys. Chem. B 2002, 106, 5029–5034. [Google Scholar] [CrossRef]
  6. Daghrir, R.; Drogui, P.; Robert, D. Modified TiO2 For Environmental Photocatalytic Applications: A Review. Ind. Eng. Chem. Res. 2013, 52, 3581–3599. [Google Scholar] [CrossRef]
  7. Tian, Y.; Tatsuma, T. Mechanisms and Applications of Plasmon-Induced Charge Separation at TiO2 Films Loaded with Gold Nanoparticles. J. Am. Chem. Soc. 2005, 127, 7632–7637. [Google Scholar] [CrossRef] [PubMed]
  8. Litter, M.I. Last advances on TiO2-photocatalytic removal of chromium, uranium and arsenic. Curr. Opin. Green Sustain. Chem. 2017, 6, 150–158. [Google Scholar] [CrossRef]
  9. Meichtry, J.M.; Brusa, M.; Mailhot, G.; Grela, M.A.; Litter, M.I. Heterogeneous photocatalysis of Cr(VI) in the presence of citric acid over TiO2 particles: Relevance of Cr(V)–citrate complexes. Appl. Catal. B 2007, 71, 101–107. [Google Scholar] [CrossRef]
  10. Meichtry, J.M.; Quici, N.; Mailhot, G.; Litter, M.I. Heterogeneous photocatalytic degradation of citric acid over TiO2: II. Mechanism of citric acid degradation. Appl. Catal. B 2011, 102, 555–562. [Google Scholar] [CrossRef]
  11. Priebe, J.B.; Radnik, J.; Lennox, A.J.J.; Pohl, M.-M.; Karnahl, M.; Hollmann, D.; Grabow, K.; Bentrup, U.; Junge, H.; Beller, M.; et al. Solar Hydrogen Production by Plasmonic Au–TiO2 Catalysts: Impact of Synthesis Protocol and TiO2 Phase on Charge Transfer Efficiency and H2 Evolution Rates. ACS Catal. 2015, 2137–2148. [Google Scholar] [CrossRef]
  12. Dozzi, M.V.; Saccomanni, A.; Selli, E. Cr(VI) photocatalytic reduction: Effects of simultaneous organics oxidation and of gold nanoparticles photodeposition on TiO2. J. Hazard. Mater. 2012, 211–212, 188–195. [Google Scholar] [CrossRef] [PubMed]
  13. Liu, X.; Lv, T.; Liu, Y.; Pan, L.; Sun, Z. TiO2–Au composite for efficient UV photocatalytic reduction of Cr(VI). Desalin. Water Treat. 2013, 51, 3889–3895. [Google Scholar] [CrossRef]
  14. Tanaka, A.; Nakanishi, K.; Hamada, R.; Hashimoto, K.; Kominami, H. Simultaneous and Stoichiometric Water Oxidation and Cr(VI) Reduction in Aqueous Suspensions of Functionalized Plasmonic Photocatalyst Au/TiO2–Pt under Irradiation of Green Light. ACS Catal. 2013, 3, 1886–1891. [Google Scholar] [CrossRef]
  15. Luo, S.; Xiao, Y.; Yang, L.; Liu, C.; Su, F.; Li, Y.; Cai, Q.; Zeng, G. Simultaneous detoxification of hexavalent chromium and acid orange 7 by a novel Au/TiO2 heterojunction composite nanotube arrays. Sep. Purif. Technol. 2011, 79, 85–91. [Google Scholar] [CrossRef]
  16. Satoh, N.; Hasegawa, H.; Tsujii, K.; Kimura, K. Photoinduced Coagulation of Au Nanocolloids. J. Phys. Chem. 1994, 98, 2143–2147. [Google Scholar] [CrossRef]
  17. Maeda, K.; Teramura, K.; Lu, D.; Saito, N.; Inoue, Y.; Domen, K. Roles of Rh/Cr2O3 (Core/Shell) Nanoparticles Photodeposited on Visible-Light-Responsive (Ga1-xZnx)(N1-xOx) Solid Solutions in Photocatalytic Overall Water Splitting. J. Phys. Chem. C 2007, 111, 7554–7560. [Google Scholar] [CrossRef]
  18. Soldat, J.; Marschall, R.; Wark, M. Improved overall water splitting with barium tantalate mixed oxide composites. Chem. Sci. 2014, 5, 3746–3752. [Google Scholar] [CrossRef] [Green Version]
  19. Busser, G.W.; Mei, B.; Muhler, M. Optimizing the Deposition of Hydrogen Evolution Sites on Suspended Semiconductor Particles using On-Line Photocatalytic Reforming of Aqueous Methanol Solutions. ChemSusChem 2012, 5, 2200–2206. [Google Scholar] [CrossRef] [PubMed]
  20. Yuan, X.; Feng, Z.; Zhao, J.; Niu, J.; Liu, J.; Peng, D.; Cheng, X. Significantly Enhanced Aqueous Cr(VI) Removal Performance of Bi/ZnO Nanocomposites via Synergistic Effect of Adsorption and SPR-Promoted Visible Light Photoreduction. Catalysts 2018, 8, 426. [Google Scholar] [CrossRef]
Figure 1. Photocatalytic mechanisms of Cr(VI) reduction by TiO2 and Au/TiO2 under UV-Vis and visible light irradiation.
Figure 1. Photocatalytic mechanisms of Cr(VI) reduction by TiO2 and Au/TiO2 under UV-Vis and visible light irradiation.
Catalysts 08 00606 g001
Figure 2. Preparation of the investigated catalyst.
Figure 2. Preparation of the investigated catalyst.
Catalysts 08 00606 g002
Figure 3. The UV-Vis diffuse reflectance measurement of photocatalysts; (a) P25, (b) Au/P25-DP, (c) Au/P25-DP after reaction under UV-Vis light, (d) Au/P25-Cal.
Figure 3. The UV-Vis diffuse reflectance measurement of photocatalysts; (a) P25, (b) Au/P25-DP, (c) Au/P25-DP after reaction under UV-Vis light, (d) Au/P25-Cal.
Catalysts 08 00606 g003
Figure 4. XPS results of Au/P25-DP after reaction under UV-Vis irradiation.
Figure 4. XPS results of Au/P25-DP after reaction under UV-Vis irradiation.
Catalysts 08 00606 g004
Figure 5. The Powder X-Ray diffraction pattern of photocatalysts, the diffraction pattern of the Au nanoparticle (NP) is marked; (a) P25, (b) P25 after reaction, (c) Au/P25-DP, (d) Au/P25-DP after reaction, (e) Au/P25-Cal.
Figure 5. The Powder X-Ray diffraction pattern of photocatalysts, the diffraction pattern of the Au nanoparticle (NP) is marked; (a) P25, (b) P25 after reaction, (c) Au/P25-DP, (d) Au/P25-DP after reaction, (e) Au/P25-Cal.
Catalysts 08 00606 g005
Figure 6. Transmission electron microscopy (TEM) for catalysts Au/P25-DP after reaction under UV-Vis irradiation.
Figure 6. Transmission electron microscopy (TEM) for catalysts Au/P25-DP after reaction under UV-Vis irradiation.
Catalysts 08 00606 g006
Figure 7. Cr(V)-CA complex detected by EPR spectroscopy (A), as well as Conversion of Cr(V)-CA (B) under UV-Vis and Vis irradiation.
Figure 7. Cr(V)-CA complex detected by EPR spectroscopy (A), as well as Conversion of Cr(V)-CA (B) under UV-Vis and Vis irradiation.
Catalysts 08 00606 g007
Figure 8. Proposed processes occurring on a bare P25 sample.
Figure 8. Proposed processes occurring on a bare P25 sample.
Catalysts 08 00606 g008
Figure 9. Proposed processes occurring on the Au/P25-DP sample.
Figure 9. Proposed processes occurring on the Au/P25-DP sample.
Catalysts 08 00606 g009
Figure 10. Proposed processes occurring on the Au/P25-DP sample.
Figure 10. Proposed processes occurring on the Au/P25-DP sample.
Catalysts 08 00606 g010
Figure 11. UV-Vis absorbance spectrum of Cr(VI) (A) and Cr(III) (B) at pH 2.
Figure 11. UV-Vis absorbance spectrum of Cr(VI) (A) and Cr(III) (B) at pH 2.
Catalysts 08 00606 g011aCatalysts 08 00606 g011b
Table 1. Conversion of Cr(VI) with different conditions. 1
Table 1. Conversion of Cr(VI) with different conditions. 1
CatalystNo LightNo CA UV-VisNo CA Vis With CA UV-VisWith CA Vis
P250740100 (70 2)92
Au/P25-DP044010075
Au/P25-Cal0481810088
no catalysts0007432
1 Conversion after a 3-h reaction time in %; molar ratio CA:Cr(VI) 25:1. CA—Citric acid, Cal—calcined, DP—Deposition-Precipitation, see Figure 2. 2 Conversion after recycling.
Table 2. Gold and chromium content on catalysts before and after reaction.
Table 2. Gold and chromium content on catalysts before and after reaction.
CatalystBefore Reaction (wt%)No CA UV-Vis (wt%)No CA Vis (wt%)With CA UV-Vis (wt%)With CA Vis (wt%)
P25Cr not determined (n.d.)Cr 0.98Cr n.d.Cr 0.72Cr 0.64
Au/P25-DPAu 0.62
Cr n.d.
Au 0.16
Cr 0.23
Au 0.07
Cr 0.11
Au 0.18
Cr 0.70
Au 0.32
Cr 0.45
Au/P25-CalAu 0.62
Cr n.d.
Au 0.51
Cr 0.61
Au 0.29
Cr 0.31
Au 0.35
Cr 0.25
Au 0.41
Cr 0.61

Share and Cite

MDPI and ACS Style

Ngo, A.B.; Nguyen, H.L.; Hollmann, D. Criticial Assessment of the Photocatalytic Reduction of Cr(VI) over Au/TiO2. Catalysts 2018, 8, 606. https://doi.org/10.3390/catal8120606

AMA Style

Ngo AB, Nguyen HL, Hollmann D. Criticial Assessment of the Photocatalytic Reduction of Cr(VI) over Au/TiO2. Catalysts. 2018; 8(12):606. https://doi.org/10.3390/catal8120606

Chicago/Turabian Style

Ngo, Anh Binh, Hong Lien Nguyen, and Dirk Hollmann. 2018. "Criticial Assessment of the Photocatalytic Reduction of Cr(VI) over Au/TiO2" Catalysts 8, no. 12: 606. https://doi.org/10.3390/catal8120606

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