Next Article in Journal
Oxidative Dehydrogenation of Methane When Using TiO2- or WO3-Doped Sm2O3 in the Presence of Active Oxygen Excited with UV-LED
Previous Article in Journal
Surface Engineering of WO3/BiVO4 to Boost Solar Water-Splitting
Previous Article in Special Issue
Recent Advances in Selective Photo-Epoxidation of Propylene: A Review
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Mechanistic Insights into Visible Light-Induced Direct Hydroxylation of Benzene to Phenol with Air and Water over Pt-Modified WO3 Photocatalyst

1
Department of Applied Chemistry, Faculty of Engineering, Osaka Institute of Technology, 5-16-1 Omiya, Asahi-ku, Osaka 535-8585, Japan
2
Graduate School of Science and Technology for Innovation, Yamaguchi University, 2-16-1 Tokiwadai, Ube 755-8611, Japan
3
Department of Chemistry and Materials Technology, Kyoto Institute of Technology, Matsugasaki, Sakyo-ku, Kyoto 606-8585, Japan
*
Author to whom correspondence should be addressed.
Catalysts 2020, 10(5), 557; https://doi.org/10.3390/catal10050557
Submission received: 2 May 2020 / Revised: 15 May 2020 / Accepted: 15 May 2020 / Published: 18 May 2020
(This article belongs to the Special Issue State-of-the-Art Photocatalytical Technology in Asia)

Abstract

:
Activation of C(sp2)-H in aromatic molecules such as benzene is one of the challenging reactions. The tungsten trioxide supported Pt nanoparticles (Pt-WO3) exhibited hydroxylation of benzene in the presence of air and H2O under visible-light (420 < λ < 540 nm) irradiation. The photocatalytic activities (yields and selectivity of phenol) were studied under several experimental conditions. Furthermore, investigations of mechanistic insight into hydroxylation of benzene have been carried out by analyses with apparent quantum yields (AQY), an H218O isotope-labeling experiment, kinetic isotope effects (KIE), electrochemical measurements and density functional theory (DFT) calculations. It was proposed that dissociation of the O–H bond in H2O is the rate-determining step. Furthermore, the substitution of the OH derived from H2O with H abstracted from benzene by photo-formed H2O2 indicated a mechanism involving a push-pull process for the hydroxylation of benzene into phenol.

Graphical Abstract

1. Introduction

Visible light-driven photocatalysts for environmental conservation, securing energy resources and selective organic synthesis have received considerable attention because they can utilize unlimited solar energy [1,2,3,4,5,6,7]. Tungsten trioxide (WO3) is one of the promising visible-light responsible photocatalyst having a direct band-gap excitation at ca. 2.7 eV. Recently, WO3 has been proven to be an effective strategy for improving the photocatalytic degradation of volatile organic compounds (VOCs) such as acetaldehyde [8,9,10,11], toluene [12], acid [13]; water splitting to form O2 in the presence of sacrificial agent [1]; as well as selective organic conversion such as oxidation of alcohol [14].
Phenols are important precursors for many chemicals and industrial products such as dyes and polymers, and they are currently produced from benzene by a three step cumene process. The cumene process for phenol formation exhibits low activity (~5% yield of phenol) and emission of large amounts of waste although the reactions require high temperature, high pressure and strong acidic conditions. Therefore, it is a great challenge to develop a one-step synthesis of phenol from benzene by using homogeneous and heterogeneous inorganic catalysts such as palladium membrane, titanium-containing mesoporous molecular sieves and vanadium-substituted phosphomolybdate by hydrogen peroxide (H2O2) [15,16,17]. Although H2O2 is often used as environmentally friendly oxidant only to produce H2O in catalytic hydroxylation of benzene, O2 is more ideal oxidant than H2O2 due to its abundance in nature.
In a recent study, one-step direct hydroxylation of benzene toward phenol has been extensively studied using homogeneous photocatalysts such as quinolinium ions and [RuII(Me2phen)3]2+ ions, [18,19,20,21], and heterogeneous semiconductor photocatalyst such as TiO2 [22,23,24,25,26,27] and WO3 [28,29] have been employed for selective photocatalytic hydroxylation of benzene to phenol under photo-irradiation ultraviolet (UV) light and/or visible-light. Yoshida et al. previously reported that platinum (Pt)-loaded TiO2 (Pt-TiO2) photocatalyst exhibited direct hydroxylation of benzene to form phenol and H2 in the absence of O2 under UV-light irradiation [22]. Also, Tomita et al. reported that the Pt-deposited WO3 (Pt-WO3) photocatalyst exhibited selective hydroxylation of benzene to phenol in the presence of O2 and H2O under light irradiation of both UV light and visible light (300 < λ < 500 nm) [28,29]. It was also confirmed that OH derived from H2O is included in phenol by employing an H218O labeling experiment [29]. However, in a previous study, photocatalytic activities and reaction mechanisms for hydroxylation of benzene on the Pt-WO3 have not been demonstrated under only irradiation of visible-light (λ > 420 nm).
In this study, we focus on understanding the photocatalytic activities for hydroxylation of benzene on the Pt-WO3 photocatalyst under irradiation of only visible light (420 < λ < 540 nm). Furthermore, investigation of mechanistic insight has been carried out by combination with apparent quantum yields (AQY), H218O isotope labeling experiment, kinetic isotope effects (C6D6, D2O), electrochemical measurements and density functional theory (DFT) calculations.

2. Results

2.1. Preparation of Pt-Deposited Tungsten Trioxide (WO3) and Its Characterization

The Pt-WO3 photocatalyst was characterized by X-ray diffraction (XRD), ultraviolet–visible (UV–Vis) spectroscopy, scanning transmittance electron microscope with energy-dispersed X-ray emission spectroscopy (STEM-EDS) and X-ray photoelectron spectroscopy (XPS) measurements. The XRD patterns of WO3 and Pt-WO3 were identified with a monoclinic structure in accordance with standard XRD profile of JCPDS cards No. 43-1035, and no other phase of Pt species was observed (see Figure 1 [I]). The WO3 exhibited absorption spectrum in the visible-light region. Bandgap of the WO3 was estimated to be ca. 2.6 eV from tauc plots υof (F(R) × )2 vs. (See Figure S1). Furthermore, the Pt-WO3 with different amounts of Pt species exhibited optical absorbance above 450 nm, which is attributable to the scattering effects from the Pt particles [30] or surface resonance [31]. The optical absorbance significantly increased as an increase of the Pt species deposited on the WO3 (See Figure 1 [II]). The scanning transmittance electron microscope with the STEM-EDS image of Pt(0.4)-WO3 confirmed that Pt nano-particles were dispersed on the WO3 surface (See Figure 1 [III]). The XPS analysis indicated that the Pt(0.4)-WO3 photocatalyst was observed to possess two different types of doublet peaks (4f5/2 and 4f7/2) at 74.8 and 71.6 eV for Pt0 as a major species, at 75.8 and 72.5 eV for Pt2+ as a minor species (See Figure 1 [IV]). These results indicated that Pt nano-particle deposition on the WO3 was successfully performed by photo-electrochemical deposition methods.

2.2. Photocatalytic Hydroxylation of Benzene on the Pt-WO3 under Visible-Light Irradiation

The photocatalytic hydroxylation of benzene in the presence of air and H2O was performed on the WO3 photocatalyst under visible-light irradiation. It was confirmed that the hydroxylation reaction does not take place under photo-irradiation without a photocatalyst nor with a photocatalyst without irradiation, i.e., both photocatalyst and irradiation are required in combination for hydroxylation reaction to occur. The WO3 exhibited very low activity for phenol formation under visible-light irradiation (See Table S1). After the photocatalytic reaction, the color of suspension was slightly bluish, suggesting that the WO3 was partially reduced and the ability for oxygen reductive reaction (ORR) was retarded. On the other hand, Pt-WO3 exhibited significant improvement of photocatalytic activities for the formation of hydroxylated products such as phenol and catechol. The yields of phenols significantly improved as an increase of loading amounts of Pt species up to 0.2 atom%, and the activity was saturated at 0.4 atom%. The selectivity of phenol was optimized at 0.1 atom% and then slightly decreased as an increase of Pt-deposition (see Figure 2 [I], Table S1). Reaction time profile for phenol evolution on the Pt(0.2)-WO3 photocatalyst was shown in Figure 2 [II]. The photocatalytic activity increased as an increase of reaction time: producing 29 μmol of phenol with 47 % selectivity after photocatalytic reaction for 20 h, and 40 μmol phenol with 64% selectivity for 70 h. (See distribution of side products shown in Table S2). After the reaction for 70 h, the amount of photo-formed H2O2 was only 0.15 μmol, which are much less than hydroxylated products assuming two electron reduction of O2 to form H2O2. It can be considered that the H2O2 may be self-decomposed on the photocatalyst and/or participate in the reaction for the hydroxylation of benzene. The role of H2O2 will be discussed later. Moreover, when the volume of solvent in the reaction became small, the yields of phenol decreased, while the selectivity of phenol improved over 70% (See Table S3). Therefore, a trade-off relation between the yields and selectivity of phenol was observed. The surface coverage of the photocatalyst by condensation of the adsorbent could be considered to prevent over oxidation of benzene leading to high selectivity of phenol, of which the similar phenomena was previously reported on the TiO2 photocatalyst [25].
Apparent quantum yields (AQY) on the Pt(0.2)-WO3 were measured. Assuming one photon producing one phenol, the AQY for phenol formation reached over 2% by excitation at 400 nm as shown in Figure 3. The AQY profile was found to be very similar with the absorption spectrum of WO3, suggesting that phenol formation is strongly correlated with the light absorbance of WO3, not with Pt scattering [30] and/or resonance absorbance [31].

2.3. Role of H2O on the Photocatalytic Reactions

In order to understand role of H2O, the H218O-labeling experiments for photocatalytic hydroxylation of benzene were performed by liquid chromatography–mass spectrometry (LC-MS), and results are shown in Table 1 and Figure S2. It is noted that the atomic exchanges between O2 and H2O, and between O2 and phenol are very slow even under photo-irradiation [23,32]. Therefore, H218O was used for tracing O species incorporated in phenol. When the photocatalytic reactions were carried out in the presence of H216O (100%), the photo-formed phenol was only observed at the mass number of 93 (m/z), which is attributed to the phenolic anion (C6H516O). Furthermore, when the photocatalytic reactions by employing H216O (90%)/H218O (10%) were carried out, the peaks at 93 (m/z) for the C6H516O as well as 95 (m/z) for C6H518O were observed. The peak intensities indicated that the ratio of 18O to 16O involved in the photo-formed phenol was 9.9% under visible-light (420 < λ < 540 nm), while 9.1% under UV-light irradiation (300 < λ < 400 nm). Assuming that H216O and H218O exhibited same activity for hydroxylation of benzene, it was concluded that hydroxyl groups in phenol is almost derived from H2O under both irradiation of visible-light and UV-light.
The rate determining step for hydroxylation of benzene was investigated by the kinetic isotope effect (KIE) using D2O and C6D6 (See Figure 4). The reaction rate constants for formation of phenol from normal benzene (C6H6) in the presence of H2O and D2O were roughly estimated to be 3.75 × 10−6 and 1.42 × 10−6 s−1 for (a) and (b), respectively. The kinetic isotope effect (kH2O/kD2O) was estimated to be ca. 2.7 (see Figure 4 [I]). On the other hand, the reaction rate constants for formation of phenol from C6H6 and C6D6 in the presence of H2O were roughly estimated to be 3.75 × 10−6 and 3.77 × 10−6 s−1 for (c) and (d), respectively. That is, the KIE of kC6H6/kC6D6 was estimated to be 1.0 (see Figure 4 [II]). These results suggest that the dissociation of O–H bond in the H2O plays an important role in the rate determining step for the hydroxylation of benzene.

2.4. Role of O2 on the Photocatalytic Reactions

When the half reaction for the photocatalytic hydroxylation of benzene was performed on the Pt-WO3 in the presence of Ag+ ions as electron scavenger instead of O2 under visible-light irradiation, no products could be detected under visible-light irradiation (Table 2). This result indicated that both H2O and O2 (air) in combination are essential for hydroxylation of benzene under visible-light irradiation.
The flat band potential of WO3 was obtained from capacitance versus voltage measurements. The WO3 powder was deposited on fluorine-doped tin oxide (FTO) coated glass. Mott–Schottky plots indicated that the flatband potential (≈conduction band) of the WO3/FTO was estimated to be +0.40 V vs. reversible hydrogen electrode (RHE) (Figure S3). LSV measurements of WO3 and Pt-WO3 were carried out in O2-saturated 0.1 M Na2SO4 aq. The cathodic current was obtained at the potential more cathodic than the conduction band (+0.40 V vs. RHE) of WO3 (See Figure S4). This cathodic current would be attributed to the reduction of WO3. Furthermore, the cathodic current significantly increased by deposition of Pt on WO3. A deposition of Pt species as a co-catalyst caused an improvement of the ORR. It is noted that multi-electron reduction of O2 to H2O2, thermodynamically occurred at E0(O2/H2O2) = +0.68 V [33]. An increase of cathodic current is attributed to the effective electron transfer to O2 via Pt species to form H2O2 [34].
It was confirmed that the hydroxylation of benzene in the presence of H2O2 did not take place under photoirradiation without a photocatalyst nor with photocatalyst without irradiation. However, both photocatalyst and visible-light irradiation are required for the hydroxylation of benzene (See Table S4). In order to understand the role of H2O2, the hydroxylation of benzene (300 μmol) in the presence of H2O2 (150 μmol) and H216O (90%)/H218O (10%) was conducted on the Pt-WO3 under visible-light irradiation. As a result, the ratio of 18O to 16O was determined to be 8.3%, suggesting that major contribution of OH groups in phenol came from H2O even in the presence of H2O2.

2.5. Reaction Mechanisms for Photocatalytic Hydroxylation of Benzene to Phenol

In order to evaluate the oxidation power of photo-induced holes, photocatalytic half reactions on the Pt(0.2)-WO3 photocatalyst were carried out. It was observed that addition of Ag+ ions instead of O2 showed no products under visible-light irradiation, while the phenolic compounds were detected under UV-light irradiation (See Table 2). These results suggest that water oxidation by UV-light irradiation induced strong oxidation power possibly to form free OH radicals, which would attack benzene directly [29], but by visible-light irradiation would not form OH radicals. In fact, we employed the tertiary butyl alcohol (TBA) as scavengers of OH radicals. As a result, the TBA added in the photocatalytic system on the Pt-WO3 under visible-light irradiation did not influence to the yields of phenol (Table 3). However, the amounts of catechol (CA) and phloroglucinol (PL) decreased by the addition of TBA (See Table 3). The di- and tri- OH groups involved in phenolic compounds may be partially derived from O2. Research is under way using an 18O2 labeling test.
As mentioned above, H2O2 was confirmed to form as intermediate species by multi-electron reduction of O2 during photocatalytic hydroxylation of benzene. The reaction mechanisms were investigated by DFT calculations. Figure 5 [Ia] showed the optimized structures of the reactant, transitional state (TS) and product by the interactions of benzene, H2O with H2O2. The O–H bond length in H2O, O–O in H2O2 and C–H in benzene were observed to increase through the reaction path. It was found that H2O2 could assist C-H bond dissociation from benzene. Subsequently, OH species derived from H2O was incorporated within phenol in the final product. Moreover, the reaction path in the presence of the Pt co-catalyst was investigated (See Figure 5 [Ib]). Strong interaction of H2O2 with Pt3 was observed to dissociate the O–O bond to form O species, and subsequently, the C–H bond in benzene was dissociated by an assist of the O species and/or Pt3. The energy changes for overall reactions were indicated in Figure 5 [II]. The hydroxylation of benzene proceeded on much lower potential energy surface with the Pt3 co-catalyst than that without Pt3 co-catalyst. Thus, we have demonstrated the possible reaction path for hydroxylation of benzene to form phenol including OH groups derived from H2O.
The reaction mechanism for the direct hydroxylation of benzene to phenol on the Pt-WO3 was proposed (See Figure 6). This reaction is initiated by visible-light irradiation of the WO3. The photo-induced holes and electrons are generated in the VB and CB of the WO3, respectively. It was assumed that the photo-induced holes can oxidize H2O to form O2 [1], as well as activated (H2O)* intermediate species of which the O–H bond length would be increased. On the other hand, the electrons effectively transfer from CB to O2 via Pt species to form peroxide species such as H2O2, which would be further activated on the catalyst surface. The activated (H2O)* species would attack the π electrons of the aromatic benzene ring, assisting the activated (H2O2)* to dissociate the C–H bond in benzene. It was, thus, proposed that the hydroxylation of benzene occurs by a mechanism involving a push-pull process, which promotes the formation of phenol.

3. Materials and Methods

3.1. Materials

Tungsten trioxide (WO3, 99.9%) from Kojundo Chemical Laboratory (Osaka, Japan), H2PtCl6∙6H2O (99.9%) from Wako Pure Chemical Industries (Osaka, Japan); and water-18O (97 atom% 18O) and titanium (IV) oxysulfate solution (99.99%) from Sigma-Aldrich (St. Louis, MO, USA) were purchased. All chemicals were used without further purification.

3.2. Photoelectrochemical Deposition of Pt Species as Co-Catalyst on the WO3

Photoelectrochemical deposition of Pt species was conducted on the WO3 surface. WO3 powder (1.0 g) was suspended in distilled water (25 mL) involving desired amounts of H2PtCl6 (4.32 × 10−6 ~ 2.16 × 10−5 mol), and the suspension was photo-irradiated by a light-emitting diode (LED) lamp (420 < λ < 540 nm) for 1 h under stirring in order to disperse photo-adsorbed Pt species. Subsequently, methanol (5 mL) as reductant was added into the suspension, and then it was continuously photo-irradiated for 4 h. The solid products were separated by centrifuge (LC-200, TOMY, 4500 rpm), followed by washing with distilled water and acetonitrile, and drying under vacuum condition at ambient temperature over night. The photocatalyst was referred to be as Pt(x)-WO3(x: atom%). The photocatalyst was kept in the ambient temperature, and the photocatalytic reactions were carried out without further treatment of photocatalyst.

3.3. Characterizations

The X-ray diffraction (XRD) patterns were obtained with a RIGAKU RINT2000 using Cu Kα radiation (λ = 1.5417 Å) (RIGAKU, Tokyo, Japan). The oxidative states of Pt species were analyzed by an X-ray photoelectron spectroscope (XPS), KRATOS, AXIS Ultra, using Al Kα radiation (E = 1486.8 eV) (Shimadzu, Kyoto, Japan). The UV–Vis spectroscopic measurements were carried out on diffuse reflectance with a UV–Vis scanning spectrophotometer (UV-3100PC, Shimadzu, Kyoto, Japan). The elemental distribution images were taken with scanning transmittance electron microscope with energy-dispersed X-ray emission spectroscopy (STEM-EDS; JEOL JSM-2100, Tokyo, Japan).

3.4. Photocatalytic Reactions

Photocatalyst (20 mg) was suspended in distilled water (2, 5 or 10 mL) in a pyrex reaction tube (volume: 20 mL) under air capped with precision seal septum. If specific description was not given, the distilled water (10 mL) was introduced. And subsequently, 26.8 μL of benzene (300 μmol) was dropped into the suspension by a syringe under vigorous stirring. The reaction cell was photo-irradiated by a blue LED (visible light: 420 < λ < 540 nm) or black light (UV light: 300 < λ < 400 nm) at 298 K. The light energy intensities were shown in Figure S5. After the reaction, the catalysts were immediately separated from the suspension by filtration through a 0.20 μm membrane filter (Dismic-25JP, Advantec, Tokyo, Japan). The solution was, then, analyzed by High Performance Liquid Chromatography (HPLC, Shimadzu LC10ATVP, Kyoto, Japan, UV–Vis detector, column: Chemcopak, mobile phase: a mixture of acetonitrile and 1.0% formic acid aqueous solution), and several products were identified (see Figure S6). The gas phase (CO2) was analyzed by Gas Chromatography–thermal conductivity detector (GC-TCD, Shimadzu GC-8A, Kyoto, Japan; column: porapak Q). Titanium sulfate (100 μL) was added to 1 mL of the reaction solution, and amounts of H2O2 formed in solution were analyzed by colorimetry from the value of absorbance at 410 nm in the UV–Vis absorption spectrum.
For the tracer experiment, H218O was used as the oxygen isotope source. The photocatalytic hydroxylation of benzene or phenol was carried out over Pt-WO3 photocatalyst in the presence of H218O (10%)/H216O (90%) as solvent and air under visible-light irradiation for 20 h. The photocatalytic hydroxylation of benzene was performed in the presence of H216O2 (100 μmol) and H218O (10%)/H216O (90%) under visible-light irradiation for 2 h. Phenol in aqueous solutions were extracted by toluene, and the molar ratios of Ph-18OH and Ph-16OH were identified by LC-MS (Shimadzu LCMS-2020 spectrometer, Kyoto, Japan).

3.5. Electrochemical Measurements

Flatband potentials of the photoelectrodes were measured by a Potentio/Galvanostat (PGSTAT204, Autolab). The electrolysis cell was constructed with three electrodes, the WO3 photoelectrode as working electrode, the platinum wire as auxiliary, and Ag/AgCl as the reference electrode. The WO3 photoelectrodes were prepared as follows: the WO3 (20 mg) was suspended in ethanol (0.2 mL) by a super-sonification to form a paste. The paste was then spread onto the FTO (active space 1.0 cm2, Aldrich) by spin coating (2200 rpm, 30 s) for 3 times, followed by heat-treatment at 773 K for 1 h in air. Prior to measurements, 0.10 M Na2SO4 aqueous solution was vigorously bubbled by N2 gas for 20 min in order to remove O2.
Linear sweep voltammetry (LSV) was conducted at the rate of 10 mV s−1. The photocatalyst (20 mg) was added into ethanol (0.2 mL), followed by a super-sonification for 5 min to disperse. The paste was casted by spin-coating (2200 rpm, 30 s) on the conductive transparent glass (FTO, 10 Ω/□, Sigma Aldrich, St. Louis, MO, USA) 3 times, followed by heat-treatment at 373 K overnight. Prior to LSV measurements, 0.10 M Na2SO4 aq. was vigorously bubbled by O2 for 20 min.

3.6. Density Functional Theory (DFT) Calculation

Gaussian 09 program [35] and the hybrid B3LYP functional [36] were used. For Pt atoms, the Los Alamos effective core potentials [37] were employed along with the corresponding valence double basis sets [38]. For other atoms, a 6-311G(d,p) basis set was used. After the transition state (TS) was characterized, the intrinsic reaction coordinate (IRC) analysis [39] was carried out for both directions, reactant and product sides. The IRC analyses were followed by normal optimization runs, and the reactant and product were optimized as local minima.

4. Conclusions

In this study, we provided an understanding of photocatalytic activities for the hydroxylation of benzene under irradiation of only visible-light (420 < λ < 540 nm). It was demonstrated that hydroxylation of benzene on the Pt-WO3 photocatalyst in the presence of air and H2O produced several products such as phenol, catechol and phloroglucinol etc., and the selectivity of phenol improved over 70%.
One of novelties in our findings was to confirm the reaction mechanisms for photocatalytic hydroxylation of benzene into phenol by experimental and theoretical studies. An investigation of the mechanistic insights has been carried out by the combination with apparent quantum yields (AQY), an H218O isotope-labeling experiment, kinetic isotope effects (C6D6, D2O), electrochemical measurements, and DFT calculations. It was proposed that the substitution of the OH derived from H2O with H abstracted from benzene by photo-formed H2O2 indicated a mechanism involving a push-pull process for the hydroxylation of benzene into phenol. Our results showed new perspectives for the enhancement of yields and selectivity of phenol, as well as deeper understanding of the reaction mechanisms for the hydroxylation of benzene into phenol on the Pt-WO3 photocatalyst.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4344/10/5/557/s1: Figure S1: Tauc plots of WO3, Figure S2: H218O isotope labeling experiments, Figure S3: Mott–Schottky plots of WO3, Figure S4: LSV measurements of WO3 and Pt-WO3, Figure S5: Photo-intensities emitted from light-emitting diode (LED) and UV lamp, Figure S6: Molecular structures of products, Table S1: Photocatalytic activities depending on the amounts of Pt-deposition, Table S2: Time profiles in the photocatalytic reaction, Table S3: Effects of volumes in the reaction on the photocatalytic activity, Table S4: Role of H2O2 for the photocatalytic reactions.

Author Contributions

Conceptualization, S.H.; investigation, Y.K., M.T.; writing—original draft preparation, Y.K.; writing—review and editing, S.H., Y.S., H.K., M.M., Y.M.; supervision, S.H.; project administration, S.H. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Y.; Suzuki, H.; Xie, J.; Tomita, O.; Martin, D.J.; Higashi, M.; Kong, D.R.; Abe, R.; Tang, J. Mimicking natural photosynthesis: Solar to renewable H2 fuel synthesis by Z-scheme water splitting systems. Chem. Rev. 2018, 118, 5201–5241. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Schneider, J.; Matsuoka, M.; Takeuchi, M.; Zhang, J.; Horiuchi, Y.; Anpo, M.; Bahnemann, D.W. Understanding TiO2 photocatalysis: Mechanisms and materials. Chem. Rev. 2014, 114, 9919–9986. [Google Scholar] [CrossRef] [PubMed]
  3. Ma, Y.; Wang, X.; Jia, Y.; Chen, X.; Han, H.; Li, C. Titanium dioxide-based nanomaterials for photocatalytic fuel generations. Chem. Rev. 2014, 114, 9987–10043. [Google Scholar] [CrossRef] [PubMed]
  4. Kou, J.; Lu, C.; Wang, J.; Chen, Y.; Xu, Z.; Varma, R.S. Selectivity enhancement in heterogeneous photocatalytic transformations. Chem. Rev. 2017, 117, 1445–1514. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Yamashita, H.; Mori, K.; Kuwahara, Y.; Kamegawa, T.; Wen, M.; Verma, P.; Che, M. Single-site and nano-confined photocatalysts designed in porous materials for environmental uses and solar fuels. Chem. Soc. Rev. 2018, 47, 8072–8096. [Google Scholar] [CrossRef]
  6. Higashimoto, S. Titanium dioxide-based visible-light sensitive photocatalysis: Mechanistic insight and applications. Catalysts 2019, 9, 201. [Google Scholar] [CrossRef] [Green Version]
  7. Asahi, R.; Morikawa, T.; Irie, H.; Ohwaki, T. Nitrogen-doped titanium dioxide as visible light-sensitive photocatalyst: Designs, developments, and prospects. Chem. Rev. 2014, 114, 9824–9852. [Google Scholar] [CrossRef]
  8. Abe, R.; Takami, H.; Murakami, N.; Ohtani, B. Pristine simple oxides as visible light driven photocatalysts: Highly efficient decomposition of organic compounds over platinum-loaded tungsten oxide. J. Am. Chem. Soc. 2008, 130, 7780–7781. [Google Scholar] [CrossRef]
  9. Arai, T.; Horiguchi, M.; Yanagida, M.; Gunji, T.; Sugihara, H.; Sayama, K. Reaction mechanism and activity of WO3-catalyzed photodegradation of organic substances promoted by a CuO cocatalyst. J. Phys. Chem. C 2009, 113, 6602–6609. [Google Scholar] [CrossRef]
  10. Arai, T.; Horiguchi, M.; Yanagida, M.; Gunji, T.; Sugihara, H.; Sayama, K. Complete oxidation of acetaldehyde and toluene over a Pd/WO3 photocatalyst under fluorescent- or visible-light irradiation. Chem. Commun. 2008, 43, 5565–5567. [Google Scholar] [CrossRef]
  11. Tomita, O.; Sugimoto, T.; Sako, K.; Hayakawa, S.; Katagiri, K.; Inumaru, K. Enhanced photocatalytic activity of Pt/WO3 photocatalyst combined with TiO2 nanoparticles by polyelectrolyte-mediated electrostatic adsorption. Catal. Sci. Technol. 2015, 5, 1163–1168. [Google Scholar]
  12. Higashimoto, S.; Katsuura, K.; Yamamoto, M.; Tahakashi, M. Photocatalytic activity for decomposition of volatile organic compound on Pt-WO3 enhanced by simple physical mixing with TiO2. Catal. Commun. 2020, 133, 105831. [Google Scholar] [CrossRef]
  13. Higashimoto, S.; Ushiroda, Y.; Azuma, M. Electrochemically assisted photocatalysis of hybrid WO3/TiO2 films: Effect of the WO3 structures on charge separation behavior. Top. Catal. 2008, 47, 148–154. [Google Scholar] [CrossRef]
  14. Tomita, O.; Otsubo, T.; Higashi, M.; Ohtani, B.; Abe, R. Partial oxidation of alcohols on visible-light-responsive WO3 photocatalysts loaded with palladium oxide cocatalyst. ACS Catal. 2016, 6, 1134–1144. [Google Scholar] [CrossRef]
  15. Niwa, S.; Eswaramoorthy, M.; Nair, J.; Raj, A.; Ito, N.; Shoji, H.; Nanba, T.; Mizukami, F. A one-step conversion of benzene to phenol with a palladium membrane. Science 2002, 295, 105–107. [Google Scholar] [CrossRef]
  16. Tanev, P.T.; Chibwe, M.; Pinnavaia, T.J. Titanium-containing mesoporous molecular sieves for catalytic oxidation of aromatic compounds. Nature 1994, 368, 321–323. [Google Scholar] [CrossRef]
  17. Chen, J.; Gao, S.; Xu, J. Direct hydroxylation of benzene to phenol over a new vanadium-substituted phosphomolybdate as a solid catalyst. Catal. Commun. 2008, 9, 728–733. [Google Scholar] [CrossRef]
  18. Ohkubo, K.; Kobayashi, T.; Fukuzumi, S. Direct oxygenation of benzene to phenol using quinolinium ions as homogeneous photocatalysts. Angew. Chem. Int. Ed. 2011, 50, 8652–8655. [Google Scholar] [CrossRef]
  19. Fukuzumi, S.; Ohkubo, K. One step selective hydroxylation of benzene to phenol. Asian J. Org. Chem. 2015, 4, 836–845. [Google Scholar] [CrossRef]
  20. Han, J.W.; Jung, J.; Lee, Y.M.; Nam, W.; Fukuzumi, S. Photocatalytic oxidation of benzene to phenol using dioxygen as an oxygen source and water as an electron source in the presence of a cobalt catalyst. Chem. Sci. 2017, 8, 7119–7125. [Google Scholar] [CrossRef] [Green Version]
  21. Fukuzumi, S.; Lee, Y.M.; Nam, W. Photocatalytic oxygenation reactions using water and dioxygen. ChemSusChem 2019, 12, 3931–3940. [Google Scholar] [CrossRef] [PubMed]
  22. Yoshida, H.; Yuzawa, H.; Aoki, M.; Otake, K.; Itoh, H.; Hattori, T. Photocatalytic hydroxylation of aromatic ring by using water as an oxidant. Chem. Commun. 2008, 38, 4634–4636. [Google Scholar] [CrossRef] [PubMed]
  23. Bui, T.D.; Kimura, A.; Ikeda, S.; Matsumura, M. Determination of oxygen sources for oxidation of benzene on TiO2 photocatalysts in aqueous solutions containing molecular oxygen. J. Am. Chem. Soc. 2010, 132, 8453–8458. [Google Scholar] [CrossRef] [PubMed]
  24. Ide, Y.; Matsuoka, M.; Ogawa, M. Efficient visible-light-induced photocatalytic activity on gold-nanoparticle-supported layered titanate. J. Am. Chem. Soc. 2010, 47, 16762–16764. [Google Scholar] [CrossRef]
  25. Ide, Y.; Torii, M.; Sano, T. Layered silicate as an excellent partner of a TiO2 photocatalyst for efficient and selective green fine-chemical synthesis. J. Am. Chem. Soc. 2013, 135, 11784–11786. [Google Scholar] [CrossRef]
  26. Zheng, Z.; Huang, B.; Qin, X.; Zhang, X.; Dai, Y.; Whangbo, M.-H. Facile in situ synthesis of visible-light plasmonic photocatalysts M@TiO2 (M = Au, Pt, Ag) and evaluation of their photocatalytic oxidation of benzene to phenol. J. Mater. Chem. 2011, 21, 9079–9087. [Google Scholar] [CrossRef]
  27. Goto, T.; Ogawa, M. Efficient photocatalytic oxidation of benzene to phenol by metal complex-clay/TiO2 hybrid photocatalyst. RSC Adv. 2016, 6, 23794–23797. [Google Scholar] [CrossRef]
  28. Tomita, O.; Abe, R.; Ohtani, B. Direct synthesis of phenol from benzene over platinum-loaded tungsten(VI) oxide photocatalysts with water and molecular oxygen. Chem. Lett. 2011, 40, 1405–1407. [Google Scholar] [CrossRef] [Green Version]
  29. Tomita, O.; Ohtani, B.; Abe, R. Highly selective phenol production from benzene on a platinum-loaded tungsten oxide photocatalyst with water and molecular oxygen: Selective oxidation of water by holes for generating hydroxyl radical as the predominant source of the hydroxyl group. Catal. Sci. Technol. 2014, 4, 3850–3860. [Google Scholar] [CrossRef] [Green Version]
  30. Chen, C.W.; Tano, D.; Akashi, M. Synthesis of platinum colloids sterically stabilized by poly(N-vinylformamide) or poly(N-vinylalkylamide) and their stability towards salt. Colloid Polym. Sci. 1999, 277, 488–493. [Google Scholar] [CrossRef]
  31. Arabatzis, I.M.; Stergiopoulos, T.; Andreeva, D. Characterization and photocatalytic activity of Au/TiO2 thin films for azo-dye degradation. J. Catal. 2003, 220, 127–135. [Google Scholar] [CrossRef]
  32. Buchachenko, A.L.; Dubinina, E.O. Photo-oxidation of water by molecular oxygen: Isotope exchange and isotope effects. J. Phys. Chem. A 2011, 115, 3196–3200. [Google Scholar] [CrossRef] [PubMed]
  33. Pourbaix, M. Atlas of Electrochemical Equilibria in Aqueous Solutions; Pergamon Press Ltd.: London, UK, 1966. [Google Scholar]
  34. Panchenko, A.; Koper, M.T.M.; Shunbina, T.E.; Mitchell, S.J.; Roduner, E. Ab Initio calculation of intermediates of oxygen reduction on low-index platinum surfaces. J. Electrochem. Soc. 2004, 151, A2016–A2027. [Google Scholar] [CrossRef] [Green Version]
  35. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G.A.; et al. Gaussian 09; Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2013. [Google Scholar]
  36. Becke, A.D. A new mixing of Hartree–Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  37. Hay, P.J.; Wadt, W.R. Ab initio effective core potentials for molecular calculations. Potentials for the transition metal atoms Sc to Hg. J. Chem. Phys. 1985, 82, 270–283. [Google Scholar] [CrossRef]
  38. Dunning, T.H.; Hay, P.J.; Schaefer, H.F. Modern Theoretical Chemistry; Schaefer, H.F., III, Ed.; Plenum Press: New York, NY, USA, 1976; Volume 3, pp. 1–28. [Google Scholar]
  39. Fukui, K.; Kato, S.; Fujimoto, H. Constituent analysis of the potential gradient along a reaction coordinate. Method and an application to methane + tritium reaction. J. Am. Chem. Soc. 1975, 97, 1–7. [Google Scholar] [CrossRef]
Figure 1. X-ray diffraction (XRD) patterns [I] of tungsten trioxide (WO3) (a) and Pt(0.4)-WO3 (platinum-tungsten trioxide) (b); ultraviolet–visible (UV–Vis) absorption spectra [II] of WO3 (c), Pt(0.1)-WO3 (d), Pt(0.2)-WO3 (e), Pt(0.4)-WO3 (f); scanning transmittance electron microscope (STEM) image (g) and energy-dispersed X-ray emission spectroscopy (EDS) (h–j) [III] of the Pt(0.4)-WO3: O (h), W (i) and Pt (j); and XPS [IV] of Pt (4f) of the Pt(0.4)-WO3.
Figure 1. X-ray diffraction (XRD) patterns [I] of tungsten trioxide (WO3) (a) and Pt(0.4)-WO3 (platinum-tungsten trioxide) (b); ultraviolet–visible (UV–Vis) absorption spectra [II] of WO3 (c), Pt(0.1)-WO3 (d), Pt(0.2)-WO3 (e), Pt(0.4)-WO3 (f); scanning transmittance electron microscope (STEM) image (g) and energy-dispersed X-ray emission spectroscopy (EDS) (h–j) [III] of the Pt(0.4)-WO3: O (h), W (i) and Pt (j); and XPS [IV] of Pt (4f) of the Pt(0.4)-WO3.
Catalysts 10 00557 g001
Figure 2. Dependence of loading amounts of Pt deposited on the WO3 [I]; and reaction time profile on the Pt(0.2)-WO3 [II] for hydroxylation of benzene to form phenol (yields: a, c; and selectivity: b and d) under visible-light irradiation.
Figure 2. Dependence of loading amounts of Pt deposited on the WO3 [I]; and reaction time profile on the Pt(0.2)-WO3 [II] for hydroxylation of benzene to form phenol (yields: a, c; and selectivity: b and d) under visible-light irradiation.
Catalysts 10 00557 g002
Figure 3. Dependence of wavelength photo-irradiated on apparent quantum yields, AQY (a) for hydroxylation of benzene to phenol, and UV–Vis absorbance of WO3 (b) and Pt(0.2)-WO3 (c).
Figure 3. Dependence of wavelength photo-irradiated on apparent quantum yields, AQY (a) for hydroxylation of benzene to phenol, and UV–Vis absorbance of WO3 (b) and Pt(0.2)-WO3 (c).
Catalysts 10 00557 g003
Figure 4. Kinetic isotope effects (KIE) on photocatalytic hydroxylation of benzene in the presence of air and H2O on Pt(0.2)-WO3 (20 mg) under visible light irradiation. Hydroxylation of C6H6 in the presence of H2O (a) and D2O (b) [I]; and that of C6H6 (c) and C6D6 (d) in the presence of H2O [II].
Figure 4. Kinetic isotope effects (KIE) on photocatalytic hydroxylation of benzene in the presence of air and H2O on Pt(0.2)-WO3 (20 mg) under visible light irradiation. Hydroxylation of C6H6 in the presence of H2O (a) and D2O (b) [I]; and that of C6H6 (c) and C6D6 (d) in the presence of H2O [II].
Catalysts 10 00557 g004
Figure 5. Optimized structures at the reactant, TS and product between benzene, H2O and H2O2 system in the absence [Ia] and presence [Ib] of Pt3 as co-catalyst for the hydroxylation of benzene to form phenol; and energy potentials for intrinsic reaction coordinates (IRC) [II]. The stabilization energy at each state was calculated by the difference from the energy of isolated system. The bond distance (Å) of O–H in H2O (blue), O–O in H2O2 (green) and C–H in benzene (red) were indicated.
Figure 5. Optimized structures at the reactant, TS and product between benzene, H2O and H2O2 system in the absence [Ia] and presence [Ib] of Pt3 as co-catalyst for the hydroxylation of benzene to form phenol; and energy potentials for intrinsic reaction coordinates (IRC) [II]. The stabilization energy at each state was calculated by the difference from the energy of isolated system. The bond distance (Å) of O–H in H2O (blue), O–O in H2O2 (green) and C–H in benzene (red) were indicated.
Catalysts 10 00557 g005
Figure 6. Reaction mechanisms for the hydroxylation of benzene into phenol in the presence of O2 and H2O on the Pt-WO3 under visible light irradiation. (*) indicates formation of activated H2O and H2O2 intermediates by photo-excitation.
Figure 6. Reaction mechanisms for the hydroxylation of benzene into phenol in the presence of O2 and H2O on the Pt-WO3 under visible light irradiation. (*) indicates formation of activated H2O and H2O2 intermediates by photo-excitation.
Catalysts 10 00557 g006
Table 1. H218O isotope labeling test for photocatalytic hydroxylation of benzene on the Pt(0.2)-WO3.
Table 1. H218O isotope labeling test for photocatalytic hydroxylation of benzene on the Pt(0.2)-WO3.
EntryProducts/µmol18O/16O 1 Ratios in Phenol/%
PHREBQCAHQPLPG
122.01.40.014.30.13.41.99.9
218.00.40.035.91.35.51.79.1
1 H216O (90%)/H218O (10%) (5 mL), benzene: 300 μmol, visible-light irradiation (entry 1) and UV-light irradiation (entry 2) for 20 h. Abbreviation: phenol (PH), resolcinol (RE), p-benzoquinone (BQ), catechol (CA), hydroquinone (HQ), phloroglucinol (PL), pyrogallol (PG).
Table 2. Effects of Ag+ ions added on the photocatalytic hydroxylation of benzene on the Pt(0.2)-WO3 photocatalyst under photo-irradiation.
Table 2. Effects of Ag+ ions added on the photocatalytic hydroxylation of benzene on the Pt(0.2)-WO3 photocatalyst under photo-irradiation.
Reaction ConditionProducts/µmolSelectivity of PH/%
PHREp-BQCAHQPLPG
Visible light/air29.01.20.218.30.319.02.441.0
Visible light/N2 (1)0000000-
UV light/N2 (1)1.10.50.50.60.26.80.0611.7
(1) presence of Ag+ (50 μmol) under N2 purge.
Table 3. Effects of tertiary butyl alcohol (TBA) added on the photocatalytic hydroxylation of benzene on the Pt(0.2)-WO3 photocatalyst under visible light irradiation for 20 h under air.
Table 3. Effects of tertiary butyl alcohol (TBA) added on the photocatalytic hydroxylation of benzene on the Pt(0.2)-WO3 photocatalyst under visible light irradiation for 20 h under air.
TBA Added/μmolProducts/µmolSelectivity of PH/%
PHREBQCAHQPLPG
029.01.20.1918.30.34192.441.0
5029.21.390.2513.30.277.95.051.0

Share and Cite

MDPI and ACS Style

Kurikawa, Y.; Togo, M.; Murata, M.; Matsuda, Y.; Sakata, Y.; Kobayashi, H.; Higashimoto, S. Mechanistic Insights into Visible Light-Induced Direct Hydroxylation of Benzene to Phenol with Air and Water over Pt-Modified WO3 Photocatalyst. Catalysts 2020, 10, 557. https://doi.org/10.3390/catal10050557

AMA Style

Kurikawa Y, Togo M, Murata M, Matsuda Y, Sakata Y, Kobayashi H, Higashimoto S. Mechanistic Insights into Visible Light-Induced Direct Hydroxylation of Benzene to Phenol with Air and Water over Pt-Modified WO3 Photocatalyst. Catalysts. 2020; 10(5):557. https://doi.org/10.3390/catal10050557

Chicago/Turabian Style

Kurikawa, Yuya, Masahiro Togo, Michihisa Murata, Yasuaki Matsuda, Yoshihisa Sakata, Hisayoshi Kobayashi, and Shinya Higashimoto. 2020. "Mechanistic Insights into Visible Light-Induced Direct Hydroxylation of Benzene to Phenol with Air and Water over Pt-Modified WO3 Photocatalyst" Catalysts 10, no. 5: 557. https://doi.org/10.3390/catal10050557

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