Next Article in Journal
Hole Doping to Enhance the Photocatalytic Activity of Bi4NbO8Cl
Previous Article in Journal
The Enhanced Performance of N-Modified Activated Carbon Promoted with Ce in Selective Catalytic Reduction of NOx with NH3
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

One-Step Catalytic or Photocatalytic Oxidation of Benzene to Phenol: Possible Alternative Routes for Phenol Synthesis?

1
Department of Industrial Engineering, University of Salerno, Via Giovanni Paolo II 132, 84084 Fisciano, Italy
2
Department of Chemistry and Biology “A. Zambelli”, University of Salerno, Via Giovanni Paolo II 132, 84084 Fisciano, Italy
*
Authors to whom correspondence should be addressed.
Catalysts 2020, 10(12), 1424; https://doi.org/10.3390/catal10121424
Submission received: 9 November 2020 / Revised: 30 November 2020 / Accepted: 3 December 2020 / Published: 5 December 2020
(This article belongs to the Section Photocatalysis)

Abstract

:
Phenol is an important chemical compound since it is a precursor of the industrial production of many materials and useful compounds. Nowadays, phenol is industrially produced from benzene by the multi-step “cumene process”, which is energy consuming due to high temperature and high pressure. Moreover, in the “cumene process”, the highly explosive cumene hydroperoxide is produced as an intermediate. To overcome these disadvantages, it would be useful to develop green alternatives for the synthesis of phenol that are more efficient and environmentally benign. In this regard, great interest is devoted to processes in which the one-step oxidation of benzene to phenol is achieved, thanks to the use of suitable catalysts and oxidant species. This review article discusses the direct oxidation of benzene to phenol in the liquid phase using different catalyst formulations, including homogeneous and heterogeneous catalysts and photocatalysts, and focuses on the reaction mechanisms involved in the selective conversion of benzene to phenol in the liquid phase.

1. Introduction

Phenol (hydroxybenzene) was discovered in coal tar and, under ambient conditions, appears as a white crystalline solid with a characteristic odor. It is an important industrial commodity, typically obtained from benzene. The use of phenol has been increasing due to its importance as a raw material from which to obtain other products [1]. The conversion of benzene to phenol possesses great relevance since phenol is a solvent and an intermediate of other kinds of industrial production, being a precursor of many materials and useful compounds [1]. Some examples of the usage of phenol are reported in the following:
(1)
PHENOLIC RESINS: by the reaction of phenol or substituted phenol with formaldehyde, phenol–formaldehyde resins or phenolic resins can be obtained. The first example was Bakelite as a commercial synthetic resin [2].
(2)
POLYCARBONATES (a very pure phenol feed is required): polycarbonates are thermoplastic polymers containing carbonate groups in their chains [3].
(3)
EPOXY RESINS: epoxy phenolic resins are resins modified at the phenolic hydroxyl group to include an epoxide functional group. This addition increases the ability of the resin to crosslink, creating a stronger polymer [4].
(4)
INTERMEDIATE FOR CAPROLACTAM (nylon production): caprolactam is a monomer for nylon production. Among the routes for its manufacture, one is via cyclohexanone and cyclohexanone oxime. Cyclohexanone can be prepared either from phenol or from cyclohexane. The phenol route is a two-stage process, in which the first stage foresees the reaction among phenol and hydrogen in the presence of a nickel catalyst at around 180 °C to form cyclohexanol, subsequently dehydrogenated at around 400 °C in the presence of a copper catalyst to yield the cyclohexanone [5].
The cumene process (also known as the Hock process) is the most important industrial process for the simultaneous production of phenol and acetone starting from benzene and propylene [6]. Oxygen from air and small amounts of a radical initiator are also reactants required for the process. The cumene process is very complex, as it consists of several stages, passing through the formation of hydroperoxide (a very reactive substance that can give runaway phenomena), which allows the indirect production of phenol and acetone [7]. The process has three main reaction steps plus, furthermore, a step of concentration of cumene hydroperoxide, and they are as follows: (i) production of cumene; (ii) conversion of cumene to cumene hydroperoxide; (iii) concentration of cumene hydroperoxide; (iv) hydrolysis of cumene hydroperoxide [1].
The scheme of the reactions is reported in Figure 1.
The production of cumene (isopropylbenzene) is a Friedel–Crafts reaction and occurs by the reaction between benzene and propene, using an acid catalyst [8].
In one process, benzene and propene (3:1 mole ratio) are passed over an acid catalyst. The excess benzene acts to limit the polyalkylations and the by-reactions of the oligomerization of propene. The zeolite is more environmentally friendly than traditional acid catalysts. The problems are related to selectivity because isomers can be produced with respect to cumene [8].
The second step, the conversion of cumene to cumene hydroperoxide, involves the use of air to give the hydroperoxide in the presence of small quantities of a radical initiator (benzoyl peroxide, for example) in slightly basic conditions.
The reaction is autocatalyzed by cumene hydroperoxide. The reaction is carried out at temperatures between 77 and 117 °C and 1–7 atm pressure, to hold the system in the liquid phase.
After the concentration of the cumene hydroperoxide, performed usually with an evaporator at the descendent film, the third and final reaction is the decomposition of cumene hydroperoxide by mixing with sulfuric acid at 40–100 °C to give, after neutralization, phenol and propanone (acetone). Then, the products are separated through distillation columns.
The economics and effectiveness of this process are related to the market of acetone, apart from phenol. Often, much more phenol is needed than the propanone that is produced at the same time. Moreover, this multistage process has a low overall yield (less than 5%), requires high energy, and the formation of by-products such as acetophenone, 2-phenylpropan-2-ol, and α-methylstyrene is encountered [9].
Today, almost 95% of the worldwide phenol production is based on the “cumene process”, despite the previously mentioned drawbacks, which are poor ecology, the formation of an explosive intermediate (cumene hydroperoxide), high capital investment, high acetone production as a co-product, which results in an oversupply in the market, and a multistep characteristic which makes it difficult to achieve high phenol yields with respect to benzene feed [10]. Therefore, it is highly desirable to develop alternative synthetic processes of phenol which are more efficient and environmentally benign. In fact, in this regard, great interest is devoted to the process in which the direct oxidation of benzene to phenol is achieved, thanks to the use of suitable catalysts and oxidant species.
Therefore, this review aims to summarize the catalytic and photocatalytic formulations studied to date for the direct conversion of benzene to phenol in the liquid phase. It is worthwhile to note that, to the best of our knowledge, only a few review articles on this topic, mainly devoted to the role of oxidant molecules (O2 or H2O2) [11,12], to the use of molecular sieves [13], and to the use of nano-biomimetic metal oxide catalysts [14], are present in the literature. Additionally, until now, no review which summarizes photocatalyst formulations suitable for the direct oxidation of benzene to phenol in the liquid phase has been developed.

2. Homogeneous and Heterogeneous Catalysts for the One-Step Catalytic Oxidation of Benzene to Phenol in Liquid Phase

As highlighted above, the development of a process able to synthesize phenol from benzene in a one-step reaction with high benzene conversion and high phenol selectivity is highly desired, both from environmental safety and economical points of view [11].
Theoretically, the high conversion of benzene to phenol through oxidation reactions (Figure 2) is possible, but the experimental results evidence that this goal is difficult to achieve. The more investigated process is the direct hydroxylation of benzene using N2O [15], O2 [16], H2O2 [17,18], and a mixture of O2 and H2 [19] as oxidant species.
In all cases, the thermodynamic analysis (carried out utilizing the data reported in Table 1) indicates that the reaction is irreversible and 100% conversion is feasible.
In particular, high selectivity to phenol could be achieved using N2O as oxidant, but high reaction temperature is required and sources of N2O are limited. On the other hand, air or oxygen is easily available. Among the oxidant agents, hydrogen peroxide is the most considered oxidant since water is the only by-product and the process is simple, green, and economic [13]. On the other hand, the hydroxylation reaction of benzene with the presence of only the oxidant is very slow and not able to oxidize benzene into phenol [20].
For this reason, to develop this process, it is essential to identify the right catalyst capable of guaranteeing high selectivity to phenol together with a high conversion of benzene. The main driving force of the development of new, efficient oxygenation catalysts is to selectively hydroxylate the non-activated C–H bonds of the benzene molecule in order to reduce the steps required in the preparation of phenol [21]. However, the direct introduction of hydroxyl into the benzene molecule is very difficult because of the low reactivity of aromatic C−H bonds and the strong nucleophilicity of hydroxyl free radicals [22]. Additionally, phenol is more reactive than benzene, resulting in further oxidation reactions of phenol and, consequently, the selectivity worsens [23], as also recently underlined in a paper dealing with the catalytic conversion of benzene to p-benzoquinone [24]. Therefore, the formulation of effective catalysts for phenol production, with a good yield and a high selectivity to the desired product, is highly desirable.
Numerous homogeneous catalysts have been tested [25,26,27], such as high active Co complexes [28], Ni complexes [29], Cu-based complexes [30,31,32,33,34], as well as nitrogen- [35,36] or oxygen-ligated [37,38] iron complexes. Additionally, Os complexes were discussed as non-trivial catalysts for benzene oxidation (oxidation with H216O2 under 18O2 gave phenol that did not contain the 18O isotope), pointing out “Os=O” as oxidizing species responsible for phenol formation [39].
Very interesting results were reported using ionic liquids with hydrogen peroxide as oxidant and ferric tri (dodecanesulfonate) as a catalyst in an aqueous solution [25]. In particular, thanks to the presence of 1-n-octyl-3-methylimidazolium tetrafluoroborate as an ionic liquid, enhanced benzene conversion (54%) together with significant phenol selectivity (90%) were observed after 6 h and at a reaction temperature of 50 °C. Additionally, the authors showed that, in the absence of ionic liquid, lower conversion and selectivity were achieved because of the formation of hydroquinone and biphenyl as by-products, underlining that the aqueous–ionic liquid biphasic reaction system is able to enhance both the benzene conversion and phenol selectivity [25]. Moreover, it was underlined that the ionic liquid containing the catalyst may be separated from the products in the aqueous phase by a simple decantation step [25].
H5PV2Mo10O40 polyoxometalate (POM) as a homogeneous catalyst is able to oxidize benzene to phenol at room temperature in the presence of only O2 as the oxidant molecule at 170 °C for 6 h [40]. With these operating conditions, hydroxylation to phenol took place. However, they underlined that the oxidation of benzene is possible even at room temperature and that phenol can be formed thanks to the formation of a benzene radicaloid species (Figure 3) in the presence of H2SO4 aqueous solution via an electron transfer mechanism and the subsequent oxidation reaction promoted by O2 present in the reaction medium [40].
Amphiphilic poly(ionic liquid) (PIL)/Wells–Dawson-type phosphovanadomolybdate (V-POM) ionic composites were also studied in benzene hydroxylation with H2O2, showing a phenol yield of 37.3% with a selectivity of 100% [41].
Other interesting homogeneous catalysts for the hydroxylation of benzene to phenol are a series of first-row transition metal complexes with Schiff base ligands or with readily available acetylacetonate ligands, which were studied for the hydroxylation of benzene to phenol at 50 °C in acetonitrile, using hydrogen peroxide as the oxidant [42]. In this context, Fe(II) complex with the N4 Schiff base ligand allowed the achievement of phenol selectivity and benzene conversion equal to 98% and 64%, respectively [42]. On the other hand, Fe(II) and Fe(III) acetylacetonate complexes evidenced 96% selectivity to phenol, but with a lower benzene conversion (20–22%). Therefore, the reported results evidenced that phenol was the main reaction product, without the formation of biphenyl as a by-product, suggesting that the reaction does not take place through a free radical mechanism [36].
A solution different to homogeneous catalysis is based on the development of solid catalysts, which offer more efficiency together with high stability under the reaction conditions. Heterogeneous catalysts, if compared to homogeneous catalysts, have some advantages, such as catalyst recovery and recycling, which is the primary goal in view of the possible industrialization of the process. Different studies deal with the oxidation of benzene to phenol by using a transition metal oxide (such as Ti, V, Mn, Fe, Co, Cr, Mo) supported on a different metal oxide (such as Al2O3, SiO2, and TiO2). Among them, it was reported that vanadium-based compounds exhibited excellent catalytic activity towards the hydroxylation of benzene to phenol [43]. For example, as reported in Table 2, Shijina et al. analyzed the oxidation reaction carried out over supported vanadia. The experimental results evidenced that, at 60 °C, the activity increases, with an increase in vanadia content up to 13.8%. After monolayer dispersion, i.e., above 14% V2O5, the percentage conversion of benzene decreased [44]. However, the weak interaction of V and support leads to the leaching of active phases from the support [45]. Furthermore, for example, in the case of vanadium supported on Al2O3, the presence of stable V2O5 disfavors the redox cycle between V5+ and V4+ [46], reflected in the low catalytic activity.
Molecular sieves (such as MCM-41, SBA15, SBA16, NaY) with incorporated transition metals have attracted much interest because of the high catalytic activity for the oxidation of organic compounds, such as benzene. Moreover, in this case, the presence of heteroatoms (such as V, Cu, and Fe) in molecular sieves is able to modify the surface properties so as to obtain highly dispersed and isolated active sites in the silica framework and, therefore, improving the catalytic activity. Figure 4 presents the mechanism suggested by Jourshabani et al. [56] over Fe-SBA16. The authors suggested that, compared with other Fe-based catalysts, the high catalytic activity of the systems may be attributed to SBA-16, which allows a guaranteed high mass transfer rate of benzene on the Fe/SBA-16 surface.
In detail, the authors proposed a reaction mechanism using H2O2 in the presence of Fe/SBA-16, underlining that only isolated species of Fe3+ are the active phases. H2O2 is activated on Fe/SBA-16 by chemisorption on the surface of the supported Fe, together with the formation of an open bi-radical form of the iron−peroxo complex. These radicals may coordinate to Fe present in Fe/SBA-16, forming an iron−peroxo complex. Similarly, a mechanism was proposed by Hu et al. [59] using vanadium-containing nitrogen-doped mesoporous carbon catalysts. Moreover, the same authors reported a further increase in phenol yield using carbon materials as the active phase for vanadium. Carbon materials, due to their high specific area, large pore volume, and hydrophobic surface properties, represent promising supports for the hydroxylation of benzene to phenol [60]. In particular, the 4V/MCN-S catalyst (Table 2) exhibited remarkable catalytic performance, with a benzene conversion of 38.2% and phenol selectivity of 96.1%, along with good reusability. Recently, as reported in Table 2, metal-doped carbon nitride [62,63,64,65] showed improved activity for the hydroxylation of benzene with H2O2 to phenol but remains unsatisfactory since phenol is further oxidized, forming other by-products [13].
On the contrary, different results were obtained by Tu et al., who synthesized by the solvothermal method an Fe-based metal-organic framework (MOF) named Fe-TBAPy (Figure 5).
Fe-TBAPy is built from [Fe(OH)(CO2)2]∞ rod-shaped SBUs (SBUs = secondary building units) and 1,3,6,8-tetrakis(p-benzoate)pyrene (TBAPy4−). The experimental tests evidenced a high phenol yield and selectivity equal to 64.5% and 92.9%, respectively. These results demonstrated the possibility to formulate MOFs possessing enhanced catalytic activity for benzene hydroxylation. However, even if, in the last few years, significant progress has been achieved, the catalytic performance is still limited by the intrinsic drawbacks of the employed catalysts. For example, the hydrophilicity of the catalyst’s surface worsens the adsorption of non-polar benzene on its surfaces and, on the contrary, favors the adsorption of polar phenol molecules, with consequently low benzene conversion and poor phenol selectivity [17]. Moreover, in some cases, the leaching of active species during the reaction is one of the main causes of deactivation of the catalyst [68].

3. Homogeneous and Heterogeneous Photocatalysts for the One-Step Catalytic Oxidation of Benzene to Phenol in Liquid Phase

The selective hydroxylation of benzene to phenol by means of photocatalysis using different oxidizing agents including O2 [69], N2O [70], or H2O2 [71] was the object of several research papers.
For this purpose, both homogeneous and heterogeneous photocatalysts were studied, although the latter evidenced some drawbacks. More specifically, in a homogeneous system, it is difficult to separate the catalyst from the reaction products [72]. In the case of homogeneous systems, Fenton’s reaction is a well-known homogeneous oxidation process in which Fe2+ is used as a catalyst and hydrogen peroxide as oxidant [73]. However, this process requires acidic conditions that lead to corrosion phenomena [74], and also more than 40% of the used hydrogen peroxide is consumed by side reactions [75].
Moreover, the literature reported the selective oxygenation of benzene to phenol with an oxygen-saturated acetonitrile solution containing benzene water and at ambient conditions using 3-cyano-1-methylquinolinium ion (QuCN+) as a homogeneous photocatalyst, showing a strong oxidizing ability towards benzene [76]. The hydroxylation of benzene to phenol under UV irradiation was also studied using alkoxohexavanadate anions and quinolinium ions, achieving high phenol selectivity (>99 %) together with a good yield (around 50%) [77].
As an alternative to homogeneous photocatalysis, the heterogeneous photocatalytic process could represent a possible green alternative for selective oxidation reactions [78,79,80,81,82,83]. Among the semiconductor photocatalysts, TiO2 is the most used material because of its chemical stability and its high oxidizing ability. It has been reported that the crystalline anatase phase of TiO2 has higher photocatalytic activity, if compared to rutile TiO2, because the anatase phase has higher levels of hydroxyl groups on its surface [84]. The anatase TiO2 presents a wide band gap of 3.2 eV and it is normally activated under ultraviolet (UV) light [85,86]. When a TiO2 photocatalyst is irradiated with energy greater than the TiO2 band gap energy, positive charge-holes are generated in the valence band, while the electrons are promoted in the conduction band. Both positive holes and electrons take part in oxidation–reduction reactions. In particular, the photogenerated positive hole is able to react with adsorbed water to produce hydroxyl radicals, whereas the electron can reduce O2, generating strongly oxidizing superoxide ions. These highly reactive species, such as hydroxyl radicals, are employed for phenol production from benzene under mild conditions through the photocatalytic process [87].
Table 3 reports some photocatalytic systems studied in the literature for the direct oxidation of benzene to phenol in the liquid phase.
Shiraishi et al. considered TiO2 with a mesoporous structure (mTiO2) synthesized through the surfactant-templating method and aggregation method consisting of the nanosized TiO2 (nTiO2) particle combination followed by sintering of the particles [88].
In particular, they underlined that the presence of “mesopores” allows a smooth diffusion of molecules, which is indispensable for the adsorption-driven activity. As can be seen from Table 3, the use of the TiO2 sample with a mesoporous structure showed a higher yield and selectivity to phenol (>80%) than TiO2 with a nonporous structure [88].
Yuzawa et al. studied the direct hydroxylation of benzene in water using a Pt/TiO2 photocatalyst [89]. In this photocatalytic reaction system, the selection of the wavelength associated with the incident light wavelength, exclusion of oxygen, and optimization of platinum loading amount were important factors to achieve selective phenol production. When Pt/TiO2 was irradiated, the formation of phenol and small amounts of biphenyl, cyclohexanol, cyclohexanone, and carbon dioxide were detected according to the following steps (Figure 6):
(1)
hydroxylation of benzene with water;
(2)
coupling of benzene;
(3)
reduction of the produced phenol and successive oxidation of the produced cyclohexanol to cyclohexanone;
(4)
decomposition of benzene with water.
Despite the fact that the phenol selectivity was high (88%), an excess amount of hydrogen (95 μmol) was also produced. However, when the irradiation wavelength was longer than 385 nm, higher phenol selectivity (91%) was obtained with no production of biphenyl, carbon dioxide, and hydrogen. Thus, the irradiation wavelength appears as a factor to be considered for obtaining the selective aromatic ring hydroxylation of benzene.
Moreover, it was shown that the presence of air or oxygen effectively promotes complete benzene oxidation to carbon dioxide and water [101], according to the reaction reported in Figure 7:
Thus, the absence of oxygen in the reactor is required to achieve high selectivity to phenol. The relationship between the platinum loading amount and the product yield was also identified. The sample with 0.1 wt% of platinum loaded onto the TiO2 surface exhibited the highest phenol yield. Thus, the optimum Pt amount for phenol production was found to be 0.1 wt% because phenol selectivity equal to 91% was obtained.
Yusuke et al. reported a versatile way to modify the efficiency and phenol selectivity of heterogeneous photocatalytic oxidation [90]. They found that the sunlight-induced photocatalytic oxidation activity of aqueous benzene to phenol on Au/TiO2 nanoparticles was improved when the reaction was carried out under CO2 atmosphere (230 kPa). In more detail, benzene is directly oxidized to phenol and the produced phenol is further converted to more oxidized products, such as catechol, hydroxyquinone, and trihydroxybenzenes, and finally mineralized to CO2. On the contrary, when the reaction is carried out under CO2 pressure, the presence of carbon dioxide probably suppressed the successive oxidation of phenol, increasing the yield and selectivity to phenol [90].
Devaraji et al. studied the photoactivity of TiO2, V-doped TiO2, and Au-V-doped TiO2 under UV irradiation. V-doped TiO2 showed benzene conversion equal to 3% with 100% phenol selectivity after 6 h of irradiation time [102]. The same authors evidenced that the level of benzene conversion increased linearly from 3% to 13% by increasing the irradiation time from 6 to 24 h, respectively. After Au deposition, the photocatalytic activity strongly increased because the incorporation of V in the TiO2 lattice generated a V5+ energy level below the conduction band of TiO2, which helped to trap the excited electrons, whereas Au deposited over V-doped TiO2 acted as an electron sink [102].
As an alternative to TiO2 as support for noble metals, WO3 has been considered. In particular, Pt/WO3 photocatalysts showed much higher selectivity to phenol than commercial TiO2 and Pt/TiO2 [91]. In particular, the photocatalytic reaction using Pt/WO3-K (Table 3) showed the highest selective phenol production, with ~74% of selectivity and with ~69% of benzene conversion after 240 min of UV irradiation time. The production of phenol was also observed with the Pt/TiO2 photocatalyst but the phenol production was saturated within 60 min of UV irradiation, with a noticeable increase in the CO2 gaseous phase, indicating that there was further oxidation of the generated phenol. Instead, after 60 min of visible light irradiation (λ > 400 nm), better phenol selectivity (~84%) was reported with Pt/WO3-K (Table 3).
In the same paper, considering an aqueous benzene solution, it has been reported that hydroxyl radicals (•OH) are produced through the reaction of photogenerated holes with water molecules adsorbed on the photocatalyst surface. The hydroxyl radical reacts with benzene to generate hydroxylated benzene radical, which is then oxidized by a positive hole on the photocatalyst surface and deprotonated, producing phenol (Figure 8).
Considering the previous papers, it appears that noble metals deposited on the TiO2 surface act as a good co-catalyst to improve the photoactivity. At the same time, these elements are quite expensive and, therefore, many efforts have devoted to replacing them with low-cost materials [91]. This led to the design of new, low-cost, and efficient photocatalysts that yield high benzene conversion and high selectivity to phenol at ambient temperature and pressure. The surface properties of TiO2 have been modified by metal impregnation in order to increase the phenol production as well as its selectivity. In this regard, Gupta et al. reported the oxidation of benzene to phenol under UV irradiation using Fe3+ (at 5 wt%) impregnated on a TiO2 catalyst [87]. The experimental tests revealed that the optimum selectivity to phenol (80-86%) was achieved during 1–2 h of UV irradiation time. The enhanced photoactivity of the Fe3+ impregnated on TiO2 is probably caused by the structural defects induced by the presence of Fe3+ ions on the TiO2 surface [103].
Some interest is also directed towards doping TiO2 with metallic elements to broaden the absorption spectrum into the visible range. In particular, the TiO2 doping with transition metal ions like Cr, Fe, and V leads to the generation of a new acceptor level in the conduction band of TiO2, which acts as an electron trapping center, increasing the number of holes available in the valence band, and simultaneously promotes the reduction of molecular oxygen to H2O2 during the photocatalytic reaction. The doping of TiO2 with Fe enhances the photocatalytic activity because it allows an increase in the electron transfer process and the electron-hole separation, thus minimizing charge carrier recombination. This minimization of charge carrier recombination is essential for redox reactions [104].
Experimental tests on the photocatalytic oxidation reaction of benzene to phenol under UV or visible light irradiation were conducted by Perumal et al. [86] using Fe-Cr-doped TiO2. It was shown that, after 12 h of UV irradiation time, 28% of benzene conversion and 90% selectivity to phenol were achieved with H2O2 as oxidant (Table 3). These results evidenced that the simultaneous presence of Cr and Fe in the TiO2 lattice led to higher photocatalytic activity towards phenol production if compared with Fe-doped TiO2 and Cr-doped TiO2. Additionally, Fe-Cr-doped TiO2 also showed visible light absorption but low benzene conversion was observed even after 24 h of visible light irradiation, suggesting that the photocatalytic activity for benzene oxidation depends on the valence band position of TiO2 and that, under visible light irradiation, a small number of holes were produced as compared to those generated upon UV light irradiation [105]. Figure 9 shows the possible reaction mechanism on Fe-Cr-doped TiO2 in which phenol is produced via two reaction paths (path-A and path-B).
In the presence of UV light, electrons are promoted from the valence band (VB) to the conduction band (CB) In path-A, the trapped electrons in the dopant level can reduce Fe3+ to Fe2+, which then is able to react with H2O2 and a proton (coming from H2O) to produce a hydroxyl radical and Fe3+. The hydroxyl radical originates from both water and H2O2 and they attack the aromatic benzene ring to form hydroxycyclohexadienyl radicals [106]. The photogenerated positive holes in the VB or Fe3+ oxidize the hydroxycyclohexadienyl radicals to phenol via a deprotonation process to restart the photocatalytic cyclic reaction [107]. In path-B, the generated holes in the VB react with benzene to produce benzene radical ions. These radical ions react with hydroxyl radicals, forming phenol, probably via deprotonation of an unstable intermediate.
To obtain higher selectivity to phenol as well as better conversion of benzene, the researchers focused their attention on more complex systems (composite photocatalysts) that showed excellent photocatalytic performance. In this regard, carbon nitride materials were extensively considered (Table 3). For example, Zhang and Park studied CuPd bimetallic alloy nanoparticle-coated holey carbon nitride materials (g-C3N4/CuPd) as photocatalyst [94]. Compared with bare g-C3N4, a significant increase in phenol selectivity is achieved when CuPd bimetallic alloy nanoparticles are uniformly dispersed on the support surface. The catalyst containing 0.5 wt% of CuPd bimetallic alloy particles showed high photocatalytic activity in the oxidation of benzene to phenol (benzene conversion: 98.1% and selectivity to phenol: 89.6%) after 90 min of solar irradiation, probably linked to the uniform distribution of CuPd bimetallic alloy nanoparticles and the synergistic effect between CuPd particles and g-C3N4. These two factors may contribute to the improvement of both solar energy use and the photo-induced electron-hole pairs, resulting in enhanced performance for the selective benzene oxidation to phenol [94].
Zhang et al. also focused their attention on the synthesis of phenol under visible light irradiation using a physical mixture of Fe salts (FeCl3 or FeCl2) and mesoporous carbon nitride (FeCl3/mpg-C3N4). The reactions were carried out with water–acetonitrile as solvents and H2O2 as oxidant using FeCl3/mpg-C3N4 samples with different FeCl3 loadings (in the range 3–20 wt%). The 5 wt% FeCl3/mpg-C3N4 photocatalyst evidenced a benzene conversion of ~38% and a phenol selectivity equal to 97%. They also investigated the possible mechanism under visible light (Figure 10). In detail, two main steps are involved in the visible light oxidation of benzene to phenol: the electrons generated by light irradiation reduce Fe3+ to Fe2+, which is able to decompose H2O2 with the formation of ·OH. The hydroxylation of benzene by ·OH produces a cyclohexadienyl radical intermediate (A) and then the positive hole (h+) of mpg-C3N4 oxidizes A to phenol.
In some cases, metal−organic framework (MOF)-based photocatalysts also evidenced significant photoactivity under visible light. For example, Wang et al. [98] studied selective benzene hydroxylation to phenol over two Fe-based MOFs (MIL-100(Fe) and MIL-68(Fe)) under visible light irradiation using H2O2 as oxidant. MIL-100(Fe) showed much intensive absorption in the visible light region compared with MIL-68(Fe), which could justify its enhanced photocatalytic performance. In fact, the MIL-100(Fe) photocatalyst evidenced a benzene conversion equal to 20.1% and higher selectivity to phenol (98%). Instead, lower benzene conversion (14%) and selectivity to phenol (90%) were observed over MIL-68(Fe).
Another studied material was a Zn2Ti-layered double hydroxide (ZnTi-LDH) photocatalyst, which showed an enhancement of photoinduced charge carrier separation due to the presence of oxygen vacancies on the LDH surface and an increase in superoxide radicals [100]. The band structure of the ZnTi-LDH photocatalyst allowed the realization of advanced activity, with selectivity to phenol equal to 87.18% in water with air as oxidant under UV-vis light irradiation.

4. Concluding Remarks and Perspectives

The one-step oxidation of benzene to phenol in the liquid phase has been extensively studied from a scientific point of view in recent years due to the importance of phenol in industrial chemistry. Several catalysts and photocatalysts (both homogenous and heterogeneous) able to work under mild conditions were proposed in the literature and special attention to the reaction mechanism was also given in most papers. Despite some catalytic formulations showing significant benzene conversion and phenol selectivity, it must be taken into account that one of the main difficulties in performing the selective oxidation of benzene is that the desired product (phenol) is easily over-oxidized to hydroxyphenols and finally oxidized into carbon dioxide. Therefore, it is still essential to study deeply and/or to develop new catalytic and photocatalytic materials able to achieve high selectivity to phenol at high conversion of benzene and, as a consequence, to ensure high phenol yield. One promising strategy could be the formulation of supports for active phases (photocatalytic or not) with high benzene adsorption capacity and with low affinity towards phenol in a manner to prevent further oxidation reactions of the desired product.
Additionally, most catalytic formulations were studied in slurry reactors. Therefore, in response to the question asked in the title of this review article, there is still a need to consider three main aspects for the possible industrialization of a catalytic reactor devoted to the one-step oxidation of benzene to phenol in the liquid phase:
  • the immobilization of the catalysts or photocatalysts on macroscopic supports (i.e., the development of structured catalysts) to avoid the separation of catalyst powders from the liquid phase containing phenol at the end of the oxidation step;
  • the development of structured catalysts or photocatalysts with high stability and which are easily recyclable;
  • the development of novel selective oxidation systems (e.g., highly efficient photoanodes for the photoelectrocatalytic oxidation of benzene to phenol);
  • the design of efficient and low-cost systems to recover the produced phenol from the liquid phase.

Author Contributions

A.M., O.S. and D.S. wrote the manuscript. O.S. and V.V. (Vincenzo Vaiano) provided the concept. V.V. (Vincenzo Venditto), D.S., A.M., O.S. and V.V. (Vincenzo Vaiano), 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. Schmidt, R.J. Industrial catalytic processes—Phenol production. Appl. Catal. A Gen. 2005, 280, 89–103. [Google Scholar] [CrossRef]
  2. Solyman, W.S.; Nagiub, H.M.; Alian, N.A.; Shaker, N.O.; Kandil, U.F. Synthesis and characterization of phenol/formaldehyde nanocomposites: Studying the effect of incorporating reactive rubber nanoparticles or Cloisite-30B nanoclay on the mechanical properties, morphology and thermal stability. J. Radiat. Res. Appl. Sci. 2017, 10, 72–79. [Google Scholar] [CrossRef]
  3. Pryde, C.; Hellman, M. Solid state hydrolysis of bisphenol-A polycarbonate. I. Effect of phenolic end groups. J. Appl. Polym. Sci. 1980, 25, 2573–2587. [Google Scholar] [CrossRef]
  4. Takeichi, T.; Furukawa, N. Epoxy Resins and Phenol-Formaldehyde Resins. In Polymer Science: A Comprehensive Reference; Elsevier BV: Amsterdam, The Netherlands, 2012; pp. 723–751. [Google Scholar]
  5. Brydson, J.A. Plastics Materials; Elsevier: Amsterdam, The Netherlands, 1999. [Google Scholar]
  6. Zakoshansky, V. The cumene process for phenol-acetone production. Pet. Chem. 2007, 47, 273–284. [Google Scholar] [CrossRef]
  7. Fortuin, J.; Waterman, H. Production of phenol from cumene. Chem. Eng. Sci. 1953, 2, 182–192. [Google Scholar] [CrossRef]
  8. The Essential Chemical Industry—Online. Available online: https://www.essentialchemicalindustry.org/chemicals/phenol.html (accessed on 2 October 2020).
  9. Park, H.; Choi, W. Photocatalytic conversion of benzene to phenol using modified TiO2 and polyoxometalates. Catal. Today 2005, 101, 291–297. [Google Scholar] [CrossRef]
  10. Molinari, R.; Poerio, T. Remarks on studies for direct production of phenol in conventional and membrane reactors. Asia Pac. J. Chem. Eng. 2010, 5, 191–206. [Google Scholar] [CrossRef]
  11. Fukuzumi, S.; Ohkubo, K. One-Step Selective Hydroxylation of Benzene to Phenol. Asian J. Org. Chem. 2015, 4, 836–845. [Google Scholar] [CrossRef]
  12. Ottenbacher, R.V.; Talsi, E.P.; Bryliakov, K.P. Recent progress in catalytic oxygenation of aromatic C–H groups with the environmentally benign oxidants H2O2 and O2. Appl. Organomet. Chem. 2020, 34, e5900. [Google Scholar] [CrossRef]
  13. Jiang, T.; Wang, W.; Han, B. Catalytic hydroxylation of benzene to phenol with hydrogen peroxide using catalysts based on molecular sieves. New J. Chem. 2013, 37, 1654–1664. [Google Scholar] [CrossRef]
  14. Wanna, W.H.; Janmanchi, D.; Thiyagarajan, N.; Ramu, R.; Tsai, Y.-F.; Yu, S.S. Selective Oxidation of Simple Aromatics Catalyzed by Nano-Biomimetic Metal Oxide Catalysts: A Mini Review. Front. Chem. 2020, 8, 589178. [Google Scholar] [CrossRef] [PubMed]
  15. Yuranov, I.; Bulushev, D.A.; Renken, A.; Kiwi-Minsker, L. Benzene to phenol hydroxylation with N2O over Fe-Beta and Fe-ZSM-5: Comparison of activity per Fe-site. Appl. Catal. A Gen. 2007, 319, 128–136. [Google Scholar] [CrossRef]
  16. Guo, H.; Chen, Z.; Mei, F.; Zhu, D.; Xiong, H.; Yin, G. Redox Inactive Metal Ion Promoted C H Activation of Benzene to Phenol with PdII (bpym): Demonstrating New Strategies in Catalyst Designs. Chem. Asian J. 2013, 8, 888–891. [Google Scholar] [CrossRef] [PubMed]
  17. Hu, L.; Wang, C.; Ye, L.; Wu, Y.; Yue, B.; Chen, X.; He, H. Direct hydroxylation of benzene to phenol using H2O2 as an oxidant over vanadium-containing mesoporous carbon catalysts. Appl. Catal. A Gen. 2015, 504, 440–447. [Google Scholar] [CrossRef]
  18. Parida, K.; Rath, D. Structural properties and catalytic oxidation of benzene to phenol over CuO-impregnated mesoporous silica. Appl. Catal. A Gen. 2007, 321, 101–108. [Google Scholar] [CrossRef]
  19. Niwa, S.-i.; Eswaramoorthy, M.; Nair, J.; Raj, A.; Itoh, N.; Shoji, H.; Namba, T.; Mizukami, F. A one-step conversion of benzene to phenol with a palladium membrane. Science 2002, 295, 105–107. [Google Scholar] [CrossRef]
  20. Al-Sabagh, A.; Yehia, F.; Eshaq, G.; ElMetwally, A. Eclectic hydroxylation of benzene to phenol using ferrites of Fe and Zn as durable and magnetically retrievable catalysts. ACS Sustain. Chem. Eng. 2017, 5, 4811–4819. [Google Scholar] [CrossRef]
  21. Monfared, H.H.; Amouei, Z. Hydrogen peroxide oxidation of aromatic hydrocarbons by immobilized iron (III). J. Mol. Catal. A Chem. 2004, 217, 161–164. [Google Scholar] [CrossRef]
  22. Lyu, Y.-J.; Qi, T.; Yang, H.-Q.; Hu, C.-W. Performance of edges on carbon for the catalytic hydroxylation of benzene to phenol. Catal. Sci. Technol. 2018, 8, 176–186. [Google Scholar] [CrossRef]
  23. Yamada, M.; Karlin, K.D.; Fukuzumi, S. One-step selective hydroxylation of benzene to phenol with hydrogen peroxide catalysed by copper complexes incorporated into mesoporous silica–alumina. Chem. Sci. 2016, 7, 2856–2863. [Google Scholar] [CrossRef] [Green Version]
  24. Wanna, W.H.; Ramu, R.; Janmanchi, D.; Tsai, Y.-F.; Thiyagarajan, N.; Yu, S.S.-F. An efficient and recyclable copper nano-catalyst for the selective oxidation of benzene to p-benzoquinone (p-BQ) using H2O2 (aq) in CH3CN. J. Catal. 2019, 370, 332–346. [Google Scholar] [CrossRef]
  25. Peng, J.; Shi, F.; Gu, Y.; Deng, Y. Highly selective and green aqueous–ionic liquid biphasic hydroxylation of benzene to phenol with hydrogen peroxide. Green Chem. 2003, 5, 224–226. [Google Scholar] [CrossRef]
  26. Nomiya, K.; Yagishita, K.; Nemoto, Y.; Kamataki, T.-a. Functional action of Keggin-type mono-vanadium (V)-substituted heteropolymolybdate as a single species on catalytic hydroxylation of benzene in the presence of hydrogen peroxide. J. Mol. Catal. A Chem. 1997, 126, 43–53. [Google Scholar] [CrossRef]
  27. Jiang, W.; Zhu, W.; Li, H.; Chao, Y.; Xun, S.; Chang, Y.; Liu, H.; Zhao, Z. Mechanism and optimization for oxidative desulfurization of fuels catalyzed by Fenton-like catalysts in hydrophobic ionic liquid. J. Mol. Catal. A Chem. 2014, 382, 8–14. [Google Scholar] [CrossRef]
  28. Anandababu, K.; Muthuramalingam, S.; Velusamy, M.; Mayilmurugan, R. Single-step benzene hydroxylation by cobalt (ii) catalysts via a cobalt (iii)-hydroperoxo intermediate. Catal. Sci. Technol. 2020, 10, 2540–2548. [Google Scholar] [CrossRef]
  29. Muthuramalingam, S.; Anandababu, K.; Velusamy, M.; Mayilmurugan, R. One step phenol synthesis from benzene catalysed by nickel (ii) complexes. Catal. Sci. Technol. 2019, 9, 5991–6001. [Google Scholar] [CrossRef]
  30. You, X.; Wei, Z.; Wang, H.; Li, D.; Liu, J.; Xu, B.; Liu, X. Synthesis of two copper clusters and their catalysis towards the oxidation of benzene into phenol. RSC Adv. 2014, 4, 61790–61798. [Google Scholar] [CrossRef]
  31. Kulakova, A.N.; Bilyachenko, A.N.; Levitsky, M.M.; Khrustalev, V.N.; Korlyukov, A.A.; Zubavichus, Y.V.; Dorovatovskii, P.V.; Lamaty, F.; Bantreil, X.; Villemejeanne, B. Si10Cu6N4 cage hexacoppersilsesquioxanes containing N ligands: Synthesis, structure, and high catalytic activity in peroxide oxidations. Inorg. Chem. 2017, 56, 15026–15040. [Google Scholar] [CrossRef]
  32. Tsuji, T.; Zaoputra, A.A.; Hitomi, Y.; Mieda, K.; Ogura, T.; Shiota, Y.; Yoshizawa, K.; Sato, H.; Kodera, M. Specific enhancement of catalytic activity by a dicopper core: Selective hydroxylation of benzene to phenol with hydrogen peroxide. Angew. Chem. Int. Ed. 2017, 129, 7887–7890. [Google Scholar] [CrossRef]
  33. Conde, A.; Diaz-Requejo, M.M.; Pérez, P.J. Direct, copper-catalyzed oxidation of aromatic C–H bonds with hydrogen peroxide under acid-free conditions. Chem. Commun. 2011, 47, 8154–8156. [Google Scholar] [CrossRef] [Green Version]
  34. Kumari, S.; Muthuramalingam, S.; Dhara, A.K.; Singh, U.; Mayilmurugan, R.; Ghosh, K. Cu (I) complexes obtained via spontaneous reduction of Cu (II) complexes supported by designed bidentate ligands: Bioinspired Cu (I) based catalysts for aromatic hydroxylation. Dalton Trans. 2020, 49, 13829–13839. [Google Scholar] [CrossRef] [PubMed]
  35. Kudrik, E.V.; Sorokin, A.B. N-Bridged Diiron Phthalocyanine Catalyzes Oxidation of Benzene with H2O2 via Benzene Oxide with NIH Shift Evidenced by Using 1, 3, 5-[D3] Benzene as a Probe. Chem. Eur. J. 2008, 14, 7123–7126. [Google Scholar] [PubMed]
  36. Raba, A.; Cokoja, M.; Herrmann, W.A.; Kühn, F.E. Catalytic hydroxylation of benzene and toluene by an iron complex bearing a chelating di-pyridyl-di-NHC ligand. Chem. Commun. 2014, 50, 11454–11457. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  37. Ramu, R.; Wanna, W.H.; Janmanchi, D.; Tsai, Y.-F.; Liu, C.-C.; Mou, C.-Y.; Yu, S.S.-F. Mechanistic study for the selective oxidation of benzene and toluene catalyzed by Fe (ClO4) 2 in an H2O2-H2O-CH3CN system. Mol. Catal. 2017, 441, 114–121. [Google Scholar] [CrossRef]
  38. Yalymov, A.I.; Bilyachenko, A.N.; Levitsky, M.M.; Korlyukov, A.A.; Khrustalev, V.N.; Shul’pina, L.S.; Dorovatovskii, P.V.; Es’kova, M.A.; Lamaty, F.; Bantreil, X. High catalytic activity of heterometallic (Fe6Na7 and Fe6Na6) cage silsesquioxanes in oxidations with peroxides. Catalysts 2017, 7, 101. [Google Scholar] [CrossRef]
  39. Vinogradov, M.M.; Kozlov, Y.N.; Nesterov, D.S.; Shul’pina, L.S.; Pombeiro, A.J.; Shul’pin, G.B. Oxidation of hydrocarbons with H2O2/O2 catalyzed by osmium complexes containing p-cymene ligands in acetonitrile. Catal. Sci. Technol. 2014, 4, 3214–3226. [Google Scholar] [CrossRef]
  40. Sarma, B.B.; Carmieli, R.; Collauto, A.; Efremenko, I.; Martin, J.M.; Neumann, R. Electron transfer oxidation of benzene and aerobic oxidation to phenol. ACS Catal. 2016, 6, 6403–6407. [Google Scholar] [CrossRef]
  41. Li, X.; Xue, H.; Lin, Q.; Yu, A. Amphiphilic poly (ionic liquid)/Wells–Dawson-type phosphovanadomolybdate ionic composites as efficient and recyclable catalysts for the direct hydroxylation of benzene with H2O2. Appl. Organomet. Chem. 2020, 34, e5606. [Google Scholar] [CrossRef]
  42. Carneiro, L.; Silva, A.R. Selective direct hydroxylation of benzene to phenol with hydrogen peroxide by iron and vanadyl based homogeneous and heterogeneous catalysts. Catal. Sci. Technol. 2016, 6, 8166–8176. [Google Scholar] [CrossRef]
  43. Dong, Y.; Niu, X.; Song, W.; Wang, D.; Chen, L.; Yuan, F.; Zhu, Y. Facile synthesis of vanadium oxide/reduced graphene oxide composite catalysts for enhanced hydroxylation of benzene to phenol. Catalysts 2016, 6, 74. [Google Scholar] [CrossRef]
  44. Shijina, A.V.; Renuka, N.K. Single step conversion of benzene to phenol using hydrogen peroxide over modified V2O5–Al2O3 systems. React. Kinet. Catal. Lett. 2009, 98, 139–147. [Google Scholar] [CrossRef]
  45. Peng, G.; Fu, Z.; Yin, D.; Zhong, S.; Yang, Y.; Yu, N.; Yin, D. A Promising Coupled Process of Pd/γ-Al2O3–NH4VO3 Catalyzing the Hydroxylation of Benzene with Hydrogen Peroxide Produced In Situ by an Anthraquinone Redox Route. Catal. Lett. 2007, 118, 270–274. [Google Scholar] [CrossRef]
  46. Jiang, W.-F.; Wang, W.; Wang, H.-L.; Li, Z.-Q. Photooxidation of benzene to phenol by Al2O3-supported Fe (III)-5-sulfosalicylic acid (ssal) complex. Catal. Lett. 2009, 130, 463–469. [Google Scholar] [CrossRef]
  47. Miyahara, T.; Kanzaki, H.; Hamada, R.; Kuroiwa, S.; Nishiyama, S.; Tsuruya, S. Liquid-phase oxidation of benzene to phenol by CuO–Al2O3 catalysts prepared by co-precipitation method. J. Mol. Catal. A Chem. 2001, 176, 141–150. [Google Scholar] [CrossRef]
  48. Masumoto, Y.-k.; Hamada, R.; Yokota, K.; Nishiyama, S.; Tsuruya, S. Liquid-phase oxidation of benzene to phenol by vanadium catalysts in aqueous solvent with high acetic acid concentration. J. Mol. Catal. A Chem. 2002, 184, 215–222. [Google Scholar] [CrossRef]
  49. Miyake, T.; Hamada, M.; Sasaki, Y.; Oguri, M. Direct synthesis of phenol by hydroxylation of benzene with oxygen and hydrogen. Appl. Catal. A Gen. 1995, 131, 33–42. [Google Scholar] [CrossRef]
  50. Lemke, K.; Ehrich, H.; Lohse, U.; Berndt, H.; Jähnisch, K. Selective hydroxylation of benzene to phenol over supported vanadium oxide catalysts. Appl. Catal. A Gen. 2003, 243, 41–51. [Google Scholar] [CrossRef]
  51. Tanarungsun, G.; Kiatkittipong, W.; Praserthdam, P.; Yamada, H.; Tagawa, T.; Assabumrungrat, S. Ternary metal oxide catalysts for selective oxidation of benzene to phenol. J. Ind. Eng. Chem. 2008, 14, 596–601. [Google Scholar] [CrossRef]
  52. Tanarungsun, G.; Kiatkittipong, W.; Praserthdam, P.; Yamada, H.; Tagawa, T.; Assabumrungrat, S. Hydroxylation of benzene to phenol on Fe/TiO2 catalysts loaded with different types of second metal. Catal. Commun. 2008, 9, 1886–1890. [Google Scholar] [CrossRef]
  53. Lee, C.W.; Lee, W.J.; Park, Y.K.; Park, S.-E. Catalytic hydroxylation of benzene over vanadium-containing molecular sieves. Catal. Today 2000, 61, 137–141. [Google Scholar] [CrossRef]
  54. He, J.; Xu, W.-p.; Evans, D.G.; Duan, X.; Li, C.-y. Role of pore size and surface properties of Ti-MCM-41 catalysts in the hydroxylation of aromatics in the liquid phase. Microporous Mesoporous Mater. 2001, 44, 581–586. [Google Scholar] [CrossRef]
  55. Gu, Y.-Y.; Zhao, X.-H.; Zhang, G.-R.; Ding, H.-M.; Shan, Y.-K. Selective hydroxylation of benzene using dioxygen activated by vanadium–copper oxide catalysts supported on SBA-15. Appl. Catal. A Gen. 2007, 328, 150–155. [Google Scholar] [CrossRef]
  56. Jourshabani, M.; Badiei, A.; Shariatinia, Z.; Lashgari, N.; Mohammadi Ziarani, G. Fe-supported SBA-16 type cagelike mesoporous silica with enhanced catalytic activity for direct hydroxylation of benzene to phenol. Ind. Eng. Chem. Res. 2016, 55, 3900–3908. [Google Scholar] [CrossRef]
  57. Abbo, H.S.; Titinchi, S.J. Di-, tri-and tetra-valent ion-exchanged NaY zeolite: Active heterogeneous catalysts for hydroxylation of benzene and phenol. Appl. Catal. A Gen. 2009, 356, 167–171. [Google Scholar] [CrossRef]
  58. Yang, J.-H.; Sun, G.; Gao, Y.; Zhao, H.; Tang, P.; Tan, J.; Lu, A.-H.; Ma, D. Direct catalytic oxidation of benzene to phenol over metal-free graphene-based catalyst. Energy Environ. Sci. 2013, 6, 793–798. [Google Scholar] [CrossRef]
  59. Hu, L.; Wang, C.; Yue, B.; Chen, X.; He, H. Direct hydroxylation of benzene to phenol using H2O2 as an oxidant over vanadium-containing nitrogen doped mesoporous carbon catalysts. RSC Adv. 2016, 6, 87656–87664. [Google Scholar] [CrossRef]
  60. Hu, L.; Wang, C.; Yue, B.; Chen, X.; He, H. Vanadium-containing mesoporous carbon and mesoporous carbon nanoparticles as catalysts for benzene hydroxylation reaction. Mater. Today Commun. 2017, 11, 61–67. [Google Scholar] [CrossRef]
  61. Pezhman, A.; Badiei, A.; Koolivand, A.; Ziarani, G.M. Direct hydroxylation of benzene to phenol over Fe3O4 supported on nanoporous carbon. Chin. J. Catal. 2011, 32, 258–263. [Google Scholar]
  62. Xu, J.; Jiang, Q.; Chen, T.; Wu, F.; Li, Y.-X. Vanadia supported on mesoporous carbon nitride as a highly efficient catalyst for hydroxylation of benzene to phenol. Catal. Sci. Technol. 2015, 5, 1504–1513. [Google Scholar] [CrossRef]
  63. Wang, C.; Hu, L.; Wang, M.; Yue, B.; He, H. Cerium promoted Vg-C3N4 as highly efficient heterogeneous catalysts for the direct benzene hydroxylation. R. Soc. Open Sci. 2018, 5, 180371. [Google Scholar] [CrossRef] [Green Version]
  64. Cao, T.; Cai, M.; Jin, L.; Wang, X.; Yu, J.; Chen, Y.; Dai, L. Amorphous Cr-doped gC 3 N 4 as an efficient catalyst for the direct hydroxylation of benzene to phenol. New J. Chem. 2019, 43, 16169–16175. [Google Scholar] [CrossRef]
  65. Yu, Z.-H.; Gan, Y.-L.; Xu, J.; Xue, B. Direct Catalytic Hydroxylation of Benzene to Phenol Catalyzed by FeCl 3 Supported on Exfoliated Graphitic Carbon Nitride. Catal. Lett. 2020, 150, 301–311. [Google Scholar] [CrossRef]
  66. Tu, T.N.; Nguyen, H.T.; Nguyen, H.T.; Nguyen, M.V.; Nguyen, T.D.; Tran, N.T.; Lim, K.T. A new iron-based metal–organic framework with enhancing catalysis activity for benzene hydroxylation. RSC Adv. 2019, 9, 16784–16789. [Google Scholar] [CrossRef] [Green Version]
  67. Zhang, T.; Zhang, D.; Han, X.; Dong, T.; Guo, X.; Song, C.; Si, R.; Liu, W.; Liu, Y.; Zhao, Z. Preassembly strategy to single Cu-N3 sites inlaid porous hollow carbonitride spheres for selective oxidation of benzene to phenol. J. Am. Chem. Soc. 2018, 140, 16936–16940. [Google Scholar] [CrossRef]
  68. Ito, S.; Mitarai, A.; Hikino, K.; Hirama, M.; Sasaki, K. Deactivation reaction in the hydroxylation of benzene with Fenton’s reagent. J. Org. Chem. 1992, 57, 6937–6941. [Google Scholar] [CrossRef]
  69. Bal, R.; Tada, M.; Sasaki, T.; Iwasawa, Y. Direct phenol synthesis by selective oxidation of benzene with molecular oxygen on an interstitial-N/Re cluster/zeolite catalyst. Angew. Chem. Int. Ed. 2006, 118, 462–466. [Google Scholar] [CrossRef]
  70. Xia, H.; Sun, K.; Sun, K.; Feng, Z.; Li, W.X.; Li, C. Direct spectroscopic observation of Fe (III)—Phenolate complex formed from the reaction of benzene with peroxide species on Fe/ZSM-5 at room temperature. J. Phys. Chem. C 2008, 112, 9001–9005. [Google Scholar] [CrossRef]
  71. Borah, P.; Ma, X.; Nguyen, K.T.; Zhao, Y. A vanadyl complex grafted to periodic mesoporous organosilica: A green catalyst for selective hydroxylation of benzene to phenol. Angew. Chem. Int. Ed. 2012, 124, 7876–7881. [Google Scholar] [CrossRef]
  72. Chen, P.; Chen, L.; Zeng, Y.; Ding, F.; Jiang, X.; Liu, N.; Au, C.-T.; Yin, S.-F. Three-dimension hierarchical heterostructure of CdWO4 microrods decorated with Bi2WO6 nanoplates for high-selectivity photocatalytic benzene hydroxylation to phenol. Appl. Catal. B Environ. 2018, 234, 311–317. [Google Scholar] [CrossRef]
  73. Chen, X.; Zhang, J.; Fu, X.; Antonietti, M.; Wang, X. Fe-g-C3N4-catalyzed oxidation of benzene to phenol using hydrogen peroxide and visible light. J. Am. Chem. Soc. 2009, 131, 11658–11659. [Google Scholar] [CrossRef]
  74. Sannino, D.; Vaiano, V.; Isupova, L.A.; Ciambelli, P. Photo-Fenton oxidation of acetic acid on supported LaFeO3 and Pt/LaFeO3 perovskites. Chem. Eng. Trans. 2011, 25, 1013–1018. [Google Scholar] [CrossRef]
  75. Zhang, P.; Gong, Y.; Li, H.; Chen, Z.; Wang, Y. Selective oxidation of benzene to phenol by FeCl3/mpg-C3N4 hybrids. RSC Adv. 2013, 3, 5121–5126. [Google Scholar] [CrossRef]
  76. Ohkubo, K.; Kobayashi, T.; Fukuzumi, S. Direct oxygenation of benzene to phenol using quinolinium ions as homogeneous photocatalysts. Angew. Chem. Int. Ed. 2011, 123, 8811–8814. [Google Scholar] [CrossRef]
  77. Gu, Y.; Li, Q.; Zang, D.; Huang, Y.; Yu, H.; Wei, Y. Light-Induced Efficient Hydroxylation of Benzene to Phenol by Quinolinium and Polyoxovanadate-Based Supramolecular Catalysts. Angew. Chem. Int. Ed. 2020. [Google Scholar] [CrossRef]
  78. Ciambelli, P.; Sannino, D.; Palma, V.; Vaiano, V.; Eloy, P.; Dury, F.; Gaigneaux, E.M. Tuning the selectivity of MoOx supported catalysts for cyclohexane photo oxidehydrogenation. Catal. Today 2007, 128, 251–257. [Google Scholar] [CrossRef]
  79. Sannino, D.; Vaiano, V.; Ciambelli, P.; Carotenuto, G.; Di Serio, M.; Santacesaria, E. Enhanced performances of grafted VOx on titania/silica for the selective photocatalytic oxidation of ethanol to acetaldehyde. Catal. Today 2013, 209, 159–163. [Google Scholar] [CrossRef]
  80. Sannino, D.; Vaiano, V.; Ciambelli, P.; Hidalgo, M.C.; Murcia, J.J.; Navío, J.A. Oxidative dehydrogenation of ethanol over Au/TiO2 photocatalysts. J. Adv. Oxid. Technol. 2012, 15, 284–293. [Google Scholar] [CrossRef]
  81. Vaiano, V.; Sarno, G.; Sacco, O.; Sannino, D. Degradation of terephthalic acid in a photocatalytic system able to work also at high pressure. Chem. Eng. J. 2017, 312, 10–19. [Google Scholar] [CrossRef]
  82. Fessi, N.; Nsib, M.F.; Cardenas, L.; Guillard, C.; Dappozze, F.; Houas, A.; Parrino, F.; Palmisano, L.; Ledoux, G.; Amans, D.; et al. Surface and Electronic Features of Fluorinated TiO2 and Their Influence on the Photocatalytic Degradation of 1-Methylnaphthalene. J. Phys. Chem. C 2020, 124, 11456–11468. [Google Scholar] [CrossRef]
  83. Žerjav, G.; Scandura, G.; Garlisi, C.; Palmisano, G.; Pintar, A. Sputtered vs. sol-gel TiO2-doped films: Characterization and assessment of aqueous bisphenol A oxidation under UV and visible light radiation. Catal. Today 2020, 357, 380–391. [Google Scholar] [CrossRef]
  84. Sclafani, A.; Herrmann, J. Comparison of the photoelectronic and photocatalytic activities of various anatase and rutile forms of titania in pure liquid organic phases and in aqueous solutions. J. Phys. Chem. 1996, 100, 13655–13661. [Google Scholar] [CrossRef]
  85. Vaiano, V.; Sacco, O.; Sannino, D.; Stoller, M.; Ciambelli, P.; Chianese, A. Photocatalytic removal of phenol by ferromagnetic N-TiO2/SiO2/Fe3O4 nanoparticles in presence of visible light irradiation. Chem. Eng. Trans. 2016, 47, 235–240. [Google Scholar] [CrossRef]
  86. Devaraji, P.; Jo, W.-K. Noble metal free Fe and Cr dual-doped nanocrystalline titania (Ti1− x− yMx+ yO2) for high selective photocatalytic conversion of benzene to phenol at ambient temperature. Appl. Catal. A Gen. 2018, 565, 1–12. [Google Scholar] [CrossRef]
  87. Gupta, N.; Bansal, P.; Pal, B. Metal ion-TiO2 nanocomposites for the selective photooxidation of benzene to phenol and cycloalkanol to cycloalkanone. J. Exp. Nanosci. 2015, 10, 148–160. [Google Scholar] [CrossRef] [Green Version]
  88. Shiraishi, Y.; Saito, N.; Hirai, T. Adsorption-driven photocatalytic activity of mesoporous titanium dioxide. J. Am. Chem. Soc. 2005, 127, 12820–12822. [Google Scholar] [CrossRef]
  89. Yuzawa, H.; Aoki, M.; Otake, K.; Hattori, T.; Itoh, H.; Yoshida, H. Reaction mechanism of aromatic ring hydroxylation by water over platinum-loaded titanium oxide photocatalyst. J. Phys. Chem. C 2012, 116, 25376–25387. [Google Scholar] [CrossRef]
  90. Ide, Y.; Nakamura, N.; Hattori, H.; Ogino, R.; Ogawa, M.; Sadakane, M.; Sano, T. Sunlight-induced efficient and selective photocatalytic benzene oxidation on TiO2-supported gold nanoparticles under CO2 atmosphere. Chem. Commun. 2011, 47, 11531–11533. [Google Scholar] [CrossRef]
  91. 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]
  92. Tanarungsun, G.; Kiatkittipong, W.; Assabumrungrat, S.; Yamada, H.; Tagawa, T.; Praserthdam, P. Multi transition metal catalysts supported on TiO2 for hydroxylation of benzene to phenol with hydrogen peroxide. J. Ind. Eng. Chem. 2007, 13, 870–877. [Google Scholar]
  93. Devaraji, P.; Jo, W.-K. Natural leaf-assisted dual-phase two-dimensional leaf TiO2 and Cu(OH)2 co-catalyst for photocatalytic conversion of benzene to phenol. Mater. Res. Bull. 2019, 110, 67–75. [Google Scholar] [CrossRef]
  94. Zhang, Y.; Park, S.-J. Stabilizing CuPd bimetallic alloy nanoparticles deposited on holey carbon nitride for selective hydroxylation of benzene to phenol. J. Catal. 2019, 379, 154–163. [Google Scholar] [CrossRef]
  95. Verma, S.; Nasir Baig, R.; Nadagouda, M.N.; Varma, R.S. Hydroxylation of Benzene via C–H activation using bimetallic CuAg@ g-C3N4. ACS Sustain. Chem. Eng. 2017, 5, 3637–3640. [Google Scholar] [CrossRef] [PubMed]
  96. Shiravand, G.; Badiei, A.; Ziarani, G.M.; Jafarabadi, M.; Hamzehloo, M. Photocatalytic synthesis of phenol by direct hydroxylation of benzene by a modified nanoporous silica (LUS-1) under sunlight. Chin. J. Catal. 2012, 33, 1347–1353. [Google Scholar] [CrossRef]
  97. Ye, X.; Cui, Y.; Qiu, X.; Wang, X. Selective oxidation of benzene to phenol by Fe-CN/TS-1 catalysts under visible light irradiation. Appl. Catal. B Environ. 2014, 152, 383–389. [Google Scholar] [CrossRef]
  98. Wang, D.; Wang, M.; Li, Z. Fe-based metal–organic frameworks for highly selective photocatalytic benzene hydroxylation to phenol. ACS Catal. 2015, 5, 6852–6857. [Google Scholar] [CrossRef]
  99. Dasireddy, V.D.; Likozar, B. Selective photocatalytic oxidation of benzene to phenol using carbon nanotube (CNT)-supported Cu and TiO2 heterogeneous catalysts. J. Taiwan Inst. Chem. Eng. 2018, 82, 331–341. [Google Scholar] [CrossRef]
  100. Li, J.; Xu, Y.; Ding, Z.; Mahadi, A.H.; Zhao, Y.; Song, Y.-F. Photocatalytic selective oxidation of benzene to phenol in water over layered double hydroxide: A thermodynamic and kinetic perspective. Chem. Eng. J. 2020, 388, 124248. [Google Scholar] [CrossRef]
  101. Bui, T.D.; Kimura, A.; Higashida, S.; Ikeda, S.; Matsumura, M. Two routes for mineralizing benzene by TiO2-photocatalyzed reaction. Appl. Catal. B Environ. 2011, 107, 119–127. [Google Scholar] [CrossRef]
  102. Devaraji, P.; Sathu, N.K.; Gopinath, C.S. Ambient oxidation of benzene to phenol by photocatalysis on Au/Ti0.98V0.02O2: Role of holes. ACS Catal. 2014, 4, 2844–2853. [Google Scholar] [CrossRef]
  103. Einaga, H.; Ibusuki, T.; Futamura, S. Improvement of catalyst durability by deposition of Rh on TiO2 in photooxidation of aromatic compounds. Environ. Sci. Technol. 2004, 38, 285–289. [Google Scholar] [CrossRef]
  104. Nishikawa, M.; Shiroishi, W.; Honghao, H.; Suizu, H.; Nagai, H.; Saito, N. Probability of Two-Step Photoexcitation of Electron from Valence Band to Conduction Band through Doping Level in TiO2. J. Phys. Chem. A 2017, 121, 5991–5997. [Google Scholar] [CrossRef] [PubMed]
  105. Choi, W.; Termin, A.; Hoffmann, M.R. The role of metal ion dopants in quantum-sized TiO2: Correlation between photoreactivity and charge carrier recombination dynamics. J. Phys. Chem. 2002, 98, 13669–13679. [Google Scholar] [CrossRef]
  106. Kunai, A.; Hata, S.; Ito, S.; Sasaki, K. The role of oxygen in the hydroxylation reaction of benzene with Fenton’s reagent. Oxygen 18 tracer study. J. Am. Chem. Soc. 1986, 108, 6012–6016. [Google Scholar] [CrossRef] [PubMed]
  107. 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]
Figure 1. The reaction steps for phenol production via the cumene process.
Figure 1. The reaction steps for phenol production via the cumene process.
Catalysts 10 01424 g001
Figure 2. One-step catalytic oxidation of benzene to phenol using different oxidants.
Figure 2. One-step catalytic oxidation of benzene to phenol using different oxidants.
Catalysts 10 01424 g002
Figure 3. Oxidation of benzene to phenol using POM in the presence of H2SO4 aqueous solution and O2 as oxidant. Adapted with permission from [40]. Copyright (2016) American Chemical Society.
Figure 3. Oxidation of benzene to phenol using POM in the presence of H2SO4 aqueous solution and O2 as oxidant. Adapted with permission from [40]. Copyright (2016) American Chemical Society.
Catalysts 10 01424 g003
Figure 4. Proposed reaction mechanism for phenol production in the presence of Fe/SBA-16. Reprinted with permission from [56]. Copyright (2016) American Chemical Society.
Figure 4. Proposed reaction mechanism for phenol production in the presence of Fe/SBA-16. Reprinted with permission from [56]. Copyright (2016) American Chemical Society.
Catalysts 10 01424 g004
Figure 5. Mechanism of catalytic benzene hydroxylation by H2O2 in the presence of Fe-TBAPy [66]. Published by The Royal Society of Chemistry.
Figure 5. Mechanism of catalytic benzene hydroxylation by H2O2 in the presence of Fe-TBAPy [66]. Published by The Royal Society of Chemistry.
Catalysts 10 01424 g005
Figure 6. Chemical reactions involving aromatic ring species under light irradiation using Pt/TiO2. Reprinted with permission from [89]. Copyright (2012) American Chemical Society.
Figure 6. Chemical reactions involving aromatic ring species under light irradiation using Pt/TiO2. Reprinted with permission from [89]. Copyright (2012) American Chemical Society.
Catalysts 10 01424 g006
Figure 7. Oxidation reaction of benzene to carbon dioxide and water under light irradiation. Reprinted with permission from [89]. Copyright (2012) American Chemical Society.
Figure 7. Oxidation reaction of benzene to carbon dioxide and water under light irradiation. Reprinted with permission from [89]. Copyright (2012) American Chemical Society.
Catalysts 10 01424 g007
Figure 8. Formation of the hydroxylated benzene radical and next deprotonation with phenol generation [91].
Figure 8. Formation of the hydroxylated benzene radical and next deprotonation with phenol generation [91].
Catalysts 10 01424 g008
Figure 9. Possible mechanism of photocatalytic benzene oxidation to phenol using Fe-Cr-doped TiO2. Reprinted from [86], Copyright (2018), with permission from Elsevier.
Figure 9. Possible mechanism of photocatalytic benzene oxidation to phenol using Fe-Cr-doped TiO2. Reprinted from [86], Copyright (2018), with permission from Elsevier.
Catalysts 10 01424 g009
Figure 10. Possible reaction mechanism for the catalytic oxidation of benzene by FeCl3/mpg-C3N4. Reproduced from [75] with permission from The Royal Society of Chemistry.
Figure 10. Possible reaction mechanism for the catalytic oxidation of benzene by FeCl3/mpg-C3N4. Reproduced from [75] with permission from The Royal Society of Chemistry.
Catalysts 10 01424 g010
Table 1. Standard heat of reaction (∆H°) and standard Gibbs free energy (∆G°) for benzene, phenol, and for the different oxidant species.
Table 1. Standard heat of reaction (∆H°) and standard Gibbs free energy (∆G°) for benzene, phenol, and for the different oxidant species.
ComponentH° at 298 K (kJ/mol)G° at 298 K (kJ/mol)
ReagentBenzene (l)48.99464124.34848
Benzene (g)82.92688129.66216
OxidantH2O2 (l)−136.10552−105.47864
N2O (g)82.04824104.1816
O2(g)
H2(g)
ProductPhenol (s)−165.01696−50.4172
Phenol (g)−96.35752−32.88624
H2O−285.82996−237.178408
Table 2. Main catalytic formulations for the one-step catalytic oxidation of benzene to phenol in liquid phase.
Table 2. Main catalytic formulations for the one-step catalytic oxidation of benzene to phenol in liquid phase.
Catalystt (h)T (°C)P (atm)Operating ConditionsBenzene Conversion (%) XPhenol Yield (%) ηSelectivity to Phenol (%) SpRef.
CuO/Al2O3-80180 vol% acetic acid,
benzene: 22.5 mmol;
ascorbic acid: 4 mmol.
-1.2-[47]
V/Al2O3-30480 vol% acetic acid;
benzene: 5.6 mmol;
ascorbic acid: 1 mmol.
-8.4-[48]
V2O5–Al2O36601Catalyst: 0.2 g (14 wt%V2O5); benzene: 1.46 mmol;
acetonitrile: 4 mL; H2O2: 11.68 mmol.
13-100[44]
Fe3+–Al2O36601Catalyst: 0.20 g; acetonitrile: 4 mL; benzene: 1.24 mmol;
H2O2: 6 mmol.
1212-[21]
Ru/SiO2
Rh/SiO2
Pd/SiO2
Ir/SiO2
Pt/SiO2
-201Catalyst: 0.5 wt.-% metal/SiO2:1.0 g;
H2/O2 = 3;
benzene: 20 mL;
acetic acid: 25 mL.
--0
99.7
88.2
64.5
63.9
[49]
Ru/SiO2
Rh/SiO2
Pd/SiO2
Ir/SiO2
Pt/SiO2
-601Catalyst: 0.5 wt.-%; metal: 20wt.-%; V2O5/SiO2: 1.0 g; H2/O2: 3;
benzene: 20 mL;
acetic acid: 25 mL.
--100
100
99.7
100
100
[49]
0.1%V/SiO2-701Catalyst: 0.204 g;
benzene: 40 mmol
benzene/H2O2
mole ratio: 1; acetonitrile: - mL.
10-81[50]
Fe5V2.5Cu2.5/TiO24301Catalyst: 0.2 g;
benzene: 11 mL;
benzene/H2O2
mole ratio: 0.5; acetonitrile: 40 mL.
9.87.15473[51]
FePt/TiO2
(5%;1%)
4301Catalyst: 0.2 g;
benzene: 11 mL;
benzene/H2O2
mole ratio: 0.5; acetonitrile: 40 mL.
6.55.9291[52]
V/MCM-41
[Si/V = 1/9.4]
6601Catalyst: 0.05 g;
benzene: 6 mL;
benzene/H2O2
mole ratio: 1/1.15; acid acetic: 6 mL.
1.4-93[53]
4%Cu/MCM-411.6301Catalyst: 0.05 g;
benzene: 1 mL;
benzene/H2O2
mole ratio: 1/2; acid acetic: 7.5 mL.
2119.794[18]
Ti-MCM-41
[Si/Ti = 25]
3.5651Catalyst: 0.05 g;
benzene: 0.045 mol;
benzene/H2O2
mole ratio: 1/3; acetone: 15 g.
98->95[54]
VOx/FeSBA-15
VOx/CoSBA-15
VOx/NiSBA-15
VOx/CrSBA-15
VOx/MnSBA-15
VOx/ZnSBA-15
VOx/AgSBA-15
VOx/CuSBA-15
5801Catalyst: 0.05 g;
benzene: 1mL; solvent (acetic acid/H2O v/v): 36 mL;
ascorbic acid: 11.9 mmol.
-12.8
11.3
15.8
10.2
17.2
17.9
18.1
24.7
-[55]
Fe/SBA-168651Catalyst: 0.1 g;
benzene: 1 mL;
H2O2: 2 mL;
acetonitrile: 20 mL;
12.111.796.4[56]
1.4wt%Cu(II)-NaY6701Catalyst: 0.025 g;
benzene: 0.02 mol;
H2O2: 0.02 mol.
33.2-100[57]
Graphene (CCG)16601Catalyst: 0.02 g;
benzene: 130 mg;
H2O2: 2.4 mL;
acetonitrile: 1.2 mL.
17.817> 99[58]
4.2V/NC-6003701Catalyst: 0.02 g;
benzene: 0.4 mL;
H2O2: 1.4 mL;
acetic acid: 5 mL.
27.726.896.7[59]
4V/MCN-S3701Catalyst: 0.02 g;
benzene: 0.4 mL;
H2O2: 1.4 mL;
acetic acid: 5 mL.
38.2 36.7 96.1[60]
Fe3O4/CMK-34601Catalyst: 0.02 g;
benzene: 1 mL;
H2O2: 2 mL;
acetonitrile: 6 mL.
18-92[61]
10V/mp-C3N43601Catalyst: 0.06 g;
benzene: 1.5 mL;
H2O2: 3 mL;
acetonitrile: 6 mL.
181893[62]
Ce0.07-0.07V-g-C3N4470 Catalyst: 0.04 g;
benzene: 1 mL;
H2O2: 3.5 mL;
acid acetic: 10 mL.
33.732.395.9[63]
Cr/g-C3N4-300765 Catalyst: 0.04 g;
benzene: 3.36 mmol;
H2O2: 1.2 mL;
acetonitrile: 2 mL.
31.130.999.5[64]
FeCl3/eg-C3N43601Catalyst: 0.05 g;
benzene: 11.2 mmol;
H2O2: 3 mL;
acetonitrile: 5 mL.
222299[65]
Fe-TBAPy3601Catalyst: 0.05 g;
benzene: 11.2 mmol;
H2O2: 3 mL;
acetonitrile: 5 mL.
-64.592.9[66]
Cu-SA/HCNS1260 Catalyst: 0.05 g;
benzene: 0.4 mL;
H2O2: 6 mL;
acetonitrile: 6 mL.
86-96.7[67]
Table 3. Main photocatalytic formulations for the one-step catalytic oxidation of benzene to phenol in liquid phase.
Table 3. Main photocatalytic formulations for the one-step catalytic oxidation of benzene to phenol in liquid phase.
Photocatalystt * (h)Light SourceOperating ConditionsBenzene Conversion (%) XPhenol Yield (%) ηSelectivity to Phenol (%) SpRef.
nTiO2
mTiO2
mTiO2
2
2
6
Hg lamp
λ > 320 nm
Photocatalyst: 10 mg + nitrogen flow
H2O: 10 mL
benzene: 20 μmol
pH 7
26
23
42
2
19
34
8
83
81
[88]
TiO26450 W Xe arc lampPhotocatalyst:25 mg
Benzene: 20 mM
[Fe3+]: 1.47 mM
[Ag+ ]: 0.98 mM
H2O2: 9.4 mM
-<196[9]
Pt-TiO21.5λ > 385 nmPhotocatalyst: 0.2 g
benzene: 0.05 mL
H2O: 4 mL
-2.191[89]
Au-P25:
in 100 kPa air
in 230 kPa
CO2
P25:
in 100 kPa air
in 230 kPa CO2
24Solar simulatorPhotocatalyst: 60 mg
aqueous benzene solution: 20 mL
C0benzene: 600 ppm
dry ice:0-200 mg closed container: 50 mL
13
14
34
31
8
13
7
7
62
89
21
22
[90]
Au-V-TiO218400 W Hg lamp
λ = 200−400 nm
Photocatalyst: 30 mg
CH3CN: 2 mL
benzene: 1 mL (25 wt%) H2O2: 2 mL
181688[86]
Pt/WO3-Ka 1
b 4
e 0.25
300 W Xe lamp
λ>300 nm
c λ>400 nm
Photocatalyst: 50 mg
C0benzene: 2.5 mmolL−1
H2O: 7.5 mL
279 K
O2
dAr
{ 22.2   a 68.9   b { 79.3     a 73.7   b [91]
Pt/WO3-K { 26.6   a ,   c 52.5   b ,   c { 83.8   a ,   c 75.1   b ,   c
WO3-K 16.4 b 84.6 b
Pt/WO3-Y 40.6a 58.8 a
Pt/WO3-S 32.4 a 48.7 a
Pt/TiO2-P25 { 38.0   a 59.1   b { 25.9   a 21.8   b
{ 13.3   a , d 33.8   b , d { 60.8   a , d 34   b , d
TiO2-P25 85.2 b 20.6 b
Pt/TiO2-M 43 a 31 a
Pt/TiO2-J. { 48.5   a 11.5   d , e 38.6   a , d { 26.5   a 63   d , e 35.4   a , d
Fe3+ impregnated TiO21–2125 W Hg lamp
UV light
Photocatalyst:
50 mg
aqueous benzene
(1 to 20 mM): 5 mL
-9–1580–86[87]
Fe-Cr-TiO212450 W mercury lamp
λ = 200–400
nm
Photocatalyst:
30 mg
CH3CN: 2 mL
benzene: 1 mL
(25 wt%) H2O2: 2 mL
28 ± 0.525.2 ± 0.590 ± 0.5[86]
Fe-V-Cu
supported on TiO2
4black light blue fluorescent bulb (8W)Photocatalyst: 0.2 g
benzene: 11 cm3
benzene/H2O2 mole ratio: 0.5
(30 wt%) H2O2: 30 cm3
solvent: 40 cm3 acetone a, acetonitrile b,
pyridine c
ascorbic acid: 0.5
18.61 a
14.27 b
7.9 c
9.68 a
9.7 b
7.11 c
52 a
68 b
90 c
[92]
LT-550
LT-750
Cu(OH)2/LT-550
Cu(OH)2/LT-750
Cu(OH)2/LT-750a
Cu(OH)2/LT-750b
6UV lightPhotocatalyst: 5 mg
Benzene: 100 μL
CH3CN: 500 μL
H2O: 13 mL
(30 wt%) H2O2: 87 μL
38.7
47.1
42
49.9
a 55.0
b 41
36.3
45.2
40.7
48.4
a 47.9
b 36.5
94
96
97
97
a 87
b 89
[93]
CuPd/g-C3N41.5solar simulatorSolution A:
-photocatalyst:
20 mg
- deionized water: 30 mL
Solution B:
- benzene: 0.5 mL
- acetonitrile:
30 mL.
(30 wt%) H2O2:
5 mmol added to the two mixed solutions.
98.187.889.6[94]
Fe2O3/g-C3N4
Pd/g-C3N4
Cu/g-C3N4
Ni/g-C3N4
Ag/g-C3N4
FePd/g-C3N4
FeCu/g-C3N4
FeAg/g-C3N4
FeNi/g-C3N4
PdCu/g-C3N4
PdNi/g-C3N4
PdAg/g-C3N4
CuNi/g-C3N4
CuAg/g-C3N4
CuAg/g-C3N4a
CuAg/g-C3N4b
CuAg/g-C3N4c
CuAg/g-C3N4d
CuAg/g-C3N4e
CuAg/g-C3N4f
12
12
12
12
12
12
12
12
12
12
12
12
12
0.5
0.5
0.5
3
0.5
0.5
0.5
Visible light
20 W
domestic bulb
Photocatalyst:
100 mg
Benzene:1 mmol
CH3CN: 5.0 mL (30 wt%) H2O2:
1.1 mmol
a50 mg of catalyst
b 25 mg of catalyst
c 15 mg of catalyst
d methanol as a solvent
e water as a solvent
f ethanol as a solvent
15
43
39
20
32
70
67
41
29
81
72
77
57
99
99 a
99 b
99 c
86 d
83 e
99 f
[95]
mpg-C3N4
3%FeCl3/mpg-C3N4
5%FeCl3/mpg-C3N4 10%FeCl3/mpg-C3N4
20%FeCl3/mpg-C3N4
5%FeCl3/mpg-C3N4a
5%FeCl3/mpg-C3N4b
5%FeCl3/mpg-C3N4c
5%FeCl3/mpg-C3N4d
5%FeCl3/mpg-C3N4e
4100 W mercury lamp
λ > 420nm
Photocatalyst: 25 mg
benzene: 4.5 mmol
(30 wt%) H2O2: 0.255 mL
60 °C
a T = 25 °C
b T = 40 °C
c T = 80 °C
d H2O2: 0.510 mL
e H2O2: 0.765 mL
2
17
38
23
25
4 a
10 b
21 c
44 d
47 e
95
98
97
94
80
99 a
96 b
81 c
85 d
60 e
[75]
g-C3N4
mpg-C3N4
FeCl3
5%Fe-g-C3N4
10%Fe-g-C3N4
20%Fe-g-C3N4Cu-g-C3N4
Ti-g-C3N4
Ni-g-C3N4
Zn-g-C3N4
Fe/SBA-15
g-C3N4/SBA-15
Fe-g-C3N4/SBA-15
4500 W Xenon lamp
λ > 420 nm
Photocatalyst:
50 mg
CH3CN: 4 mL
benzene: 0.8 mL
H2O: 4 mL
(30 wt%)H2O2:
0.51 mL
0
2.0
0.5
1.8
4.8
2.5
1.4
0.1
0.1
0.1
1.0
0.1
11.9
[73]
10%Fe-g-C3N4
20%Fe-g-C3N4
30%Fe-g-C3N4
10%Fe-g-C3N4-LUS-1
20%Fe-g-C3N4-LUS-1
4sunlightPhotocatalyst: 0.05 g
benzene: 1 mL
CH3CN: 4 mL
H2O2: 0.5 mL
T = 60 °C
6.5
8
10.5
10
16
>90
~90
~ 90
>90
>90
[96]
Fe-CN
TS-1
Fe-CN/TS-1–1 a
Fe-CN/TS-1–2 b
Fe-CN/TS-1–3 c
Fe-CN/TS-1–4 d
Fe-CN/TS-1–5 e
Fe-CN/TS-1–6 f
Fe-CN/TS-1–2 g
Fe-CN/TS-1–7 h
Fe-CN/TS-1–8 i
Fe/TS-1
4300 W Xenon lamp
λ> 420 nm
CH3CN: 4 mL
benzene: 0.8 mL
H2O: 4 mL
(30 wt%) H2O2:
0.51 mL
60 °C pH = 7
Fe-CN/TS-1-X
a X = 1 for 10% dicyandiamide/TS-1
b X = 2 for 20% dicyandiamide/TS-1
c X = 3 for 50% dicyandiamide/TS-1
d X = 4 for 100% dicyandiamide/TS-1
e X = 5 for 200% dicyandiamide/TS-1
Fe-CN/TS-1-X
f X = 6 for 5% FeCl3/dicyandiamide
g X = 2 for 10% FeCl3/dicyandiamide
h X = 7 for 20% FeCl3/dicyandiamide
i X = 8 for 50% FeCl3/dicyandiamide
1.1
2.4
2.8 a
10 b
8.8 c
1.3 d
0.1 e
1.4 f
10 g
5 h
1.6 i
7.6
[97]
MIL-100(Fe)
MIL-68(Fe)i
8Visible light irradiation
λ≥ 420 nm
Photocatalyst: 10 mg
H2O2: 0.5 mmol
Solvent: 4 mL
a CH3CN solvent
H2O2:benzene(1:2)
b Acetone solvent
H2O2:benzene(1:2)
c H2O solvent
H2O2:benzene(1:2)
d DMF solvent
H2O2:benzene(1:2)
e CH3CN:H2O (1:1)
H2O2:benzene(1:2)
f CH3CN:H2O (1:1)
H2O2:benzene(3:4)
g CH3CN:H2O (1:1)
H2O2:benzene(2:2)
h CH3CN:H2O (1:1)
H2O2:benzene(3:2)
i CH3CN:H2O (1:1)
H2O2:benzene(3:4)
10.3 a
2.4 b
8.3 c
3.3 d
13.6 e
20.1 f
21.7 g
22.5 h
14i
10.3 a
2.38 b
7.1 c
2.5 d
13.3 e
14.77 f
20.8 g
31.05 h
9.45 i
>99 a
99 b
85 c
76 d
98 e
98 f
96 g
92 h
90 i
[98]
Ti/CNT
Cu/Ti/CNT
0.75Low-pressure mercury lampPhotocatalyst:
100 mg
benzene: 20 mL
H2O: 20 mL
53.8
68.3
35.1
51.8
65.3
75.8
[99]
Zn-Ti-LDH3300 W Xenon lampPhotocatalyst: 20 mg
Benzene: 0.2 mmol
H2O: 20 mL
5.654.5987.18[100]
Bi2WO6/CdWO4 composite3300 W Xe lamp
λ ≥4 00nm
Photocatalyst:
50 mg
benzene:
0.5 mmol
CH3CN: 3 mL
H2O: 100 μL
O2: 3 mL min−1
5.8>99[72]
QuCN+ ion1500 W xenon lamp
λ = 290–600
nm
[QuCN+]: 2.0 mM
[C6H6]: 30 mM
[H2O]: 3.0 M
313098[76]
* t = irradiation time, min.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mancuso, A.; Sacco, O.; Sannino, D.; Venditto, V.; Vaiano, V. One-Step Catalytic or Photocatalytic Oxidation of Benzene to Phenol: Possible Alternative Routes for Phenol Synthesis? Catalysts 2020, 10, 1424. https://doi.org/10.3390/catal10121424

AMA Style

Mancuso A, Sacco O, Sannino D, Venditto V, Vaiano V. One-Step Catalytic or Photocatalytic Oxidation of Benzene to Phenol: Possible Alternative Routes for Phenol Synthesis? Catalysts. 2020; 10(12):1424. https://doi.org/10.3390/catal10121424

Chicago/Turabian Style

Mancuso, Antonietta, Olga Sacco, Diana Sannino, Vincenzo Venditto, and Vincenzo Vaiano. 2020. "One-Step Catalytic or Photocatalytic Oxidation of Benzene to Phenol: Possible Alternative Routes for Phenol Synthesis?" Catalysts 10, no. 12: 1424. https://doi.org/10.3390/catal10121424

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