Next Article in Journal
Anti-Adipogenic Effects of Salicortin from the Twigs of Weeping Willow (Salix pseudolasiogyne) in 3T3-L1 Cells
Previous Article in Journal
Utilization of Agro-Industrial Residues in the Rearing and Nutritional Enrichment of Zophobas atratus Larvae: New Food Raw Materials
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Highly Dispersed Vanadia Anchored on Protonated g-C3N4 as an Efficient and Selective Catalyst for the Hydroxylation of Benzene into Phenol

1
College of Materials & Environmental Engineering, Hangzhou Dianzi University, Hangzhou 310036, China
2
Key Lab of Applied Chemistry of Zhejiang Province, Department of Chemistry, Zhejiang University, Hangzhou 310027, China
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(20), 6965; https://doi.org/10.3390/molecules27206965
Submission received: 9 September 2022 / Revised: 9 October 2022 / Accepted: 10 October 2022 / Published: 17 October 2022

Abstract

:
The direct hydroxylation of benzene is a green and economical-efficient alternative to the existing cumene process for phenol production. However, the undesired phenol selectivity at high benzene conversion hinders its wide application. Here, we develop a one-pot synthesis of protonated g-C3N4 supporting vanadia catalysts (V-pg-C3N4) for the efficient and selective hydroxylation of benzene. Characterizations suggest that protonating g-C3N4 in diluted HCl can boost the generation of amino groups (NH/NH2) without changing the bulk structure. The content of surface amino groups, which determines the dispersion of vanadia, can be easily regulated by the amount of HCl added in the preparation. Increasing the content of surface amino groups benefits the dispersion of vanadia, which eventually leads to improved H2O2 activation and benzene hydroxylation. The optimal catalyst, V-pg-C3N4-0.46, achieves 60% benzene conversion and 99.7% phenol selectivity at 60 oC with H2O2 as the oxidant.

Graphical Abstract

1. Introduction

Phenol, one of the most valuable organic intermediates for fine chemical manufacturing [1,2], is produced industrially by a three-step cumene process from benzene [3]. Unfortunately, this process suffers from high energy consumption, relatively low yield, and large amount of acetone as by-products [4,5]. To address these problems, extensive efforts have been devoted to developing an economical efficient and green approach for phenol production. Among various methods, the direct hydroxylation of benzene to phenol using H2O2 as an oxidant has been recognized as a promising alternative as it can be operated under mild conditions and only produces water as a co-product [4,6]. However, because phenol is more reactive to oxidation than benzene, obtaining high selectivity to phenol at high benzene conversion is difficult [7,8]. To this end, a key topic for the catalytic hydroxylation of benzene is the development of advanced catalysts to achieve high activity and benzene selectivity simultaneously.
Recently, graphitic carbon nitride (g-C3N4) has emerged as a fascinating material to load active metals for benzene hydroxylation [9,10,11,12]. As an analog of graphite, g-C3N4 possesses a stacked 2D structure with π-conjugated planar layers, which endow it with high capability in activating aromatic molecules (e.g., benzene) [13]. Meanwhile, the rich uncondensed aliphatic amines, in the form of –NH2 and –NH– groups, provide abundant anchor sites for active species [14,15,16]. Because the planar adsorption of benzene on the surface of g-C3N4 is stronger than the non-planer adsorption of phenol [17], g-C3N4-supported catalysts usually exhibit good phenol selectivity in benzene hydroxylation, regardless of the supported transition metals (Cu, Fe, V, Co, Ni). The key problem for g-C3N4-based catalysts then becomes “how to improve the benzene conversion”. A frequently used strategy is to optimize the type and loading amount of metal sites. For example, Ding et al. [4] synthesized a series of metal-doped g-C3N4 catalysts and found that V-g-C3N4 was the most efficient one. Under optimal conditions, V-g-C3N4 can achieve 18.2% benzene conversion and 100% phenol selectivity. Wang et al. [18] optimized the loading of vanadia and found that 8V/g-C3N4 exhibited superior activity with a benzene conversion of 24.6% and phenol selectivity of 99.2%. Recently, Wang et al. [8] doped cerium into V/g-C3N4 catalysts by a simple co-assembly strategy and found that Ce0.07/0.07V/g-C3N4 greatly improved catalytic activity with 33.7% benzene conversion and 95.9% phenol selectivity.
At a fixed loading amount of vanadium, the structure of g-C3N4 also affects the catalytic performance of V-g-C3N4. For example, using mesoporous carbon nitride as a support material can increase the surface area of g-C3N4 and the dispersion of vanadia to improve the benzene conversion [19,20]. Recently, Xu et al. reported that exfoliated and protonated g-C3N4 can provide more anchor sites for the immobilization of active species [21]. They proposed that the chemical environment (especially its nitrogen species) of g-C3N4 is critical to the loading of vanadia. Notably, the protonation of g-C3N4 was usually conducted in strong acidic conditions, which not only change the local chemical environment but also the surface area. To investigate the intrinsic influence of the chemical environment of g-C3N4 on the vanadia dispersion and the catalytic performance, a synthetic method that can regulate the surface structure of g-C3N4 without changing the bulk structures is highly desired.
Herein, we develop a one-pot synthesis of V-pg-C3N4 to regulate the surface amino groups of g-C3N4 without changing the bulk structure and surface area. Characterizations and catalytic tests suggest that the content of surface amino groups is a decisive factor for the dispersion of vanadia and can quasi-linearly influence the catalytic performance of V-pg-C3N4 in the direct hydroxylation of benzene to phenol. A maximum of 60% benzene conversion and 99.7% selectivity to phenol can be achieved at 60 oC on V-pg-C3N4-0.46, which is attributed to the high dispersion of V species stabilized by amino groups, the presence of V4+/V5+ redox pairs, and the cooperative benzene-activation capability of g-C3N4. The mass specific activity of V-pg-C3N4-0.46 is as high as 4.26 and 2.11 g gcat−1 h−1, which exceeds most reported results (Table 1).

2. Results and Discussion

2.1. Bulk Structures and Surface Structures of V-g-C3N4 and V-pg-C3N4

Unlike traditional methods that protonate g-C3N4 before loading vanadia [21], the one-pot synthesis of V-pg-C3N4 was conducted in a mixture of HCl, g-C3N4, and vanadyl acetylacetonate where the protonation of g-C3N4 and the loading of vanadia took place simultaneously. V-g-C3N4 was synthesized in the same way as V-pg-C3N4-0.46 except that no HCl was added. As shown in Figure 1a, the diffraction patterns of V-g-C3N4 and V-pg-C3N4-0.46 are very similar to those of g-C3N4, suggesting that vanadia loading and protonation treatment exert little influence on the bulk structure of C3N4. Characteristic peaks at 27.5° and 13.6° correspond to (002) and (001) planes of graphitic C3N4, which are in accordance with the interlayer stacking structure of aromatic systems and the in-plane reflection of the tri-s-triazine motifs, respectively [31]. N2 sorption measurements show that g-C3N4, V-g-C3N4, and V-pg-C3N4-0.46 have similar surface areas (ca. 11 m2 g−1), excluding the exfoliation of bulk g-C3N4 during the protonation treatment in the one-pot synthesis. No distinct peaks related to vanadia-containing phases are observed in V-g-C3N4 and V-pg-C3N4-0.46, likely due to the low mass loading and high dispersion of vanadia species [32].
Consistent with the similar bulk structure, the samples exhibit very similar FT-IR spectra (Figure 1b). The broad bands at ~3156 cm−1 can be ascribed to the stretching vibration of uncondensed amino functional groups (such as -NH- and -NH2) on the graphitic sheets and the O-H groups of the adsorbed water [33,34]. The characteristic bands at 1200–1650 cm−1 correspond to the stretching and rotation vibration of C-N and C=N in heterocycles [35]. The sharp peaks at 808 cm−1 are usually treated as the breathing modes of triazine units (C6N7, the building blocks of the g-C3N4 structure) [36,37,38]. Notably, though the peak location and shape are similar, the intensity varies, indicating their different local structures. Specifically, the intensity of the bands at 1200–1650 cm−1 ascribed to aromatic heterocyclic rings obviously weakens on V-pg-C3N4 in comparison with that of V-g-C3N4 while the intensity of the bands at 3156 cm−1 is almost identical. These results indicate that the addition of HCl in the preparation would cleave some heterocycles in g-C3N4. According to the literature [21,39], such cleavage usually promotes the formation of defects sites and terminal N species (e.g., C-NHx). Considering the similar bulk structures of V-g-C3N4 and V-pg-C3N4, the cleavage of heterocycles likely occurred only on the catalyst surface.
To further reveal the influence of protonation on the surface structure of V-pg-C3N4, we performed N1s and V2p X-ray photoelectron spectroscopy (XPS) analysis for V-g-C3N4 and V-pg-C3N4-0.46. As shown in Figure 2a, the high-resolution spectra of N1s can be deconvoluted into three peaks. The major peak with a binding energy of 398.4 eV is ascribed to triazine nitrogen (C-N=C, Na in Figure 2b) [40]. The second peak at 399.3 eV corresponds to sp3-hybridized three-coordinate N species (N-(C)3 and C-N(-C)-H, Nb in Figure 2b) [41,42], and the peak with the highest binding energy can be assigned to sp3-hybridized surface amino groups (e.g., -NH2 and -NH-, Nc in Figure 2b) at the edge of g-C3N4 sheets [21,43]. Interestingly, the relative intensities of these peaks are different for V-pg-C3N4-0.46 and V-g-C3N4, suggesting the different composition of N species. Specifically, the percentage of Na decreases from 0.72 on V-g-C3N4 to 0.59 on V-pg-C3N4-0.46. This result is consistent with the reduced FT-IR band intensity at 1200–1650 cm−1, suggesting the presence of HCl in the preparation process indeed changes the surface heterocycle structure of C3N4. The percentage of Nb increases from 16% for V-g-C3N4 to 26% for V-pg-C3N4-0.46 while the percentage of Nc increases from 12% for V-g-C3N4 to 14% for V-pg-C3N4-0.46, suggesting the cleaved heterocycles were converted into groups with sp3-hybridized N species. The changed surface structure of g-C3N4 support eventually alters the surface state of supported vanadia. As shown in Figure 2c, the V2p3/2 spectra can be deconvoluted into two peaks at 515.3 and 517.1 eV, which are characteristic of V4+ and V5+ species, respectively [19]. Notably, the molar ratio of V4+/V5+ changes from 0.31/0.69 for V-g-C3N4 to 0.21/0.79 for V-pg-C3N4-0.46. This is likely due to the stronger interaction between vanadia and protonated g-C3N4 support in V-pg-C3N4-0.46 [22]. Specifically, the abundant Nb and Nc species in V-pg-C3N4-0.46 provide more anchoring sites for vanadia species, which improve the dispersion of vanadia and promote the electron transfer between vanadia and the g-C3N4 support.
To quantify the influence of HCl addition on the surface structure of V-pg-C3N4-X (X corresponds to mol HCl/mol N), we further investigated the XPS spectra for other V-pg-C3N4-X catalysts and plotted the percentage of (Nb + Nc) as a function of X. As shown in Figure 2d, the percentage of (Nb + Nc) exhibits a volcano plot versus X. Specifically, (Nb + Nc)/N increases from 28% for V-g-C3N4 (X = 0) and 36% for V-pg-C3N4-0.09 (X = 0.09) to 41% for V-pg-C3N4-0.46 (X = 0.46), and then decreases to 31% for V-pg-C3N4-0.93 (X = 0.93). The maximum (Nb + Nc)/N is achieved at X = 0.46, suggesting a proper amount of HCl addition is critical to the surface structure. When X is lower than 0.46, the amount of HCl is too small to change the bulk structure of g-C3N4 (Figure 1a). The increase in X would promote the hydrolysis of the aromatic CN heterocycles on the surface of g-C3N4 and therefore lead to increased (Nb + Nc)/N, but when X is larger than 0.46, the excessive HCl not only hydrolyzes the aromatic CN heterocycles but also promotes the exfoliation of bulk g-C3N4. The further increase in X would expose more Na on the surface and therefore decrease (Nb + Nc)/N. As previously reported in the literature [3,44], the amino groups on the surface can react with VO(acac)2 to immobilize vanadia species. It is therefore expected that higher vanadia dispersion would be achieved by V-pg-C3N4 with larger (Nb + Nc)/N. To confirm this hypothesis, we calculated the surface V/N molar ratio. Interestingly, it first increases from to 0.8% for V-pg-C3N4-0.05 to 2.1% for V-pg-C3N4-0.46 and then decreases to 1.2% for V-pg-C3N4-0.92. The similar trend of (Nb + Nc)/N and V/N molar ratio confirms that high surface amino groups are beneficial for vanadia dispersion.

2.2. Microstructures of V-g-C3N4 and V-pg-C3N4

The different surface chemical states are associated with the different microstructures of V-g-C3N4 and V-pg-C3N4-0.46 (Figure 3). As shown in Figure 3a and 3b, V-pg-C3N4-0.46 presents a thin lamellar and platelet-like structure, which is similar to that of g-C3N4. Elemental mapping of V-pg-C3N4 shows a homogeneous distribution of C, N, and V (Figure 3e). No obvious aggregation of vanadia on the surface of g-C3N4 was observed, confirming the high dispersion of vanadia species over the support. In stark contrast, the transmission electron microscope (TEM) and scanning electron microscope (SEM) images of V-g-C3N4 exhibit two distinct characters (Figure 3c and 3d). The thin lamellar and platelet-like structures are characteristic of g-C3N4 while the nanorods are vanadia. Elemental mapping of V-g-C3N4 (Figure 3f) shows an inhomogeneous distribution of C, N, and V elements, which further confirms the phase separation observed by TEM and SEM. The distinct microstructures between V-g-C3N4 and V-pg-C3N4-0.46 likely originate from the different surface structure and interfacial interaction. According to our previous publications, strong interfacial interaction benefits the high dispersion of precursors on the support surface, eventually leading to the high dispersion of supported nanomaterials. For example, strong electronic interaction between Bi3+ and ZnO (or TiO2) motivates high dispersion of Bi precursors on ZnO (or TiO2) [45]. During the following transformation of Bi3+ into BiOI or Bi2O3, the highly dispersed Bi3+ favors bounded nucleation and growth, which eventually lead to the high dispersion of BiOI or Bi2O3 on the supports [46,47]. For g-C3N4, the interaction between amino groups and vanadium species is significantly higher than that between triazine nitrogen and vanadium species [4]. When HCl is added to the preparation of V-pg-C3N4-0.46, the surface of g-C3N4 is protonated and can provide more amino groups to anchor vanadium species. The bounded vanadium precursors in situ convert into vanadia and therefore achieve high dispersion on the surface of protonated g-C3N4. In the absence of HCl, the surface amino groups are limited and the interaction between g-C3N4 and vanadium species is too weak to drive the high dispersion of vanadium precursors. The free nucleation and growth of vanadium species eventually produce vanadia nanorods that are separated from the g-C3N4 support. It is important to highlight that in traditional synthetic methods, the high dispersion of vanadia in V-g-C3N4 was usually achieved by increasing the surface area of g-C3N4 (e.g., exfoliation or mesopores) [19,21]. In this study, however, the high dispersion of vanadia in V-pg-C3N4 was achieved without changing the bulk structure of g-C3N4. The similar bulk structure, surface area but distinct vanadia dispersion between V-g-C3N4 and V-g-C3N4 suggest the decisive factor for high vanadia dispersion is the surface structure of g-C3N4 rather than the surface area.

2.3. Catalytic Performance of V-pg-C3N4

The catalytic performance of the catalysts was tested in direct hydroxylation of benzene to phenol using H2O2 as the oxidant. According to the literature, H2O2 was relatively stable in weak acidity conditions [48]. To this end, a mixture of acetonitrile and acetic acid was used as a solvent. Control experiment suggests that both g-C3N4 and protonated g-C3N4 display no activity in the absence of vanadia species, which is associated with its poor activity in H2O2 activation [4]. Interestingly, once vanadia was supported, V-g-C3N4 exhibited considerable conversion of benzene, suggesting vanadia species can activate H2O2 to hydroxylate benzene. Before evaluating the catalytic performances of different vanadia catalysts, we first optimized the reaction conditions. The optimal reaction temperature is 333 K as it achieves high phenol yield and avoids the volatilization of benzene (boiling point of 353 K). Figure 4a plots benzene conversion and phenol selectivity as a function of reaction time over V-pg-C3N4-0.46. The selectivity of phenol is always higher than 95%, which is similar to that for other reported C3N4-based catalysts (Table 1). According to the literature [4], the high selectivity of phenol is closely related to the unique structure of C3N4. Xu et al. [19] carried out benzene temperature-programmed desorption experiments over g-C3N4, V2O5, and V/g-C3N4 and found that benzene preferred to adsorb on g-C3N4 instead of vanadia. Notably, g-C3N4 features aromatic s-triazine rings and π electrons, which facilitate a strong planar adsorption of benzene on the surface of g-C3N4. However, once the adsorbed benzene is hydroxylated to phenol by H2O2, the aromatic ring alters the symmetry of the molecular orbital [17]. The resulting non-planer adsorption of phenol is much weaker than the planer adsorption of benzene. As a result, phenol tends to desorb from g-C3N4. The conversion of benzene rapidly increases from 25% to 62% along with the reaction time extension from 1 h to 5 h. Further increasing the reaction to 8 h, the conversion slightly increases to ~70%, likely due to the runout of H2O2. To this end, the products of 5-h reaction were used to evaluate the catalytic performance of different catalysts.
Figure 4b plots the catalytic performance of V-pg-C3N4-X as a function of X, i.e., the amount of HCl (mol HCl/mol N) used in the preparation of V-pg-C3N4-X. When X = 0, it is also called V-g-C3N4. Interestingly, all V-pg-C3N4 catalysts exhibit higher benzene conversion than V-g-C3N4, suggesting that the addition of HCl in the preparation benefits the catalytic performance. In particular, the benzene conversion over V-pg-C3N4-0.46 is 62%, which is significantly higher than that of V-g-C3N4 (42%). Considering that these two catalysts have similar XRD patterns, surface area, and vanadia loading amount, the mass transfer performances should be the same. Their different catalytic performances most likely originate from their distinct surface structures, especially their different amino contents and vanadia dispersion which influence the activation of H2O2. Second, the catalytic performance of V-pg-C3N4-X exhibits a volcano plot versus X, suggesting the amount of HCl added in the preparation is critical. Interestingly, the plot follows the same trend as that of the volcano plot of (Nb + Nc)/N versus X. Specifically, when X is lower than 0.46, (Nb + Nc)/N and benzene conversion increase along with the increase in X, but when X exceeds 0.46, they decrease. A maximum of 62% benzene conversion with 60.1% phenol yield is achieved at 333 K over V-pg-C3N4-0.46, a catalyst with the highest (Nb + Nc)/N. The mass specific activity of V-pg-C3N4-0.46 is as high as 2.11 g gcat−1 h−1, which exceeds most reported results (Table 1). Moreover, when we plot benzene conversion as a function of the percentage of (Nb + Nc), a quasi-linear relationship is obtained (Figure 4c). These results are consistent with the hypothesis that high vanadia dispersion enabled by abundant surface amino groups benefits the hydroxylation of benzene to phenol over V-pg-C3N4.

2.4. Reaction Mechanism of Benzene Hydroxylation over V-pg-C3N4

According to the literature [18,19,49], the hydroxylation of benzene to phenol over V-pg-C3N4 follows a synergistic mechanism (Figure 4d). Specifically, benzene is adsorbed and activated on the g-C3N4 support while H2O2 is involved in the V4+/V5+ redox cycle. Considering that phenol is produced by the reaction between activated benzene and V5+–O–O, the overall reaction rate of benzene hydroxylation depends on both activation reactions. In general, due to the abundant tri-s-triazine moieties and the facile electron transfer from C3N4 to benzene, the adsorption and activation of benzene over g-C3N4 can easily occur. In contrast, the oxidation of V4+ into V5+–O–O by H2O2 may be limited by the number of active vanadia sites. In this study, the surface structure (amino groups and vanadia dispersion) of V-pg-C3N4 was regulated without changing the bulk structures, providing an ideal model to investigate the influence of surface structure on the catalytic performance of benzene hydroxylation. According to the above discussion, the addition of HCl in the preparation of V-pg-C3N4-X leads to the protonation of g-C3N4. The resultant increase in surface amino groups provides more anchoring sites for vanadia species and therefore achieves higher vanadia dispersion. Considering that V-pg-C3N4-0.46 has a similar surface area and vanadia loading amount as V-g-C3N4, the higher vanadia dispersion means it has more active vanadia sites for H2O2 activation, which therefore leads to superior catalytic performance in benzene hydroxylation to phenol.

3. Materials and Methods

3.1. Preparation of g-C3N4

g-C3N4 was synthesized by the direct pyrolysis of melamine. Briefly, 5.0 g of melamine was placed in a crucible with a cover and then heated at 550 °C for 2 h with a ramping rate of 5 °C min−1.

3.2. Preparation of V-pg-C3N4

Typically, 0.3 g of as-synthesized g-C3N4 was dispersed into 10 mL of water containing 0.032 g of vanadyl acetylacetonate. After vigorous stirring for 20 min, a certain amount of HCl (ca. 37 wt%) was added to the above mixture dropwise. The mixture was stirred at room temperature for 2 h and subsequently heated at ~80 °C until the solvent was totally evaporated. The obtained product was calcined in the crucible with a cover at 300 °C for 2 h with a ramping rate of 2 °C min−1. The obtained sample was labeled as V-pg-C3N4-X, where X (X = 0.05, 0.09, 0.28, 0.46, 0.70, and 0.93) is the molar ratio between HCl and N atom in g-C3N4 (mol HCl/mol N). V-g-C3N4 was synthesized in the same way as V-pg-C3N4-X except that no HCl was added. Inductively coupled plasma mass spectrometry analysis confirmed that the mass loading of V was around 1.5 wt.%.

3.3. Characterizations

Scanning electron microscopy (SEM) images and elemental mapping were obtained using a Hitachi S4800 SEM microscopy equipped with an energy dispersive X-ray spectroscopy (EDS). Transmission electron microscopy investigations were performed on a HT7700 electron microscopy. X-ray diffraction patterns were collected using a Rigaku Ultimate IV diffractometer (Cu Kα radiation, 40 kV and 30 mA). Fourier transform infrared (FT-IR) spectra were recorded on a Thermo Nicolet 380 spectrometer. X-ray photoelectron spectroscopy (XPS) measurements were performed on a VG Scientific ESCALAB Mark II spectrometer.

3.4. Catalytic Performance Evaluation

The direct hydroxylation of benzene to phenol was performed in a three-neck round-bottom flask equipped with a reflux condenser. Typically, 1 mL of benzene, 4.8 mL of acetonitrile, 1.2 mL of 80 wt.% acetic acid, and 60 mg of catalyst were added to a 25 mL three-necked flask. The solution was stirred at 60 °C for 20 min to ensure adsorption equilibrium of benzene on the catalyst. Then, 3 mL of aqueous H2O2 solution (30 wt.%, 29.5 mmol) was dropwise added into the reactor within 2 min under vigorous stirring. The reaction was conducted at 60 °C for 1–8 h. After the reaction, the mixture was separated and the liquid products were analyzed by Shimadzu LC-20AD HPLC with an Ultimate XB-C18 column.
Conversion of benzene (Conv.) and selectivity to phenol (Sel.) were calculated as follows:
C o n v . = n p h e n o l + n B Q + n H Q + n C A n b e n z e n e + n p h e n o l + n B Q + n H Q + n C A × 100 %
S e l . = n p h e n o l n p h e n o l + n B Q + n H Q + n C A × 100 %
where nbenzene, nphenol, nBQ, nHQ, nCA are the molar amount (mol) of benzene, phenol, benzoquinone (BQ), hydroquinone (HQ), and catechol (CA).
The mass specific activity for each catalyst was calculated as follows:
S p e c i f i c   a c t i v i t y   g   h 1 g c a t a l . 1 = n p h e n o l × M p h e n o l W c a t a l . × t
where Mphenol, Wcatal., and t represent the formula weight (g mol−1) of phenol, the mass of overall catalyst (g), and the reaction time (h), respectively.

4. Conclusions

We developed a one-pot synthesis of V-pg-C3N4 which protonates g-C3N4 and loads vanadia simultaneously. Characterizations suggest that the protonation of g-C3N4 in diluted HCl solution facilitates the generation of amino groups (NH/NH2) on the surface without changing its bulk structure. The generated amino groups serve as anchoring sites to stabilize the highly dispersed vanadia species, which efficiently activate H2O2 into radical-containing V5+ species to react with the adjacent benzene activated over g-C3N4. Interestingly, the percentage of surface amino groups and the vanadia dispersion can be easily regulated by the amount of HCl added in the preparation. The optimal catalyst, V-pg-C3N4-0.46, achieves 60% benzene conversion and 99.7% phenol selectivity at 60 oC with H2O2 as the oxidant. The mass-specific activity of V-pg-C3N4-0.46 is as high as 2.11 g gcat−1 h−1, which exceeds most reported results.

Author Contributions

Conceptualization, J.L. and S.Z.; methodology, J.L.; validation, J.L. and Q.N.; investigation, J.L. and H.Y.; writing—original draft preparation, J.L.; writing—review and editing, S.Z.; visualization, J.L.; supervision, J.L. and S.Z.; project administration, J.L. and S.Z.; funding acquisition, J.L. and S.Z. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the National Natural Science Foundation of China (21703050, 21802122) and the Zhejiang Province Natural Science Foundation (LY22B030010).

Data Availability Statement

The data presented in this study are available on request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds V-g-C3N4 and V-pg-C3N4 are available from the authors.

References

  1. 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] [PubMed]
  2. 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, 56, 7779–7782. [Google Scholar] [CrossRef]
  3. 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, 51, 7756–7761. [Google Scholar] [CrossRef] [PubMed]
  4. Ding, G.; Wang, W.; Jiang, T.; Han, B.; Fan, H.; Yang, G. Highly Selective Synthesis of Phenol from Benzene over a Vanadium-Doped Graphitic Carbon Nitride Catalyst. ChemCatChem 2013, 5, 192–200. [Google Scholar] [CrossRef]
  5. Guo, B.; Zhu, L.; Hu, X.; Zhang, Q.; Tong, D.; Li, G.; Hu, C. Nature of vanadium species on vanadium silicalite-1 zeolite and their stability in hydroxylation reaction of benzene to phenol. Catal. Sci. Technol. 2011, 1, 1060–1067. [Google Scholar] [CrossRef]
  6. Lee, C.-H.; Lin, T.-S.; Mou, C.-Y. 2+ ions immobilized on functionalized surface of mesoporous silica and their activity toward the hydroxylation of benzene. J. Phys. Chem. B 2003, 107, 2543–2551. [Google Scholar] [CrossRef]
  7. 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]
  8. Wang, C.; Hu, L.; Wang, M.; Yue, B.; He, H. Cerium promoted Vg-C3N4 as highly efficient heterogeneous catalysts for the direct benzene hydroxylation. Roy. Soc. Open Sci. 2018, 5, 180371. [Google Scholar] [CrossRef] [Green Version]
  9. Long, Z.; Zhou, Y.; Chen, G.; Ge, W.; Wang, J. C3N4-H5PMo10V2O40: A dual-catalysis system for reductant-free aerobic oxidation of benzene to phenol. Sci. Rep. 2014, 4, 1–5. [Google Scholar] [CrossRef] [Green Version]
  10. Borah, P.; Datta, A.; Nguyen, K.T.; Zhao, Y. VOPO4·2H2O encapsulated in graphene oxide as a heterogeneous catalyst for selective hydroxylation of benzene to phenol. Green Chem. 2016, 18, 397–401. [Google Scholar] [CrossRef]
  11. 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]
  12. 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]
  13. Shah Bacha, R.u.; Li, L.; Guo, Y.-R.; Jing, L.; Pan, Q.-J. Actinyl-modified g-C3N4 as CO2 activation materials for chemical conversion and environmental remedy via an artificial photosynthetic route. Inorg. Chem. 2020, 59, 8369–8379. [Google Scholar] [CrossRef] [PubMed]
  14. Jin, X.; Balasubramanian, V.V.; Selvan, S.T.; Sawant, D.P.; Chari, M.A.; Lu, G.e.Q.; Vinu, A. Highly ordered mesoporous carbon nitride nanoparticles with high nitrogen content: A metal-free basic catalyst. Angew. Chem. Int. Ed. 2009, 48, 7884–7887. [Google Scholar] [CrossRef] [PubMed]
  15. Xu, J.; Shen, K.; Xue, B.; Li, Y.-X. Microporous carbon nitride as an effective solid base catalyst for Knoevenagel condensation reactions. J. Mol. Catal. A 2013, 372, 105–113. [Google Scholar] [CrossRef]
  16. Xu, J.; Chen, T.; Jiang, Q.; Li, Y.X. Utilization of Environmentally Benign Dicyandiamide as a Precursor for the Synthesis of Ordered Mesoporous Carbon Nitride and its Application in Base-Catalyzed Reactions. Chem. Asian J. 2014, 9, 3269–3277. [Google Scholar] [CrossRef]
  17. Wang, Y.; Yao, J.; Li, H.; Su, D.; Antonietti, M. Highly Selective Hydrogenation of Phenol and Derivatives over a Pd@Carbon Nitride Catalyst in Aqueous Media. J. Am. Chem. Soc. 2011, 133, 2362–2365. [Google Scholar] [CrossRef]
  18. Wang, C.; Hu, L.; Wang, M.; Ren, Y.; Yue, B.; He, H. Vanadium supported on graphitic carbon nitride as a heterogeneous catalyst for the direct oxidation of benzene to phenol. Chin. J. Catal. 2016, 37, 2003–2008. [Google Scholar] [CrossRef]
  19. 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]
  20. Xu, J.; Jiang, Q.; Shang, J.-K.; Wang, Y.; Li, Y.-X. A Schiff-base-type vanadyl complex grafted on mesoporous carbon nitride: A new efficient catalyst for hydroxylation of benzene to phenol. RSC Adv. 2015, 5, 92526–92533. [Google Scholar] [CrossRef]
  21. Xu, J.; Chen, Y.; Hong, Y.; Zheng, H.; Ma, D.; Xue, B.; Li, Y.-X. Direct catalytic hydroxylation of benzene to phenol catalyzed by vanadia supported on exfoliated graphitic carbon nitride. Appl. Catal. A 2018, 549, 31–39. [Google Scholar] [CrossRef]
  22. Wang, H.; Zhao, M.; Zhao, Q.; Yang, Y.; Wang, C.; Wang, Y. In-situ immobilization of H5PMo10V2O40 on protonated graphitic carbon nitride under hydrothermal conditions: A highly efficient and reusable catalyst for hydroxylation of benzene. Ind. Eng. Chem. Res. 2017, 56, 2711–2721. [Google Scholar] [CrossRef]
  23. Xue, B.; Chen, Y.; Hong, Y.; Ma, D.-Y.; Xu, J.; Li, Y.-X. Facile synthesis of Fe-containing graphitic carbon nitride materials and their catalytic application in direct hydroxylation of benzene to phenol. Chin. J. Catal. 2018, 39, 1263–1271. [Google Scholar] [CrossRef]
  24. Qin, L.; Feng, Z.; Zhang, Q.; Mao, H.; Cheng, F.; Shi, S. Enhanced Hydroxylation of Benzene to Phenol with Hydrogen Peroxide over g-C3N4 Quantum Dots-Modified Fe-SBA-15 Catalysts: Synergistic Effect Among Fe Species, g-C3N4 QDs, and Porous Structure. Ind. Eng. Chem. Res. 2021, 60, 13876–13885. [Google Scholar] [CrossRef]
  25. Zhu, Y.; Dong, Y.; Zhao, L.; Yuan, F. Preparation and characterization of Mesopoous VOx/SBA-16 and their application for the direct catalytic hydroxylation of benzene to phenol. J. Mol. Catal. A 2010, 315, 205–212. [Google Scholar] [CrossRef]
  26. Nemati Kharat, A.; Moosavikia, S.; Tamaddoni Jahromi, B.; Badiei, A. Liquid phase hydroxylation of benzene to phenol over vanadium substituted Keggin anion supported on amine functionalized SBA-15. J. Mol. Catal. A 2011, 348, 14–19. [Google Scholar] [CrossRef]
  27. Xu, D.; Jia, L.; Guo, X. Direct Hydroxylation of Benzene to Phenol Over Mixed-Crystal Particles of Mesoporous VOx/TiO2 Catalyst Mixed-Crystal VOx/TiO2 for Benzene Hydroxylation. Catal. Lett. 2012, 142, 1251–1261. [Google Scholar] [CrossRef]
  28. Li, Y.; Wang, Z.; Chen, R.; Wang, Y.; Xing, W.; Wang, J.; Huang, J. The hydroxylation of benzene to phenol over heteropolyacid encapsulated in silica. Catal. Commun. 2014, 55, 34–37. [Google Scholar] [CrossRef]
  29. 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 2015, 504, 440–447. [Google Scholar] [CrossRef]
  30. Wang, C.; Hu, L.; Hu, Y.; Ren, Y.; Chen, X.; Yue, B.; He, H. Direct hydroxylation of benzene to phenol over metal oxide supported graphene oxide catalysts. Catal. Commun. 2015, 68, 1–5. [Google Scholar] [CrossRef]
  31. Muniandy, L.; Adam, F.; Mohamed, A.R.; Iqbal, A.; Rahman, N.R.A. Cu2+ coordinated graphitic carbon nitride (Cu-g-C3N4) nanosheets from melamine for the liquid phase hydroxylation of benzene and VOCs. Appl. Surf. Sci. 2017, 398, 43–55. [Google Scholar] [CrossRef]
  32. Liu, J.; Wu, X.-P.; Zou, S.; Dai, Y.; Xiao, L.; Gong, X.-Q.; Fan, J. Origin of the High Activity of Mesoporous CeO2 Supported Monomeric VOx for Low-Temperature Gas-Phase Selective Oxidative Dehydrogenation of Benzyl Alcohol: Role As an Electronic “Hole”. J. Phys. Chem. C 2014, 118, 24950–24958. [Google Scholar] [CrossRef]
  33. Maity, N.; Barman, S.; Minenkov, Y.; Ould-Chikh, S.; Abou-Hamad, E.; Ma, T.; Qureshi, Z.S.; Cavallo, L.; D’Elia, V.; Gates, B.C.; et al. A Silica-Supported Monoalkylated Tungsten Dioxo Complex Catalyst for Olefin Metathesis. ACS Catal. 2018, 8, 2715–2729. [Google Scholar] [CrossRef]
  34. Talapaneni, S.N.; Anandan, S.; Mane, G.P.; Anand, C.; Dhawale, D.S.; Varghese, S.; Mano, A.; Mori, T.; Vinu, A. Facile synthesis and basic catalytic application of 3D mesoporous carbon nitride with a controllable bimodal distribution. J. Mater. Chem. 2012, 22, 9831–9840. [Google Scholar] [CrossRef]
  35. Bojdys, M.J.; Müller, J.O.; Antonietti, M.; Thomas, A. Ionothermal synthesis of crystalline, condensed, graphitic carbon nitride. Chem. Eur. J. 2008, 14, 8177–8182. [Google Scholar] [CrossRef] [PubMed]
  36. Liu, J.; Zhang, T.; Wang, Z.; Dawson, G.; Chen, W. Simple pyrolysis of urea into graphitic carbon nitride with recyclable adsorption and photocatalytic activity. J. Mater. Chem. 2011, 21, 14398–14401. [Google Scholar] [CrossRef]
  37. Huang, Z.; Sun, Q.; Lv, K.; Zhang, Z.; Li, M.; Li, B. Effect of contact interface between TiO2 and g-C3N4 on the photoreactivity of g-C3N4/TiO2 photocatalyst: (001) vs (101) facets of TiO2. Appl. Catal. B 2015, 164, 420–427. [Google Scholar] [CrossRef]
  38. Wang, X.; Blechert, S.; Antonietti, M. Polymeric graphitic carbon nitride for heterogeneous photocatalysis. ACS Catal. 2012, 2, 1596–1606. [Google Scholar] [CrossRef]
  39. Zhu, Z.; Pan, H.; Murugananthan, M.; Gong, J.; Zhang, Y. Visible light-driven photocatalytically active g-C3N4 material for enhanced generation of H2O2. Appl. Catal. B 2018, 232, 19–25. [Google Scholar] [CrossRef]
  40. Su, Q.; Sun, J.; Wang, J.; Yang, Z.; Cheng, W.; Zhang, S. Urea-derived graphitic carbon nitride as an efficient heterogeneous catalyst for CO2 conversion into cyclic carbonates. Catal. Sci. Technol. 2014, 4, 1556–1562. [Google Scholar] [CrossRef]
  41. Huang, Z.; Li, F.; Chen, B.; Lu, T.; Yuan, Y.; Yuan, G. Well-dispersed g-C3N4 nanophases in mesoporous silica channels and their catalytic activity for carbon dioxide activation and conversion. Appl. Catal. B 2013, 136, 269–277. [Google Scholar] [CrossRef]
  42. Lan, D.-H.; Wang, H.-T.; Chen, L.; Au, C.-T.; Yin, S.-F. Phosphorous-modified bulk graphitic carbon nitride: Facile preparation and application as an acid-base bifunctional and efficient catalyst for CO2 cycloaddition with epoxides. Carbon 2016, 100, 81–89. [Google Scholar] [CrossRef]
  43. Xu, J.; Long, K.-Z.; Wang, Y.; Xue, B.; Li, Y.-X. Fast and facile preparation of metal-doped g-C3N4 composites for catalytic synthesis of dimethyl carbonate. Appl. Catal. A 2015, 496, 1–8. [Google Scholar] [CrossRef]
  44. Barman, S.; Maity, N.; Bhatte, K.; Ould-Chikh, S.; Dachwald, O.; Haeßner, C.; Saih, Y.; Abou-Hamad, E.; Llorens, I.; Hazemann, J.-L.; et al. Single-Site VOx Moieties Generated on Silica by Surface Organometallic Chemistry: A Way To Enhance the Catalytic Activity in the Oxidative Dehydrogenation of Propane. ACS Catal. 2016, 6, 5908–5921. [Google Scholar] [CrossRef]
  45. Zou, S.; Liu, J.; Kobayashi, H.; Chen, C.; Qiao, P.; Li, R.; Xiao, L.; Fan, J. Boosting Hydrogen Evolution Activities by Strong Interfacial Electronic Interaction in ZnO@Bi(NO3)3 Core–Shell Structures. J. Phys. Chem. C 2017, 121, 4343–4351. [Google Scholar] [CrossRef]
  46. Liu, J.; Zou, S.; Lou, B.; Chen, C.; Xiao, L.; Fan, J. Interfacial electronic interaction induced engineering of ZnO-BiOI heterostructures for efficient visible-light photocatalysis. Inorg. Chem. 2019, 58, 8525–8532. [Google Scholar] [CrossRef] [PubMed]
  47. Zou, S.; Lou, B.; Yang, K.; Yuan, W.; Zhu, C.; Zhu, Y.; Du, Y.; Lu, L.; Liu, J.; Huang, W. Grafting nanometer metal/oxide interface towards enhanced low-temperature acetylene semi-hydrogenation. Nat. Commun. 2021, 12, 1–11. [Google Scholar] [CrossRef]
  48. Yan, Y.; Chen, C.; Zou, S.; Liu, J.; Xiao, L.; Fan, J. High H2O2 Utilization Promotes Selective Oxidation of Methane to Methanol at Low Temperature. Front. Chem. 2020, 8, 252. [Google Scholar] [CrossRef]
  49. Nomiya, K.; Nemoto, Y.; Hasegawa, T.; Matsuoka, S. Multicenter active sites of vanadium-substituted polyoxometalate catalysts on benzene hydroxylation with hydrogen peroxide and two reaction types with and without an induction period. J. Mol. Catal. A 2000, 152, 55–68. [Google Scholar] [CrossRef]
Figure 1. (a) X-ray diffraction (XRD) patterns and (b) Fourier transform infrared spectroscopy (FT-IR) spectra of as-synthesized samples.
Figure 1. (a) X-ray diffraction (XRD) patterns and (b) Fourier transform infrared spectroscopy (FT-IR) spectra of as-synthesized samples.
Molecules 27 06965 g001
Figure 2. XPS analysis of V-g-C3N4 and V-pg-C3N4-0.46 samples. (a) N1s; (b) a possible structure of g-C3N4 reproduced with permission from [21]. Copyright 2018, Elsevier; (c) V2p; (d) the percentage of (Nb + Nc) as a function of X in V-pg-C3N4-X.
Figure 2. XPS analysis of V-g-C3N4 and V-pg-C3N4-0.46 samples. (a) N1s; (b) a possible structure of g-C3N4 reproduced with permission from [21]. Copyright 2018, Elsevier; (c) V2p; (d) the percentage of (Nb + Nc) as a function of X in V-pg-C3N4-X.
Molecules 27 06965 g002
Figure 3. TEM (a) and SEM (b) images of V-pg-C3N4-0.46; TEM (c) and SEM (d) images of V-g-C3N4. Elemental mapping of V-pg-C3N4-0.46 (e) and (f) V-g-C3N4.
Figure 3. TEM (a) and SEM (b) images of V-pg-C3N4-0.46; TEM (c) and SEM (d) images of V-g-C3N4. Elemental mapping of V-pg-C3N4-0.46 (e) and (f) V-g-C3N4.
Molecules 27 06965 g003
Figure 4. (a) Catalytic activity as a function of reaction time over V-pg-C3N4-0.46; (b) Catalytic activity (5 h) as a function of the HCl amount used in the preparation of V-pg-C3N4-X. Reaction conditions: benzene (1.0 mL, 11.3 mmol), hydrogen peroxide (30 wt%, 3.0 mL, 29.6 mmol), acetonitrile (4.8 mL), acetic acid (1.2 mL), catalyst (0.06 g, 0.022 mmol V), T = 333 K. (c) Benzene conversion as a function of (Nb + Nc)/N over V-pg-C3N4-X. (d) A possible reaction mechanism for the direct hydroxylation of benzene to phenol over V-pg-C3N4 catalysts.
Figure 4. (a) Catalytic activity as a function of reaction time over V-pg-C3N4-0.46; (b) Catalytic activity (5 h) as a function of the HCl amount used in the preparation of V-pg-C3N4-X. Reaction conditions: benzene (1.0 mL, 11.3 mmol), hydrogen peroxide (30 wt%, 3.0 mL, 29.6 mmol), acetonitrile (4.8 mL), acetic acid (1.2 mL), catalyst (0.06 g, 0.022 mmol V), T = 333 K. (c) Benzene conversion as a function of (Nb + Nc)/N over V-pg-C3N4-X. (d) A possible reaction mechanism for the direct hydroxylation of benzene to phenol over V-pg-C3N4 catalysts.
Molecules 27 06965 g004
Table 1. The catalytic activity of various vanadia- and g-C3N4-based catalysts for benzene hydroxylation.
Table 1. The catalytic activity of various vanadia- and g-C3N4-based catalysts for benzene hydroxylation.
CatalystVbenzene
(mL)
VH2O2
(mL)
Wcatal.
(mg)
T
(K)
t
(h)
Conv.
(%)
Sel.
(%)
Specific Activity
(h−1) a
Ref.
0.4V-g-C3N41.0360333617.71000.52[4]
H5PMo10V2O40/pg-C3N41.04.1100333825.899.70.34[22]
VO/MCM-41-NH20.91100333158.618.51.03[6]
8V-g-C3N41.03.540343424.699.21.61[18]
Fe-g-C3N41350333217.599.01.84[23]
V2O5-mp-C3N41.5360333318.795.91.59[19]
VO-peg-C3N41.0375333411.797.90.40[21]
g-C3N4 QD/Fe-SBA-1536120333341,798.83.01[24]
VOx-SBA-160.31.510333413.897.51.06[25]
HPMoV/NH2-SBA-151.03.0100333620.095.00.33[26]
VOx-TiO22.66180333526.390.00.72[27]
PMoV2/SiO21.03.0150333621.61000.25[28]
V-C-6000.41.420343331.894.92.12[29]
VOx-GO1.03.540338323.198.41.99[30]
Ce0.07-0.07 V-g-C3N41.03.540343433.795.92.13[8]
V-pg-C3N4-0.461.03.060333562.097.02.11p.w. b
V-pg-C3N4-0.461.03.060333125.097.04.26
a Specific activity is calculated based on the mass of synthesized phenol per gram of whole catalyst per hour. b Present work.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Liu, J.; Yin, H.; Nie, Q.; Zou, S. Highly Dispersed Vanadia Anchored on Protonated g-C3N4 as an Efficient and Selective Catalyst for the Hydroxylation of Benzene into Phenol. Molecules 2022, 27, 6965. https://doi.org/10.3390/molecules27206965

AMA Style

Liu J, Yin H, Nie Q, Zou S. Highly Dispersed Vanadia Anchored on Protonated g-C3N4 as an Efficient and Selective Catalyst for the Hydroxylation of Benzene into Phenol. Molecules. 2022; 27(20):6965. https://doi.org/10.3390/molecules27206965

Chicago/Turabian Style

Liu, Juanjuan, Haoyong Yin, Qiulin Nie, and Shihui Zou. 2022. "Highly Dispersed Vanadia Anchored on Protonated g-C3N4 as an Efficient and Selective Catalyst for the Hydroxylation of Benzene into Phenol" Molecules 27, no. 20: 6965. https://doi.org/10.3390/molecules27206965

Article Metrics

Back to TopTop