Next Article in Journal
Optimization Study of Glucose Synthesis to 5-Hydroxymethylfurfural (5-HMF) by Using Low Transition-Temperature Mixtures (LTTM)
Next Article in Special Issue
Preparation of High Geometric Filling Factor Perovskite Module and Feasibility Study on Electrocatalytic Hydrogen Production
Previous Article in Journal
Balance between Catalyst Acidity and Hydrophilicity in Biofuel Production from Fatty Acid Esterification over Al-SBA-15
Previous Article in Special Issue
Recent Progress of MIL MOF Materials in Degradation of Organic Pollutants by Fenton Reaction
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Atmospheric and Efficient Selective Oxidation of Ethylbenzene Catalyzed by Cobalt Oxides Supported on Mesoporous Carbon Nitride

Jiangsu Key Laboratory of Advanced Catalytic Materials and Technology, School of Petrochemical Engineering, Changzhou University, Gehu Middle Road 21, Changzhou 213164, China
*
Authors to whom correspondence should be addressed.
Catalysts 2023, 13(5), 828; https://doi.org/10.3390/catal13050828
Submission received: 29 March 2023 / Revised: 23 April 2023 / Accepted: 28 April 2023 / Published: 30 April 2023
(This article belongs to the Special Issue Recent Progress of Catalysis in “Dual Carbon Targets”)

Abstract

:
Mesoporous carbon nitride (mpg-C3N4) was prepared by using cyanamide as a precursor and colloidal nanosilica as a template. Then, the mpg-C3N4 was used as a catalytic support to load CoOx. The physicochemical properties of the synthesized CoOx/mpg-C3N4 materials were elucidated by multiple characterization methods, and the catalytic activities were examined in the selective oxidation of ethylbenzene (EB) under atmospheric pressure by using tert-butyl hydrogen peroxide (TBHP) as an oxidant. It was found that mpg-C3N4 possessed a higher specific surface area than other carbon nitride materials, and its abundant Nb species were able to interact with Co (II) species. When the dosages of EB and TBHP were 10 mmol and 30 mmol, respectively, the reaction temperature was 100 °C, and the reaction time was 10 h, the conversion rate of ethylbenzene was 62%, and the selectivity of AP was 84.7%.

1. Introduction

Liquid-phase selective oxidation of arenes is one of the most significant strategies for the synthesis of aromatic compounding, including aromatic alcohols, aldehydes, ketones, acids, etc. [1]. From the viewpoints of both fundamental research and industrial application, it is of great interest to develop an efficient and highly selective route for selective oxidation reactions [2]. Among these products, acetophenone (AP) is the simplest aromatic ketone and is widely used to manufacture pharmaceuticals, resins, drugs, perfumes, aldehydes, flavoring agents, etc. Conventionally, AP is produced as a byproduct in the decomposition of cumene hydroperoxide. The Friedel–Crafts acylation reaction, using acid halides or anhydrides in the presence of Lewis acids, is also a well-established process for the production of AP. However, these two processes are environmentally harmful because of the generation of a large amount of corrosive waste [3].
Alternatively, liquid-phase direct oxidation of ethylbenzene (EB) is a practical approach for the synthesis of AP. By now, the most widely reported oxidants include molecular oxygen (O2), hydrogen peroxide (H2O2), and tert-butyl hydroperoxide (TBHP). Although O2 is widely available and inexpensive, the oxidative reaction in the use of O2 always demands high temperatures and pressures. The use of H2O2 can enable the reaction to proceed under atmospheric pressure. However, the low conversion of EB is still a practical issue [4,5]. By comparison, adopting TBHP as an oxidant is a preferable method for the synthesis of AP as the process can be conducted under milder reaction conditions and also offer relatively higher productivity of AP [5].
Until now, the catalysts reported for the TBHP-involved oxidation in the synthesis of AP are mainly supported by Cu (II) [2], Co (II) [1], Cr (III) [6], Ag (I) [7], Pd (II) [8] compounds, etc. Despite their convenience in catalyst separation and product purification, the catalyst supports are mainly silica (or mesoporous silica) and pure carbon materials. Due to the low interaction, the supported metal species might separate from the catalyst support, and thus, leach out into the liquid phase [9,10]. In this sense, it is of interest to explore a new material as catalyst support, which not only has a large surface area but can also anchor the active metal components in the catalytic reactions.
Graphitic carbon nitride (g-C3N4) is a novel inorganic material, which because of its combination of multiple unique physicochemical properties, is widely used in photocatalysis [11,12], fuel cell [13], gas storage [14], etc. In the field of thermal catalysis, owing to its abundant nitrogen-containing groups, g-C3N4, especially mesoporous g-C3N4 with a large surface area, has been widely reported to be capable of dispersing and anchoring many metal species including K [15,16], Zn [17,18], Co [19,20], Au [21], Pd [22,23], etc. Therefore, g-C3N4 material has been regarded as a promising catalytic support in metal-catalyzed heterogeneous reactions [8]. Given the shortcomings in the use of the traditional catalyst supports, in this work, mesoporous carbon nitride (mpg-C3N4) was prepared by a hard-templating method, and then, utilized as a support to load cobalt oxide. The synthesized CoOx/mpg-C3N4 materials as heterogeneous catalysts showed good and stable catalytic activity in the selective oxidation of EB to AP by TBHP. The characterization results revealed that there was a possible interaction between Co (II) and the nitrogen-containing species of mpg-C3N4.

2. Results and Discussions

2.1. Materials Characterization

Figure 1 shows XRD patterns of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials. The mpg-C3N4 support has two obvious diffraction peaks at 2 θ of 12.7° and 27.6°, indexed as the intralayer and interlayer structure of graphite-like material, i.e., the (100) and (002) planes [24]. The positions of the two peaks show no obvious change from mpg-C3N4 and 3CoOx/mpg-C3N4-T, while the intensities of the two peaks become weaker because the mpg-C3N4 material was prepared by thermal condensation of cyanamide and contains a certain amount of incomplete condensed nitrogen-containing species [25,26]. During the preparation procedure of CoOx/mpg-C3N4, the mpg-C3N4 support suffered a second heating treatment, in which the incomplete condensed nitrogen-containing species would decompose into small-molecule gases, such as NH3. The generated gas inevitably deteriorated the graphic structure of mpg-C3N4, thus, leading to the decline of peak intensity in XRD patterns. Nevertheless, the comparison of XRD patterns of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials suggests that the incorporation of CoOx has not altered the overall graphitic structures of the mpg-C3N4 support. Furthermore, no additional peaks were detected, implying that the supported CoOx species might disperse better on the surface of mpg-C3N4.
The N2 adsorption–desorption isothermal curves of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials are described in Figure 2A. The mpg-C3N4 material exhibited a typical type-Ⅳ isothermal curve, along with an H3 hysteresis loop in the range of relative pressure (p/p0) = 0.6–0.95, indicating that the mpg-C3N4 material possesses mesoporous structures. Similar to mpg-C3N4, the supported 3CoOx/mpg-C3N4-T materials also have type-Ⅳ isothermal curves. The adsorption quantity of such 3CoOx/mpg-C3N4-0.6-T materials is associated with the preparation temperature. Among them, 3CoOx/mpg-C3N4-400 possesses the highest adsorption quantity, i.e., the largest total pore volume. The corresponding pore size distributions (Figure 2B) indicate that the pore sizes of both mpg-C3N4 and 3CoOx/mpg-C3N4-T materials are centered at ca. 12 nm, which is in agreement with the mean particle size (~12 nm) of the Ludox silica template. This suggests that by using the hard-templating procedure, the mpg-C3N4-0.6 material has successfully negatively replicated the structure of the silica nanoparticles.
Table 1 lists the specific surface area and porous properties of the above materials. The surface area (SBET) of mpg-C3N4 is 84 m2∙g−1, and the pore volume is 0.30 cm3∙g−1. For the 3CoOx/mpg-C3N4-T materials, which were prepared by heating treatment of mpg-C3N4 and Co(NO3)2, in addition to the generation of CoOx from Co(NO3)2, the mpg-C3N4 support also suffers second calcination. Wherein, a slightly higher temperature could induce the decomposition of the skeleton of C3N4 of mpg-C3N4, which is responsible for the larger surface area and pore volume of 3CoOx/mpg-C3N4-400 than 3CoOx/mpg-C3N4-300. However, an extra high heating temperature might induce the partial collapse of the mesoporous structures.
The microscopic morphology of the materials was analyzed by SEM (Figure 3). The image of mpg-C3N4 revealed coral-like morphology. In the case of 3CoOx/mpg-C3N4, the SEM image also exhibited a coral-like structure, whereas directly judging the image images, the particles of 3CoOx/mpg-C3N4 look much smaller than those of mpg-C3N4. The energy dispersive X-ray spectroscopy (EDX) mapping (Figure 4) of 3CoOx/mpg-C3N4 indicates that carbon, nitrogen, oxygen, and cobalt elements are evenly distributed in this material.
TEM was used to further analyze the microscopic morphology of the mpg-C3N4 and 3CoOx/mpg-C3N4 materials. As displayed in Figure 5, mpg-C3N4 features disordered foam-like pores (especially at the edge of the particles) with diameters in a range of 10–15 nm, similar to the reported mesoporous g-C3N4 samples prepared using nanoparticles as hard templates [26,27]. These TEM images, in conjunction with the results of the N2 adsorption–desorption (Table 1), provide evidence that the disordered mesoporous structure of mpg-C3N4 is the negative replica of the silica templates. The image of the 3CoOx/mpg-C3N4 also reveals a disordered foam-like porous structure, some of which, however, seem to be blocked and aggregated. As described above, this is due to the collapse of the mesostructures of mpg-C3N4 during its second calcination. The CoOx particles are found on the surface of mpg-C3N4. Rough statistics show that the mean size of these CoOx particles is ca. 15 nm (Figure S1).
The chemical functionalities of mpg-C3N4 and the supported materials were analyzed by FT-IR spectroscopy (Figure S2). The spectra of all the materials exhibit sharp and strong transmittance peaks centered ca. 812 cm−1, characteristic bands corresponding to the breathing mode of conjugated heptazine units [26]. The multiple bands in the range of 1200–1600 cm−1 are attributed to the stretching vibration of the C–N and C=N groups in the heterocyclic compounds. In addition, a broad band is detected at ca. 3152 cm−1 in each material. This transmittance signal is ascribed to the O–H groups in the adsorbed water and the amino group in mpg-C3N4. The above bands in the three regions are characteristic of g-C3N4. Comparing all the spectra from the 3CoOx/mpg-C3N4 materials with various preparation temperatures and loading amounts (Figure S3) of CoOx, no additional bands were observed, suggesting that the loading of CoOx has not changed the overall chemical functionalities of mpg-C3N4.
The surface chemical bonding state was probed by XPS analysis. Figure 6 is the survey spectra of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials. Three peaks appear at the binding energies of ca. 288, 398, and 531 eV, which signify the carbon, nitrogen, and oxygen elements, respectively. In the cases of the 3CoOx/mpg-C3N4-T materials, in addition to the three peaks, peaks were detected at ca. 781 eV, indexed as Co 2p orbits. Based on the peak areas, the molar ratios of carbon to nitrogen in the materials were calculated. As listed in Table S1, the molar ratio obtained from mpg-C3N4 was 0.85:1, much higher than the theoretical value of ideal g-C3N4 (0.75:1) [24]. Compared to the bulk g-C3N4, prepared without templating procedure, the C3N4 wall of mpg-C3N4 was much thinner; thus, resulting in an easier decomposition of the nitrogen-containing groups during the heating step of mpg-C3N4 [27]. As for the 3CoOx/mpg-C3N4-T materials, the molar ratios of carbon to nitrogen are close to the value of mpg-C3N4.
As described above, supported Co materials have recently been widely studied as heterogeneous catalysts in the selective oxidation of EB, and therein, the chemical bonding state of cobalt species is an important factor in deciding the catalytic performance [28]. Given this point, the fine Co 2p3/2 spectra of 3CoOx/mpg-C3N4-T were deconvoluted (note: the deconvolution of Co 2p1/2 spectra was not performed as the corresponding peaks were very weak). As illustrated in Figure 7, the Co 2p3/2 signals can be separated into four peaks. The peaks with binding energies of ca. 780.4 and 785.2 eV are assigned to Co (III) and Co (II), respectively [9,29,30]. The weak peaks centered at ca. 788.2 eV are attributed to shake-up (satellite) signals in the Co 2p3/2 orbit [31,32], and the low-intensity signals in the range of 774–778 eV are attributed to Co0. Such metallic cobalt species could come from the reduction of cobalt cations by some nitrogen-containing species of mpg-C3N4 during the preparation procedure (i.e., heating treatment), which has also been reported in cobalt supported by nitrogen-doped carbon materials [33,34].
On the other hand, g-C3N4 and its derived materials contain abundant nitrogen-containing groups and particular nitrogen pots surrounded by nonbonding sp2 orbitals of nitrogen atoms [35]. Comparable to porphyrin and phthalocyanine, such unique structures enable g-C3N4 and its related materials to be able to include metal species [18]. In view of this point, the fine N 1s spectra of various materials were analyzed and presented in Figure 8. The spectra of mpg-C3N4 and 3CoOx/mpg-C3N4-T are deconvoluted into three independent peaks. The major peaks with the lowest binding energies of 397.8 eV are attributed to sp2-hybridized nitrogen atoms (Na) in the conjugated heptazine units. This type of nitrogen accounts for over 65% of the total nitrogen species. The peaks at 398.9 and 400.2 eV correspond to bridging nitrogen atoms (Nb) that connect several adjacent heptazine units, and incomplete uncondensed amino groups (–NH2 and –NH–, Nc), respectively [26,36]. Accordingly, the distributions of various nitrogen species in the materials were calculated (Table 2). It can be found that the introduction of CoOx affects the distribution of the nitrogen groups, leading to the transformation of a small number of Na to Nb species. The variation is probably derived from the partial destruction of the heptazine rings (Na) in mpg-C3N4, which occurs in the preparation steps of 3CoOx/mpg-C3N4-T and has actually been reported in previous literature about metal-included g-C3N4 materials [37,38]. Moreover, among the three 3CoOx/mpg-C3N4-T materials, 3CoOx/mpg-C3N4-400 owns the highest percentage of Nb (25.5%). It should be noted that such a type of nitrogen has a superior capacity to Na and Nc when interacting with a transition-metal cation [39]; thus, probably facilitating the dispersion of Co species in the mpg-C3N4 support.
The interaction between Co and mpg-C3N4 has also been evidenced by the UV–vis spectra. As depicted in Figure S4, the spectrum of the mpg-C3N4 material displays photo-absorption in the UV region, ascribed to the bandgap between HOMO and LUMO in the polymeric heptazine units of g-C3N4. The spectra of 3CoOx/mpg-C3N4-T materials show similar absorption intensities in the UV region to the support. Notwithstanding, compared with mpg-C3N4, the supported materials demonstrate enhanced absorption in the visible light region. Such a variation in the electronic structure of mpg-C3N4 is attributed to the inclusion of the metal cation in the framework of g-C3N4, i.e., d–p repulsion between N 2p and Co 3d orbits, which has also been found in metal-doped g-C3N4 materials [18,40,41,42]. Moreover, in the cases of CoOx/mpg-C3N4 materials with various loading amounts of CoOx, it can be found that as the loading amount is increased, the increase in the absorption intensity in the visible region becomes more obvious. In fact, the change in the electronic structure has also been reflected by the apparent colors of the materials. The color of mpg-C3N4 is pale yellow, whereas the 3CoOx/mpg-C3N4 materials are green. Combining the analytic results of UV–vis and XPS spectra, it can be further induced that the Co species could interact with the mpg-C3N4 materials. Namely, the loading of CoOx onto mpg-C3N4 is not merely via physical dispersion.

2.2. Catalyst Activity

The above materials were evaluated in the selective oxidation of EB using TBHP as an oxidant and acetonitrile as a solvent (Table 3). In the blank test without any catalysts, only 1.6% of EB is converted. The major product, BA accounting for 55.9%, originated from the deep oxidation of AP. After the addition of 100 mg of 3CoOx/g-C3N4, the conversion of EB is drastically improved, and the AP becomes the main product. Compared with 3CoOx/g-C3N4, 3CoOx/eg-C3N4 exhibits higher catalytic activity. Obviously, the supported CoOx is the key site in this selective oxidation and the eg-C3N4 support with the large surface area could facilitate the exposure of more catalytic sites. Besides g-C3N4 and eg-C3N4, the CoOx catalysts supported on carbon nanotube (CNT) and ordered mesoporous silica FDU-12 have also been prepared and examined under the same reaction conditions. Unfortunately, the corresponding conversions of EB and selectivity to AP are all less than the values gained in 3CoOx/eg-C3N4.
By contrast, the use of mpg-C3N4 supports the results of higher catalytic activity than eg-C3N4. Moreover, in the cases of 3CoOx/mpg-C3N4 catalysts, the EB conversions are found to be related to the supports. Among them, 3CoOx/mpg-C3N4-0.6 and 3CoOx/mpg-C3N4-0.8 demonstrate superior activity to 3CoOx/mpg-C3N4-0.4 and 3CoOx/mpg-C3N4-1.2. The obtained maximum conversion of EB is 62.0%, together with 84.7% of AP. The difference between various mpg-C3N4 supports is derived by adding the amounts of the siliceous templates (i.e., Ludox HS40). The N2 adsorption–desorption characterization results (Figure S5) confirm that all the 3CoOx/mpg-C3N4 catalysts are mesoporous materials, along with the relatively concentrated pore size distributions of ca. 10 nm. As listed in Table S2, both the surface areas and pore volumes of 3CoOx/mpg-C3N4-0.6 and 3CoOx/mpg-C3N4-0.8 are larger than those of the other two materials, probably responsible for their higher catalytic activities. Additionally, we evaluated the catalytic activity of 3CoOx/mpg-C3N4-0.6 using molecular oxygen as an oxidant instead of TBHP. As shown in Table 3, a small amount of EB (10.7%) can be also converted.
In addition to the preparation conditions of the mpg-C3N4 support, we have also prepared a series of 3CoOx/mpg-C3N4 with various heating temperatures (300–500 °C), and the corresponding catalytic performances are summarized in Table 4. It can be seen that the EB conversion and the selectivity to AP increase alongside the elevation of the heating temperature. As revealed in the above XPS analytical results (Table 2), among the several 3CoOx/mpg-C3N4-T (350–450 °C) materials, 3CoOx/mpg-C3N4-400 has the highest percentage of Co (II) cation and largest surface area. As reported by Jie and Yang et al. [29,43], Co (II) instead of Co (III) is proposed as the crucial catalytic active site in the oxidation of EB. Alongside Co (II), the abundant Nb species are able to anchor Co (II) and restrain its further oxidation. For the 3CoOx/mpg-C3N4 materials prepared in higher temperatures, the obtained EB conversions are much higher than the values gained by 3CoOx/mpg-C3N4-400. Nevertheless, it should be noted that due to the high preparation temperature, the masses of the two catalysts were indeed very low (Table 1).
Table S3 shows the catalytic performances of CoOx/mpg-C3N4 catalysts with various loading amounts of CoOx. Adopting a high loading amount results in an obvious increase in EB conversion along with selectivity to AP. However, as the loading amount is greater than 5 wt%, no significant improvement in catalytic conversion was received. This might be due to the possible agglomeration of the CoOx species on the mpg-C3N4 support.
The 3CoOx/mpg-C3N4 catalyst was chosen as a representative to further explore the effect of the reaction conditions on the catalytic performance. As shown in Figure 9A, under a reaction temperature of 80 °C, the EB conversion is only 28.8% and the selectivity to AP is 73.7%. At 120 °C, the conversion could reach 98.6% but the selectivity levels off. As mentioned above, using molecular oxygen instead of TBHP, the corresponding conversion of EB could also reach up to 10.7% (Table 3). The much higher conversion acquired at 120 °C could be due to the corporative oxidation of EB by TBHP and oxygen. The influence of reaction time on the catalytic performance is plotted in Figure 9B. It can be seen that prolonging the reaction could achieve higher catalytic conversions of EB. However, after a reaction of 10 h, the conversion exhibits an equilibrium. That is, as time increases, the conversion does not increase progressively. Figure S6 describes the catalytic performances with various feeding doses of 3CoOx/mpg-C3N4. As the weight of the catalyst is greater than 100 mg, the conversion shows no obvious improvement.
The recycling test has also been conducted to test the recyclability of the 3CoOx/mpg-C3N4. After four consecutive runs, the EB conversion and selectivity to AP are 60.3% and 81.2%, respectively (Table 3). Namely, there is only a slight loss in catalytic activity after several runs. The physicochemical properties of the recycled catalyst were analyzed by N2 adsorption–desorption and XPS, and the corresponding results confirm that the surface areas (Table 1) and surface chemical compositions (Table S1 and Table 2) of the catalyst have not undergone apparent change after the recycling experiments, suggesting good recyclability and stability of the catalyst in the selective oxidation of EB.
Table 5 lists the recently reported catalytic performances of the supported metal catalysts (heterogeneous system) in the selective oxidation of EB to AP by TBHP. The catalytic active metals mainly include Pd, Co, Cu, and Mn. From the viewpoint of reaction conditions, the reaction temperatures are basically in the range of 80–100 °C. Overall, the most efficient catalysts are the supported precious catalysts, which can convert EB smoothly under 80 °C. However, the major shortcoming in the use of such catalysts might be their high cost. For the other metal catalysts, their catalyst supports are mainly functionalized mesoporous materials, including SBA-15 [2], SBA-1 [44], TUD-1 [45], and organometallic-SiO2/Al2O3 [46], where the preparation of these materials requires expensive organic surfactants as soft templates and the whole processes for the final supported catalysts are relatively complicated. In terms of catalytic performance, the 3CoOx/mpg-C3N4 catalyst in this work demonstrates moderate catalytic activity under relatively mild reaction conditions. In this sense, the present work provides a convenient and low-cost heterogeneous catalyst for atmospheric and efficient selective oxidation of EB to AP.

3. Experimental Section

3.1. Catalyst Preparation

3.1.1. Synthesis of Mpg-C3N4

Mesoporous C3N4 materials were prepared according to the previous hard-templating approach reported by Goettmann et al. [48]. In brief, 4 g of cyanamide was added to 4–12 g of a colloidal silica particle (~12 nm) dispersion (Ludox HS40, 40 wt%, Sigma-Aldrich, Shanghai, China), stirred for 2 h, and ultrasonicated for 20 min. The mixture was heated at 60 °C until completely dried. Next, the white solid was ground, transferred into a crucible with a lid, heated with a ramping rate of 2.5 °C∙min−1 to reach 550 °C, and tempered for this temperature for another 3 h. The obtained pale-yellow powder was labeled as mpg-C3N4-SiO2.
Mpg-C3N4-SiO2 was added to 200 mL of NH4HF2 (Aladdin, Shanghai, China) aqueous solution (4 mol∙L−1) and stirred at room temperature for 24 h to remove the silica template. After that, the suspension was filtrated and rinsed with water several times until the pH value of the supernatant was close to 7. The obtained yellow solid was further rinsed with ethanol two times, and dried overnight at 60 °C. The resultant solid was designated as mpg-C3N4-r, where r indicated the ratio of silica in Ludox to cyanamide.

3.1.2. Preparation of CoOx/Mpg-C3N4

Mpg-C3N4 (600 mg) was added to 20 mL of cobalt nitrate aqueous solution containing 30–207 mg of Co(NO3)2∙6H2O (Aladdin, Shanghai, China). The suspension was ultrasonicated for 30 min, stirred for 3 h, and heated with stirring to 80 °C, until completely dried. The obtained, orange-colored solid was ground, transferred into a silica boat in a tubular furnace with N2 flow (ca. 20 mL∙min−1), and heated at a rate of 3 °C∙min−1 to a desired temperature (300–500 °C), and then, tempered at the temperature for another 3 h. The resultant dark green powder was labeled as mCoOx/mpg-C3N4-r-T, where r and T indicated the loading amount of Co and heating temperature, respectively. Unless otherwise specified, r and T were 0.6 and 400, respectively.

3.1.3. Preparation of Other Supported CoOx Catalysts

Similar to the preparation procedure of CoOx/mpg-C3N4, other supported CoOx catalysts were prepared by using exfoliated g-C3N4 (eg-C3N4), carbon nanotube, and ordered mesoporous FDU-12 as catalytic supports. The preparation methods of eg-C3N4 and FDU-12 are described in the supplementary material.

3.2. Material Characterization

The specific surface areas and porous properties of the materials were analyzed by N2 adsorption–desorption isotherms at −196 °C using an ASAP 2020 (Micromeritics, Norcross, GA, US) instrument. Before the analysis, the samples were pretreated in a vacuum at 150 °C for 6 h. The surface areas were calculated using the Brunauer–Emmet–Teller (BET) method.
X-ray diffraction (XRD) patterns were recorded on a D/max 2500 X-ray diffractometer (Rigaku, Tokyo, Japan) using a graphite monochromator (40 kV, 40 mA) equipped with Ni-filtered Cu-Kα radiation.
Fourier-transform infrared (FT-IR) spectra were tested in a Tensor 27 (Bruker, Billerica, MA, US) spectrometer based on the transmission mode with a resolution of 4 cm−1. Each spectrum was based on 32 scans (4000–400 cm−1).
UV–vis diffuse reflectance spectra were carried out on a UV-3600 spectrophotometer (Shimadzu, Tokyo, Japan) using BaSO4 as a standard reference.
X-ray photoelectron spectroscopy (XPS) measurements were performed using an ESCALAB 250 XI spectrometer (Thermo–Fisher, Waltham, MA, US) working in a constant energy mode with Mg Kα radiation as the excitation source.
Scanning electron microscopy (SEM) images were obtained on a Gemini SEM 300 (ZEISS, Oberkochen, Germany) microscope.
Transmission electron microscopy (TEM) experiments were conducted with a Talos F200X G2 (FEI, Hillsboro, OR, USA) electron microscope.

3.3. Catalytic Evaluation

The atmospheric selective oxidation of EB was carried out in a three-necked round-bottom flask equipped with a condenser. EB (10 mmol) and acetonitrile (4 mL) were mixed well, followed by the addition of 100 mg of catalyst. After the mixture was heated up to 100 °C, under continuous stirring (500 rpm), 30 mmol of TBHP (70 wt%, a.q.) was added into the flask through a peristaltic pump (the feeding time lasted for ca. 30 min). The reaction proceeded for 1–12 h. During the process, a small amount of reaction mixture (ca. 0.1 mL) was collected periodically and centrifuged. The liquid phase was analyzed by a gas chromatograph (Shangdong Rainbow Chemical Co., Ltd., Zaozhuang, China) equipped with a capillary column (FFAP). The conversion (conv.) of EB and selectivity (sel.) to AP were calculated using an area-normalization method, and the detailed calculation equations were as follows:
C o n v . ( % ) = A AP × f AP + A PE × f PE + A BA × f BA A EB × f EB + A AP × f AP + A PE × f PE + A BA × f BA
S e l . ( % ) = A AP × f AP A AP × f AP + A PE × f PE + A BA × f BA
where A and f were the peak area and response factor, respectively, for each component analyzed by GC. BA and PE stood for benzaldehyde and phenethyl alcohol, respectively.

4. Conclusions

In summary, mpg-C3N4 material was prepared using cyanamide as a precursor and colloidal silica as the template, which was then used as catalyst support to load cobalt oxide (CoOx). The characterization results show that the mesoporous structures of mpg-C3N4 remained after the incorporation of CoOx. Co (II) species disperse on the surface of the support and there is a probable interaction between Co (II) and the nitrogen species in mpg-C3N4. The heating temperatures during the catalyst preparation can adjust the distributions of the Co cations and nitrogen species. In the selective oxidation of EB by TBHP, the 3CoOx/mpg-C3N4-400 catalyst exhibited higher catalytic activity than other catalysts.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/catal13050828/s1, Preparation of eg-C3N4 and FDU-12; Figure S1: Particle size histogram of CoOx in 3CoOx/mpg-C3N4-400; Figure S2: FT-IR spectra of mpg-C3N4 (a), 3CoOx/mpg-C3N4-300 (b), 3CoOx/mpg-C3N4-350 (c), 3CoOx/mpg-C3N4-400 (d), 3CoOx/mpg-C3N4-450 (e), and 3CoOx/mpg-C3N4-500 (f) materials; Figure S3: FT-IR spectra of mCoOx/mpg-C3N4 materials; Figure S4: UV–vis spectra of mpg-C3N4 and CoOx/mpg-C3N4 materials; Figure S5: N2 adsorption–desorption isotherms (A) of 3CoOx/mpg-C3N4-r materials and the corresponding pore size distributions (B); Figure S6: Catalytic performances with various feeding doses of 3CoOx/mpg-C3N4. Reaction conditions:10 mmol of EB, 30 mmol of TBHP, 4 mL of acetonitrile, T = 100 °C, and t = 10 h; Table S1: Molar ratios of C/N of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials; Table S2: Surface areas and porous properties of 3CoOx/mpg-C3N4-r materials; Table S3: Catalytic performances of various mCoOx/mpg-C3N4 catalysts in the selective oxidation of EB.

Author Contributions

Conceptualization and methodology, J.X.; writing—original draft preparation, Y.Z.; writing—review and editing, J.X. and B.X.; investigation and test, Y.Z., X.-W.Z. and F.W. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (21878027).

Data Availability Statement

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

Acknowledgments

This work was supported by Advanced Catalysis and Green Manufacturing Collaborative Innovation Center.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chaudhary, V.; Sharma, S. Study of ethylbenzene oxidation over polymer-silica hybrid supported Co (II) and Cu (II) complexes. Catal. Today 2021, 375, 601–613. [Google Scholar] [CrossRef]
  2. Chaudhary, V.; Sharma, S. Synthesis of polymer-silica hybrid-supported catalysts for solvent-free oxidation of ethylbenzene with TBHP. Asia Pac. J. Chem. Eng. 2020, 15, e2441. [Google Scholar] [CrossRef]
  3. Nandanwar, S.U.; Rathod, S.; Bansal, V.; Bokade, V.V. A review on selective production of acetophenone from oxidation of ethylbenzene over heterogeneous catalysts in a decade. Catal. Lett. 2021, 151, 2116–2131. [Google Scholar] [CrossRef]
  4. Wang, K.; Zhao, S.; Ma, L.; Yang, M.; Qin, J.; Huang, X.; Gong, L.; Xiong, Y.; Li, R. A 3D-honeycomb-like catalyst: A nitrogen-doped carbon material with cobalt and manganese-oxide for C–H bond oxidation. Inorg. Chem. Front. 2019, 6, 3270–3287. [Google Scholar] [CrossRef]
  5. Liu, J.; Wang, W.; Jian, P.; Wang, L.; Yan, X. Promoted selective oxidation of ethylbenzene in liquid phase achieved by hollow CeVO4 microspheres. J. Colloid Interface Sci. 2022, 614, 102–109. [Google Scholar] [CrossRef]
  6. Selvaraj, M.; Park, D.W.; Kim, I.; Kawi, S.; Ha, C.S. Highly active mesoporous chromium silicate catalysts in side-chain oxidation of alkylaromatics. Dalton Trans. 2012, 41, 14204–14210. [Google Scholar] [CrossRef] [PubMed]
  7. Raji, V.; Chakraborty, M.; Parikh, P.A. Catalytic Performance of Silica-Supported Silver Nanoparticles for Liquid-Phase Oxidation of Ethylbenzene. Ind. Eng. Chem. Res. 2012, 51, 5691–5698. [Google Scholar] [CrossRef]
  8. Nilforoushan, S.; Ghiaci, M.; Hosseini, S.M.; Laurent, S.; Muller, R.N. Selective liquid phase oxidation of ethyl benzene to acetophenone by palladium nanoparticles immobilized on a g-C3N4–rGO composite as a recyclable catalyst. New J. Chem. 2019, 43, 6921–6931. [Google Scholar] [CrossRef]
  9. Li, Y.; Jie, S.; Li, K.; Liu, Z. Synthesis of efficient Co and N co-doped carbon catalysts with high surface areas for selective oxidation of ethylbenzene. New J. Chem. 2018, 42, 12677–12683. [Google Scholar] [CrossRef]
  10. Zhang, H.; Cai, A.; He, H.; Zhang, Q.; Zhang, F.; Zhang, G.; Fan, X.; Peng, W.; Li, Y. Nitrogen-doped 3D hollow carbon spheres for efficient selective oxidation of C–H bonds under mild conditions. New J. Chem. 2022, 46, 9727–9734. [Google Scholar] [CrossRef]
  11. Wang, X.; Blechert, S.; Antonietti, M. Polymeric Graphitic Carbon Nitride for Heterogeneous Photocatalysis. ACS Catal. 2012, 2, 1596–1606. [Google Scholar] [CrossRef]
  12. Wang, Y.; Di, Y.; Antonietti, M.; Li, H.; Chen, X.; Wang, X. Excellent Visible-Light Photocatalysis of Fluorinated Polymeric Carbon Nitride Solids. Chem. Mater. 2010, 22, 5119–5121. [Google Scholar] [CrossRef]
  13. Chakraborty, I.; Ghosh, N.; Ghosh, D.; Dubey, B.; Pradhan, D.; Ghangrekar, M. Application of synthesized porous graphitic carbon nitride and its composite as excellent electrocatalysts in microbial fuel cell. Int. J. Hydrogen Energy 2020, 45, 31056–31069. [Google Scholar] [CrossRef]
  14. Wang, P.; Xia, K.; Chen, Y.; Tian, Q.; Xiong, R.; Han, B.; Gao, Q.; Zhou, C.; Yu, D. Acid-assisted synthesis of nitrogen-deficient mesoporous graphitic carbon nitride for hydrogen storage. Mater. Lett. 2021, 301, 130347. [Google Scholar] [CrossRef]
  15. Wu, M.; Yan, J.; Tang, X.; Zhao, M.; Jiang, Q. Synthesis of Potassium-Modified Graphitic Carbon Nitride with High Photocatalytic Activity for Hydrogen Evolution. ChemSusChem 2014, 7, 2654–2658. [Google Scholar] [CrossRef] [PubMed]
  16. Guo, Y.; Chen, T.; Liu, Q.; Zhang, Z.; Fang, X. Insight into the Enhanced Photocatalytic Activity of Potassium and Iodine Codoped Graphitic Carbon Nitride Photocatalysts. J. Phys. Chem. C 2016, 120, 25328–25337. [Google Scholar] [CrossRef]
  17. Chen, D.; Wang, K.; Ren, T.; Ding, H.; Zhu, Y. Synthesis and characterization of the ZnO/mpg-C3N4 heterojunction photocatalyst with enhanced visible light photoactivity. Dalton Trans. 2014, 43, 13105–13114. [Google Scholar] [CrossRef]
  18. Wang, X.; Chen, X.; Thomas, A.; Fu, X.; Antonietti, M. Metal-Containing Carbon Nitride Compounds: A New Functional Organic-Metal Hybrid Material. Adv. Mater. 2009, 21, 1609–1612. [Google Scholar] [CrossRef]
  19. Ghafuri, H.; Jafari, G.; Rashidizadeh, A.; Manteghi, F. Co2+ immobilized on highly ordered mesoporous graphitic carbon nitride (ompg-C3N4/Co2+) as an efficient and recyclable heterogeneous catalyst for one-pot tandem selective photo-oxidation/Knoevenagel condensation. Mol. Catal. 2019, 475, 110491. [Google Scholar] [CrossRef]
  20. Yang, C.; Fu, L.; Zhu, R.; Liu, Z. Influence of cobalt species on the catalytic performance of Co-NC/SiO2 for ethylbenzene oxidation. Phys. Chem. Chem. Phys. 2016, 18, 4635–4642. [Google Scholar] [CrossRef]
  21. Chen, Z.; Mitchell, S.; Vorobyeva, E.; Leary, R.K.; Hauert, R.; Furnival, T.; Ramasse, Q.M.; Thomas, J.M.; Midgley, P.A.; Dontsova, D.; et al. Stabilization of Single Metal Atoms on Graphitic Carbon Nitride. Adv. Funct. Mater. 2017, 27, 1605785. [Google Scholar] [CrossRef]
  22. Gong, Y.; Zhang, P.; Xu, X.; Li, Y.; Li, H.; Wang, Y. A novel catalyst Pd@ompg-C3N4 for highly chemoselective hydrogenation of quinoline under mild conditions. J. Catal. 2013, 297, 272–280. [Google Scholar] [CrossRef]
  23. Yi, X.-T.; Zhao, T.; Wang, F.; Xu, J.; Xue, B. Palladium Nanoparticles Supported on Exfoliated g-C3N4 as Efficient Catalysts for Selective Oxidation of Benzyl Alcohol by Molecular Oxygen. New J. Chem. 2021, 45, 13519–13528. [Google Scholar] [CrossRef]
  24. 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]
  25. Zhang, Y.; Mori, T.; Ye, J. Polymeric Carbon Nitrides: Semiconducting Properties and Emerging Applications in Photocatalysis and Photoelectrochemical Energy Conversion. Sci. Adv. Mater. 2012, 4, 282–291. [Google Scholar] [CrossRef]
  26. Thomas, A.; Fischer, A.; Goettmann, F.; Antonietti, M.; Müller, J.-O.; Schlögl, R.; Carlsson, J.M. Graphitic carbon nitride materials: Variation of structure and morphology and their use as metal-free catalysts. J. Mater. Chem. 2008, 18, 4893–4908. [Google Scholar] [CrossRef]
  27. Xu, J.; Wu, H.-T.; Wang, X.; Xue, B.; Li, Y.-X.; Cao, Y. A new and environmentally benign precursor for the synthesis of mesoporous g-C3N4 with tunable surface area. Phys. Chem. Chem. Phys. 2013, 15, 4510–4517. [Google Scholar] [CrossRef]
  28. Shi, S.; Chen, C.; Wang, M.; Ma, J.; Gao, J.; Xu, J. Mesoporous strong base supported cobalt oxide as a catalyst for the oxidation of ethylbenzene. Catal. Sci. Technol. 2014, 4, 3606–3610. [Google Scholar] [CrossRef]
  29. Jie, S.; Yang, C.; Chen, Y.; Liu, Z. Facile synthesis of ultra-stable Co-N-C catalysts using cobalt porphyrin and peptides as precursors for selective oxidation of ethylbenzene. Mol. Catal. 2018, 458, 1–8. [Google Scholar] [CrossRef]
  30. Jie, S.; Lin, X.; Chao, Z.; Liu, Z. Effective ternary copper-cerium-cobalt catalysts synthesized via a modified pechini method for selective oxidation of ethylbenzene. Mater. Chem. Phys. 2018, 214, 239–246. [Google Scholar] [CrossRef]
  31. Jiang, Y.; Zhang, C.; Li, Y.; Jiang, P.; Jiang, J.; Leng, Y. Solvent-free aerobic selective oxidation of hydrocarbons catalyzed by porous graphitic carbon encapsulated cobalt composites. New J. Chem. 2018, 42, 16829–16835. [Google Scholar] [CrossRef]
  32. Gao, L.; Zhuge, W.; Feng, X.; Sun, W.; Sun, X.; Zheng, G. Co/rGO synthesized via the alcohol-thermal method as a heterogeneous catalyst for the highly efficient oxidation of ethylbenzene with oxygen. New J. Chem. 2019, 43, 8189–8194. [Google Scholar] [CrossRef]
  33. Fu, L.; Chen, Y.; Zhao, S.; Liu, Z.; Zhu, R. Sulfur-mediated synthesis of N-doped carbon supported cobalt catalysts derived from cobalt porphyrin for ethylbenzene oxidation. RSC Adv. 2016, 6, 19482–19491. [Google Scholar] [CrossRef]
  34. Zhang, L.; Jie, S.; Liu, Z. Bicontinuous mesoporous Co, N co-doped carbon catalysts with high catalytic performance for ethylbenzene oxidation. New J. Chem. 2019, 43, 7275–7281. [Google Scholar] [CrossRef]
  35. Oh, Y.; Hwang, J.O.; Lee, E.; Yoon, M.; Le, V.; Kim, Y.; Kim, D.H.; Kim, S.O. Divalent Fe Atom Coordination in Two-Dimensional Microporous Graphitic Carbon Nitride. ACS Appl. Mater. Interfaces 2016, 8, 25438–25443. [Google Scholar] [CrossRef]
  36. 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]
  37. Yu, Z.-H.; Gan, Y.-L.; Xu, J.; Xue, B. Direct Catalytic Hydroxylation of Benzene to Phenol Catalyzed by FeCl3 Supported on Exfoliated Graphitic Carbon Nitride. Catal. Lett. 2020, 150, 301–311. [Google Scholar] [CrossRef]
  38. Zhu, J.; Wei, Y.; Chen, W.; Zhao, Z.; Thomas, A. Graphitic carbon nitride as a metal-free catalyst for NO decomposition. Chem. Commun. 2010, 46, 6965–6967. [Google Scholar] [CrossRef]
  39. Bahuguna, A.; Kumar, A.; Chhabra, T.; Kumar, A.; Krishnan, V. Potassium-Functionalized Graphitic Carbon Nitride Supported on Reduced Graphene Oxide as a Sustainable Catalyst for Knoevenagel Condensation. ACS Appl. Nano Mater. 2018, 1, 6711–6723. [Google Scholar] [CrossRef]
  40. 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]
  41. Chen, P.-W.; Li, K.; Yu, Y.-X.; Zhang, W.-D. Cobalt-doped graphitic carbon nitride photocatalysts with high activity for hydrogen evolution. Appl. Sur. Sci. 2017, 392, 608–615. [Google Scholar] [CrossRef]
  42. Yue, B.; Li, Q.; Iwai, H.; Kako, T.; Ye, J. Hydrogen production using zinc-doped carbon nitride catalyst irradiated with visible light. Sci. Technol. Adv. Mater. 2011, 12, 034401. [Google Scholar] [CrossRef] [PubMed]
  43. Yang, C.; Jie, S.; Zhu, R.; Zhang, N.; Wang, J.; Liu, Z. Co-N-C Catalysts Synthesized via Pyrolyzing the Ionic Liquids Solution Dissolved with Casein and Cobalt Porphyrin for Ethylbenzene Oxidation. ChemistrySelect 2017, 2, 4255–4260. [Google Scholar] [CrossRef]
  44. Imran, G.; Maheswari, R. Mn-incorporated SBA-1 cubic mesoporous silicates: Synthesis and characterization. Mater. Chem. Phys. 2015, 161, 237–242. [Google Scholar] [CrossRef]
  45. Pachamuthu, M.P.; Rajalakshmi, R.; Maheswari, R.; Ramanathan, A. Direct glycol assisted synthesis of an amorphous mesoporous silicate with framework incorporated Co2+: Characterization and catalytic application in ethylbenzene oxidation. RSC Adv. 2014, 4, 29909–29916. [Google Scholar] [CrossRef]
  46. Habibi, D.; Faraji, A.R.; Arshadi, M.; Fierro, J.L.G. Characterization and catalytic activity of a novel Fe nano-catalyst as efficient heterogeneous catalyst for selective oxidation of ethylbenzene, cyclohexene, and benzylalcohol. J. Mol. Catal. A 2013, 372, 90–99. [Google Scholar] [CrossRef]
  47. Kalita, L.; Saikia, L. Palladium-Supported Nanoceria: A Highly Efficient Catalyst for Solvent-Free Selective Oxidation of Ethylbenzene to Acetophenone. ChemistrySelect 2020, 5, 4848–4855. [Google Scholar] [CrossRef]
  48. Goettmann, F.; Fischer, A.; Antonietti, M.; Thomas, A. Chemical synthesis of mesoporous carbon nitrides using hard templates and their use as a metal-free catalyst for Friedel-Crafts reaction of benzene. Angew. Chem. Int. Ed. 2006, 45, 4467–4471. [Google Scholar] [CrossRef]
Figure 1. XRD patterns of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
Figure 1. XRD patterns of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
Catalysts 13 00828 g001
Figure 2. N2 adsorption–desorption isotherms (A) of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials and the corresponding pore size distributions (B).
Figure 2. N2 adsorption–desorption isotherms (A) of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials and the corresponding pore size distributions (B).
Catalysts 13 00828 g002
Figure 3. SEM images of mpg-C3N4 (A) and 3CoOx/mpg-C3N4 (B).
Figure 3. SEM images of mpg-C3N4 (A) and 3CoOx/mpg-C3N4 (B).
Catalysts 13 00828 g003
Figure 4. EDX mapping of 3CoOx/mpg-C3N4.
Figure 4. EDX mapping of 3CoOx/mpg-C3N4.
Catalysts 13 00828 g004
Figure 5. TEM images of mpg-C3N4 (A) and 3CoOx/mpg-C3N4 (B).
Figure 5. TEM images of mpg-C3N4 (A) and 3CoOx/mpg-C3N4 (B).
Catalysts 13 00828 g005
Figure 6. XPS survey of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
Figure 6. XPS survey of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
Catalysts 13 00828 g006
Figure 7. Co 2p3/2 fine spectra of 3CoOx/mpg-C3N4-T materials.
Figure 7. Co 2p3/2 fine spectra of 3CoOx/mpg-C3N4-T materials.
Catalysts 13 00828 g007
Figure 8. N 1s fine spectra of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
Figure 8. N 1s fine spectra of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
Catalysts 13 00828 g008
Figure 9. Catalytic performances of 3CoOx/mpg-C3N4 under various reaction temperatures (A) and reaction time (B). Reaction conditions: 10 mmol of EB, 30 mmol of TBHP, Wcatal. = 100 mg, 4 mL of acetonitrile, t = 10 h (A), and T = 100 °C (B).
Figure 9. Catalytic performances of 3CoOx/mpg-C3N4 under various reaction temperatures (A) and reaction time (B). Reaction conditions: 10 mmol of EB, 30 mmol of TBHP, Wcatal. = 100 mg, 4 mL of acetonitrile, t = 10 h (A), and T = 100 °C (B).
Catalysts 13 00828 g009
Table 1. Specific surface areas and porous properties of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
Table 1. Specific surface areas and porous properties of mpg-C3N4 and 3CoOx/mpg-C3N4-T materials.
SampleSBET (m2·g−1)Pore Size (nm) aPore Volume (cm3·g−1)Mass (g) b
mpg-C3N48413.20.302.44
3CoOx/mpg-C3N4-300448.80.150.57
3CoOx/mpg-C3N4-3503210.20.150.53
3CoOx/mpg-C3N4-4007010.20.260.45
3CoOx/mpg-C3N4-4505610.50.870.35
3CoOx/mpg-C3N4-5006012.50.320.24
3CoOx/mpg-C3N4-400-R7712.10.28
a Determined by the adsorption branches. b Mass of the synthesized materials.
Table 2. Molar percentages of various cobalt and nitrogen species a.
Table 2. Molar percentages of various cobalt and nitrogen species a.
MaterialCo (II)Co (III)Co0NaNbNc
mpg-C3N474.114.011.9
3CoOx/mpg-C3N4-35032.762.05.370.120.19.8
3CoOx/mpg-C3N4-40041.556.02.464.825.59.7
3CoOx/mpg-C3N4-45035.361.63.168.022.99.1
3CoOx/mpg-C3N4-400R37.659.23.266.024.99.1
a Calculated by deconvoluted Co 2p3/2 and N 1s spectra.
Table 3. Catalytic performances of various catalysts in the selective oxidation of EB a.
Table 3. Catalytic performances of various catalysts in the selective oxidation of EB a.
CatalystCon. (%)Sel. (%)
APPEBA
/1.626.617.555.9
3CoOx/g-C3N427.675.214.510.3
3CoOx/eg-C3N432.974.69.515.9
3CoOx/CNT19.663.518.817.7
3CoOx/FDU-128.671.911.216.9
3CoOx/mpg-C3N4-0.437.879.46.813.8
3CoOx/mpg-C3N4-0.662.084.74.810.6
3CoOx/mpg-C3N4-0.857.483.65.211.8
3CoOx/mpg-C3N4-1.235.975.515.29.3
3CoOx/mpg-C3N4-0.6 b10.767.827.15.1
3CoOx/mpg-C3N4-0.6-R60.381.25.912.9
a Reaction conditions: 10 mmol of EB, 30 mmol of TBHP, Wcatal. = 100 mg, 4 mL of acetonitrile, T = 100 °C, and t = 10 h. b Reaction conditions: 10 mmol of EB, O2 (10 mL∙min−1), Wcatal. = 100 mg, 4 mL of acetonitrile, T = 120 °C, and t = 10 h.
Table 4. Catalytic performances of various 3CoOx/mpg-C3N4-T catalysts in the selective oxidation of EB a.
Table 4. Catalytic performances of various 3CoOx/mpg-C3N4-T catalysts in the selective oxidation of EB a.
CatalystCon. (%)Sel. (%)
APPEBA
3CoOx/mpg-C3N4-30011.769.27.623.3
3CoOx/mpg-C3N4-35023.869.09.121.9
3CoOx/mpg-C3N4-40062.084.74.810.6
3CoOx/mpg-C3N4-45071.190.02.87.2
3CoOx/mpg-C3N4-50076.691.41.86.8
a Reaction conditions: 10 mmol of EB, 30 mmol of TBHP, Wcatal. = 100 mg, 4 mL of acetonitrile, T = 100 °C, and t = 10 h.
Table 5. Comparison of catalytic performances of supported metal catalysts in selective oxidation of EB by TBHP.
Table 5. Comparison of catalytic performances of supported metal catalysts in selective oxidation of EB by TBHP.
CatalystnTBHP
(mmol)
nEB
(mmol)
Wcatal.
(mg)
T
(°C)
t
(h)
Conv.
(%)
Sel.
(%)
Pd/g-C3N4–rGO [8]411080246797
Pd/CeO2 [47]112080410079
Co-Cu/SAPS-15 [2]301030100697100
MnSBA-1 [44]10 1008082057
CoTUD-1 [45]10101008083874
SiO2/Al2O3-APTMS [46]995050242774
LDH-[NAPABA–Cu(II)] 3913100120781100
3CoOx/mpg-C3N4 3010100100106285
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Zhu, Y.; Zhang, X.-W.; Wang, F.; Xue, B.; Xu, J. Atmospheric and Efficient Selective Oxidation of Ethylbenzene Catalyzed by Cobalt Oxides Supported on Mesoporous Carbon Nitride. Catalysts 2023, 13, 828. https://doi.org/10.3390/catal13050828

AMA Style

Zhu Y, Zhang X-W, Wang F, Xue B, Xu J. Atmospheric and Efficient Selective Oxidation of Ethylbenzene Catalyzed by Cobalt Oxides Supported on Mesoporous Carbon Nitride. Catalysts. 2023; 13(5):828. https://doi.org/10.3390/catal13050828

Chicago/Turabian Style

Zhu, Ye, Xue-Wen Zhang, Fei Wang, Bing Xue, and Jie Xu. 2023. "Atmospheric and Efficient Selective Oxidation of Ethylbenzene Catalyzed by Cobalt Oxides Supported on Mesoporous Carbon Nitride" Catalysts 13, no. 5: 828. https://doi.org/10.3390/catal13050828

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