Next Article in Journal
Thermal Creep Behavior and Creep Crystallization of Al-Mg-Si Aluminum Alloys
Previous Article in Journal
Evaluation of the Fatigue Behavior and Failure Mechanisms of 4340 Steel Coated with WIP-C1 (Ni/CrC) by Cold Spray
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Nitrogen-Doped Porous Carbon from Biomass with Efficient Toluene Adsorption and Superior Catalytic Performance

1
School of Energy Science and Engineering, Central South University, Changsha 410083, China
2
Hunan Ecological and Environmental Affairs Center, Changsha 410014, China
*
Authors to whom correspondence should be addressed.
Materials 2022, 15(22), 8115; https://doi.org/10.3390/ma15228115
Submission received: 8 October 2022 / Revised: 4 November 2022 / Accepted: 14 November 2022 / Published: 16 November 2022
(This article belongs to the Section Catalytic Materials)

Abstract

:
The chemical composition and surface groups of the carbon support affect the adsorption capacity of toluene. To investigate the effect of catalyst substrate on the catalytic performance, two different plant biomasses, banana peel and sugarcane peel, were used as carbon precursors to prepare porous carbon catalyst supports (Cba, Csu, respectively) by a chemical activation method. After decorating PtCo3 nanoparticles onto both carbon supports (Cba, Csu), the PtCo3-su catalyst demonstrated better catalytic performance for toluene oxidation (T100 = 237 °C) at a high space velocity of 12,000 h−1. The Csu support possessed a stronger adsorption capacity of toluene (542 mg g−1), resulting from the synergistic effect of micropore volume and nitrogen-containing functional groups, which led to the PtCo3-su catalyst exhibiting a better catalytic performance. Moreover, the PtCo3-su catalyst also showed excellent stability, good water resistance properties, and high recyclability, which can be used as a promising candidate for practical toluene catalytic combustion.

1. Introduction

Growing volatile organic compounds (VOCs) from combustion and many different industry processes have aroused more concern for their damage to the atmospheric environment and human health [1,2]. As the most common volatile organic compound in the air, toluene is widely used as a solvent for paints, coatings, and adhesives and an important ingredient in various industrial and household products [3]. Long-term exposure to toluene may cause fatal harm to the human respiratory system, nervous system, and hematopoietic system. Therefore, there is an imminent need to develop technologies to dispel poisonous VOCs.
Many removal techniques are presented to improve air quality, including carbon adsorption [4], membrane separation [5], photocatalytic oxidation [6], non-thermal plasma oxidation [7], biological treatment [8], and catalytic oxidation [9,10,11]. Catalytic combustion is considered one of the most progressive and practical technologies for VOCs removal on account of its high efficiency, stability, and reliability. Supported noble metal catalysts (Pd and Pt) could achieve satisfactory catalytic activity under low temperatures. For example, Chen et al. synthesized a series of size-controllable Pt nanoparticles supported on ZSM-5. The Pt particles of 1.9 nm size could completely oxidize toluene at 155 °C [12]. The TiO2-supported 0.20 wt% Pt catalyst prepared by Wang et al. showed a high catalytic performance for toluene oxidation with T50 and T90 at 173 °C and 183 °C, respectively [13]. However, the high price of noble metal catalysts made them unable to be extensively used, and researchers were committed to finding alternatives with excellent performance. Fortunately, the addition of a second metal (Mn, Co, Fe, and Cu) to Pt made it reasonably priced, which may be a good alternative for developing oxidation catalysts [14,15].
Besides, the catalytic activity of supported catalysts was inextricably linked to the support. Generally speaking, catalyst supports could be divided into the following two categories [16,17]: one was active support, which had a positive catalytic effect on the oxidative combustion of VOCs, such as CeO2, Co3O4, MnOx, etc. [18,19,20]; the other was inert support, which on its own did not contribute to the catalytic removal of VOCs (such as C, Al2O3, molecular sieves, ceramics, etc.) [21,22,23]. However, the research on the role of supported catalysts and their substrates in catalytic oxidation remained elusive and needed to be further explored. Among the supports, biomass materials have become very promising support candidates due to their low cost, abundant pore structure, and large specific surface area and excellent adsorption capacity [24,25]. Zhao et al. prepared shaddock peel porous carbon with a specific surface area of 2398.74 m2 g−1 and the maximum adsorption capacity of the sample for methylene blue was 869.57 mg g−1 [26]. Li et al. prepared N-doped porous carbon from biomass lotus root as a precursor to becoming an efficient CO2 adsorbent [27].
In this work, two different biomass wastes, banana peel and sugarcane peel, were selected as carbon sources, zinc chloride and basic magnesium carbonate as double templates, and urea as a nitrogen source to prepare biomass porous carbon. Then PtCo3 nanoparticles were modified on these carbon supports, and the catalytic performance of the alloy was evaluated on the catalytic reaction testbed to explore the effects of carbon supports with different microstructures and chemical compositions on the performance of the catalyst. Furthermore, with the aid of multiple characterizations, the structure-property relationship and the potential mechanism were explored.

2. Experimental Methodology

2.1. Raw Materials

Biomass raw materials, banana peel and sugarcane peel, were gathered from Central South University in China. ZnCl2 (>98%), urea, toluene (95%), and hexane (>97%) were acquired from Sinopharm. Magnesium carbonate basic was purchased from Macklin. Hydrochloric acid (HCl, 36–38%) came from Xingkong, China. Chemicals used for the synthesis of catalysts such as cobalt acetylacetonate (Co(acac)3, 98%), platinum acetylacetonate (Pt(acac)2, 97%), oleic acid (AR), tetradecenediol (TDD, 90%), dichlorobenzene (>99.9%), benzyl ether (95%) and oleylamine (80–90%) were obtained from Aladdin. All the gases used in the experiments, such as nitrogen (99.99%), oxygen (99.99%), etc., were provided by High-Tech gas.

2.2. Materials Preparation

Both biomass raw materials were carefully washed with distilled water and then dried in a blast drying oven at 100 °C for 10 h. The dried biomass materials were crushed in a high-speed multifunctional crusher (800Y, Kemanshi, Jinhua, China) for 5 min. Each biomass, urea, basic magnesium carbonate, and zinc chloride were mixed at a mass ratio of 2:1:1:1 and ground in a mortar for 15 min. Then the mixture was heated to 900 °C in N2 atmosphere at a heating rate of 5 °C min−1 and carbonized at 900 °C for 1 h. The carbonized samples were washed in hydrochloric acid (80 mL, 2.0 M) for 12 h, after which they were cleaned with distilled water until the pH was neutral, and then dried at 60 °C for 12 h. The acquired samples were named Cba, Csu, respectively (Figure 1).
The preparation of PtCo3 catalysts supported on biomass carbon was prepared by organic phase reduction method. The mixture of 0.58 mmol Co(acac)3, 0.2 mL oleylamine, 0.2 mL oleic acid, 100 mg TDD and 15 mL benzyl ether in a three-necked flask was heated to 100 °C in Ar atmosphere and maintained for 1 h to dislodge moisture. When the temperature rose to 200 °C, 0.1 mmol Pt(acac)2 dissolved in 0.7 mL dichlorobenzene were added to reaction vessel and the reaction continued for 1 h. When the solution was cooled to room temperature, 25 mL of absolute ethanol was added to the solution and centrifuged at 10,000 rpm for 10 min to obtain PtCo3 nanoparticles (NPs). In order to decorate PtCo3 NPs on the carbon supports, 30 mL hexane and biomass carbon were mixed with PtCo3 NPs, then the compound was sonicated for 30 min. Then the samples were centrifuged (10,000 revolutions, 10 min). The PtCo3-ba and PtCo3-su samples were obtained after drying in a vacuum drying chamber at 60 °C for 18 h.

2.3. Materials Characterization

The morphologies of biomass carbons were visualized by scanning electron microscope (SEM, JSM-7900F, JEOL, Tokyo, Japan). The components and functional groups of the sample were identified by X-ray photoelectron spectroscopy (XPS, Thermo Scientific K-Alpha, Waltham, MA, USA). The crystal phase type of the samples was explored by X-ray diffraction (XRD, Rigaku SmartLab SE, Tokyo, Japan). Transmission electron microscope (TEM, Titan G260-300, FEI, Hillsboro, OR, USA) was used to characterize the surface morphology of catalysts. The surface area and pore structure of the samples were determined by using a JW-BK132Z static volumetric analyzer (Beijing JWGB Sci & Tech Co., Ltd., Beijing, China). The thermal stability of as-prepared samples was studied by thermal gravimetric analysis (TGA5500, TA, New Castle, DE, USA). The in-situ DRIFT test was carried out by FT-IR spectrometer (Thermo Scientific Nicolet iS50, Waltham, MA, USA). Additionally, the content of Pt and Co in the samples was quantified by inductively coupled plasma mass spectrometry (ICP-MS, Agilent 7700, Santa Clara, CA, USA).

2.4. Catalytic Tests

The activity and thermal stability of the catalyst were evaluated on the catalytic performance test platform. In total, 30 mg of catalyst and 1 g of quartz sand were mixed equably and placed into the quartz tube, which can be heated to a certain temperature by an external heating furnace, as shown in Figure S1. Moreover, all ventilation ducts were exteriorized with an insulation layer set to reduce heat transfer and eliminate environmental disturbances. The total reaction gas flow was 160 mL min−1, containing x ppm (x = 500, 1000, 1500) of toluene, 20% O2, and nitrogen, giving the gas hourly space velocity (GHSV) at 12,000 h−1. Toluene in different concentrations in a gas mixture was produced by using an N2 aerator through a bottle filled with pure toluene liquid, which was kept at 0 °C by a thermostatic water bath. The reactants and products were quantitatively analyzed by using an online gas chromatograph (GC-9860-5CNJ, Haoerpu, Nanjing, China) equipped with a flame ionization detector (FID) and TCD. Mass flowmeters and injection pumps were used to control gas concentrations when studying the effects of water vapor on the catalyst, respectively.

2.5. Calculations

Density functional theory (DFT) calculations were conducted via the Perdew–Burke–Ernzerh (PBE) of the generalized gradient approximation (GGA) method, and Materials Studio was the software used in calculations [28]. The convergence parameters set in the model optimization were as follows: gradient 2 × 10−3 Ha/Å, energy 1 × 10−5 Ha, and displacement 5 × 10−3 Å. The toluene adsorption energies onto the pure graphite and N-doped surface were derived in accordance with Equation (1) as follows:
Eads = Esurface+toluene − (Esurface + Etoluene)
where Eads, Esurface+toluene, Esurface, Etoluene represents the adsorption energy and total energy of adsorbent-toluene complex, adsorbent surface, and free toluene, respectively.

3. Results and Discussion

3.1. Catalytic Performance

The catalytic performance of two PtCo3 catalysts for toluene combustion at 500, 1000, and 1500 ppm was assessed, and the results are shown in Figure 2. Two biomass carbon supports had no significant catalytic effect on the conversion of toluene (Figure S2) and possessed good thermal stability (Figure S3). When the toluene concentration was 500 ppm, toluene conversions of PtCo3-ba and PtCo3-su achieved 100% at 220 °C and 236 °C, respectively. The PtCo3-su and PtCo3-ba catalysts can completely remove toluene at 237 °C at a toluene concentration of 1000 ppm. Moreover, it was clear that the PtCo3-su sample can remove 91% of toluene at 226 °C, while that of the PtCo3-ba was 66%. When the concentration value of toluene was 1500 ppm, the PtCo3-su sample still showed a higher catalytic activity. The above results indicated that the PtCo3-su sample possessed an excellent catalytic performance for toluene at low reaction temperatures (Table S1).
Catalytic durability was also a momentous issue in its holistic performance [29]. The stability of the PtCo3-su catalyst in consecutive reactions was examined at a reaction temperature of 237 °C and a GHSV of 12,000 h−1, and the results are shown in Figure 3. The catalyst maintained over 95% conversion to toluene, exhibiting an outstanding catalytic stability within a 96 h reaction time.
Numerous studies indicated that the existence of vapor had a negative impact on catalytic properties, leading to catalyst poisoning and deactivation. Inspired by Qi et al. [30], the effects of water on PtCo3-su catalyst performance were studied when the toluene conversion was 80% (Figure 4). The experiments were performed within 18 h as one cycle with an intermittent injection of water vapor. Obviously, as the water vapor content increased from 3.5% to 20%, the harmful effect of water vapor on the catalytic behavior became stronger. However, from the perspective of overall efficiency, it could be considered that there was no significant impact within 84 h. Meanwhile, the capability for toluene oxidation was quickly restored to a satisfactory value when the water vapor was cut off. The above-mentioned results make it clear that the PtCo3-su catalyst had excellent properties against water vapor.

3.2. Characterization of Biomass Carbon

In terms of catalyst support, SEM was used to learn about the surface morphology of Cba and Csu. Both samples exhibited a rough surface and obvious pores (Figure S4). The N2 adsorption/desorption analysis was used to further explore the micropore structure of the prepared supports (Figure 5a).
The BET-specific surface area values and pore structure parameters of the supports were presented in Table 1. The two samples displayed a canonical type IV isotherm based on IUPAC classification [31,32,33], confirming the presence of mesopores. Additionally, the H3 hysteresis loop was observed in Cba and Csu. Figure 5b displays the pore size distribution (PSD) of two samples obtained from the non-local density functional theory (NLDFT) model [34,35]. The Csu sample had more micropores at 0.8–1.4 nm than the Cba sample. However, the Cba sample (1029 m2 g−1) possessed a higher BET-specific surface area than the Csu (960 m2 g−1).
Before loading alloy particles, the adsorption capacity of the support to toluene was explored, as shown in Figure 5c, which showed the order of PtCo3-su (542 mg g−1) > PtCo3-ba (387 mg g−1). Although the adsorption capacity of toluene was influenced by the BET-specific surface area and total pore volume, the micropore volume was the factor most closely related to it [36,37]. The micropore volume of the Csu was 0.377 cm3 g−1 while that of the Cba sample was 0.365 cm3 g−1. Besides, surface chemistry was also an important factor affecting the adsorption capacity [38]. As shown in Table 1, it was clear that the nitrogen content of Csu was higher than that of the Cba sample. Figure 5d,e shows the N1s spectrum and the four peaks obtained by its deconvolution. Four separated peaks were located at 398.3 eV, 399.8 eV, 400.9 eV, and 402.9 eV, respectively, referring to pyridinic-N, pyrrolic-N, graphitic-N, and oxidized-N [39]. Interestingly, the toluene adsorption capacity of Csu was approximately 1.4 times that of Cba, while the contents of pyridinic-N, pyrrolic-N, and graphitic-N showed a similar multiplier relationship. This could be interpreted in terms of the synergistic effect of micropore volume and nitrogen-containing functional groups. The microporous structure provided active sites for toluene adsorption, while the nitrogen functional group acted as the adsorption center and generated new active sites.
To further investigate the intrinsic mechanism of the interaction between the N-containing functional group with toluene, N-containing functional groups were assessed by DFT calculations. Figure 6 shows the optimized geometry and the binding energies (BEs) of different nitrogen-containing functional groups in the carbon framework. It can be seen that the adsorption energies of all N-containing functional groups for toluene were higher than those on the pristine graphite model (−44.71 kJ mol−1). For three types of N-containing functional groups, the greatest BEs were obtained by pyrrolic-N (−74.30 kJ mol−1), followed by graphitic-N (−59.63 kJ mol−1) and pyridinic-N (−51.72 kJ mol−1). These results evidenced that the introduction of nitrogen-containing functional groups strengthened the electrostatic interaction, specifically referring to the π electrons formed on biochar that could form cation—π bonds via the electrostatic attraction on the benzene ring, thereby increasing the affinity of carbon materials with toluene molecules, in which the pyrrolic-N exhibited the strongest promoting effect [40,41]. This again confirmed that higher nitrogen content was one reason for the stronger adsorption capacity of Csu to toluene.

3.3. Characterization of Catalyst

As shown in Figure 7 and Table 2, both PtCo3-ba and PtCo3-su exhibited a similar specific surface area (68 m2 g−1 and 74 m2 g−1). The PtCo3-su obtained more mesopores of 6∼15 nm, presenting a larger total pore volume of 0.176 cm3 g−1, while the value for the PtCo3-ba was 0.147 cm3 g−1. The above results showed that PtCo3 loading caused a reduction in surface area, which mainly occupied pores with a diameter of 0.5–4 nm for two catalysts. The toluene adsorption and desorption properties of samples were determined by the static adsorption method and C7H8-TPD. Consistent with the previous adsorption on the supports, the PtCo3-su sample possessed a higher adsorption. Figure 7d illustrates that the peak center of PtCo3-su fell in a higher temperature area than that of PtCo3-ba. Such differences presented that PtCo3-su and toluene possessed a stronger interaction with each other than PtCo3-ba. This trend showed good consistency with the toluene conversion temperature of both catalysts, which suggested that the adsorption properties of toluene on the substrates had a significant impact on the catalytic activity.
The porous structure of the substrate, particle size, and structure of bimetallic PtCo3 nanoparticles were investigated intensively by TEM techniques. As shown in Figure 8a,e, the well-dispersed alloy nanoparticles could be observed from TEM images of PtCo3-ba and PtCo3-su, where the black dots represented the PtCo3 NPs and the gray transparent areas represented the support. Figure 8b,f showed the results of the size distribution analysis, which indicated that the PtCo3 were mostly in the form of tiny nanoparticles with a diameter of 2–3 nm. Figure 8c,g showed the images of Fourier fast transform (FFT) and inverse Fourier fast transform (IFFT) of HRTEM, indicating the presence of (111) crystal planes in the catalyst. The lattice distance of metal particles in the PtCo3-ba and PtCo3-su catalyst was 2.18 Å and 2.19 Å, which lay between the interplane distance of Pt (111) (2.27 Å) and Co (111) (1.98 Å), proving the alloy phase formation [40]. The alloy structural feature was further identified by elemental mapping (Figure 8d,h). The actual metal contents of PtCo3-ba and PtCo3-su were acquired via ICP-MS (Table 2). Additionally, the data from ICP revealed that the ratio of Pt to Co atoms in both PtCo3-ba and PtCo3-su was close to 37:84, implying PtCo3 was successfully synthesized.
Figure 9a showed the XRD results of the support and two catalyst samples. For the former, the strong diffraction peak located approximately at 24.1° and the weak diffraction peak located at 43.8° corresponded to the (002) and (100) crystal facets of graphite crystals, respectively. After loading the metal, the characteristic peaks of Pt (111) (2θ of 39.5°) and Co (111) (2θ of 44.2°) were hardly observed in the two samples, and (002) crystal planes of graphite crystals were also hardly detected. The position of the peak changed to 42.1°, with a negative offset of graphite crystal and a positive shift of alloy, which may be the result of the combined action of (111) and (002) [30].
The XPS analysis was undertaken to find out the oxidation conditions of surface species, and the full XPS spectrum (Figure S5) of PtCo3-ba and PtCo3-su confirmed the presence of Pt, Co, C, N, and O. As indicated in Figure 9b, for the O 1s XPS spectra, two peaks at 532.2 and 533.8 eV corresponded to the surface oxygen species (Oads, O2−, O 2 2 , O, or CO 3 2 ) with the absence of lattice oxygen [42,43]. Moreover, there was no large difference in the content of oxygen elements in the two catalysts. According to the results in Table 2, when the oxygen content of the PtCo3-ba sample was 4.562, the PtCo3-su was 4.314. The surface oxygen species would be involved in the later toluene oxidation process, which was an important element impacting the oxidation reaction.
Figure 9c,d showed the Pt 4f and Co 2p XPS spectra of the samples. For Pt 4f, the two peaks located at 71.62 and 74.68 eV corresponded to the Pt 4f7/2 and Pt 4f5/2 orbitals, respectively. The binding energy peaks of Pt 4f7/2 for Pt0 and Pt2+ on powdered metals were at roughly 74.9 and 75.9 eV, while Pt 4f5/2 were at 71.5 and 72.4 eV [44]. It was noticeable that the Pt species on two PtCo3 samples were mainly Pt0. As for Co 2p, the peaks of Co0 were located at 780.0 eV and 795.4 eV, and those of Co2+ are at 782.0 eV and 797.3 eV. Besides, the peaks located at 786.3 eV and 802.9 eV refer to the satellite peaks of Co 2p1/2 and Co 2p3/2 [45,46]. Moreover, the peak positions of Pt 4f and Co 2p of the PtCo3-su sample after the reaction were essentially the same as the fresh sample.
Besides, the PtCo3-su samples that had been tested for water resistance or thermal stability were also characterized by XPS (Figure S6). It turned out that among the tested samples, the XPS spectra of platinum and cobalt still displayed no apparent difference with fresh samples, further confirming their stable physicochemical properties.

3.4. Apparent Activation Energy

In order to deeply investigate the catalytic activity of two samples, a number of experiments were carried out to test the internal reaction speed. The amount of catalyst was reduced (15 mg) to reach a <20% removal rate of toluene with the goal of dislodging mass and thermal transfer restrictions [47]. The apparent activation energy (Ea) of PtCo3 samples for toluene oxidation was determined by the Arrhenius formula, as follows [48]:
lnr = −Ea/R + lnA
where r is the reaction rate (mol s−1 g−1), Ea represents the apparent activation energy (kJ mol−1), R is a constant with the value 8.314 × 10−3, T refers to the absolute temperature in the reaction (K), while A is a pre-exponential factor. The reaction rate of toluene oxidation was obtained based on Equation (3) as follows [9]:
r = (CToluene,InCToluene,Out)/W
where CToluene,In and CToluene,Out represents the toluene concentration values at the inlet and outlet of the reaction device respectively (mol s−1), and W represents the weight of catalysts (g).
Figure 10 showed the Arrhenius diagram and Ea values of two samples. The Ea value (55.9 kJ mol−1) of the PtCo3-su catalyst was relatively lower than that of PtCo3-ba (63.4 kJ mol−1), thereby exhibiting a more superior catalytic activity, which was consistent with the results in Figure 2.

3.5. Reaction Mechanism

In situ DRIFTS were taken to elucidate the possible avenue of toluene oxidation, with the test consequences given in Figure 11a. The strong bands at around 1540 and 1508 cm−1 were caused by the skeletal vibration of aromatic rings [49]. The band at around 1749 cm−1 was relevant with C=O stretching vibration, which was a signal for the existence of species such as esters, aldehydes, or carboxylic acids [50]. Besides, the band at 1673 cm−1 can be assigned as the stretching vibration of the aldehydes, evidencing the production of benzaldehyde as the reaction proceeded [51]. The peak located at around 1789 cm−1 was assigned to COO stretching vibration, implying the presence of benzoate species. Along with the reaction, all intermediate components continued to be oxidized on the catalyst surface until they were finally converted into CO2 and H2O.
Figure 11b recapitulates the probable toluene oxidation mechanism on the PtCo3-su catalyst. Firstly, the involved toluene molecules were adsorbed on the surface of the PtCo3-su catalyst. With the increase in temperature, the surface-adsorbed oxygen species of bimetallic PtCo3 catalysts facilitated the disintegration of the C-H bond on methyl of toluene molecules. Then the products were continued to be oxidized to benzaldehyde and benzoic acid successively by surface active oxygen. Finally, the benzene ring was broken, and the molecules were quickly oxidized to H2O and CO2. Meanwhile, due to the consumption of surface-absorbed oxygen, the derived oxygen vacancies can be complemented by gas-phase O2, and thus active oxygen species will be formed [52]. The reaction proceeded until the toluene was completely removed from the gas. The excellent toluene adsorption capacity of the Csu sample enabled tight adsorption of toluene, benefiting the catalytic oxidation process of toluene. In a nutshell, the lower Ea and excellent adsorption property of the PtCo3-su sample were important reasons for the excellent catalytic performance.

4. Conclusions

In summary, two PtCo3 catalysts supported on two different kinds of carbon supports have been prepared by organic phase reduction and employed in toluene catalytic oxidative removal. Under identical external conditions, the PtCo3-su catalyst manifested better catalytic activity, which can accomplish the elimination of toluene at a temperature of 237 °C and achieve a conversion rate of 90% faster than PtCo3 NPs supported on Cba. Comparative characterizations revealed that higher N-containing levels of PtCo3-su were beneficial for the reactant adsorption, drastically boosting the catalytic activity. It was noticed that the PtCo3-su catalyst held remarkable stability, good water resistance properties as well as favorable recyclability. This work not merely provided an efficient catalyst for low-temperature toluene catalytic oxidation but also provided an effective strategy to improve the catalytic performance.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15228115/s1, Figure S1: Schematic diagram of catalyst performance test equipment; Table S1: Catalytic oxidation of toluene of bimetallic catalysts; Figure S2: Performance test results of two biomass carbon supports; Figure S3: TGA curve of two biomass carbon supports; Figure S4: SEM of Cba (a), Csu (b); Figure S5: XPS spectrum of PtCo3-ba and PtCo3-su; Figure S6: XPS spectrum of PtCo3-su after H2O and stabilization test: (a) Pt 4f and (b) Co 2p. References [53,54,55,56,57] are cited in the Supplementary Materials.

Author Contributions

J.Z. (Jing Zhang): conceptualization, methodology, writing—original draft. J.Z. (Jianwu Zou): methodology, validation. X.X.: software, investigation. Z.L.: formal analysis, investigation. Z.Z.: supervision, writing—review and editing. L.L.: supervision, writing—review and editing, funding acquisition. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Key R&D Program of China (Nos. 2019YFC0214302 (Recipient: Liqing Li), 2019YFC0214303 (Recipient: Zheng Zeng)) and the National Natural Science Foundation of China (No. 21878338 (Recipient: Liqing Li)).

Data Availability Statement

Data are contained within the manuscript or Supplementary Material.

Acknowledgments

The authors gratefully acknowledge the financial support from the National Key R&D Program of China (Nos. 2019YFC0214302, 2019YFC0214303) and the National Natural Science Foundation of China (No. 21878338).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Li, B.; Yang, Q.; Peng, Y.; Chen, J.; Deng, L.; Wang, D.; Hong, X.; Li, J. Enhanced low-temperature activity of LaMnO3 for toluene oxidation: The effect of treatment with an acidic KMnO4. Chem. Eng. J. 2019, 366, 92–99. [Google Scholar] [CrossRef]
  2. Xiong, C.; Li, B.; Liu, H.; Zhao, W.; Duan, C.; Wu, H.; Ni, Y. A smart porous wood-supported flower-like NiS/Ni conjunction with vitrimer co-effect as a multifunctional material with reshaping, shape-memory, and self-healing properties for applications in high-performance supercapacitors, catalysts, and sensors. J. Mater. Chem. A 2020, 8, 10898–10908. [Google Scholar] [CrossRef]
  3. Lan, L.; Huang, Y.; Dan, Y.; Jiang, L. Conjugated porous polymers for gaseous toluene adsorption in humid atmosphere. React. Funct. Polym. 2020, 159, 104804. [Google Scholar] [CrossRef]
  4. Jia, L.; Shi, J.; Long, C.; Lian, F.; Xing, B. VOCs adsorption on activated carbon with initial water vapor contents: Adsorption mechanism and modified characteristic curves. Sci. Total Environ. 2020, 731, 139184. [Google Scholar] [CrossRef] [PubMed]
  5. Ranjbaran, F.; Kamio, E.; Matsuyama, H. Toluene vapor removal using an inorganic/organic double-network ion gel membrane. Sep. Sci. Technol. 2018, 53, 2840–2851. [Google Scholar] [CrossRef]
  6. Tan, Y.-X.; Chai, Z.-M.; Wang, B.-H.; Tian, S.; Deng, X.-X.; Bai, Z.-J.; Chen, L.; Shen, S.; Guo, J.-K.; Cai, M.-Q.; et al. Boosted Photocatalytic Oxidation of Toluene into Benzaldehyde on CdIn2S4-CdS: Synergetic Effect of Compact Heterojunction and S-Vacancy. ACS Catal. 2021, 11, 2492–2503. [Google Scholar] [CrossRef]
  7. Lu, Z.; Wang, C.; Chen, X.; Song, M.; Xia, W. Effects of buffer gas on N-doped graphene in a non-thermal plasma process. Diam. Relat. Mater. 2021, 118, 108548. [Google Scholar] [CrossRef]
  8. Muccee, F.; Ejaz, S.; Riaz, N. Toluene degradation via a unique metabolic route in indigenous bacterial species. Arch. Microbiol. 2019, 201, 1369–1383. [Google Scholar] [CrossRef]
  9. Du, Y.; Zou, J.; Guo, Y.; Xu, X.; Chen, H.; Su, C.; Zeng, Z.; Li, L. A novel viewpoint on the surface adsorbed oxygen and the atom doping in the catalytic oxidation of toluene over low-Pt bimetal catalysts. Appl. Catal. A Gen. 2020, 609, 117913. [Google Scholar] [CrossRef]
  10. Zhang, X.; Bi, F.; Zhao, Z.; Yang, Y.; Li, Y.; Song, L.; Liu, N.; Xu, J.; Cui, L. Boosting toluene oxidation by the regulation of Pd species on UiO-66: Synergistic effect of Pd species. J. Catal. 2022, 413, 59–75. [Google Scholar] [CrossRef]
  11. Guo, M.; Ma, P.; Wang, J.; Xu, H.; Zheng, K.; Cheng, D.; Liu, Y.; Guo, G.; Dai, H.; Duan, E.; et al. Synergy in Au-CuO Janus Structure for Catalytic Isopropanol Oxidative Dehydrogenation to Acetone. Angew. Chem. Int. Ed. 2022, 61, e202203827. [Google Scholar] [CrossRef]
  12. Chen, C.; Chen, F.; Zhang, L.; Pan, S.; Bian, C.; Zheng, X.; Meng, X.; Xiao, F.-S. Importance of platinum particle size for complete oxidation of toluene over Pt/ZSM-5 catalysts. Chem. Commun. 2015, 51, 5936–5938. [Google Scholar] [CrossRef] [PubMed]
  13. Wang, Z.; Yang, H.; Liu, R.; Xie, S.; Liu, Y.; Dai, H.; Huang, H.; Deng, J. Probing toluene catalytic removal mechanism over supported Pt nano- and single-atom-catalyst. J. Hazard. Mater. 2020, 392, 122258. [Google Scholar] [CrossRef] [PubMed]
  14. Zou, J.; Du, Y.; Fang, R.; Duan, X.; Liu, Y.; Mao, J.; Yu, L.; Xu, X.; Zeng, Z.; Li, L. Low platinum alloy catalyst PtCo3 obtaining high catalytic activity and stability with great water and CO2 resistance for catalytic oxidation of toluene. Fuel 2021, 307, 121794. [Google Scholar] [CrossRef]
  15. Miao, H.; Hu, S.; Ma, K.; Sun, L.; Wu, F.; Wang, H.; Li, H. Synthesis of PtCo nanoflowers and its catalytic activity towards nitrobenzene hydrogenation. Catal. Commun. 2018, 109, 33–37. [Google Scholar] [CrossRef]
  16. Guo, Y.; Wen, M.; Li, G.; An, T. Recent advances in VOC elimination by catalytic oxidation technology onto various nanoparticles catalysts: A critical review. Appl. Catal. B Environ. 2020, 281, 119447. [Google Scholar] [CrossRef]
  17. Yang, M.; Liu, J.; Lee, S.; Zugic, B.; Huang, J.; Allard, L.F.; Flytzani-Stephanopoulos, M. A common single-site Pt(II)-O(OH)x- species stabilized by sodium on “active” and “inert” supports catalyzes the water-gas shift reaction. J. Am. Chem. Soc. 2015, 137, 3470–3473. [Google Scholar] [CrossRef]
  18. Deng, J.; He, S.; Xie, S.; Yang, H.; Liu, Y.; Guo, G.; Dai, H. Ultralow Loading of Silver Nanoparticles on Mn2O3 Nanowires Derived with Molten Salts: A High-Efficiency Catalyst for the Oxidative Removal of Toluene. Environ. Sci. Technol. 2015, 49, 11089–11095. [Google Scholar] [CrossRef]
  19. Tan, H.; Wang, J.; Yu, S.; Zhou, K. Support Morphology-Dependent Catalytic Activity of Pd/CeO2 for Formaldehyde Oxidation. Environ. Sci. Technol. 2015, 49, 8675–8682. [Google Scholar] [CrossRef]
  20. Yan, Z.; Xu, Z.; Cheng, B.; Jiang, C. Co3O4 nanorod-supported Pt with enhanced performance for catalytic HCHO oxidation at room temperature. Appl. Surf. Sci. 2017, 404, 426–434. [Google Scholar] [CrossRef]
  21. Deng, H.; Kang, S.; Wang, C.; He, H.; Zhang, C. Palladium supported on low-surface-area fiber-based materials for catalytic oxidation of volatile organic compounds. Chem. Eng. J. 2018, 348, 361–369. [Google Scholar] [CrossRef]
  22. Zuo, S.; Wang, X.; Yang, P.; Qi, C. Preparation and high performance of rare earth modified Pt/MCM-41 for benzene catalytic combustion. Catal. Commun. 2017, 94, 52–55. [Google Scholar] [CrossRef]
  23. Qin, C.; Guo, M.; Zheng, Y.; Yu, R.; Huang, J.; Dang, X.; Yan, D. Two-component zeolite-alumina system for toluene trapping with subsequent nonthermal plasma mineralization. J. Ind. Eng. Chem. 2020, 95, 215–223. [Google Scholar] [CrossRef]
  24. Serafin, J.; Ouzzine, M.; Cruz, O.F.; Sreńscek-Nazzal, J.; Gómez, I.C.; Azar, F.-Z.; Mafull, C.A.R.; Hotza, D.; Rambo, C.R. Conversion of fruit waste-derived biomass to highly microporous activated carbon for enhanced CO2 capture. Waste Manag. 2021, 136, 273–282. [Google Scholar] [CrossRef]
  25. Zeng, L.; Thiruppathi, A.R.; van der Zalm, J.; Li, X.; Chen, A. Biomass-derived amorphous carbon with localized active graphite defects for effective electrocatalytic N2 reduction. Appl. Surf. Sci. 2022, 575, 151630. [Google Scholar] [CrossRef]
  26. Zhao, H.; Zhong, H.; Jiang, Y.; Li, H.; Tang, P.; Li, D.; Feng, Y. Porous ZnCl2-Activated Carbon from Shaddock Peel: Methylene Blue Adsorption Behavior. Materials 2022, 15, 895. [Google Scholar] [CrossRef]
  27. Li, Q.; Lu, T.; Wang, L.; Pang, R.; Shao, J.; Liu, L.; Hu, X. Biomass based N-doped porous carbons as efficient CO2 adsorbents and high-performance supercapacitor electrodes. Sep. Purif. Technol. 2021, 275, 119204. [Google Scholar] [CrossRef]
  28. Zhou, X.; Chu, W.; Sun, W.; Zhou, Y.; Xue, Y. Enhanced interaction of nickel clusters with pyridinic-N (B) doped graphene using DFT simulation. Comput. Theor. Chem. 2017, 1120, 8–16. [Google Scholar] [CrossRef]
  29. Li, C.; Nakagawa, Y.; Yabushita, M.; Nakayama, A.; Tomishige, K. Guaiacol Hydrodeoxygenation over Iron–Ceria Catalysts with Platinum Single-Atom Alloy Clusters as a Promoter. ACS Catal. 2021, 11, 12794–12814. [Google Scholar] [CrossRef]
  30. Qi, Y.; Yang, Z.; Jiang, Y.; Han, H.; Wu, T.; Wu, L.; Liu, J.; Wang, Z.; Wang, F. Platinum−Copper Bimetallic Nanoparticles Supported on TiO2 as Catalysts for Photo−thermal Catalytic Toluene Combustion. ACS Appl. Nano Mater. 2022, 5, 1845–1854. [Google Scholar] [CrossRef]
  31. Tsai, W.-T.; Yang, J.-M.; Hsu, H.-C.; Lin, C.-M.; Lin, K.-Y.; Chiu, C.-H. Development and characterization of mesoporosity in eggshell ground by planetary ball milling. Microporous Mesoporous Mater. 2008, 111, 379–386. [Google Scholar] [CrossRef]
  32. Sun, Y.; Su, G.; He, Z.; Wei, Y.; Hu, J.; Liu, H.; Liu, G. Porous Carbons Derived from Desiliconized Rice Husk Char and Their Applications as an Adsorbent in Multivalent Ions Recycling for Spent Battery. J. Chem. 2022, 2022, 8225088. [Google Scholar] [CrossRef]
  33. Madhu, J.; Ramakrishnan, V.M.; Santhanam, A.; Natarajan, M.; Palanisamy, B.; Velauthapillai, D.; Chi, N.T.L.; Pugazhendhi, A. Comparison of three different structures of zeolites prepared by template-free hydrothermal method and its CO2 adsorption properties. Environ. Res. 2022, 214, 113949. [Google Scholar] [CrossRef]
  34. Liu, B.; Shi, R.; Ma, X.; Chen, R.; Zhou, K.; Xu, X.; Sheng, P.; Zeng, Z.; Li, L. High yield nitrogen-doped carbon monolith with rich ultramicropores prepared by in-situ activation for high performance of selective CO2 capture. Carbon 2021, 181, 270–279. [Google Scholar] [CrossRef]
  35. Wedler, C.; Span, R. Micropore Analysis of Biomass Chars by CO2 Adsorption: Comparison of Different Analysis Methods. Energy Fuels 2021, 35, 8799–8806. [Google Scholar] [CrossRef]
  36. Li, B.; Hu, J.; Xiong, H.; Xiao, Y. Application and Properties of Microporous Carbons Activated by ZnCl2: Adsorption Behavior and Activation Mechanism. ACS Omega 2020, 5, 9398–9407. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, H.; Gao, J.; Xu, X.; Liu, B.; Yu, L.; Ren, Y.; Shi, R.; Zeng, Z.; Li, L. Adsorption of Volatile Organic Compounds (VOCs) on Oxygen-rich Porous Carbon Materials Obtained from Glucose/Potassium Oxalate. Chem. -Asian J. 2021, 16, 1118–1129. [Google Scholar] [CrossRef]
  38. Xu, X.; Guo, Y.; Shi, R.; Chen, H.; Du, Y.; Liu, B.; Zeng, Z.; Yin, Z.; Li, L. Natural Honeycomb-like structure cork carbon with hierarchical Micro-Mesopores and N-containing functional groups for VOCs adsorption. Appl. Surf. Sci. 2021, 565, 150550. [Google Scholar] [CrossRef]
  39. Ma, X.; Li, L.; Chen, R.; Wang, C.; Zhou, K.; Li, H. Porous carbon materials based on biomass for acetone adsorption: Effect of surface chemistry and porous structure. Appl. Surf. Sci. 2018, 459, 657–664. [Google Scholar] [CrossRef]
  40. Meng, X.; Yang, L.; Jiang, W.; Yao, L. Adsorption of acetone and toluene by N-functionalized porous carbon derived from ZIF-8. J. Ind. Eng. Chem. 2022, 111, 137–146. [Google Scholar] [CrossRef]
  41. Shen, Y.; Zhang, N. A facile synthesis of nitrogen-doped porous carbons from lignocellulose and protein wastes for VOCs sorption. Environ. Res. 2020, 189, 109956. [Google Scholar] [CrossRef] [PubMed]
  42. Xie, S.; Deng, J.; Zang, S.; Yang, H.; Guo, G.; Arandiyan, H.; Dai, H. Au–Pd/3DOM Co3O4: Highly active and stable nanocatalysts for toluene oxidation. J. Catal. 2014, 322, 38–48. [Google Scholar] [CrossRef]
  43. Arandiyan, H.; Dai, H.; Deng, J.; Liu, Y.; Bai, B.; Wang, Y.; Li, X.; Xie, S.; Li, J. Three-dimensionally ordered macroporous La0.6Sr0.4MnO3 with high surface areas: Active catalysts for the combustion of methane. J. Catal. 2013, 307, 327–339. [Google Scholar] [CrossRef]
  44. Li, Q.; Odoom-Wubah, T.; Fu, X.; Mulka, R.; Sun, D.; Zheng, Y.; Jia, L.; Huang, J.; Li, Q. Photoinduced Pt-Decorated Expanded Graphite toward Low-Temperature Benzene Catalytic Combustion. Ind. Eng. Chem. Res. 2020, 59, 11453–11461. [Google Scholar] [CrossRef]
  45. Dong, J.; Zhang, X.; Huang, J.; Hu, J.; Chen, Z.; Lai, Y. In-situ formation of unsaturated defect sites on converted CoNi alloy/Co-Ni LDH to activate MoS2 nanosheets for pH-universal hydrogen evolution reaction. Chem. Eng. J. 2021, 412, 128556. [Google Scholar] [CrossRef]
  46. Zhang, Q.; Li, X.L.; Tao, B.X.; Wang, X.H.; Deng, Y.H.; Gu, X.Y.; Li, L.J.; Xiao, W.; Li, N.B.; Luo, H.Q. CoNi based alloy/oxides@N-doped carbon core-shell dendrites as complementary water splitting electrocatalysts with significantly enhanced catalytic efficiency. Appl. Catal. B Environ. 2019, 254, 634–646. [Google Scholar] [CrossRef]
  47. He, J.; Chen, D.; Li, N.; Xu, Q.; Li, H.; He, J.; Lu, J. Pt–Pd bimetallic nanoparticles anchored on uniform mesoporous MnO2 sphere as an advanced nanocatalyst for highly efficient toluene oxidation. Green Energy Environ. 2022, 7, 1349–1360. [Google Scholar] [CrossRef]
  48. Du, X.; Li, C.; Zhang, J.; Zhao, L.; Li, S.; Lyu, Y.; Zhang, Y.; Zhu, Y.; Huang, L. Highly efficient simultaneous removal of HCHO and elemental mercury over Mn-Co oxides promoted Zr-AC samples. J. Hazard. Mater. 2020, 408, 124830. [Google Scholar] [CrossRef]
  49. Wang, Y.; Arandiyan, H.; Liu, Y.; Liang, Y.; Peng, Y.; Bartlett, S.; Dai, H.; Rostamnia, S.; Li, J. Template-free Scalable Synthesis of Flower-like Co3−xMnxO4 Spinel Catalysts for Toluene Oxidation. ChemCatChem 2018, 10, 3429–3434. [Google Scholar] [CrossRef]
  50. Shao, J.; Lin, F.; Wang, Z.; Liu, P.; Tang, H.; He, Y.; Cen, K. Low temperature catalytic ozonation of toluene in flue gas over Mn-based catalysts: Effect of support property and SO2/water vapor addition. Appl. Catal. B Environ. 2020, 266, 118662. [Google Scholar] [CrossRef]
  51. Xu, W.; Wang, N.; Chen, Y.; Chen, J.; Xu, X.; Yu, L.; Chen, L.; Wu, J.; Fu, M.; Zhu, A.; et al. In situ FT-IR study and evaluation of toluene abatement in different plasma catalytic systems over metal oxides loaded γ-Al2O3. Catal. Commun. 2016, 84, 61–66. [Google Scholar] [CrossRef]
  52. Rong, S.; Zhang, P.; Liu, F.; Yang, Y. Engineering Crystal Facet of α-MnO2 Nanowire for Highly Efficient Catalytic Oxidation of Carcinogenic Airborne Formaldehyde. ACS Catal. 2018, 8, 3435–3446. [Google Scholar] [CrossRef]
  53. Shi, Y.; Wang, J.; Wang, Y.; Kong, F.; Zhou, R. Synergistic effect in bimetal Cr-Pt/ZSM-5 catalysts with high activity for CVOCs low-temperature removal. J. Environ. Chem. Eng. 2022, 10, 107629. [Google Scholar] [CrossRef]
  54. Yao, S.; Fang, W.; Wang, B.; Zeng, Y.; Chen, L.; Yan, X.; Bai, G.; Li, Y. Rh1Cu3/ZSM-5 as an Efficient Bifunctional Catalyst/Adsorbent for VOCs Abatement. Catal. Lett. 2021, 152, 771–780. [Google Scholar] [CrossRef]
  55. Wang, Z.; Zhang, L.; Ji, J.; Wu, Y.; Cai, Y.; Yao, X.; Gu, X.; Xiong, Y.; Wan, H.; Dong, L.; et al. Catalytic enhancement of small sizes of CeO2 additives on Ir/Al2O3 for toluene oxidation. Appl. Surf. Sci. 2022, 571, 151200. [Google Scholar] [CrossRef]
  56. Niftaliyeva, A.; Karaduman, A.; Kalwar, N.H.; Avci, A.; Pehlivan, E. Synthesis of novel metal/bimetal nanoparticle-modified ZSM-5 zeolite nanocomposite catalysts and application on toluene methylation. Res. Chem. Intermed. 2021, 48, 145–165. [Google Scholar] [CrossRef]
  57. Zhang, W.; Descorme, C.; Valverde, J.L.; Giroir-Fendler, A. Cu-Co mixed oxide catalysts for the total oxidation of toluene and propane. Catal. Today 2022, 384–386, 238–245. [Google Scholar] [CrossRef]
Figure 1. Diagram of PtCo3 catalyst synthesis method.
Figure 1. Diagram of PtCo3 catalyst synthesis method.
Materials 15 08115 g001
Figure 2. Conversion rates of as-synthesized samples at 500 ppm (a), 1000 ppm (b), 1500 ppm (c).
Figure 2. Conversion rates of as-synthesized samples at 500 ppm (a), 1000 ppm (b), 1500 ppm (c).
Materials 15 08115 g002
Figure 3. Reaction stability with time for toluene oxidation over PtCo3-su catalyst.
Figure 3. Reaction stability with time for toluene oxidation over PtCo3-su catalyst.
Materials 15 08115 g003
Figure 4. Reaction stability with time for toluene oxidation over PtCo3-su catalyst with 3.5 vol% to 20 vol% water vapor contents.
Figure 4. Reaction stability with time for toluene oxidation over PtCo3-su catalyst with 3.5 vol% to 20 vol% water vapor contents.
Materials 15 08115 g004
Figure 5. N2 adsorption and desorption (a), pore size distribution (b), toluene adsorption isotherms (c), N1s spectrum of the Cba, Csu (d,e), and atomic ratio of N species of Cba and Csu (f).
Figure 5. N2 adsorption and desorption (a), pore size distribution (b), toluene adsorption isotherms (c), N1s spectrum of the Cba, Csu (d,e), and atomic ratio of N species of Cba and Csu (f).
Materials 15 08115 g005
Figure 6. Optimized geometry and stable adsorption configurations of toluene on various nitrogen-containing functional groups, including pure graphite (a), graphitic-N (b), pyridinic-N (c), and pyrrolic-N (d).
Figure 6. Optimized geometry and stable adsorption configurations of toluene on various nitrogen-containing functional groups, including pure graphite (a), graphitic-N (b), pyridinic-N (c), and pyrrolic-N (d).
Materials 15 08115 g006
Figure 7. N2 adsorption and desorption (a), pore size distribution (b), toluene adsorption isotherms (c), and C7H8-TPD profiles (d) of catalysts.
Figure 7. N2 adsorption and desorption (a), pore size distribution (b), toluene adsorption isotherms (c), and C7H8-TPD profiles (d) of catalysts.
Materials 15 08115 g007
Figure 8. TEM images of PtCo3-ba (a,b), PtCo3-su (e,f). HRTEM of PtCo3-ba (c), PtCo3-su (g). Elemental mappings of PtCo3-ba (d), PtCo3-su (h).
Figure 8. TEM images of PtCo3-ba (a,b), PtCo3-su (e,f). HRTEM of PtCo3-ba (c), PtCo3-su (g). Elemental mappings of PtCo3-ba (d), PtCo3-su (h).
Materials 15 08115 g008
Figure 9. XRD patterns of samples (a), high-resolution XPS spectra of the O 1s (b), Pt 4f (c), and Co 2p (d) of as-prepared catalysts.
Figure 9. XRD patterns of samples (a), high-resolution XPS spectra of the O 1s (b), Pt 4f (c), and Co 2p (d) of as-prepared catalysts.
Materials 15 08115 g009
Figure 10. Arrhenius diagram of toluene oxidation on PtCo3-ba and PtCo3-su.
Figure 10. Arrhenius diagram of toluene oxidation on PtCo3-ba and PtCo3-su.
Materials 15 08115 g010
Figure 11. In situ DRIFTS spectrum of PtCo3-su (a) and probable toluene oxidation mechanism on PtCo3-su catalyst (b).
Figure 11. In situ DRIFTS spectrum of PtCo3-su (a) and probable toluene oxidation mechanism on PtCo3-su catalyst (b).
Materials 15 08115 g011
Table 1. Structural properties and elemental compositions of biomass carbon support.
Table 1. Structural properties and elemental compositions of biomass carbon support.
SamplesSBET
m2 g−1
Vtotal
cm3 g−1
Vmic
cm3 g−1
Vmes
cm3 g−1
Micropore Percentage
%
Average Pore Width
nm
C
at.%
O
at.%
N
at.%
Cba10290.7870.3650.42246.383.05888.344.267.40
Csu9600.7730.3770.39648.773.21187.434.148.43
Table 2. Structural properties and elemental compositions of catalysts.
Table 2. Structural properties and elemental compositions of catalysts.
SamplesSBET
m2 g−1
Vtotal
cm3 g−1
PtCoO
PtCo3-ba680.1471.8474.2084.562
PtCo3-su740.1761.8414.2154.314
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Zhang, J.; Zou, J.; Xu, X.; Li, Z.; Zeng, Z.; Li, L. Nitrogen-Doped Porous Carbon from Biomass with Efficient Toluene Adsorption and Superior Catalytic Performance. Materials 2022, 15, 8115. https://doi.org/10.3390/ma15228115

AMA Style

Zhang J, Zou J, Xu X, Li Z, Zeng Z, Li L. Nitrogen-Doped Porous Carbon from Biomass with Efficient Toluene Adsorption and Superior Catalytic Performance. Materials. 2022; 15(22):8115. https://doi.org/10.3390/ma15228115

Chicago/Turabian Style

Zhang, Jing, Jianwu Zou, Xiang Xu, Zhuang Li, Zheng Zeng, and Liqing Li. 2022. "Nitrogen-Doped Porous Carbon from Biomass with Efficient Toluene Adsorption and Superior Catalytic Performance" Materials 15, no. 22: 8115. https://doi.org/10.3390/ma15228115

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