Next Article in Journal
Composite Materials Based on a Zr4+ MOF and Aluminosilicates for the Simultaneous Removal of Cationic and Anionic Dyes from Aqueous Media
Next Article in Special Issue
Efficient and Selective Removal of Organic Cationic Dyes by Peel of Brassica juncea Coss. var. gemmifera Lee et Lin-Based Biochar
Previous Article in Journal
The Expanded Vermiculite Was Quickly Prepared by the Catalytic Action of Manganese Dioxide on Hydrogen Peroxide and Its Adsorption Properties to Cd
Previous Article in Special Issue
Plant and Microbial Approaches as Green Methods for the Synthesis of Nanomaterials: Synthesis, Applications, and Future Perspectives
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Sweet-Potato-Vine-Based High-Performance Porous Carbon for Methylene Blue Adsorption

1
Chongqing Key Laboratory of Economic Plant Biotechnology, College of Landscape Architecture and Life Science/Institute of Special Plants, Chongqing University of Arts and Sciences, Yongchuan, Chongqing 402160, China
2
College of Food Science, Southwest University, Beibei, Chongqing 400716, China
3
College of Biology and Food Engineering, Chongqing Three Gorges University, Wanzhou, Chongqing 404199, China
*
Author to whom correspondence should be addressed.
Submission received: 21 December 2022 / Revised: 8 January 2023 / Accepted: 12 January 2023 / Published: 13 January 2023
(This article belongs to the Special Issue Carbon-Based Materials for Sustainable Chemistry)

Abstract

:
In this study, sweet-potato-vine-based porous carbon (SPVPC) was prepared using zinc chloride as an activating and pore-forming agent. The optimised SPVPC exhibited abundant porous structures with a high specific surface area of 1397.8 m2 g−1. Moreover, SPVPC exhibited excellent adsorption characteristics for removing methylene blue (MB) from aqueous solutions. The maximum adsorption capacity for MB reached 653.6 mg g−1, and the reusability was satisfactory. The adsorption kinetics and isotherm were in good agreement with the pseudo-second-order kinetics and Langmuir models, respectively. The adsorption mechanism was summarised as the synergistic effects of the hierarchically porous structures in SPVPC and various interactions between SPVPC and MB. Considering its low cost and excellent adsorption performance, the prepared porous carbon is a promising adsorbent candidate for dye wastewater treatment.

1. Introduction

Dye wastewater has attracted wide attention, with approximately 100 tons of dye discharged into natural water bodies annually, causing serious water pollution and ecosystem destruction [1]. Methylene blue (MB) is a synthetic cationic dye and the most commonly used organic dye in artificial synthesis [2]. It poses serious threats to human health, such as cyanosis, heart disease, eye burns, tissue necrosis, nausea, diarrhoea, emesis, gastritis [3,4], and negatively affects the growth of aquatic plant and animals [5]. Moreover, MB is very stable and poorly degradable in wastewater under natural conditions [6]. Therefore, there is an urgent need to eliminate MB from wastewater using simple and efficient methods.
Currently available techniques, such as membrane separation, biodegradation, adsorption, electrochemical, photocatalysis, oxidation, and flocculation, have been employed to purify dye wastewater [7]. In contrast, adsorption technology is widely applied because of its high efficiency, low-cost, easy operation and the extensiveness of available materials [1]. Various adsorbents, such as carbon materials, metal nanomaterials, biomass, hydrogels, graphene, polymers, metal–organic frameworks, and their derivatives, have been explored for dye adsorption [8]. Among them, porous carbon (PC) is considered a potential dye adsorbent owing to its abundant pore structure, high specific surface area, and excellent adsorbability [9]. In recent years, the preparation of PC through renewable and cost-effective biomass as a basic material has arisen great interest, for example, rice husk [10], wheat straw [11], pomelo peel [12], cotton stalk [13], bagasse [14], etc. [15].
Sweet potato (Ipomoea batatas L.) is planted worldwide, and its tuberous root is usually used as a food source [16]. Sweet potato vine (SPV) is the aboveground biomass of sweet potatoes as a by-product of a farm. A large amount of sweet potato vine is produced annually in the field. A small part of them is used as animal feed, and most of them are discarded as agricultural waste and underutilised [17,18], which leads to resource waste and environmental pollution. Dried SPV is rich in cellulose (33.01%), hemicellulose (12.25%), and lignin (7.85%) [18]. Therefore, the use of SPV as a raw material to develop PC is worth considering. Gao et al. prepared SPV-based N- and S-doped carbons at 800 °C to catalyse the oxygen reduction reaction [19]. However, their preparation temperature was relatively high, and the porous morphology and specific surface area required further improvement. To the best of our knowledge, high-quality SPV-based PC (SPVPC) as an MB adsorbent has not yet been reported. In this study, as shown in Figure 1, SPVPC was developed using zinc chloride (ZnCl2) as an activating and pore-forming agent, and its structural characteristics and adsorption performance towards MB were analysed in detail.

2. Results and Discussion

2.1. Preparation Parameters of SPVPC

In this study, SPVPC was prepared using ZnCl2 as the pore-forming agent. ZnCl2 can catalyse the degradation, dehydration, and carbonisation of cellulose-rich materials, leading to the charring and aromatisation of the carbon skeleton and the formation of pore structures [20]. Therefore, the preparation impact factors of the mass ratio (MR) of ZnCl2 to SPV, as well as the carbonisation temperature (T) and time (t), were investigated using a single factor methodology to obtain high-performance SPVPC. As seen in Figure S1A, qe increased sharply with the increasing MR until 2:1, with the highest value being 299.6 mg g−1, but it slightly decreased when the MR was over 2:1. Moreover, the effect of T on qe showed a similar pattern and qe reached a maximum value of 299.2 mg g−1 at 500 °C (Figure S1B). A higher temperature of over 500 °C reduced qe because it could damage the pore structure of carbon, reducing the specific surface area [21]. Additionally, when t was prolonged to 1 h, qe reached its highest value of 299.7 mg g−1 and then decreased (Figure S1C). Therefore, MR = 2:1, T = 500 °C, and t = 1 h were selected as the optimal preparation conditions for SPVPC.

2.2. Physicochemical Properties of SPVPC

The surface morphology of the optimised SPVPC was observed using SEM. As shown in Figure 2A, SPVPC exhibited a porous grid-pipe-like shape with abundant macropore structures. Furthermore, the mesopore and micropore properties, pore size distribution, and specific surface area were studied using the Brunauer–Emmett–Teller (BET) and density function theory (DFT) methods. Figure 2B showed that the nitrogen adsorption–desorption isotherm of SPVPC exhibited a typical IV adsorption isotherm according to the IUPAC classification, indicating the existence of micropores and mesopores in SPVPC [4]. An adsorption hysteresis loop in the curve at 0.4 < P/P0 < 1.0 demonstrated that SPVPC had abundant mesopores [12]. Meanwhile, the curve increased sharply at P/P0 < 0.01, suggesting the presence of rich micropores in the SPVPC [1]. The DFT pore size distribution also revealed that SPVPC had generous micropores and mesopores with the average pore diameter of 2.9 nm (Figure 2C). Combining the SEM results, SPVPC possessed a large number of pore structures with a total pore volume of 1.03 cm3 g−1. Notably, the pore structures could accelerate the diffusion of MB molecules on the surface of SPVPC rapidly and entrap more MB molecules [22]. Surprisingly, the specific surface area of SPVPC reached 1397.8 m2 g−1 owing to its rich porous structure and was higher than those of other recently reported PCs [12,13,23,24,25], especially the reported SPV-derived carbon (884.9 m2 g−1) [19].
The crystallographic characteristics of the SPVPC were explored using the XRD pattern shown in Figure 2D. The sample exhibited two broad diffraction peaks at around 26° and 44°, consistent with the (002) and (100)/(101) crystal planes of typical graphitic carbon (JCPDS card no. 75-1621) [26]. The (002) diffraction revealed that the carbons were randomly oriented carbon layers with many defects, while the (100)/(101) diffraction indicated a turbostratic structure with low crystallinity [26,27]. These turbostratic structures favoured the adsorption of dye molecules [28]. To further investigate the structure of the SPVPC, its Raman spectrum was recorded (Figure 2E). The peak at 1356 cm−1 (D-band) is generally attributed to disordered and defective carbons with the breathing mode of the A1g symmetry, and 1595 cm−1 (G-band) corresponds to ordered graphitic structures with the zone centre of the E2g mode [19,26]. Furthermore, the calculated intensity ratio of the D and G bands (ID/IG) was 0.7, which was less than that of pristine graphene (approximately 1.4), indicating that SPVPC had abundant amorphous carbon structures [12,19]. Additionally, the full width at half-maximum of the D band was much larger than that of the G band, demonstrating many defects in SPVPC [26], which was associated with the XRD analysis. The surface functional groups of the SPVPC were analysed using FTIR spectroscopy. Figure 2F shows the characteristic bands at 3385 cm−1, 2926 cm−1 and 2860 cm−1, 1697 cm−1, 1568 cm−1, 1165 cm−1 and 1070 cm−1, 889 cm−1 and 812 cm−1, ascribed to the stretching vibrations of O–H (hydroxyl), C–H (methyl and methylene), C=O (carbonyl and carboxylic), C=C (benzene ring), C–O (ester, ether and phenol), and C–H (benzene ring), respectively [1,29,30,31]. Therefore, it can be concluded that SPVPC possessed abundant functional groups, ascribed to the carbonisation of SPV during the heat-treatment process at a relatively low temperature of 500 °C [12,32]. However, the higher temperature (600–700 °C) could destroy most functional groups of porous carbon according to our previous report [32]. Importantly, these groups played key roles in dye adsorption via various interactions [33].

2.3. Adsorption and Reusability Properties of SPVPC

Based on previous reports on MB adsorption, pH was considered a critical factor influencing the adsorption process by altering the chemical structure and surface of MB and the adsorbent [34]. As shown in Figure 3A, with the solution pH increasing from 2 to 12, qe gradually increased from 218.8 mg g−1 to 299.4 mg g−1. MB is a cationic dye with a positive charge in an aqueous solution [12], and the SPVPC surface is negative at a pH > 4.1 (pHzero charge point, Figure S2A) [35,36]. As the pH increased from 4.1 to 12, qe increased sharply owing to the reinforced electrostatic interaction between SPVPC and MB. When pH < 4.1, electrostatic repulsion between the positively charged SPVPC surface and MB occurred [35,36], and excess H+ competed with MB for the surface adsorption sites of SPVPC, leading to a lower qe [2]. However, qe could reach 218.8 mg g−1 at pH 2, indicating that adsorption was regulated by electrostatic interaction and other interactions, such as π–π attraction, and hydrogen bonding between SPVPC and MB molecules. Moreover, qe increased with increasing c0 (Figure S2B), because a high c0 could improve the contact area between the MB molecules and SPVPC and provide a large driving force to accelerate MB mass transfer [2]. Additionally, a high adsorption temperature promoted qe to a certain extent, involving an endothermic process (Figure S2C). Surprisingly, the MB adsorption on SPVPC was quick (209.5 mg g−1 min−1) within 1 min (Figure 3B), ascribed to the numerous unoccupied adsorption sites and the decrease of MB diffusion resistance by pores. Subsequently, the adsorption rate decreased gradually owing to the occupancy of adsorption sites on the SPVPC surface by MB molecules, which slowed the internal diffusion [26]. After 40 min, the adsorption reached an equilibrium with qe of 298.5 mg g−1, revealing that SPVPC possessed a high adsorption efficiency for MB.
The PFO and PSO kinetics were analysed to investigate the adsorption process of MB by SPVPC (Figure 3C,D; Table S1). Compared with PFO, PSO had a higher correlation coefficient (R2 = 0.9993), and its calculated qe,cal was consistent with the experimental qe,exp, revealing that PSO was more suitable for explaining the adsorption process, and chemisorption mainly dominated the adsorption behaviour of SPVPC towards MB [1]. Furthermore, the Langmuir and Freundlich isotherms were employed to determine the adsorption characteristics at equilibrium (Figure 3E and Figure S2D; Table S2). It was found that the R2 (0.9995) of Langmuir was much higher than that of Freundlich (0.8423), suggesting that the adsorption process fit the Langmuir isotherm and a monolayer coverage of MB molecules occurred on the SPVPC surface, which was consistent with previously reported PCs [1,12,32]. From the Langmuir isotherm, the calculated qm of 653.6 mg g−1 was higher than that of most reported adsorbents (Table S3). Additionally, the reusability of SPVPC is an important factor in assessing their practical application. As shown in Figure 3F, the regenerated SPVPC maintained a high qe of 97.3 mg g−1 after five cycles, and the qe decreased by only 1.5%, indicating that SPVPC had excellent reusability and stability for MB removal. Therefore, the prepared SPVPC can be used as a potential regenerable adsorbent for dye wastewater cleaning.

2.4. Adsorption Mechanism of SPVPC towards MB

The MB adsorption mechanism of the SPVPC is summarised in Figure 4. According to the above PSO kinetics and Langmuir isotherm results, the adsorption process of SPVPC towards MB was mainly controlled by chemical forces and a monolayer adsorption [1]. Importantly, the rich hierarchical porous structures of SPVPC provided sufficient active adsorption sites to trap more MB molecules and offered available channels to decrease the migration resistance and improve the mass transfer process [12,22,30]. The adsorption was driven by the strong electrostatic attraction between the negatively charged functional groups on the SPVPC surface and the positively charged MB molecules [1,33]. Furthermore, hydrogen bonding interaction was a possible factor for adsorption, which occurred between the H atoms on the SPVPC surface and the N atoms on MB molecules [33]. In addition, the π–π interaction formed by the aromatic turbostratic structures of SPVPC and the benzene rings of MB molecules was favourable for the adsorption process [1,12,32]. Therefore, the adsorption mechanism of SPVPC was clarified as the various interactions between SPVPC and the MB dye and the rich pores in SPVPC.

3. Materials and Methods

3.1. Materials

Sweet potato vines were collected from a sweet potato field in Yongchuan (Chongqing, China), dried, milled, and passed through a 60 mesh sieve to obtain a powder with a particle size ≤0.25 mm for the carbonisation experiments. MB was purchased from the Shanghai Aladdin Chemical Company (Shanghai, China). ZnCl2, hydrochloric acid (HCl), sodium hydroxide (NaOH), and other analytical reagents were purchased from the Chengdu Kelong Chemical Company (Sichuan, China).

3.2. Preparation of SPVPC

SPVPC was prepared using the ZnCl2 activation method. Typically, 1.0 g of SPV powder and 0.5–4 g of ZnCl2 were dispersed in 50 mL of water with vigorous magnetic stirring for 1 h. The mixtures were then dried using a vacuum freeze dryer. The dried samples were heated and carbonised in a tube furnace at 300–700 °C for 0.5–2.5 h at a heating rate of 5 °C min−1 under a nitrogen flow. After natural cooling, the obtained samples were ground and washed with a HCl solution (1 mol L−1) and ultrapure water to remove inorganic impurities [19,20]. Finally, the samples were dried at 50 °C in a vacuum drying oven.

3.3. Characterisations of SPVPC

The morphology was observed using a JSM-7800F scanning electron microscope (SEM, JEOL, Tokyo, Japan) at 10 kV. The specific surface area and pore characteristics were determined using a porosity analyser (Quantachrome, Boynton Beach, FL, USA). The crystalline structure was determined using an Ultima IV X-ray diffractometer (XRD, Rigaku, Tokyo, Japan) with Cu-Kα radiation (λ = 0.15 nm) and a scan speed of 4° min−1. The microcrystalline structure of graphite was characterised using a Renishaw inVia plus micro-Raman spectroscopy system (Renishaw, Wotton-under-Edge, UK). The surface groups were recorded on a Nicolet 6700 Fourier transform infrared spectrophotometer (FTIR, Thermo Fisher, Waltham, MA, USA).

3.4. Adsorption of SPVPC

Batch adsorption experiments were performed by adding 10 mg SPVPC into 10 mL of MB aqueous solutions with different initial concentrations (100–700 mg L−1), pH values (2–12), times (1–120 min), and temperatures (25–50 °C) in an incubation shaker at a speed of 200 rpm. The mixtures were then centrifuged, and their supernatant absorbances were measured using a TU-1901 UV–vis spectrophotometer (Persee, Beijing, China) at 664 nm. The final concentrations of MB were determined using the standard curve of absorbance and concentration. The formulas of the adsorption capacity, pseudo-first-order (PFO) and pseudo-second-order (PSO) kinetics, and Langmuir and Freundlich isotherms are as follows.
q e = ( c 0 c e ) V m
log ( q e q t )   = log q e - k 1 t 2 . 303
t q t = 1 k 2 q e 2 + t q e
c e q e = c e q m + 1 b q m
log q e = log k + 1 n log c e
where qe (mg g−1) and qt (mg g−1) are the adsorption capacities of SPVPC toward MB at equilibrium and any time t (min), respectively; c0 (mg L−1) and ce (mg L−1) are the initial and equilibrium concentrations of MB, respectively; V (L) is the MB solution volume; m (g) is the SPVPC mass; k1 (min−1) and k2 (g mg−1 min−1) are the rate constants of the PFO and PSO kinetics, respectively; qm (mg g−1) is the max adsorption capacity; b (L mg−1) and k ((mg g−1(L mg−1)1/n) are the constants of the Langmuir and Freundlich isotherms, respectively; and 1/n is the heterogeneity factor of the Freundlich isotherm.

3.5. Reusability of SPVPC

Typically, 10 mg of SPVPC was mixed with 10 mL of a 100 mg L−1 MB solution at 200 rpm for 60 min (pH = 12, 25 °C). After adsorption, the MB-adsorbed SPVPC was separated by centrifugation, and 10 mL of ethanol (pH = 2) was added for desorption through ultrasound treatment for 5 min. Finally, the desorbed SPVPC was isolated by centrifugation for the next cycle. This process was repeated five times.

4. Conclusions

In this study, SPVPC was successfully prepared using a facile ZnCl2 activation strategy. The obtained SPVPC possessed a high specific surface area, rich pores, strong adsorption ability, and excellent reusability for MB removal. The adsorption process fit pseudo-second-order kinetics and Langmuir isotherm models. The adsorption mechanism was considered to be the integrated effect of hierarchical pores in SPVPC and the various interactions between SPVPC and MB molecules. This study provides a novel and promising absorbent material for treating dye wastewater. Prospectively, the microorganism and enzyme are possible alternatives and ecofriendly porogens to treat the biomass precursor for preparing PC.

Supplementary Materials

The following are available online at https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/molecules28020819/s1. Figure S1: Effect of the preparation parameters of (A) MR (T: 500 °C, t: 1 h), (B) T (MR: 2:1, t: 1 h), and (C) t (MR: 2:1, T: 500 °C) of SPVPC towards MB adsorption (c0: 300 g L−1, pH = 12, 25 °C, 2 h), Figure S2: (A) Zero-charge point of SPVPC. Effect of (B) c0, (C) T, and (D) Freundlich isotherm for MB adsorption on SPVPC, Table S1: PFO and PSO kinetics parameters, Table S2: Langmuir and Freundlich isotherm parameters, Table S3: Comparison of the max adsorption capacities of MB on various adsorbents. References from [37,38,39,40,41,42,43,44,45] are from Supplementary Materials.

Author Contributions

Performing the experiments and writing the initial draft, W.Z.; processing data, Y.Z. and Z.L.; revising the manuscript and validation, Q.L.; supervision and supply of the research idea, D.J. and J.T. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by Chongqing University of Arts and Sciences Natural Science Major Cultivation Project (P2022LS14) and Special Grant for Chongqing Postdoctoral Researcher Research Project (2021XM3050).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors have no conflict of interest.

References

  1. Tan, Y.; Wang, X.; Xiong, F.; Ding, J.; Qing, Y.; Wu, Y. Preparation of lignin-based porous carbon as an efficient absorbent for the removal of methylene blue. Ind. Crop. Prod. 2021, 171, 113980. [Google Scholar] [CrossRef]
  2. Luo, L.; Wu, X.; Li, Z.; Zhou, Y.; Chen, T.; Fan, M.; Zhao, W. Synthesis of activated carbon from biowaste of fir bark for methylene blue removal. R. Soc. Open Sci. 2019, 6, 190523. [Google Scholar] [CrossRef] [Green Version]
  3. Charola, S.; Patel, H.; Chandna, S.; Maiti, S. Optimization to prepare porous carbon from mustard husk using response surface methodology adopted with central composite design. J. Clean. Prod. 2019, 223, 969–979. [Google Scholar] [CrossRef]
  4. 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]
  5. Sharma, K.; Sharma, S.; Sharma, V.; Mishra, P.K.; Ekielski, A.; Sharma, V.; Kumar, V. Methylene blue dye adsorption from wastewater using hydroxyapatite/gold nanocomposite: Kinetic and thermodynamics studies. Nanomaterials 2021, 11, 1403. [Google Scholar] [CrossRef]
  6. Zhang, N.; Cheng, N.; Liu, Q. Functionalized biomass carbon-based adsorbent for simultaneous removal of Pb2+ and MB in wastewater. Materials 2021, 14, 3537. [Google Scholar] [CrossRef]
  7. Katheresan, V.; Kansedo, J.; Lau, S.Y. Efficiency of various recent wastewater dye removal methods: A review. J. Environ. Chem. Eng. 2018, 6, 4676–4697. [Google Scholar] [CrossRef]
  8. Badawi, A.K.; Abd Elkodous, M.; Ali, G.A. Recent advances in dye and metal ion removal using efficient adsorbents and novel nano-based materials: An overview. RSC Adv. 2021, 11, 36528–36553. [Google Scholar] [CrossRef]
  9. ben Mosbah, M.; Mechi, L.; Khiari, R.; Moussaoui, Y. Current state of porous carbon for wastewater treatment. Processes 2020, 8, 1651. [Google Scholar] [CrossRef]
  10. Wu, Z.; Meng, Z.; Yao, C.; Deng, Y.; Zhang, G.; Wang, Y. Rice husk derived hierarchical porous carbon with lightweight and efficient microwave absorption. Mater. Chem. Phys. 2022, 275, 125246. [Google Scholar] [CrossRef]
  11. Sharma, P.; Singh, D.; Minakshi, M.; Quadsia, S.; Ahuja, R. Activation-induced surface modulation of biowaste-derived hierarchical porous carbon for supercapacitors. ChemPlusChem 2022, 87, e202200126. [Google Scholar] [CrossRef]
  12. Zhang, W.; Liu, M.; Zhao, Y.; Liao, Q. Facile preparation of porous carbon derived from pomelo peel for efficient adsorption of methylene blue. Molecules 2022, 27, 3096. [Google Scholar] [CrossRef]
  13. Tian, J.; Zhang, T.; Talifu, D.; Abulizi, A.; Ji, Y. Porous carbon materials derived from waste cotton stalk with ultra-high surface area for high performance supercapacitors. Mater. Res. Bull. 2021, 143, 111457. [Google Scholar] [CrossRef]
  14. Deng, Q.; Liu, H.; Zhou, Y.; Luo, Z.; Wang, Y.; Zhao, Z.; Yang, R. N-doped three-dimensional porous carbon materials derived from bagasse biomass as an anode material for K-ion batteries. J. Electroanal. Chem. 2021, 899, 115668. [Google Scholar] [CrossRef]
  15. Matsagar, B.M.; Yang, R.X.; Dutta, S.; Ok, Y.S.; Wu, K.C.W. Recent progress in the development of biomass-derived nitrogen-doped porous carbon. J. Mater. Chem. A 2021, 9, 3703–3728. [Google Scholar] [CrossRef]
  16. Jo, Y.; Kim, S.M.; Choi, H.; Yang, J.W.; Lee, B.C.; Cho, W.K. Sweet potato viromes in eight different geographical regions in Korea and two different cultivars. Sci. Rep. 2020, 10, 1–16. [Google Scholar] [CrossRef] [Green Version]
  17. Dai, T.; Wang, J.; Dong, D.; Yin, X.; Zong, C.; Jia, Y.; Shao, T. Effects of brewers’ spent grains on fermentation quality, chemical composition and in vitro digestibility of mixed silage prepared with corn stalk, dried apple pomace and sweet potato vine. Ital. J. Anim. Sci. 2022, 21, 198–207. [Google Scholar] [CrossRef]
  18. Wang, T.; Dong, X.; Jin, Z.; Su, W.; Ye, X.; Dong, C.; Lu, Q. Pyrolytic characteristics of sweet potato vine. Bioresour. Technol. 2015, 192, 799–801. [Google Scholar] [CrossRef]
  19. Gao, S.; Li, L.; Geng, K.; Wei, X.; Zhang, S. Recycling the biowaste to produce nitrogen and sulfur self-doped porous carbon as an efficient catalyst for oxygen reduction reaction. Nano Energy 2015, 16, 408–418. [Google Scholar] [CrossRef]
  20. Angin, D. Production and characterization of activated carbon from sour cherry stones by zinc chloride. Fuel 2014, 115, 804–811. [Google Scholar] [CrossRef]
  21. Liu, J.; Li, H.; Zhang, H.; Liu, Q.; Li, R.; Li, B.; Wang, J. Three-dimensional hierarchical and interconnected honeycomb-like porous carbon derived from pomelo peel for high performance supercapacitors. J. Solid State Chem. 2018, 257, 64–71. [Google Scholar] [CrossRef]
  22. Zhang, W.; Zhang, L.Y.; Zhao, X.J.; Zhou, Z. Citrus pectin derived porous carbons as a superior adsorbent toward removal of methylene blue. J. Solid State Chem. 2016, 243, 101–105. [Google Scholar] [CrossRef]
  23. Zheng, J.; Yan, B.; Feng, L.; Zhang, Q.; Zhang, C.; Yang, W.; Han, J.; Jiang, S.; He, S. Potassium citrate assisted synthesis of hierarchical porous carbon materials for high performance supercapacitors. Diam. Relat. Mater. 2022, 128, 109247. [Google Scholar] [CrossRef]
  24. Vafaeinia, M.; Khosrowshahi, M.S.; Mashhadimoslem, H.; Emrooz, H.B.M.; Ghaemi, A. Oxygen and nitrogen enriched pectin-derived micro-meso porous carbon for CO2 uptake. RSC Adv. 2022, 12, 546–560. [Google Scholar] [CrossRef]
  25. Miao, Q.; Xu, Y.; Kang, R.; Peng, H. Foaming waste wine yeast mud into nitrogen doped porous carbon framework by recyclable activator for high specific energy supercapacitor. J. Energy Storage 2022, 55, 105590. [Google Scholar] [CrossRef]
  26. Zhu, G.; Liu, Q.; Cao, F.; Qin, Q.; Jiao, M. Silkworm cocoon derived N, O-codoped hierarchical porous carbon with ultrahigh specific surface area for efficient capture of methylene blue with exceptionally high uptake: Kinetics, isotherm, and thermodynamics. RSC Adv. 2019, 9, 33872–33882. [Google Scholar] [CrossRef]
  27. Guo, H.; Li, Y.; Wang, C.; He, L.; Li, C.; Guo, Y.; Zhou, Y. Effect of the air oxidation stabilization of pitch on the microstructure and sodium storage of hard carbons. New Carbon Mater. 2021, 36, 1073–1078. [Google Scholar] [CrossRef]
  28. Zhang, S.; Zeng, M.; Li, J.; Li, J.; Xu, J.; Wang, X. Porous magnetic carbon sheets from biomass as an adsorbent for the fast removal of organic pollutants from aqueous solution. J. Mater. Chem. A 2014, 2, 4391–4397. [Google Scholar] [CrossRef]
  29. Cheng, S.; Zhang, L.; Ma, A.; Xia, H.; Peng, J.; Li, C.; Shu, J. Comparison of activated carbon and iron/cerium modified activated carbon to remove methylene blue from wastewater. J. Environ. Sci. 2018, 65, 92–102. [Google Scholar] [CrossRef]
  30. Zhang, Y.; Ji, G.; Li, C.; Wang, X.; Li, A. Templating synthesis of hierarchical porous carbon from heavy residue of tire pyrolysis oil for methylene blue removal. Chem. Eng. J. 2020, 390, 124398. [Google Scholar] [CrossRef]
  31. Hao, Y.; Wang, Z.; Wang, Z.; He, Y. Preparation of hierarchically porous carbon from cellulose as highly efficient adsorbent for the removal of organic dyes from aqueous solutions. Ecotox. Environ. Safe. 2019, 168, 298–303. [Google Scholar] [CrossRef] [PubMed]
  32. Zhang, W.; Li, H.; Tang, J.; Lu, H.; Liu, Y. Ginger straw waste-derived porous carbons as effective adsorbents toward methylene blue. Molecules 2019, 24, 469. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Jawad, A.H.; Abdulhameed, A.S.; Wilson, L.D.; Syed-Hassan, S.S.A.; ALOthman, Z.A.; Khan, M.R. High surface area and mesoporous activated carbon from KOH-activated dragon fruit peels for methylene blue dye adsorption: Optimization and mechanism study. Chin. J. Chem. Eng. 2021, 32, 281–290. [Google Scholar] [CrossRef]
  34. Chen, S.; Chen, G.; Chen, H.; Sun, Y.; Yu, X.; Su, Y.; Tang, S. Preparation of porous carbon-based material from corn straw via mixed alkali and its application for removal of dye. Colloids Surfaces A 2019, 568, 173–183. [Google Scholar] [CrossRef]
  35. Novais, R.M.; Caetano, A.P.; Seabra, M.P.; Labrincha, J.A.; Pullar, R.C. Extremely fast and efficient methylene blue adsorption using eco-friendly cork and paper waste-based activated carbon adsorbents. J. Clean. Prod. 2018, 197, 1137–1147. [Google Scholar] [CrossRef]
  36. Bedin, K.C.; Souza, I.P.; Cazetta, A.L.; Spessato, L.; Ronix, A.; Almeida, V.C. CO2-spherical activated carbon as a new adsorbent for methylene blue removal: Kinetic, equilibrium and thermodynamic studies. J. Mol. Liq. 2018, 269, 132–139. [Google Scholar] [CrossRef]
  37. Lonappan, L.; Rouissi, T.; Das, R.K.; Brar, S.K.; Ramirez, A.A.; Verma, M.; Surampalli, R.Y.; Valero, J.R. Adsorption of methylene blue on biochar microparticles derived from different waste materials. Waste Manag. 2016, 49, 537–544. [Google Scholar] [CrossRef] [PubMed]
  38. Hua, Y.; Xiao, J.; Zhang, Q.; Cui, C.; Wang, C. Facile synthesis of surface-functionalized magnetic nanocomposites for effectively selective adsorption of cationic dyes. Nanoscale Res. Lett. 2018, 13, 99. [Google Scholar] [CrossRef]
  39. Xing, Y.; Liu, D.; Zhang, L.-P. Enhanced adsorption of Methylene Blue by EDTAD-modified sugarcane bagasse and photocatalytic regeneration of the adsorbent. Desalination 2010, 259, 187–191. [Google Scholar] [CrossRef]
  40. Qiao, X.-Q.; Hu, F.-C.; Tian, F.-Y.; Hou, D.-F.; Li, D.-S. Equilibrium and kinetic studies on MB adsorption by ultrathin 2D MoS2nanosheets. RSC Adv. 2016, 6, 11631–11636. [Google Scholar] [CrossRef]
  41. Bentahar, S.; Dbik, A.; El Khomri, M.; El Messaoudi, N.; Lacherai, A. Adsorption of methylene blue, crystal violet and congo red from binary and ternary systems with natural clay: Kinetic, isotherm, and thermodynamic. J. Environ. Chem. Eng. 2017, 5, 5921–5932. [Google Scholar] [CrossRef]
  42. Bradder, P.; Ling, S.K.; Wang, S.; Liu, S. Dye Adsorption on Layered Graphite Oxide. J. Chem. Eng. Data 2011, 56, 138–141. [Google Scholar] [CrossRef]
  43. Ji, Y.; Xu, F.; Wei, W.; Gao, H.; Zhang, K.; Zhang, G.; Xu, Y.; Zhang, P. Efficient and fast adsorption of methylene blue dye onto a nanosheet MFI zeolite. J. Solid State Chem. 2021, 295, 121917. [Google Scholar] [CrossRef]
  44. Sun, Z.; Qu, K.; Cheng, Y.; You, Y.; Huang, Z.; Umar, A.; Ibrahim, Y.S.A.; Algadi, H.; Castañeda, L.; Colorado, H.A.; et al. Corncob-derived Activated Carbon for Efficiently Adsorption Dye in Sewage. ES Food Agrofor. 2021, 4, 61–73. [Google Scholar] [CrossRef]
  45. Fan, X.; Wang, X.; Cai, Y.; Xie, H.; Han, S.; Hao, C. Functionalized cotton charcoal/chitosan biomass-based hydrogel for capturing Pb2+, Cu2+ and MB. J. Hazard. Mater. 2022, 423, 127191. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the preparation and adsorption of SPVPC.
Figure 1. Schematic representation of the preparation and adsorption of SPVPC.
Molecules 28 00819 g001
Figure 2. (A) SEM image, (B) nitrogen adsorption–desorption isotherm, (C) DFT pore size distribution curve, (D) XRD pattern, (E) Raman spectrum, and (F) FTIR spectrum of SPVPC.
Figure 2. (A) SEM image, (B) nitrogen adsorption–desorption isotherm, (C) DFT pore size distribution curve, (D) XRD pattern, (E) Raman spectrum, and (F) FTIR spectrum of SPVPC.
Molecules 28 00819 g002
Figure 3. Effect of (A) pH (c0: 300 mg L−1, 120 min, 25 °C), (B) adsorption time (c0: 300 mg L−1, pH: 12, 25 °C), (C) PFO, and (D) PSO kinetics; (E) Langmuir isotherm and (F) reusability of MB adsorption on SPVPC.
Figure 3. Effect of (A) pH (c0: 300 mg L−1, 120 min, 25 °C), (B) adsorption time (c0: 300 mg L−1, pH: 12, 25 °C), (C) PFO, and (D) PSO kinetics; (E) Langmuir isotherm and (F) reusability of MB adsorption on SPVPC.
Molecules 28 00819 g003
Figure 4. Adsorption mechanism diagram of SPVPC towards MB.
Figure 4. Adsorption mechanism diagram of SPVPC towards MB.
Molecules 28 00819 g004
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

Zhang, W.; Zhao, Y.; Liao, Q.; Li, Z.; Jue, D.; Tang, J. Sweet-Potato-Vine-Based High-Performance Porous Carbon for Methylene Blue Adsorption. Molecules 2023, 28, 819. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28020819

AMA Style

Zhang W, Zhao Y, Liao Q, Li Z, Jue D, Tang J. Sweet-Potato-Vine-Based High-Performance Porous Carbon for Methylene Blue Adsorption. Molecules. 2023; 28(2):819. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28020819

Chicago/Turabian Style

Zhang, Wenlin, Yuhong Zhao, Qinhong Liao, Zhexin Li, Dengwei Jue, and Jianmin Tang. 2023. "Sweet-Potato-Vine-Based High-Performance Porous Carbon for Methylene Blue Adsorption" Molecules 28, no. 2: 819. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules28020819

Article Metrics

Back to TopTop