Next Article in Journal
Extreme Hydrometeorological Conditions of Sediment Waves’ Formation and Migration in Peter the Great Bay (The Sea of Japan)
Next Article in Special Issue
Municipal Sewage Treatment Technology: A2/O-VMBR Integrated Technology for Municipal Treatment and Improved Pollutant Removal
Previous Article in Journal
Unveiling Distribution, Hydrogeochemical Behavior and Environmental Risk of Chromium in Tannery Wastewater
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis, Characterization and Performance Evaluation of Burmese Grape (Baccaurea ramiflora) Seed Biochar for Sustainable Wastewater Treatment

1
Department of Chemical Engineering, Bangladesh University of Engineering and Technology, Dhaka 1000, Bangladesh
2
Sanitary Environmental Engineering Division (SEED), Department of Civil Engineering, University of Salerno, via Giovanni Paolo II 132, 84084 Fisciano, SA, Italy
3
Department of Energy Science and Engineering, Khulna University of Engineering and Technology, Khulna 9203, Bangladesh
4
Hubei Provincial Engineering Laboratory for Clean Production and High Value Utilization of Bio-Based Textile Materials, Wuhan Textile University, Wuhan 430200, China
5
Department of Chemistry, Bangladesh University of Engineering and Technology, Dhaka 1000, Bangladesh
*
Authors to whom correspondence should be addressed.
Water 2023, 15(3), 394; https://doi.org/10.3390/w15030394
Submission received: 29 December 2022 / Revised: 12 January 2023 / Accepted: 14 January 2023 / Published: 18 January 2023
(This article belongs to the Special Issue Advances in Food and Textile Industry Wastewater Treatment)

Abstract

:

Highlights

What are the main findings?
  • FTIR analysis showed multiple surface functional groups, e.g., R-OH, OH, -C=O, -COOH, etc., on the biochar surface.
  • FESEM confirmed the randomized and porous tunnel-like structures of biochar.
  • The BET specific surface area was 19.90 ± 1.20 m2/g.
  • Kinetics and isotherm analysis showed alignment with pseudo-second-order kinetic and Langmuir Isotherm models.
What is the implication of the main finding?
  • ~85% removal and 166.30 mg/g sorption capacity were obtained.
  • The possible interaction of MB and BGS-derived biochar can be characterized by elec-trostatic attraction, H-bond and π–π conjugation.

Abstract

Biochar prepared from different bio-sources serves as a feasible solution for the decontamination of dye-contaminated wastewater. In this study, biochar was synthesized from a sustainable source, i.e., local fruit waste, Burmese grape seeds (BGSs). The seeds were collected from a local market, washed, pre-teated and finally converted into biochar by pyrolysis in a N2 furnace. The removal efficiency of the synthesized biochar was evaluated towards a cationic industrial azo dye, methylene blue (MB). The phosphoric acid (H3PO4) and potassium hydroxide (KOH) pretreated BGS were pyrolized at 500 °C for 3 h in a N2 furnace at a heating rate of 10 °C/min. The spectroscopic analysis confirmed the presence of multiple surface functional groups, e.g., R-OH, OH, -C=O, -COOH, etc. The surface of the biochar was randomized with porous tunnel-like structures. The specific surface area and pore volume obtained from BET analysis were 19.90 ± 1.20 m2/g and 5.85 cm3/g. The MB concentration (mg/L), contact duration (min) and pH were varied to assess the MB sorption phenomena. The optimum pH was found to be 8. During the first 20 min of contact time, adsorption was rapid and equilibrium was reached after 75 min. The adsorption was best described by pseudo-first-order kinetics with a good fit (R2 = 0.99). The maximum removal percentage was ~85%, and per gram of BGS can adsorb 166.30 mg of MB, which supports the Langmuir adsorption isotherm model. The obtained results were compared with the reported literature, and BGS showed its excellent candidacy to be industrially utilized in the tertiary stage of wastewater treatment plants.

1. Introduction

Chemical industries now comprise a significant component of the global economy in the 21st century. Bangladesh has embraced industrialization that is focused on exports. Its main export industries include leather goods, fish, seafood, shipbuilding, jute and textiles, with the textile industry dominating in recent years [1]. However, these textile industries have caused numerous environmental hazards, with water pollution being one of the most severe problems [2]. The effluents from these industries contain numerous untreated toxic chemicals [3]. Organic dyes and highly hazardous carcinogenic contamination are among the major pollutants in the effluent of such industries. Approximately 100,000 commercially available dyes are used to manufacture approximately 7 × 105 tonnes of reactive dyes per year, according to estimates [4]. The textile, leather, garments and food industries are considered to be the primary sources of dye wastewater. The most used dye for silk, cotton and wool is methylene blue (MB). Eye burns brought on by MB have the potential to permanently harm both human and animal eyes. It produces a burning sensation when consumed via the mouth and may cause methemoglobinemia, vomiting, nausea, increased sweating, gastritis and diarrhoea. It can also cause short periods of fast or difficult breathing when inhaled [5]. The treatment of wastewater carrying such colour is crucial because it has an aesthetic impact on receiving rivers.
Over the years, several methods have been developed by researchers, the most reported ones being sedimentation [6], membrane separation [7], biological treatment [8], chemical oxidation [9], coagulation–flocculation [10], photocatalysis [11,12,13,14], advanced oxidation [15,16] and adsorption [17,18,19,20,21,22,23,24]. Out of these techniques, the adsorption process is considered superior and is an extensively used technique for the advantages of high removal efficiency, low initial and maintenance cost, ease of operation, regeneration of adsorbent and simple design [25].
The choice of adsorbents mainly depends on the availability, affordability, adsorption capacity and kinetics, selectivity of particular pollutants and ease of regeneration [26]. Due to its low cost, straightforward construction, the convenience of use, insensitivity to hazardous pollutants and low concentrations of dangerous compounds, adsorption has been viewed as an appealing way to remove MB from aqueous systems [27]. Palm kernel fiber [28], fly ash [29], swelling clays [30], montmorillonite [31], pyrolyzed petrified sediment [32], mesoporous metal-oxide-doped silica nanocomposites [33,34,35], recycled nano-silica [36], silkworm exuviae [37], iron terephthalate [38], walnut shell [39], coal-based materials [40], carbon nanotubes (CNTs) [41], graphene materials [42] and metal–organic frameworks (MOFs) [43] have all been studied for the treatment of dye-contaminated wastewater. However, the reported sorbents are not always cost-friendly and reusable [44]. In recent years, biomass-based materials have been utilized in different applications [45]. Biochar, due to its remarkable effectiveness in eliminating a variety of contaminants, has drawn a lot of interest from researchers. Biochar has several applications in the field of wastewater treatment, energy production, CO2 sequestration, soil remediation, supercapacitor generation, etc. [46,47,48,49,50,51]. Biochar has a porous and extremely amorphous structure. Its microscopic and macroscopic structure exhibits high levels of surface reactivity, mechanical stability and chemical stability [21]. Its surface activity, which depends on its pore properties and the existence of surfaces with functional groups, such as -C=O, -COOH, -OH and lactone groups, has a significant impact on its capacity to adsorb [52]. These functional groups account for the surface structure’s strong attraction to different dyes and other substances. As a result, it ranks among the best and most versatile adsorbents for water filtration [53]. Hydrogen bonding, electrostatic interaction, pi–pi interactions and hydrophilic–hydrophobic interaction are all probable mechanisms by which the reactive dye molecules engage with the biochar [54]. The primary source’s elemental composition also affects how well biochar performs as an adsorbent [52]. Utilizing readily accessible agricultural by-products as a raw material has recently been shown to be highly economically appealing. In recent years, several bio-waste materials, such as coconut shell [55], banana peel [56], pineapple stem [57], moringa oleifera seeds [22,23], rice straw [58], orange peel [59], wheat straw [60], etc., have been studied for this purpose. The source material and manufacturing technique significantly impact how the activated biochar’s pores develop and, therefore, how it ultimately behaves [61]. Hence, this study implemented the use of BGS, which is a bio-waste generated from the fruits of Burmese Grapes. It belongs to the family Phyllanthaceae, primarily grown in South Asian countries, e.g., Burma, India, Bangladesh, Vietnam and Malaysia. This abundantly available bio-waste can be highly economically beneficial in producing an efficient adsorbent for water purification. Thus, the choice of this bio-source as a raw material for biochar production can be justified by its availability in the local market, cheap or no cost and seed to fruit ratio.
The Burmese grape (Baccaurea ramiflora), a green tree with an overhead and thin skin that grows slowly and may reach heights of 25 m, is a member of the Phyllanthaceae family. This plant is widely cultivated in India and Malaysia and can be found across South Asia [62]. The fruit is gathered and used to make wine, edible fruit and medication to cure skin conditions. The wood, roots and bark are gathered because they offer therapeutic qualities. Traditional Chinese Dai medicine has traditionally used the whole plant [63]. Numerous scientific studies have demonstrated the presence of various bioactive substances in plants and their products, such as fruits, peels, leaves and seeds, which provide myriad health benefits and protection against degenerative illnesses [64]. However, until now, no study has been found to utilize its waste for sustainable wastewater treatment. The biodegradable seeds of this fruit can be utilized to be transformed into carbonaceous material, which can be used for organic contaminant removal from aqueous solution. The non-toxic material is not harmful to the aquatic environment and can be an eminent source of adsorbent for industrial application.
With this goal in mind, this work focuses on using biochar made from Burmese grape seeds to remove MB from an aqueous medium. As per our knowledge, this work is the first attempt to use and characterize BGS-derived biochar and assess its applicability for wastewater treatment.In order to determine the ideal operating conditions for dye removal, process parameters, including initial dye concentration, contact duration and pH, were changed. For an in-depth analysis on the sorption system and its mechanism, kinetic and isotherm studies were executed. Comparative analysis of MB removal by different biochar and the probable mechanism of MB removal onto BGS-derived biochar were examined and discussed.

2. Materials and Methods

Burmese grape seeds were collected from the local market. The other reagents used were: (i) phosphoric acid (H3PO4), (ii) potassium hydroxide (KOH) and (iii) methylene blue (MB). All of these reagents were purchased from Sigma Aldrich, and DI water was used to prepare the solutions.

2.1. Preparation of BGS Biochar

The collected BGSs were chopped into small parts. Then, the sample was air-dried for 2 days to remove the surface-captured moisture. BGS was then mixed with the solution of H3PO4 (40%) in a weight ratio of 1:2 followed by drying in an air oven at 80 °C for 3 h. The dried samples were then mixed with 50% (w/v) KOH solution to increase the functional groups in the structure. Then, it was dried at 100 °C for 2 h. Finally, the particles were carbonized under a nitrogen gas flow at a temperature of 500 °C for 3 h to be converted into biochar. The samples were further used for characterizations and adsorption experiments.

2.2. Characterizations

To check the chemical functionality of the prepared sample, it was examined using Fourier-transform infrared spectroscopy (FTIR). To create pellets, a small quantity of the produced material was combined with KBr powder. FTIR spectra were captured in transmission mode using an FTIR-8400 analyzer in the 4000–400 cm−1 region by a spectrometer from Shimadzu, Japan. SEM pictures were taken using a field-emission scanning electron microscope (FESEM). An FESEM instrument procured from JEOL Ltd. in Tokyo, Japan, was used to assess the surface morphological characteristics and elemental composition analysis. Brunauer–Emmett–Teller (BET) surface area analysis was conducted in a PulseChemiSorb 2705 instrument.

2.3. Experimental Design

The process variables, e.g., dye solution concentration (mg/L), contact duration (min) and pH, were tuned to assess the MB removal characteristics of the BGS biochar. The biochar dose for each experiment was 5 mg. The influence of pH on adsorption was investigated by altering the solution pH between 3.0 and 9.0. To alter the pH of the solution, 0.1 M NaOH and HCl solution was utilized. Kinetic tests were performed using an MB dye concentration of 60 mg/L at ambient temperature (27 ± 2 °C) and examined at a time ranging from 5 to 90 min. With initial dye concentrations of Co = 10–80 mg/L for 90 min, concentration dependency and isotherm tests were performed. According to the Equations, the dye’s removal (%) and adsorption capacity (q, mg/g) were determined. According to Equations (1) and (2) [65]:
Removal , % = ( C i C e ) × 100 C i
Adsorption   capacity ,   q = ( C i C e ) × V m
Specifically, Ci and Ce represent the initial and equilibrium concentrations, whereas m is the adsorbent dosage and V is the solution volume.
The most widely accepted kinetic models for studying the adsorption processes are the pseudo-first-order and pseudo-second-order models. These models were utilized to comprehend the kinetic behaviour of the adsorption process of the BGS-derived biochar particles. Adsorption is predicated on chemisorption in the pseudo-second-order model, whereas the pseudo-first-order model implies that the adsorption is regulated by the molecular diffusion on the surface and pores of the sorbents.
The pseudo-first-order model considers diffusion to be the dominating factor for adsorption. The following Equation (3) is an illustration of the integrated pseudo-first-order equation:
ln ( q e q t ) = ln q e k 1 t   i
In this equation, the 1st-order rate constant, k1 (min−1), can be calculated by taking the sorption data, e.g., sorption capacity at equilibrium, qe (mg/g), and at any time, t (min), qt (mg/g).
The pseudo-second-order model considers chemisorption as the prime factor for adsorption. It can be described by Equation (4),
t q t = 1 k 2 q e 2 + t q e  
The calculated parameter, k2, is the pseudo-second-order rate constant, qe and qt, are the same as defined previously.
To characterize the dye-removal process by BGS-generated biochar, isotherm analysis was carried out. In order to explain the adsorption behaviour, two key isotherm models, Langmuir and Freundlich, were employed to linearize the experimental data. Monolayer adsorption is the main consideration in the Langmuir isotherm model, whereas multilayer adsorption on a heterogenous adsorption site is the prime consideration of the Freundlich model. The following are the linear expressions for the Langmuir (5) and Freundlich (6) isotherms:
C e q e = 1 b L q m   + C e q m    
ln q e = 1 n ln C e + ln k f  
In Equations (5) and (6), equilibrium dye concentration, Ce (mg/L), and equilibrium sorption capacity, qe (mg/g), were used. The Langmuir sorption constant, bL and maximum sorption capacity, qm (mg/g), in Equation (5), and Freundlich adsorption constant, kf, and heterogeneity factor, n, in Equation (6) are the calculated parameters, respectively.

3. Results and Discussions

3.1. Biochar Characterizations

Pyrolysis is a practical approach to use wastes, e.g., municipal solid wastes (MSWs), and biowastes in an environment-friendly technique to prepare biochar. Herein, dried seeds were reacted with H3PO4 and KOH to activate and reveal their functional groups. The impurities (metals and organics) are leached in the H3PO4, and KOH helps the carbonaceous structure to disintegrate. Then, the pre-treated BGSs were pyrolyzed to become biochar via the following steps shown in Figure 1.
The diffusion of stack gases (CO, H2, N2) perforated the biochar structure producing highly macro- and mesoporous biochar (micropores are hard to identify). The FTIR spectra of the material were captured using a spectrophotometer and are shown in Figure 2A to verify the synthesis of biochar.
The broad and strong peaks within 3100–3600 cm−1 correspond to the O–H stretching mode of hydroxyl groups present in organic compounds, such as alcohols, phenols and carboxylic acids [66]. The C–H stretching (symmetric and asymmetric) vibrations for different aliphatic groups may account for the peak at roughly 2900 cm−1 [67]. The aromatic C=C and C=O stretching of ketones and quinones may provide peaks in a range of 1550 cm−1 to 1680 cm−1 [68]. The peak demonstrates the lactone structure O=C-O vibration at 1400 cm−1 [69]. In contrast to the 700 cm−1 (broad) cis C–H out-of-plane bend, bands between 970 and 960 cm−1 correspond to trans C–H out-of-plane bend vibrations. The peak around 1220–1180 cm−1 may be related to the stretching modes of the hydrogen-bonded P=O, the O–C stretching vibrations in the P–O–C linkage and the P=OOH that arises owing to the presence of residual H3PO4 in the structure during pyrolysis [70]. A peak at 1300 cm−1 indicates the deformation of the aliphatic CH3. Near the apex, which is between 1100 and 1240 cm−1, fragrant CO stretches. Aliphatic ether and alcohols exhibit symmetric C–O stretching, which is shown by the peak at 1021 cm−1. The aromatic C–H deformation modes are shown by peaks about 670 cm−1 bands [67]. The functional groups present in the BGS biochar surface highly influenced the sorption process [71].
The FESEM image in Figure 2B demonstrates that the prepared biochar had a hierarchical porosity and a three-dimensional cage-like structure. The biochar exhibited a complicated and changeable structure. The image shows that the biochar surface was porous and rough, with various-sized cracks, fissures and holes. The alignment of these pores was irregular, and a limited number of holes with various diameters likely appeared in the generated biochar as a result of the volatilization and dehydration of moisture and organics [72]. In the pre-treatment process, utilization of KOH may initiate the following reactions with residual KOH [73].
4 KOH + C → K2CO3 + K2O + 2 H2
Dehydration of KOH took place at temperatures lower than 500 °C after the reaction shown in Reaction 2.
2 KOH → K2O + 2 H2O
The reforming processes were maintained by the water vapor produced ((R3) and (R4)),
H2O + C → CO + 2 H2
CO + H2O → CO2 + 2 H2
According to reaction theory, the potassium oxide and biochar dioxide subsequently interacted to generate potassium carbonate (R5).
K2O + CO2 → K2CO3
These SEM images show that the surface of the biochar sample was rough, porous and ruptured because of the release of large volatile molecules during the calcination process [74]. The pores on the biochar surface resemble circles and ellipses most of the time. Among the types of pore size, they are more likely macropores. In Figure 2B, complicated coral-structure macropores are observed on the surface of the BGS-derived biochar. It may be possible that these macropores and associated tunnel structures are formed due to the hydrothermal conversion stage [75]. The sorption of MB by BGS-derived biochar may be highly influenced by these macropores and rough surface of the biochar [76] in the pore diffusion stage.
Table 1 presents the elemental composition and BET analysis of the BGS-derived biochar.
The C, N, O, K and P content of BGS-derived biochar obtained from EDS analysis was found to be 70.02, 17.05, 12.08, 0.35 and 0.50%, respectively. The presence of K and P shows the effect of pre-activation. The O and N content may be associated with oxygenated and nitrous functional groups in the biochar surface. The specific surface area and pore volume obtained from BET analysis were 19.90 ± 1.20 m2/g and 5.85 cm3/g, respectively. However, the literature reported a much higher surface area for some synthesized biochar than the obtained surface area in our case. Therefore, the sorption of MB by BGS-derived biochar was possibly controlled combinedly by the effect of MB interaction with surface functional groups of biochar and pore diffusion of MB through biochar mesopores.

3.2. Effect of pH, Time and Concentration

The dye solution’s pH level significantly impacts the adsorption phenomena by changing the adsorbents’ surface-active sites and the ionization of the molecules. Figure 3A displays the removal rate of MB by BGS-derived AC at various pH values, ranging from 3 to 9. Because the negatively charged OH ion predominates on the adsorbent surface, previous studies have demonstrated that positively charged (cationic) dyes may adsorb more easily at higher pH values. It encourages the cationic dye molecules to engage electrostatically with the active/reactive regions of the adsorbent [77]. In this investigation, a similar effect was shown.
The maximum amount of dye was removed around pH 8.0, denoting that the electrostatic attraction between the cationic portion of the dye and the negatively charged sites (-OH, -COOH) on the surface of the functionalized biochar was the primary factor in the adsorption mechanism. The removal rate increased quickly in a pH range of 4–7. Positively charged dye molecules make more electrostatic interaction with negatively charged adsorbent surfaces with a rise in pH, leading to a corresponding increase in adsorption [78]. The competing adsorption between the proton and dye reduces the dye adsorption at comparatively lower pH levels [79]. The electrostatic interaction here between positively charged MB dye and the negatively charged sorbent surface rises with a pH increase, leading to a corresponding increase in adsorption [80]. The adsorption efficiency almost stayed the same when pH rose from 8.0 to 9.0.
Between 5 and 90 min, the adsorption of MB by AC generated from BGS was examined. According to Figure 3B, the adsorption capacity rose sharply in the first 30 min before gradually rising over time. After 75 min, a balance was attained. A reduction in adsorption rate at greater concentrations occurs because the adsorbates eventually occupy unoccupied active sites that were once free to occupy [81]. After 30 min, the sluggish adsorption rate might be described by a process in which the dye molecules first come into contact with the boundary layer before slowly diffusing over the top of the AC to penetrate into the pore structure [82]. With time, the molecular diffusion between the dye molecules and the AC surface slowed. The system finally reached saturation, suggesting that all of the active sites were engaged and that the system had reached an equilibrium state [83]. Once the equilibrium had been reached, the impact of contact time was not apparent.
Figure 4A displays the results of the adsorption of MB onto BGS-derived biochar at varied initial dye solution concentrations. The adsorption capacity of BGS-generated biochar rose from 42.75 mg/L to 150.75 mg/L when the starting dye concentration increased from 10 to 80 mg/L. This is a result of the driving factor between adsorbents (biochar) and dyes (MB) becoming stronger as the initial concentration rises [84]. Figure 4B represents the relationship between Ce and qe. At a certain Ce, the adsorption capacity reaches an optimum value due to the saturation of the active sites.

3.3. Kinetics and Isotherms Analysis

Kinetics and isotherm studies are performed in order to fully comprehend the mechanism governing the adsorption of MB onto BGS-generated biochar. To predict the interaction between MB and biochar, the kinetic behaviour of the adsorption procedure was examined. Unlike the pseudo-first-order model, the pseudo-second-order model can forecast how the system will behave throughout the entire contact duration. The model’s advantage is that it can determine the preliminary adsorption rate and capacity instead of first identifying any unidentified parameters [52]. The linearized plots are shown in Figure 5. The estimated correlation coefficients (R2), R2Pseudo 1st = 0.98 and R2Pseudo 2nd = 0.99 indicate that the pseudo-second-order model more closely matches our experiment than the pseudo-first-order model. The agreement of our experimental results with a pseudo-second-order model showed that chemisorption, which is characterized by the exchange or sharing of electrons, dominates adsorption. The attractive force between the positively charged MB and the negatively charged surface of the BGS-derived biochar may be used to explain the chemisorption process. Probably, the functional groups, e.g., -OH and -COOH, of the BGS-derived biochar surface interacted with the dye molecules, leading to the adsorption process. Several reported bio-materials have been found to follow pseudo-second-order kinetics for the removal of MB [85]. However, the close value of R2 for the two models in our case represented the dual mechanism of the adsorption process, which meant diffusion and chemical interaction both played import roles in dye capture on the adsorbent [86].
To characterize the MB adsorption by the produced BGS-derived biochar, adsorption isotherm studies were carried out. The experimental adsorption data were analyzed using two popular isotherm models, the Langmuir and Freundlich models, which are frequently employed to characterize adsorption behavior (Figure 6). According to Langmuir, when a dye molecule is at an active site on a surface, no more adsorption occurs there. The model has a dynamic equilibrium point when adsorption and desorption rates are identical. The physical importance of this occurrence is that it has reached saturation at all active spots on the surface [87]. The correlation coefficient value (R2) was found to be 0.99 for the Langmuir model and 0.95 for the Freundlich model. However, Using the Langmuir model, the maximum adsorption capacity (qm) was determined, which was 166.3 mg/g. This calculated qm is very close to the experimental saturation adsorption capacity (qe), which was 150.75 mg/g. For the dye adsorption isotherm, it can be argued that the Langmuir model fits better, suggesting that the adsorption process is monolayer adsorption on the heterogeneous deformed graphitic structure of BGS-generated biochar. The surface charges and porous cracks may play a role in the monolayer adsorption of dye particles [88].
The MB sorption results of different biochar prepared from diverse sources are presented in Table 2.
In Table 2, several sources of biochar synthesis, different preparation conditions, experimental conditions for MB removal and maximum sorption capacity are listed. Wide-ranging studies are reported in the literature; however, research on biochar potentiality for organic contaminant sequestration is still ongoing. The synthesized BGS-derived biochar showed excellent potentiality compared to different reported biochar for MB removal. Different straws, fruit shells, seeds, leaves, solid wastes, etc., have been applied to prepare biochar, as the choice of the parent source for biochar preparation depends on the availability, function ability, cost feasibilities, etc. Therefore, low-cost and highly abundant BGS is an excellent source for biochar synthesis and, in the future, its different applicability, e.g., efficiency in real waste water, photocatalytic activity, nitrogen/phosphorus co-doping, capacitive deionization potential, extraction of radioactive materials, etc., must be assessed [98,99].

4. Adsorption Mechanism

The prediction of the mechanism or interaction routes for the adsorbent–contaminant system is critical to develop an improved version of the catalyst (adsorbent) [100,101]. Figure 7 illustrates various adsorption routes, which include pore saturation via the mesopores of the biochar, H bonds, e-static interactions and π–π conjugation of the surface functional groups and aromatic moieties of the biochar with the MB molecules [24]. Interfacial layers, which may be split into the surface of the biochar and MB, from where the adsorptive motive generates, are the frequent points for adsorption. Relying on the type of forces involved, they may be categorized as either physical or chemical [102].
The micropore and mesopore filling process provides diffusion pathways for dye molecules to move into surface pores. The SEM image detects that the surface of BGS-derived biochar is rough and disordered with coral-shaped mesopores. The higher adsorption capability in BGS-derived biochar may be caused by the MB molecules diffusing through these pores [103]. The hydroxyl groups in the attachment of -COOH groups may produce the C=O groups of the biochar surface. As an alternative, the oligosaccharides or monosaccharides would undergo intramolecular dehydration, which would result in C=C bonds and aromatic character. These species are associated with the functional groups of surface-oxygenated functional groups [104]. The surface of the biochar exhibited several functional groups, including R-OH, -COOH and -OH, and structures, such as phenol, quinone and esters, according to the FTIR spectra. These functional groups and aromatic compounds directly influence the adsorption process, which also serve as the building blocks for different interactions [105]. In this work, a basic solution with a pH of 8 was shown to promote the greatest dye adsorption onto biochar made from BGS. The oxygenated biochar surface promotes the deprotonation of carboxylic, hydroxyl and ester groups at basic pH because -OH groups are present in the solution [106], which enhanced the electrostatic force of +ve MB molecules and −ve biochar surface [107]. The electronegative N atom (MB molecules) and polar H atoms in the carboxylic (-COOH), hydroxyl (-OH), epoxy (-O-), carbonyl C=O), etc., groups of the biochar surface may also contribute to the sorption process [108]. The reported literature on carbonaceous materials with post-activation of KOH also supports similar observations as ours for the sorption of organic components [109]. Similar interactions are commonly seen when MB dye is adsorbed by different carbon materials, such as low-rank coal [110] and bamboo chips [111], etc.

5. Conclusions

In this work, biochar was synthesized from a bio-source, Burmese grape seeds (BGSs), and characterized with FT-IR and FESEM. To determine the produced biochar’s’ adsorptive characteristics, adsorption tests were conducted. It was discovered that changes in process variables, such as contact duration, dye concentration and pH, significantly influenced adsorption. The sorption rate was very high for the first 20 min, and equilibrium was established with a clearance percentage of around 85%. The adsorption process is boosted at a pH value of 8 by improved interactions (+ve/−ve interactions, electronegative bonding and aromatic conjugations) between the MB molecules and biochar surface. However, after the blocking and saturation of the sorption sites of the biochar surface, sorption achieves a maximum value (qm). The sorption kinetics followed the pseudo-second-order kinetic nature, and the Langmuir isotherm model described the sorption phenomenon well. The highest sorption of 166.30 mg/g was obtained from the Langmuir model, and from experimental data, it was 150.75 mg/g. Therefore, the synthesis of biochar from BGS has advantages of simultaneous utilization of biowaste and wastewater treatment, which can be realized on an industrial scale in the future.

Author Contributions

Conceptualization, H.R., M.N.P., S.H.F. and M.S.I.; methodology, H.R., D.S. and S.P.; formal analysis, H.R. and M.N.P.; resources, V.N., Y.C., S.H.F. and M.S.I.; writing—original draft preparation, H.R., D.S. and S.P.; writing—review and editing, H.R., M.N.P., V.N., Y.C. and S.H.F.; supervision, M.N.P., V.N., S.H.F. and M.S.I., funding acquisition, M.N.P. and V.N. All authors have read and agreed to the published version of the manuscript.

Funding

The research is funded by support from the Sanitary Environmental Engineering Division (SEED) and grants (FARB projects) from the University of Salerno, Italy, coordinated by V. Naddeo. Grant Number: 300393FRB22NADDE.

Data Availability Statement

The raw data will be available on reasonable request from the corresponding author (M.S.I).

Acknowledgments

The authors acknowledge the Department of Chemical Engineering, BUET, for providing the lab facilities and the Department of Chemistry for supporting the characterizations.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Suhan, M.B.K.; Shuchi, S.B.; Al-Mamun, M.R.; Roy, H.; Islam, M.S. Enhanced UV light-driven photocatalytic degradation of methyl orange using MoO3/WO3-fluorinated TiO2 nanocomposites. Environ. Nanotechnol. Monit. Manag. 2022, 19, 100768. [Google Scholar] [CrossRef]
  2. Lellis, B.; Fávaro-Polonio, C.Z.; Pamphile, J.A.; Polonio, J.C. Effects of textile dyes on health and the environment and bioremediation potential of living organisms. Biotechnol. Res. Innov. 2019, 3, 275–290. [Google Scholar] [CrossRef]
  3. Ilyas, M.; Ahmad, W.; Khan, H.; Yousaf, S.; Yasir, M.; Khan, A. Environmental and health impacts of industrial wastewater effluents in Pakistan: A review. Rev. Environ. Health 2019, 34, 171–186. [Google Scholar] [CrossRef] [PubMed]
  4. Ghosh, D.; Bhattacharyya, K.G. Adsorption of methylene blue on kaolinite. Appl. Clay Sci. 2002, 20, 295–300. [Google Scholar] [CrossRef]
  5. Özer, A.; Dursun, G. Removal of methylene blue from aqueous solution by dehydrated wheat bran carbon. J. Hazard. Mater. 2007, 146, 262–269. [Google Scholar] [CrossRef]
  6. Gadekar, M.R.; Ahammed, M.M. Use of water treatment residuals for colour removal from real textile dye wastewater. Appl. Water Sci. 2020, 10, 160. [Google Scholar] [CrossRef]
  7. Thamaraiselvan, C.; Noel, M. Membrane Processes for Dye Wastewater Treatment: Recent Progress in Fouling Control. Crit. Rev. Environ. Sci. Technol. 2015, 45, 1007–1040. [Google Scholar] [CrossRef]
  8. Kornaros, M.; Lyberatos, G. Biological treatment of wastewaters from a dye manufacturing company using a trickling filter. J. Hazard. Mater. 2006, 136, 95–102. [Google Scholar] [CrossRef]
  9. Pan, Y.; Zhu, T.; He, Z. Enhanced Removal of Azo Dye by a Bioelectrochemical System Integrated with a Membrane Biofilm Reactor. Ind. Eng. Chem. Res. 2018, 57, 16433–16441. [Google Scholar] [CrossRef]
  10. Moghaddam, S.S.; Moghaddam, M.R.; Arami, M. Coagulation/flocculation process for dye removal using sludge from water treatment plant: Optimization through response surface methodology. J. Hazard. Mater. 2010, 175, 651–657. [Google Scholar] [CrossRef]
  11. Akter, S.; Islam, M.S.; Kabir, M.H.; Shaikh, M.A.A.; Gafur, M.A. UV/TiO2 photodegradation of metronidazole, ciprofloxacin and sulfamethoxazole in aqueous solution: An optimization and kinetic study. Arab. J. Chem. 2022, 15, 103900. [Google Scholar] [CrossRef]
  12. Al-Mamun, M.R.; Islam, M.S.; Hossain, M.R.; Kader, S.; Islam, M.S.; Khan, M.Z.H. A novel and highly efficient Ag and GO co-synthesized ZnO nano photocatalyst for methylene blue dye degradation under UV irradiation. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100495. [Google Scholar]
  13. Al-Mamun, M.R.; Kader, S.; Islam, M.S. Solar-TiO2 immobilized photocatalytic reactors performance assessment in the degradation of methyl orange dye in aqueous solution. Environ. Nanotechnol. Monit. Manag. 2021, 16, 100514. [Google Scholar] [CrossRef]
  14. Al-Mamun, M.R.; Karim, M.N.; Nitun, N.A.; Kader, S.; Islam, M.S.; Khan, M.Z.H. Photocatalytic performance assessment of GO and Ag co-synthesized TiO2 nanocomposite for the removal of methyl orange dye under solar irradiation. Environ. Technol. Innov. 2021, 22, 101537. [Google Scholar] [CrossRef]
  15. Shuchi, S.B.; Suhan, M.B.K.; Humayun, S.B.; Haque, M.E.; Islam, M.S. Heat-activated potassium persulfate treatment of Sudan Black B dye: Degradation kinetic and thermodynamic studies. J. Water Process Eng. 2021, 39, 101690. [Google Scholar] [CrossRef]
  16. Suhan, M.B.K.; Shuchi, S.B.; Anis, A.; Haque, Z.; Islam, M.S. Comparative degradation study of remazol black B dye using electro-coagulation and electro-Fenton process: Kinetics and cost analysis. Environ. Nanotechnol. Monit. Manag. 2020, 14, 100335. [Google Scholar] [CrossRef]
  17. Islam, M.S.; Kwak, J.-H.; Nzediegwu, C.; Wang, S.; Palansuriya, K.; Kwon, E.E.; Naeth, M.A.; El-Din, M.G.; Ok, Y.S.; Chang, S.X. Biochar heavy metal removal in aqueous solution depends on feedstock type and pyrolysis purging gas. Environ. Pollut. 2021, 281, 117094. [Google Scholar] [CrossRef] [PubMed]
  18. Islam, M.S.; McPhedran, K.N.; Messele, S.A.; Liu, Y.; El-Din, M.G. Isotherm and kinetic studies on adsorption of oil sands process-affected water organic compounds using granular activated carbon. Chemosphere 2018, 202, 716–725. [Google Scholar] [CrossRef]
  19. Islam, M.S.; Roy, H.; Aftose, S. Phosphoric acid surface modified Moringa oleifera leaves biochar for the sequestration of methyl orange from aqueous solution: Characterizations, isotherm, and kinetics analysis. Remediation 2022, 32, 281–298. [Google Scholar] [CrossRef]
  20. Jahan, N.; Roy, H.; Reaz, A.H.; Arshi, S.; Rahman, E.; Firoz, S.H.; Islam, M.S. A comparative study on sorption behavior of graphene oxide and reduced graphene oxide towards methylene blue. Case Stud. Chem. Environ. Eng. 2022, 6, 100239. [Google Scholar] [CrossRef]
  21. Kwak, J.-H.; Islam, M.S.; Wang, S.; Messele, S.A.; Naeth, M.A.; El-Din, M.G.; Chang, S.X. Biochar properties and lead (II) adsorption capacity depend on feedstock type, pyrolysis temperature, and steam activation. Chemosphere 2019, 231, 393–404. [Google Scholar] [CrossRef] [PubMed]
  22. Roy, H.; Islam, M.S.; Arifin, M.T.; Firoz, S.H. Chitosan-ZnO decorated Moringa oleifera seed biochar for sequestration of methylene blue: Isotherms, kinetics, and response surface analysis. Environ. Nanotechnol. Monit. Manag. 2022, 18, 100752. [Google Scholar] [CrossRef]
  23. Roy, H.; Islam, M.S.; Arifin, M.T.; Firoz, S.H. Synthesis, Characterization and Sorption Properties of Biochar, Chitosan and ZnO-Based Binary Composites towards a Cationic Dye. Sustainability 2022, 14, 14571. [Google Scholar] [CrossRef]
  24. Roy, H.; Prantika, T.R.; Riyad, M.; Paul, S.; Islam, M.S. Synthesis, characterizations, and RSM analysis of Citrus macroptera peel derived biochar for textile dye treatment. South Afr. J. Chem. Eng. 2022, 41, 129–139. [Google Scholar] [CrossRef]
  25. Mahmoodi, N.M.; Salehi, R.; Arami, M. Binary system dye removal from colored textile wastewater using activated carbon: Kinetic and isotherm studies. Desalination 2011, 272, 187–195. [Google Scholar] [CrossRef]
  26. Amin, M.T.; Alazba, A.A.; Manzoor, U. A Review of Removal of Pollutants from Water/Wastewater Using Different Types of Nanomaterials. Adv. Mater. Sci. Eng. 2014, 2014, 825910. [Google Scholar] [CrossRef] [Green Version]
  27. Sharma, Y.C.; Uma, A.S.K.; Sinha, S.N. Upadhyay, Characterization and Adsorption Studies of Cocos nucifera L. Activated Carbon for the Removal of Methylene Blue from Aqueous Solutions. J. Chem. Eng. Data 2010, 55, 2662–2667. [Google Scholar] [CrossRef]
  28. El-Sayed, G.O. Removal of methylene blue and crystal violet from aqueous solutions by palm kernel fiber. Desalination 2011, 272, 225–232. [Google Scholar] [CrossRef]
  29. Rastogi, K.; Sahu, J.; Meikap, B.; Biswas, M. Removal of methylene blue from wastewater using fly ash as an adsorbent by hydrocyclone. J. Hazard. Mater. 2008, 158, 531–540. [Google Scholar] [CrossRef]
  30. Li, Z.; Chang, P.-H.; Jiang, W.-T.; Jean, J.-S.; Hong, H. Mechanism of methylene blue removal from water by swelling clays. Chem. Eng. J. 2011, 168, 1193–1200. [Google Scholar] [CrossRef]
  31. Almeida, C.; Debacher, N.; Downs, A.; Cottet, L.; Mello, C. Removal of methylene blue from colored effluents by adsorption on montmorillonite clay. J. Colloïd Interface Sci. 2009, 332, 46–53. [Google Scholar] [CrossRef]
  32. Aroguz, A.Z.; Gulen, J.; Evers, R. Adsorption of methylene blue from aqueous solution on pyrolyzed petrified sediment. Bioresour. Technol. 2008, 99, 1503–1508. [Google Scholar] [CrossRef]
  33. Gomaa, H.; Hussein, M.A.; Motawea, M.M.; Aboraia, A.M.; Cheira, M.F.; Alotaibi, M.T.; El-Bahy, S.M.; Ali, H.M. A hybrid mesoporous CuO@ barley straw-derived SiO2 nanocomposite for adsorption and photocatalytic degradation of methylene blue from real wastewater. Colloids Surf. A Physicochem. Eng. Asp. 2022, 644, 128811. [Google Scholar] [CrossRef]
  34. Hussein, M.A.T.; Motawea, M.M.; Elsenety, M.M.; El-Bahy, S.M.; Gomaa, H. Mesoporous spongy Ni–Co oxides@wheat straw-derived SiO2 for adsorption and photocatalytic degradation of methylene blue pollutants. Appl. Nanosci. 2022, 12, 1519–1536. [Google Scholar] [CrossRef]
  35. Kassem, K.O.; Hussein, M.A.; Motawea, M.M.; Gomaa, H.; Alrowaili, Z.; Ezzeldien, M. Design of mesoporous ZnO@ silica fume-derived SiO2 nanocomposite as photocatalyst for efficient crystal violet removal: Effective route to recycle industrial waste. J. Clean. Prod. 2021, 326, 129416. [Google Scholar] [CrossRef]
  36. Seaf El-Nasr, T.A.; Gomaa, H.; Emran, M.Y.; Motawea, M.M.; Ismail, A.-R.A. Recycling of nanosilica from agricultural, electronic, and industrial wastes for wastewater treatment. In Waste Recycling Technologies for Nanomaterials Manufacturing; Springer: Berlin/Heidelberg, Germany, 2021; pp. 325–362. [Google Scholar]
  37. Chen, H.; Zhao, J.; Dai, G. Silkworm exuviae—A new non-conventional and low-cost adsorbent for removal of methylene blue from aqueous solutions. J. Hazard. Mater. 2011, 186, 1320–1327. [Google Scholar] [CrossRef] [PubMed]
  38. Haque, E.; Jun, J.W.; Jhung, S.H. Adsorptive removal of methyl orange and methylene blue from aqueous solution with a metal-organic framework material, iron terephthalate (MOF-235). J. Hazard. Mater. 2011, 185, 507–511. [Google Scholar] [CrossRef]
  39. Kumari, S.; Rajput, V.D.; Minkina, T.; Rajput, P.; Sharma, P.; Verma, A.K.; Agarwal, S.; Garg, M.C. Application of RSM for Bioremoval of Methylene Blue Dye from Industrial Wastewater onto Sustainable Walnut Shell (Juglans regia) Biomass. Water 2022, 14, 3651. [Google Scholar] [CrossRef]
  40. Wang, J.; Ma, J.; Sun, Y. Adsorption of Methylene Blue by Coal-Based Activated Carbon in High-Salt Wastewater. Water 2022, 14, 3576. [Google Scholar] [CrossRef]
  41. Madrakian, T.; Afkhami, A.; Ahmadi, M.; Bagheri, H. Removal of some cationic dyes from aqueous solutions using magnetic-modified multi-walled carbon nanotubes. J. Hazard. Mater. 2011, 196, 109–114. [Google Scholar] [CrossRef]
  42. Liu, T.; Li, Y.; Du, Q.; Sun, J.; Jiao, Y.; Yang, G.; Wang, Z.; Xia, Y.; Zhang, W.; Wang, K. Adsorption of methylene blue from aqueous solution by graphene. Colloids Surf. B Biointerfaces 2012, 90, 197–203. [Google Scholar] [CrossRef]
  43. Wang, S.; Zhang, L.; Zhang, M.; Xu, L.; Hu, Q.; Yang, T.; Tu, K.; Wu, M.; Yu, D. Enhanced Methylene Blue Adsorption by Cu-BTC Metal-Organic Frameworks with Engineered Particle Size Using Surfactant Modulators. Water 2022, 14, 1864. [Google Scholar] [CrossRef]
  44. Jjagwe, J.; Olupot, P.W.; Menya, E.; Kalibbala, H.M. Synthesis and application of Granular activated carbon from biomass waste materials for water treatment: A review. J. Bioresour. Bioprod. 2021, 6, 292–322. [Google Scholar] [CrossRef]
  45. Ali, S.H.; Emran, M.Y.; Gomaa, H. Rice husk-derived nanomaterials for potential applications. In Waste Recycling Technologies for Nanomaterials Manufacturing; Springer: Berlin/Heidelberg, Germany, 2021; pp. 541–588. [Google Scholar]
  46. Feng, L.; Yan, B.; Zheng, J.; Chen, J.; Wei, R.; Jiang, S.; Yang, W.; Zhang, Q.; He, S. Soybean protein-derived N, O co-doped porous carbon sheets for supercapacitor applications. New J. Chem. 2022, 46, 10844–10853. [Google Scholar] [CrossRef]
  47. Guo, S.; Li, Y.; Wang, Y.; Wang, L.; Sun, Y.; Liu, L. Recent advances in biochar-based adsorbents for CO2 capture. Carbon Capture Sci. Technol. 2022, 4, 100059. [Google Scholar] [CrossRef]
  48. Lee, M.-H.; Chang, E.-H.; Lee, C.-H.; Chen, J.-Y.; Jien, S.-H. Effects of biochar on soil aggregation and distribution of organic carbon fractions in aggregates. Process 2021, 9, 1431. [Google Scholar] [CrossRef]
  49. Obey, G.; Adelaide, M.; Ramaraj, R. Biochar derived from non-customized matamba fruit shell as an adsorbent for wastewater treatment. J. Bioresour. Bioprod. 2022, 7, 109–115. [Google Scholar] [CrossRef]
  50. Roy, H.; Alam, S.R.; Bin-Masud, R.; Prantika, T.R.; Pervez, M.N.; Islam, M.S.; Naddeo, V. A Review on Characteristics, Techniques, and Waste-to-Energy Aspects of Municipal Solid Waste Management: Bangladesh Perspective. Sustainability 2022, 14, 10265. [Google Scholar] [CrossRef]
  51. Yan, B.; Feng, L.; Zheng, J.; Zhang, Q.; Jiang, S.; Zhang, C.; Ding, Y.; Han, J.; Chen, W.; He, S. High performance supercapacitors based on wood-derived thick carbon electrodes synthesized via green activation process. Inorg. Chem. Front. 2022, 9, 6108–6123. [Google Scholar] [CrossRef]
  52. Foo, K.Y.; Hameed, B. An overview of dye removal via activated carbon adsorption process. Desalination Water Treat. 2010, 19, 255–274. [Google Scholar] [CrossRef] [Green Version]
  53. Abu-Nada, A.; Abdala, A.; McKay, G. Removal of phenols and dyes from aqueous solutions using graphene and graphene composite adsorption: A review. J. Environ. Chem. Eng. 2021, 9, 105858. [Google Scholar] [CrossRef]
  54. Al-Degs, Y.S.; El-Barghouthi, M.I.; El-Sheikh, A.H.; Walker, G.M. Effect of solution pH, ionic strength, and temperature on adsorption behavior of reactive dyes on activated carbon. Dye. Pigment. 2008, 77, 16–23. [Google Scholar] [CrossRef]
  55. Zhao, F.; Zou, G.; Shan, Y.; Ding, Z.; Dai, M.; He, Z. Coconut shell derived biochar to enhance water spinach (Ipomoea aquatica Forsk) growth and decrease nitrogen loss under tropical conditions. Sci. Rep. 2019, 9, 20291. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  56. Zhang, P.; O’Connor, D.; Wang, Y.; Jiang, L.; Xia, T.; Wang, L.; Tsang, D.C.; Ok, Y.S.; Hou, D. A green biochar/iron oxide composite for methylene blue removal. J. Hazard. Mater. 2020, 384, 121286. [Google Scholar] [CrossRef] [PubMed]
  57. Mahmud, K.N.; Wen, T.H.; Zakaria, Z.A. Activated carbon and biochar from pineapple waste biomass for the removal of methylene blue. Environ. Toxicol. Manag. 2021, 1, 30–36. [Google Scholar] [CrossRef]
  58. Wang, S.; Kwak, J.-H.; Islam, M.S.; Naeth, M.A.; El-Din, M.G.; Chang, S.X. Biochar surface complexation and Ni (II), Cu (II), and Cd (II) adsorption in aqueous solutions depend on feedstock type. Sci. Total Environ. 2020, 712, 136538. [Google Scholar] [CrossRef]
  59. Amin, M.; Alazba, A.; Shafiq, M. Comparative study for adsorption of methylene blue dye on biochar derived from orange peel and banana biomass in aqueous solutions. Environ. Monit. Assess. 2019, 191, 735. [Google Scholar] [CrossRef]
  60. Naeem, M.A.; Imran, M.; Amjad, M.; Abbas, G.; Tahir, M.; Murtaza, B.; Zakir, A.; Shahid, M.; Bulgariu, L.; Ahmad, I. Batch and Column Scale Removal of Cadmium from Water Using Raw and Acid Activated Wheat Straw Biochar. Water 2019, 11, 1438. [Google Scholar] [CrossRef] [Green Version]
  61. Saleem, J.; Shahid, U.B.; Hijab, M.; Mackey, H.; McKay, G. Production and applications of activated carbons as adsorbents from olive stones. Biomass Convers. Biorefin. 2019, 9, 775–802. [Google Scholar] [CrossRef] [Green Version]
  62. Rahman, E.; Roy, H.; Ahmed, S.; Firoz, S.H. The seed of Burmese grape (BACCAUREA RAMIFLORA) as low cost bio-adsorbent for removal of methylene blue from wastewater. In Proceedings of the International Conference on Civil Engineering for Sustainable Development, Khulna, Bangladesh, 7–8 February 2020. [Google Scholar]
  63. Ullah, M.O.; Urmi, K.F.; Howlader, M.A.; Hossain, M.; Ahmed, M.T.; Hamid, K. Hypoglycemic, hypolippidemic and antioxidant effects of leaves methanolic extract of baccaurea ramiflora. Int. J. Pharm. Pharm. Sci. 2012, 4, 266–269. [Google Scholar]
  64. Ahmad, I.; Das, T.T.; Yasin, M. Afzal Hossain, Study on Biochemical Compounds, Antioxidant Activity and Organoleptic Taste of Some Spice Tea. Agric. Food Sci. Res. 2016, 3, 53–58. [Google Scholar]
  65. Suhaimi, N.; Kooh, M.R.R.; Lim, C.M.; Chou Chao, C.-T.; Chou Chau, Y.-F.; Mahadi, A.H.; Chiang, H.-P.; Haji Hassan, N.H.; Thotagamuge, R. The Use of Gigantochloa Bamboo-Derived Biochar for the Removal of Methylene Blue from Aqueous Solution. Adsorpt. Sci. Technol. 2022, 2022, 8245797. [Google Scholar] [CrossRef]
  66. Fanning, P.E.; Vannice, M.A. A DRIFTS study of the formation of surface groups on carbon by oxidation. Carbon 1993, 31, 721–730. [Google Scholar] [CrossRef]
  67. Özçimen, D.; Ersoy-Meriçboyu, A. Characterization of biochar and bio-oil samples obtained from carbonization of various biomass materials. Renew. Energy 2010, 35, 1319–1324. [Google Scholar] [CrossRef]
  68. Shaaban, A.; Se, S.-M.; Mitan, N.M.M.; Dimin, M. Characterization of biochar derived from rubber wood sawdust through slow pyrolysis on surface porosities and functional groups. Procedia Eng. 2013, 68, 365–371. [Google Scholar] [CrossRef] [Green Version]
  69. He, R.; Peng, Z.; Lyu, H.; Huang, H.; Nan, Q.; Tang, J. Synthesis and characterization of an iron-impregnated biochar for aqueous arsenic removal. Sci. Total Environ. 2018, 612, 1177–1186. [Google Scholar] [CrossRef]
  70. Prahas, D.; Kartika, Y.; Indraswati, N.; Ismadji, S. Activated carbon from jackfruit peel waste by H3PO4 chemical activation: Pore structure and surface chemistry characterization. Chem. Eng. J. 2008, 140, 32–42. [Google Scholar] [CrossRef]
  71. Elkady, M.; Shokry, H.; Hamad, H. New Activated Carbon from Mine Coal for Adsorption of Dye in Simulated Water or Multiple Heavy Metals in Real Wastewater. Materials 2020, 13, 2498. [Google Scholar] [CrossRef]
  72. Nasir, M.; Rahmawati, T.; Dara, F. Synthesis and characterization of biochar from crab shell by pyrolysis. IOP Conf. Ser. Mater. Sci. Eng. 2019, 553, 012031. [Google Scholar] [CrossRef]
  73. Yang, H.; Zhang, D.; Chen, Y.; Ran, M.; Gu, J. Study on the application of KOH to produce activated carbon to realize the utilization of distiller’s grains. IOP Conf. Ser. Earth Environ. Sci. 2017, 69, 012051. [Google Scholar] [CrossRef] [Green Version]
  74. Silva, T.L.; Cazetta, A.L.; Souza, P.S.; Zhang, T.; Asefa, T.; Almeida, V.C. Mesoporous activated carbon fibers synthesized from denim fabric waste: Efficient adsorbents for removal of textile dye from aqueous solutions. J. Clean. Prod. 2018, 171, 482–490. [Google Scholar] [CrossRef]
  75. Gonçalves, N.P.; Lourenço, M.A.; Baleuri, S.R.; Bianco, S.; Jagdale, P.; Calza, P. Biochar waste-based ZnO materials as highly efficient photocatalysts for water treatment. J. Environ. Chem. Eng. 2022, 10, 107256. [Google Scholar] [CrossRef]
  76. Ajeng, A.A.; Abdullah, R.; Junia, A.; Lau, B.F.; Ling, T.C.; Ismail, S. Evaluation of palm kernel shell biochar for the adsorption of Bacillus cereus. Phys. Scr. 2021, 96, 105004. [Google Scholar] [CrossRef]
  77. Roy, H.; Shakil, R.; Tarek, Y.A.; Firoz, S.H. Study of the Removal of Basic Blue-41 from Simulated Wastewater by Activated Carbon Prepared from Discarded Jute Fibre. ECS Trans. 2022, 107, 8407. [Google Scholar] [CrossRef]
  78. Roy, H.; Rahman, M.M.; Tarek, Y.A.; Firoz, S.H. Encapsulation of Industrial Cationic Pollutants from Aqueous Solution by Nanocrystalline Cellulose and It’s Modified Forms. In Proceedings of the International Exchange and Innovation Conference on Engineering & Sciences, Fukuoka, Japan, 21–22 October 2021; Volume 7. [Google Scholar]
  79. Aragaw, T.A.; Alene, A.N. A comparative study of acidic, basic, and reactive dyes adsorption from aqueous solution onto kaolin adsorbent: Effect of operating parameters, isotherms, kinetics, and thermodynamics. Emerg. Contam. 2022, 8, 59–74. [Google Scholar] [CrossRef]
  80. Wong, S.; Ghafar, N.A.; Ngadi, N.; Razmi, F.A.; Inuwa, I.M.; Mat, R.; Amin, N.A.S. Effective removal of anionic textile dyes using adsorbent synthesized from coffee waste. Sci. Rep. 2020, 10, 2928. [Google Scholar] [CrossRef] [Green Version]
  81. Aljeboree, A.M.; Alshirifi, A.N.; Alkaim, A.F. Kinetics and equilibrium study for the adsorption of textile dyes on coconut shell activated carbon. Arab. J. Chem. 2017, 10, S3381–S3393. [Google Scholar] [CrossRef] [Green Version]
  82. Liu, Q.-X.; Zhou, Y.-R.; Wang, M.; Zhang, Q.; Ji, T.; Chen, T.-Y.; Yu, D.-C. Adsorption of methylene blue from aqueous solution onto viscose-based activated carbon fiber felts: Kinetics and equilibrium studies. Adsorpt. Sci. Technol. 2019, 37, 312–332. [Google Scholar] [CrossRef] [Green Version]
  83. Ko, D.C.; Tsang, D.H.; Porter, J.F.; McKay, G. Applications of multipore model for the mechanism identification during the adsorption of dye on activated carbon and bagasse pith. Langmuir 2003, 19, 722–730. [Google Scholar] [CrossRef]
  84. Banerjee, S.; Chattopadhyaya, M.C. Adsorption characteristics for the removal of a toxic dye, tartrazine from aqueous solutions by a low cost agricultural by-product. Arab. J. Chem. 2017, 10, S1629–S1638. [Google Scholar] [CrossRef] [Green Version]
  85. Gomaa, H.; El-Monaem, A.; Eman, M.; Eltaweil, A.S.; Omer, A.M. Efficient removal of noxious methylene blue and crystal violet dyes at neutral conditions by reusable montmorillonite/NiFe2O4@ amine-functionalized chitosan composite. Sci. Rep. 2022, 12, 15499. [Google Scholar]
  86. Agboola, O.D.; Benson, N.U. Physisorption and Chemisorption Mechanisms Influencing Micro (Nano) Plastics-Organic Chemical Contaminants Interactions: A Review. Front. Environ. Sci. 2021, 9, 167. [Google Scholar] [CrossRef]
  87. Foo, K.Y.; Hameed, B.H. Insights into the modeling of adsorption isotherm systems. Chem. Eng. J. 2010, 156, 2–10. [Google Scholar] [CrossRef]
  88. Kaur, S.; Rani, S.; Mahajan, R.K. Adsorption Kinetics for the Removal of Hazardous Dye Congo Red by Biowaste Materials as Adsorbents. J. Chem. 2013, 2013, 628582. [Google Scholar] [CrossRef]
  89. Li, G.; Zhu, W.; Zhang, C.; Zhang, S.; Liu, L.; Zhu, L.; Zhao, W. Effect of a magnetic field on the adsorptive removal of methylene blue onto wheat straw biochar. Bioresour. Technol. 2016, 206, 16–22. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  90. Zhu, Y.; Yi, B.; Yuan, Q.; Wu, Y.; Wang, M.; Yan, S. Removal of methylene blue from aqueous solution by cattle manure-derived low temperature biochar. RSC Adv. 2018, 8, 19917–19929. [Google Scholar] [CrossRef] [Green Version]
  91. Mubarak, N.; Fo, Y.; Al-Salim, H.S.; Sahu, J.; Abdullah, E.; Nizamuddin, S.; Jayakumar, N.; Ganesan, P. Removal of methylene blue and orange-G from waste water using magnetic biochar. Int. J. Nanosci. 2015, 14, 1550009. [Google Scholar] [CrossRef]
  92. Franciski, M.A.; Peres, E.C.; Godinho, M.; Perondi, D.; Foletto, E.L.; Collazzo, G.C.; Dotto, G.L. Development of CO2 activated biochar from solid wastes of a beer industry and its application for methylene blue adsorption. Waste Manag. 2018, 78, 630–638. [Google Scholar] [CrossRef]
  93. Liu, S.; Li, J.; Xu, S.; Wang, M.; Zhang, Y.; Xue, X. A modified method for enhancing adsorption capability of banana pseudostem biochar towards methylene blue at low temperature. Bioresour. Technol. 2019, 282, 48–55. [Google Scholar] [CrossRef]
  94. Yu, K.L.; Lee, X.J.; Ong, H.C.; Chen, W.-H.; Chang, J.-S.; Lin, C.-S.; Show, P.L.; Ling, T.C. Adsorptive removal of cationic methylene blue and anionic Congo red dyes using wet-torrefied microalgal biochar: Equilibrium, kinetic and mechanism modeling. Environ. Pollut. 2021, 272, 115986. [Google Scholar] [CrossRef]
  95. Mu, Y.; Ma, H. NaOH-modified mesoporous biochar derived from tea residue for methylene Blue and Orange II removal. Chem. Eng. Res. Des. 2021, 167, 129–140. [Google Scholar] [CrossRef]
  96. Sahu, S.; Pahi, S.; Tripathy, S.; Singh, S.K.; Behera, A.; Sahu, U.K.; Patel, R.K. Adsorption of methylene blue on chemically modified lychee seed biochar: Dynamic, equilibrium, and thermodynamic study. J. Mol. Liq. 2020, 315, 113743. [Google Scholar] [CrossRef]
  97. Ji, B.; Wang, J.; Song, H.; Chen, W. Removal of methylene blue from aqueous solutions using biochar derived from a fallen leaf by slow pyrolysis: Behavior and mechanism. J. Environ. Chem. Eng. 2019, 7, 103036. [Google Scholar] [CrossRef]
  98. Abdien, H.G.; Cheira, M.; Abd-Elraheem, M.; El-Naser, T.A.S.; Zidan, I.H. Extraction and pre-concentration of uranium using activated carbon impregnated trioctyl phosphine oxide. Elixir Appl. Chem 2016, 100, 43462–43469. [Google Scholar]
  99. Zhang, H.; Wang, C.; Zhang, W.; Zhang, M.; Qi, J.; Qian, J.; Sun, X.; Yuliarto, B.; Na, J.; Park, T. Nitrogen, phosphorus co-doped eave-like hierarchical porous carbon for efficient capacitive deionization. J. Mater. Chem. A 2021, 9, 12807–12817. [Google Scholar] [CrossRef]
  100. Zhu, S.; Chen, Y.; Khan, M.A.; Xu, H.; Wang, F.; Xia, M. In-depth study of heavy metal removal by an etidronic acid-functionalized layered double hydroxide. ACS Appl. Mater. Interfaces 2022, 14, 7450–7463. [Google Scholar] [CrossRef]
  101. Zhu, S.; Khan, M.A.; Kameda, T.; Xu, H.; Wang, F.; Xia, M.; Yoshioka, T. New insights into the capture performance and mechanism of hazardous metals Cr3+ and Cd2+ onto an effective layered double hydroxide based material. J. Hazard. Mater. 2022, 426, 128062. [Google Scholar]
  102. Weber, W.J.J. Physicochemical Processes for Water Quality Control; Wiley-Interscience: New York, NY, USA, 1972; 640p. [Google Scholar]
  103. Hu, B.; Ai, Y.; Jin, J.; Hayat, T.; Alsaedi, A.; Zhuang, L.; Wang, X. Efficient elimination of organic and inorganic pollutants by biochar and biochar-based materials. Biochar 2020, 2, 47–64. [Google Scholar] [CrossRef] [Green Version]
  104. Al-Ghouti, M.A.; Khraisheh, M.A.; Ahmad, M.N.; Allen, S. Adsorption behaviour of methylene blue onto Jordanian diatomite: A kinetic study. J. Hazard. Mater. 2009, 165, 589–598. [Google Scholar] [CrossRef]
  105. Luo, Z.; Yao, B.; Yang, X.; Wang, L.; Xu, Z.; Yan, X.; Tian, L.; Zhou, H.; Zhou, Y. Novel insights into the adsorption of organic contaminants by biochar: A review. Chemosphere 2022, 287, 132113. [Google Scholar] [CrossRef]
  106. Samsuri, A.W.; Sadegh-Zadeh, F.; Seh-Bardan, B.J. Characterization of biochars produced from oil palm and rice husks and their adsorption capacities for heavy metals. Int. J. Environ. Sci. Technol. 2014, 11, 967–976. [Google Scholar] [CrossRef]
  107. Chen, Q.; Zhang, Q.; Yang, Y.; Wang, Q.; He, Y.; Dong, N. Synergetic effect on methylene blue adsorption to biochar with gentian violet in dyeing and printing wastewater under competitive adsorption mechanism. Case Stud. Therm. Eng. 2021, 26, 101099. [Google Scholar] [CrossRef]
  108. Wang, Y.; Zhang, Y.; Li, S.; Zhong, W.; Wei, W. Enhanced methylene blue adsorption onto activated reed-derived biochar by tannic acid. J. Mol. Liq. 2018, 268, 658–666. [Google Scholar] [CrossRef]
  109. 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]
  110. Jawad, A.H.; Ismail, K.; Ishak, M.A.M.; Wilson, L.D. Conversion of Malaysian low-rank coal to mesoporous activated carbon: Structure characterization and adsorption properties. Chin. J. Chem. Eng. 2019, 27, 1716–1727. [Google Scholar] [CrossRef]
  111. Jawad, A.H.; Abdulhameed, A.S. Statistical modeling of methylene blue dye adsorption by high surface area mesoporous activated carbon from bamboo chip using KOH-assisted thermal activation. Energy Ecol. Environ. 2020, 5, 456–469. [Google Scholar] [CrossRef]
Figure 1. Schematic procedure for BGS-derived biochar preparation.
Figure 1. Schematic procedure for BGS-derived biochar preparation.
Water 15 00394 g001
Figure 2. FTIR analysis (A,B) SEM image of Burmese grape-seed-derived biochar.
Figure 2. FTIR analysis (A,B) SEM image of Burmese grape-seed-derived biochar.
Water 15 00394 g002
Figure 3. The influences of (A) pH and (B) time on removal (%) of MB by BGS-derived biochar.
Figure 3. The influences of (A) pH and (B) time on removal (%) of MB by BGS-derived biochar.
Water 15 00394 g003
Figure 4. Effect of initial concentration (Co) on equilibrium adsorption capacity qe (A) and equilibrium adsorption capacity vs. equilibrium concentration (Ce) (B).
Figure 4. Effect of initial concentration (Co) on equilibrium adsorption capacity qe (A) and equilibrium adsorption capacity vs. equilibrium concentration (Ce) (B).
Water 15 00394 g004
Figure 5. Pseudo-first-order (A) and pseudo-second-order (B) linear data fit for MB removal BGS-derived biochar.
Figure 5. Pseudo-first-order (A) and pseudo-second-order (B) linear data fit for MB removal BGS-derived biochar.
Water 15 00394 g005
Figure 6. Langmuir (A) and Freundlich (B) linear data fit for MB removal on BGS-derived biochar.
Figure 6. Langmuir (A) and Freundlich (B) linear data fit for MB removal on BGS-derived biochar.
Water 15 00394 g006
Figure 7. The adsorption mechanism of MB dyes onto BGS-derived biochar.
Figure 7. The adsorption mechanism of MB dyes onto BGS-derived biochar.
Water 15 00394 g007
Table 1. Elemental composition and BET analysis of the BGS biochar.
Table 1. Elemental composition and BET analysis of the BGS biochar.
Elemental Composition
SampleC (%)N (%)O (%)K (%)P (%)
BGS-biochar70.0217.0512.080.350.50
BET Analysis
SampleSurface Area(m2/g)Pore Volume(cm3/g)
BGS-biochar19.90 ± 1.205.85
Table 2. Methylene blue (MB) sorption results of biochar derived from different sources.
Table 2. Methylene blue (MB) sorption results of biochar derived from different sources.
Sources of BiocharConditions of Biochar PreparationExperimental ConditionsAdsorption Capacity (mg/g)References
Wheat strawPyrolyzed over 550 °C in N2 atmosphere for 3 h and activated with HClpH = 11, T = 25 °C, Dye = 80 mg/L, Contact time = 150 min.61.6[89]
Cattle manure biocharPyrolyzed over 200 °C in N2 atmosphere and activated with HClpH = 10, T = 25 °C, Dye = 200 mg/L, Contact time = 1.25 h241.99[90]
Palm oil fruit bunchMicrowaved at 800 W, reaction time of 30 min and a ratio of impregnation of 0.5 with FeCl3pH = 10, T = 25 °C, Dye = 30 mg/L, Contact time = 120 min31.25[91]
Solid wastes (beer industry)Prepared through N2 pyrolysis followed by CO2 activation at 800 °C for 60 minpH = 6, T = 25 °C, Dye = 50 mg/L, Contact time = 2 h161[92]
Banana pseudo-stemPyrolyzed with a rate of 5 °C/min with phosphomolybdic acidpH = 7, T = 25 °C, Dye = 50 mg/L, Contact time = 72 h146.23[93]
Wet-torrefied microalgaePyrolyzed in a N2 -purged inert condition using a modified household microwave at 800 W irradiationpH = 8, T = 30 °C, Dye = 210 mg/L, Contact time = 7 h113.00[94]
Tea residuePyrolyzed in a muffle furnace at 700 °C with heating rate at 10 °C/min for 4 h and NaOH activationpH = 7, T = 25 °C, Dye = 50 mg/L, Contact time = 60 min.105.27[95]
Moringa oleifera biocharPyrolyzed in a N2 furnace at 300 °C at 5 °C/min for 3 hpH = 8, T = 25 °C, Dye = 60 mg/L, Contact time = 120 min67[22]
Lychee seedsPyrolyzed in a muffle furnace at 700 °C, KOH impregnation and microwave irradiation at 600 W of 10 minpH = 6, T = 25 °C, Dye = 50 mg/L, Contact time = 120 min124.5[96]
Magnolia grandiflora leafPyrolyzed in a muffle furnace at 500 °C with heating rate at 3 °C/min for 3 hpH = 12, T = 25 °C, Dye = 50 mg/L, Contact time = 180 min.101.27[97]
Burmese grape seedsPyrolyzed at 500 °C, and preactivated with H3PO4 and KOHpH = 8, T = 27 °C, Dye = 60 mg/L, Contact time = 70 min166.3This study
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

Roy, H.; Sarkar, D.; Pervez, M.N.; Paul, S.; Cai, Y.; Naddeo, V.; Firoz, S.H.; Islam, M.S. Synthesis, Characterization and Performance Evaluation of Burmese Grape (Baccaurea ramiflora) Seed Biochar for Sustainable Wastewater Treatment. Water 2023, 15, 394. https://doi.org/10.3390/w15030394

AMA Style

Roy H, Sarkar D, Pervez MN, Paul S, Cai Y, Naddeo V, Firoz SH, Islam MS. Synthesis, Characterization and Performance Evaluation of Burmese Grape (Baccaurea ramiflora) Seed Biochar for Sustainable Wastewater Treatment. Water. 2023; 15(3):394. https://doi.org/10.3390/w15030394

Chicago/Turabian Style

Roy, Hridoy, Dipayan Sarkar, Md. Nahid Pervez, Shuvo Paul, Yingjie Cai, Vincenzo Naddeo, Shakhawat H. Firoz, and Md. Shahinoor Islam. 2023. "Synthesis, Characterization and Performance Evaluation of Burmese Grape (Baccaurea ramiflora) Seed Biochar for Sustainable Wastewater Treatment" Water 15, no. 3: 394. https://doi.org/10.3390/w15030394

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