Next Article in Journal
Developing a New Bursting Liability Index Based on Energy Evolution for Coal under Different Loading Rates
Previous Article in Journal
Stakeholder Analysis and Their Attitude towards PPP Success
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

How the Carbonization Time of Sugarcane Biomass Affects the Microstructure of Biochar and the Adsorption Process?

by
Gabriel Cabral da Fonseca
,
Marilene Silva Oliveira
,
Carlos Vinicius Costa Martins
and
João Carlos Perbone de Souza
*
Goiano Federal Institute of Education, Science and Technology, Campus Rio Verde, Rio Verde 75901-970, Brazil
*
Author to whom correspondence should be addressed.
Sustainability 2022, 14(3), 1571; https://doi.org/10.3390/su14031571
Submission received: 17 December 2021 / Revised: 18 January 2022 / Accepted: 25 January 2022 / Published: 28 January 2022
(This article belongs to the Section Sustainable Materials)

Abstract

:
Biochars (BCs) are very versatile adsorbents, mainly, in the effectiveness of adsorption of organic and inorganic compounds in aqueous solutions. Here, the sugarcane biomass (SCB) was used to produce biochar at different carbonization times: 1, 2, 3, 4, and 5 h, denominated as BC1, BC2, BC3, BC4, and BC5, respectively. The superficial reactivity was studied with adsorption equilibrium experiments and kinetics models; Methylene Blue (MB) was used as adsorbate at different pH values, concentrations, and temperatures. In summary, the carbonization time provides the increase of superficial area, with exception of BC4, which decreased. Equilibrium studies showed inflection points and fluctuations with different initial dye concentration and temperature; SCB showed the best adsorption capacity compared to the BCs at the three temperatures tested, varying with the increase of MB concentration, suggesting the dependence of these two main factors on the adsorption process. The proposed adsorption mechanism suggests the major influence of Coulomb interactions, H-bonding, and π-interactions on the adsorption of MB onto adsorbents, evidencing that the adsorption is led by physical adsorption. Therefore, the results led to the use of the SCB without carbonization at 200 °C, saving energy and more adsorbent mass, considering that the carbonization influences weight loss. This study has provided insights of the use of SCB in MB dye adsorption as a low-cost and eco-friendly adsorbent.

1. Introduction

The presence of dyes in water bodies poses a major threat to human health and aquatic organisms. Methylene Blue (MB) is an organic dye that has been widely used in printing industries for coloring paper, clothes, and plastics [1,2,3]; however, its improper disposal in the environment can cause a series of problems for the environment and human health due to its potential mutagenic and carcinogenic effect and great ability to color ecosystems [4,5].
Adsorption is a technique widely used in water treatment processes and has been broadly studied in the application for removing organic dyes in water [6], for being an easy and economically viable method [1,7]. Among the challenges to be faced, there is the choice of a renewable source for the production of adsorbent materials, since these can be produced from different sources.
The production of materials and composites from sustainable sources has attracted attention due to the range of application possibilities [8]. Lignocellulosic biomass is a residue of agro-industrial activities that has aroused great interest in the development of new adsorbent materials, precisely because of its large quantity and availability. Lignocellulosic waste, such as rice husk [9], palm fiber [10], and sugarcane biomass (SCB) [11], has been used for the production of adsorbent materials as well as activated carbon [12,13] and biochar (BC) [14,15,16]. BC is a product of the thermochemical transformation of lignocellulosic biomass, heated in a closed system with limited oxygen [17], and have been employed in the removal of organic compounds from wastewater, such as dye [18], medicines [19], and pesticides [9].
The combustion of biomass for BC production, a process denominated carbonization or pyrolysis, generates a material rich in carbon, considering SCB as a source of waste. SCB is constituted mostly of cellulose (1,4-β-poly-anhydroglucose), which is a polymer of D-glucose monomers; hemicellulose, composed of polymers of pentoses, such as galactans and xylans; and lignin, in which the structure varies depending on the source, generally constituted of coniferyl alcohol and others polymers [20]. The carbonization process provides the modification of the functional groups of lignocellulosic material, forming hydroxyl groups (O-H), aromatic chains (C=C), ethers (C-O), and others, on the surface (Figure S1). These functional groups have been widely studied, due to their presence on the surface being considered as an important factor in the adsorption process.
Moreover, there are several methodologies for BC production, such as hydrothermal carbonization [21], gasification [22], and pyrolysis [10]. The pyrolysis process involves high temperatures for carbonization (>300 °C), under a flow of nitrogen gas [23]. SCB was turned in BC under a nitrogen flow with temperatures up to 300 °C [11], and the cool down to room temperature was under a nitrogen flow too; the capacity for the cooper sorption was 40.5 mg g−1 for the 350 °C pyrolytic temperature of BC. A pyrolyzed rice husk was used for malachite green dye removal, the material was prepared at temperatures above 400 °C in a nitrogen atmosphere, and the authors found maximum adsorption of 37.47 mg g−1 for adsorbate concentration equal to 20 mg L−1 [24]. Biswas et al. 2020 performed removal tests of MB dye onto SCB; the biomass was pyrolyzed at 500 °C under a nitrogen flow. The results showed maximum adsorption of 43.97 mg g−1 for adsorbate concentration equal to 20 mg L−1. However, for this process high temperatures have been used under nitrogen flow (for an inert atmosphere), inducing a very expensive process, increasing the costs of the final product.
Thus, it is important to search for new alternatives for the production of low-cost BC, i.e., using low temperatures in the carbonization process without an inert atmosphere. There is a lack of information and research that concerns the influence of time in the carbonization of SCB for adsorption. Furthermore, research with biomass in natura showed very good adsorption capacities for dyes in several studies [25,26,27,28]; this brings the possibility of using the biomass itself as an adsorbent. Herein, BCs from SCB were prepared at different carbonization times (1–5 h) at 200 °C and characterized by microscopy and spectroscopy techniques, that provide insights into the structure. The novelty is related to evaluating the influence of carbonization of the SCB and identifying if it is an appropriate approach to improve the adsorption performance or the changes in the structure do not improve the adsorption. We performed batch adsorption with SCB and BCs adsorbents at various conditions, such as different initial dye concentrations, contact time, adsorbent dosage, MB solution pH, and temperature. Furthermore, we propose the mechanism of adsorption of MB onto SCB and BCs produced.

2. Materials and Methods

2.1. Synthesis and Characterization of the Adsorbents

The precursor SCB waste was collected in a sugar-alcohol industry, then dried in an oven with air circulation for 24 h at 100 °C and milled in a Willye micro mill. To analyze the thermal stability of SCB, a thermogravimetric analysis (TGA/DTG) was performed in Perkin Elmer Pyris 1 equipment. The initial mass was 3.167 mg, heated in the range from 65 to 750 °C (10 °C min−1), under nitrogen flow (20 mL min−1). The result is presented in the SI.
To produce the BCs, the SCB was carbonized in a muffle furnace (200 °C) at different times, as follows: 1, 2, 3, 4, and 5 h, denominated as BC1, BC2, BC3, BC4, and BC5, respectively. Subsequently, they were sieved in a granulometric mesh (0.15 mm).
The characterization setups of the SCB and BCs structures were: the functional groups of the surfaces of adsorbents were studied using the Fourier-transform infrared spectroscopy (FTIR/NIR, Perkin Elmer Frontier) with attenuated total reflectance (ATR) mode, covering the spectral range from 650 to 4000 cm−1. Points of zero charge (pHPZC) were determined as follows: 10 mg of each adsorbent was stirred for 24 h, at room temperature, with 30 mL of distilled water with initial pH (pHi) between 2 and 12; then, the final pH (pHf) value was measured. We plotted a graph of pHi versus pHf. The pHPZC is defined as the point where the curve intersects the line pHi = pHf. Moreover, the morphological analysis was performed with scanning electron microscopy (SEM, JEOL IT300LV). The surface area was measured by the N2 physisorption technique (Quantachrome Instruments, NOVA 1000e model) and Brunauer–Emmett–Teller (BET) method, with a pre-treatment under vacuum at 240 °C for 2 h.

2.2. Adsorption Studies of MB Removal onto SCB and BCs

The six types of adsorbents (SCB and BCs) were used in the batch adsorption experiments to investigate the removal of MB at different conditions. The equipment setup was by using a Solab Shaker for the experiments at room temperature and an incubator for the experiments at 35 and 45 °C (150 rpm for all studies). The contact time was studied ranging the agitation time from 0 to 1440 min. Adsorbent dosage was performed only for BC5, the mass ranged from 5 to 30 mg at pH 7, and temperature of 25 °C. The influence of positive and negative charges in the adsorption process was evaluated by adjusting the pH of MB solution (20 mg L−1) to 2, 4, 6, 8, and 10 with NaOH (0.1 M) and HCl (0.1 M). To understand the adsorption equilibrium, the initial concentration of MB ranged from 10 to 270 mg L−1 with a volume equal to 10 mL and an adsorbent mass of 5 mg, at 25, 35, and 45 °C; this experiment was conducted in duplicate to reduce experimental error.
The final concentration of MB was determined using an MB standard curve (0.1–10 mg L−1) developed in a UV-Vis spectrophotometer at 664 nm (R2 = 0.999). The amounts of MB adsorbed onto the adsorbents and removal efficiency were calculated according to the following equations:
q e = ( C 0   C e ) m × V
r e m o v a l   e f f i c i e n y = ( C 0   C e ) C 0 × 100 %
where q e (mg g−1) is the maximum amount of MB adsorbed by mass of adsorbent, C 0 and C e (mg L−1) is the initial and equilibrium MB concentration, respectively; m (g) is the adsorbent mass, and V (L) is the MB volume.

3. Results and Discussion

3.1. FTIR Spectroscopy

FTIR analysis was carried out to evaluate the chemical surface of the SCB and BCs produced at different carbonization times (Figure 1). Likewise, Table S1 shows the functional groups identified with the principal bands. SCB has broadband between 3500 and 3100 cm−1, in a range attributed to hydroxyl groups present of water adsorbed and linked to cellulose, hemicellulose, and lignin. To the BC1 sample, we can observe the increase of this band intensity, due to the liberation of water adsorbed, a result also observed in the DTG curve up to 150 °C; however, this band decreases when the carbonization time increases, and this is associated with dehydration. Bands related to the CO and CO2 groups are observed at 1350 cm−1, characteristic from products of slow cellulose pyrolysis, these molecules are adsorbed on the BC surface. Vibrational stretching bands between 1740 and 1026 cm−1 are all associated with cellulose, hemicellulose, and lignin, the decrease in the intensity of these bands means an evident precursor degradation, generating several functional groups on the surface; however, the peak at approximately 1026 cm−1 almost disappears, suggesting the almost complete decomposition of cellulose and hemicellulose [29].
The generation of functional groups on the surfaces of the adsorbents can directly affect the adsorption capacity, and the interactions that occur on the surface can be of a chemical and/or physical nature. Some of the potential adsorption mechanisms can be described by surface adsorption by covalent bonds, Coulomb interactions, hydrogen bonds, and π-interactions [30]. Carbonized adsorbents materials at 200 °C, after 1 h presents functionality characteristic of adsorbent, considering the formation of amorphous carbon, and the presence of superficial functional groups can contribute to remove organic molecules more efficiently. Moreover, the morphological study can identify the influence of residence time on pore formation.

3.2. Topography Study and Surface Area of SCB and BCs

The SEM images of the adsorbents were obtained to understand how the residence time affects the development of the pore on the surface of the adsorbents (Figure 2). To SCB, in the micrograph images, no pores are observed, just a rigid structure. The BC1 and BC2 micrographs did not show an increase in surface pores, even with the increased carbonization time.
The increase in pore quantity can be observed in the adsorbents BC3, BC4, and BC5, also showing roughness aspects, corroborating with the FTIR spectra, in which the increased carbonization time favors the degradation of the cellulose, hemicellulose, and lignin structures. Likewise, the degradation of the polymer increases the formation of internal pores, providing even more adsorption sites when added to the superficial ones, favoring the removal of the contaminants in the solution. However, the structures have the same fiber appearance, related to lignin, which is highly resistant to thermal oxidation [31]. The formation and enlargement of internal and superficial pores can be important in the adsorption process; to better understand the changes in the surfaces of the adsorbents, we studied the influence of the carbonization time in the surface area.
The results of the surface area are presented in Table S2. We observed that the surface area increases up to three hours of carbonization to 66.05 m2 g−1, then decreases with four hours to 37.98 m2 g−1, and then increases again with five hours to 90.64 m2 g−1. The decrease in the area of BC1 to the precursor SCB may be due to the narrowing of the pores by the formation of oxygen groups, in the pore entrances and walls, with one hour of carbonization [18]. Then, we see that the surface area increases in BC2 and BC3, which may be associated with an increase in the carbonization time, which degrades the structures of the cellulose, hemicellulose, and lignin polymers, forming internal and superficial pores.
BC4 reduces its surface area by almost half, from 66.05 to 37.98 m2 g−1; we can associate this feature with the conversion of micro and mesoporous to macropores due to the conditions of thermal oxidation, i.e., four hours of carbonization [18]. Though, with the increase in carbonization time to five hours, the surface area increases to 90.64 m2 g−1, suggesting the formation of more internal and superficial micropores in BC5 [32]. We can conclude that for the formation of a larger surface area, the ideal carbonization time is five hours; moreover, adsorption studies have been carried out to better understand the relationship between surface area and adsorption capacity, as well as the influence of functional groups on the adsorption process.

3.3. Adsorption Study

3.3.1. Effect of BC5 Dosages

The adsorbent dosage is an important parameter to study since it directly implies the cost-benefit of the material when applied in water treatment [33]. To investigate this parameter, we used BC5 as a representative sample; the mass of the adsorbent varied between 5 and 30 mg, with 25 mL of MB (25 mg L−1) (Figure S2). With 5 mg of adsorbent, we noticed a removal efficiency of 83.8% with an adsorption capacity of 104.8 mg g−1. However, when we increased the adsorbent mass to 10 mg, the removal efficiency and adsorption capacities were 84.2% and 52.6 mg g−1, respectively. This behavior can be explained by the increase in active sites for adsorption with the increase in mass, where the adsorption capacity decreases by increasing the mass ratio of adsorbate/mass of adsorbent (mg g−1) [21].
The increase in mass to 30 mg, the removal efficiency increased to 97.1% and the adsorption capacity decreased to 20.2 mg g−1. This result may be due to the adsorbent–adsorbate interactions that can occur with high concentrations of adsorbent, which reduces the area available for adsorption and, consequently, the adsorption capacity [34]. With the increase in mass from 5 to 15 mg, the removal efficiency increases only 1.3%, while the adsorption capacity decreases from 104.8 to 35.5 mg g−1; in this case, the ideal dosage of the mass to be used, aiming at a better cost–benefit ratio, is 5 mg (0.2 g L−1).

3.3.2. pHPZC and Effect of the Solution pH on the MB Adsorption

The pHPZC is determinate as the pH value where the surface charge is zero; for the adsorbents in this study, the pHPZC varied between 6.1 and 7.7 (Figure 3a). These results show that the variation in the carbonization time may be influencing the presence and/or amount of functional groups on the surface of the adsorbents, as we noticed in the FTIR spectra, because when pHPZC < 7 represents the predominance of acid groups and above 7 the presence of basic functional groups [35]. When the pH value of the solution is below pHPZC, the adsorbent surface is positively charged, being more efficient in attracting anions [36]; therefore, there will be repulsion between the adsorbent surface and the MB molecules, since MB is a cationic dye.
The results of the removal efficiency as a function of pH value show that at pH 2 and 4, the removal is lower than at pH 6, 8, and 10 (Figure 3b). For example, SCB had a removal efficiency of 37.1 and 89.0% at pH 2 and 10, respectively. This is due to the presence of H+ ions in the solution, which competes for the active sites with the MB molecules, contributing to the relatively low removal efficiency [36]. However, for BCs the removal efficiency did not change significantly with the change in pH values, demonstrating that electrostatic interactions are not the main factor in the removal of MB. It is important to consider other parameters when explaining the adsorption mechanism of MB, including the bonds that take place in the adsorption process, such as π-bonding and Coulomb interactions.
Additionally, MB solutions (20 mg L−1) with pH 8 and 10 decreased after the adsorption process, keeping the pH value between 6 and 7.3 (Table S3). This decrease probably occurs due to the interactions between H+ ions and the hydrolyzed form of MB (MB+) are weak; therefore, most of the H+ adsorbed on the surfaces of the adsorbents are released in solution [37]. These results suggest the predominance of Coulomb interactions between MB molecules and functional groups with negative partial charges, after dissociation of H+ in solution.
The removal efficiency for adsorbents, on average, was higher for solutions with pH 8 (63.7%); this result can be understood by the protonation and deprotonation of the MB in solution, in the MB and MB+ forms [27]. El-Ahmady and Rabei reported that at pH 8, the seaweed biomass showed greater removal of MB, indicating that the surface of the adsorbent used also has protonation and deprotonation behavior, depending on the pH of the solution [27].
The removal efficiency of BCs increased by a rate of 7.8% with a change in pH from 2 to 10. This low increase can be associated with the influence of other interactions between BCs and MB molecules, such as the π-π interactions, suggesting that the dominant interaction in the adsorption mechanism [38]. As seen previously, FTIR spectra show the presence of -OH, -COOH, and C=C functional groups on the surfaces of the adsorbents. The -COOH group has pKa values between 2 and 4; therefore, its dissociation contributes to the MB removal by electrostatic interactions at acid pH; as well as the dissociation of protons from the phenolic groups -OH (pKa between 7–10) contributes to the removal at basic pH [39].

3.3.3. Effect of Contact Time and Kinetic Models

To evaluate the effect of the contact time on the adsorption capacity, we calculated the MB concentration at different contact times, varying between 0 and 1440 min. The removal efficiency is shown in Figure 4a. With only 60 min of contact, the SCB and BC4 had a removal capacity of 90%, while the BC1 with 1440 min of contact removed only 60% of the MB in solution. The decrease in the efficiency of SCB removal for BC1 may be due to the low availability of active sites for adsorption caused by the decrease in surface area from 6.54 to 2.29 m2 g−1 after one hour of carbonization, which implies a low availability of adsorption sites.
At pH between 8 and 10, we saw that adsorption is favored by the decrease of H+ on the surface of the adsorbents, suggesting the predominance of Coulomb interactions between MB molecules and functional groups with negative partial charges, after dissociation of H+ in solution. Hydrogen bonds occur when hydrogen is donated to a hydrogen acceptor, such as nitrogen atoms, for example. The MB molecule contains three nitrogen atoms that can donate electrons to form a bond with functional groups on the surface of the adsorbents. This suggests that these interactions occur between the groups -OH and -COOH, which are present as seen in the FTIR spectra (Figure 1).
BC2, BC3, BC4, and BC5 adsorbents had a removal rate > 96% within 360 min and gradually decreased, resulting from the diffusion of MB molecules on the surface of the particles [40], reaching equilibrium with 720 min of contact. This behavior, with high removal efficiency in a short time, maybe associated with several factors, such as (i) high availability of sites available for MB adsorption in the beginning [41], decreasing as it reaches equilibrium; and (ii) a large number of functional groups on the surface of the adsorbents and MB molecules in the initial stage, leading to greater adsorption due to the ability to conduct the mass transfer [42].
For different BCs, the increase in removal efficiency is seen with the increase in carbonization time. After 1440 min of contact, BC2 reached 97.6% removal and increased to 99.1 and 99.8% for BC3 and BC5, respectively. This increase may be due to the enlargement of the surface area (Table S2), a result of the increase in carbonization time. Wang et al. (2018) also observed an increase in the adsorption capacity of MB with an increase in the surface area of reed-derived biochar [43]. The authors investigated the effect of tannic acid on activated BC to remove MB; the surface area increased from 26.0 to 37.5 m2 g−1 after treatment, increasing the adsorption capacity from 27.2 to 37.2 mg g−1. Based on the results above, the contact time of 360 min was chosen for the studies of adsorption isotherms, considering the best cost–benefit and economy of energy demand.
The kinetic models used were Pseudo-First-Order (PFO), Pseudo-Second-Order (PSO), and the Elovich model to analyze the adsorption capacity over time. The equations of the models used and their respective kinetic parameters are presented in Table 1, and Figure 4a–c present the model fits. According to the results of the correlation factor (R2), experimental data were well described by all the models used, and the PSO model was better compared to the PFO model, this can be seen by the values of R2 and q e qe that were more similar to the experimental values. The Elovich and PSO model were the two models that better describe the adsorption kinetics for all adsorbents, with R2 ≥ 0.98, with exception of BC1 in the PSO, which presented an R2 of 0.91; this result brings the possibility of the BC1 in following a different adsorption mechanism.
The Elovich model is generally applied to chemosorption data [47] and describes heterogeneous diffusion processes that are regulated by two parameters: diffusion factor and reaction rate [48]. Thus, the good correlation of the data with the Elovich model suggests that the process of adsorption of MB in adsorbents occurs by heterogeneous diffusion. Additionally, the PSO model reveals that adsorption is a rate-limiting step; this process can be associated with chemical interactions between MB molecules and adsorbents, with changes in valence forces and sharing or exchange of electrons [49]. Therefore, MB molecules may be adsorbed on the surface through chemical bonds, which usually take a little longer to reach equilibrium [49].

3.3.4. Adsorption Mechanism

The adsorption of MB onto SBC and BCs was a dynamic process and presented different variations depending on several factors, such as pH of the solution, initial concentration of MB, and temperature, which we can consider to propose an adsorption mechanism (Figure 5). The kinetic results suggest that the adsorption process takes place by chemical bonds; however, we can deduce that covalent bonds may not be one of the main interactions that occur between MB and adsorbent molecules, due to the possibility of weak interaction forces, as discussed previously. Thus, the main interactions that occur in the studied adsorption process are Coulomb interactions, hydrogen bonds, and π-interactions, which are considered for physical adsorption processes [30].
The adsorption process, although generally endothermic, can be exothermic; when the process is endothermic, there is an increase in adsorption raising the temperature, as in the results discussed in the previous section. Moreover, when the process is exothermic, increasing temperature will decrease the adsorption capacity [25].
Figure 6 shows the effect of temperature on MB adsorption in SCB and BCs at different concentrations. In general, the increase in temperature had a positive influence on the increase in the adsorption of the dye in the adsorbents, such as BC5, which at a concentration of 230 mg L−1, at 25 °C had an adsorption capacity of only 3.6 mg g−1, increasing to 69. 2 mg g−1 at 45 °C. A significant increase was also observed for BC4 with MB concentration of 110 mg L−1, increasing from 4.2 to 45.6 mg g−1 at 25 and 45 °C, respectively. This may be because with increasing temperature there is a decrease in the viscosity of the solution, thus increasing the mobility of MB molecules [50].
We observed that in high concentrations of MB, the SCB decreases its adsorption capacity, with an increase in temperature from 25 to 45 °C (Figure 6), indicating that at high temperatures the bonding forces become weaker for this adsorbent. With an initial concentration of 230 mg L−1 at 35 °C the adsorption capacity was 67.14 mg g−1, and decreased to 43.1 mg g−1 at 45 °C. This result may be associated with a decrease in interactions between the active sites and the MB molecules, as discussed in the study by Biswas et al. (2020) [34]. The authors investigated the adsorption of MB in composites prepared from SCB biochar and observed that with an increase in temperature from 30 to 60 °C, the removal increased from 86 to 88.93%, respectively.
Notwithstanding the decrease, SCB was still a better adsorbent than BCs under the same concentration conditions at 25 and 35 °C. Figure 6 also reveals that the increase in temperature had a greater influence on the adsorption capacity of BC5, which proved to be the best adsorbent among BCs; however, the average adsorption capacity for BC5 at 25, 35, and 45 °C were 13.9, 32.9, and 44.6 mg g−1, respectively, while for SCB it was 46.2, 49.4, and 46.9 mg g−1. This demonstrates that despite the carbonization time of five hours, increasing the surface area by 14× to SCB is still proved to be a better adsorbent for MB dye under the studied conditions. Besides, the similar SCB adsorption capacity between 25 and 45 °C increases its applicability and is not dependent on the temperature of the effluent contaminated with MB dye.
An interesting behavior observed in the SCB adsorption capacity in some concentrations (30, 60, 230, and 270 mg L−1) is that it increases from 25 to 35 °C, and then decreases to 45 °C. This may be related to factors, such as (i) the existence of weak bonding forces between the SCB and MB molecules, which, at high temperatures (45 °C), are weakened and break, decreasing the adsorption capacity [51]; and (ii) the increase in the kinetic energy of MB molecules with increasing temperature [52]. These results suggest the dependence on factors, such as the adsorbate concentration and temperature, as being the main influencers in the adsorption process, defining the endothermic or exothermic nature of the process.
To investigate the influence of the initial MB concentration, a stock solution of 1000 mg L−1 was prepared and then diluted in different concentrations (10, 30, 60, 110, 150, 190, 230, and 270 mg L−1). The values calculated for q e , which is the maximum degree of surface coverage, were calculated and plotted against the initial concentration in the dye; the value here is presented as the maximum adsorption capacity. There are different inflection points with increasing concentration causing fluctuations in the adsorption capacity of MB by the adsorbents; this behavior is observed for all adsorbents at different temperatures (Figure S4a). The Figure also shows that the maximum adsorption capacity was observed in different concentrations studied for the adsorbents, being 230, 230, 190, 60, 270, and 60 mg L−1 for SCB, BC1, BC2, BC3, BC4, and BC5, respectively. Two tests were conducted to reject the alternative of the inflection point be due to agglomeration of the MB molecules or by interference in the filtration; they are in the SI.
The SCB adsorption capacity increases between 10 and 60 mg L−1, decays to 150 mg L−1, and increases again up to 230 mg L−1, followed by a decrease to 270 mg L−1 at 25 °C. BC1 increases remain almost constant between 10 and 110 mg L−1, decreases up to 150 mg L−1, increases up to 230 mg L−1, and decreases again by 270 mg L−1 at 25 °C. These results are similar to those of Al-Ghouti and Al-Absi (2020) who obtained fluctuations in removal efficiencies (in percentage) with increasing concentration. The authors studied the removal of MB by green and black olive stones and attributed the observed behavior to several factors as discussed by Albroomi et al. (2017) [53]. The major factors are (i) with low initial dye concentrations, the availability of adsorption sites is great; nevertheless, the amount adsorbed becomes low, as well as the mass transfer; (ii) with the increase of the initial concentration the mass transfer tends to increase, implying a high in the adsorption on active sites; and (iii) further increasing the concentration, the proportion of MB molecules and adsorption sites reaches levels where the mass transfer is not supported.
The fluctuations and inflection points are less intense for BC1 (between 10 and 150 mg L−1) and BC5 (between 110 and 230 mg L−1) at 25 °C (Figure S4a). These differences may be associated with the heterogeneity of the adsorption process, with different interactions between the MB molecules and the adsorbents, due to the functional groups present on the surfaces [25]. The FTIR (Figure 1) spectra present many differences of functional groups in the adsorbent’s structures, which can be a factor that influences the adsorption at high MB concentrations, as we have discussed previously. For example, for BC1, the adsorption capacity was below 20 mg g−1 to 150 mg L−1 at 25 °C, which was expected since this adsorbent has a lower surface area; however, there was an increase to 42.0 and 34.7 mg g−1 at 230 and 270 mg L−1, respectively. In these two concentrations, BC1 outperformed the other biochar. This result may be associated with the increase in interactions between the functional groups on the surface of the adsorbent in high concentrations of dye; the reason for being greater than the other BCs may be due to the greater amount of groups -COOH and -OH, which retain the MB molecules through hydrogen bonds and electrostatic interactions [40]. BC4 showed a similar result; between 110 and 270 mg L−1, the adsorption capacity increased from 9.3 to 28.8 mg g−1, which can be associated with a greater driving force, by increasing the concentration, which overcomes resistance in the process of mass transfer of the dye from the solution to the adsorbent [36].
Furthermore, another interaction that may be influencing the adsorption process is the π-interactions. Figure S1 shows the presence of aromatic groups that can be formed in the carbonization process, and in the FTIR spectra, we can see the presence of C=C stretch that is associated with the presence of aromatic groups (Table S1). The C=C double bond is π-systems attractive to polar molecules and other π-systems; in addition, π-systems are electronegative nature, implying greater attraction of cations [30]. Thus, we can imply that these interactions occurred predominantly in this adsorption study.

4. Conclusions

In this study, BCs were produced at different carbonization times at 200 °C to compare their adsorption capacities with SCB of the MB dye. Although there was an increase in superficial area with an increase in carbonization time, the results show that the SCB was still a better adsorbent than any BCs produced. We show that by varying the MB initial concentration, the adsorption capacity shows inflection points and fluctuations, even at different temperatures, which predict the endothermic or exothermic nature of the adsorption process. SCB showed the better adsorbent capacity at all initial concentrations (10, 30, 60, 110, 150, 190, 230, and 270 mg L−1) at 45 °C, and contact time of 360 min. With the proposed adsorption mechanism, there is a predominance of physical adsorption through Coulomb interactions, hydrogen bonds, and π-interactions. Therefore, the SCB is a low-cost alternative adsorbent for use in the treatment of wastewater contaminated with MB dye.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/su14031571/s1, Figure S1: On the left side, from top to bottom: two D-glucose molecules representing the cellulose polymer; two xylan molecules representing hemicelluloses; and a molecule of coniferyl alcohol representing the lignin. On the right side, adapted from Fanning et al. 1993: functional groups on the carbon surface. (a) aromatic stretching C=C, (b) ketene, (c,d) carbonates, (e) carboxylic acid, (f,g) lactones, (h) cyclic ethers, (i) cyclic anhydrides, (j) quinone, (k) phenol, and (l) alcohol. Figure S2: TGA ( Sustainability 14 01571 i028) and DTG ( Sustainability 14 01571 i029) of SCB in an inert atmosphere (N2), with a heating rate of 10 °C min−1, and flux of 20 mL min−1. Figure S3: MB adsorption: effect of adsorbent dosage for BC5. Figure S4: Effect of MB initial concentration in the adsorption: (a) maximum adsorption capacity and (b) removal efficiency. SCB ( Sustainability 14 01571 i030), BC1 ( Sustainability 14 01571 i031), BC2 ( Sustainability 14 01571 i032), BC3 ( Sustainability 14 01571 i033), BC4 ( Sustainability 14 01571 i034), and BC5 ( Sustainability 14 01571 i035); at 25 °C; Table S1: Function groups on the SCB and BC’s surfaces. Table S2: The surface area of the six types of adsorbents used in the study. Table S3: Initial and final pH of the test of pH effect on the MB adsorption. Table S4: Solution concentration before and after filtration and agitation.

Author Contributions

G.C.d.F. was carried out the experiments to obtain the biochar, the isotherms and carried out FTIR spectroscopy analysis. M.S.O. was carried TGA and MEV analysis. C.V.C.M. was carried out the experiments to obtain the surface area of biochar. J.C.P.d.S. conceived of the study, designed the study, coordinated the study and helped draft the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Goiano Federal Institute of Education, Science and Technology, Campus Rio Verde and National Council for Scientific and Technological Development (CNPq) providing an undergraduate research scholarship for G.C.d.F. National Post-Doctoral Program PNPD/CAPES for the scholarship No. 88887.342460/2019-00 of M.S.O.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The datasets supporting this article have been uploaded as part of the Supplementary Information.

Acknowledgments

The authors are thankful to the Goiano Federal Institute Campus Rio Verde for the research infrastructure. To the State University of Goias—CCET-UEG (Anápolis Campus of Exact and Technological Sciences Henrique Santillo, Innovation and Technology Center of UEG—CAiTec, Coordination of Superior Level Staff Improvement (CAPES), National Council for Scientific and Technological Development (CNPq), Research Support Foundation of the State of Goiás, (FAPEG)), Multi-User Analytical Center of the Master Program in Agrochemistry (PPGAq).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Wang, Z.; Gao, M.; Li, X.; Ning, J.; Zhou, Z.; Li, G. Efficient adsorption of methylene blue from aqueous solution by graphene oxide modified persimmon tannins. Mater. Sci. Eng. C 2020, 108, 110196. [Google Scholar] [CrossRef] [PubMed]
  2. Yin, Z.; Liu, N.; Bian, S.; Li, J.; Xu, S.; Zhang, Y. Enhancing the adsorption capability of areca leaf biochar for methylene blue by K2FeO4-catalyzed oxidative pyrolysis at low temperature. RSC Adv. 2019, 9, 42343–42350. [Google Scholar] [CrossRef] [Green Version]
  3. Liu, Y.; Jia, J.; Gao, T.; Wang, X.; Yu, J.; Wu, D.; Li, F. Rapid, Selective Adsorption of Methylene Blue from Aqueous Solution by Durable Nanofibrous Membranes. J. Chem. Eng. Data 2020, 65, 3998–4008. [Google Scholar] [CrossRef]
  4. Lei, H.; Hao, Z.; Chen, K.; Chen, Y.; Zhang, J.; Hu, Z.; Song, Y.; Rao, P.; Huang, Q. Insight into Adsorption Performance and Mechanism on Efficient Removal of Methylene Blue by Accordion-like V2CTx MXene. J. Phys. Chem. Lett. 2020, 11, 4253–4260. [Google Scholar] [CrossRef]
  5. 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]
  6. Li, Y.; Wang, S.; Shen, Z.; Li, X.; Zhou, Q.; Sun, Y.; Wang, T.; Liu, Y.; Gao, Q. Gradient Adsorption of Methylene Blue and Crystal Violet onto Compound Microporous Silica from Aqueous Medium. ACS Omega 2020, 5, 28382–28392. [Google Scholar] [CrossRef]
  7. Shen, Y.; Ni, W.-X.; Li, B. Porous Organic Polymer Synthesized by Green Diazo-Coupling Reaction for Adsorptive Removal of Methylene Blue. ACS Omega 2021, 6, 3202–3208. [Google Scholar] [CrossRef]
  8. Mohanty, A.K.; Vivekanandhan, S.; Pin, J.M.; Misra, M. Composites from renewable and sustainable resources: Challenges and innovations. Science 2018, 362, 536–542. [Google Scholar] [CrossRef] [Green Version]
  9. Mandal, A.; Singh, N.; Purakayastha, T.J. Characterization of pesticide sorption behaviour of slow pyrolysis biochars as low cost adsorbent for atrazine and imidacloprid removal. Sci. Total Environ. 2017, 577, 376–385. [Google Scholar] [CrossRef]
  10. Selvarajoo, A.; Oochit, D. Effect of pyrolysis temperature on product yields of palm fibre and its biochar characteristics. Mater. Sci. Energy Technol. 2020, 3, 575–583. [Google Scholar] [CrossRef]
  11. Hass, A.; Lima, I.M. Effect of feed source and pyrolysis conditions on properties and metal sorption by sugarcane biochar. Environ. Technol. Innov. 2018, 10, 16–26. [Google Scholar] [CrossRef]
  12. Brito, M.J.P.; Veloso, C.M.; Santos, L.S.; Bonomo, R.C.F.; Fontan, R.D.C.I. Adsorption of the textile dye Dianix® royal blue CC onto carbons obtained from yellow mombin fruit stones and activated with KOH and H3PO4: Kinetics, adsorption equilibrium and thermodynamic studies. Powder Technol. 2018, 339, 334–343. [Google Scholar] [CrossRef]
  13. Açikyildiz, M.; Gürses, A.; Karaca, S. Preparation and characterization of activated carbon from plant wastes with chemical activation. Microporous Mesoporous Mater. 2014, 198, 45–49. [Google Scholar] [CrossRef]
  14. Titiladunayo, I.F.; McDonald, A.G.; Fapetu, O.P. Effect of temperature on biochar product yield from selected lignocellulosic biomass in a pyrolysis process. Waste Biomass Valorization 2012, 3, 311–318. [Google Scholar] [CrossRef]
  15. López, J.E.; Builes, S.; Salgado, M.A.H.; Tarelho, L.A.C.; Arroyave, C.; Aristizábal, A.; Chavez, E. Adsorption of Cadmium Using Biochars Produced from Agro-Residues. J. Phys. Chem. C 2020, 124, 14592–14602. [Google Scholar] [CrossRef]
  16. Tayibi, S.; Monlau, F.; Fayoud, N.E.; Abdeljaoued, E.; Hannache, H.; Zeroual, Y.; Oukarroum, A.; Barakat, A. Production and Dry Mechanochemical Activation of Biochars Derived from Moroccan Red Macroalgae Residue and Olive Pomace Biomass for Treating Wastewater: Thermodynamic, Isotherm, and Kinetic Studies. ACS Omega 2021, 6, 159–171. [Google Scholar] [CrossRef]
  17. Chen, S.; Rotaru, A.E.; Shrestha, P.M.; Malvankar, N.S.; Liu, F.; Fan, W.; Nevin, K.P.; Lovley, D.R. Promoting interspecies electron transfer with biochar. Sci. Rep. 2014, 4, 5019. [Google Scholar] [CrossRef] [Green Version]
  18. Güzel, F.; Sayğılı, H.; Sayğılı, G.A.; Koyuncu, F.; Yılmaz, C. Optimal oxidation with nitric acid of biochar derived from pyrolysis of weeds and its application in removal of hazardous dye methylene blue from aqueous solution. J. Clean. Prod. 2017, 144, 260–265. [Google Scholar] [CrossRef]
  19. Shang, J.G.; Kong, X.R.; He, L.L.; Li, W.H.; Liao, Q.J.H. Low-cost biochar derived from herbal residue: Characterization and application for ciprofloxacin adsorption. Int. J. Environ. Sci. Technol. 2016, 13, 2449–2458. [Google Scholar] [CrossRef] [Green Version]
  20. Ravve, A. Principles of Polymer Chemistry, 3rd ed.; Springer Science & Business Media: New York, NY, USA, 2013; ISBN 978-1-4614-2211-2. [Google Scholar]
  21. Wu, J.; Yang, J.; Huang, G.; Xu, C.; Lin, B. Hydrothermal carbonization synthesis of cassava slag biochar with excellent adsorption performance for Rhodamine B. J. Clean. Prod. 2020, 251, 119717. [Google Scholar] [CrossRef]
  22. Igalavithana, A.D.; Choi, S.W.; Dissanayake, P.D.; Shang, J.; Wang, C.H.; Yang, X.; Kim, S.; Tsang, D.C.W.; Lee, K.B.; Ok, Y.S. Gasification biochar from biowaste (food waste and wood waste) for effective CO2 adsorption. J. Hazard. Mater. 2020, 391, 121147. [Google Scholar] [CrossRef] [PubMed]
  23. Oginni, O.; Singh, K. Influence of high carbonization temperatures on microstructural and physicochemical characteristics of herbaceous biomass derived biochars. J. Environ. Chem. Eng. 2020, 8, 104169. [Google Scholar] [CrossRef]
  24. Ganguly, P.; Sarkhel, R.; Das, P. Synthesis of pyrolyzed biochar and its application for dye removal: Batch, kinetic and isotherm with linear and non-linear mathematical analysis. Surf. Interfaces 2020, 20, 100616. [Google Scholar] [CrossRef]
  25. Al-Ghouti, M.A.; Al-Absi, R.S. Mechanistic understanding of the adsorption and thermodynamic aspects of cationic methylene blue dye onto cellulosic olive stones biomass from wastewater. Sci. Rep. 2020, 10, 15928. [Google Scholar] [CrossRef] [PubMed]
  26. Bouhadjra, K.; Lemlikchi, W.; Ferhati, A.; Mignard, S. Enhancing removal efficiency of anionic dye (Cibacron blue) using waste potato peels powder. Sci. Rep. 2021, 11, 2090. [Google Scholar] [CrossRef] [PubMed]
  27. El-Naggar, N.E.A.; Rabei, N.H. Bioprocessing optimization for efficient simultaneous removal of methylene blue and nickel by Gracilaria seaweed biomass. Sci. Rep. 2020, 10, 17439. [Google Scholar] [CrossRef]
  28. Uddin, M.K.; Nasar, A. Walnut shell powder as a low-cost adsorbent for methylene blue dye: Isotherm, kinetics, thermodynamic, desorption and response surface methodology examinations. Sci. Rep. 2020, 10, 7983. [Google Scholar] [CrossRef]
  29. Janković, B.; Manić, N.; Dodevski, V.; Radović, I.; Pijović, M.; Katnić, Đ.; Tasić, G. Physico-chemical characterization of carbonized apricot kernel shell as precursor for activated carbon preparation in clean technology utilization. J. Clean. Prod. 2019, 236, 117614. [Google Scholar] [CrossRef]
  30. Tong, Y.; McNamara, P.J.; Mayer, B.K. Adsorption of organic micropollutants onto biochar: A review of relevant kinetics, mechanisms and equilibrium. Environ. Sci. Water Res. Technol. 2019, 5, 821–838. [Google Scholar] [CrossRef]
  31. Chen, W.H.; Lu, K.M.; Lee, W.J.; Liu, S.H.; Lin, T.C. Non-oxidative and oxidative torrefaction characterization and SEM observations of fibrous and ligneous biomass. Appl. Energy 2014, 114, 104–113. [Google Scholar] [CrossRef]
  32. Li, W.; Yang, K.; Peng, J.; Zhang, L.; Guo, S.; Xia, H. Effects of carbonization temperatures on characteristics of porosity in coconut shell chars and activated carbons derived from carbonized coconut shell chars. Ind. Crops Prod. 2008, 28, 190–198. [Google Scholar] [CrossRef]
  33. Yaashikaa, P.R.; Kumar, P.S.; Varjani, S.J.; Saravanan, A. Advances in production and application of biochar from lignocellulosic feedstocks for remediation of environmental pollutants. Bioresour. Technol. 2019, 292, 122030. [Google Scholar] [CrossRef] [PubMed]
  34. Biswas, S.; Mohapatra, S.S.; Kumari, U.; Meikap, B.C.; Sen, T.K. Batch and continuous closed circuit semi-fluidized bed operation: Removal of MB dye using sugarcane bagasse biochar and alginate composite adsorbents. J. Environ. Chem. Eng. 2020, 8, 103637. [Google Scholar] [CrossRef]
  35. Patawat, C.; Silakate, K.; Chuan-Udom, S.; Supanchaiyamat, N.; Hunt, A.J.; Ngernyen, Y. Preparation of activated carbon fromDipterocarpus alatusfruit and its application for methylene blue adsorption. RSC Adv. 2020, 10, 21082–21091. [Google Scholar] [CrossRef]
  36. Li, X.Y.; Han, D.; Xie, J.F.; Wang, Z.B.; Gong, Z.Q.; Li, B. Hierarchical porous activated biochar derived from marine macroalgae wastes (Enteromorpha prolifera): Facile synthesis and its application on Methylene Blue removal. RSC Adv. 2018, 8, 29237–29247. [Google Scholar] [CrossRef] [Green Version]
  37. Wang, P.; Wu, C.; Guo, Y.; Wang, C. Experimental and theoretical studies on methylene blue and methyl orange sorption by wheat straw-derived biochar with a large surface area. Phys. Chem. Chem. Phys. 2016, 18, 30196–30203. [Google Scholar] [CrossRef]
  38. Lyu, H.; Gao, B.; He, F.; Zimmerman, A.R.; Ding, C.; Tang, J.; Crittenden, J.C. Experimental and modeling investigations of ball-milled biochar for the removal of aqueous methylene blue. Chem. Eng. J. 2018, 335, 110–119. [Google Scholar] [CrossRef]
  39. Xu, X.; Cao, X.; Zhao, L.; Zhou, H.; Luo, Q. Interaction of organic and inorganic fractions of biochar with Pb(ii) ion: Further elucidation of mechanisms for Pb(ii) removal by biochar. RSC Adv. 2014, 4, 44930–44937. [Google Scholar] [CrossRef]
  40. 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]
  41. Wang, Y.; Zhou, Y.; Jiang, G.; Chen, P.; Chen, Z. One-step fabrication of carbonaceous adsorbent from corncob for enhancing adsorption capability of methylene blue removal. Sci. Rep. 2020, 10, 12515. [Google Scholar] [CrossRef]
  42. Liu, C.; Wang, W.; Wu, R.; Liu, Y.; Lin, X.; Kan, H.; Zheng, Y. Preparation of Acid- And Alkali-Modified Biochar for Removal of Methylene Blue Pigment. ACS Omega 2020, 5, 30906–30922. [Google Scholar] [CrossRef] [PubMed]
  43. 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]
  44. Lagergren, S. Zur theorie der sogenannten adsorption geloster stoffe. K. Sven. Vetensk. Handl. 1898, 24, 1–39. [Google Scholar]
  45. Blanchard, G.; Maunaye, M.; Martin, G. Removal of heavy metals from waters by means of natural zeolites. Water Res. 1984, 18, 1501–1507. [Google Scholar] [CrossRef]
  46. Aharoni, C.; Tompkins, F.C. Kinetics of Adsorption and Desorption and the Elovich Equation. Adv. Catal. 1970, 21, 1–49. [Google Scholar] [CrossRef]
  47. McLintock, I.S. The elovich equation in chemisorption kinetics. Nature 1967, 216, 1204–1205. [Google Scholar] [CrossRef]
  48. Fan, S.; Tang, J.; Wang, Y.; Li, H.; Zhang, H.; Tang, J.; Wang, Z.; Li, X. Biochar prepared from co-pyrolysis of municipal sewage sludge and tea waste for the adsorption of methylene blue from aqueous solutions: Kinetics, isotherm, thermodynamic and mechanism. J. Mol. Liq. 2016, 220, 432–441. [Google Scholar] [CrossRef]
  49. Chen, S.; Qin, C.; Wang, T.; Chen, F.; Li, X.; Hou, H.; Zhou, M. Study on the adsorption of dyestuffs with different properties by sludge-rice husk biochar: Adsorption capacity, isotherm, kinetic, thermodynamics and mechanism. J. Mol. Liq. 2019, 285, 62–74. [Google Scholar] [CrossRef]
  50. Mosoarca, G.; Vancea, C.; Popa, S.; Gheju, M.; Boran, S. Syringa vulgaris leaves powder a novel low-cost adsorbent for methylene blue removal: Isotherms, kinetics, thermodynamic and optimization by Taguchi method. Sci. Rep. 2020, 10, 17676. [Google Scholar] [CrossRef]
  51. Sharma, P.; Kaur, H. Sugarcane bagasse for the removal of erythrosin B and methylene blue from aqueous waste. Appl. Water Sci. 2011, 1, 135–145. [Google Scholar] [CrossRef] [Green Version]
  52. Mohammad, N.; Atassi, Y. Adsorption of methylene blue onto electrospun nanofibrous membranes of polylactic acid and polyacrylonitrile coated with chloride doped polyaniline. Sci. Rep. 2020, 10, 13412. [Google Scholar] [CrossRef] [PubMed]
  53. Albroomi, H.I.; Elsayed, M.A.; Baraka, A.; Abdelmaged, M.A. Batch and fixed-bed adsorption of tartrazine azo-dye onto activated carbon prepared from apricot stones. Appl. Water Sci. 2017, 7, 2063–2074. [Google Scholar] [CrossRef] [Green Version]
Figure 1. FTIR spectra of the six adsorbents. SCB ( Sustainability 14 01571 i001), BC1 ( Sustainability 14 01571 i002), BC2 ( Sustainability 14 01571 i003), BC3 ( Sustainability 14 01571 i004), BC4 ( Sustainability 14 01571 i005), and BC5 ( Sustainability 14 01571 i006). The intense band of C-O and O-H (1100–1300 cm−1) in the BC1 spectra decreases with increasing the carbonization time, evidencing the degradation of cellulose, hemicellulose, and lignin.
Figure 1. FTIR spectra of the six adsorbents. SCB ( Sustainability 14 01571 i001), BC1 ( Sustainability 14 01571 i002), BC2 ( Sustainability 14 01571 i003), BC3 ( Sustainability 14 01571 i004), BC4 ( Sustainability 14 01571 i005), and BC5 ( Sustainability 14 01571 i006). The intense band of C-O and O-H (1100–1300 cm−1) in the BC1 spectra decreases with increasing the carbonization time, evidencing the degradation of cellulose, hemicellulose, and lignin.
Sustainability 14 01571 g001
Figure 2. SEM images of the adsorbents. The arrows are indicating some pores on the BCs surfaces.
Figure 2. SEM images of the adsorbents. The arrows are indicating some pores on the BCs surfaces.
Sustainability 14 01571 g002
Figure 3. (a) pHPZC of the adsorbents: SCB ( Sustainability 14 01571 i007), BC1 ( Sustainability 14 01571 i008), BC2 ( Sustainability 14 01571 i009), BC3 ( Sustainability 14 01571 i010), BC4 ( Sustainability 14 01571 i011), and BC5 ( Sustainability 14 01571 i012). (b) effect of the solution pH in the MB adsorption onto SCB and BCs: SCB ( Sustainability 14 01571 i013), BC1 ( Sustainability 14 01571 i014), BC2 ( Sustainability 14 01571 i015), BC3 ( Sustainability 14 01571 i016), BC4 ( Sustainability 14 01571 i017), and BC5 ( Sustainability 14 01571 i018).
Figure 3. (a) pHPZC of the adsorbents: SCB ( Sustainability 14 01571 i007), BC1 ( Sustainability 14 01571 i008), BC2 ( Sustainability 14 01571 i009), BC3 ( Sustainability 14 01571 i010), BC4 ( Sustainability 14 01571 i011), and BC5 ( Sustainability 14 01571 i012). (b) effect of the solution pH in the MB adsorption onto SCB and BCs: SCB ( Sustainability 14 01571 i013), BC1 ( Sustainability 14 01571 i014), BC2 ( Sustainability 14 01571 i015), BC3 ( Sustainability 14 01571 i016), BC4 ( Sustainability 14 01571 i017), and BC5 ( Sustainability 14 01571 i018).
Sustainability 14 01571 g003
Figure 4. Effect of contact time in the removal efficiency of MB onto SCB and BCs. (a) Experimental data with contact time varying between 0 and 1440 min. (b) Pseudo-First-Order kinetic model, (c) Pseudo-Second-Order kinetic model, and (d) Elovich kinetic model. SCB ( Sustainability 14 01571 i019), BC1 ( Sustainability 14 01571 i020), BC2 ( Sustainability 14 01571 i021), BC3 ( Sustainability 14 01571 i022), BC4 ( Sustainability 14 01571 i023), and BC5 ( Sustainability 14 01571 i024).
Figure 4. Effect of contact time in the removal efficiency of MB onto SCB and BCs. (a) Experimental data with contact time varying between 0 and 1440 min. (b) Pseudo-First-Order kinetic model, (c) Pseudo-Second-Order kinetic model, and (d) Elovich kinetic model. SCB ( Sustainability 14 01571 i019), BC1 ( Sustainability 14 01571 i020), BC2 ( Sustainability 14 01571 i021), BC3 ( Sustainability 14 01571 i022), BC4 ( Sustainability 14 01571 i023), and BC5 ( Sustainability 14 01571 i024).
Sustainability 14 01571 g004
Figure 5. Different adsorbents have different adsorption capacities, the pH influences the increase or decrease of the maximum adsorption in different concentrations; that is associated with different interactions in the adsorption process of MB onto SCB and BCs.
Figure 5. Different adsorbents have different adsorption capacities, the pH influences the increase or decrease of the maximum adsorption in different concentrations; that is associated with different interactions in the adsorption process of MB onto SCB and BCs.
Sustainability 14 01571 g005
Figure 6. Effect of temperature on the adsorption of MB onto SCB and BCs at different initial MB concentrations: 25 °C ( Sustainability 14 01571 i025), 35 °C ( Sustainability 14 01571 i026), and 45 °C ( Sustainability 14 01571 i027).
Figure 6. Effect of temperature on the adsorption of MB onto SCB and BCs at different initial MB concentrations: 25 °C ( Sustainability 14 01571 i025), 35 °C ( Sustainability 14 01571 i026), and 45 °C ( Sustainability 14 01571 i027).
Sustainability 14 01571 g006
Table 1. Kinetics models parameters of MB onto SCB and BCs.
Table 1. Kinetics models parameters of MB onto SCB and BCs.
ModelEquation/Ref.ParameterValue
SCBBC1BC2BC3BC4BC5
PFO q t = q 1   ( 1 e k 1 t )
[44]
q1 (mg g−1)18.389.4618.7319.0918.7918.56
k1 (min−1)0.340.200.040.040.140.12
R21.000.850.960.960.970.96
PSO q t = k 2 q 2 2 t 1 + k 2 q 2 t
[45]
q2 (mg g−1)18.7710.5919.4219.6819.7119.57
k2 (g mg−1 min−1)0.040.010.0030.0040.010.01
R21.000.910.980.980.990.99
Elovich q t = 1 β ln ( 1 + α β t )
[46]
α (mg g−1 min−1)101244.112.318.13177700
β (mg g−1)1.840.920.430.440.710.63
R20.990.980.990.980.980.99
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Fonseca, G.C.d.; Oliveira, M.S.; Martins, C.V.C.; de Souza, J.C.P. How the Carbonization Time of Sugarcane Biomass Affects the Microstructure of Biochar and the Adsorption Process? Sustainability 2022, 14, 1571. https://doi.org/10.3390/su14031571

AMA Style

Fonseca GCd, Oliveira MS, Martins CVC, de Souza JCP. How the Carbonization Time of Sugarcane Biomass Affects the Microstructure of Biochar and the Adsorption Process? Sustainability. 2022; 14(3):1571. https://doi.org/10.3390/su14031571

Chicago/Turabian Style

Fonseca, Gabriel Cabral da, Marilene Silva Oliveira, Carlos Vinicius Costa Martins, and João Carlos Perbone de Souza. 2022. "How the Carbonization Time of Sugarcane Biomass Affects the Microstructure of Biochar and the Adsorption Process?" Sustainability 14, no. 3: 1571. https://doi.org/10.3390/su14031571

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