Next Article in Journal
Spectroscopic Ellipsometry Studies on Solution-Processed OLED Devices: Optical Properties and Interfacial Layers
Previous Article in Journal
Correlation between Laboratory-Accelerated Corrosion and Field Exposure Test for High-Strength Stainless Steels
Previous Article in Special Issue
Amino-Functionalized Titanate Nanotubes: pH and Kinetic Study of a Promising Adsorbent for Acid Dye in Aqueous Solution
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Biochar Obtained from Caryocar brasiliense Endocarp for Removal of Dyes from the Aqueous Medium

by
André L. F. C. Melo
1,
Marcelo T. Carneiro
1,
Ariane M. S. S. Nascimento
2,
Alan I. S. Morais
2,
Roosevelt D. S. Bezerra
3,
Bartolomeu C. Viana
2,
Josy A. Osajima
2 and
Edson C. Silva-Filho
2,*
1
Federal Institute of Piauí, Floriano Campus, Floriano 64808-475, PI, Brazil
2
Interdisciplinary Laboratory for Advanced Materials, Teresina 64049-550, PI, Brazil
3
Federal Institute of Piauí, Teresina-Central Campus, Teresina 64000-040, PI, Brazil
*
Author to whom correspondence should be addressed.
Materials 2022, 15(24), 9076; https://doi.org/10.3390/ma15249076
Submission received: 24 September 2022 / Revised: 17 November 2022 / Accepted: 21 November 2022 / Published: 19 December 2022

Abstract

:
Given the increase in environmental pollution, especially of water, the emergence of studies that seek to develop strategies to mitigate/treat such effects have gained prominence in the world scientific community. Among the numerous adsorption processes, those made from biochar production stand out. This study analyzed the adsorption properties of the blue methylene model dye in the aqueous solution of biochar and activated biochar developed from pequi (Caryocar brasiliense) endocarp. The biochar was characterized, before and after adsorption, by infrared spectroscopy (FTIR), scanning electron microscopy (SEM), X-ray diffractometry (XRD), and thermogravimetric analysis (TG). The surface load of the materials was performed by the point of zero charge (pHPZC) method. The study also included analyses of contact time parameters and adsorbed concentration in the adsorption process. Morphological analysis showed that a more significant and profound number of fissures and pores appeared in the activated biochar compared to the biochar. Residual mass analysis evidenced that biochar lost about 15% more mass than the activated biochar, indicating that activation occurred satisfactorily. The adsorption process was well adjusted by pseudo-second-order kinetics and Langmuir’s isothermal model. The activated biochar achieved an excellent adsorption capacity of 476.19 mg.g−1, thus demonstrating to be a sound system for removing dyes from an aqueous medium.

Graphical Abstract

1. Introduction

The current context, marked by the advance of economic activity and worldwide population growth, has led to increased environmental pollution, especially water pollution. It has generated concern and received attention from society and the scientific community [1]. In particular, the growth of industrial activities has contaminated streams, rivers, and seas. Therefore, the academic community and society need to search for innovations that help combat such socioeconomic problems [2,3,4].
A significant contaminant of effluents is wastewater from the dye and paint industries, which should be classified as possible environmental pollutants. The dyes widely used in the textile industry are primarily synthetic and challenging to study or recover. They can generate a series of mutagenic and carcinogenic actions in humans and living organisms in aquatic environments. Pollution caused by dyes, when they react with complexes of chemicals present in the environment, becomes even more resistant to degradation. In particular, studies have shown that Methylene Blue, the dye used in this study, is widely used in the local textile industry in dyeing fabrics [5]. Thus, removing these organic compounds has become a significant technological challenge for the current scientific community [6].
The indiscriminate environmental pollution of dyes by the textile, pharmaceutical, cosmetic, and food industries has become a concern for the scientific community in recent years, which has led to research on techniques for removing these pollutants, with an emphasis on adsorption. The adsorption process is a versatile and affordable alternative for treating various effluents, including those of the textile industry [7,8,9].
One of the most used materials in dye adsorption processes is activated carbon or biochar. It can effectively remove pollutants, mainly organic, in adsorption processes and by the economic viability of production mechanisms and less polluting and generating fewer intermediaries [10,11,12].
Biochars can be produced from materials of plant origin (biomass), which are lignocellulosic, renewable, abundant, and have some degree of porosity [13]. The biomass residues studied for the production of activated carbon include the pequi endocarp (Caryocar brasiliense), whose carbonization of biomass via pyrolysis is a necessary thermal modification process [14,15].
The Cerrado biome of Brazil, one of the biomes richest in plant resources, is still minimally used in the daily life of Brazilians. This biome has diverse fruit-producing plant species that humans and animals can consume. Among the fruits of Cerrado, the pequi (Caryocar brasiliense) stands out because it has significant economic value through its fruits both in cooking and in the manufacture of liqueurs and medicinal syrups [16,17].
Despite the numerous applications of pequi, only some studies are available about the socioeconomic importance of using its mesocarp, endocarp, and almond. Because it is a typical fruit of the Cerrado, pequi is abundant in several states of Brazil. Finding productive uses for its residues is essential; therefore, research using these as biochar is relevant to the scientific community [18,19,20]. Works involving biochars used in the adsorption process produced from lychee seed, banana, and Moringa oleifera seed can be cited [21,22,23].
This work aimed to develop biochar and activated biochar from biomass residues (pequi endocarp), characterizing them by technical trust and applying them in removing a model organic dye through adsorption tests, varying the time, pH, and concentration parameters.

2. Materials and Methods

2.1. Materials and Reagents

For this work, the following reagents were used: methylene blue (C16H18SN3Cl, Dinâmica, Brasilpaís), Sodium Hydroxide (NaOH, Dinâmica, 98%, Brazil), Hydrochloric Acid (HCl, Dinâmica, 38%, Brazil), Sodium Chloride (NaCl, Dinâmica, 99%, Brazil), Distilled Water (Brazil). All reagents were analytical in degree, and no previous purification was required. The pequi endocarp residues were obtained at the free fair of the municipality of Floriano (PI), Avenida Bucar Neto, Centro, Floriano (PI), CEP 64800-000.

2.2. Biochar Preparation

The biochar was produced in five steps from the base of the pequi endocarp. Initially, pequi was collected, and any other residue was physically removed. In the second step, each part of the pequi (mesocarp, endocarp, and almond) was separated, and the endocarp was dried for 2 h by sunlight (30 ± 2 °C). Then, dried endocarp was crushed in a mill. Third, drying was performed in an oven for 6 h (80 ± 2 °C). In the fourth step, the dry material was placed in porcelain crucibles and charred in a muffle with a heating ratio of 10 °C min−1 until reaching the temperature of 500 °C. The vacuum muffle is composed of a closed system, with temperature control, heating, and cooling speed, therefore not allowing oxygen flow. Then, it was cooled to room temperature with a cooling ramp 10 °C min−1. The biochar obtained from this process was called biochar of endocarp (BE).
In the fifth step, the BE underwent a chemical activation with sodium hydroxide (NaOH), in which the ratio between BE and NaOH was 1:3 (m:m). The activation process involved mixing the biochar with the precursor (NaOH). The mixture was submitted to an activation heat treatment, with a heating rate of 5 °C min−1 until the temperature of 800 °C. Soon after, the material was submitted to a cooling process with a rate of 5 °C min−1. Finally, the material that underwent the activation heat treatment was placed in an agitator with distilled water for 1 h and then filtered the material on qualitative paper, drying it in an oven for 12 h at 105 °C. The activated material was called activated biochar (ABE). NaOH was chosen as a precursor because of its ease of acquisition, economic advantage, and use in works involving activated biochar [21,24,25,26].

2.3. Biochar Characterizations

The biochars were characterized by Scanning Electron Microscopy (SEM), Infrared Spectroscopy (FTIR), and Thermogravimetry (TG). The infrared spectra were obtained using an Vertex 70 FTIR spectrophotometer (Bruker, MA, USA), by the insert method, with 60 scans in the range of 600 to 4000 cm−1. The TG curves were obtained using a thermal analyzer (Shimadzu, TGA-50) (Quioto, Japan) using a heating rate of 10 °C min−1 between 25 and 800 °C in an inert nitrogen atmosphere. X-ray diffraction was performed using a Shimadzu instrument, model Labx-XDR 6000 (Quioto, Japan) [3,27,28,29,30].

2.4. Point of Zero Charge (pHPZC)

The pHPZC method determines the pH by balancing loads between the adsorbent surface and a solution. In the test, 20.0 mg samples of activated carbons were added to 20.0 mL of NaCl (0.1 mol.L−1). The pH was adjusted to 2 to 12, adding HCl solution at 1.0 mol.L−1 and/or NaOH at 1.0 mol.L−1. The initial pH was read using one pH meter. For 24 h, the mixtures were left under constant agitation at 140 rpm at 25 °C. After the described period, the solutions were centrifuged for 10 min at 5000 rpm, and the final pH of the solution was measured. The difference between the initial and final pH was calculated (ΔpH) (Equation (1)) and plotted from the obtained data, and then the intersection where ΔpH = 0 was determined, corresponding to the point of zero charge [31].
Δ p H = p H 0 p H f

2.5. Influence of pH

The effect of pH on dye adsorption was studied at pHs 4, 7, and 10 in triplicate, in which pH was adjusted with NaOH and HCl solutions. Biochar samples were in contact with 40.0 mL of a solution containing the methylene blue model dye (MB), with a concentration of 300.0 mg.L−1. The system was agitated for 24 h. After contact time, the solutions were centrifuged for 1 min at 14,000 rpm and diluted. The new concentrations were determined by reading the aliquots in a UV-vis spectrophotometer (Agilent, Cary 60, Santa Clara, CA, USA) using a pre-established calibration curve (665 nm). Finally, the adsorbed amount qe and (mg.g−1) and adsorption efficiency R (%) were calculated according to Equations (2) and (3), respectively [6,26].
q e = C i C f m · V
R = C i C f C i × 100 %
C i   and C f   represent the initial and final concentrations (mg. L−1) of the dye, respectively; m corresponds to the mass of the adsorbent in grams; and V is equivalent to the volume in liters of the dye solution used.

2.6. Adsorption Kinetics

The kinetic adsorption assay was performed in triplicate at 25 °C, with pH adjusted according to the best adsorption capacity. Different masses of biochar were added in different vials containing 40.0 mL of the dye solution, with a concentration of 100 mg.L-1 for biochar (BE) and 800 mg.L−1 for activated biochar (ABE). The solutions were constantly shaken at 140 rpm in 10, 20, 40, 60, 120, 220, 260, and 300 min intervals. Next, the samples were centrifuged, and the resulting solutions were measured in a UV-vis spectrophotometer to calculate the concentrations, according to Equation (2). The experimental data were adjusted to two kinetic models (pseudo-first-order and pseudo-second-order) to verify the best fit. The linearized form of the pseudo-first and pseudo-second-order kinetic models are expressed in Equations (4) and (5), respectively [32,33].
l n q e q t = l n q e k 1 . t
t q t = 1 k 2 . q e 2 + t q e
where qe and qt are the amount of dye adsorbed (mg.g−1) at equilibrium and time t (min), respectively; k1 represents the adsorption constant of the pseudo-first-order model (min−1), and k2 is the constant of the pseudo-second-order kinetic model (mg.g−1min−1).

2.7. Adsorption Isotherm

The test was performed in triplicate at a temperature of 25 °C, with pH adjusted according to the best adsorption. First, 70.0 mg of biochar was added to 40.0 mL of the dye in solution, varying the concentrations between 50 and 1000 mg.L−1, which agitated the best equilibrium time found in the kinetics experiment. Then, the samples were centrifuged, and measurements of UV-vis spectrophotometer solutions were performed to calculate the equilibrium concentrations resulting from adsorption, according to Equation (2). Finally, we investigated which adsorption isotherm model best fits the experimental data (Langmuir or Freundlich). Equation 6 expresses the linearized form of the Langmuir equation, and Equation (7) presents the linearized form of the Freundlich equation [34,35].
C q = 1 K L · q 0 + C q 0
l n q = l n K F + 1 n l n C
where q is the amount of species adsorbed by mass of the bioadsorbent (mg.g−1), q0 is the maximum amount of species adsorbed by mass of the bioadsorbent (mg.g−1), C represents the equilibrium concentration of the adsorbent (mg.L−1), KL is the Langmuir adsorption constant related to the chemical balance between adsorbent and adsorbent (mg.L−1), KF is the Freundlich adsorption constant related to adsorption capacity, n is the parameter related to the intensity of the adsorption process.

3. Results and Discussion

Pequi endocarp was used to generate the biochars, which have yellowish placement, according to the methodology presented, characterizing them as a fine black powder with low density, characteristic of this type of material. The color change after the thermal process indicates the production of biochar. Visually, it was impossible to verify differences between biochars after activation, so several characterizations were made to verify the properties of the materials obtained before and after adsorption.

3.1. Characterization of Biochars

Figure 1 presented the results of XRD, FTIR, thermogravimetric analysis, and pHPZC for the studied biochar samples of the endocarp (BE) and activated biochar (ABE).
The FTIR spectra of BE and ABE are presented in Figure 1A. The band around 1600 cm−1 is attributed to C=O vibrations in aromatic structures for biochar samples produced from the BE. After activation, the ABE presented a band around 1000 cm−1, usually found in oxidized biochars, and attributed to the elongation of C-O vibrations in acids, ethanols, phenols, ethers, and esters. Another band around 690 cm−1 is related to Na-O vibration from changes in the activation process, which indicates structural changes in the biochar [36,37,38].
The thermogravimetric analysis (TG) of the biochars is presented in Figure 1B. The TG curves of BE and ABE are b1 and b2, respectively, while those derived from the thermogravimetric curves (DTG) are b3 and b4, respectively. The thermal profile of both is divided into two stages. The first stage occurs around 45 °C for the BE and 53 °C for the ABE. This mass loss is related to water output and physically adsorbed gases. The second stage occurs around 599 °C for BE, with the output of molecules present in the material’s pores, indicating the need for activation. At a temperature above 700 °C and a maximum of 877 °C for ABE, an event with significant degradation velocity indicates the presence of aromatic compounds in the material after the first pyrolysis. Analysis of the residual mass verified that the BE material lost about 15% more mass than the ABE curve, indicating once again that the activation occurred successfully and the need to activate biochars to remove impurities and molecules present in the materials.
The pH of the medium affects the surface charge of the adsorbent and its degree of ionization and affects the adsorbed species. The pH at the point of zero charges (pHPZC) determines an index at which a surface tends to be positively or negatively charged as a function of pH. Figure 1C presents the pHPZC curves of the biochar obtained. The pHPZC value of the BE was 5.7, while the pHPZC value of the ABE was 6.4. For pH values below pHPZC of biochars BE (5.7) and ABE (6.4), the surface charge is positive, indicating that adsorption is favored for anionic species, while values above pHPZC favor adsorption of cationic species, which is the case of the Methylene Blue dye, used in the research. The results confirm that the medium’s pH influences the biochar’s surface. Ions (H+ or OH) present in the solution can interact with the active sites of the biochars, thus altering their charge balance [6,39]. The TG curves indicate that the presence of molecules, which were removed with activation, caused a greater acidity on the surface of the BE [40].
The study also presents the results of XRD characterizations before and after adsorption. The results can be found in the Supplementary Material (Figure S1).
Figure 2 shows the SEM images for the BE and activated carbon sample of the ABE studied with a size of 50 μm (A), 20 μm (B), and 10 μm. These images of the morphologies of biochars in different magnitudes will be used to analyze the shape and surface properties of adsorbents.
The BE images are presented in Figure 2A,B when compared to the ABE. Figure 2D,E have the morphology of three-dimensional particles in the form of nonhomogeneous blocks but with a lower pore presentation, which corroborates the data presented in the TG and the pHPZC values, indicating the filling with organic molecules present. Figure 2C,F of the ABE illustrate larger and deeper cracks and pores than in the BE, again indicating the successful activation and corroborating the techniques described above. The nonhomogeneous structures of biochar (ABE) are irregular and may positively influence adsorption [41,42].
Figure 3 shows the relationship between the amount of dye adsorbed by BE (a1) and ABE (a2) as a function of pH. It illustrates the resulting behavior of the amount adsorbed as a function of pH by studying the pH variation of the MB solution. The BE observed a significant increase in the adsorbed amount of MB at pH 10; thus, we decided to use pH 10 for the following tests of this biochar. On the other hand, the most effective removal of MB by the ABE at pH 7 and 10 were similar, so we chose to use the pH 7 of MB for the following tests of that biochar to optimize the process further. In addition, the adsorption was much higher after activation, proving the success of the activation process of the biochar obtained from the endocarp of pequi. As shown in pHPZC before, the increase in pH causes an increase in the negative charge, increasing electrostatic interactions between MB, the cationic dye, and the adsorbents with a negative charge [43].
The kinetic adsorption studies of BE and ABE are presented in Figure 4A and were adjusted by the pseudo-first-order (PFO) and pseudo-second-order (PSO) models. The linear form of these models, Figure S1, was analyzed using the correlation coefficient (R2), whose adjustment parameters are presented in Table 1.
Table 1 shows the PSO model describing the process of adsorption of MB by the biochars in both cases due to the high correlation coefficient (R2), 0.98920 for BE and 0.99990 for ABE. The results of qe calculated by the PSO model were close to the experimental values of qe, thus indicating that the adsorption of the MB dye involved chemical interactions. The analysis showed that the ABE adsorption had a value of qe equal to 476.19 mg.g−1, much higher than that of the BE, with q equal to 7.78 mg.g−1.
The adsorption isotherms of biochars are presented in Figure 4B, indicating a maximum adsorption capacity with 65.5 mg.g−1 for BE and 432.3 mg.g−1 for ABE, indicating a significant increase in adsorption capacity after activation. The isotherms were adjusted against the Langmuir and Freundlich models. The linear form of Langmuir and Freundlich models is presented in Figure S2 and was analyzed using the correlation coefficient (R2), whose statistical parameters are presented in Table 2.
Table 2 demonstrates that the process of MB adsorption by biochars is better described by the Langmuir isothermal model in both cases due to the values of the correlation coefficient (R2), 0.8784 for the BE and 0.9962 for the ABE. Thus, the analysis suggests that the adsorption process occurs homogeneously on the adsorbent surface in the monolayer and by chemical interactions, corroborating the kinetic data.
The experimental values for each biochar (Figure 4B) follow a concave curve marked by a rapid increase in initial concentrations, indicating a strong affinity of the dye to the materials, following a saturation, characteristic of Langmuir isotherms. Balance adsorption isotherms provide relevant information about this process, with concave isotherms being the most commonly used for microporous biochars. The results also indicate that the R2 adjustments were similar for both Langmuir and Freundlich in the case of BE; this suggests the presence of residues that promote a mixture between chemical and physical adsorption processes. As for the ABE biochar, which was better adjusted to the Langmuir model, the result suggests chemical adsorption in a monolayer [44,45].

3.2. Characterization of Biochars after Adsorption

After the dye adsorption, the materials were characterized to verify possible changes in characterizations and to prove the fixation of the dye in the biochars. The same acronyms were used to facilitate identification, adding A after the acronyms of the material, BE and ABE for biochar after adsorption, and activated biochar after adsorption, respectively.
The FTIR spectra of the materials after adsorption (Figure 5A) presented significant changes in the results. The band around 1600 cm−1 attributed to C=O vibrations in aromatic structures was only found for the BE. For ABE, the bands remained around 1000 cm−1, which indicates the elongation of C-O vibrations in acids, ethanols, phenols, ethers, and esters, and around 690 cm−1 related to Na-O vibration [46].
The thermal analysis of the BE biochar before and after adsorption is presented in Figure 5B. Figure 5B shows that the thermal profile of the biochar after adsorption remained similar to that presented before adsorption. The difference of about 0.95% refers to the dye mass that was adsorbed, which is a small percentage proportional to the amount adsorbed. It corroborates the low adsorption efficiency of the biochar mentioned above, indicating the need for surface activation. The BE and ABE curves (Figure 5C) verify that it has undergone a significant change in mass loss without changing the temperature range. The increase is proportional to the amount of dye adsorbed, proving by this technique that the dye was fixed on the surface and that the amount adsorbed after activation was much higher than that of biochar without activation, resulting in a difference in the mass loss in the final residue of about 18.18%. Thus, the analysis after adsorption strongly indicates the interaction of the dye with the biochars that reinforce the adsorption data and corroborate other techniques used [47,48].

4. Conclusions

Biochar of the pequi endocarp was successfully prepared by pyrolysis of biomass at 500 °C. The biochar underwent an efficient activation process, with an activating agent NaOH at 800 °C, proven by the characterizations with alteration in structure, change of thermal profile by removing some impurities present before activation, and alteration of some FTIR bands, due to the removal of some groups with activation. Both were applied in the adsorption of a model dye, with a better fit to the kinetic model of pseudo-second-order, Langmuir isotherm, and adsorption capacity at pH 10 for BE and 7 and 10 for ABE. The activated material, derived from biomass, renewable, and often discarded in nature without any proven use, is a promising adsorbent for the removal of MB from an aqueous environment with a preferential adsorption capacity of 476.19 mg.g−1, which revealed the feasibility of using it as a sustainable source to produce high performance activated carbon. It is noticed that the use of biochar produced from the endocarp of pequi presented itself as an excellent option in the adsorption process. Therefore, future work is suggested to analyze the production of biochar from the other compounds of pequi, almond, and mesocarp.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma15249076/s1, Figure S1. XRD for the studied samples of biochar of endocarp (BE) (a1) and activated biochar of the endocarp (ABE) (a2) (A); XRD pattern for the materials BE (b1) and ABE (b2) (B). Figure S2. Adjustment to the kinetic model of pseudo-first-order (A) and pseudo-second-order (B) of biochar (BE); Fit to the kinetic model of pseudo-first-order (C) and pseudo-second-order (D) of biochar (ABE). Figure S3. Adjustment to the isothermal model of Langmuir (A) and Freundlich (B) of biochar (BE); Fit to the isothermal model of biochar Langmuir (C) and Freundlich (D) (ABE).

Author Contributions

A.L.F.C.M.: methodology and writing—original draft; M.T.C.: writing—original draft; A.M.S.S.N.: conceptualization, visualization, and writing—review and editing; A.I.S.M.: methodology; B.C.V.: writing—review and editing; R.D.S.B.: writing—review and editing; J.A.O.: writing—review; E.C.S.-F.: supervision, project administration, visualization, and writing—review. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by CAPES and CNPq agencies (306176/2019-0), and IFPI and UFPI (Convênio 45/2018—FADEX).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors are grateful to CAPES and CNPq agencies (306176/2019-0) for financial support, and IFPI and UFPI (Convênio 45/2018—FADEX).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Silva, A.B.C.; Andrade, R.M.D.F.; Freire, F.B.; Nagalli, A.; Carvalho, K.Q.D.; Passig, F.H.; Kreutz, C. Application of red clay as adsorbent in the removal of textile dye Direct Blue of an aqueous solution. Mater.-Rio Jan. 2017, 22, 983–999. [Google Scholar]
  2. Mehra, S.; Singh, M.; Chadha, P. Adverse impact of textile dyes on the aquatic environment as well as on human beings. Toxicol. Int. 2021, 28, 165–176. [Google Scholar]
  3. Castro, A.S.; Franco, C.R.; Cidade, M.J.A. Adsorção de Corantes Azul Indosol, Laranja Indosol e Vermelho Drimaren em Solução Aquosa por Argila Branca Adsorption of Dyes Indosol Blue, Indosol Orange and Drimarene Red in Aqueous Solution by White Clay. Rev. Virtual Quim. 2018, 10, 1502–1515. [Google Scholar] [CrossRef]
  4. Silva, L.S.; Ferreira, F.J.; Silva, M.S.; Citó, A.M.; Meneguin, A.; Sábio, R.; Barud, H.S.; Bezerra, R.D.; Osajima, J.A.; Filho, E.C.S. Potential of amino-functionalized cellulose as an alternative sorbent intended to remove anionic dyes from aqueous solutions. Int. J. Biol. Macromol. 2018, 116, 1282–1295. [Google Scholar] [CrossRef] [PubMed]
  5. 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] [PubMed]
  6. Carneiro, M.T.; Barros, A.Z.B.; Morais, A.I.S.; Melo, A.L.F.C.; Bezerra, R.D.S.; Osajima, J.A.; Silva-Filho, E.C. Application of Water Hyacinth Biomass (Eichhornia crassipes) as an Adsorbent for Methylene Blue Dye from Aqueous Medium: Kinetic and Isothermal Study. Polymers 2022, 14, 2732. [Google Scholar] [CrossRef]
  7. Shindhal, T.; Rakholiya, P.; Varjani, S.; Pandey, A.; Ngo, H.H.; Guo, W.; Ng, H.Y.; Taherzadeh, M.J. A critical review on advances in the practices and perspectives for the treatment of dye industry wastewater. Bioengineered 2020, 12, 70–87. [Google Scholar] [CrossRef]
  8. Chai, W.S.; Cheun, J.Y.; Kumar, P.S.; Mubashir, M.; Majeed, Z.; Banat, F.; Ho, S.-H.; Show, P.L. A review on conventional and novel materials towards heavy metal adsorption in wastewater treatment application. J. Clean. Prod. 2021, 296, 126589. [Google Scholar] [CrossRef]
  9. Liu, S.; Wei, W.; Wu, S.; Zhang, F. Preparation of hierarchical porous activated carbons from different industrial lignin for highly efficient adsorption performance. J. Porous Mater. 2020, 27, 1523–1533. [Google Scholar] [CrossRef]
  10. Ghosh, A.; Razzino, C.D.A.; Dasgupta, A.; Fujisawa, K.; Vieira, L.H.S.; Subramanian, S.; Costa, R.S.; Lobo, A.O.; Ferreira, O.P.; Robinson, J.; et al. Structural and electrochemical properties of babassu coconut mesocarp-generated activated carbon and few-layer graphene. Carbon 2019, 145, 175–186. [Google Scholar] [CrossRef]
  11. Heylmann KK, A.; Lopes, B.V.; Afonso, T.F.; Demarco, C.F.; Cadaval Junior, T.R.; Quadro, M.S.; Andreazza, R. Production, characterization, and application of activated charcoal from peach kernel in textile effluent treatment. Eng. Sanit. E Ambient. 2021, 26, 485–494. [Google Scholar]
  12. Ferreira, A.A.; Guimarães, E.R.; Contador, J.C. Patente como instrumento competitivo e como fonte de informação tecnológica. Gest. Prod. 2009, 16, 209–221. [Google Scholar] [CrossRef] [Green Version]
  13. Leng, L.; Xiong, Q.; Yang, L.; Li, H.; Zhou, Y.; Zhang, W.; Jiang, S.; Li, H.; Huang, H. An overview on engineering the surface area and porosity of biochar. Sci. Total Environ. 2020, 763, 144204. [Google Scholar] [CrossRef] [PubMed]
  14. Ghesti, G.F.; Silveira, E.A.; Guimarães, M.G.; Evaristo, R.B.; Costa, M. Towards a sustainable waste-to-energy pathway to pequi biomass residues: Biochar, syngas, and biodiesel analysis. Waste Manag. 2022, 143, 144–156. [Google Scholar] [CrossRef]
  15. Scapin, E.; Rambo, M.K.D.; Viana, G.C.C.; Borges, M.S.; Rambo, M.C.; Carneiro, C. Production of Furanic Compounds and Organic Acids from Brazilian Pequi (Caryocar brasiliensis Camb.) Residues Using Green Chemistry. J. Braz. Chem. Soc. 2020, 31, 1383–1391. [Google Scholar] [CrossRef]
  16. da Luz, G.R.; Rodrigues, P.M.S.; Menino, G.C.O.; Coutinho, E.S.; Nunes, Y.R.S. Caracterização física de frutos e putâmens e taxa de ataque por carmenta sp. a pequizeiros (Ccaryocar brasiliense camb.) no norte de Minas Gerais. Rev. Bras. Frutic. 2011, 33, 746–756. [Google Scholar] [CrossRef] [Green Version]
  17. Fábio, F.B.F.; Wanessa, F.A.; Mônica, R.F.M.; Ariel, E.S.; Liliane, N.; Sandra, A.B.-R.; Frederico, A.G.G.; Andréia, V.C.D.A. Therapeutic applications of Caryocar brasiliense: Systematic review. J. Med. Plants Res. 2021, 15, 380–389. [Google Scholar] [CrossRef]
  18. Bailão, E.F.L.C.; de Oliveira, M.G.; de Almeida, L.M.; Amaral, V.C.S.; Chen, L.C.; Caramori, S.S.; de Paula, J.A.M.; de Melo Cruvinel, W.; Borges, L.L. Food Composition Data: Edible Plants in Cerrado. In Local Food Plants of Brazil; Springer: Cham, Switzerland, 2021; pp. 179–224. [Google Scholar]
  19. Qiu, R.; Lin, M.; Qin, B.; Xu, Z.; Ruan, J. Environmental-friendly recovery of non-metallic resources from waste printed circuit boards: A review. J. Clean. Prod. 2021, 279, 123738. [Google Scholar] [CrossRef]
  20. Oliveira, C.S.; Gonçalves, L.E.N.; Coutinho, M.P.; Peixoto, N.; Gatto, A. Aspectos Socioambientais da Comercialização de Pequi em Goiás. Floresta E Ambient. 2017, 24, e00058213. [Google Scholar] [CrossRef]
  21. Pai, S.; Kini, M.S.; Selvaraj, R. A review on adsorptive removal of dyes from wastewater by hydroxyapatite nanocomposites. Environ. Sci. Pollut. Res. 2021, 28, 11835–11849. [Google Scholar] [CrossRef]
  22. Roy, H.; Shahinoor Islam, M.d.; Tanvir Arifin, M.; 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. 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]
  24. Borba, L.L.; Cuba, R.M.F.; Terán, F.J.C.; Castro, M.N.; Mendes, T.A. Use of Adsorbent Biochar from Pequi (Caryocar brasiliense) Husks for the Removal of Commercial Formulation of Glyphosate from Aqueous Media. Braz. Arch. Biol. Technol. 2019, 62, e19180450. [Google Scholar] [CrossRef]
  25. Gao, Y.; Yue, Q.; Gao, B.; Li, A. Insight into activated carbon from different kinds of chemical activating agents: A review. Sci. Total Environ. 2020, 746, 141094. [Google Scholar] [CrossRef] [PubMed]
  26. Cazetta, A.L.; Vargas, A.M.M.; Nogami, E.M.; Kunita, M.H.; Guilherme, M.R.; Martins, A.C.; Silva, T.L.; Moraes, J.C.G.; Almeida, V.C. NaOH-activated carbon of high surface area produced from coconut shell: Kinetics and equilibrium studies from the methylene blue adsorption. Chem. Eng. J. 2011, 174, 117–125. [Google Scholar] [CrossRef]
  27. Amoim, F.A.C. Estudo de Remoção de Cu(II) em Meio Aquoso Utilizando Carvão Preparado a partir da Casca do Licurí (Syagrus coronata). Rev. Virtual Quim. 2017, 9, 2121–2134. [Google Scholar]
  28. Costa, L.F.; de Oliveira, D.G.; Moreira, F.M.S.; de Urzedo, A.P.F.M.; Cestarolli, D.T.; Bernardes-Silva, A.C. Use of Biochar and Advanced Oxidative Processes for the Removal of Propranolol from Simulated Aqueous Effluents. Rev. Virtual Quim. 2018, 10, 295–312. [Google Scholar] [CrossRef]
  29. Lin, Y.; Ma, J.; Liu, W.; Li, Z.; He, K. Efficient removal of dyes from dyeing wastewater by powder activated charcoal/titanate nanotube nanocomposites: Adsorption and photoregeneration. Environ. Sci. Pollut. Res. 2019, 26, 10263–10273. [Google Scholar] [CrossRef]
  30. Zhang, X.; Mao, X.; Pi, L.; Wu, T.; Hu, Y. Adsorptive and capacitive properties of the activated carbons derived from pig manure residues. J. Environ. Chem. Eng. 2019, 7, 103066. [Google Scholar] [CrossRef]
  31. Ferreira, F.J.; Silva, L.S.; da Silva, M.S.; Osajima, J.A.; Meneguin, A.B.; Santagneli, S.H.; Barud, H.S.; Bezerra, R.D.; Silva-Filho, E.C. Understanding kinetics and thermodynamics of the interactions between amitriptyline or eosin yellow and aminosilane-modified cellulose. Carbohydr. Polym. 2019, 225, 115246. [Google Scholar] [CrossRef]
  32. Lagergren, S.K. About the theory of so-called adsorption of soluble substances. Sven. Vetenskapsakad. Handingarl. 1898, 24, 1–39. [Google Scholar]
  33. Ho, Y.S.; McKay, G. Pseudo-second order model for sorption processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  34. Langmuir, I. The adsorption of gases on plane surfaces of glass, mica and platinum. J. Am. Chem. Soc. 1918, 40, 1361–1403. [Google Scholar] [CrossRef] [Green Version]
  35. Freundlich, H. Über die adsorption in lösungen. Z. Für Phys. Chem. 1907, 57, 385–470. [Google Scholar] [CrossRef]
  36. Mokhtar, M. Application of Synthetic Layered Sodium Silicate Magadiite Nanosheets for Environmental Remediation of Methylene Blue Dye in Water. Materials 2017, 10, 760. [Google Scholar] [CrossRef] [PubMed]
  37. Rodríguez-Zapién, K.V.; Torres-Pérez, J.; Reyes-López, S.Y. Environmental application of quartz-based construction waste: Tartrazine removal from aqueous media. Int. J. Environ. Sci. Technol. 2021, 19, 10381–10392. [Google Scholar] [CrossRef]
  38. Khalit, W.N.A.W.; Marliza, T.S.; Taufiq-Yap, Y.H. Synthesis and Characterizations of Nickel Supported on Activated Charcoal and its Application for Green Fuel Production. Mater. Sci. Forum 2020, 1010, 424–430. [Google Scholar] [CrossRef]
  39. Qiu, B.; Tao, X.; Wang, H.; Li, W.; Ding, X.; Chu, H. Biochar as a low-cost adsorbent for aqueous heavy metal removal: A review. J. Anal. Appl. Pyrolysis 2021, 155, 105081. [Google Scholar] [CrossRef]
  40. Doğan, M.; Sabaz, P.; Bicil, Z.; Kizilduman, B.K.; Turhan, Y. Activated carbon synthesis from tangerine peel and its use in hydrogen storage. J. Energy Inst. 2020, 93, 2176–2185. [Google Scholar] [CrossRef]
  41. Ouhammou, M.; Lahnine, L.; Mghazli, S.; Hidar, N.; Bouchdoug, M.; Jaouad, A.; Mandi, L.; Mahrouz, M. Valorisation of cellulosic waste basic cactus to prepare activated carbon. J. Saudi Soc. Agric. Sci. 2019, 18, 133–140. [Google Scholar] [CrossRef]
  42. Sanou, Y.; Alfa-Sika, M.S.L.; Vunain, E.; Pare, S. Preparation and Characterization of Husk Based Carbons: Effect of the Temperature. J. Environ. Treat. Tech. 2019, 7, 150–157. [Google Scholar]
  43. Bezerra, R.D.S.; Morais, A.I.S.; Osajima, J.A.; Nunes, L.C.C.; Silva Filho, E.C. Development of new phosphated cellulose for application as an efficient biomaterial for the incorporation/release of amitriptyline. Int. J. Biol. Macromol. 2016, 86, 362–375. [Google Scholar] [CrossRef] [PubMed]
  44. Yan, Q.; Li, J.; Cai, Z. Preparation and characterization of chars and activated carbons from wood wastes. Carbon Lett. 2021, 31, 941–956. [Google Scholar] [CrossRef]
  45. Gaydukova, A.M.; Nenasheva, A.S.; Kolesnikov, V.A.; Vetlugin, N.A. The Extraction of Activated Carbon from Aqueous Solutions in the Presence of Organic and Inorganic Compounds by the Electroflotation Method. J. Water Chem. Technol. 2021, 43, 116–122. [Google Scholar] [CrossRef]
  46. Jawad, A.H.; Abdulhameed, A.S.; Mastuli, M.S. Mesoporous Crosslinked Chitosan-Activated Charcoal Composite for the Removal of Thionine Cationic Dye: Comprehensive Adsorption and Mechanism Study. J. Polym. Environ. 2020, 28, 1095–1105. [Google Scholar] [CrossRef]
  47. Yin, Y.; Zhou, T.; Luo, H.; Geng, J.; Yu, W.; Jiang, Z. Adsorption of arsenic by activated charcoal coated zirconium-manganese nanocomposite: Performance and mechanism. Colloids Surf. A Physicochem. Eng. Asp. 2019, 575, 318–328. [Google Scholar] [CrossRef]
  48. del Hombre Bueno, B.R.; Boluda-Botella, N.; Rico, D.P. Removal of emerging pollutants in water treatment plants: Adsorption of methyl and propylparaben onto powdered activated carbon. Adsorption 2019, 25, 983–999. [Google Scholar] [CrossRef]
Figure 1. FTIR (A) for the studied biochar (BE) (a1) and biochar (ABE) (a2) (B); TG and DTG (B) curves for the studied samples of biochar (BE), (b1) and (b2), respectively, and curves for the studied samples of biochar (ABE) (b3) and (b4), respectively; (C) pHPZC result for the studied samples of biochar (BE) (c1), and biochar (ABE) (c2).
Figure 1. FTIR (A) for the studied biochar (BE) (a1) and biochar (ABE) (a2) (B); TG and DTG (B) curves for the studied samples of biochar (BE), (b1) and (b2), respectively, and curves for the studied samples of biochar (ABE) (b3) and (b4), respectively; (C) pHPZC result for the studied samples of biochar (BE) (c1), and biochar (ABE) (c2).
Materials 15 09076 g001
Figure 2. SEM images for the biochar (BE) sample with a size of 50 μm (A), 20 μm (B), and 20 μm (C), and for a sample of biochar (ABE) with a size of 50 μm (D), 20 μm (E), and 20 μm (F).
Figure 2. SEM images for the biochar (BE) sample with a size of 50 μm (A), 20 μm (B), and 20 μm (C), and for a sample of biochar (ABE) with a size of 50 μm (D), 20 μm (E), and 20 μm (F).
Materials 15 09076 g002
Figure 3. Relationship between the amount of dye adsorbed by the biochar (BE) (a1) and activated biochar (ABE) (a2) depending on the pH.
Figure 3. Relationship between the amount of dye adsorbed by the biochar (BE) (a1) and activated biochar (ABE) (a2) depending on the pH.
Materials 15 09076 g003
Figure 4. Adsorption kinetics for biochar (BE) (a1) and activated biochar (ABE) (a2) (A) and isotherm concentration for biochar (BE) (b1) and activated biochar (ABE) (b2) (B).
Figure 4. Adsorption kinetics for biochar (BE) (a1) and activated biochar (ABE) (a2) (A) and isotherm concentration for biochar (BE) (b1) and activated biochar (ABE) (b2) (B).
Materials 15 09076 g004
Figure 5. FTIR (A) for the studied biochar BE (a2), ABE (a1), BE (a4), and ABE (a3); TG (B) curves for the studied samples of biochar BE (b3), ABE (b4) and DTG (B) of biochar BE (b2) and ABE (b1) samples; TG (C) curves for the studied biochar samples BE (c3), ABE (c4), and DTG (C) of biochar samples BE (c1) and ABE (c2).
Figure 5. FTIR (A) for the studied biochar BE (a2), ABE (a1), BE (a4), and ABE (a3); TG (B) curves for the studied samples of biochar BE (b3), ABE (b4) and DTG (B) of biochar BE (b2) and ABE (b1) samples; TG (C) curves for the studied biochar samples BE (c3), ABE (c4), and DTG (C) of biochar samples BE (c1) and ABE (c2).
Materials 15 09076 g005
Table 1. Adjustment to the kinetic model of pseudo-first-order and pseudo-second-order for the biochar (BE) and activated biochar (ABE).
Table 1. Adjustment to the kinetic model of pseudo-first-order and pseudo-second-order for the biochar (BE) and activated biochar (ABE).
BiocharPseudo-First-OrderPseudo-Second-Order
qe
(mg.g−1)
K1
(min−1)
R2qe
(mg.g−1)
K2
(mg.g−1.min−1)
R2
BE11.20−0.009820.075187.780.139920.9892
ABE13.46−0.007000.91000476.191.615380.9999
Table 2. Adjustment to Langmuir and Freundlich isotherm model for biochar (BE) and activated biochar (ABE).
Table 2. Adjustment to Langmuir and Freundlich isotherm model for biochar (BE) and activated biochar (ABE).
BiocharLangmuirFreundlich
qoKLR2nKFR2
BE70.920.0039950.87843.0575.930.8749
ABE500.00.0098090.99624.0912.530.9105
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Melo, A.L.F.C.; Carneiro, M.T.; Nascimento, A.M.S.S.; Morais, A.I.S.; Bezerra, R.D.S.; Viana, B.C.; Osajima, J.A.; Silva-Filho, E.C. Biochar Obtained from Caryocar brasiliense Endocarp for Removal of Dyes from the Aqueous Medium. Materials 2022, 15, 9076. https://doi.org/10.3390/ma15249076

AMA Style

Melo ALFC, Carneiro MT, Nascimento AMSS, Morais AIS, Bezerra RDS, Viana BC, Osajima JA, Silva-Filho EC. Biochar Obtained from Caryocar brasiliense Endocarp for Removal of Dyes from the Aqueous Medium. Materials. 2022; 15(24):9076. https://doi.org/10.3390/ma15249076

Chicago/Turabian Style

Melo, André L. F. C., Marcelo T. Carneiro, Ariane M. S. S. Nascimento, Alan I. S. Morais, Roosevelt D. S. Bezerra, Bartolomeu C. Viana, Josy A. Osajima, and Edson C. Silva-Filho. 2022. "Biochar Obtained from Caryocar brasiliense Endocarp for Removal of Dyes from the Aqueous Medium" Materials 15, no. 24: 9076. https://doi.org/10.3390/ma15249076

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