Next Article in Journal
Synthesis of Aminoalkyl Sclareolide Derivatives and Antifungal Activity Studies
Next Article in Special Issue
Improved Viability of Probiotics via Microencapsulation in Whey-Protein-Isolate-Octenyl-Succinic-Anhydride-Starch-Complex Coacervates
Previous Article in Journal
UPLC-QE-Orbitrap-Based Cell Metabolomics and Network Pharmacology to Reveal the Mechanism of N-Benzylhexadecanamide Isolated from Maca (Lepidium meyenii Walp.) against Testicular Dysfunction
Previous Article in Special Issue
Use of Piranha Solution as An Alternative Route to Promote Bioactivation of PEEK Surface with Low Functionalization Times
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Novel Adsorbent Material from Plinia cauliflora for Removal of Cationic Dye from Aqueous Solution

by
Natalia Nara Janner
1,
Luana Vaz Tholozan
1,
Guilherme Kurz Maron
2,
Neftali Lenin Villarreal Carreno
2,
Alaor Valério Filho
2 and
Gabriela Silveira da Rosa
1,*
1
Chemical Engineering, Federal University of Pampa, Bagé 96413-172, Brazil
2
Graduate Program in Materials Science and Engineering, Technology Development Center, Federal University of Pelotas, Pelotas 96010-610, Brazil
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(10), 4066; https://doi.org/10.3390/molecules28104066
Submission received: 20 March 2023 / Revised: 8 May 2023 / Accepted: 9 May 2023 / Published: 12 May 2023

Abstract

:
The food industry is responsible for the generation of large amounts of organic residues, which can lead to negative environmental and economic impacts when incorrectly disposed of. The jaboticaba peel is an example of organic waste, widely used in industry due to its organoleptic characteristcs. In this study, residues collected during the extraction of bioactive compounds from jaboticaba bark (JB) were chemically activated with H3PO4 and NaOH and used to develop a low-cost adsorbent material for the removal of the cationic dye methylene blue (MB). For all adsorbents, the batch tests were carried out with the adsorbent dosage of 0.5 g L−1 and neutral pH, previously determined by 22 factorial design. In the kinetics tests, JB and JB-NaOH presented a fast adsorption rate, reaching equilibrium in 30 min. For JB-H3PO4, the equilibrium was reached in 60 min. JB equilibrium data were best represented by the Langmuir model and JB-NaOH and JB-H3PO4 data by the Freundlich model. The maximum adsorption capacities from JB, JB-NaOH, and JB-H3PO4 were 305.81 mg g−1, 241.10 mg g−1, and 122.72 mg g−1, respectively. The results indicate that chemical activations promoted an increase in the volume of large pores but interacted with functional groups responsible for MB adsorption. Therefore, JB has the highest adsorption capacity, thus presenting as a low-cost and sustainable alternative to add value to the product, and it also contributes to water decontamination studies, resulting in a zero-waste approach.

1. Introduction

Food industry production has been growing at a very fast pace, which can be related to population and urbanization growth, leading to concerns regarding the management of the residues produced by this sector [1,2]. Thus, the negative environmental, economic, and social impacts caused by food waste have been increasing over the years. To overcome this scenario, numerous studies involving the reuse of these wastes as value-added compounds have been recently reported [3,4,5,6,7].
Jaboticaba (Plinia cauliflora) is a Brazilian fruit originating from the Atlantic Forest that presents spherical and black-colored peels and is naturally adapted to adhere to the stem of the jaboticaba tree. The jaboticaba peel has a high concentration of phenolic compounds, which is responsible for the antimicrobial and antioxidant properties of the fruit [8]. Due to its organoleptic characteristics and high productivity, jaboticaba is used in the food industry for the production of jellies, sweets, and beverages [9,10].
The industrial processing of these products generates high amounts of residues, such as peels and seeds, which need to be treated and correctly disposed to prevent potential environmental impacts, resulting in costs for the industries. There are several reports describing the reuse of agro-industrial residues for the production of natural extracts. However, in most cases, the solid residues generated during the extraction processes are discarded, rejecting lignocellulosic materials with a high potential for the production of bioadsorbents. As a promising alternative, the conversion of wastes into valued-added materials has been studied for numerous applications, such as activated carbons for adsorption, biochars, and active compounds for food packaging [5,11,12,13,14,15]. Thus, recent studies have been investigating the use of industrial residues for the production of new adsorbent materials as an alternative to commercial activated carbon (AC), considering that AC is responsible for 70% of the costs involved in the adsorption operation due to the high energy consumption needed during thermal treatment [16,17].
The production of low-cost bioadsorbents can be carried out through chemical or physical activation. The chemical activation is usually cheaper and is normally based on processes that employs a dehydrating agent followed by neutralization, resulting in a porous material with high specific surface area [18]. The chemical activation can be carried out using various reagents, including H3PO4—which is preferred due to environmental and economic aspects as well as its ability to develop macro- and mesopores in the adsorbent [19], although it can improve the ash content of the material developed—and NaOH, which is widely used due to its low cost and capacity to produce high surface area adsorbents [20,21]. Accordingly, it is reported the use of organic wastes as environmentally friendly, sustainable, and renewable sources for the development of bioadsorbents applied in the removal of contaminants from aqueous solutions, such as golden trumpet tree bark, mulberry branches, and oxidized chitosans [22,23,24]. However, the reuse of jaboticaba peels as an alternative source has not been reported.
Since water decontamination has been widely investigated due to the increase in pollution by industrial activities, the use of the residues generated in the extraction of jaboticaba peels for the production of a bioadsorbent for treatment of aqueous effluents can be an interesting alternative for the valorization of this biomass, as it purposes the reuse of a residue generated by a product that was already developed using a residue as raw material—the jaboticaba peel [24]. The textile industry is responsible for a high percentage of this pollution, discarding high amounts of pigments in watercourses [11,25]. One of the most popular textile contaminants is the cationic dye methylene blue (MB) whose process of degradation is complex and has high toxicity [26]. Among the processes used in the removal of contaminants from effluents, adsorption is one of the most efficient [25,27,28,29], comprising a surface phenomenon that consists in transferring the contaminants from an aqueous solution to a solid phase. Besides high efficiency, this process is also considered excellent due to its low cost, availability, profitability, and ease of operation [30].
Herein, we describe for the first time an investigation involving the reuse of the solid waste produced in the process of extraction of phytochemical compounds from the jaboticaba peels, which is also considered a residue, as novel adsorbents of MB from simulated textile effluents, revealing a zero-waste approach. Additionally, the purpose of this work brings the development of materials with a cheaper preparation process and higher adsorption capacity when compared to materials with complex and expensive syntheses [31,32]. The adsorbents were characterized through scanning electron microscopy, thermogravimetric analysis, X-ray diffraction, and Fourier-transform infrared spectroscopy. The adsorbent–MB interaction was evaluated through experimental design and kinetic and isothermal models.

2. Results and Discussion

2.1. Characterization of JB, JB-H3PO4, and JB-NaOH Bioadsorbents

Figure 1a–c illustrate the sequence of modification of the raw material, represented by the jaboticaba fruit after harvest (a), the dried JB (b), and the residue generated in the extraction of biocompounds (c) [11]. During the extraction of phenolic compounds from the JB, there is a break between the bonds of different functional groups, leaving the surface of the bioadsorbent more susceptible to interaction with MB molecules [11]. The SEM images of JB, JB-H3PO4, and JB-NaOH are exhibited in Figure 1d–f, respectively, indicating that the three samples can be classified as a fine particulate material with irregular and heterogeneous surfaces, exhibiting cavities and protuberances. Moreover, the JB-H3PO4 and JB-NaOH samples presented larger cavities when compared to JB due to the chemical activation with NaOH and H3PO4, which results in the degradation of components such as lignin and hemicellulose [6,33]. These characteristics agree with previous studies that characterized adsorbents from lignocellulosic wastes [34].
Table 1 describes the results of the physical properties of JB, JB-H3PO4, and JB-NaOH. The data were subjected to the Tukey test at a significance level of 0.05, and there are no significant differences between the three adsorbents. The values of ρb, ρr, and ε are in accordance with the study described by Costa et al. (2015), which characterized açai pulp and obtained the value of 1.02 g cm−3 for real density [35].
Thermogravimetric analysis (TGA) was performed to evaluate the thermal stability of the samples, and the results are shown in Figure 2a. The TGA curves for all samples show four well-defined stages, and an initial weight loss at temperatures lower than 100 °C was observed for all samples due to the evaporation of moisture and low molecular compounds, representing less than 10% of the mass loss [36]. The DTG curves (Figure 2b) show peaks at approximately 280 °C for JB-H3PO4 and near to 320 °C for JB and JB-NaOH, which represents the simultaneous degradation of lignin (100–900 °C), hemicellulose (235–315 °C), and cellulose (315–370 °C) [37]. The degradation observed after 350 °C is characterized by the absence of peaks, which indicates the constant degradation of lignin [38].
In order to identify the functional groups attached to the surface of the samples, FTIR measurements were performed, and the spectrum recorded for JB, JB-H3PO4, and JB-NaOH are displayed in Figure 2c. All samples presented an intense band between 3600 and 3400 cm−1, which is attributed to the presence of alcoholic and phenolic groups (O-H) [30,39,40]. The slight bands around 2900 cm−1 are related to sp2 C-H from cellulose and sp3 C-H stretching from hemicellulose. The bands in the range of 1400–1700 cm−1 represent non-conjugated C=O stretching of hemicellulose structure [41], and the peaks at 1050 cm−1 are related to C-O-C of the β–glycosidic linkages and represent the stretch of lignocellulosic compounds [42]. The peak at 672 cm−1 observed in JB spectra is related to C-H stretching [38]. The FTIR spectra show a decrease in the peaks of O-H groups after activation, which can be related to the hydroxylation in the surface for the basic treatment. The JB-H3PO4 showed a decrease in the O-H peak and the disappearance of the C-H peak, related to the dehydration promoted by the treatment [43]. The modification in the bands indicates that the activation was able to modify chemical characteristics of the adsorbents. Furthermore, functional groups on the bioadsorbent surface often play a major role in the adsorption of various contaminants [26]. The main mechanism of MB adsorption is based on the electrostatic interaction between the dye molecules and the negatively charged C=O group. Additionally, the π–π interaction between the aromatic rings of MB molecule and the adsorbent can be considered a mechanism of MB adsorption [44]. The presence of these functional groups is considered very important in the adsorption of dyes, as they promote anchorage points that interact by electrostatic means [45].
The crystal structure of the materials was characterized by X-ray diffraction (XRD), and the XRD patterns of JB, JB-H3PO4, and JB-NaOH are shown in Figure 2d. The characteristic diffractograms of lignocellulosic materials with peaks in the region of 10–25° can be observed for the three samples [46]. Thus, a decrease in the cellulose peak for JB-H3PO4 when compared to JB was noticed, indicating that the acid treatment provided a disturbance in the crystal structure of the material, promoting the breakage of crystal chains and the widening of the amorphous structure. It suggests that a fraction of the crystalline cellulose was converted into amorphous cellulose [47]. The increase in the intensity of the peak at 20.13° for JB-NaOH can be attributed to the alkaline treatment, causing an increase in the crystallinity of the fibers due to the hydrolyzation of the amorphous regions during the NaOH treatment [48]. The absence of other peaks confirms that the three samples have a predominantly amorphous structure, which may favor the adsorption of methylene blue dye [26,49,50].

2.2. Factorial Design

A 22 factorial design was used to determine the optimized conditions for the adsorption of MB onto the adsorbents developed in this study. The conditions and results of the experimental design are shown in Table 2 and Figure 3, as well as the values of Ad (g L−1) and pH. The three samples exhibited that satisfactory results in terms of q and R values. The adsorbent that provided the best results was JB, followed by JB-H3PO4 and JB-NaOH. The tests performed with the JB reached a q value greater than 125 mg g−1 (run 3) and removal percentages above 96% (runs 6 and 7). The JB-H3PO4 reached a q of 90 mg g−1 (run 3) and a removal percentage greater than 97% (run 4), while the JB-NaOH presented an adsorption capacity of approximately 80 mg g−1 (run 3) and a 77% removal percentage (run 4). The Pareto charts showed that Ad and pH variables presented significant effects on both responses for the three samples, as they are located at the right side of the significance level (p ≤ 0.05). The interaction between the parameters had a significant effect only on the adsorption capacity of JB and JB-NaOH as well as in the removal percentage of JB. The adsorbent dosage was the parameter that had the greatest effects on the responses, and this effect was negative for the adsorption capacity for all adsorbents, which means that, the lower the adsorbent dosage, the greater the adsorption capacity. This relation can be explained by the number of available sites in each case [51,52]. When the Ad is low, there are fewer free sites, which results in a high adsorption capacity. On the other hand, when the dosage is high, there is a high number of free sites available, promoting a greater removal percentage [30]. Furthermore, a high adsorption dosage can cause aggregation of particles and reduce the adsorbent surface area.
For removal percentage, the adsorbent dosage had a positive effect, indicating that the greater the adsorbent dosage, the greater the removal percentage obtained, since at higher adsorption dosages, there will be more available sites and the accessibility to these pores become easier [53]. The pH had a positive effect on all responses, which can be attributed to the fact that MB is a cationic dye. This parameter has great importance to the adsorption process of dyes because it controls the intensity of electrostatic forces of charges provided by the dye molecule [54]. At a pH lower than 5, the H+ molecules of the dye will repulse the negatively charged adsorbent surface, which results in the protonation of the functional groups located at the adsorbent surface and, consequently, reduces the removal efficiency [55,56]. The high pH increases the electrostatic forces that are responsible for the attraction of the anions attached on the solid surface and the cations that form the dye molecules [57]. Furthermore, the removal of this dye is also favored in basic solutions, as at low pH, there is a high amount of positive ions in the solution, which can compete with the MB positive ions by occupying the available sites on the surface of the adsorbent [58]. After investigating the pH effects of the MB solutions on q and R values, it was selected to maintain the neutral pH of the MB solutions, since the effluents contaminated with MB are normally found in a neutral pH.
The ANOVA analyses (Table S2) showed a regression significance, which is confirmed by the Fvalue > Ftabled for both parameters for all adsorbents. These values are considered satisfactory, indicating that the models can predict the adsorption capacity and removal efficiency responses for the conditions used in this study. Equations (1)–(6) describe the dependent variables q and R for JB, JB-NaOH, and JB-H3PO4, respectively.
q = 76.2058 38.6597 A d + 1.9602 pH 1.7804 A d pH + 0.5411
q = 50.2224 20.1008 A d + 3.9109 pH 2.1300 A d pH + 0.7451
q = 57.7496 18.9842 A d + 6.8587 pH 19.1100
R = 93.2779 3.2894 A d + 1.6662 pH 1.7236 A d pH + 1.3014
R = 64.0535 9.5716 A d + 4.1260 pH 1.1318
R = 74.7555 16.4647 A d + 7.4046 pH 20.6382

2.3. Kinetic and Equilibrium Studies

Figure 4 illustrates the results obtained in the kinetic experiments of MB removal through the bioadsorbents. JB (Figure 4a) and JB-NaOH (Figure 4c) showed faster adsorption, while JB-H3PO4 (Figure 4b) presented a slightly slower adsorption. In total, 80% of MB was adsorbed by JB within 30 min, reaching equilibrium with approximately 90% of removal after 60 min. The JB-NaOH removed about 50% in the first 15 min of adsorption, reaching equilibrium at 30 min with 57% removal. Finally, JB-H3PO4 adsorbed approximately 40% of the dye in the first 30 min, reaching equilibrium with 50% removal in 60 min. Royer et al. (2009) used a Brazilian pine fruit shell for the production of an adsorbent material applied in the removal of MB, reaching equilibrium in 240 min [59]. Jawad et al. (2018) performed MB removal studies in adsorbent materials obtained from pomegranate (Punica granatum) peels, in which the equilibrium was reached in 120 min [60]. These results indicate that the studies carried out with JB, JB-NaOH, and JB-H3PO4 showed good adsorbent–adsorbate interaction, requiring a shorter time for the contaminant removal. The values obtained in the adjustments of the adsorption kinetic models for the three samples are presented in Table 3.
The fitting for JB occurred for the Elovich model, indicated by the highest value of R2 and lowest values of ARE and 2. This model indicates that the adsorption decreases with the increasing adsorbent–adsorbate contact time, which is attributed to the coverage of the adsorbent surface by the contaminant [61]. For JB-NaOH, the best fit occurred for the pseudo-second-order model, suggesting that the mechanism occurred was the chemisorption, characterized by a chemical reaction between the adsorbent and adsorbate, which also agrees with the good fit for the Elovich model [62,63]. The pseudo-first- and pseudo-second-order models presented a good fit for JB-H3PO4. The pseudo-first-order model indicates that van der Waals forces and H bonding can occur in the MB adsorption [64,65]. The intraparticle diffusion model is used to indicate the limiting steps of adsorption. The first step indicates the diffusion of the contaminant onto the adsorbent surface, while the intraparticle diffusion occurs at the second step. The third and last step represent the equilibrium plateau region, where the ions of the contaminant use all the available sites of the adsorbent [66]. Figure 4 shows that, for all materials, the linear portion of intra-particle diffusion data did not pass through the origin, which represents a variation of mass transfer in the initial ad final stages of the adsorption process. This indicates that the adsorption of MB onto JB, JB-H3PO4, and JB-NaOH was a multi-step process with adsorption on the external surface and diffusion into the interior [67].
According to Figure 5, all adsorbent materials presented favorable isotherms, and the convex format indicates that low concentrations of adsorbate promoted high adsorption [68]. The isotherms obtained are classified as type IV and V, representing a monolayer adsorption process, which is common on microporous materials [14,69]. Table 4 presents the values of the parameters obtained in the adjustment of the equilibrium isotherm models.
The results show that the best fit for JB occurred in the Temkin model, which was the one that presented the higher values of R2 and lower values of ARE and X2. This model indicates that the heat during the adsorption decreases linearly during the process due to the adsorbent–adsorbate interaction. The values of AT and B represent the binding energy and adsorption heat, respectively. It is possible to observe that the value of B was positive, which indicates that the adsorption process was exothermic [70]. Additionally, this model is an indication that the molecules of adsorbate are attracted on the surface with strong intermolecular attraction [71]. Consequently, the adsorption stops when the absorbent surface saturates [71,72]. The kL parameter of 0.07 L mg−1 obtained for JB is related to the free energy of adsorption, i.e., the affinity between the adsorbent and adsorbate.
JB-H3PO4 and JB-NaOH had a better fit to the Freundlich model, indicating that the adsorbent–adsorbate interaction is heterogeneous, which promotes an exponential distribution of free energy of adsorption. Thus, it is assumed that the adsorption occurs infinitely with the increase in the concentration of adsorbate in the solution, with no equilibrium plateau (characteristic of Langmuir model) [72,73]. In this regard, there is no qmax parameter in this model. For comparison with the literature, the maximum experimental values of 122.72 mg g−1 and 241.10 mg g−1 for JB-H3PO4 and JB-NaOH are going to be used, respectively.
Compared to the literature, the residue from the extraction of the bioactive compounds from jaboticaba bark presents satisfactory results when used in the removal of MB. Table 5 presents a comparison of the maximum adsorption capacity of various similar adsorbents for MB. These results confirm the high potential of the proposed methodology to produce an alternative MB adsorbent material from residues. Comparing the qmax values of JB, JB-NaOH, and JB-H3PO4, the best result was obtained with JB, which was not submitted to chemical activation, reaching a qmax of 314 mg g−1. The high value of qmax obtained for JB can be related to the origin of the raw material, where the extraction might have caused an increase in the porosity of the material [74]. The lower values for JB-NaOH and JB-H3PO4 could be attributed to a deterioration of the structure caused by the chemical activation [75]. This result can be considered very promising, considering that the activation process has some disadvantages such as corrosion of the equipment, high cost of reagents, and difficult recovery of chemicals [76]. Additionally, the materials developed in this study present characteristics of catalysts, which are also used in water decontamination [77]. In this sense, the process investigated in the present study might be related to a catalysis, since this process consists of the migration of the contaminant molecules to the surface of the catalyst. When the contaminant molecule is bigger than the pore diameter, the migration occurs through intra-particle diffusion [78]. In addition, the low cost of the production of JB generates a smaller environmental impact by reducing the release of chemicals in the environment and avoiding the washing step, which requires a large amount of water and is commonly needed in chemical activation.

2.4. Adsorption Mechanism

The adsorption process is very complex and involves multiple mechanisms. These mechanisms are important to understand the basic principles of the adsorption process. The possible interactions that occur between the MB molecules (cationic dye) and JB, JB-H3PO4, and JB-NaOH can be of three types: H bonding, π–π bonds, and electrostatic interactions with the functional groups attached on the surface of the adsorbents [87]. Figure 6 presents a scheme of these interactions. Electrostatic interactions can occur by the attraction between the cations of the MB and the surface of the adsorbent when the adsorbent has a negative charge. The π–π interactions can occur due to the presence of the aromatic rings of lignin in the structures of the adsorbents and MB.
From the FTIR analysis, some functional groups present in the adsorbents (-OH and C=O) were found to be derived from the main biomass components (cellulose, hemicellulose, and lignin). These groups may have interacted by hydrogen bonds as well as Van der Waals forces [87,88]. Among the functional groups, JB has the highest amount of the OH and C=O groups on the surface. The interaction of the OH group by hydrogen bonding between the aromatic N of MB probably aimed for a higher result for qmax. From JB-NaOH and JB-H3PO4, the analysis presented in Figure 2, Figure 3 and Figure 4 suggests that the chemical activation caused a reorganization in the JB structure, degrading the smaller pores that had been formed in the extraction of the phenolic compounds, thus forming larger pores [89]. Additionally, the chemical activation interacted with functional groups that could aid in MB adsorption [75].

3. Materials and Methods

3.1. Materials

The solid residues from jaboticaba bark were obtained from a study involving the extraction of bioactive compounds [11]. Prior to the chemical activation process, the residues were dried at 105 °C for 24 h, and here they are entitled JB. All reagents used were purchased from Sigma-Aldrich (Burlington, MA, USA).

3.2. Chemical Activation

JB residues were chemically activated with phosphoric acid (H3PO4) and sodium hydroxide (NaOH). The activation with phosphoric acid was performed according to the methodology proposed by Portinho et al. (2017) [6]. In a typical experiment, 10 g of dry biomass was mixed with the H3PO4 (85% w/w) in a 4:1 ratio. The mixture was agitated on a shaker at 100 rpm for 24 h at room temperature. For the activation with NaOH, the methodology proposed by Silva et al. (2020) was applied [33]. In this procedure, 10 g of JB was soaked in a NaOH solution (20% w/v) at a 1:3 ratio for 24 h at room temperature. Both activated samples were then oven-dried at 105 °C for 24 h. Finally, the materials were washed with distilled water until reaching a neutral pH, dried in an oven, and sieved through an electromagnetic sieve shaker (Bertel, model 4830) to obtain the particle < 495 μm [30]. The samples are here called JB-H3PO4 (H3PO4-activated jaboticaba bark) and JB-NaOH (NaOH-activated jaboticaba bark).

3.3. Characterization

The real density (ρr) values were obtained through helium pycnometer using a gas pycnometer (Quantachrome Corporation, Boynton Beach, FL, USA) at 18 psig and 19.5 °C. The values of the bulk density were obtained with the use of a graduated cylinder. The porosity of the particle bed ( ε ) was calculated according to Equation (7), which relates the ρr (g cm−3) and ρb (g cm−3).
ε = 1 ρ b ρ r
The thermal decomposition of the materials was studied through thermogravimetric analysis, using a thermal analyzer (TA 60WS Shimadzu, Japan) and a thermobalance (TGA 50 Shimadzu, Kyoto, Japan). Analysis was carried out using approximately 8 mg of sample, heated from 25 °C to 700 °C with a heating rate of 10 °C min−1 in a 50 mL min−1 N2 atmosphere. X-ray diffraction technique (XRD, Rigaku Ultima IV diffractometer, Tokyo, Japan) with Cu-Kα radiation at a voltage of 40 kV was used to investigate the crystalline structure of the adsorbents. The XRD patterns were recorded in the 2θ range from 5° to 70°.
Fourier-transform infrared spectroscopy (FTIR) was carried out to identify the main functional groups on the surface of JB, JB-H3PO4, and JB-NaOH. The results were obtained in a Shimadzu 01722, IR Prestige, Japan spectrophotometer, using diffuse reflectance with KBr. The spectra were obtained over the range of 400–4000 cm−1 and a resolution of 4 cm−1. The morphological characteristics of the materials were examined by scanning electron microscopy (SEM, JEOL, JSM 6610 L V, Akishima, Japan) at an acceleration voltage of 5 Kv and at magnifications of 100, 300, and 500×.

3.4. Adsorption of MB onto JB, JB-H3PO4, and JB-NaOH

The MB concentrations in the liquid phase were quantified by a UV–VIS spectrophotometer (Kazuaki, II-226, Hangzhou, China) at 665 nm, using a standard curve with concentrations ranging from 0.5 to 50. The adsorption trials were carried out in batch, and the operating range conditions were obtained from preliminary trials. For the batch studies, 25 mL of MB solution (at a concentration of 70) was added to 0.5 g L−1 of adsorbent dosage. Then, the adsorbent–adsorbate mixtures were agitated in a shaker (NOVA ÉTICA, model 109-1, Minas Gerais, Brazil) for a varied time at 120 rpm and centrifuged for 10 min at 3000 rpm to separate the adsorbents from the MB solution. The best condition for adsorption capacity and percentage of MB removal onto the jaboticaba bioadsorbents was evaluated by the experimental design carried out through a 22 factorial with 4 tests and 3 central points performed in duplicate. The levels and real values of the evaluated parameters are presented in Table 6. The remaining MB concentrations were quantified in a spectrophotometer UV-Vis, and the adsorption capacity (q) and removal percentage (R) were calculated according to Equations (8) and (9), respectively,
q = C i C f M   V
R = C i C f C i 100
where Ci is the initial concentration (mg L−1), Cf is the final concentration (mg L−1), M is the mass of adsorbent (g), and V is the volume of solution (L).
Statistical analysis of experimental data was performed with the use of the 10th version of STATISTICA. The Pareto diagram was used to determine the effect and verify the significant variables for q and R responses. The adequacy of the models was analyzed through a variance analysis (ANOVA) with F-test (Fisher test). The response surface was generated from the values predicted by the models and used to define the conditions of adsorption tests. The first-order model that predicts the responses is represented by Equation (10).
Y = β o + i = 1 k β i X i + i < j   β i j X i X j
where βo is the constant coefficient, βi is the linear coefficient, βii is the quadratic coefficient, βij is the interaction coefficient, Xi is the independent variable level, k is the number of variables, e is the experimental error, and Y is the model response.
The kinetic experiments were performed with the adsorbent dosage, MB concentration, and pH predicted by experimental design. The samples were agitated in a shaker at 150 rpm and collected at different times (2–120 min). The adsorption kinetic data obtained were fitted to pseudo-first-order [90] pseudo-second-order [91], Elovich [92], and Weber and Morris models [93]. In the adsorption equilibrium experiments, the MB solutions at different concentration ratios (30–700) were kept in contact with the adsorbent until equilibrium, previously determined by the kinetic studies. The equilibrium data were fitted to Langmuir [72] and Freundlich [73] models. Table S1 (Supplementary Materials) presents the adsorption isotherm and kinetic equations used in this study.
The kinetic and isotherm parameters were evaluated through non-linear regression. The adsorption capacities obtained in the models were analyzed based on the values of average relative error (ARE) [94] and Chi-square (X2) [95], represented by Equations (11) and (12), respectively. These analyses were applied as fit quality indicators.
A R E = 100 n n q e x p q p r e d q e x p
X 2 = q e x p q p r e d 2 n n N N
where qexp is the experimental value, qpred is the value predicted by the model, nn is the number of experiments observed, and NN is the number of parameters in the model.

4. Conclusions

This work described the production of a novel adsorbent material for MB from jaboticaba bark residues. The JB residues were chemically activated with H3PO4 and NaOH, aiming to evaluate their effect on the adsorption properties of the material. The results were in agreement with previous reports described in the literature for similar materials. The TGA and XRD analyses showed that the activated materials had slightly lower cellulose peaks compared to JB, indicating that chemical activation did not have a significant influence on the degradation of this compound. Thus, JB and JB-H3PO4 reached equilibrium in 60 min, whilst JB-NaOH presented the fastest rate, reaching equilibrium in 30 min. The kinetic models that best fit the experimental data for JB and JB-NaOH were the Elovich and pseudo-second-order models, respectively. JB-H3PO4 best fit to pseudo-second-order and intraparticle diffusion models. The maximum adsorption capacity values obtained for JB, JB-NaOH, and JB-H3PO4 were 305.81 mg g−1, 241.10 mg g−1, and 122.72 mg g−1, respectively. The Langmuir isotherm model best described the experimental data for JB. For JB-NaOH and JB-H3PO4, the best fit occurred for the Freundlich model. The best adsorption capacity was obtained with the adsorbent material without chemical treatment. This is a promising outcome, resulting in an effective, cheap, and eco-friendly product. This research reveals a zero-waste approach, since the raw material used for the fabrication of the adsorbent was a residue from the agroindustry in which the extraction of biocompounds had already been carried out, promoting the complete use of fruit.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28104066/s1, Table S1: Kinetics and isotherm models; Table S2 ANOVA for the factorial design.

Author Contributions

Conceptualization, N.N.J., A.V.F. and G.S.d.R.; methodology, N.N.J., A.V.F. and G.S.d.R. software, N.N.J. and A.V.F.; validation, N.N.J., A.V.F. and G.S.d.R.; formal analysis, N.N.J.; investigation, N.N.J., A.V.F. and G.S.d.R.; resources, G.S.d.R.; data curation, N.N.J. and A.V.F.; writing—original draft preparation, L.V.T., A.V.F. and G.K.M.; writing—review and editing, L.V.T., A.V.F., G.K.M., N.L.V.C. and G.S.d.R.; visualization, A.V.F.; supervision, N.L.V.C. and G.S.d.R.; project administration, G.S.d.R.; funding acquisition, N.L.V.C. and G.S.d.R. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Coordination for the Improvement of Higher Education Personnel (CAPES), grant number 001 and Foundation for Research Support of the State of Rio Grande do Sul (FAPERGS)/Brazilian Service of Support for Micro and Small Enterprises (SEBRAE) 03/2022—PDEmp (22/2551-0001076-8-6).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

The authors would like to thank the support from National Council for Scientific and Technological Development (CNPq).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples used in this study are available upon request through the corresponding author.

Abbreviations

JB: jaboticaba bark; MB, methylene blue; AC, activated carbon; thermogravimetric analysis (TGA); thermogravimetric derivative (DTG); X-ray diffraction (XRD); H3PO4-activated jaboticaba bark (JB- H3PO4); NaOH-activated jaboticaba bark (JB-NaOH); real density (ρr); porosity of the particle bed (ε); bulk density (ρb); Fourier-transform infrared spectroscopy (FTIR); scanning electron microscopy (SEM); average relative error (ARE); Chi-square (X2).

References

  1. Parmar, A.; Nema, P.K.; Agarwal, T. Biochar Production from Agro-Food Industry Residues: A Sustainable Approach for Soil and Environmental Management. Curr. Sci. 2014, 107, 1673–1682. [Google Scholar]
  2. Kibler, K.M.; Reinhart, D.; Hawkins, C.; Motlagh, A.M.; Wright, J. Food Waste and the Food-Energy-Water Nexus: A Review of Food Waste Management Alternatives. Waste Manag. 2018, 74, 52–62. [Google Scholar] [CrossRef] [PubMed]
  3. Antonino, R.S.C.M.D.Q.; Fook, B.R.P.L.; Lima, V.A.D.O.; Rached, R.Í.D.F.; Lima, E.P.N.; Lima, R.J.D.S.; Covas, C.A.P.; Fook, M.V.L. Preparation and Characterization of Chitosan Obtained from Shells of Shrimp (Litopenaeus Vannamei Boone). Mar. Drugs 2017, 15, 141. [Google Scholar] [CrossRef] [PubMed]
  4. Uzun, B.B.; Pütün, A.E.; Pütün, E. Composition of Products Obtained via Fast Pyrolysis of Olive-Oil Residue: Effect of Pyrolysis Temperature. J. Anal. Appl. Pyrolysis 2007, 79, 147–153. [Google Scholar] [CrossRef]
  5. Martiny, T.R.; Raghavan, V.; de Moraes, C.C.; da Rosa, G.S.; Dotto, G.L. Bio-Based Active Packaging: Carrageenan Film with Olive Leaf Extract for Lamb Meat Preservation. Foods 2020, 9, 1759. [Google Scholar] [CrossRef]
  6. Portinho, R.; Zanella, O.; Féris, L.A. Grape Stalk Application for Caffeine Removal through Adsorption. J. Environ. Manag. 2017, 202, 178–187. [Google Scholar] [CrossRef]
  7. Valério Filho, A.; Avila, L.B.; Lacorte, D.H.; Martiny, T.R.; Rosseto, V.; Moraes, C.C.; Dotto, G.L.; Carreno, N.L.V.; da Rosa, G.S. Brazilian Agroindustrial Wastes as a Potential Resource of Bioative Compounds and Their Antimicrobial and Antioxidant Activities. Molecules 2022, 27, 6876. [Google Scholar] [CrossRef]
  8. Silva, M.C.; de Souza, V.B.; Thomazini, M.; da Silva, E.R.; Smaniotto, T.; Carvalho, R.A.d.; Genovese, M.I.; Favaro-Trindade, C.S. Use of the Jabuticaba (Myrciaria Cauliflora) Depulping Residue Toproduce a Natural Pigment Powder with Functional Properties. LWT Food Sci. Technol. 2014, 55, 203–209. [Google Scholar] [CrossRef]
  9. Wu, S.B.; Long, C.; Kennelly, E.J. Phytochemistry and Health Benefits of Jaboticaba, an Emerging Fruit Crop from Brazil. Food Res. Int. 2013, 54, 148–159. [Google Scholar] [CrossRef]
  10. Avila, L.B.; Barreto, E.R.C.; de Souza, P.K.; Silva, B.D.Z.; Martiny, T.R.; Moraes, C.C.; Morais, M.M.; Raghavan, V.; da Rosa, G.S. Carrageenan-Based Films Incorporated with Jaboticaba Peel Extract: An Innovative Material for Active Food Packaging. Molecules 2020, 25, 5563. [Google Scholar] [CrossRef]
  11. Avila, L.B.; Fontes, M.R.V.; Zavareze, E.d.R.; Moraes, C.C.; Morais, M.M.; da Rosa, G.S. Recovery of Bioactive Compounds from Jaboticaba Peels and Application into Zein Ultrafine Fibers Produced by Electrospinning. Polymers 2020, 12, 2916. [Google Scholar] [CrossRef] [PubMed]
  12. Demiral, H.; Güngör, C. Adsorption of Copper(II) from Aqueous Solutions on Activated Carbon Prepared from Grape Bagasse. J. Clean. Prod. 2016, 124, 103–113. [Google Scholar] [CrossRef]
  13. Cunha, M.R.; Lima, E.C.; Lima, D.R.; da Silva, R.S.; Thue, P.S.; Seliem, M.K.; Sher, F.; dos Reis, G.S.; Larsson, S.H. Removal of Captopril Pharmaceutical from Synthetic Pharmaceutical-Industry Wastewaters: Use of Activated Carbon Derived from Butia Catarinensis. J. Environ. Chem. Eng. 2020, 8, 104506. [Google Scholar] [CrossRef]
  14. Zazycki, M.A.; Godinho, M.; Perondi, D.; Foletto, E.L.; Collazzo, G.C.; Dotto, G.L. New Biochar from Pecan Nutshells as an Alternative Adsorbent for Removing Reactive Red 141 from Aqueous Solutions. J. Clean. Prod. 2018, 171, 57–65. [Google Scholar] [CrossRef]
  15. Raupp, Í.N.; Filho, A.V.; Arim, A.L.; Muniz, A.R.C.; da Rosa, G.S. Development and Characterization of Activated Carbon from Olive Pomace: Experimental Design, Kinetic and Equilibrium Studies in Nimesulide Adsorption. Materials 2021, 14, 6820. [Google Scholar] [CrossRef]
  16. Li, W.; Mu, B.; Yang, Y. Feasibility of Industrial-Scale Treatment of Dye Wastewater via Bio-Adsorption Technology. Bioresour. Technol. 2019, 277, 157–170. [Google Scholar] [CrossRef] [PubMed]
  17. Tholozan, L.V.; Valério Filho, A.; Maron, G.K.; Carreno, N.L.V.; da Rocha, C.M.; Bordin, J.; da Rosa, G.S. Sphagnum Perichaetiale Hampe Biomass as a Novel, Green, and Low-Cost Biosorbent in the Adsorption of Toxic Crystal Violet Dye. Environ. Sci. Pollut. Res. 2023, 30, 52472–52484. [Google Scholar] [CrossRef]
  18. Hu, Z.; Srinivasan, M.P.; Ni, Y. Novel Activation Process for Preparing Highly Microporous and Mesoporous Activated Carbons. Carbon N. Y. 2001, 39, 877–886. [Google Scholar] [CrossRef]
  19. Yorgun, S.; Yıldız, D. Preparation and Characterization of Activated Carbons from Paulownia Wood by Chemical Activation with H3PO4. J. Taiwan Inst. Chem. Eng. 2015, 53, 122–131. [Google Scholar] [CrossRef]
  20. Liu, J.; Yang, X.; Liu, H.; Cheng, W.; Bao, Y. Modification of Calcium-Rich Biochar by Loading Si/Mn Binary Oxide after NaOH Activation and Its Adsorption Mechanisms for Removal of Cu(II) from Aqueous Solution. Colloids Surfaces A Physicochem. Eng. Asp. 2020, 601, 124960. [Google Scholar] [CrossRef]
  21. Linares-Solano, A.; Lillo-Ródenas, M.A.; Marco-Lozar, J.P.; Kunowsky, M.; Romero-Anaya, A.J. NaOH and KOH for Preparing Activated Carbons Used in Energy and Environmental Applications. Int. J. Energy Environ. Econ. 2012, 20, 59–91. [Google Scholar]
  22. Hernandes, P.T.; Oliveira, M.L.S.; Georgin, J.; Franco, D.S.P.; Allasia, D.; Dotto, G.L. Adsorptive Decontamination of Wastewater Containing Methylene Blue Dye Using Golden Trumpet Tree Bark (Handroanthus Albus). Environ. Sci. Pollut. Res. 2019, 26, 31924–31933. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, L.; Gao, Z.Y.; Su, X.P.; Chen, X.; Jiang, L.; Yao, J.M. Adsorption Removal of Dyes from Single and Binary Solutions Using a Cellulose-Based Bioadsorbent. ACS Sustain. Chem. Eng. 2015, 3, 432–442. [Google Scholar] [CrossRef]
  24. León, O.; Muñoz-Bonilla, A.; Soto, D.; Pérez, D.; Rangel, M.; Colina, M.; Fernández-García, M. Removal of Anionic and Cationic Dyes with Bioadsorbent Oxidized Chitosans. Carbohydr. Polym. 2018, 194, 375–383. [Google Scholar] [CrossRef] [PubMed]
  25. Bounaas, M.; Bouguettoucha, A.; Chebli, D.; Gatica, J.M.; Vidal, H. Role of the Wild Carob as Biosorbent and as Precursor of a New High-Surface-Area Activated Carbon for the Adsorption of Methylene Blue. Arab. J. Sci. Eng. 2021, 46, 325–341. [Google Scholar] [CrossRef]
  26. Silva, E.O.; Valério Filho, A.; de Araujo, E.B.; Andrade, T.D.; dos Santos, M.C.; Zottis, R.; da Rosa, G.S.; de Almeida, A.R.F. Application of Lolium multiflorum as an Efficient Raw Material in the Production of Adsorbent for Removal of Methylene Blue. C 2023, 9, 44. [Google Scholar] [CrossRef]
  27. El-Qanni, A.; Nassar, N.N.; Vitale, G.; Hassan, A. Maghemite Nanosorbcats for Methylene Blue Adsorption and Subsequent Catalytic Thermo-Oxidative Decomposition: Computational Modeling and Thermodynamics Studies. J. Colloid Interface Sci. 2016, 461, 396–408. [Google Scholar] [CrossRef]
  28. Russo, V.; Hmoudah, M.; Broccoli, F.; Iesce, M.R.; Serio, M.D.; National, P.; Korea, S.; National, A. Applications of Metal Organic Frameworks in Wastewater Treatment: A Review on Adsorption and Photodegradation. Front. Chem. Eng. 2020, 2, e581487. [Google Scholar] [CrossRef]
  29. Abuhatab, S.; El-qanni, A.; Marei, N.N.; Hmoudah, M.; El-hamouz, A. Blue and Acid Red 88 from Synthetic Wastewater Experimental and Computational Modeling Studies. RSC Adv. 2019, 9, 35483–35498. [Google Scholar] [CrossRef]
  30. Valério Filho, A.; Kulman, R.X.; Janner, N.N.; Tholozan, L.V.; de Almeida, A.R.F.; da Rosa, G.S. Optimization of Cationic Dye Removal Using a High Surface Area-Activated Carbon from Water Treatment Sludge. Bull. Mater. Sci. 2021, 44, 41. [Google Scholar] [CrossRef]
  31. Georgin, J.; da Boit Martinello, K.; Franco, D.S.P.; Netto, M.S.; Piccilli, D.G.A.; Foletto, E.L.; Silva, L.F.O.; Dotto, G.L. Efficient Removal of Naproxen from Aqueous Solution by Highly Porous Activated Carbon Produced from Grapetree (Plinia Cauliflora) Fruit Peels. J. Environ. Chem. Eng. 2021, 9, 106820. [Google Scholar] [CrossRef]
  32. Maroldi, W.V.; Maciel, G.M.; Rossetto, R.; Bortolini, D.G.; Fernandes, I.d.A.A.; Haminiuk, C.W.I. Biosorption of Phenolic Compounds from Plinia Cauliflora Seeds in Residual Yeast: Kinetic, Equilibrium, and Bioaccessibility Studies. J. Food Process. Preserv. 2022, 46, e17156. [Google Scholar] [CrossRef]
  33. da Silva, E.O.; dos Santos, V.D.; de Araujo, E.B.; Guterres, F.P.; Zottis, R.; Flores, W.H.; de Almeida, A.R.F. Removal of Methylene Blue from Aqueous Solution by Ryegrass Straw. Int. J. Environ. Sci. Technol. 2020, 17, 3723–3740. [Google Scholar] [CrossRef]
  34. Georgin, J.; Drumm, F.C.; Grassi, P.; Franco, D.; Allasia, D.; Dotto, G.L. Potential of Araucaria Angustifolia Bark as Adsorbent to Remove Gentian Violet Dye from Aqueous Effluents. Water Sci. Technol. 2018, 78, 1693–1703. [Google Scholar] [CrossRef]
  35. Costa, R.G.; Andreola, K.; de Andrade Mattietto, R.; de Faria, L.J.G.; Taranto, O.P. Effect of Operating Conditions on the Yield and Quality of Açai (Euterpe Oleracea Mart.) Powder Produced in Spouted Bed. LWT Food Sci. Technol. 2015, 64, 1196–1203. [Google Scholar] [CrossRef]
  36. Rovani, S.; Rodrigues, A.G.; Medeiros, L.F.; Cataluña, R.; Lima, É.C.; Fernandes, A.N. Synthesis and Characterisation of Activated Carbon from Agroindustrial Waste—Preliminary Study of 17β-Estradiol Removal from Aqueous Solution. J. Environ. Chem. Eng. 2016, 4, 2128–2137. [Google Scholar] [CrossRef]
  37. Rosas, J.M.; Bedia, J.; Rodríguez-Mirasol, J.; Cordero, T. HEMP-Derived Activated Carbon Fibers by Chemical Activation with Phosphoric Acid. Fuel 2009, 88, 19–26. [Google Scholar] [CrossRef]
  38. Yang, H.; Yan, R.; Chen, H.; Lee, D.H.; Zheng, C. Characteristics of Hemicellulose, Cellulose and Lignin Pyrolysis. Fuel 2007, 86, 1781–1788. [Google Scholar] [CrossRef]
  39. Calvete, T.; Lima, E.C.; Cardoso, N.F.; Vaghetti, J.C.P.; Dias, S.L.P.; Pavan, F.A. Application of Carbon Adsorbents Prepared from Brazilian-Pine Fruit Shell for the Removal of Reactive Orange 16 from Aqueous Solution: Kinetic, Equilibrium, and Thermodynamic Studies. J. Environ. Manage. 2010, 91, 1695–1706. [Google Scholar] [CrossRef]
  40. Calvete, T.; Lima, E.C.; Cardoso, N.F.; Dias, S.L.P.; Ribeiro, E.S. Removal of Brilliant Green Dye from Aqueous Solutions Using Home Made Activated Carbons. Clean Soil Air Water 2010, 38, 521–532. [Google Scholar] [CrossRef]
  41. Lütke, S.F.; Igansi, A.V.; Pegoraro, L.; Dotto, G.L.; Pinto, L.A.A.; Cadaval, T.R.S. Preparation of Activated Carbon from Black Wattle Bark Waste and Its Application for Phenol Adsorption. J. Environ. Chem. Eng. 2019, 7, 103396. [Google Scholar] [CrossRef]
  42. Fontana, K.B.; Chaves, E.S.; Sanchez, J.D.S.; Watanabe, E.R.L.R.; Pietrobelli, J.M.T.A.; Lenzi, G.G. Textile Dye Removal from Aqueous Solutions by Malt Bagasse: Isotherm, Kinetic and Thermodynamic Studies. Ecotoxicol. Environ. Saf. 2016, 124, 329–336. [Google Scholar] [CrossRef] [PubMed]
  43. Ateş, A. The Effect of Microwave and Ultrasound Activation on the Characteristics of Biochar Produced from Tea Waste in the Presence of H3PO4 and KOH. Biomass Convers. Biorefinery 2021, 1–20. [Google Scholar] [CrossRef]
  44. Wang, P.; Tang, Y.; Liu, Y.; Wang, T.; Wu, P.; Lu, X.Y. Halloysite Nanotube@carbon with Rich Carboxyl Groups as a Multifunctional Adsorbent for the Efficient Removal of Cationic Pb(Ii), Anionic Cr(vi) and Methylene Blue (MB). Environ. Sci. Nano 2018, 5, 2257–2268. [Google Scholar] [CrossRef]
  45. Mishra, S.B.; Mishra, A.K. Bio-and Nanosorbents from Natural Resources; Springer: Berlin/Heidelberg, Germany, 2018; ISBN 9783319687070. [Google Scholar]
  46. Franco, D.S.P.; Tanabe, E.H.; Bertuol, D.A.; Dos Reis, G.S.; Lima, É.C.; Dotto, G.L. Alternative Treatments to Improve the Potential of Rice Husk as Adsorbent for Methylene Blue. Water Sci. Technol. 2017, 75, 296–305. [Google Scholar] [CrossRef] [PubMed]
  47. Gao, C.; Xiong, G.Y.; Luo, H.L.; Ren, K.J.; Huang, Y.; Wan, Y.Z. Dynamic Interaction between the Growing Ca-P Minerals and Bacterial Cellulose Nanofibers during Early Biomineralization Process. Cellulose 2010, 17, 365–373. [Google Scholar] [CrossRef]
  48. Rahnama, N.; Mamat, S.; Shah, U.K.M.; Ling, F.H.; Rahman, N.A.A.; Ariff, A.B. Effect of Alkali Pretreatment of Rice Straw on Cellulase and Xylanase Production by Local Trichoderma Harzianum SNRS3 under Solid State Fermentation. BioResources 2013, 8, 2881–2896. [Google Scholar] [CrossRef]
  49. de Azevedo, A.C.N.; Vaz, M.G.; Gomes, R.F.; Pereira, A.G.B.; Fajardo, A.R.; Rodrigues, F.H.A. Starch/Rice Husk Ash Based Superabsorbent Composite: High Methylene Blue Removal Efficiency. Iran. Polym. J. Engl. Ed. 2017, 26, 93–105. [Google Scholar] [CrossRef]
  50. de Souza Macedo, J.; da Costa Júnior, N.B.; Almeida, L.E.; da Silva Vieira, E.F.; Cestari, A.R.; de Fátima Gimenez, I.; Villarreal Carreño, N.L.; Barreto, L.S. Kinetic and Calorimetric Study of the Adsorption of Dyes on Mesoporous Activated Carbon Prepared from Coconut Coir Dust. J. Colloid Interface Sci. 2006, 298, 515–522. [Google Scholar] [CrossRef]
  51. Georgin, J.; Franco, D.S.P.; Grassi, P.; Tonato, D.; Piccilli, D.G.A.; Meili, L.; Dotto, G.L. Potential of Cedrella Fissilis Bark as an Adsorbent for the Removal of Red 97 Dye from Aqueous Effluents. Environ. Sci. Pollut. Res. 2019, 26, 19207–19219. [Google Scholar] [CrossRef]
  52. Georgin, J.; Marques, B.S.; Peres, E.C.; Allasia, D.; Dotto, G.L. Biosorption of Cationic Dyes by Pará Chestnut Husk (Bertholletia Excelsa). Water Sci. Technol. 2018, 77, 1612–1621. [Google Scholar] [CrossRef] [PubMed]
  53. Jain, S.N.; Gogate, P.R. Efficient Removal of Acid Green 25 Dye from Wastewater Using Activated Prunus Dulcis as Biosorbent: Batch and Column Studies. J. Environ. Manag. 2018, 210, 226–238. [Google Scholar] [CrossRef]
  54. Mahmoud, D.K.; Salleh, M.A.M.; Karim, W.A.W.A.; Idris, A.; Abidin, Z.Z. Batch Adsorption of Basic Dye Using Acid Treated Kenaf Fibre Char: Equilibrium, Kinetic and Thermodynamic Studies. Chem. Eng. J. 2012, 181–182, 449–457. [Google Scholar] [CrossRef]
  55. Hu, L.; Guang, C.; Liu, Y.; Su, Z.; Gong, S.; Yao, Y.; Wang, Y. Adsorption Behavior of Dyes from an Aqueous Solution onto Composite Magnetic Lignin Adsorbent. Chemosphere 2020, 246, 125757. [Google Scholar] [CrossRef]
  56. de Salomón, Y.L.O.; Georgin, J.; Franco, D.S.P.; Netto, M.S.; Foletto, E.L.; Allasia, D.; Dotto, G.L. Application of Seed Residues from Anadenanthera Macrocarpa and Cedrela Fissilis as Alternative Adsorbents for Remarkable Removal of Methylene Blue Dye in Aqueous Solutions. Environ. Sci. Pollut. Res. 2021, 28, 2342–2354. [Google Scholar] [CrossRef]
  57. Das Saha, P.; Chakraborty, S.; Chowdhury, S. Batch and Continuous (Fixed-Bed Column) Biosorption of Crystal Violet by Artocarpus Heterophyllus (Jackfruit) Leaf Powder. Colloids Surf. B Biointerfaces 2012, 92, 262–270. [Google Scholar] [CrossRef] [PubMed]
  58. Weng, C.H.; Lin, Y.T.; Tzeng, T.W. Removal of Methylene Blue from Aqueous Solution by Adsorption onto Pineapple Leaf Powder. J. Hazard. Mater. 2009, 170, 417–424. [Google Scholar] [CrossRef] [PubMed]
  59. Royer, B.; Cardoso, N.F.; Lima, E.C.; Vaghetti, J.C.P.; Simon, N.M.; Calvete, T.; Veses, R.C. Applications of Brazilian Pine-Fruit Shell in Natural and Carbonized Forms as Adsorbents to Removal of Methylene Blue from Aqueous Solutions-Kinetic and Equilibrium Study. J. Hazard. Mater. 2009, 164, 1213–1222. [Google Scholar] [CrossRef] [PubMed]
  60. Jawad, A.H.; Waheeb, A.S.; Rashid, R.A.; Nawawi, W.I.; Yousif, E. Equilibrium Isotherms, Kinetics, and Thermodynamics Studies of Methylene Blue Adsorption on Pomegranate (Punica Granatum) Peels as a Natural Low-Cost Biosorbent. Desalin. Water Treat. 2018, 105, 322–331. [Google Scholar] [CrossRef]
  61. Grassi, P.; Drumm, F.C.; Georgin, J.; Franco, D.S.P.; Dotto, G.L.; Foletto, E.L.; Jahn, S.L. Application of Cordia Trichotoma Sawdust as an Effective Biosorbent for Removal of Crystal Violet from Aqueous Solution in Batch System and Fixed-Bed Column. Environ. Sci. Pollut. Res. 2021, 28, 6771–6783. [Google Scholar] [CrossRef]
  62. Al-Rashdi, B.; Tizaoui, C.; Hilal, N. Copper Removal from Aqueous Solutions Using Nano-Scale Diboron Trioxide/Titanium Dioxide (B2O3/TiO2) Adsorbent. Chem. Eng. J. 2012, 183, 294–302. [Google Scholar] [CrossRef]
  63. Vijayaraghavan, K.; Joshi, U.M. Application of Ulva Sp. Biomass for Single and Binary Biosorption of Chromium(III) and Manganese(II) Ions: Equilibrium Modeling. Environ. Prog. Sustain. Energy 2013, 33, 676–680. [Google Scholar]
  64. Jauris, I.M.; Matos, C.F.; Zarbin, A.J.G.; Umpierres, C.S.; Saucier, C.; Lima, E.C.; Fagan, S.B.; Zanella, I.; Machado, F.M. Adsorption of Anti-Inflammatory Nimesulide by Graphene Materials: A Combined Theoretical and Experimental Study. Phys. Chem. Chem. Phys. 2017, 19, 22099–22110. [Google Scholar] [CrossRef] [PubMed]
  65. Saucier, C.; Adebayo, M.A.; Lima, E.C.; Cataluña, R.; Thue, P.S.; Prola, L.D.T.; Puchana-Rosero, M.J.; Machado, F.M.; Pavan, F.A.; Dotto, G.L. Microwave-Assisted Activated Carbon from Cocoa Shell as Adsorbent for Removal of Sodium Diclofenac and Nimesulide from Aqueous Effluents. J. Hazard. Mater. 2015, 289, 18–27. [Google Scholar] [CrossRef] [PubMed]
  66. Qiu, H.; Lv, L.; Pan, B.C.; Zhang, Q.J.; Zhang, W.M.; Zhang, Q.X. Critical Review in Adsorption Kinetic Models. J. Zhejiang Univ. Sci. A 2009, 10, 716–724. [Google Scholar] [CrossRef]
  67. Pathania, D.; Sharma, S.; Singh, P. Removal of Methylene Blue by Adsorption onto Activated Carbon Developed from Ficus Carica Bast. Arab. J. Chem. 2017, 10, S1445–S1451. [Google Scholar] [CrossRef]
  68. McCabe, W.L.; Smith, J.C.; Harriott, P. Unit Operations of Chemical Engineering; McGraw-hill: New York, NY, USA, 1993. [Google Scholar]
  69. Chakraborty, A.; Sun, B. An Adsorption Isotherm Equation for Multi-Types Adsorption with Thermodynamic Correctness. Appl. Therm. Eng. 2014, 72, 190–199. [Google Scholar] [CrossRef]
  70. Saxena, M.; Sharma, N.; Saxena, R. Highly Efficient and Rapid Removal of a Toxic Dye: Adsorption Kinetics, Isotherm, and Mechanism Studies on Functionalized Multiwalled Carbon Nanotubes. Surf. Interfaces 2020, 21, 100639. [Google Scholar] [CrossRef]
  71. Valério Filho, A.; Tholozan, L.V.; Arim, A.L.; de Almeida, A.R.F.; da Rosa, G.S. High-Performance Removal of Anti-Inflammatory Using Activated Carbon from Water Treatment Plant Sludge: Fixed-Bed and Batch Studies. Int. J. Environ. Sci. Technol. 2022, 19, 3633–3644. [Google Scholar] [CrossRef]
  72. 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]
  73. Freundlich, H.M.F. Over the Adsorption in Solution. J. Phys. Chem. 1906, 57, 385–471. [Google Scholar]
  74. Seikova, I.; Mintchev, A. Varying Porosity Kinetic Model for Direct and Reactive Solid-Liquid Extraction. Hungarian J. Ind. Chem. 2002, 30, 95–101. [Google Scholar]
  75. Peláez-Cid, A.A.; Velázquez-Ugalde, I.; Herrera-González, A.M.; García-Serrano, J. Textile Dyes Removal from Aqueous Solution Using Opuntia Ficus-Indica Fruit Waste as Adsorbent and Its Characterization. J. Environ. Manag. 2013, 130, 90–97. [Google Scholar] [CrossRef] [PubMed]
  76. Cha, J.S.; Park, S.H.; Jung, S.C.; Ryu, C.; Jeon, J.K.; Shin, M.C.; Park, Y.K. Production and Utilization of Biochar: A Review. J. Ind. Eng. Chem. 2016, 40, 1–15. [Google Scholar] [CrossRef]
  77. Iwanow, M.; Gärtner, T.; Sieber, V.; König, B. Activated Carbon as Catalyst Support: Precursors, Preparation, Modification and Characterization. Beilstein J. Org. Chem. 2020, 16, 1188–1202. [Google Scholar] [CrossRef]
  78. Adel, M.; Ahmed, M.A.; Mohamed, A.A. Synthesis and Characterization of Magnetically Separable and Recyclable Crumbled MgFe2O4/Reduced Graphene Oxide Nanoparticles for Removal of Methylene Blue Dye from Aqueous Solutions. J. Phys. Chem. Solids 2021, 149, 109760. [Google Scholar] [CrossRef]
  79. Georgin, J.; Franco, D.S.P.; Netto, M.S.; Allasia, D.; Oliveira, M.L.S.; Dotto, G.L. Treatment of Water Containing Methylene by Biosorption Using Brazilian Berry Seeds (Eugenia Uniflora). Environ. Sci. Pollut. Res. 2020, 27, 20831–20843. [Google Scholar] [CrossRef]
  80. Raja, Y.S.; Samsudin, M.F.R.; Sufian, S. Development of the Low-Cost and Green Hibiscus Cannabinus Bioadsorbent for the Removal of Dye in Wastewater. Arab. J. Sci. Eng. 2021, 46, 6349–6358. [Google Scholar] [CrossRef]
  81. Bhatti, H.N.; Akhtar, N.; Saleem, N. Adsorptive Removal of Methylene Blue by Low-Cost Citrus Sinensis Bagasse: Equilibrium, Kinetic and Thermodynamic Characterization. Arab. J. Sci. Eng. 2012, 37, 9–18. [Google Scholar] [CrossRef]
  82. Yagub, M.T.; Sen, T.K.; Ang, M. Removal of Cationic Dye Methylene Blue (MB) from Aqueous Solution by Ground Raw and Base Modified Pine Cone Powder. Environ. Earth Sci. 2014, 71, 1507–1519. [Google Scholar] [CrossRef]
  83. Han, R.; Zou, W.; Yu, W.; Cheng, S.; Wang, Y.; Shi, J. Biosorption of Methylene Blue from Aqueous Solution by Fallen Phoenix Tree’s Leaves. J. Hazard. Mater. 2007, 141, 156–162. [Google Scholar] [CrossRef] [PubMed]
  84. Azzaz, A.A.; Jellali, S.; Akrout, H.; Assadi, A.A.; Bousselmi, L. Optimization of a Cationic Dye Removal by a Chemically Modified Agriculture By-Product Using Response Surface Methodology: Biomasses Characterization and Adsorption Properties. Environ. Sci. Pollut. Res. 2017, 24, 9831–9846. [Google Scholar] [CrossRef]
  85. Vadivelan, V.; Vasanth Kumar, K. Equilibrium, Kinetics, Mechanism, and Process Design for the Sorption of Methylene Blue onto Rice Husk. J. Colloid Interface Sci. 2005, 286, 90–100. [Google Scholar] [CrossRef]
  86. Belala, Z.; Jeguirim, M.; Belhachemi, M.; Addoun, F.; Trouvé, G. Biosorption of Basic Dye from Aqueous Solutions by Date Stones and Palm-Trees Waste: Kinetic, Equilibrium and Thermodynamic Studies. Desalination 2011, 271, 80–87. [Google Scholar] [CrossRef]
  87. Bortoluz, J.; Ferrarini, F.; Bonetto, L.R.; da Silva Crespo, J.; Giovanela, M. Use of Low-Cost Natural Waste from the Furniture Industry for the Removal of Methylene Blue by Adsorption: Isotherms, Kinetics and Thermodynamics. Cellulose 2020, 27, 6445–6466. [Google Scholar] [CrossRef]
  88. Cusioli, L.F.; Quesada, H.B.; Baptista, A.T.A.; Gomes, R.G.; Bergamasco, R. Soybean Hulls as a Low-Cost Biosorbent for Removal of Methylene Blue Contaminant. Environ. Prog. Sustain. Energy 2020, 39, e13328. [Google Scholar]
  89. Valério Filho, A.; Tholozan, L.V.; da Silva, E.O.; Meili, L.; de Almeida, A.R.F.; da Rosa, G.S. Perspectives of the Reuse of Agricultural Wastes from the Rio Grande Do Sul, Brazil, as New Adsorbent Materials. In Biomass-Derived Materials for Environmental Applications; Elsevier: Amsterdam, The Netherlands, 2022; pp. 243–266. ISBN 9788490225370. [Google Scholar]
  90. Largegren, S. About the Theory of So-Called Adsorption of Soluble Substances. K. Suensk Vetensk. Handl. 1898, 24, 1–39. [Google Scholar]
  91. Ho, Y.S.; McKay, G. Pseudo-Second Order Model for Sorption Processes. Process Biochem. 1999, 34, 451–465. [Google Scholar] [CrossRef]
  92. Elovich, S.Y.; Larinov, O.G. Theory of Adsorption from Solutions of Non Electrolytes on Solid (I) Equation Adsorption from Solutions and the Analysis of Its Simplest Form,(II) Verification of the Equation of Adsorption Isotherm from Solutions. Izv. Akad. Nauk. SSSR 1962, 2, 209–216. [Google Scholar]
  93. Weber, W.; Morris, J.C. Kinetics of Adsorption on Carbon from Solutions. J. Sanit. Eng. Div. 1963, 89, 31–60. [Google Scholar] [CrossRef]
  94. Pérez-Marín, A.B.; Zapata, V.M.; Ortuño, J.F.; Aguilar, M.; Sáez, J.; Lloréns, M. Removal of Cadmium from Aqueous Solutions by Adsorption onto Orange Waste. J. Hazard. Mater. 2007, 139, 122–131. [Google Scholar] [CrossRef] [PubMed]
  95. Ncibi, M.C. Applicability of Some Statistical Tools to Predict Optimum Adsorption Isotherm after Linear and Non-Linear Regression Analysis. J. Hazard. Mater. 2008, 153, 207–212. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Post-harvest jaboticaba (a); dried jaboticaba bark (b); jaboticaba extract residue (c); and SEM micrographs for JB (d), JB-H3PO4 (e), and JB-NaOH (f).
Figure 1. Post-harvest jaboticaba (a); dried jaboticaba bark (b); jaboticaba extract residue (c); and SEM micrographs for JB (d), JB-H3PO4 (e), and JB-NaOH (f).
Molecules 28 04066 g001
Figure 2. Characterization JB, JB-H3PO4, and JB-NaOH. Thermogravimetric (a), thermogravimetric derivative (b), FTIR spectra (c), and XRD pattern (d).
Figure 2. Characterization JB, JB-H3PO4, and JB-NaOH. Thermogravimetric (a), thermogravimetric derivative (b), FTIR spectra (c), and XRD pattern (d).
Molecules 28 04066 g002
Figure 3. Response surface and Pareto charts for adsorption capacity and percent removal of JB (a,b), JB-H3PO4 (c,d), and JB-NaOH (e,f).
Figure 3. Response surface and Pareto charts for adsorption capacity and percent removal of JB (a,b), JB-H3PO4 (c,d), and JB-NaOH (e,f).
Molecules 28 04066 g003
Figure 4. Kinetic models for JB (a), JB-H3PO4 (b), and JB-NaOH (c); volume of dye solution = 25 mL; temperature = 25 °C; adsorbent dosage 0.5 mg; pH = 7; agitation speed = 150 rpm.
Figure 4. Kinetic models for JB (a), JB-H3PO4 (b), and JB-NaOH (c); volume of dye solution = 25 mL; temperature = 25 °C; adsorbent dosage 0.5 mg; pH = 7; agitation speed = 150 rpm.
Molecules 28 04066 g004
Figure 5. Results of the equilibrium experiments for JB (a), JB-H3PO4 (b), and JB-NaOH (c) volume of dye solution = 25 mL; temperature = 25 °C; adsorbent dosage 0.5 mg; pH = 7; agitation speed = 150 rpm.
Figure 5. Results of the equilibrium experiments for JB (a), JB-H3PO4 (b), and JB-NaOH (c) volume of dye solution = 25 mL; temperature = 25 °C; adsorbent dosage 0.5 mg; pH = 7; agitation speed = 150 rpm.
Molecules 28 04066 g005
Figure 6. Proposed interaction of MB with the JB adsorbent.
Figure 6. Proposed interaction of MB with the JB adsorbent.
Molecules 28 04066 g006
Table 1. Results of physical characterization of JB, JB-H3PO4, and JB-NaOH samples.
Table 1. Results of physical characterization of JB, JB-H3PO4, and JB-NaOH samples.
ParameterJBJB-H3PO4JB-NaOH
ρb1.5133 ± 0.0012 a1.5118 ± 0.0005 a1.5613 ± 0.0018 a
ρr0.6091 ± 0.0023 a0.6001 ± 0.0056 a0.4774 ± 0.0054 a
ε0.59750.60300.6942
a Mean values in each row with different letters are significantly different (Tukey test, p < 0.05).
Table 2. Experimental results for adsorption capacity and removal percentage of MB.
Table 2. Experimental results for adsorption capacity and removal percentage of MB.
JBJB-H3PO4JB-NaOH
AdpHqRqRqR
(g L−1) (mg g−1)(%)(mg g−1)(%)(mg g−1)(%)
10.57117.72 ± 0.4484.43 ± 1.9969.81 ± 2.1049.89 ± 2.3068.22 ± 1.2149.52 ± 1.27
21.5743.96 ± 0.1894.45 ± 0.3838.50 ± 2.1082.51 ± 0.3632.28 ± 0.1769.17 ± 0.01
30.511125.20 ± 1.1991.21 ± 0.2190.19 ± 6.2664.39 ± 3.9580.30 ± 0.2258.28 ± 1.08
41.51144.32 ± 0.2894.34 ± 0.4745.56 ± 0.2797.63 ± 0.0635.84 ± 0.2476.91 ± 1.03
51.0967.07 ± 0.0195.24 ± 0.2056.44 ± 2.3380.14 ± 3.4745.71 ± 0.5865.69 ± 0.19
61.0967.84 ± 0.2696.72 ± 0.1753.39 ± 4.2276.09 ± 5.5744.42 ± 0.1664.22 ± 0.22
71.0967.32 ± 0.5996.55 ± 0.3150.35 ± 0.6772.65 ± 0.8244.77 ± 0.7164.58 ± 0.12
Table 3. Kinetic parameters for JB, JB-H3PO4, and JB-NaOH.
Table 3. Kinetic parameters for JB, JB-H3PO4, and JB-NaOH.
JBJB-H3PO4JB-NaOH
Pseudo-first orderqe112.1067.5775.86
k10.660.060.71
R20.910.980.98
ARE8.198.863.51
2112.210.019.92
Pseudo-second
order
qe117.6778.3178.24
k20.010.00090.02
R20.960.980.99
ARE6.059.721.33
253.228.741.57
Elovichα24,043.2511.074,056,450
β0.100.060.23
R20.990.970.99
ARE1.587.342.20
24.6116.093.97
Intraparticle
Diffusion
kdi19.546.287.18
C63.3019.4858.52
R20.990.970.99
ARE62.695.8031.56
20.323.450.04
Table 4. Equilibrium parameters for JB, JB-H3PO4, and JB-NaOH.
Table 4. Equilibrium parameters for JB, JB-H3PO4, and JB-NaOH.
JBJB-H3PO4JB-NaOH
Langmuirqmax305.81116.78283.55
kL0.070.030.01
R20.970.770.96
ARE7.1115.9214.13
X2288.61197.48225.78
FreundlichKF75.4724.3918.19
nF4.134.022.41
R20.960.960.98
ARE8.126.547.16
X2355.7930.47105.73
TemkinB48.6818.3553.25
AT1.630.840.13
R20.990.850.95
ARE4.72729.3317.11
X256.7398.91236.81
Table 5. Comparison of adsorption capacity of MB between this work and other related adsorbents.
Table 5. Comparison of adsorption capacity of MB between this work and other related adsorbents.
AdsorbentActivationCo
(mg L−1 )
Contact TimeTemperature (°C)pHAd
(g L−1)
qmax
(mg g−1)
Ref.
JB-30–70030 min2580.5305.81This work
JB-NaOHNaOH30–70030 min2580.5241.10This work
Golden trumpet tree bark-25–30025 min55100.5232.25[22]
Cedrela fissilis-50–30060 min2581.0230.06[56]
Anadenanthera macrocarpa-50–30060 min3581.0227.57[56]
Brazilian berry seeds-25–50120 min5580.8189.6[79]
Hibiscus cannabinusKOH10–5060 min35110.2154[80]
Citrus sinensis-5024 h3571.0147.4[81]
Pine coneNaOH20–80120 min309.170.003142.25[82]
JB-H3PO4H3PO430–70060 min2570.5122.72This work
Fallen phoenix tree’s leaf-30–180180 min2272.080.9[83]
Orange tree sawdustNaOH40–100180 min2061.078.74[84]
Ryegrass strawNaOH15060 min25820.067.19[33]
Rice husk-10–12048 h3280.540.5[85]
Palm tree-70–700240 min706.31.039.5[86]
Table 6. Independent variables and design levels.
Table 6. Independent variables and design levels.
Levels
Factors−101
Ad (g L−1)0.511.5
pH7911
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Janner, N.N.; Tholozan, L.V.; Maron, G.K.; Carreno, N.L.V.; Valério Filho, A.; da Rosa, G.S. Novel Adsorbent Material from Plinia cauliflora for Removal of Cationic Dye from Aqueous Solution. Molecules 2023, 28, 4066. https://doi.org/10.3390/molecules28104066

AMA Style

Janner NN, Tholozan LV, Maron GK, Carreno NLV, Valério Filho A, da Rosa GS. Novel Adsorbent Material from Plinia cauliflora for Removal of Cationic Dye from Aqueous Solution. Molecules. 2023; 28(10):4066. https://doi.org/10.3390/molecules28104066

Chicago/Turabian Style

Janner, Natalia Nara, Luana Vaz Tholozan, Guilherme Kurz Maron, Neftali Lenin Villarreal Carreno, Alaor Valério Filho, and Gabriela Silveira da Rosa. 2023. "Novel Adsorbent Material from Plinia cauliflora for Removal of Cationic Dye from Aqueous Solution" Molecules 28, no. 10: 4066. https://doi.org/10.3390/molecules28104066

Article Metrics

Back to TopTop