Next Article in Journal
Effect of Process Parameters on the Formability, Microstructure, and Mechanical Properties of Laser-Arc Hybrid Welding of Q355B Steel
Next Article in Special Issue
The Effect of Hydrothermal, Microwave, and Mechanochemical Treatments of Tin Phosphate on Sorption of Some Cations
Previous Article in Journal
Thermoelectric Properties of the Corbino Disk in Graphene
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Novel High-Efficiency Natural Biosorbent Material Obtained from Sour Cherry (Prunus cerasus) Leaf Biomass for Cationic Dyes Adsorption

Faculty of Industrial Chemistry and Environmental Engineering, Politehnica University Timisoara, Bd. V. Parvan No. 6, 300223 Timisoara, Romania
*
Authors to whom correspondence should be addressed.
Materials 2023, 16(12), 4252; https://doi.org/10.3390/ma16124252
Submission received: 8 May 2023 / Revised: 4 June 2023 / Accepted: 5 June 2023 / Published: 8 June 2023
(This article belongs to the Special Issue Environmentally Friendly Adsorption Materials)

Abstract

:
The present study aimed to investigate the potential of a new lignocellulosic biosorbent material derived from mature leaves of sour cherry (Prunus cerasus L.) for removing methylene blue and crystal violet dyes from aqueous solutions. The material was first characterized using several specific techniques (SEM, FTIR, color analysis). Then, the adsorption process mechanism was investigated through studies related to adsorption equilibrium, kinetics, and thermodynamics. A desorption study was also performed. Results showed that the Sips isotherm provided the best fit for the adsorption process of both dyes, with a maximum adsorption capacity of 168.6 (mg g−1) for methylene blue and 524.1 (mg g−1) for crystal violet, outperforming the capacity of other similar adsorbents. The contact time needed to reach equilibrium was 40 min for both studied dyes. The Elovich equation is the most suitable model for describing the adsorption of methylene blue, while the general order model is better suited for the adsorption of crystal violet dye. Thermodynamic analyses revealed the adsorption process to be spontaneous, favorable, and exothermic, with physical adsorption involved as the primary mechanism. The obtained results suggest that sour cherry leaves powder can be a highly efficient, eco-friendly, and cost-effective adsorbent for removing methylene blue and crystal violet dyes from aqueous solutions.

1. Introduction

Water is an essential resource for sustaining life on Earth. Industrial development, urbanization, and population growth have led to an increase in the water requirement. Pollution of water sources, underground and surface, has become a global problem that requires special attention [1,2,3,4].
Among the compounds playing a major role in water pollution are organic substances. Of these, dyes generate significant water pollution [1,3,5,6]. Industries that release considerable amounts of colored wastewater into the environment are textiles, pulp and paper, plastic, leather, cosmetics, pharmaceuticals, rubber, food processing, etc. The dyes have a complex aromatic structure, are stable to light, heat and oxidizing agents, presenting toxic mutagenic, teratogenic and carcinogenic effects on living organisms. Therefore, the elimination of these compounds from wastewater is a necessity [1,5,7,8,9,10,11].
Cationic dyes are more toxic compared to anionic and non-ionic ones due to their ability to interact with negatively charged cell membranes, and present a higher risk for human health [1,5]. Nowadays, methylene blue (MB) and crystal violet (CV) are used in numerous industrial activities, having also important human and veterinary medicine applications. However, their presence in natural waters has a negative impact on aquatic life. They can cause various adverse effects on people, such as irritation of the skin and gastrointestinal system, respiratory problems and high blood pressure, cyanosis, and cancer [1,2,6,9,12].
Many methods (physical, chemical, biological) have been used to remove dyes from aqueous solutions. Adsorption is a popular method widely used in this scope due to its high efficiency, selectivity, and flexibility, as well as its simple design, easy operation, and low cost [13,14,15,16,17,18,19,20,21,22,23,24]. The total cost of the adsorption process is largely determined by the adsorbent material, prompting researchers to seek out inexpensive materials derived from or based on industrial and agricultural waste, minerals, and vegetable materials [13,15,25,26,27,28,29,30].
Numerous plant-based wastes and biomasses have been demonstrated to be effective at retaining dyes. The advantage of these materials compared to other adsorbent categories results from the fact that they are easily-available in large quantities every year in most regions of the globe, are cheap, and do not require additional or preliminary treatment or activation [13,15,18,19,20,22,23,27,31,32].
Cellulose, hemicellulose, pectins, and lignin are the major components of plant leaves. These compounds contain many types of functional groups, including carboxyl, hydroxyl, carbonyl, amino, and nitro able to interact with the dyes functional groups [27].
The sour cherry (Prunus cerasus) is a fruit tree belonging to the Rosaceae family that can grow up to a height of 6–10 m. It is native to Europe and Southwest Asia, but it is widely distributed throughout the temperate zones of the globe with well-differentiated seasons. The sour cherry is very well adapted to the winter low temperatures, and the summer drought and high temperatures. Its fruits can be eaten fresh or processed in many forms: juices, jams, compotes, dried fruits, alcoholic beverages, etc. Turkey, the United States, Iran, Italy, Spain, Chile, and Eastern European countries are the main sour cherry growers. Fruits contain significant amounts of sugars, anthocyanin, proteins, mineral salts, vitamins, pectin, and organic acids. They have numerous therapeutic effects, improving the chemical composition of the blood and contributing to delaying the aging process, to curing kidney, diabetic, liver, and cardiovascular diseases. It also has beneficial effects on alleviating mental stress and anemia [33,34,35,36].
Sour cherry leaves are a cost-effective and abundant natural resource that can be found in many areas, making them an excellent option for adsorption that has not yet been reported in the scientific literature. The goal of this study was to show that the adsorbent obtained from this material (without chemical or thermal treatment) can be an effective, economical, and eco-friendly adsorption material for removing methylene blue and crystal violet from aqueous solutions. This was demonstrated by characterizing the materials with FTIR, SEM, and color analysis. Additionally, the effect of various parameters on the adsorption process was analyzed, and studies on equilibrium, kinetics, thermodynamics, and desorption were conducted.

2. Materials and Methods

The Prunus cerasus L. mature leaves were collected from a sour cherry tree located in a private garden in Cerneteaz village, Timis County, Romania. The leaves were washed with distilled water, dried at room temperature for 5 days, and then placed in an air oven at 90 °C for 24 h. The dried leaves were then ground into a fine powder material with an electric mill, passed through a 2 mm sieve, and washed again with distilled water to remove any turbidity and color. The washed powder material was then dried in an air oven at 105 °C for 8 h.
A Shimadzu Prestige-21 FTIR spectrophotometer (Shimadzu, Kyoto, Japan), a Quanta FEG 250 microscope (FEI, Eindhoven, The Netherlands), and a Cary-Varian 300 Bio UV-VIS colorimeter (Varian Inc., Mulgrave, Australia) were used to carry out FTIR (Fourier-transform infrared spectroscopy), SEM (Scanning Electron Microscopy), and color analysis, respectively. For FTIR analysis, the adsorbent sample was mixed with KBr and formed it into a pellet, while the SEM micrograph was taken at 3000× magnification. The color analysis was conducted under D65 (natural light) illumination and with 10 observer angles. The point of zero charge (pHPZC) was identified using the solid addition method [37].
To investigate the adsorption process of each dye, an individual batch system was used. The experiments were carried out in three independent replicates at a constant stirring speed. The pH of the solutions was adjusted with dilute solutions of hydrochloric acid (HCl) and sodium hydroxide (NaOH), both at a concentration of 0.1 (mol dm−3), while the effect of ionic strength was tested by adding sodium chloride (NaCl). Finally, the methylene blue and crystal violet concentrations were measured with a UV-VIS spectrophotometer (Specord 200 PLUS UV-VIS spectrophotometer, Analytik Jena, Jena, Germany), at a wavelength of 664 nm and 590 nm, respectively. Limit of Detection (LOD) and Limit of Quantitation (LOQ) for methylene blue concentration determination were 0.21 (mg L−1) and 0.61 (mg L−1), respectively. For the crystal violet concentration determination, the values for this parameters were LOD = 0.16 (mg L−1) and LOQ = 0.49 (mg L−1).
Five different isotherm models and five kinetic models were used to analyze the equilibrium and kinetics of adsorption. These models and their equations [38,39] are detailed in the Supplementary Materials, Table S1. The suitability of the tested models was evaluated by determining the value of the determination coefficient (R2) and the sum of square error (SSE), chi-square (χ2), and average relative error (ARE) [39]. The equations for these error parameters are described in the Supplementary Materials, Table S2. The experimentally obtained results at temperatures of 283, 297, and 317 K were used to calculate the thermodynamic parameters, whose equations [38] are listed in Table S3 of the Supplementary Material.
The desorption process was conducted using three different substances, distilled water, 0.1 (mol dm−3), HCl and 0.1 (mol dm−3) NaOH, in a batch system with constant stirring for a period of two hours. The desorption efficiency was then calculated using the equation presented in Table S4 in the Supplementary Material.

3. Results and Discussion

3.1. Adsorbent Material Characteriation

Figure 1 presents the FTIR spectra of the sour cherry leaves powder (before adsorption). This spectrum shows specific peaks corresponding to different functional groups (Table 1). Analysis of the spectrum indicates that the primary constituents of the adsorbent are cellulose, hemicellulose, and lignin. This fact highlights its affinity to bind dye molecules [27].
After dye adsorption, only two peaks were shifted as follows: 3282 cm−1 shifted to 3120 cm−1 (methylene blue adsorption) and 3227 cm−1 (crystal violet adsorption), respectively; 1422 cm−1 shifted to 1370 cm−1 at both dye adsorption. These observations suggest that O–H and C–H bonds may be involved in dye retention. The rest of the peaks kept their initial positions and no new ones appeared, indicating no breaking or formation of new bonds after adsorption; therefore, physical adsorption is the main mechanism involved in the process [50,51,52].
The SEM images of the adsorbent material are displayed in Figure 2. Before adsorption, the adsorbent surface appears to be irregular and complex, with pores, crevices, and empty spaces of various sizes and shapes that suggest it is suitable for capturing dyes. After the adsorption process, the adsorbent surface became more uniform, smoother, and consistent, which indicates that the dye molecules filled up the pores and covered up the surface irregularities (Figure 2B,C).
The adsorption process can be characterized by analyzing the initial and final color of the adsorbent using the CIELab* color parameters. During the adsorption process, the color of the dye in the solution is transferred to the sour cherry leaves powder (Figure 3). This causes the luminosity of the adsorbent to decrease and the color parameters a* and b* to change. Point (1), which describes the initial color of the sour cherry leaves, becomes point (4) after adsorption and shifts into the color area of methylene blue, which was initially represented by point (2). The same observation can be made for the absorption of crystal violet dyes: point (1) becomes point (5) after adsorption and shifts into the color area of crystal violet, which was initially represented by point (3).
The point of zero charge (pHPZC) is a measure of the adsorbent surface charge. When the pH is below the pHPZC, the surface of the adsorbent becomes positively charged, and when the pH is above the pHPZC, the surface becomes negatively charged. The surface charge affects the adsorption of cationic dyes, as a negatively charged surface is more favorable for adsorption [14,23]. According to Figure 4, the pHPZC of the sour cherry leaves powder was determined to be 5.5, meaning that a pH above this value is suitable for the adsorption of methylene blue and violet crystal dyes.

3.2. Effect of pH, Ionic Strength, and Adsorbent Dose on Cationic Dyes Adsorption

The pH, ionic strength, and adsorbent dose are parameters that significantly influence the dye’s adsorption process. Figure 5 illustrates the effect of these parameters on methylene blue and crystal violet adsorption on sour cherry leaves powder.
As expected, the adsorption capacity was positively influenced when the solutions pH were higher than pHPZC, the electrostatic attraction between the cationic dye molecules and the negatively charged adsorbent surface favoring the adsorption process. Similar results were recorded for methylene blue adsorption on pineapple leaf powder [46], citrus limetta peel [13], and lotus leaf powder [53], and for crystal violet dye adsorption on Ananas comosus leaves [20], Ocotea puberula bark [54], and Terminalia arjuna sawdust [14].
The presence of other ions in the dyeing wastewater can have a negative effect on the adsorption process. As illustrated in Figure 5, when the ionic strength is increased, due to the addition of NaCl, the adsorption capacity decreases because the sodium ions are competing with the dye cations for the available adsorption sites on the material surface. A similar effect of ionic strength on the methylene blue and crystal violet adsorption was observed in other studies in which similar adsorbents were used, such as: Daucus carota leaves [37], phoenix tree’s leaves [55], potato leaves [56], Ananas comosus leaves [46], lotus leaves [53], Arundo donax L. [57], and Artocarpus odoratissimus leaf-based cellulose [48].
The data in Figure 5 show that higher adsorbent dosages lead to an increase in the adsorption efficiency, based on a larger adsorption surface area and a higher number of active adsorption sites. The decrease in the adsorption capacity is probably due to the fact that many of these sites remain unsaturated and also to the agglomeration of adsorbent material particles [13,55,58,59]. Other researchers previously observed that the amount of adsorbent used had the same effect on the adsorption capacity and removal efficiency of methylene blue and crystal violet [13,14,23,53,54,55].

3.3. Equilibrum Study

The equilibrium adsorption process was evaluated using the non-linear isotherms Langmuir, Freundlich, Temkin, Sips, and Redlich–Peterson. After analyzing the fitted isotherm curves (Figure 6 and Figure 7) and the corresponding error parameters (Table 2), it was found that the applicability of the five isotherms for the obtained experimental data follows the order: Sips > Redlich–Peterson > Langmuir > Freundlich > Temkin for the methylene blue adsorption. For crystal violet adsorption, the order of applicability is slightly modified: Sips > Freundlich > Redlich–Peterson > Langmuir > Temkin.
Previous studies showed that the Sips isotherm best characterized the adsorption process of methylene blue on Maclura pomifera biomass [60], bilberry leaves [61], raspberry leaves [62], dicarboxymethyl cellulose [63], and the adsorption process of crystal violet dye on Artocarpus altilis skin [64], Eragrostis plana Nees [65], and motherwort biomass [42].
Table 3 presents a comparison of the maximum absorption capacities of various similar absorbents used for the absorption of methylene blue and crystal violet dyes from aqueous solutions. Analyzing the presented data, it can be seen that the sour cherry leaves powder has a superior adsorption capacity compared to many other similar adsorbents, indicating the practical utility of the new adsorbent proposed in this study.

3.4. Kinetic Study

The effect of contact time on adsorption capacity for methylene blue and crystal violet retention using sour cherry powder as adsorbent material is shown in Figure 8 and Figure 9. During the first 5–10 min of the adsorption process, the capacity of the adsorbent to retain the dyes increased at a rapid rate. As the contact time increased, active adsorption sites gradually filled up, resulting in a slower increase in adsorption capacity. Finally, after 40 min, an equilibrium was reached in which the amount of dye absorbed had stabilized. This suggests that dye diffusion occurred in the pores of the adsorbent and that a monolayer of dye was formed on its surface, resulting in a decrease in the adsorption rate [23,53,86], therefore, the value of the adsorption capacity remained constant.
Table 4 shows comparatively the time taken to reach equilibrium during the adsorption of methylene blue and crystal violet on various similar adsorbents obtained from plant biomass.
The kinetic data for both dyes adsorption were modeled using five different nonlinear kinetic models. Analyzing these models plots (Figure 8 and Figure 9), the constants, and their corresponding error functions (Table 5), it is concluded that the Elovich model is the most appropriate to describe the methylene blue adsorption, while for the adsorption of the crystal violet dye, a more suitable model is the general order. The coefficient of determination (R2) values for some tested kinetic models were very similar, however, the lower values for χ2, SSE, and ARE are the main arguments that ultimately led to the final conclusion.

3.5. Thermodynamic Study

The thermodynamic parameters, calculated from the experimental results obtained at temperatures of 283, 297, and 317 K, are depicted in Table 6. These parameters indicate that the process is spontaneous, favorable, and exothermic, as evidenced by the negative values of the standard Gibbs energy change (ΔG0) and the standard enthalpy change (ΔH0). Similar results were obtained by other researchers who studied the adsorption of methylene blue on Salix babylonica leaves [23], Daucus carota leaves [37], potato leaves [56], Maclura Pomifera biomass [60], and Typha angustifolia (L.) leaves [68] and, respectively, the adsorption of crystal violet on pineapple leaf [21], Ocotea puberula bark powder [54], Arundo donax L. [57], Moringa oleifera pod husk [78], and jackfruit leaf powder [80].
The positive value of the standard entropy change (ΔS0) suggests that there is an increased randomness at the solid-liquid interface [13,53,69]. The values of ΔG0, both for the adsorption of methyl blue and crystal violet, fall within the range −20 to 0 (kJ mol–1). In addition, the ΔH0 value is less than 40 (kJ mol–1). These two observations indicate that the primary mechanism involved in the absorption is physisorption [23,31,87]. The value of ΔH0 lower than 20 (kJ mol−1) indicates that van der Waals forces are implied and have an important role in the physical adsorption process [52,88,89].

3.6. Desorption Study

The data obtained in this study are illustrated in Figure 10. The highest methylene blue desorption efficiency was obtained when HCl was used as desorption agent (Figure 10A). The regenerated adsorbent was reused for methylene blue adsorption, but the obtained adsorption capacity was approximately 50% lower. In conclusion, it can be stated that the regeneration of the adsorbent material is not justified from both technical and economic point of view.
The desorption efficiency of the crystal violet dye was less than 20% regardless of the desorption agent used (Figure 10B). In this case, the regeneration of the exhausted absorbent cannot be considered as feasible.
The fact that sour cherry leaves are a low-cost and readily available material in large quantities in nature compensate for this disadvantage. Furthermore, due to the combustion properties of plant leaves, the incineration of the exhausted adsorbent can be a simple and efficient reuse solution. Another possible use is as a foaming agent to produce ceramic or glass foams. During the combustion process, a large amount of gas results, which makes it an ideal sporogenous precursor for this type of materials [61,62].

4. Conclusions

This study proposes a new natural adsorbent material, obtained from mature sour cherry (Prunus cerasus L.) leaves, suitable for methylene blue and crystal violet dyes removal from aqueous solutions. This material was characterized and then subjected to adsorption experiments to evaluate its effectiveness in dye removal. The FTIR analysis shows that the adsorbent contains different functional groups specific for cellulose, hemicellulose, and lignin, able to bind dyes. The structure of the adsorbent surface was studied using SEM images, both before and after adsorption, highlighting the importance of the adsorbent porous structure. The dyes retention was indicated using color analysis, dyes color being transferred from the initial solution to the sour cherry leaves powder. pH, ionic strength, and adsorbent dose on cationic dyes adsorption were identified as key factors influencing the effectiveness of the adsorbent. The Sips isotherm best describes the adsorption processes for both studied dyes, with a maximum adsorption capacity of 168.6 (mg g−1) for methylene blue adsorption and 524.1 (mg g−1) for crystal violet adsorption, superior to other similar adsorbents. The contact time needed to reach equilibrium was 40 min for both studied dyes. The Elovich model is the most appropriate to describe the methylene blue adsorption, while for the adsorption of the crystal violet dye more suitable model is general order. Thermodynamic analyses reveal a spontaneous, favorable, and exothermic process, the calculated values for ΔG0 and ΔH0 suggesting physisorption as the primary mechanism involved in the absorption process for both dyes.
Regenerating the absorbent is not a viable option, but this fact is compensated by its very low price.
All results indicate sour cherry leaves powder as an affordable, readily available, environmentally friendly, and efficient adsorbent to remove cationic dyes from aqueous solutions.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ma16124252/s1, Table S1: The non-linear equations of the adsorption isotherms and kinetic models used to assess the adsorption process, Table S2: The calculation equations for error parameters R2, χ2, SSE, and ARE, Table S3: The calculation equations of thermodynamic parameters, Table S4: The calculation equation of the desorption efficiency.

Author Contributions

Conceptualization, G.M., C.V., S.P. and M.D.; methodology, G.M. and M.D.; software, G.M. and C.V.; validation, G.M. and M.D.; formal analysis, G.M. and S.P.; investigation, G.M., S.P., M.D. and S.B.; resources, G.M. and M.D.; data curation, G.M.; writing—original draft preparation, G.M., S.P., C.V. and S.B.; writing—review and editing, G.M., C.V., S.P. and M.D.; visualization, G.M.; supervision, G.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All the experimental data obtained are presented, in the form of table and/or figure, in the article and in the Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Algarni, T.S.; Al-Mohaimeed, A.M.; Al-Odayni, A.-B.; Abduh, N.A.Y. Activated Carbon/ZnFe2O4 Nanocomposite Adsorbent for Efficient Removal of Crystal Violet Cationic Dye from Aqueous Solutions. Nanomaterials 2022, 12, 3224. [Google Scholar] [CrossRef] [PubMed]
  2. Birniwa, A.H.; Mahmud, H.N.M.E.; Abdullahi, S.S.; Habibu, S.; Jagaba, A.H.; Ibrahim, M.N.M.; Ahmad, A.; Alshammari, M.B.; Parveen, T.; Umar, K. Adsorption Behavior of Methylene Blue Cationic Dye in Aqueous Solution Using Polypyrrole-Polyethylenimine Nano-Adsorbent. Polymers 2022, 14, 3362. [Google Scholar] [CrossRef] [PubMed]
  3. Block, I.; Günter, C.; Duarte Rodrigues, A.; Paasch, S.; Hesemann, P.; Taubert, A. Carbon Adsorbents from Spent Coffee for Removal of Methylene Blue and Methyl Orange from Water. Materials 2021, 14, 3996. [Google Scholar] [CrossRef]
  4. Marciniak, M.; Goscianska, J.; Norman, M.; Jesionowski, T.; Bazan-Wozniak, A.; Pietrzak, R. Equilibrium, Kinetic, and Thermodynamic Studies on Adsorption of Rhodamine B from Aqueous Solutions Using Oxidized Mesoporous Carbons. Materials 2022, 15, 5573. [Google Scholar] [CrossRef] [PubMed]
  5. Diaz-Uribe, C.; Angulo, B.; Patiño, K.; Hernández, V.; Vallejo, W.; Gallego-Cartagena, E.; Romero Bohórquez, A.R.; Zarate, X.; Schott, E. Cyanobacterial Biomass as a Potential Biosorbent for the Removal of Recalcitrant Dyes from Water. Water 2021, 13, 3176. [Google Scholar] [CrossRef]
  6. Xie, L.-Q.; Jiang, X.-Y.; Yu, J.-G. A Novel Low-Cost Bio-Sorbent Prepared from Crisp Persimmon Peel by Low-Temperature Pyrolysis for Adsorption of Organic Dyes. Molecules 2022, 27, 5160. [Google Scholar] [CrossRef]
  7. Ouakouak, A.; Abdelhamid, M.; Thouraya, B.; Chahinez, H.-O.; Hocine, G.; Hamdi, N.; Syafiuddin, A.; Boopathy, R. Development of a Novel Adsorbent Prepared from Dredging Sediment for Effective Removal of Dye in Aqueous Solutions. Appl. Sci. 2021, 11, 10722. [Google Scholar] [CrossRef]
  8. Soldatkina, L.; Yanar, M. Equilibrium, Kinetic, and Thermodynamic Studies of Cationic Dyes Adsorption on Corn Stalks Modified by Citric Acid. Colloids Interfaces 2021, 5, 52. [Google Scholar] [CrossRef]
  9. Akbari, M.; Jafari, H.; Rostami, M.; Mahdavinia, G.R.; Sobhani Nasab, A.; Tsurkan, D.; Petrenko, I.; Ganjali, M.R.; Rahimi-Nasrabadi, M.; Ehrlich, H. Adsorption of Cationic Dyes on a Magnetic 3D Spongin Scaffold with Nano-Sized Fe3O4 Cores. Mar. Drugs 2021, 19, 512. [Google Scholar] [CrossRef]
  10. Hachem, C.; Bocquillon, F.; Zahraa, O.; Bouchy, M. Decolorization of Textile Industry Wastewater by Photo Catalytic Degradation Process. Dyes Pigm. 2001, 49, 117–125. [Google Scholar] [CrossRef]
  11. Khamparia, S.; Jaspal, D. Technologies for Treatment of Colored Wastewater from Different Industries. In Handbook of Environmental Materials Management; Hussain, C., Ed.; Springer: Cham, Switzerland, 2018; pp. 1–11. [Google Scholar]
  12. Ahmad, A.; Jamil, S.N.A.M.; Choong, T.S.Y.; Abdullah, A.H.; Faujan, N.H.; Adeyi, A.A.; Daik, R.; Othman, N. Removal of Cationic Dyes by Iron Modified Silica/Polyurethane Composite: Kinetic, Isotherm and Thermodynamic Analyses, and Regeneration via Advanced Oxidation Process. Polymers 2022, 14, 5416. [Google Scholar] [CrossRef]
  13. Shakoor, S.; Nasar, A. Removal of methylene blue dye from artificially contaminated water using citrus limetta peel waste as a very low cost adsorbent. J. Taiwan Inst. Chem. Eng. 2016, 66, 154–163. [Google Scholar] [CrossRef]
  14. Shakoor, S.; Nasar, A. Adsorptive decontamination of synthetic wastewater containing crystal violet dye by employing Terminalia arjuna sawdust waste. Groundw. Sustain. Dev. 2018, 7, 30–38. [Google Scholar] [CrossRef]
  15. Sharma, K.; Sharma, S.; Sharma, V.; Mishra, P.K.; Ekielski, A.; Sharma, V.; Kumar, V. Methylene blue dye adsorption from wastewater using hydroxyapatite/gold nanocomposite: Kinetic and thermodynamics studies. Nanomaterials 2021, 11, 1403. [Google Scholar] [CrossRef] [PubMed]
  16. El-Sayed, G.O. Removal of methylene blue and crystal violet from aqueous solutions by palm kernel fiber. Desalination 2011, 272, 225–232. [Google Scholar] [CrossRef]
  17. Alvarez-Torrellas, S.; Boutahala, M.; Boukhalfa, N.; Munoz, M. Effective Adsorption of Methylene Blue dye onto Magnetic Nanocomposites. Modeling and Reuse Studies. Appl. Sci. 2019, 9, 4563. [Google Scholar] [CrossRef] [Green Version]
  18. 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] [Green Version]
  19. Kadhom, M.; Albayati, N.; Alalwan, H.; Al-Furaiji, M. Removal of dyes by agricultural waste. Sustain. Chem. Pharm. 2020, 16, 100259. [Google Scholar] [CrossRef]
  20. Chakraborty, S.; Chowdhury, S.; Das, P. Insight into biosorption equilibrium, kinetics and thermodynamics of crystal violet onto Ananas comosus (pineapple) leaf powder. Appl. Water Sci. 2012, 2, 135–141. [Google Scholar] [CrossRef] [Green Version]
  21. Aysu, T.; Küçük, M.M. Removal of crystal violet and methylene blue from aqueous solutions by activated carbon prepared from Ferula orientalis. Int. J. Environ. Sci. Technol. 2014, 12, 2273–2284. [Google Scholar] [CrossRef] [Green Version]
  22. Pang, X.; Sellaoui, L.; Franco, D.; Netto, M.S.; Georgin, J.; Dotto, G.L.; Abu Shayeb, M.K.; Belmabrouk, H.; Bonilla-Petriciolet, A.; Li, Z. Preparation and characterization of a novel mountain soursop seeds powder adsorbent and its application for the removal of crystal violet and methylene blue from aqueous solutions. Chem. Eng. J. 2020, 391, 123617. [Google Scholar] [CrossRef]
  23. Khodabandehloo, A.; Rahbar-Kelishami, A.; Shayesteh, H. Methylene blue removal using Salix babylonica (Weeping willow) leaves powder as a low-cost biosorbent in batch mode: Kinetic, equilibrium, and thermodynamic studies. J. Mol. Liq. 2017, 244, 540–548. [Google Scholar] [CrossRef]
  24. Mrkajic, N.S.; Hama, J.R.; Strobel, B.W.; Hansen, H.C.B.; Rasmussen, L.H.; Pedersen, A.K.; Christensen, S.C.B.; Hedegaard, M.J. Removal of phytotoxins in filter sand used for drinking water treatment. Water Res. 2021, 205, 117610. [Google Scholar] [CrossRef] [PubMed]
  25. Aljar, M.A.A.; Rashdan, S.; Abd El-Fattah, A. Environmentally Friendly Polyvinyl Alcohol−Alginate/Bentonite Semi-Interpenetrating Polymer Network Nanocomposite Hydrogel Beads as an Efficient Adsorbent for the Removal of Methylene Blue from Aqueous Solution. Polymers 2021, 13, 4000. [Google Scholar] [CrossRef] [PubMed]
  26. Al-Gorair, A.S.; Sayed, A.; Mahmoud, G.A. Engineered Superabsorbent Nanocomposite Reinforced with Cellulose Nanocrystals for Remediation of Basic Dyes: Isotherm, Kinetic, and Thermodynamic Studies. Polymers 2022, 14, 567. [Google Scholar] [CrossRef]
  27. Bulgariu, L.; Escudero, L.B.; Bello, O.S.; Iqbal, M.; Nisar, J.; Adegoke, K.A.; Alakhras, F.; Kornaros, M.; Anastopoulos, I. The utilization of leaf-based adsorbents for dyes removal: A review. J. Mol. Liq. 2019, 276, 728–747. [Google Scholar] [CrossRef] [Green Version]
  28. Rahmi, R.; Lelifajri, L.; Iqbal, M.; Fathurrahmi, F.; Jalaluddin, J.; Sembiring, R.; Farida, M.; Iqhrammullah, M. Preparation, Characterization and Adsorption Study of PEDGE-Cross-linked Magnetic Chitosan (PEDGE-MCh) Microspheres for Cd2+ Removal. Arab. J. Sci. Eng. 2023, 48, 159–167. [Google Scholar] [CrossRef]
  29. Marlina; Iqhrammullah, M.; Saleha, S.; Fathurrahmi; Maulina, F.P.; Idroes, R. Polyurethane film prepared from ball-milled algal polyol particle and activated carbon filler for NH3–N removal. Heliyon 2020, 6, e04590. [Google Scholar] [CrossRef]
  30. Julinawati, J.; Febriani, F.; Mustafa, I.; Fathurrahmi, F.; Rahmi, R.; Sheilatina, S.; Ahmad, K.; Puspita, K.; Iqhrammullah, M. Tryptophan-Based Organoclay for Aqueous Naphthol Blue Black Removal—Preparation, Characterization, and Batch Adsorption Studies. J. Ecol. Eng. 2023, 24, 274–284. [Google Scholar] [CrossRef]
  31. Loulidi, I.; Boukhlifi, F.; Ouchabi, M.; Amar, A.; Jabri, M.; Kali, A.; Chraibi, S.; Hadey, C.; Aziz, F. Adsorption of Crystal Violet onto an Agricultural Waste Residue: Kinetics, Isotherm, Thermodynamics, and Mechanism of Adsorption. Sci. World. J. 2020, 2020, 5873521. [Google Scholar] [CrossRef]
  32. Hamad, H.N.; Idrus, S. Recent Developments in the Application of Bio-Waste-Derived Adsorbents for the Removal of Methylene Blue from Wastewater: A Review. Polymers 2022, 14, 783. [Google Scholar] [CrossRef]
  33. Butu, M.; Rodino, S. Fruit and Vegetable-Based Beverages—Nutritional Properties and Health Benefits. In Natural Beverages, 1st ed.; Grumezescu, A.M., Holban, A.M., Eds.; Academic Press: Amsterdam, The Netherlands, 2019; Volume 13, pp. 303–338. [Google Scholar]
  34. Kirakosyan, A.; Seymour, E.M.; Kaufman, P.B.; Bolling, S.F. Tart cherry fruits: Implications for human health. In Bioactive Food as Dietary Interventions for Arthritis and Related Inflammatory Diseases: Bioactive Food in Chronic; Watson, R.R., Preedy, V.R., Eds.; Academic Press: Cambridge, MA, USA, 2013; pp. 473–484. [Google Scholar]
  35. Albuquerque, T.G.; Silva, M.A.; Oliveira, M.B.P.P.; Costa, M.B.P.P.; Oliveira, H.S. Analysis, Identification, and Quantification of Anthocyanins. In Fruit Juices; Rajauria, G., Tiwari, B.K., Eds.; Academic Press: San Diego, CA, USA, 2018; Volume 8, pp. 693–737. [Google Scholar]
  36. Mitre, V.; Mitre, I. Lean fructification phenomenon in sour-cherry. Agricultura 2007, 61, 19. [Google Scholar]
  37. Kushwaha, A.K.; Gupta, N.; Chattopadhyaya, M.C. Removal of cationic methylene blue and malachite green dyes from aqueous solution by waste materials of Daucus carota. J. Saudi Chem. Soc. 2014, 18, 200–207. [Google Scholar] [CrossRef]
  38. Piccin, J.S.; Cadaval, T.R.S.; de Pinto, L.A.A.; Dotto, G.L. Adsorption Isotherms in Liquid Phase: Experimental, Modeling, and Interpretations. In Adsorption Processes for Water Treatment and Purification; Bonilla-Petriciolet, A., Mendoza-Castillo, D., Reynel-Avila, H., Eds.; Springer: Cham, Switzerland, 2017; pp. 19–51. [Google Scholar]
  39. Dotto, G.L.; Salau, N.P.G.; Piccin, J.S.; Cadaval, T.R.S.; de Pinto, L.A.A. Adsorption Kinetics in Liquid Phase: Modeling for Discontinuous and Continuous Systems. In Adsorption Processes for Water Treatment and Purification; Bonilla-Petriciolet, A., Mendoza-Castillo, D., Reynel-Avila, H., Eds.; Springer: Cham, Switzerland, 2017; pp. 53–76. [Google Scholar]
  40. Triki, A.; Dittmer, J.; Hassen, M.B.; Arous, M.; Bulou, A.; Gargourie, M. Spectroscopy Analyses of Hybrid Unsaturated Polyester Composite Reinforced by Alfa, Wool, and Thermo-Binder Fibres. Polym. Sci. Ser. A 2016, 58, 255–264. [Google Scholar] [CrossRef]
  41. Peng, H. Physicochemical characterization of hemicelluloses from bamboo (Phyllostachys pubescens Mazel) stem. Ind. Crop. Prod. 2012, 37, 41–50. [Google Scholar] [CrossRef]
  42. Mosoarca, G.; Vancea, C.; Popa, S.; Dan, M.; Boran, S. Crystal Violet Adsorption on Eco-Friendly Lignocellulosic Material Obtained from Motherwort (Leonurus cardiaca L.) Biomass. Polymers 2022, 14, 3825. [Google Scholar] [CrossRef] [PubMed]
  43. Özgenç, Ö.; Durmaz, S.; Kuştaş, S. Chemical Analysis of Tree Barks using ATR-FTIR Spectroscopy and Conventional Techniques. Bioresources 2017, 12, 9143–9151. [Google Scholar] [CrossRef]
  44. Chang, H.T.; Yeh, T.F.; Hsu, F.L.; Kuo-Huang, L.L.; Lee, C.M.; Huang, Y.S.; Chang, S.T. Profiling the chemical composition and growth strain of giant bamboo (Dendrocalamus giganteus Munro). Bioresources 2015, 10, 1260–1270. [Google Scholar] [CrossRef] [Green Version]
  45. Sahoo, S.; Seydibeyoglu, M.O.; Mohanty, A.K.; Misra, M. Characterization of industrial lignins for their utilization in future value added applications. Biomass. Bioenergy 2011, 35, 4230–4237. [Google Scholar] [CrossRef]
  46. 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]
  47. Saravanakumar, S.S.; Kumaravel, A.; Nagarajan, T.; Moorthy, I.G. Effect of chemical treatments on physicochemical properties of Prosopis juliflora fibers. Int. J. Polym. Anal. Charact. 2014, 19, 383–390. [Google Scholar] [CrossRef]
  48. Zaidi, N.A.H.M.; Lim, L.B.L.; Usman, A. Artocarpus odoratissimus leaf-based cellulose as adsorbent for removal of methyl violet and crystal violet dyes from aqueous solution. Cellulose 2018, 25, 3037–3049. [Google Scholar] [CrossRef]
  49. Meisam, T.M. Biosorption of lanthanum and cerium from aqueous solutions using tangerine (Citrus reticulata) peel: Equilibrium, kinetic, and thermodynamic studies. Chem. Ind. Chem. Eng. Q 2013, 19, 79–88. [Google Scholar]
  50. Silva, V.C.; Araújo, M.E.B.; Rodrigues, A.M.; Vitorino, M.d.B.C.; Cartaxo, J.M.; Menezes, R.R.; Neves, G.A. Adsorption Behavior of Crystal Violet and Congo Red Dyes on Heat-Treated Brazilian Palygorskite: Kinetic, Isothermal and Thermodynamic Studies. Materials 2021, 14, 5688. [Google Scholar] [CrossRef]
  51. Hussin, Z.M.; Talib, N.; Hussin, N.M.; Hanafiah, M.A.K.M.; Khalir, W.K.A.W.M. Methylene Blue Adsorption onto NaOH Modified Durian Leaf Powder: Isotherm and Kinetic Studies. Am. J. Environ. Eng. 2015, 5, 38–43. [Google Scholar]
  52. Mosoarca, G.; Vancea, C.; Popa, S.; Boran, S. Bathurst Burr (Xanthium spinosum) Powder—A New Natural Effective Adsorbent for Crystal Violet Dye Removal from Synthetic Wastewaters. Materials 2021, 14, 5861. [Google Scholar] [CrossRef] [PubMed]
  53. Han, X.; Wang, W.; Ma, X. Adsorption characteristics of methylene blue onto low cost biomass material lotus leaf. Chem. Eng. J. 2011, 171, 1–8. [Google Scholar] [CrossRef]
  54. Georgin, J.; Franco, D.S.P.; Netto, M.S.; Allasia, D.; Oliveira, M.L.S.; Dotto, G.L. Evaluation of Ocotea puberula bark powder (OPBP) as an effective adsorbent to uptake crystal violet from colored effluents: Alternative kinetic approaches. Environ. Sci. Pollut. Res. 2020, 27, 25727–25739. [Google Scholar] [CrossRef]
  55. 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]
  56. Gupta, N.; Kushwaha, A.K.; Chattopadhyaya, M.C. Application of potato (Solanum tuberosum) plant wastes for the removal of methylene blue and malachite green dye from aqueous solution. Arab. J. Chem. 2016, 9, S707–S716. [Google Scholar] [CrossRef] [Green Version]
  57. Krika, F.; Krika, A.; Azizi, A. Arundo donax L. as a low cost and promising biosorbent for the removal of crystal violet from aqueous media: Kinetic, isotherm and thermodynamic investigations. Chem. Rev. Lett. 2019, 2, 59–68. [Google Scholar]
  58. Chowdhury, S.; Chakraborty, S.; Das, P. Removal of Crystal Violet from Aqueous Solution by Adsorption onto Eggshells: Equilibrium, Kinetics, Thermodynamics and Artificial Neural Network Modeling. Waste Biomass Valorization 2012, 4, 655–664. [Google Scholar] [CrossRef]
  59. Grassi, P.; Reis, C.; Drumm, F.C.; Georgin, J.; Tonato, D.; Escudero, L.B.; Kuhn, R.; Jahn, S.L.; Dotto, G.L. Biosorption of crystal violet dye using inactive biomass of the fungus Diaporthe Schini. Water Sci. Technol. 2019, 79, 709–717. [Google Scholar] [CrossRef] [PubMed]
  60. Bounaas, M.; Bouguettoucha, A.; Chebli, D.; Reffas, A.; Gatica, J.M.; Amrane, A. Batch adsorption of synthetic dye by Maclura Pomifera, a new eco-friendly waste biomass: Experimental studies and modeling. Int. J. Chem. React. Eng. 2019, 17, 20180063. [Google Scholar] [CrossRef]
  61. Mosoarca, G.; Vancea, C.; Popa, S.; Dan, M.; Boran, S. The Use of Bilberry Leaves (Vaccinium myrtillus L.) as an Efficient Adsorbent for Cationic Dye Removal from Aqueous Solutions. Polymers 2022, 14, 978. [Google Scholar] [CrossRef]
  62. Mosoarca, G.; Popa, S.; Vancea, C.; Dan, M.; Boran, S. Removal of Methylene Blue from Aqueous Solutions Using a New Natural Lignocellulosic Adsorbent—Raspberry (Rubus idaeus) Leaves Powder. Polymers 2022, 14, 1966. [Google Scholar] [CrossRef] [PubMed]
  63. Gago, D.; Chagas, R.; Ferreira, L.M.; Velizarov, S.; Coelhoso, I. A novel cellulose-based polymer for efficient removal of methylene blue. Membranes 2020, 10, 13. [Google Scholar] [CrossRef] [Green Version]
  64. Lim, L.B.L.; Priyantha, N.; Mansor, N.H.M. Artocarpus altilis (breadfruit) skin as a potential low-cost biosorbent for the removal of crystal violet dye: Equilibrium, thermodynamics and kinetics studies. Environ. Earth Sci. 2014, 73, 3239–3247. [Google Scholar] [CrossRef]
  65. Filho, A.C.D.; Mazzocato, A.C.; Dotto, G.L.; Thue, P.S.; Pavan, F.A. Eragrostis plana Nees as a novel eco-friendly adsorbent for removal of crystal violet from aqueous solutions. Environ. Sci. Pollut. Res. 2017, 24, 19909–19919. [Google Scholar] [CrossRef]
  66. Guo, D.; Li, Y.; Cui, B.; Hu, M.; Luo, S.; Ji, B.; Liu, Y. Natural adsorption of methylene blue by waste fallen leaves of Magnoliaceae and its repeated thermal regeneration for reuse. J. Clean. Prod. 2020, 267, 121903. [Google Scholar] [CrossRef]
  67. Krishni, R.R.; Foo, K.Y.; Hameed, B.H. Adsorptive removal of methylene blue using the natural adsorbent-banana leaves. Desalin. Water Treat. 2014, 52, 6104–6112. [Google Scholar] [CrossRef]
  68. Boumaza, S.; Yenounne, A.; Hachi, W.; Kaouah, F.; Bouhamidi, Y.; Trari, M. Application of Typha angustifolia (L.) Dead leaves waste as biomaterial for the removal of cationic dye from aqueous solution. Int. J. Environ. Res. 2018, 12, 561–573. [Google Scholar] [CrossRef]
  69. Setiabudi, H.D.; Jusoh, R.; Suhaimi, S.F.R.M.; Masrur, S.F. Adsorption of methylene blue onto oil palm (Elaeis guineensis) leaves: Process optimization, isotherm, kinetics and thermodynamic studies. J. Taiwan Inst. Chem. Eng. 2016, 63, 363–370. [Google Scholar] [CrossRef] [Green Version]
  70. Singh, R.; Singh, T.S.; Odiyo, J.O.; Smith, J.A.; Edokpayi, J.N. Evaluation of methylene blue sorption onto low-cost biosorbents: Equilibrium, kinetics, and thermodynamics. J. Chem. 2020, 2020, 8318049. [Google Scholar] [CrossRef] [Green Version]
  71. Vadivelan, V.; Kumar, K.V. 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]
  72. Mahadlek, J.; Mahadlek, J. Investigation of various factors affecting methylene blue adsorption on agricultural waste: Banana stalks. Sci. Eng. Health Stud. 2020, 14, 47–61. [Google Scholar]
  73. Bhattacharyya, K.G.; Sharma, A. Kinetics and thermodynamics of methylene blue adsorption on neem (Azadirachta indica) leaf powder. Dyes Pigm. 2005, 65, 51–59. [Google Scholar] [CrossRef]
  74. Al-Azabi, K.; Al-Marog, S.; Abukrain, A.; Sulyman, M. Equilibrium, isotherm studies of dye adsorption onto orange peel powder. Chem. Res. J. 2018, 3, 45–59. [Google Scholar]
  75. Fiaz, R.; Hafeez, M.; Mahmood, R. Ficcus palmata leaves as a low-cost biosorbent for methylene blue: Thermodynamic and kinetic studies. Water Environ. Res. 2019, 91, 689–699. [Google Scholar] [CrossRef]
  76. Kulkarni, M.R.; Revanth, T.; Acharya, A.; Bhat, P. Removal of Crystal Violet dye from aqueous solution using water hyacinth: Equilibrium, kinetics and thermodynamics study. Resour. Technol. 2017, 3, 71–77. [Google Scholar]
  77. Mosoarca, G.; Vancea, C.; Popa, S.; Boran, S. Optimization of crystal violet adsorption on common lilac tree leaf powder as natural adsorbent material. Glob. Nest J. 2022, 24, 87–96. [Google Scholar]
  78. Keereerak, A.; Chinpa, W. A potential biosorbent from Moringa oleifera pod husk for crystal violet adsorption: Kinetics, isotherms, thermodynamic and desorption studies. Sci. Asia 2020, 46, 186–194. [Google Scholar] [CrossRef]
  79. Pavan, F.A.; Camacho, E.S.; Lima, E.C.; Dotto, G.L.; Branco, V.T.; Dias, S. Formosa papaya seed powder (FPSP): Preparation, characterization and application as an alternative adsorbent for the removal of crystal violet from aqueous phase. J. Environ. Chem. Eng. 2014, 2, 230–238. [Google Scholar] [CrossRef]
  80. Das, 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]
  81. Ghazali, A.; Shirani, M.; Semnania, A.; Zare-Shahabadic, V.; Nekoeiniad, M. Optimization of crystal violet adsorption onto Date palm leaves as a potent biosorbent from aqueous solutions using response surface methodology and ant colony. J. Environ. Chem. Eng. 2018, 6, 3942–3950. [Google Scholar] [CrossRef]
  82. Ahmad, R. Studies on adsorption of crystal violet dye from aqueous solution onto coniferous pinus bark powder (CPBP). J. Hazard. Mater. 2009, 171, 767–773. [Google Scholar] [CrossRef]
  83. Khan, F.A.; Ahad, A.; Shah, S.S.; Farooqui, M. Adsorption of crystal violet dye using Platanus orientalis (Chinar tree) leaf powder and its biochar: Equilibrium, kinetics and thermodynamics study. Int. J. Environ. Anal. Chem. 2021, 1–21. [Google Scholar] [CrossRef]
  84. Gemici, B.T.; Ozel, H.U.; Ozel, H.B. Adsorption behaviors of crystal violet from aqueous solution using Anatolian black pine (Pinus nigra Arnold.): Kinetic and equilibrium studies. Sep. Sci. Technol. 2020, 55, 406–414. [Google Scholar] [CrossRef]
  85. Ali, H.; Muhammad, S.K. Biosorption of Crystal Violet from Water on Leaf Biomass of Calotropis procera. J. Environ. Sci. Technol. 2008, 1, 143–150. [Google Scholar] [CrossRef] [Green Version]
  86. Song, J.; Zou, W.; Bian, Y.; Su, F.; Han, R. Adsorption characteristics of methylene blue by peanut husk in batch and column modes. Desalination 2011, 265, 119–125. [Google Scholar] [CrossRef]
  87. Zehra, T.; Priyantha, N.; Lim, L.B.L. Removal of crystal violet dye from aqueous solution using yeast-treated peat as adsorbent: Thermodynamics, kinetics, and equilibrium studies. Environ. Earth Sci. 2016, 75, 357. [Google Scholar] [CrossRef]
  88. Jiang, Z.; Hu, D. Molecular mechanism of anionic dyes adsorption on cationized rice husk cellulose from agricultural wastes. J. Mol. Liq. 2019, 276, 105–114. [Google Scholar] [CrossRef]
  89. Wakkel, M.; Khiari, B.; Zagrouba, F. Textile wastewater treatment by agro-industrial waste: Equilibrium modelling, thermodynamics and mass transfer mechanisms of cationic dyes adsorption onto low-cost lignocellulosic adsorbent. J. Taiwan Inst. Chem. Eng. 2019, 96, 439–452. [Google Scholar] [CrossRef]
Figure 1. The FTIR spectra of sour cherry (Prunus cerasus) leaves powder, before and after dyes adsorption: (SCLP)—sour cherry leaf powder, (SCLP + MB)—sour cherry leaf powder after methylene blue adsorption, (SCLP + CV)—sour cherry leaf powder after crystal violet adsorption.
Figure 1. The FTIR spectra of sour cherry (Prunus cerasus) leaves powder, before and after dyes adsorption: (SCLP)—sour cherry leaf powder, (SCLP + MB)—sour cherry leaf powder after methylene blue adsorption, (SCLP + CV)—sour cherry leaf powder after crystal violet adsorption.
Materials 16 04252 g001
Figure 2. The SEM images of sour cherry (Prunus cerasus) leaves powder: (A) before adsorption, (B) after methylene blue adsorption, (C) after crystal violet adsorption (cathode voltage 25 kV; working distance 10.4 mm, magnification 3000×).
Figure 2. The SEM images of sour cherry (Prunus cerasus) leaves powder: (A) before adsorption, (B) after methylene blue adsorption, (C) after crystal violet adsorption (cathode voltage 25 kV; working distance 10.4 mm, magnification 3000×).
Materials 16 04252 g002
Figure 3. CIEL*a*b* color parameters of sour cherry (Prunus cerasus) leaves powder before and after adsorption of methylene blue and crystal violet: 1—(SCLP) sour cherry leaf powder, 2—(MB) methylene blue, 3—(CV) crystal violet, 4—(SCLP + MB) sour cherry leaf powder after methylene blue adsorption, 5—(SCLP + CV) sour cherry leaf powder after crystal violet adsorption).
Figure 3. CIEL*a*b* color parameters of sour cherry (Prunus cerasus) leaves powder before and after adsorption of methylene blue and crystal violet: 1—(SCLP) sour cherry leaf powder, 2—(MB) methylene blue, 3—(CV) crystal violet, 4—(SCLP + MB) sour cherry leaf powder after methylene blue adsorption, 5—(SCLP + CV) sour cherry leaf powder after crystal violet adsorption).
Materials 16 04252 g003
Figure 4. pHPZC determination for sour cherry (Prunus cerasus) leaves powder.
Figure 4. pHPZC determination for sour cherry (Prunus cerasus) leaves powder.
Materials 16 04252 g004
Figure 5. Effect of pH (A), ionic strength (B), and adsorbent dose (C) on methylene blue (MB) and crystal violet (CV) dyes adsorption (the circles group the curves, and the arrows indicate the ordinate to which they relate).
Figure 5. Effect of pH (A), ionic strength (B), and adsorbent dose (C) on methylene blue (MB) and crystal violet (CV) dyes adsorption (the circles group the curves, and the arrows indicate the ordinate to which they relate).
Materials 16 04252 g005
Figure 6. Adsorption isotherms used to assess the adsorption behavior of methylene blue on sour cherry (Prunus cerasus) leaves powder.
Figure 6. Adsorption isotherms used to assess the adsorption behavior of methylene blue on sour cherry (Prunus cerasus) leaves powder.
Materials 16 04252 g006
Figure 7. Adsorption isotherms used to assess the adsorption behavior of crystal violet on sour cherry (Prunus cerasus) leaves powder.
Figure 7. Adsorption isotherms used to assess the adsorption behavior of crystal violet on sour cherry (Prunus cerasus) leaves powder.
Materials 16 04252 g007
Figure 8. Kinetic models used to assess the adsorption behavior of methylene blue on sour cherry (Prunus cerasus) leaves powder.
Figure 8. Kinetic models used to assess the adsorption behavior of methylene blue on sour cherry (Prunus cerasus) leaves powder.
Materials 16 04252 g008
Figure 9. Kinetic models used to assess the adsorption behavior of crystal violet on sour cherry (Prunus cerasus) leaves powder.
Figure 9. Kinetic models used to assess the adsorption behavior of crystal violet on sour cherry (Prunus cerasus) leaves powder.
Materials 16 04252 g009
Figure 10. The desorption efficiency for the all tested desorption agents: (A) methylene blue loaded adsorbent, (B) crystal violet loaded adsorbent.
Figure 10. The desorption efficiency for the all tested desorption agents: (A) methylene blue loaded adsorbent, (B) crystal violet loaded adsorbent.
Materials 16 04252 g010
Table 1. FTIR bands that were assigned to different functional groups specific to the main components (cellulose, hemicellulose, and lignin) of the adsorbed material.
Table 1. FTIR bands that were assigned to different functional groups specific to the main components (cellulose, hemicellulose, and lignin) of the adsorbed material.
FTIR BandsAssignmentReference
3282 cm−1stretching vibration of the O–H bonds[40]
2933 cm−1–CH stretching of CH2[41]
2350 cm−1aromatic ring C=C bond[42]
1605 cm−1aromatic skeletal and C=O stretch vibrations characteristic of lignin[43]
1422 cm−1–C–H deformation in lignin[44,45]
1255 cm−1–C–O stretching and CH or OH bending of hemicellulose structures[46,47]
1057 cm−1C–O–C stretching of cellulose[23,48]
625 cm−1bending modes of aromatic compounds[49]
Table 2. Parameters of the adsorption isotherms used to assess the dyes adsorption behavior on sour cherry (Prunus cerasus) leaves powder.
Table 2. Parameters of the adsorption isotherms used to assess the dyes adsorption behavior on sour cherry (Prunus cerasus) leaves powder.
Adsorption IsothermParametersValue
MB AdsorptionCV Adsorption
Langmuir non-linearKL (L mg−1)0.0026 ± 0.00050.0041 ± 0.0008
qmax (mg g−1)543.2 ± 16.8229.8 ± 11.2
R20.98460.9982
χ29.881.04
SSE317.5119.96
ARE (%)20.727.34
Freundlich non-linearKf (mg g−1) (L mg−1)1/n1.88 ± 0.272.27 ± 0.34
1/n0.89 ± 0.080.72 ± 0.05
R20.97560.9994
χ211.180.18
SSE439.625.62
ARE (%)21.882.63
Temkin non-linearKT (L mg−1)0.31 ± 0.030.14 ± 0.02
b (kJ g−1)94.49 ± 8.3485.87 ± 6.75
R20.84710.9411
χ269.5165.19
SSE4071.57600.68
ARE (%)41.90114.97
Sips non-linearQsat (mg g−1)168.64 ± 7.41524.1 ± 16.25
KS (L mg−1)0.0004 ± 0.00010.0033 ± 0.0004
n1.86 ± 0.110.81 ± 0.08
R20.99990.9995
χ20.090.05
SSE0.813.83
ARE (%)2.261.04
Redlich-Peterson
non-linear
KRP (L g−1)1.34 ± 0.271.25 ± 0.24
aRP (L mg−1)0.0003 ± 0.00010.049 ± 0.007
βRP0.74 ± 0.150.62 ± 0.07
R20.99280.9994
χ27.760.20
SSE191.416.12
ARE (%)18.243.07
qm and Qsat are the maximum absorption capacities; KL, KF, KT, KS, and KRP are the Langmuir, Freundlich, Temkin, Sips, and Redlich–Peterson isotherms constants; 1/nF is an empirical constant indicating the intensity of adsorption; b is Temkin constant which related to the adsorption heat; n is Sips isotherm exponent; aRP is Redlich–Peterson isotherm constant, βRP is Redlich–Peterson exponent; R2 is determining the value of the determination coefficient; SSE is the sum of square error; χ2 is chi-square and ARE is average relative error.
Table 3. Comparison of the maximum absorption capacities (qm) of various similar absorbents use for methylene blue and crystal violet removal from aqueous solutions.
Table 3. Comparison of the maximum absorption capacities (qm) of various similar absorbents use for methylene blue and crystal violet removal from aqueous solutions.
Methylene Blue Adsorption
Adsorbentqm (mg g−1)Reference
raspberry leaves244.6[62]
citrus limetta peel227.3[13]
lotus leaf221.7[53]
bilberry leaves200.4[61]
sour-cherry leaves168.6This study
Magnolia grandiflora leaves149.2[66]
banana leaves109.9[67]
Typha angustifolia leaves106.7[68]
Elaeis guineensis leaves103.0[69]
palm kernel fiber95.4[16]
Pará chestnut husk83.8[18]
phoenix tree’s leaves80.9[55]
Salix babylonica60.97[23]
Daucus carota leaves66.7[37]
potato leaves52.6[56]
Ginkgo biloba leaves48.1[70]
rice husk40.5[71]
banana stalks20.8[72]
Azadirachta indica leaves19.6[73]
orange peel18.6[74]
Ficcus Palmata leaves6.8[75]
Crystal Violet Adsorption
Adsorbentqm (mg g−1)Reference
sour-cherry leaves524.1This study
water hyacinth root322.5[76]
Ocotea puberula bark272.1[54]
lilac tree leaf196.7[77]
bathurst burr biomass164.1[52]
Moringa oleifera pod husk156.2[78]
Artocarpus altilis skin145.8[64]
motherwort biomass125.6[42]
papaya seeds powder85.9[79]
Pará chestnut husk83.6[18]
palm kernel fiber78.9[16]
Ananasus comosus leaf78.2[20]
Eragrostis plana nees60.1[65]
jackfruit leaf powder43.3[80]
date palm leaves37.7[81]
pinus bark powder32.7[82]
Platanus orientalis leaf25.8[83]
Arundo donax L.19.6[57]
anatolian black pine12.3[84]
almond shells12.2[31]
Calotropis procera leaf4.14[85]
Table 4. The values of the equilibrium time reported in the scientific literature for the adsorption of methylene blue and crystal violet on various adsorbents obtained from plant biomass.
Table 4. The values of the equilibrium time reported in the scientific literature for the adsorption of methylene blue and crystal violet on various adsorbents obtained from plant biomass.
Methylene Blue Adsorption
AdsorbentEquilibrium Time (min)Reference
potato leaves24[56]
Daucus carota leaves30[37]
orange peel30[74]
sour-cherry leaves40This study
bilberry leaves40[61]
raspberry leaves40[62]
Pará chestnut husk40[18]
Magnolia grandiflora leaves60[66]
Typha angustifolia leaves60[68]
palm kernel fiber60[16]
banana stalks60[72]
Azadirachta indica leaves60[73]
Ficcus Palmata leaves80[75]
Ginkgo biloba leaves100[70]
Salix babylonica120[23]
banana leaves120[67]
phoenix tree’s leaves150[55]
rice husk150[70]
citrus limetta peel180[13]
lotus leaf240[53]
Crystal Violet Adsorption
AdsorbentEquilibrium Time (min)Reference
date palm leaves21[81]
Moringa oleifera pod husk24[78]
Arundo donax L.30[57]
bathurst burr biomass30[52]
Platanus orientalis leaf30[83]
sour-cherry leaves40This study
Pará chestnut husk40[18]
lilac tree leaf40[77]
motherwort biomass50[42]
palm kernel fiber60[16]
papaya seeds powder60[79]
Calotropis procera leaf60[85]
almond shells90[31]
Ocotea puberula bark120[54]
water hyacinth root120[76]
jackfruit leaf powder120[80]
pinus bark powder120[82]
anatolian black pine120[84]
Eragrostis plana nees180[65]
Artocarpus altilis skin210[64]
Table 5. Parameters of the kinetic models used to assess the dyes adsorption behavior on sour cherry (Prunus cerasus) leaves powder.
Table 5. Parameters of the kinetic models used to assess the dyes adsorption behavior on sour cherry (Prunus cerasus) leaves powder.
Kinetic ModelParametersValue
MB AdsorptionCV Adsorption
Pseudo-first orderk1 (min−1)1.33 ± 0.070.41 ± 0.05
qe,calc (mg g−1)15.33 ± 0.7816.46 ± 0.81
R20.96050.9636
χ20.591.35
SSE8.7711.86
ARE (%)5.127.46
Pseudo-second orderk2 (min−1)0.164 ± 0.0140.034 ± 0.004
qe,calc (g mg−1 min−1)15.90 ± 0.6417.65 ± 0.95
R20.98830.9912
χ20.170.29
SSE2.582.70
ARE (%)2.943.74
Elovicha (g mg−1)0.98 ± 0.170.41 ± 0.06
b (mg g−1 min−1)(20.54 ± 0.26) × 10476.01 ± 5.87
R20.99900.9828
χ20.040.40
SSE0.235.25
ARE (%)0.794.09
General orderkN (min−1 (g mg−1)n–1)0.0004 ± 0.00010.006 ± 0.001
qn (mg g−1)16.43 ± 0.3518.11 ± 0.68
n3.89 ± 0.122.29 ± 0.18
R20.99700.9959
χ20.050.09
SSE0.791.22
ARE (%)1.441.99
AvramikAV (min−1)0.63 ± 0.050.77 ± 0.08
qAV (mg g−1)15.33 ± 0.3416.45 ± 0.47
nAV0.95 ± 0.090.53 ± 0.06
R20.96050.9636
χ20.591.35
SSE8.7711.86
ARE (%)5.117.46
qt is the dye amount adsorbed at time t; k1, k2, kn, and kAV are the rate constants of pseudo-first-order, pseudo-second-order, general order, and Avrami kinetic models; qe, qn, and qAV are the theoretical values for the adsorption capacity; a is the desorption constant of Elovich model; b is the initial velocity; n is the general order exponent and nAV is a fractional exponent; R2 is determining the value of the determination coefficient; SSE is the sum of square error; χ2 is chi-square and ARE is average relative error.
Table 6. The thermodynamic parameters used to assess the dyes adsorption process.
Table 6. The thermodynamic parameters used to assess the dyes adsorption process.
DyeΔG0 (kJ mol−1)ΔH0 (kJ mol−1)ΔS0 (J mol−1 K−1)
283 K297 K317 K
MB adsorption−16.34−17.05−17.81−0.515.16
CV adsorption−17.91−18.34−19.55−0.495.83
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

Mosoarca, G.; Vancea, C.; Popa, S.; Dan, M.; Boran, S. A Novel High-Efficiency Natural Biosorbent Material Obtained from Sour Cherry (Prunus cerasus) Leaf Biomass for Cationic Dyes Adsorption. Materials 2023, 16, 4252. https://doi.org/10.3390/ma16124252

AMA Style

Mosoarca G, Vancea C, Popa S, Dan M, Boran S. A Novel High-Efficiency Natural Biosorbent Material Obtained from Sour Cherry (Prunus cerasus) Leaf Biomass for Cationic Dyes Adsorption. Materials. 2023; 16(12):4252. https://doi.org/10.3390/ma16124252

Chicago/Turabian Style

Mosoarca, Giannin, Cosmin Vancea, Simona Popa, Mircea Dan, and Sorina Boran. 2023. "A Novel High-Efficiency Natural Biosorbent Material Obtained from Sour Cherry (Prunus cerasus) Leaf Biomass for Cationic Dyes Adsorption" Materials 16, no. 12: 4252. https://doi.org/10.3390/ma16124252

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