Acessibilidade / Reportar erro

Moringa oleifera seed husks for methylene blue dye adsorption: kinetic, equilibrium, and thermodynamic analyses

Cascas da semente de Moringa oleifera para adsorção do corante azul de metileno: análise cinética, de equilíbrio e termodinâmica

Abstract

M. ole·ifera se·ed hu·sks (H-M·O) w·ere u·sed as adsorbent for the removal of methylene blue (MB) dye from an aqueous solution. The ads·orbent was char·acterized by SE·M, ED·X, X·RD, F·TIR, B·ET, and p·HPZC. To evaluate adsorption capacity, the influences of p·H, kinetics, isotherms, and thermodynamic prop·erties were anal·yzed. Characteri·zation tech·niques indic·ated that H-MO has het·erogeneous morp·hological charac·teristics wi·th a spe·cific sur·face a·rea of 3.24 m2 g-1, mai·nly comp·osed of ce·llulose distr·ibuted in an amor·phous stru·cture. MB adsorption was favorable at pH values higher than pHPZC of 6.6, using an adsorbent dosage of 1.0 g L-1. Equilibrium was achieved in the first 240 min, and the pseudo-second-order model was suitable for describing the kinetic data. Freu·ndlich was the mo·st adeq·uate model for descri·bing the isoth·erm cur·ves, pred·icting a max·imum adsor·ption capa·city of 122.7 mg g-1 at 24°C, hig·her tha·n oth·er natu·ral ad·sorbents. The ads·orption proc·ess w·as spo·ntaneous and exoth·ermic, indic·ating that lo·wer temper·atures fav·or the ad·sorption and th·at it is cont·rolled by p·hysical fo·rces. The da·ta presen·ted indic·ate the M·B has t·he pot·ential to be succ·essfully trea·ted by the eco-fr·iendly and l·ow-cost ads·orbent M. oleif·era seed h·usks.

Keywords:
agricultural wastes; low-cost adsorbent; physisorption

Resumo

Cas·cas de sem·entes de M. oleifera (H-MO) fora·m usadas com·o adsorve·nte para remo·ver o coran·te az·ul de metile·no (MB) de uma solu·ção aqu·osa. O adsorve·nte foi caracte·rizado por ME·V, ED·S, DRX, FT·IR, B·ET e pH·PCZ. Para aval·iar a capacid·ade de adsor·ção, foi ana·lisada a influ·ência do pH, ciné·tica, isoter·mas e propr·iedades termo·dinâmicas. Técnic·as de caracter·ização indic·aram que a H-MO poss·ui caracter·ísticas morfológ·icas hetero·gêneas com á·rea supe·rficial espe·cífica de 3,24 m2 g-1, co·mposta prin·cipalmente por celu·lose distr·ibuída em estr·utura am·orfa. A adso·rção de MB foi favo·rável em va·lores de pH supe·riores a pHP·CZ de 6,6, utili·zando uma dosa·gem de ads·orvente de 1,0 g L-1. O equ·ilíbrio foi alca·nçado nos prim·eiros 240 min, e o mode·lo de pseu·do seg·unda or·dem foi adequa·do para desc·rever os dad·os cin·éticos. Fre·undlich foi o mo·delo mais ade·quado para des·crever as curvas isot·érmicas, preve·ndo uma cap·acidade máx·ima de ads·orção de 122,7 mg g-1 a 24°C, supe·rior a ou·tros adsor·ventes nat·urais. O pro·cesso de adsor·ção foi esp·ontâneo e exoté·rmico, indi·cando que temper·aturas mais ba·ixas favor·ecem a adsor·ção e que ela é con·trolada por for·ças físi·cas. Os d·ados aprese·ntados indi·cam que o M·B tem· po·tencial para ser tr·atado com suc·esso pelas ca·scas de semen·tes de M. oleif·era, um adsorvente ecol·ogicamente favo·rável e de ba·ixo custo.

Palavras-chave:
adsorventes de baixo custo; fisissorção; resíduos agrícolas

1. INTRODUCTION

In the last decades, a growing interest in the use of alternative, low-cost adsorbents (LCAs) has been observed. LCAs are classified as natural materials (e.g., wood, coal, clays), industrial wastes/by-products (e.g., fly ash, red mud), and agricultural wastes/by-products (e.g., seed, seed husk, corn cob waste) (Gupta et al., 2009GUPTA, V. K.; CARROTT, P. J. M.; RIBEIRO CARROTT, M. M. L.; SUHAS. Low-cost adsorbents: growing approach to wastewater treatment - a review. Critical Reviews in Environmental Science and Technology, v. 39, n. 10, p. 783-842, 2009. https://doi.org/10.1080/10643380801977610
https://doi.org/10.1080/1064338080197761...
). In this sense, the seed husk of Moringa oleifera is included as agricultural waste. This plant has several applications in the food, pharmaceutical, cosmetics, and food nutrition industries due to its properties (Ueda Yamaguchi et al., 2021UEDA YAMAGUCHI, N.; CUSIOLI, L. F.; QUESADA, H. B.; CAMARGO FERREIRA, M. E.; FAGUNDES-KLEN, M. R.; SALCEDO VIEIRA, A. M. et al. A review of Moringa oleifera seeds in water treatment: Trends and future challenges. Process Safety and Environmental Protection, v. 147, p. 405-420, 2021. https://doi.org/10.1016/j.psep.2020.09.044
https://doi.org/10.1016/j.psep.2020.09.0...
). Also, for many years, several studies have been investigating the potential of its seeds in drinking water treatment (Ribeiro et al., 2019RIBEIRO, J. V. M.; ANDRADE, P. V.; REIS A. G. Moringa oleifera seed as a natural coagulant to treat low-turbidity water by in-line filtration. Revista Ambiente & Água, v. 14, n. 6, p. e2442, 2019. https://doi.org/10.4136/ambi-agua.2442
https://doi.org/10.4136/ambi-agua.2442...
), and in domestic (Vega Andrade et al., 2021VEGA ANDRADE, P.; PALANCA, C. F.; OLIVEIRA, M. A. C.; ITO, C. Y. K.; REIS, A. G. Use of Moringa oleifera seed as a natural coagulant in domestic wastewater tertiary treatment: Physicochemical, cytotoxicity and bacterial load evaluation. Journal of Water Process Engineering, v. 40, p. 101859, 2021. https://doi.org/10.1016/j.jwpe.2020.101859
https://doi.org/10.1016/j.jwpe.2020.1018...
) and industrial wastewater treatment (Villaseñor-Basulto et al., 2018VILLASEÑOR-BASULTO, D. L.; ASTUDILLO-SÁNCHEZ, P. D.; REAL-OLVERA, J.; BANDALA, E. R. Wastewater treatment using Moringa oleifera Lam seeds: A review. Journal of Water Process Engineering, v. 23, p. 151-164, 2018. https://doi.org/10.1016/j.jwpe.2018.03.017
https://doi.org/10.1016/j.jwpe.2018.03.0...
) due to its coagulant properties. In addition to its use as a coagulant, M. oleifera seed has also been investigated as an alternative in the treatment of drinking water and wastewater as an adsorbent, for example, in the removal of heavy metals (Araújo et al., 2013ARAÚJO, C. S. T; CARVALHO, D. C; REZENDE, H. C.; ALMEIDA, I. L. S.; COELHO, L. M.; COELHO, N. M. M. et al. Bioremediation of Waters Contaminated with Heavy Metals Using Moringa oleifera Seeds as Biosorbent. In: PATIL, Y. B.; RAO, P. (eds.). Applied Bioremediation - Active and Passive Approaches. London: IntechOpen, 2013. http://dx.doi.org/10.5772/56157
http://dx.doi.org/10.5772/56157...
) and dyes (Reck et al., 2018RECK, I. M.; PAIXÃO, R. M.; BERGAMASCO, R.; VIEIRA, M. F.; VIEIRA, M. A. S. Removal of tartrazine from aqueous solutions using adsorbents based on activated carbon and Moringa oleifera seeds. Journal of Cleaner Production, v. 171, n. 10, p. 85-97, 2018. https://doi.org/10.1016/j.jclepro.2017.09.237
https://doi.org/10.1016/j.jclepro.2017.0...
). However, the use of the M. oleifera seed generates the seed’s husk waste. The potential advancement in the use of this seed as a coagulant and adsorbent on a commercial scale requires better management of its waste (husk) through the evaluation of possible scientific and technological applications.

The use of M. oleifera seed husk has not proved to be efficient as a coagulant (Kansal and Kumari, 2014KANSAL, S. K.; KUMARI, A. Potential of M. oleifera for the treatment of water and wastewater. Chemical Reviews, v. 114, n. 9, p. 4993-5010, 2014. https://doi.org/10.1021/cr400093w
https://doi.org/10.1021/cr400093w...
). However, recent studies have been conducted to evaluate the possibility of using the husk as an adsorbent to remove pollutants from water, such as atrazine herbicide (Cusioli et al., 2019CUSIOLI, L. F.; BEZERRA, C. O.; QUESADA, H. B.; BAPTISTA, A. T. A.; NISHI, L.; VIEIRA, M. F. et al. Modified Moringa oleifera Lam. Seed husks as low-cost biosorbent for atrazine removal. Environmental Technology, v. 42, p. 1092-1103, 2019. https://doi.org/10.1080/09593330.2019.1653381
https://doi.org/10.1080/09593330.2019.16...
), acetaminophen analgesic (Quesada et al., 2019QUESADA, H. B.; CUSIOLI, L. F.; BEZERRA, C. O.; BAPTISTA, A. T. A.; NISHI, L.; GOMES, R. G. et al. Acetaminophen adsorption using a low-cost adsorbent prepared from modified residues of Moringa oleifera Lam. seed husks. Journal of Chemical Technology and Biotechnology, v. 94, p. 3147-3157, 2019. https://doi.org/10.1002/jctb.6121
https://doi.org/10.1002/jctb.6121...
), diuron pesticide (Bezerra et al., 2018BEZERRA, C. O.; CUSIOLI, L. F.; QUESADA, H. B.; NISHI, L.; MANTOVANI, D.; VIEIRA, M. F. et al. Assessment of the use of Moringa oleifera seed husks for removal of pesticide diuron from contaminated water. Environmental Technology, v. 41, n. 2, p. 191-201, 2018. https://doi.org/10.1080/09593330.2018.1493148
https://doi.org/10.1080/09593330.2018.14...
), trihalomethanes (Okoya et al., 2020OKOYA, A. A.; OLAIYA, O. O.; AKINYELE, A. B.; OCHOR, N. O. Efficacy of Moringa oleifera Seed Husk as Adsorptive Agent for Trihalomethanes from a Water Treatment Plant in Southwestern, Nigeria. Journal of Chemistry, v. 2020, p. 3450954, 2020. https://doi.org/10.1155/2020/3450954
https://doi.org/10.1155/2020/3450954...
), and cyanobacterial hepatotoxin microcystin-LR (Warhurst et al., 1997WARHURST, A. M.; RAGGETT, S. L.; MCCONNACHIE, G. L.; POLLARD, S. J. T.; CHIPOFYA, V.; CODD, G.A. Adsorption of the cyanobacterial hepatotoxin microcystin-LR by a low-cost activated carbon from the seed husks of the pan-tropical tree, Moringa oleifera. Science of The Total Environment, v. 207, n. 2-3, p. 207-211, 1997. https://doi.org/10.1016/S0048-9697(97)00260-X
https://doi.org/10.1016/S0048-9697(97)00...
). The capacity of M. oleifera seed husk to remove those pollutants shows its potential application as a LCA in the treatment of drinking water and wastewater, and new studies for removal of different problematic pollutants are required, such as dyes from textile wastewaters.

Dye industry effluents, when not properly treated, can cause an oxygen imbalance in ecosystems by dissolved oxygen depletion, in addition to preventing sunlight penetration, thereby altering the photosynthetic activity of the environment, which results in water quality deterioration and, consequently, in adverse effects to the local fauna and flora (Yagub et al., 2014YAGUB, M. T.; SEN, T. K.; AFROZE, S.; ANG, H. M. Dye and its removal from aqueous solution by adsorption: a review. Advances in Colloid and Interface Science, v. 209, p. 172-184, 2014. https://doi.org/10.1016/j.cis.2014.04.002
https://doi.org/10.1016/j.cis.2014.04.00...
). Methylene blue (MB) dye is a synthetic and cationic dye, from the group of thiazines, soluble in water and widely used in the textile industry to color silk, cotton, wool, leather, and paper coating. This dye shows toxicity such as teratogenicity, mutagenicity, neurotoxicity, nucleic acid damage, etc. (Sabnis, 2010SABNIS, R. W. Biological Dyes and Stains Synthesis and Industrial Applications. Hoboken : John Wiley & Sons, 2010.) An aggravating factor is that dyes are resistant to degradation, and this property becomes an obstacle to effluent treatment, since they are substances that have highly stable chemical structures, making their removal difficult whether by biological, chemical, or physical process (Yagub et al., 2014YAGUB, M. T.; SEN, T. K.; AFROZE, S.; ANG, H. M. Dye and its removal from aqueous solution by adsorption: a review. Advances in Colloid and Interface Science, v. 209, p. 172-184, 2014. https://doi.org/10.1016/j.cis.2014.04.002
https://doi.org/10.1016/j.cis.2014.04.00...
). Several methods are used for the removal of dyes from wastewater, such as adsorption, coagulation, advanced oxidation, and membrane separation (Gupta et al., 2009GUPTA, V. K.; CARROTT, P. J. M.; RIBEIRO CARROTT, M. M. L.; SUHAS. Low-cost adsorbents: growing approach to wastewater treatment - a review. Critical Reviews in Environmental Science and Technology, v. 39, n. 10, p. 783-842, 2009. https://doi.org/10.1080/10643380801977610
https://doi.org/10.1080/1064338080197761...
). Many textile industries use commercial activated carbon for the treatment of dye waste, as adsorption is one of the most effective processes of advanced wastewater treatment (Yagub et al., 2014YAGUB, M. T.; SEN, T. K.; AFROZE, S.; ANG, H. M. Dye and its removal from aqueous solution by adsorption: a review. Advances in Colloid and Interface Science, v. 209, p. 172-184, 2014. https://doi.org/10.1016/j.cis.2014.04.002
https://doi.org/10.1016/j.cis.2014.04.00...
). Several types of adsorbents have been evaluated for MB dye removal as a cost-effective alternative to commercial activated carbons, such as the agricultural wastes/by-products LCAs apple pomace (Bonetto et al., 2021BONETTO, L. R.; CRESPO, J. S.; GUÉGAN, R.; ESTEVES, V. I.; GIOVANELA, M. Removal of methylene blue from aqueous solutions using a solid residue of the apple juice industry: Full factorial design, equilibrium, thermodynamics and kinetics aspects. Journal of Molecular Structure, v. 1224, n. 15, p. 129296, 2021. https://doi.org/10.1016/j.molstruc.2020.129296
https://doi.org/10.1016/j.molstruc.2020....
), Brazilian berry seeds (E. uniflora) (Georgin et al., 2020GEORGIN, 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). Environmental Science and Pollution Research, v. 27, p. 20831-20843, 2020. https://doi.org/10.1007/s11356-020-08496-8
https://doi.org/10.1007/s11356-020-08496...
), and sorghum husk (Adeogun et al., 2019ADEOGUN, A. I.; AKANDE, J. A.; IDOWU, M. A.; KAREEM, S. O. Magnetic tuned sorghum husk biosorbent for effective removal of cationic dyes from aqueous solution: isotherm, kinetics, thermodynamics and optimization studies. Applied Water Science, v. 9, n. 160, 2019. https://doi.org/10.1007/s13201-019-1037-2
https://doi.org/10.1007/s13201-019-1037-...
).

In this context, considering the need for better use of M. oleifera seed husk as a low-cost adsorbent from agricultural waste and the risks associated with the discharge of MB dye from an industrial effluent, a systematic investigation, including equilibrium, thermodynamics, kinetics, and mechanism of the adsorption of MB dye by M. oleifera seed husk, is a subject of interest and is discussed in this paper.

2. MATERIAL AND METHODS

2.1. Adsorbent preparation and characterization

Moringa oleifera seeds were purchased from Arbocenter, and the seeds were harvested in the city of Araçatuba (20°56′19.72″ S, 50°40′6.17″ W), São Paulo - Brazil. First, the seeds were manually peeled. Seed kernels were segregated for future studies related to oil extraction and further use as a coagulant. The husks were then washed with deionized water and dried in a SolidSteel’s 42L SSDc oven at 60°C for 24 h. After drying, the husks were ground using a commercial blender and sieved to be standardized on a maximum particle size of 600 µm. The powder obtained was used as adsorbent, and hereinafter will be called “H-MO”.

The surface functional groups of H-MO were determined by Fourier Transform Infrared (FTIR) Spectroscopy (Shimadzu IRAffinity-1), in which the samples were scanned in the region of 4000 - 650 cm-1, in transmittance mode. The structural properties of the powder samples were determined by X-ray diffractometer (Rigaku Ultima IV), equipped with CuKα radiation (λ = 0.154 nm) at 40 kV, 30 mA and 2θ scan ranging from 5° to 80° with 2° min-1 step. Surface morphology was evaluated by scanning electron microscopy (SEM) (Tescan Vega 3), with 15.0 kV acceleration voltage. The energy dispersive X-ray spectroscopy (EDX) provided the elemental analysis of the H-MO through the SEM coupled with the EDX detector (Oxford X-act). For this analysis, the samples were previously covered with gold at a thickness of approximately 25 nm. The point of zero charge (pHPZC) was measured by the salt addition method with 0.1M NaNO3 (purity ≥ 98%, Êxodo), proposed by Bakatula et al. (2018)BAKATULA, E. N.; RICHARD, D.; NECULITA, C. M.; ZAGURI, G. J. Determination of point of zero charge of natural organic materials. Environmental Science and Pollution Research, v. 25, p. 7823-7833, 2018. https://doi.org/10.1007/s11356-017-1115-7
https://doi.org/10.1007/s11356-017-1115-...
. To determine the specific surface area, the BET method was used by adsorption/desorption isotherms of N2 at -196°C in a Nova 4200e (Quantachrome). Before the adsorption of N2, the samples were dried for 24 h at 60°C and subjected to degassing for 3 h at 90°C.

2.2. H-MO adsorption studies

The adsorption assays were conducted using methylene blue dye (MB) hydrate (purity ≥ 82%, Neon) as adsorbate. From a MB stock solution (1 g L-1), dilutions with demineralized water were made for the adsorption study. The experiments were conducted in triplicate in a batch mode on a thermostatic stirring bath (Dubnoff 304-TPA) at 100 rpm. After adsorption time, the solutions were then centrifuged (Excelsa II 206-BL) at 3,000 rpm for 5 min and the MB concentration was measured by UV-Vis spectrophotometer (Shimadzu UV 1800), operating at λmax = 664 nm wavelength. The percentage of dye removal (R, % - Equation 1), adsorption capacity at any time (q t , mg g−1 - Equation 2) and at equilibrium (q e , mg g−1 - Equation 3) were calculated as follows:

R = ( C 0 - C t ) 100 C 0 (1)

q t = ( C 0 - C t ) V m (2)

q e = ( C 0 - C e ) V m (3)

Where C 0 is the initial concentration of MB, C t is the concentration of MB at any time, C e is the equilibrium concentration of MB, V is the volume of solution, and m is the mass of the adsorbent.

The H-MO dosage effect was analyzed for the values of 0.2, 0.6, 1.0, 1.6, 2.0, 2.6, 3.0, and 4.0 g L-1. For these assays, the pH was set at 7.0. The different H-MO dosages were put in contact with 50 mL of MB solution with an initial concentration of 50 mg L-1. Since the equilibrium time was not determined yet, solutions were stirred during 1,200 min at 24°C. Based on the percentage of dye removal (R), the ideal dosage of adsorbent for the next tests was determined.

The pH effect test was conducted for the pH values of 5.0, 6.0, 7.0, 8.0, and 9.0, using 1.0 g L-1 of H-MO adsorbent, previously defined by the dosage effect. The pH adjustments were made with HNO3 0.1M (purity ≥ 66%, LS Chemicals) or NaOH 0.1M (purity 99.67%, Neon). H-MO adsorbent was added to 50 mL of MB solutions with initial concentration of 50 mg L-1, and the solutions were stirred for 1,200 min at 24°C.

The kinetic adsorption experiments were conducted using the best conditions found in the pH (9.0) and dosage (1.0 g L-1) tests. H-MO adsorbent was added to 50 mL of MB solutions with an initial concentration of 50 mg L-1, and the solutions were agitated at 24°C. Samples were collected at predefined times of 5, 20, 40, 60, 120, 240, 360, 480, 600, 720, and 1440 min. Through adsorption kinetics, the equilibrium time was determined, which was used in the adsorption isotherm test.

The adsorption isotherms were determined by varying the initial MB concentration (5, 10, 15, 20, 30, 40, 50, 60, 70, 80 and 90 mg L-1) in 50 mL of solution with dosage of adsorbent (1.0 g L-1), pH value (pH = 9), and equilibrium time (240 min) previously determined. The temperatures evaluated were 24°C, 34°C, and 44°C.

3. RESULTS AND DISCUSSION

3.1. H-MO adsorbent characterization

The morphology H-MO's SEM images are shown in Figure 1, where non-uniform complex fiber matrix with no particular shape can be observed, in agreement with other authors (Araujo et al., 2018ARAUJO, L. A.; BEZERRA, C. O.; CUSIOLI, L. F.; SILVA, M. F.; NISHIB, L.; GOMES, R. G. et al. Moringa oleifera biomass residue for the removal of pharmaceuticals from water. Journal of Environmental Chemical Engineering, v. 6, n. 6, p. 7192-7199, 2018. https://doi.org/10.1016/j.jece.2018.11.016
https://doi.org/10.1016/j.jece.2018.11.0...
; Quesada et al., 2019QUESADA, H. B.; CUSIOLI, L. F.; BEZERRA, C. O.; BAPTISTA, A. T. A.; NISHI, L.; GOMES, R. G. et al. Acetaminophen adsorption using a low-cost adsorbent prepared from modified residues of Moringa oleifera Lam. seed husks. Journal of Chemical Technology and Biotechnology, v. 94, p. 3147-3157, 2019. https://doi.org/10.1002/jctb.6121
https://doi.org/10.1002/jctb.6121...
). The overall H-MO chemical composition, as a result of semi-quantitative EDX spectrometry, presented a major presence of carbon (69.4% of weight) and oxygen (30.4% of weight), which are related to the natural organic structure of the lignocellulosic structure of the material, such as cellulose, hemicellulose, and lignin (Bezerra et al., 2018BEZERRA, C. O.; CUSIOLI, L. F.; QUESADA, H. B.; NISHI, L.; MANTOVANI, D.; VIEIRA, M. F. et al. Assessment of the use of Moringa oleifera seed husks for removal of pesticide diuron from contaminated water. Environmental Technology, v. 41, n. 2, p. 191-201, 2018. https://doi.org/10.1080/09593330.2018.1493148
https://doi.org/10.1080/09593330.2018.14...
; Garcia-Fayos et al., 2016GARCIA-FAYOS, B.; ARNAL, J. M.; PIRIS, J.; SANCHO, M. Valorization of Moringa oleifera seed husk as biosorbent: isotherm and kinetics studies to remove cadmium and copper from aqueous solutions. Desalination and Water Treatment, v. 57, n. 48-49, p. 23382-23396, 2016. https://doi.org/10.1080/19443994.2016.1180473
https://doi.org/10.1080/19443994.2016.11...
).

Figure 1.
SEM images of the H-MO adsorbent at different magnifications: a) × 2000 e b) × 6000.

The N2 adsorption-desorption isotherm of H-MO adsorbent is presented in Figure 2, in which one can notice a typical behavior of a Type II isotherm with Type H3 hysteresis. Type II isotherms are given by the physisorption of most gases on nonporous or microporous adsorbents. The shape is the result of unrestricted monolayer-multilayer adsorption up to high P/P0. Since the curvature in Figure 2 is more gradual, this is an indication of a great amount of overlap of monolayer coverage and the onset of multilayer adsorption. When P/P0 = 1, the thickness of the adsorbed multilayer usually appears to increase without limit (Thommes et al., 2015THOMMES, M.; KANEKO, K.; NEIMARK, A. V.; OLIVIER, J. P.; RODRIGUEZ-REINOSO, F.; ROUQUEROL, J. et al. Physisorption of gasses, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure and Applied Chemistry, v. 87, p. 1051-1069, 2015. https://doi.org/10.1515/pac-2014-1117
https://doi.org/10.1515/pac-2014-1117...
). The open-loop presented in the N2 adsorption-desorption isotherm (Figure 2) is probably related to the pressure-dependent elastic deformation behavior, which is in accordance with low surface area adsorbents (Tang et al., 2017TANG, J.; YANG, M.; YU, F.; CHEN, X.; TAN, L.; WANG, G. 1-Octadecanol@ hierarchical porous polymer composite as a novel shape-stability phase change material for latent heat thermal energy storage. Applied Energy, v. 187, p. 514-522, 2017. https://doi.org/10.1016/j.apenergy.2016.11.043
https://doi.org/10.1016/j.apenergy.2016....
).

Figure 2.
BET N2 adsorption-desorption isotherm at -196°C for H-MO.

The BET textural parameters of H-MO indicate a low specific surface area and total pore volume of 3.24 m2 g-1 and 0.010 cm3 g-1, respectively. These results are expected, as the seed husks have not been activated, and are in accordance with the literature relating to in natura M. oleifera seed husks. Cusioli et al. (2019)CUSIOLI, L. F.; BEZERRA, C. O.; QUESADA, H. B.; BAPTISTA, A. T. A.; NISHI, L.; VIEIRA, M. F. et al. Modified Moringa oleifera Lam. Seed husks as low-cost biosorbent for atrazine removal. Environmental Technology, v. 42, p. 1092-1103, 2019. https://doi.org/10.1080/09593330.2019.1653381
https://doi.org/10.1080/09593330.2019.16...
reported 1.52 m2 g-1 and 0.021 cm3 g-1, and Quesada et al. (2019)QUESADA, H. B.; CUSIOLI, L. F.; BEZERRA, C. O.; BAPTISTA, A. T. A.; NISHI, L.; GOMES, R. G. et al. Acetaminophen adsorption using a low-cost adsorbent prepared from modified residues of Moringa oleifera Lam. seed husks. Journal of Chemical Technology and Biotechnology, v. 94, p. 3147-3157, 2019. https://doi.org/10.1002/jctb.6121
https://doi.org/10.1002/jctb.6121...
reported 1.2 m2 g-1 and 0.002 cm3 g-1 of specific surface area and total pore volume, respectively. The results are also in agreement with the behavior of a Type II isotherm, shown in Figure 2, of nonporous adsorbents.

The XRD pattern was conducted to determine the degree of crystalline or amorphous nature of the H-MO adsorbent and is presented in Figure 3(a). One can see, from the wide shape of the pattern, the material presents a predominance of an amorphous characteristic, suggesting a more disordered structure of the carbonaceous matrix. The main peak is detected at the 2θ value of around 22°, and the small peaks at around 16° and 35°, and they are attributed to crystalline cellulose (Barnette et al., 2012BARNETTE, A. L.; LEE, C.; BRADLEY, L. C.; SCHREINER, E. P.; PARK, Y. B.; SHIN, H. et al. Quantification of crystalline cellulose in lignocellulosic biomass using sum frequency generation (SFG) vibration spectroscopy and comparison with other analytical methods. Carbohydrate Polymers, v. 89, n. 3, p. 802-809, 2012. https://doi.org/10.1016/j.carbpol.2012.04.014
https://doi.org/10.1016/j.carbpol.2012.0...
). Similar peaks have been also found for modified M. oleifera seed husk (Bezerra et al., 2018BEZERRA, C. O.; CUSIOLI, L. F.; QUESADA, H. B.; NISHI, L.; MANTOVANI, D.; VIEIRA, M. F. et al. Assessment of the use of Moringa oleifera seed husks for removal of pesticide diuron from contaminated water. Environmental Technology, v. 41, n. 2, p. 191-201, 2018. https://doi.org/10.1080/09593330.2018.1493148
https://doi.org/10.1080/09593330.2018.14...
) and Brazilian berry seeds (Eugenia uniflora) (Georgin et al., 2020GEORGIN, 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). Environmental Science and Pollution Research, v. 27, p. 20831-20843, 2020. https://doi.org/10.1007/s11356-020-08496-8
https://doi.org/10.1007/s11356-020-08496...
).

The FTIR spectrum of H-MO is shown in Figure 3(b), exhibiting characteristic bands of lignocellulosic materials (Li et al., 2018LI, X.; WEI, Y.; XU, J.; HE, Y. Quantitative visualization of lignocellulose components in transverse sections of moso bamboo based on FTIR macro- and micro-spectroscopy coupled with chemometrics. Biotechnology for Biofuels, v. 11, n. 263, 2018. https://doi.org/10.1186/s13068-018-1251-4
https://doi.org/10.1186/s13068-018-1251-...
). The broad peak in 3370 cm-1 indicates hydroxyl bonds O-H. Smaller peak in 2922 cm-1 could describe symmetric and asymmetric C-H stretching vibrations. The 1638 cm-1 peak can appear due to the interactions between water and C=O or N-H stretching vibrations. Finally, the strong peak at 1032 cm-1 can be originated by C-O bonds. The FTIR spectra also agrees with the literature relating to in natura M. oleifera seed husks (Tavares et al., 2017TAVARES, F. O.; PINTO, L. A. M.; BASSETTI, F. J.; VIEIRA, M. F.; BERGAMASCO, R.; VIEIRA, M. A. S. Environmentally friendly biosorbents (husks, pods and seeds) from Moringa oleifera for Pb(II) removal from contaminated water. Environmental Technology, v. 38, n. 24, p. 3145-3155, 2017. https://doi.org/10.1080/09593330.2017.1290150
https://doi.org/10.1080/09593330.2017.12...
; Garcia-Fayos et al., 2016GARCIA-FAYOS, B.; ARNAL, J. M.; PIRIS, J.; SANCHO, M. Valorization of Moringa oleifera seed husk as biosorbent: isotherm and kinetics studies to remove cadmium and copper from aqueous solutions. Desalination and Water Treatment, v. 57, n. 48-49, p. 23382-23396, 2016. https://doi.org/10.1080/19443994.2016.1180473
https://doi.org/10.1080/19443994.2016.11...
). One can thus conclude that, according to the FTIR spectra found, the surface of the adsorbent possesses a wide variety of functional groups that favor the adsorptive process.

Figure 3.
a) X-ray diffractogram at 40 kV and 30 mA and b) FTIR spectrum in the region of 4000 - 650 cm-1 of H-MO.

3.2. Effect of adsorbent dosage on adsorption capacity

The effect of H-MO adsorbent dosage on the percentage removal of the adsorbate MB is shown in Figure 4. Initially, the percentage of MB removal increases sharply as the adsorbent dosage increases. Such an increase might be attributed to the higher number of available adsorbent active sites for adsorption (Shah et al., 2015SHAH, I.; ADNAN, R.; NGAH, W. S. W; MOHAMED, N. Iron impregnated activated carbon as an efficient adsorbent for the removal of methylene blue: regeneration and kinetics studies. PLoS One, v. 10, p. e0122603, 2015 https://doi.org/10.1371/journal.pone.0122603
https://doi.org/10.1371/journal.pone.012...
). A maximum removal value of 98% is achieved at a dosage of 4 g L-1. However, considering that in the dosage of 1 g L-1 an MB removal greater than 90% was achieved, and that the increase in this dosage did not bring significant increases in removal, the dosage of 1 g L-1 of adsorbent is justified for economic purposes and was chosen for this study.

Figure 4.
Effect of dosage on adsorption capacity of MB dye onto H-MO (C0 = 50 mg L-1, 24°C, V = 50 mL, pH = 7.0, 1,200 min, 100 rpm).

3.3. Point of zero charge (pHPZC) and effect of the solution pH on adsorption capacity

Figure 5(a) shows the influence of pH value on the adsorption of MB by H-MO. One can observe that pH values increase from 5 to 9 to enhance the adsorption capacity. This effect is a result of the MB characteristics, as well as the H-MO surface properties when submitted to solutions with different pH values. Figure 5(b) shows that the pHPZC of H-MO is 6.6. The value of pHPZC is an important property to understand the adsorption process, since when the solution is at a pH value above pHPZC, the functional groups of the adsorbent are deprotonated, resulting in a negatively charged surface, and contributing to the favorable adsorption of the cationic adsorbates. MB has a pK a value of 2.6, so MB dye is in the cationic form at pH value greater than 2.6 (Sabnis, 2010SABNIS, R. W. Biological Dyes and Stains Synthesis and Industrial Applications. Hoboken : John Wiley & Sons, 2010.). This effect enhances the electrostatic attraction between the negatively charged surface of H-MO and positively charged MB dye. Considering the results obtained for the pH effect and pHPZC, the maximum value of adsorption capacity on this study is reached at pH 9 (qe = 52.02 mg g-1), and this pH value was set for the other tests.

Figure 5.
Adsorption of MB on H-MO: a) pH effect (C0 = 50 mg L−1, 24°C, V = 50 mL, H-MO dosage = 1 g L−1, 1,200 min, 100 rpm) and b) point of zero charge (pHPZC) by the salt addition method with 0.1M NaNO3.

3.4. Adsorption kinetics and modeling

Adsorption kinetics were determined to evaluate the effect of contact time on MB dye adsorption on H-MO adsorbent. To better understand the MB dye adsorption process, the kinetic models of pseudo-first-order (Lagergren, 1898LAGERGREN, S. About the theory of so-called adsorption of soluble substances. Kungl. Svenska vetenskapsakademiens handlingar, v. 24, n. 4, p. 1-39, 1898.) and pseudo-second-order (Blanchard et al., 1984BLANCHARD, G.; MAUNAYE, M.; MARTIN, G. Removal of heavy metals from waters by means of natural zeolites. Water Research, v. 18 n. 12, p. 1501-1507, 1984. https://doi.org/10.1016/0043-1354(84)90124-6
https://doi.org/10.1016/0043-1354(84)901...
) were fitted to the experimental data using nonlinear method (Equations 4 and 5, respectively) and are represented in Figure 6.

q t = q e ( 1 - e - k 1 t ) (4)

q t = q e 2 k 2 t 1 + k 2 q e t (5)

Where qt is the amount of adsorbate uptake per mass of adsorbent at time t, qe is the amount of adsorbate uptake per mass of adsorbent at equilibrium, k1 is the rate constant of the pseudo-first-order equation, and k2 is the rate constant of the pseudo-second-order equation.

Figure 6.
Kinetic curves for the adsorption of MB dye onto H-MO (C0 = 50 mg L−1, 24°C, V = 50 mL, H-MO dosage = 1 g L−1, pH = 9.0, 100 rpm).

The adsorption kinetics presented in Figure 6 shows that H-MO adsorption capacity increases sharply at the beginning of the process, and it can be associated with the large number of active sites available. But the adsorption capacity reaches stability with a contact time of 240 min (qe = 48.6 mg g-1). This stability of qe values indicates that equilibrium was achieved. This may be explained by the fact that almost all empty sites of the H-MO surface adsorbed the MB molecules available on the solution, making it difficult for new MB molecules to be adsorbed.

Correlation coefficients and parameters of the kinetic models for MB dye adsorption on H-MO are presented in Table 1. The pseudo-second-order was better fitted to the experimental data, since the coefficient of determination (R2) value was higher (0.946); the chi-square (χ2) value, related to the error, was lower (0.326); and the adsorption capacity calculated (48.4 mg g-1) strongly agreed with that obtained experimentally (48.6 mg g-1). This means it is likely that both external and intraparticle diffusion steps control the total kinetics of the adsorption process (Ho and McKay, 1999HO, Y. S.; MCKAY, G. Pseudo-second order model for sorption processes. Process Biochemistry, v. 34, n. 5, p. 451-465, 1999. https://doi.org/10.1016/S0032-9592(98)00112-5
https://doi.org/10.1016/S0032-9592(98)00...
). The pseudo-second-order model is a good fit to the kinetics data for most adsorption processes involving dyes (Yagub et al., 2014YAGUB, M. T.; SEN, T. K.; AFROZE, S.; ANG, H. M. Dye and its removal from aqueous solution by adsorption: a review. Advances in Colloid and Interface Science, v. 209, p. 172-184, 2014. https://doi.org/10.1016/j.cis.2014.04.002
https://doi.org/10.1016/j.cis.2014.04.00...
).

Table 1.
Kinetic parameters of non-linear models for MB dye onto H-MO.

3.5. Adsorption isotherms

The adsorption isotherms were determined to better understand the interaction mechanisms between MB dye and the H-MO adsorbent. The isotherm models of Langmuir (Langmuir, 1918)LANGMUIR, I. The adsorption of gasses on plane surfaces of glass, mica and platinum. Journal of the American Chemical Society, v. 40, n. 9, p. 1361-1403, 1918. and Freundlich (Freundlich, 1906FREUNDLICH, H. Über die Adsorption in Lösungen. Zeitschrift für Physikalische Chemie, v. 57, p. 385-471, 1906.) were fitted to the experimental data using nonlinear methods (Equations 6 and 7, respectively) and are represented in Figure 7.

q e = q m a x K L C e 1 + K L C e (6)

q e = K F C e 1 n (7)

Where q e is the amount of adsorbate uptake at equilibrium, C e is the equilibrium concentration of adsorbate, q max is the maximum adsorption capacity, K L is adsorbent/adsorbate interaction constant, K F is Freundlich adsorption capacity constant, and 1/n is the Freundlich intensity parameter, which indicates the magnitude of the adsorption driving force or the surface heterogeneity.

Figure 7.
Adsorption isotherm for the adsorption of MB dye onto H-MO (V = 50 mL, H-MO dosage = 1 g L−1, pH = 9.0, 240 min, 100 rpm).

The calculated parameters for such isotherms are shown in Table 2.

Table 2.
Parameters of non-linear equilibrium isotherms models representing equilibrium adsorption data for MB onto H-MO.

The separation factor or equilibrium parameter R L is derived from the Langmuir model and can be calculated according to Equation 8 (Hall et al., 1966HALL, K. R.; EAGLETON, L. C.; ACRIVOS, A.; VERMEULEN, T. Pore- and solid-diffusion kinetics in fixed-bed adsorption under constant-pattern conditions. Industrial & Engineering Chemistry Fundamentals, v. 5, n. 2, p. 212-223, 1966. https://doi.org/10.1021/i160018a011
https://doi.org/10.1021/i160018a011...
):

R L = 1 1 + K L C 0 (8)

Where K L is the Langmuir equilibrium constant, and C0 is the initial adsorbate concentration. The calculated values of R L are in the range of 0.06 to 0.56 (24°C), 0.07 to 0.59 (34°C), and 0.11 to 0.72 (44°C).

According to Figure 7, the adsorption capacity of MB dye increased with increasing equilibrium concentration (Ce), with the adsorption isotherms showing a typical concave shape. In addition to that, 0 < R L < 1 and 0 < 1/n < 1 (Table 2). All these parameters indicate the adsorption is favorable for the studied temperatures (24°C, 34°C and 44°C).

The correlation coefficients and parameters of the adsorption models for MB dye adsorption on H-MO presented in Table 2 indicate that both models were well adjusted for the adsorption processes in the studied temperatures, with a coefficient of determination R2 ranging from 0.97 to 0.99. However, the chi-square (χ 2) values were lower from the Freundlich model (6.25 to 13.23) when compared to Langmuir model (11.29 to 21.98) for the three temperatures studied, indicating that the Freundlich model was better fitted to the experimental data. The Freundlich model assumes the adsorbent surface energy is heterogeneous with the strongest binding sites being occupied first and the binding force decreasing with the increasing degree of site occupancy (Febrianto et al., 2009FEBRIANTO, J.; KOSASIH, A. N.; SUNARSO, J.; JU, Y. H.; INDRASWATI, N.; ISMADJI, S. Equilibrium and kinetic studies in adsorption of heavy metals using biosorbent: a summary of recent studies. Journal of Hazardous Materials, v. 162, n. 2-3, p. 616-645, 2009. https://doi.org/10.1016/j.jhazmat.2008.06.042
https://doi.org/10.1016/j.jhazmat.2008.0...
). A better suitability of the Freundlich model compared to the Langmuir model for the adsorption of MB dye was also reported by Tuli et al. (2020)TULI, F. J.; HOSSAIN, A.; FAZLE KIBRIA, A. K. M.; TAREQ, A. R. M.; MAMUN, S. M. M. A.; ATIQUE ULLAH, A. K. M. Removal of methylene blue from water by low-cost activated carbon prepared from tea waste: A study of adsorption isotherm and kinetics. Environmental Nanotechnology, Monitoring & Management, v. 14, p. 100354, 2020. https://doi.org/10.1016/j.enmm.2020.100354
https://doi.org/10.1016/j.enmm.2020.1003...
using activated tea waste carbon, and by Khodaie et al. (2013)KHODAIE, M.; GHASEMI, N.; MORADI, B.; RAHIMI, M. Removal of methylene blue from wastewater by adsorption onto ZnCl2 activated corn husk carbon equilibrium studies. Journal of Chemistry, v. 2013, p. 383985, 2013. https://doi.org/10.1155/2013/383985
https://doi.org/10.1155/2013/383985...
using activated corn husk carbon.

From the temperature effect on the equilibrium curves shown in Figure 7 and Table 2, one may notice the temperature increase provided a little decrease in adsorption capacity (qmax), from 122.7 (24°C) to 119.8 mg g−1 (44°C).

The adsorption capacity of various adsorbents is commonly compared by the parameter q max . Table 3 shows the maximum adsorption capacity of different adsorbents used for MB dye removal. The value obtained for q max in this study is higher than those of other adsorbents produced with agricultural wastes. M. oleifera seed husk is shown to be a good adsorbent for MB dye, although the result for q max is quite lower than those obtained with activated carbon. However, the adsorbent used in this study does not require high temperatures during its processing or activation steps, meaning that this material has a lower operating cost than activated carbon.

Table 3.
Comparison of adsorption capacities of different adsorbents for MB dye removal.

3.6. Thermodynamic parameters

Thermodynamic adsorption plays a key role in predicting viability, spontaneity, and adsorptive mechanisms (Tran et al., 2016TRAN, H. N.; YOU, S. J.; CHAO, H. P. Thermodynamic parameters of cadmium adsorption onto orange peel calculated from various methods: A comparison study. Journal of Environmental Chemical Engineering, v. 4, n. 3, p. 2671-2682, 2016. https://doi.org/10.1016/j.jece.2016.05.009
https://doi.org/10.1016/j.jece.2016.05.0...
). The thermodynamic parameters can be computed according to the laws of thermodynamics, using the following Equation 9:

G ° = - R T l n K C (9)

The relationship between the three thermodynamic parameters, Gibbs energy (ΔG°), enthalpy (ΔH°), and entropy (ΔS°) is described by Equation 10:

G ° = Δ H ° - T S ° (10)

By substituting Equation 9 in Equation 10, one can obtain Equation 11, known as van’t Hoff equation.

l n K C = - H ° R T + S ° R (11)

Where R is the universal constant (8.314 J mol-1 K-1) and T is the solution temperature (K). Enthalpy (ΔH°) and entropy (ΔS°) changes were determined from the slope and intercept, respectively, of a plot of lnK C versus 1/T (Equation 11). Gibbs energy change (ΔG°) was directly calculated from Equation 9.

The dimensionless thermodynamic equilibrium constant K C in Equations 9 and 11 was calculated from the Freundlich adsorption isotherm constant K F , since this model was better fitted to the experimental data, through Equation 12 (Tran et al., 2016TRAN, H. N.; YOU, S. J.; CHAO, H. P. Thermodynamic parameters of cadmium adsorption onto orange peel calculated from various methods: A comparison study. Journal of Environmental Chemical Engineering, v. 4, n. 3, p. 2671-2682, 2016. https://doi.org/10.1016/j.jece.2016.05.009
https://doi.org/10.1016/j.jece.2016.05.0...
; Ghosal and Gupta, 2015GHOSAL, P. S.; GUPTA, A. K. An insight into thermodynamics of adsorptive removal of fluoride by calcined Ca-Al-(NO3) layered double hydroxide. RSC Advances, v. 5, n. 128, p. 105889-105900, 2015. https://doi.org/10.1039/C5RA20538G
https://doi.org/10.1039/C5RA20538G...
).

K C = K F ρ 1000 10 6 ρ ( 1 - 1 / n ) (12)

Where ρ is the density of pure water (assumed as ~ 1 g mL-1) and 1/n is the Freundlich intensity parameter.

The MB adsorption thermodynamic parameters onto H-MO are shown in Table 4. Evidence that suggests the appropriate application of the Freundlich constant KF in estimating the thermodynamic parameters is the high determination coefficient of the van’t Hoff equation (R2 = 0.991). A critical factor for determining the degree of spontaneity of the adsorption process is the Gibbs energy variation: the negative values of Gibbs energy (ΔG°) shown in Table 4 denote that the adsorption is energetically favorable; in other words, the process tends to be spontaneous. In addition, the decreasing of the magnitude of ΔG° with temperature means the process tends to a less thermodynamically favorable process at higher temperatures, which can be confirmed by the enthalpy value. Similar results were found for the adsorption of MB by apple pomace (Bonetto et al., 2021BONETTO, L. R.; CRESPO, J. S.; GUÉGAN, R.; ESTEVES, V. I.; GIOVANELA, M. Removal of methylene blue from aqueous solutions using a solid residue of the apple juice industry: Full factorial design, equilibrium, thermodynamics and kinetics aspects. Journal of Molecular Structure, v. 1224, n. 15, p. 129296, 2021. https://doi.org/10.1016/j.molstruc.2020.129296
https://doi.org/10.1016/j.molstruc.2020....
).

Table 4.
Thermodynamic parameters for adsorption of MB dye onto H-MO.

The negative value of the enthalpy parameter (ΔH°) shown in Table 4 indicates the adsorption process is exothermic, which means the decrease in temperature results in an increase in adsorption capacity (qe ) (Table 2) and in the equilibrium constant (KC ) (Table 4). Furthermore, the type of interaction can be assigned, to a certain extent, by the magnitude of enthalpy change. Physisorption, such as van der Waals interactions, is usually lower than 20 kJ mol-1. Electrostatic interaction ranges from 20-80 kJ mol-1 and they are also frequently classified as physisorption, while a value in the range of 80-450 kJ mol-1 indicates a chemisorption (Alencar et al., 2012ALENCAR, W. S.; LIMA, E. C.; ROYER, B.; DOS SANTOS, B. D.; CALVETE, T.; DA SILVA, E. A. et al. Application of aqai stalks as biosorbents for the removal of the dye procion blue MX-R from aqueous solution. Separation Science and Technology, v. 47 n. 3, p. 513-526, 2012. https://doi.org/10.1080/01496395.2011.616568
https://doi.org/10.1080/01496395.2011.61...
; Ghosh et al., 2021GHOSH, I.; KAR, S.; CHATTERJEE, T.; BAR, N.; DAS, S. K. Removal of methylene blue from aqueous solution using Lathyrus sativus husk: Adsorption study, MPR and ANN modeling. Process Safety and Environmental Protection, v. 149, n. 345-361, 2021. https://doi.org/10.1016/j.psep.2020.11.003
https://doi.org/10.1016/j.psep.2020.11.0...
). Therefore, the typical physisorption enthalpy value for this study (ΔH° = -45.8 kJ mol-1) may be related to the sum of different molecular interactions, such as hydrogen bonding and electrostatic interactions between active sites on the adsorbent and the cationic dye. An exothermic adsorption process was also reported elsewhere (Ghosh et al., 2021GHOSH, I.; KAR, S.; CHATTERJEE, T.; BAR, N.; DAS, S. K. Removal of methylene blue from aqueous solution using Lathyrus sativus husk: Adsorption study, MPR and ANN modeling. Process Safety and Environmental Protection, v. 149, n. 345-361, 2021. https://doi.org/10.1016/j.psep.2020.11.003
https://doi.org/10.1016/j.psep.2020.11.0...
). Finally, the negative sign in entropy (ΔS°) indicates the organization of the adsorbate at the solid/solution interface during the adsorption process becomes less random, also implying the adsorption phenomenon involves an associative mechanism.

3.7. Adsorption mechanism

Figure 8 shows the proposed interaction between the MB and the H-MO. Based on the dataset discussed in this study, one could show that MB dye removal by H-MO occurs mainly due to electrostatic interactions, which is evidenced by the adsorption capacity enhancement in pH values above pHPZC. Similar conclusions were reported by other authors who investigated the adsorption of MB by agricultural wastes (Bonetto et al., 2021BONETTO, L. R.; CRESPO, J. S.; GUÉGAN, R.; ESTEVES, V. I.; GIOVANELA, M. Removal of methylene blue from aqueous solutions using a solid residue of the apple juice industry: Full factorial design, equilibrium, thermodynamics and kinetics aspects. Journal of Molecular Structure, v. 1224, n. 15, p. 129296, 2021. https://doi.org/10.1016/j.molstruc.2020.129296
https://doi.org/10.1016/j.molstruc.2020....
; Ghosh et al., 2021GHOSH, I.; KAR, S.; CHATTERJEE, T.; BAR, N.; DAS, S. K. Removal of methylene blue from aqueous solution using Lathyrus sativus husk: Adsorption study, MPR and ANN modeling. Process Safety and Environmental Protection, v. 149, n. 345-361, 2021. https://doi.org/10.1016/j.psep.2020.11.003
https://doi.org/10.1016/j.psep.2020.11.0...
). Additionally, the value found for ΔH° (-45.8 kJ mol-1) reinforces that these interactions are in the range of electrostatic interaction, also classified as physisorption, with no sharing or exchange of electrons.

Figure 8
Proposed interaction for MB dye and H-MO adsorption system.

Moreover, H-MO composition presents several oxygenated functional groups, as presented by crystalline cellulose shown in XRD pattern (Figure 3(a)), and by O-H, C=O, N-H and C-O groups by FTIR spectrum (Figure 3(b)), which can also interact with MB molecules through dipole-dipole hydrogen bonds (Tran et al. 2017TRAN, H. N.; YOU, S. J.; NGUYEN, T. V.; CHAO, H. P. Insight into adsorption mechanism of cationic dye onto biosorbents derived from agricultural wastes. Chemical Engineering Communications, v. 204, n. 9, p. 1020-1036, 2017. https://doi.org/10.1080/00986445.2017.1336090
https://doi.org/10.1080/00986445.2017.13...
). The incidence of aromatic rings, both in the chemical structure of H-MO (cellulose) and in MB dye molecules, can also favor π- π interactions, contributing to the adsorption process (Cemin et al., 2021CEMIN, A.; FERRARINI, F.; POLETTO, M.; BONETTO, L. R.; BORTOLUZ, J.; LEMÉE, L. et al. Characterization and use of a lignin sample extracted from Eucalyptus grandis sawdust for the removal of methylene blue dye. International Journal of Biological Macromolecules, v. 170, n. 15, p. 375-389, 2021. https://doi.org/10.1016/j.ijbiomac.2020.12.155
https://doi.org/10.1016/j.ijbiomac.2020....
).

4. CONCLUSIONS

This study evaluated the adsorption capacity of methylene blue (MB) dye using Moringa oleifera seed husks (H-MO). The kinetic model that showed the best fit was the pseudo-second-order model, and the isotherm that presented the best fit was a Freundlich model, indicating a favorable adsorption process (1/n from 0.60 to 0.63) with a maximum adsorption capacity of MB by H-MO of 122.7 mg g-1 at 24°C, higher than some recently published studies in the literature. The thermodynamic data indicated that the adsorption process is spontaneous (ΔG° from −3.88 to −1.06 kJ mol−1) and exothermic (ΔH° of -45.80 KJ mol−1), showing that lower temperatures favor MB adsorption by H-MO and that adsorption is controlled by physical forces. The disadvantage of the adsorbent preparation method of presenting low surface area (3.24 m2 g-1) and total pore volume (0.010 cm3 g-1) were expected, as the seed husks have not been previously activated, but were compensated by the electrostatic attraction at pH values above 6.6 (pHPZC) and by dipole-dipole hydrogen bonds and π-π interactions due to its surface composition, at any pH, resulting in a favorable adsorption process. Thus, from the results obtained, one can conclude that H-MO has promise for use in the adsorption process to remove MB in aqueous solution. Considering that seed husks are an agricultural waste, their use as adsorbents of agricultural waste is a promising and low-cost option, which can also minimize the environmental impacts of their improper disposal.

5. ACKNOWLEDGEMENTS

The authors are grateful to National Council for Scientific and Technological Development (CNPq) - ENT#091;grant PIBIC 123798/2020-5ENT#093; for the partial financial support.

6. REFERENCES

  • ADEOGUN, A. I.; AKANDE, J. A.; IDOWU, M. A.; KAREEM, S. O. Magnetic tuned sorghum husk biosorbent for effective removal of cationic dyes from aqueous solution: isotherm, kinetics, thermodynamics and optimization studies. Applied Water Science, v. 9, n. 160, 2019. https://doi.org/10.1007/s13201-019-1037-2
    » https://doi.org/10.1007/s13201-019-1037-2
  • ALENCAR, W. S.; LIMA, E. C.; ROYER, B.; DOS SANTOS, B. D.; CALVETE, T.; DA SILVA, E. A. et al. Application of aqai stalks as biosorbents for the removal of the dye procion blue MX-R from aqueous solution. Separation Science and Technology, v. 47 n. 3, p. 513-526, 2012. https://doi.org/10.1080/01496395.2011.616568
    » https://doi.org/10.1080/01496395.2011.616568
  • ARAÚJO, C. S. T; CARVALHO, D. C; REZENDE, H. C.; ALMEIDA, I. L. S.; COELHO, L. M.; COELHO, N. M. M. et al. Bioremediation of Waters Contaminated with Heavy Metals Using Moringa oleifera Seeds as Biosorbent. In: PATIL, Y. B.; RAO, P. (eds.). Applied Bioremediation - Active and Passive Approaches. London: IntechOpen, 2013. http://dx.doi.org/10.5772/56157
    » http://dx.doi.org/10.5772/56157
  • ARAUJO, L. A.; BEZERRA, C. O.; CUSIOLI, L. F.; SILVA, M. F.; NISHIB, L.; GOMES, R. G. et al. Moringa oleifera biomass residue for the removal of pharmaceuticals from water. Journal of Environmental Chemical Engineering, v. 6, n. 6, p. 7192-7199, 2018. https://doi.org/10.1016/j.jece.2018.11.016
    » https://doi.org/10.1016/j.jece.2018.11.016
  • BAKATULA, E. N.; RICHARD, D.; NECULITA, C. M.; ZAGURI, G. J. Determination of point of zero charge of natural organic materials. Environmental Science and Pollution Research, v. 25, p. 7823-7833, 2018. https://doi.org/10.1007/s11356-017-1115-7
    » https://doi.org/10.1007/s11356-017-1115-7
  • BARNETTE, A. L.; LEE, C.; BRADLEY, L. C.; SCHREINER, E. P.; PARK, Y. B.; SHIN, H. et al. Quantification of crystalline cellulose in lignocellulosic biomass using sum frequency generation (SFG) vibration spectroscopy and comparison with other analytical methods. Carbohydrate Polymers, v. 89, n. 3, p. 802-809, 2012. https://doi.org/10.1016/j.carbpol.2012.04.014
    » https://doi.org/10.1016/j.carbpol.2012.04.014
  • BEZERRA, C. O.; CUSIOLI, L. F.; QUESADA, H. B.; NISHI, L.; MANTOVANI, D.; VIEIRA, M. F. et al. Assessment of the use of Moringa oleifera seed husks for removal of pesticide diuron from contaminated water. Environmental Technology, v. 41, n. 2, p. 191-201, 2018. https://doi.org/10.1080/09593330.2018.1493148
    » https://doi.org/10.1080/09593330.2018.1493148
  • BLANCHARD, G.; MAUNAYE, M.; MARTIN, G. Removal of heavy metals from waters by means of natural zeolites. Water Research, v. 18 n. 12, p. 1501-1507, 1984. https://doi.org/10.1016/0043-1354(84)90124-6
    » https://doi.org/10.1016/0043-1354(84)90124-6
  • BONETTO, L. R.; CRESPO, J. S.; GUÉGAN, R.; ESTEVES, V. I.; GIOVANELA, M. Removal of methylene blue from aqueous solutions using a solid residue of the apple juice industry: Full factorial design, equilibrium, thermodynamics and kinetics aspects. Journal of Molecular Structure, v. 1224, n. 15, p. 129296, 2021. https://doi.org/10.1016/j.molstruc.2020.129296
    » https://doi.org/10.1016/j.molstruc.2020.129296
  • CEMIN, A.; FERRARINI, F.; POLETTO, M.; BONETTO, L. R.; BORTOLUZ, J.; LEMÉE, L. et al. Characterization and use of a lignin sample extracted from Eucalyptus grandis sawdust for the removal of methylene blue dye. International Journal of Biological Macromolecules, v. 170, n. 15, p. 375-389, 2021. https://doi.org/10.1016/j.ijbiomac.2020.12.155
    » https://doi.org/10.1016/j.ijbiomac.2020.12.155
  • CUSIOLI, L. F.; BEZERRA, C. O.; QUESADA, H. B.; BAPTISTA, A. T. A.; NISHI, L.; VIEIRA, M. F. et al. Modified Moringa oleifera Lam. Seed husks as low-cost biosorbent for atrazine removal. Environmental Technology, v. 42, p. 1092-1103, 2019. https://doi.org/10.1080/09593330.2019.1653381
    » https://doi.org/10.1080/09593330.2019.1653381
  • FEBRIANTO, J.; KOSASIH, A. N.; SUNARSO, J.; JU, Y. H.; INDRASWATI, N.; ISMADJI, S. Equilibrium and kinetic studies in adsorption of heavy metals using biosorbent: a summary of recent studies. Journal of Hazardous Materials, v. 162, n. 2-3, p. 616-645, 2009. https://doi.org/10.1016/j.jhazmat.2008.06.042
    » https://doi.org/10.1016/j.jhazmat.2008.06.042
  • FREUNDLICH, H. Über die Adsorption in Lösungen. Zeitschrift für Physikalische Chemie, v. 57, p. 385-471, 1906.
  • GARCIA-FAYOS, B.; ARNAL, J. M.; PIRIS, J.; SANCHO, M. Valorization of Moringa oleifera seed husk as biosorbent: isotherm and kinetics studies to remove cadmium and copper from aqueous solutions. Desalination and Water Treatment, v. 57, n. 48-49, p. 23382-23396, 2016. https://doi.org/10.1080/19443994.2016.1180473
    » https://doi.org/10.1080/19443994.2016.1180473
  • 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). Environmental Science and Pollution Research, v. 27, p. 20831-20843, 2020. https://doi.org/10.1007/s11356-020-08496-8
    » https://doi.org/10.1007/s11356-020-08496-8
  • GHOSH, I.; KAR, S.; CHATTERJEE, T.; BAR, N.; DAS, S. K. Removal of methylene blue from aqueous solution using Lathyrus sativus husk: Adsorption study, MPR and ANN modeling. Process Safety and Environmental Protection, v. 149, n. 345-361, 2021. https://doi.org/10.1016/j.psep.2020.11.003
    » https://doi.org/10.1016/j.psep.2020.11.003
  • GHOSAL, P. S.; GUPTA, A. K. An insight into thermodynamics of adsorptive removal of fluoride by calcined Ca-Al-(NO3) layered double hydroxide. RSC Advances, v. 5, n. 128, p. 105889-105900, 2015. https://doi.org/10.1039/C5RA20538G
    » https://doi.org/10.1039/C5RA20538G
  • GUPTA, V. K.; CARROTT, P. J. M.; RIBEIRO CARROTT, M. M. L.; SUHAS. Low-cost adsorbents: growing approach to wastewater treatment - a review. Critical Reviews in Environmental Science and Technology, v. 39, n. 10, p. 783-842, 2009. https://doi.org/10.1080/10643380801977610
    » https://doi.org/10.1080/10643380801977610
  • HALL, K. R.; EAGLETON, L. C.; ACRIVOS, A.; VERMEULEN, T. Pore- and solid-diffusion kinetics in fixed-bed adsorption under constant-pattern conditions. Industrial & Engineering Chemistry Fundamentals, v. 5, n. 2, p. 212-223, 1966. https://doi.org/10.1021/i160018a011
    » https://doi.org/10.1021/i160018a011
  • HO, Y. S.; MCKAY, G. Pseudo-second order model for sorption processes. Process Biochemistry, v. 34, n. 5, p. 451-465, 1999. https://doi.org/10.1016/S0032-9592(98)00112-5
    » https://doi.org/10.1016/S0032-9592(98)00112-5
  • KANSAL, S. K.; KUMARI, A. Potential of M. oleifera for the treatment of water and wastewater. Chemical Reviews, v. 114, n. 9, p. 4993-5010, 2014. https://doi.org/10.1021/cr400093w
    » https://doi.org/10.1021/cr400093w
  • KHODAIE, M.; GHASEMI, N.; MORADI, B.; RAHIMI, M. Removal of methylene blue from wastewater by adsorption onto ZnCl2 activated corn husk carbon equilibrium studies. Journal of Chemistry, v. 2013, p. 383985, 2013. https://doi.org/10.1155/2013/383985
    » https://doi.org/10.1155/2013/383985
  • LAGERGREN, S. About the theory of so-called adsorption of soluble substances. Kungl. Svenska vetenskapsakademiens handlingar, v. 24, n. 4, p. 1-39, 1898.
  • LANGMUIR, I. The adsorption of gasses on plane surfaces of glass, mica and platinum. Journal of the American Chemical Society, v. 40, n. 9, p. 1361-1403, 1918.
  • LI, X.; WEI, Y.; XU, J.; HE, Y. Quantitative visualization of lignocellulose components in transverse sections of moso bamboo based on FTIR macro- and micro-spectroscopy coupled with chemometrics. Biotechnology for Biofuels, v. 11, n. 263, 2018. https://doi.org/10.1186/s13068-018-1251-4
    » https://doi.org/10.1186/s13068-018-1251-4
  • OKOYA, A. A.; OLAIYA, O. O.; AKINYELE, A. B.; OCHOR, N. O. Efficacy of Moringa oleifera Seed Husk as Adsorptive Agent for Trihalomethanes from a Water Treatment Plant in Southwestern, Nigeria. Journal of Chemistry, v. 2020, p. 3450954, 2020. https://doi.org/10.1155/2020/3450954
    » https://doi.org/10.1155/2020/3450954
  • QUESADA, H. B.; CUSIOLI, L. F.; BEZERRA, C. O.; BAPTISTA, A. T. A.; NISHI, L.; GOMES, R. G. et al. Acetaminophen adsorption using a low-cost adsorbent prepared from modified residues of Moringa oleifera Lam. seed husks. Journal of Chemical Technology and Biotechnology, v. 94, p. 3147-3157, 2019. https://doi.org/10.1002/jctb.6121
    » https://doi.org/10.1002/jctb.6121
  • RECK, I. M.; PAIXÃO, R. M.; BERGAMASCO, R.; VIEIRA, M. F.; VIEIRA, M. A. S. Removal of tartrazine from aqueous solutions using adsorbents based on activated carbon and Moringa oleifera seeds. Journal of Cleaner Production, v. 171, n. 10, p. 85-97, 2018. https://doi.org/10.1016/j.jclepro.2017.09.237
    » https://doi.org/10.1016/j.jclepro.2017.09.237
  • RIBEIRO, J. V. M.; ANDRADE, P. V.; REIS A. G. Moringa oleifera seed as a natural coagulant to treat low-turbidity water by in-line filtration. Revista Ambiente & Água, v. 14, n. 6, p. e2442, 2019. https://doi.org/10.4136/ambi-agua.2442
    » https://doi.org/10.4136/ambi-agua.2442
  • SABNIS, R. W. Biological Dyes and Stains Synthesis and Industrial Applications. Hoboken : John Wiley & Sons, 2010.
  • SHAH, I.; ADNAN, R.; NGAH, W. S. W; MOHAMED, N. Iron impregnated activated carbon as an efficient adsorbent for the removal of methylene blue: regeneration and kinetics studies. PLoS One, v. 10, p. e0122603, 2015 https://doi.org/10.1371/journal.pone.0122603
    » https://doi.org/10.1371/journal.pone.0122603
  • TANG, J.; YANG, M.; YU, F.; CHEN, X.; TAN, L.; WANG, G. 1-Octadecanol@ hierarchical porous polymer composite as a novel shape-stability phase change material for latent heat thermal energy storage. Applied Energy, v. 187, p. 514-522, 2017. https://doi.org/10.1016/j.apenergy.2016.11.043
    » https://doi.org/10.1016/j.apenergy.2016.11.043
  • TAVARES, F. O.; PINTO, L. A. M.; BASSETTI, F. J.; VIEIRA, M. F.; BERGAMASCO, R.; VIEIRA, M. A. S. Environmentally friendly biosorbents (husks, pods and seeds) from Moringa oleifera for Pb(II) removal from contaminated water. Environmental Technology, v. 38, n. 24, p. 3145-3155, 2017. https://doi.org/10.1080/09593330.2017.1290150
    » https://doi.org/10.1080/09593330.2017.1290150
  • THOMMES, M.; KANEKO, K.; NEIMARK, A. V.; OLIVIER, J. P.; RODRIGUEZ-REINOSO, F.; ROUQUEROL, J. et al. Physisorption of gasses, with special reference to the evaluation of surface area and pore size distribution (IUPAC Technical Report). Pure and Applied Chemistry, v. 87, p. 1051-1069, 2015. https://doi.org/10.1515/pac-2014-1117
    » https://doi.org/10.1515/pac-2014-1117
  • TRAN, H. N.; YOU, S. J.; CHAO, H. P. Thermodynamic parameters of cadmium adsorption onto orange peel calculated from various methods: A comparison study. Journal of Environmental Chemical Engineering, v. 4, n. 3, p. 2671-2682, 2016. https://doi.org/10.1016/j.jece.2016.05.009
    » https://doi.org/10.1016/j.jece.2016.05.009
  • TRAN, H. N.; YOU, S. J.; NGUYEN, T. V.; CHAO, H. P. Insight into adsorption mechanism of cationic dye onto biosorbents derived from agricultural wastes. Chemical Engineering Communications, v. 204, n. 9, p. 1020-1036, 2017. https://doi.org/10.1080/00986445.2017.1336090
    » https://doi.org/10.1080/00986445.2017.1336090
  • TULI, F. J.; HOSSAIN, A.; FAZLE KIBRIA, A. K. M.; TAREQ, A. R. M.; MAMUN, S. M. M. A.; ATIQUE ULLAH, A. K. M. Removal of methylene blue from water by low-cost activated carbon prepared from tea waste: A study of adsorption isotherm and kinetics. Environmental Nanotechnology, Monitoring & Management, v. 14, p. 100354, 2020. https://doi.org/10.1016/j.enmm.2020.100354
    » https://doi.org/10.1016/j.enmm.2020.100354
  • UEDA YAMAGUCHI, N.; CUSIOLI, L. F.; QUESADA, H. B.; CAMARGO FERREIRA, M. E.; FAGUNDES-KLEN, M. R.; SALCEDO VIEIRA, A. M. et al. A review of Moringa oleifera seeds in water treatment: Trends and future challenges. Process Safety and Environmental Protection, v. 147, p. 405-420, 2021. https://doi.org/10.1016/j.psep.2020.09.044
    » https://doi.org/10.1016/j.psep.2020.09.044
  • VEGA ANDRADE, P.; PALANCA, C. F.; OLIVEIRA, M. A. C.; ITO, C. Y. K.; REIS, A. G. Use of Moringa oleifera seed as a natural coagulant in domestic wastewater tertiary treatment: Physicochemical, cytotoxicity and bacterial load evaluation. Journal of Water Process Engineering, v. 40, p. 101859, 2021. https://doi.org/10.1016/j.jwpe.2020.101859
    » https://doi.org/10.1016/j.jwpe.2020.101859
  • VILLASEÑOR-BASULTO, D. L.; ASTUDILLO-SÁNCHEZ, P. D.; REAL-OLVERA, J.; BANDALA, E. R. Wastewater treatment using Moringa oleifera Lam seeds: A review. Journal of Water Process Engineering, v. 23, p. 151-164, 2018. https://doi.org/10.1016/j.jwpe.2018.03.017
    » https://doi.org/10.1016/j.jwpe.2018.03.017
  • YAGUB, M. T.; SEN, T. K.; AFROZE, S.; ANG, H. M. Dye and its removal from aqueous solution by adsorption: a review. Advances in Colloid and Interface Science, v. 209, p. 172-184, 2014. https://doi.org/10.1016/j.cis.2014.04.002
    » https://doi.org/10.1016/j.cis.2014.04.002
  • WARHURST, A. M.; RAGGETT, S. L.; MCCONNACHIE, G. L.; POLLARD, S. J. T.; CHIPOFYA, V.; CODD, G.A. Adsorption of the cyanobacterial hepatotoxin microcystin-LR by a low-cost activated carbon from the seed husks of the pan-tropical tree, Moringa oleifera. Science of The Total Environment, v. 207, n. 2-3, p. 207-211, 1997. https://doi.org/10.1016/S0048-9697(97)00260-X
    » https://doi.org/10.1016/S0048-9697(97)00260-X

Publication Dates

  • Publication in this collection
    07 Mar 2022
  • Date of issue
    2022

History

  • Received
    04 Nov 2021
  • Accepted
    11 Jan 2022
Instituto de Pesquisas Ambientais em Bacias Hidrográficas Instituto de Pesquisas Ambientais em Bacias Hidrográficas (IPABHi), Estrada Mun. Dr. José Luis Cembranelli, 5000, Taubaté, SP, Brasil, CEP 12081-010 - Taubaté - SP - Brazil
E-mail: ambi.agua@gmail.com