Open access peer-reviewed chapter

Sugarcane Bagasse As Potentially Low-Cost Biosorbent

Written By

Honória de Fátima Gorgulho, Viviane Vasques da Silva Guilharduci and Patrícia Benedini Martelli

Submitted: 23 May 2017 Reviewed: 02 November 2017 Published: 20 December 2017

DOI: 10.5772/intechopen.72153

From the Edited Volume

Sugarcane - Technology and Research

Edited by Alexandre Bosco de Oliveira

Chapter metrics overview

1,552 Chapter Downloads

View Full Metrics

Abstract

Sugarcane bagasse (SB) is one of the major residues obtained from agriculture, every year millions of tons of SB have been produced by the sugarcane agribusiness. This abundant residue has been showed potential as biosorbent in wastewater treatment. SB, in nature or chemically modified, has been widely reported as a promising sorbent for the removal of dyes or heavy metals from aqueous medium. The application of SB in oil sorption is rarer, especially for the treatment of used motor oil wastewater. However, in this chapter, we show that this material has good oil sorption capacity when compared to other commercial and natural sorbents. This study evaluates the effect of several coupling agents over SB in used motor oil sorption as well as the influence of surfactant in this process.

Keywords

  • engine washing wastewater
  • oil sorption
  • sugarcane bagasse
  • dye sorption
  • biomass

1. Introduction

Agricultural waste by-products, such as sugarcane bagasse (SB), rice husk, coconut husk, sisal, and so on, have been extensively studied as a potential sorbent material for removing contaminants from water and wastewater [1]. These materials have many advantages such as the abundance, low cost, floatability, good flexibility and mechanical strength, and environmentally friendly properties. Besides, the potential use of several agricultural by-products as biosorbent is supported by its native adsorption capacity derived from their main constituents such as cellulose, hemicellulose, and lignin. These are polymeric structures with high content of hydroxyl and carboxyl groups, which have a strong influence in the adsorption capacity of different chemicals present in the aqueous medium.

SB, in nature or chemically modified, has been reported as a potential renewable sorbent for wastewater treatment [2, 3]. Millions of tons of SB have been generated every year by the sugarcane agribusiness, encouraging its reuse and recycling. The sugarcane industry is based mainly on the production of sugar and ethanol, which generates huge volumes of SB and sugarcane trash [4, 5]. After sugarcane is milled for juice extraction, bagasse is obtained as a residue, which corresponds to about 25% of the total weight, and is composed of approximately 40% cellulose, 24% hemicellulose, and 25% lignin. The hydroxyl groups are the most abundant and reactive sites in this biopolymer and are used to attach a variety of functional groups [6]. The higher content of cellulose in the SB biomass favors its hydrophilicity, which improves its interaction with cationic species in aqueous medium. As a result, SB has been widely used for the removal of heavy metals and dyes from wastewater [7, 8, 9].

1.1. Sugarcane bagasse as a sorbent for heavy metals removal

The utilization of unmodified or modified SB as an adsorbent have been described as a cheaper and effective technology for the removal of metal ions from wastewater [10, 11]. The metal ion-binding mechanism of adsorption on SB is attributed to its abundance of hydroxyl groups from cellulose, in which aqueous medium favors ion exchange or complexation with metal ions. Batch studies [12] using natural SB as a sorbent for removal of Cd(II) show the maximum adsorption at pH 6. The pH dependence of Cd(II) uptake was linked to both the surface functional groups and the metal ion species predominant in aqueous solution. The species Cd2+ and Cd(OH)+ are predominant at pH lower than 6, while the groups on surface are protonated and cannot bind to metal ions in solution. Besides, at very low pH, the surface groups are associated with the hydronium ions (H3O+), negatively affecting the interaction with the metal cations. When the pH increases, the surface affinity with the metal also increases, and adsorption is improved.

The metal ion-binding capacity of SB can be intensified by the introduction of surface groups with capacity chelating as carboxylate or amine [13, 14, 15]. The introduction of carboxylic functions (−COOH) on cellulosic fiber, lignin, and hemicellulose can be performed via cyclic anhydride reaction. Figure 1 displays an example of succinylation [16] reaction as an alternative route to attach carboxylic acid onto the cellulose.

Figure 1.

Scheme of possible synthetic route for introduction of carboxylic groups.

Pereira et al. describe the chemical modification of SB by ethylenediaminetetraacetic (EDTA) dianhydride (EDTAD) in order to improve Zn2+ adsorption [14]. The EDTA molecule is a chelating group that enhanced the metal complexation on fiber surface. Other compounds such as citric acid and phthalic anhydrides have also been used for SB modification, resulting then in an increase of adsorption capacity for these fibers [9, 13, 17, 18]. Table 1 summarizes some studies where SB in nature or modified form is used to remove heavy metals from aqueous environment.

Metal ion SB/modification Sorption capacity (mg g−1) References
Cd2+ SB without modification 69.06 [12]
Zn2+ Ethylenediaminetetraacetic dianhydride 105.26 [14]
45.45
Pb2+ Without modification 26.67 [17]
Pleurotus ostreatus (U2–11) 36.31
Lentinula edodes (U6–1), 27.68
Basidiomycetes 39.93
Ganoderma lucidum (U12–6) 36.00
Co2+ Trimellitic anhydride 1.153 [3]
Cu2+ 0.979
Ni2+ 0.849
Cu2+ Tetraethylenepentamine 0.016 [18]
Cu2+ Succinic anhydride 114 [15]
Cd2+ 196
Pb2+ 189
Co2+ Phthalic anhydride 0.561 [3]
Cu2+ 0.935
Ni2+ 0.932
Cr3+ Without modification 16.21 [13]
NaOH 26.41
Citric acid 17.2
NaOH/acylation with citric acid 31.28

Table 1.

SB in nature or modified as metal sorbent reported in the studies.

1.2. Sugarcane bagasse as an adsorbent for removal of dyes

With the expansion of textile sector, dyes in wastewater have become a serious environmental problem. Dyes are organic chemical compounds that appear colored due to the presence of chromophore groups such as nitrous, azo, and carbonyl [19, 20]. The release of dye waste into water bodies affects the life in aquatic environments, causing the ruining of soils and poisoning of drinking water. Besides, dyes cannot be removed by conventional treatment methods, and are resistant to aerobic digestion. As an alternative method, physical removal of dyes from effluent through biosorption has been extensively studied [8, 21]. A dye molecule is characterized by the presence of chromophore groups, which are responsible for producing the color, and also by groups known as auxochromes such as carboxylic acid, sulfonic acid, amino, and hydroxyl groups. These auxochromes are responsible for impacting or shifting of a particular color when attached to a chromophore, and also used to influence the dye solubility. In fact, the auxochromes are important to enhance the affinity of the dye toward the fibers. As a result, natural fiber, as SB, presents greater potential to remove dyes from wastewater.

Basically, dyes can be classified as cationic or anionic. Figure 2 shows, for example, the molecules of erythrosin B (EB) and methylene blue (MB). Cationic dyes carry a positive charge in their molecule, while anionic dyes carry a negative charge [22, 23, 24]. In aqueous solution, the dye molecules will present positively or negatively charged as a function of pH, and the electrostatic interaction with the fiber surface will direct the adsorption process. Therefore, the dyes adsorption route by SB can be described in a similar way as observed for metal ions. At low pH, the surface of SB became positively charged and the cationic dye adsorption will decrease, while for anionic dyes the reverse process occurs. In contrast, at high pH, the cationic dye removal will increase because the surface appears negatively charged and the anionic dye adsorption became inhibited. Table 2 summarizes some results of adsorption of dyes using SB, natural or modified, as a sorbent.

Figure 2.

Example of cationic (erythrosin B) and anionic (methylene blue) dyes.

Dyes SB/modification Sorption capacity (mg g−1) References
Crystal violet Oxidation with H3PO4-NaNO2 1018.2 [8]
Auramine O 571.8
Methylene blue Maleic anhydride 30.4 [25]
Rhodamine B Without modification 51.3 [2]
Methylene blue 28.0
Acid Alizarin Violet N Without modification 20.8 [26]
Methyl red Phosphoric acid 11.0 [21]
Without modification 5.7
Congo red Without modification 39.8 [24]
Methylene blue Ethylenediaminetetraacetic dianhydride 115.3 [27]
Crystal violet Meldrum’s acid 552.7 [7]
Methylene blue Ethylenediaminetetraacetic dianhydride 192.3 [19]
Gentian violet 357.1
Methylene blue Succinic anhydride 434.3 [16]
Gentian violet 1133.7
Erythrosin B Without modification 333.3 [22]
Methylene blue 1000.0

Table 2.

Examples of dyes adsorption by SB.

Advertisement

2. Sugarcane bagasse-based sorbents for motor oil removal

Research involving oil sorbents was firstly encouraged by the great environmental accidents generated by oil spills at sea [28, 29, 30]. In these cases, the adsorption processes are more suitable to remove and recover the oil. The sorbent material facilitates a transformation from liquid to solid phase, and then oil can be removed together with the sorbent. The main characteristic of crude oil sorbent material is the hydrophobicity and oleophilicity in order to attract oil preferentially to water. However, the amount of sorbents added to an oil polluted environment is a critical factor, because the inappropriate and excessive use can present subsequent waste disposal problems. It is especially important when organic synthetic products are used as sorbents. Synthetic sorbents, as polypropylene, do not degrade and are very expensive. Therefore, agricultural waste by-products were firstly used as an alternative oil sorbent to replace the conventional and nondegradable sorbent used to clean up oil spills. These biosorbents are biodegradable, renewable, abundant, and low cost. Teas et al. [31] compared the oil sorption capacity of cellulose with the expanded perlite and polypropylene in artificial seawater containing crude oil. These authors observed that for crude oil, the sorption capacity of cellulose overtakes the other sorbents. When light cycle oil and light gas oil were used in artificial seawater, they observed lower sorption by cellulose in relation to polypropylene, but similar behavior to expand perlite. The oil sorption capacity of vegetal fibers observed by several authors has been attributed to the interaction with hydrophobic sites in the biomass. Lignocellulosic fibers contain both hydrophilic and hydrophobic groups; however, the cellulose structure has hydrophilic nature with excellent wettability. The chemical functionalization of cellulose can increase its hydrophobic character, which is possible by changing the hydrophilic groups, hydroxyl (−OH) in the raw coir cellulose to hydrophobic hydrocarbons [32, 33, 34]. The biomass acetylation has been extensively used to increase its oil sorption capacity. Sun et al. [35] observed that SB acetylated presents greater machine oil sorption capacity (13.5–20.2 g g1) than polypropylene fibers (10 g g−1). Table 3 summarizes other studies that have been demonstrated the potential sorbent of agricultural by-products to oil removal and the effect of biomass modification in the sorption capacity.

Adsorption/sorption conditions Sorbent/treatment Sorption capacity (g g−1) References
Crude oil in dry conditions Dacryodes edulis leaf Unmodified leaf 3.440 [33]
Acetylated with acetic anhydride 4.990
Raw cotton 30.5 [36]
Crude oil in distilled water Sisal (Agave sisalana) 3.0–6.4 [28]
Coir fiber (Cocos nucifera) 1.8–5.4
Sponge-gourd (Luffa cylindrica) 1.9–4.6
Crude oil in artificial seawater Mangrove barks Untreated bark 1.666 [37]
Treated with oleic acid 3.333
Treated with palmitic acid 3.333
Machine oil in distilled water Sugarcane bagasse Untreated bagasse 8.9 [35]
Acetylated with acetic anhydride 11.4–16.5
Sugarcane bagasse modified with acetic anhydride 13.5–20.2 [38]
Synthetic effluent of crude petroleum Sugarcane bagasse 6.65 [39]
Motor oil in water Natural wood fibers 33–43 [40]
Wastewater of used motor oil Natural wood fibers 5.56 [41]
Engine oil in dry conditions Ceiba pentandra (L.) Gaertn. fibers (Kapok) packing density 0.02 g mL−1) 47.4 [42]
Used engine oil in dry conditions 50.8

Table 3.

Examples of oil sorption by biomass.

In spite of crude oil sorption favored by hydrophobic surface, in the case of used motor oil, this behavior is not completely precise. The crude oil and fresh motor oil composition is based on a mixture of hydrocarbons, and hence present affinity for hydrophobic sorbent. However, the composition of used motor oil has been changed through thermal degradation and contamination from generated waste in the engine. This process leads to the formation of low-molecular weight compounds and oxidation products [43]. Recently, Guilharduci et al. [44] showed that introduction of hydrophilic groups on SB improve the oil sorption from engine washing wastewater. Besides, it was also verified that surfactant present in wastewater affect the motor oil sorption.

The biomass acylation is well established in the literature, and the method is based on the reaction of hydroxyl groups (−OH) of the fiber surface with acyl groups (RCO–). Considering that the reaction with fiber is a heterogeneous reaction, not all hydroxyl groups is esterified, and a nonuniform product may be obtained [32, 35, 38, 45]. In the case of SB, the acetylation with acetic anhydride should introduce the methyl group and increase hydrophobicity at surface. Other acylating agents such as maleic [46], phthalic [9], and succinic [16] anhydrides have been used to introduce carboxylic groups on fibers and increase its surface hydrophilicity.

Another class of coupling agents recognized as efficient to promote the fiber hydrophobicity or hydrophilicity are the silanes. They have been extensively used to modify the surface of natural fibers to produce composites with thermoplastics or thermosets [47]. The aminosilanes, as the aminopropyltriethoxysilane (APS), are the most reported coupling agents for natural fibers. Recently, APS was successful used to modify the SB surface in order to improve its hydrophilic properties, and increase its oils sorption capacity [44].

Despite the great demand for treatment of effluents containing used oil, there is a scarcity of literature on this problem. In this study, this subject is explored in order to verify the potential of SB fibers for oil wastewater treatment. The effect of several coupling agents over SB capacity for sorption of used motor oil is investigated as well as the influence of surfactant in this process.

2.1. Methodology

2.1.1. Reagents

Acetic anhydride (99%), maleic anhydride (96%), phthalic anhydride (99%), succinic anhydride (97%), and aminopropyltriethoxysilane (APS) (97%) were purchased from Sigma-Aldrich and used without further purification. The other analytical grade reagents were obtained from Merck (Brazil) and were used as received. The used motor oil used in this study was obtained from Retifica del Rei (São João del Rei, MG, Brazil). SB was supplied by Cachaça Coqueiro (Nazareno, MG, Brazil). The bagasse was washed repeatedly with distilled water to remove all the dirt particles. The washed fiber was then dried in an oven (Q314M, Quimis) at 60°C for 24 h under a flow of air. It was subsequently ground and sieved through 30 mesh sieves (TE648, Tecnal). The resulting natural fiber was designated as SBN and was used as the starting material to produce the modified SB.

2.1.2. Fiber modification

Thermal treatment of SBN at 200°C for 24 h was performed to obtain the sample SB-200. The SBN acylation was obtained based on previous reports [44]. Briefly, the fiber was firstly soaked in 10% NaOH solution (using a ratio of 1 g/1 mL) at 0°C for 2 h. The bagasse was then washed with Milli-Q water until neutral pH and dried at 60°C for 24 h. Following this procedure, 70 g of the cleaned bagasse was placed in a 1 L round-bottom flask containing 300 mL of acylating agents and 200 mL of acetic acid. The mixture was acidified by adding 1 mL of H2SO4 and maintained under agitation for 24 h at 60°C. At the end of this period, the solution was filtered and the product was rinsed first with ethanol and then with water until the pH reached around pH 7.0. After this procedure, the sample was dried at 60°C for 24 h. The acylating agents used were acetic anhydride, maleic anhydride, phthalic anhydride, and succinic anhydride, which produce the samples SB-Acet, SB-M, SB-P, and SB-S, respectively.

The procedure of SB silanization was based on previous report [44]. Basically, the SB is firstly washed with Milli-Q water at 60°C until no color was observed in the washed water, next the fibers are mixed with APS, using a 5/2 mass ratio of SBN/APS, and dispersed into 400 mL of acetone. The suspension was placed in a bottle with glass spheres and left on a roller-conveyor (TE500/1, Tecnal) for 24 h at 200 rpm. The excess of reagents were Soxhlet-extracted with acetone for 24 h. Subsequently, the fiber was dried in an oven at 60°C for 24 h, under a flow of air. The result sample was denominated as SB-APS.

2.1.3. Fibers characterization

Fourier transform infrared (FTIR) spectra were obtained in the range 400–4000 cm−1, using a Perkin-Elmer 1720 FTIR spectrometer. The samples were mixed with KBr (Merck, spectroscopy grade) in an approximate ratio of 10/1. The resulting mixture was pressed into pellets and the spectra were acquired using 300 scans with resolution of 4 cm−1. Determination of C and N contents was carried out with a Thermo Fisher Scientific, Flash 2000 Analyzer.

2.1.4. Sorption experiments

Sorption test was performed using a synthetic used motor oil/water mixture with an oil concentration of 10.0 g L−1. The used motor oil sorption was determined by immersing 1.0 g of fiber in 100 mL of the solution of oil/water mixture. After 24 h, the sorbent was removed, and dried for 24 h at room temperature. The oil sorption was calculated using the following equation:

Q = m f m i m i E1

where Q is the oil sorption capacity (g/g), mf is the total mass of dry sorbent (g) after sorption, and mi is the mass of dry sorbent before sorption (g). The water sorption was determined by the same procedure but using only water.

The experiments of oil sorption were also carried out in solutions containing sodium dodecylbenzenesulfonate in a proportion of 0.05–0.30% in the dispersion of used motor oil/water. The surfactant concentration was determined following the methodology established by the American Public Health Association (Water Environment Federation & APHA 2005).

2.2. Results and discussion

Table 4 summarizes the samples prepared in this study. Results from elemental analysis of nitrogen and carbon in each sample are also presented in Table 4. It is noticed that the highest content of nitrogen obtained for SB-APS in relation to SBN fiber, which is attributed to amino groups from APS.

Sample Treatment Elemental analysis
N% C%
SBN In nature 0.18 45.7
SB-200 24 h at 200°C 0.28 48.2
SB-Acet Modified with acetic anhydride Not detected 45.1
SB-M Modified with maleic anhydride 0.12 44.8
SB-P Modified with phthalic anhydride 0.14 46.1
SB-S Modified with succinic anhydride Not detected 43.1
SB-APS Modified with aminopropyltriethoxysilane 2.15 42.5

Table 4.

SB-based sorbents produced in this study.

The modification introduced on the SBN was evaluated by FTIR and results are shown in Figures 3 and 4. For SBN spectra, four peaks are clearly observed between 3600 and 3307 cm−1 characteristic of stretching vibration of hydroxyl groups, which are associated with lignocellulose structure. Upon acylation with acetic anhydride, the intensity of ─O─H absorption band decrease because of hydroxyl groups reduction after reaction. In the same region of spectra, for SB-M only a diffuse band, centered at 3387 cm−1, is observed, and can be associated to hydrogen bonding in the hydroxyl groups. Similar profile is observed for SB-P and SB-S, in which, it is possible to identify two peaks less intense, at around 3600 and 3186 cm−1, also attributed to hydrogen bonding in the hydroxyl groups. These results confirm the modification of surface, which decrease the free hydroxyl groups at fiber surface. Successful esterification can be supported by three important absorption band around 1736, 1367, and 1242 cm−1, correspondent to the stretching vibration of carbonyl groups (C═O), C─H stretching, and C─O stretching characteristic of ester molecules. These bands can be clearly observed for SB-Acet, SB-S, and SB-M, however, with less intensity for SB-P, which can be associated with possible lower acylation obtained for this sample.

Figure 3.

FTIR spectra of (a) SBN, (b) SB-Acet (c) SB-P, (d) SB-S, and (e) SB-M.

Figure 4.

FTIR spectra of (a) SBN and (b) SB-APS.

For sample SB-APS, the FTIR is very similar to SBN (Figure 4). The characteristics of absorption bands of ─Si─O─Si─ at 700, 1030, 1145, and 1187 cm−1 overlapping with groups present in the biomass surface hinder its identification [47]. However, in the region of 3600–3000 cm−1, the APS modification leads to emerging of strong band, centered at 3398 cm−1, attributed to hydrogen bonding from hydroxyl groups. This behavior suggesting that hydrophilic groups (─NH2) from APS could be interacting with free hydroxyl groups.

The effect of SB modifications over oil and water sorption capacity can be observed in Figure 5. It is possibly observed that samples SB-APS, SBN, and SB-200 show higher affinity with water and also with used motor oil. This result suggests that the introduction of hydrophilic groups on SBN improve the used motor oil sorption. Guilharduci et al. [44] showed that used motor oil presents more affinity to hydrophilic surface than for crude oil. As previously discussed in the introduction, this behavior can be attributed to the chemical differences between used motor oil and new motor oil. Used motor oil can contain sludge, metal residues, and various other contaminants. Therefore, the presence of polar amino end groups (NH2) from silane structure favors the interaction with the constituents of the used motor oil. In contrast, the acylation of SBN was not effective to improve affinity with the constituents of used motor oil resulting in decreased adsorption capacity.

Figure 5.

Oil and water sorption capacity for SB samples.

In spite of the higher oil sorption obtained for SB-APS (0.71 g g−1) in relation to SBN (0.60 g g−1), the difference was about 15%, which is relatively small. It is important to take in account that the fiber modification with silane groups is a costly and complex process. Considering the abundance, low cost and efficiency of SBN, this material presented the best cost benefit for use as sorbent of wastewater of used motor oil.

Previous studies showed that the surfactants in used motor oil wastewater can affect its sorption by natural fibers [44]. Surfactants are widely used in various industrial processes for oil recovery, because of their physicochemical characteristics for emulsification, dispersion, and solubilization [48]. They have the ability to reduce the oil-water interfacial tension, increase the capillarity, and change the wettability of the adsorbent surface. Then, in order to evaluate the impact of surfactants on the SBN sorption capacity, batch experiments were carried out using an anionic surfactant, sodium dodecylbenzenesulfonate, in a dispersion of oil (1.0 g)/water (0.1 L). The results are summarized in Figure 6(a, b).

Figure 6.

(a) The effect of anionic surfactant in the oil sorption and (b) surfactant sorption in dispersion with and without oil.

The used motor oil sorption shows slight increase (13%) in low concentration of surfactant, and afterward, the sorption decreases almost 42%. When the surfactant concentration increases, the hydrophobicity of fiber in aqueous solution decreased as a result of lower interfacial energy between water and fibers. In parallel, the oil-water interface is improved although micelles or microemulsions formation, which probably affects the oil sorption by SBN and improve surfactant interaction with the fiber. In this condition, the surfactant sorption is likely increased by SBN surface. This suggestion can be evaluated by the analyses of surfactant sorption in the water dispersion with and without oil. The results of these experiments are depicted in Figure 6(b). Based on this result, it is noticed that surfactant sorption not only increases when its concentration increases but also is enhanced in the oil/water dispersion.

Advertisement

3. Conclusion

In this chapter, the great potential of SB in the natural and chemically modified form is shown to be applied as sorbent material for oil removal from aqueous medium. The performance of SBN to remove used motor oil from wastewater was around 0.6 g g−1. This value increases around 13% in the presence of surfactant at 0.05–0.15% in water/oil dispersion. SB also showed efficient sorption of anionic surfactant in water dispersion with or without oil. The results support the use of SB as a natural sorbent as a substitute for commercial synthetic oil sorbents in the treatment of wastewater of used motor oil.

References

  1. 1. Nguyen TAH, Ngo HH, Guo WS, Zhang J, Liang S, Yue QY, Li Q, Nguyen TV. Applicability of agricultural waste and by-products for adsorptive removal of heavy metals from wastewater. Bioresource Technology. 2013;148:574-585
  2. 2. Zhang Z, O’Hara IM, Kent GA, Doherty WOS. Comparative study on adsorption of two cationic dyes by milled sugarcane bagasse. Industrial Crops and Products. 2013;42:41-49
  3. 3. Ramos SNC, Xavier ALP, Teodoro FS, Elias MMC, Goncalves FJ, Gil LF, Freitas RP, Gurgel LVA. Modeling mono- and multi-component adsorption of cobalt(II), copper(II), and nickel(II) metal ions from aqueous solution onto a new carboxylated sugarcane bagasse. Part I: Batch adsorption study. Industrial Crops and Products. 2015;74:357-371
  4. 4. Szczerbowski D, Pitarelo AP, Filho AZ, Ramo LP. Sugarcane biomass for biorefineries: Comparative composition of carbohydrate and non-carbohydrate components of bagasse and straw. Carbohydrate Polymers. 2014;114:95-101
  5. 5. Rezende CA, Lima MA, Maziero P, Azevedo ER. Chemical and morphological characterization of sugarcane bagasse submitted to a delignification process for enhanced enzymatic digestibility. Biotechnology for Biofuels. 2011;4:54-63
  6. 6. Chandel AK, Silva SS, Carvalho V, Singh OV. Sugarcane bagasse and leaves: Foreseeable biomass of biofuel and bio-products. Journal of Chemical Technology and Biotechnology. 2012:11-20
  7. 7. Ferreira BVS, Teodoro FS, Mageste AB, Gil LF, Freitas RP, Gurgel LVA. Application of a new carboxylate-functionalized sugarcane bagasse for adsorptive removal of crystal violet from aqueous solution: Kinetic, equilibrium and thermodynamic studies. Industrial Crops and Products. 2015;65:521-534
  8. 8. Martins LR, Rodrigues JAV, Adarme OFH, Melo TMS, Gurgel LVA, Gil LF. Optimization of cellulose and sugarcane bagasse oxidation: Application for adsorptive removal of crystal violet and auramine-O from aqueous solution. Journal of Colloid and Interface Science. 2017;494:223-241
  9. 9. Ramos SNC, Xavier ALP, Teodoro FS, Elias MMC, Goncalves FJ, Gil LF, Freitas RP, Gurgel LVA. Removal of cobalt(II), copper(II), and nickel(II) ions from aqueous solutions using phthalate-functionalized sugarcane bagasse: Mono- and multicomponent adsorption in batch mode. Industrial Crops and Products. 2016;79:116-130
  10. 10. Putra WP, Kamari A, Yusoff SNM, Ishak CF, Mohamed A, Hashim N. Biosorption of Cu(II), Pb(II) and Zn(II) ions from aqueous solutions using selected waste materials: Adsorption and characterisation studies. Journal of Encapsulation and Adsorption Sciences. 2014;4:25-35
  11. 11. Nguyen TAH, Ngo HH, Guo WS, Zhang J, Liang S, Yue QY, Li Q, Nguyen TV. Applicability of agricultural waste and by-products for adsorptive removal of heavy metals from wastewater. Bioresource Technology. 2013;148:574-585
  12. 12. Garg U, Kaur MP, Jawa GK, Sud D, Garg VK. Removal of cadmium(II) from aqueous solutions by adsorption on agricultural waste biomass. Journal of Hazardous Materials. 2008;154:1149-1157
  13. 13. Santos VCG, Salvado ADPA, Dragunski DC, DNC P. Highly improved chromium(III) uptake capacity in modified sugarcane bagasse using different chemical treatments. Quimica Nova. 2012;35(8):1606-1611
  14. 14. Pereira FV, Gurgel LVA, Gil LF. Removal of Zn2+ from aqueous single metal solutions and electroplating wastewater with wood sawdust and sugarcane bagasse modified with EDTA dianhydride (EDTAD). Journal of Hazardous Materials. 2010;176:856-863
  15. 15. Karnitz O, Gurgel LVA, Melo JCP, Botaro VR, Melo TMS, Gil RPF, Gil LF. Adsorption of heavy metal ion from aqueous single metal solution by chemically modified sugarcane bagasse. Bioresource Technology. 2007;98:1291-1297
  16. 16. Gusmão GAG, Gurgel LVA, Melo TMS, Gil LF. Application of succinylated sugarcane bagasse as adsorbent to remove methylene blue and gentian violet from aqueous solutions—Kinetic and equilibrium studies. Dyes and Pigments. 2012;92:967-974
  17. 17. Palin D Jr, Rufato KB, Linde GA, Colauto NB. Evaluation of Pb(II) biosorption utilizing sugarcane bagasse colonized by Basidiomycetes. Environmental Monitoring and Assessment. 2016;188(279):3-14
  18. 18. Chen J, Yu J, Wang F, Tang J, Zhang Y, Xu Y, Chi R. Selective adsorption and recycle of Cu2+ from aqueous solution by modified sugarcane bagasse under dynamic condition. Environmental Science and Pollution Research. 2017;24:9202-9209. DOI: 10.1007/s11356-017-8608-2
  19. 19. Gusmão GAG, Gurgel LVA, Melo TMS, Gil LF. Adsorption studies of methylene blue and gentian violet on sugarcane bagasse modified with EDTA dianhydride (EDTAD) in aqueous solutions: Kinetic and equilibrium aspects. Journal of Environmental Management. 2013;118:135-143
  20. 20. Salleh MAM, Mahmoud DK, Karim WAWA, Idris A. Cationic and anionic dye adsorption by agricultural solid wastes: A comprehensive review. Desalination. 2011;280:1-13
  21. 21. Saad SA, Isa KM, Bahari R. Chemically modified sugarcane bagasse as a potentially low-cost biosorbent for dye removal. Desalination. 2010;264:123-128
  22. 22. Sharma P, Kaur H. Sugarcane bagasse for the removal of erythrosin B and methylene blue from aqueous waste. Applied Water Science. 2011;1:135-145
  23. 23. Tahir H, Sultan M, Akhtar N, Hameed U, Abid T. Application of natural and modified sugar cane bagasse for the removal of dye from aqueous solution. Journal of Saudi Chemical Society. 2016;20:S115-S121
  24. 24. Zhang Z, Moghaddam L, O’Hara IM, Doherty WOS. Congo red adsorption by ball-milled sugarcane bagasse. Chemical Engineering Journal. 2011;178:122-128
  25. 25. Ge M, Du M, Zheng L, Wang B, Zhou X, Jia Z, Hu G, Alam SMJ. A maleic anhydride grafted sugarcane bagasse adsorbent and its performance on the removal of methylene blue from related wastewater. Materials Chemistry and Physics. 2017;192:147-155
  26. 26. Mitter EK, Santos GC, Almeida EJR, Morão LG, Rodrigues HDP, Corso CR. Analysis of acid alizarin violet N dye: Removal using sugarcane bagasse as adsorbent. Water, Air, and Soil Pollution. 2012;223:765-770. DOI: 10.1007/s11270-011-0900-z
  27. 27. Xing Y, Liu D, Zhang L-P. Enhanced adsorption of methylene blue by EDTAD-modified sugarcane bagasse and photocatalytic regeneration of the adsorbent. Desalination. 2010;259:187-191
  28. 28. Annunciado TR, Sydenstricker THD, Amico SC. Experimental investigation of various vegetable fibers as sorbent materials for oil spills. Marine Pollution Bulletin. 2005;50:1340-1346
  29. 29. Deschamps G, Cruel H, Borredon ME, Albasi C, Vignoles C. Oil removal from water by sorption on hydrophobic cotton fibers. 2. Study of sorption properties in dynamic mode. Environmental Science & Technology. 2003;37:5034-5039
  30. 30. Chol HM, Cloud RM. Natural sorbents in oil spill cleanup. Environmental Science & Technology. 1992;26:772-776
  31. 31. Teas C, Kalligeros S, Zanikos F, Stournas S, Lois E. Investigation of the effectiveness of absorbent materials in oil spills clean up. Desalination. 2001;140:259-264
  32. 32. Cerqueira DA, Filho GR, Meireles CS. Optimization of sugarcane bagasse cellulose acetylation. Carbohydrate Polymers. 2007;69:579-582
  33. 33. Nnaji NJN, Onuegbu TU, Edokwe O, Ezeh GC, Ngwu AP. An approach for the reuse of Dacryodes edulis leaf: Characterization, acetylation and crude oil sorption studies. Journal of Environmental Chemical Engineering. 2016;4:3205-3216
  34. 34. Olaru N, Olaru L, Vasile C, Ander P. Surface modified cellulose obtained by acetylation without solvents of bleached and unbleached kraft pulps. Polimery. 2011;56:11-12
  35. 35. Sun XF, Sun RC, Sun JX. A convenient acetylation of sugarcane bagasse using NBS as a catalyst for the preparation of oil sorption-active materials. Journal of Materials Science. 2003;38:3915-3923
  36. 36. Singh V, Kendall RJ, Hake K, Ramkumar S. Crude oil sorption by raw cotton. Industrial and Engineering Chemistry Research. 2013;52:6277-6281. DOI: dx.doi.org/10.1021/ie4005942
  37. 37. Asadpour R, Sapari NB, Hasnain M, Orj UK. Enhancing the hydrophobicity of mangrove bark by esterification for oil adsorption. Water Science & Technology. 2014:1220-1228. DOI: 10.2166/wst.2014.355
  38. 38. Sun XF, Sun RC, Sun JX. Acetylation of sugarcane bagasse using NBS as catalyst under mild reaction conditions for the production of oil sorption-active materials. Bio/Technology. 2004;95:343-350
  39. 39. Boni HT, Oliveira D, Souza AAU, Souza SMAGU. Bioadsorption by sugarcane bagasse for the reduction in oil and grease content in aqueous effluent. International Journal of Environmental Science and Technology. 2016;13:1169-1176. DOI: 10.1007/s13762-016-0962-y
  40. 40. Rajakovic V, Aleksic G, Rajakovic L. Governing factors for motor oil removal from water with different sorption materials. Journal of Hazardous Materials. 2008;154:558-563
  41. 41. Rajakovic V, Aleksic G, Radetic M, Rajakovic L. Efficiency of oil removal from real wastewater with different sorbent materials. Journal of Hazardous Materials. 2007;143:494-499
  42. 42. Abdullah MA, Rahmah AU, Man Z. Physicochemical and sorption characteristics of Malaysian Ceiba pentandra (L) Gaertn. as a natural oil sorbent. Journal of Hazardous Materials. 2010;177:683-691
  43. 43. Mohammed RR, Ibrahim IAR, Taha AH, McKay G. Waste lubricating oil treatment by extraction and adsorption. Chemical Engineering Journal. 2013;220:343-351
  44. 44. Guilharduci VVS, Martelli PB, Gorgulho HF. Efficiency of sugarcane bagasse-based sorbents for oil removal from engine washing wastewater. Water Science and Technology. 2017;75(1):173-181. DOI: 10.2166/wst.2016.476
  45. 45. Chung S, Suidan MT, Venosa AD. Partially acetylated sugarcane bagasse for wicking oil from contaminated wetlands. Chemical Engineering and Technology. 2011;34:1989-1996
  46. 46. Ge M, Du M, Zheng L, Wang B, Zhou X, Jia Z, Hu G, Alam J. A maleic anhydride grafted sugarcane bagasse adsorbent and its performance on the removal of methylene blue from related wastewater. Materials Chemistry and Physics. 2017;192:147-155
  47. 47. Xie Y, Hill CAS, Xiao Z, Militz H, Mai C. Silane coupling agents used for natural fiber/polymer composites: A review. Composites: Part A. 2010;41:806-819
  48. 48. Bera A, Kumar T, Ojha K, Mandal A. Adsorption of surfactants on sand surface in enhanced oil recovery: Isotherms, kinetics and thermodynamic studie. Applied Surface Science. 2013;284:87-99

Written By

Honória de Fátima Gorgulho, Viviane Vasques da Silva Guilharduci and Patrícia Benedini Martelli

Submitted: 23 May 2017 Reviewed: 02 November 2017 Published: 20 December 2017