The contamination of water resources by heavy metals (HMs), mainly introduced by anthropogenic resources, has been considered as a serious environmental issue in the recent era. Therefore, the conventional methods for heavy metal removal have various drawbacks in terms of cost, toxicity, and efficiency. The novel study investigated the biosorption intensity of raw and chemically modified sawdust (SD) and corn husk (CH) to eliminate chromium (III) from the aqueous solutions. SD and CH were used as biosorbents and were treated chemically with sulfuric acid (H2SO4), sodium hydroxide (NaOH), and detergent powder. The biosorption potential was estimated based on percentage removal efficiency (% RE) of chromium (III) and the adsorption intensity (qmax). The characterization of surface morphology and functional groups of biosorbents was studied by SEM and FTIR. Moreover, the adsorption isotherm models (Langmuir and Freundlich), kinetic models (pseudo-first-order kinetics and pseudo-second-order kinetics), and thermodynamic parameters (ΔG°, ΔH°, and ΔS°) were also assessed to predict the biosorption process. The results of % RE revealed that detergent-treated SD (DTSD) and detergent-treated CH (DTCH) showed highest % RE of 9.27 ± 0.15 (%) and 99.16 ± 0.08 (%) for elimination of chromium (III), respectively. Similarly, base-treated SD (BTSD) and base-treated CH (BTCH) exhibited 95.53 ± 0.18 (%) and 92.43 ± 0.22 (%) of % RE compared with 77.87 ± 1.64 (%) and 81.96 ± 0.34 (%) removal efficiency of acid-treated SD (ATSD) and acid-treated CH (ATCH), respectively. The raw SD (RSD) and raw CH (RCH) showed lower % RE of 23.68 ± 1.52 (%) and 35.52 ± 4.74 (%) for Cr (III) removal. Further, the porosity on the biosorbent's surface and attached functional groups increased after chemical treatment and this variation resulted in increased removal efficiency after chemical treatments. The Langmuir isotherm model and the Freundlich isotherm model were employed to predict the biosorption process, and both these models are best-fitted. The pseudo-second-order kinetic model was promising representative of the biosorption data. The process was endothermic and the parameters indicated that Cr (III) biosorbtion was thermodynamically favorable. Therefore, the obtained results highlighted that detergent-treatment enhanced the % RE, and DTSD and DTCH are highly efficient biosorbents for eliminating Cr (III) from aqueous solutions. Thus, detergent-treated biosorbents are proved to be a promising sustainable, eco-friendly, and cost-effective way to treat HMs from aqueous solutions as they utilize natural agriculture waste products and also handle issues related to excessive solid waste management.

  • Cr (III) ions from wastewater were efficiently removed using raw and chemically treated sawdust and corn husk as a biosorbent.

  • Cr (III) ions biosorption on the detergent treated sawdust and corn husk is effective, rapid and spontaneous.

  • The detergent-treated sawdust and corn husk efficiently removed 99.27 ± 0.15 (%) and 99.16 ± 0.08 (%) of Cr (III) ions in 120 min.

Graphical Abstract

Graphical Abstract
Graphical Abstract
World Health Organization

WHO

National Environmental Quality Standards

NEQS

United States Environmental Protection Agency

US EPA

heavy metals

HMs

sawdust

SD

corn husk

CH

Sustainable Development Study Center

SDSC

Center for Advanced Studies in Physics

CASP

raw sawdust

RSD

raw corn husk

RCH

acid-treated sawdust

ATSD

acid-treated corn husk

ATCH

base-treated sawdust

BTSD

base-treated corn husk

BTCH

detergent-treated sawdust

DTSD

detergent-treated corn husk

DTCH

atomic absorption spectroscopy

AAS

scanning electron microscope

SEM

Fourier transform infrared

FTIR

percentage removal efficiency

% RE

pseudo-first-order

PFO

pseudo-second-order

PSO

Environmental defilement has been considered as a cardinal issue accelerating with the rise in global world population. Water-related illnesses are also increasing and about 1.1% of world population does not have access to safe water for drinking reported by United Nations association report (Muhaisen 2017). Heavy metals (HMs) contamination in water is becoming a serious concern and gaining attention for the last decade (Berenjkar et al. 2019; Kou et al. 2020). The increasing expansion in industrialization and urbanization resulted into vast quantity of release of heavy metals in the environment (Sewwandi et al. 2012). HMs such as copper (Cu), lead (Pb), nickel (Ni), chromium (Cr), arsenic (As), cadmium (Cd), zinc (Zn), aluminum (Al), and mercury (Hg) enter water from various industrial sources. They are highly destructive due to their persistence and non-biodegradability in the environment. Thus, the elimination of HMs is necessary to lessen their threatening effects due to their toxic nature, recalcitrant behavior, and bioaccumulation in living organisms (Ahmed et al. 2019; Chen et al. 2019).

Among the HMs, chromium (Cr) is a substantial environmental concern and abundant metal used in different industrial processes (Miretzky & Cirelli 2010; Vinodhini & Das 2010). Chromium, in the environment, exists in six oxidation states, but the most significant ones are the trivalent chromium Cr (III) and hexavalent chromium Cr (VI) (Miretzky & Cirelli 2010). Chromium (III) is 300 times less deadly in nature than chromium (VI), and its minute quantity is essential for both plants and animals’ normal metabolic functioning. However, Cr (III) can cause health hazards (Miretzky & Cirelli 2010). Cr (III) mainly enters from weathering of Cr containing rocks, industrial discharge, leather tanning industry, electroplating, alloys, and processing units. In water, the pH is the main factor for Cr (III) solubility. Under basic pH, Cr (III) precipitates out, and at acidic pH, it solubilizes. Cr accumulates mainly in roots and shoots of plants and bio-accumulates in human beings (Kimbrough et al. 1999). According to World Health Organization (WHO) guidelines and National Environmental Quality Standards (NEQS) of Pakistan, the allowable limit for chromium in portable water is 0.05 μg/ L. Similarly, in the United States Environmental Protection Agency (US EPA) guideline, the permissible limit for chromium in drinking water is 0.1 μg/L (Raza et al. 2017).

Different treatment technologies can remove HMs from wastewater. However, each treatment has its limitations in terms of installment and maintenance cost, sludge production, removal efficiency, ions specificity, and toxicity. So suitable treatments for the removal of HMs from wastewater should be required (Salman et al. 2015; Al-Qodah et al. 2017). Among these, biosorption utilizes dead or inactive biomass for HMs removal from wastewater. Biosorption is termed as ‘the capability of biological matter to uptake HMs from aqueous solutions via metabolically mediated (by the utilization of ATP) or rapid physicochemical ways of assimilation (not utilizing ATP), or the ability of certain forms of inactive, non-living biomass that adsorb HMs from wastewater’ (Volesky 2003; Joo et al. 2010; Shamim 2018).

The major advantage of bio-adsorption is that it is based on natural organic biological materials, which have the properties of sequestering HMs and reducing the concentration of HMs from ppm to ppb (Lim & Aris 2014; Salman et al. 2015), and it is easily available and easy to process. Furthermore, agriculture waste products are considered an efficient bio-sorbent because these waste products are enriched with lignocellulosic materials and are found to be an inexpensive, eco-friendly, and readily available alternative material for HMs removal (Arunkumar et al. 2014).

The mechanism of biosorption includes the physical binding (electrostatic force of attraction or Van der Waals force of attraction) or chemical binding (exchange of ions) of the metal ions, chelation, complexation, reduction, and precipitation (Kanamarlapudi et al. 2018). The adsorption properties of biosorbents can be enhanced by chemical modifications. The chemical modification increases the functional groups attached to the surface of biosorbents, which help to sequester HMs and increase the biosorption capacity (Pothan et al. 2002; Hasar 2003).

Sawdust (SD) is produced by sawmills when the wood is processed for different purposes. The wooden powder, which is considered as a residue, is known as SD. In addition, SD is used to warm houses as a minor fuel (Singh et al. 2011). Corn husk (CH) is the leafy part of corn that is peeled off from the corn ear. It is different from a leaf in respect to structure and function. It is considered a valuable bio-sorbent because of its adsorption capacity for HMs and dyes. Rafatullah et al. (2009) experimented, the biosorption of Cu (II), Ni (II), Pb (II), and Cr (III) by using SD and investigated the effects of operational parameters and kinetics studies. Gunatilake (2016) studies the elimination of Cr (III) by using biochar produced from SD and rice husk. Similarly, Yu et al. (2001) performed analysis for eradication of Cu and Pb by using maple sawdust and showed 90% removal efficiency. A study for the elimination of Cr (VI), Pb (II), Cu (II), and Hg (II) used sawdust. Another examination was carried by Lin et al. (2019) showed corn husk (CH) had bio-sorption ability for mercury (Hg). Although many studies have utilized SD and CH to treat Cr (III) from wastewater, the usage of detergent powder for activation of SD and CH has not been investigated yet. Therefore, this study aims to explore the influence of administering chemical modifications to the SD and CH on enhancing the biosorption of Cr (III) from the aqueous solutions. The chemical treatment utilized sulfuric acid (H2SO4), sodium hydroxide (NaOH), and detergent powder. The study aimed at: (a) Removal of Cr (III) from wastewater to reduce its harmful impacts on the public, aquatic life, and environment. (b) Chemical treatment of SD and CH with an acid, a base, and a detergent powder to activate attached functional groups. (c) Characterization of biosorption capacity of sawdust and corn husk, before and after treatment with chemicals (d) Effects of operating parameters (pH, contact time, biosorbent dose, and temperature). (e) Estimating comparative isotherm models to understand the biosorption process. (f) Estimating the adsorption kinetic models and thermodynamic parameters for elimination of Cr (III) from aqueous solution.

The schematic overview of the entire procedure was adopted by El-Saied et al. (2017) and Duru et al. (2019), but due to changed environmental conditions, the process was modified to some extent.

Adsorbents

An SD sample was collected from a saw mill located in a small town named Bahoru, district Nankana Sahib, Pakistan. A CH sample was collected from a street vendor of corn in Muslim Town, Lahore, Pakistan. Both these samples were carried to the laboratory of Sustainable Development Study Center (SDSC) of Government College University, Lahore, for further processing.

Preparation of raw biosorbents

SD and corn husk CH were used as biosorbent materials, as both were readily available. Both SD and CH samples were thoroughly rinsed with deionized water many times to take off impurities and dust particles. After that, the samples were oven-dried (Oven Memmert UM200, Germany, available at the laboratory) at 70 °C for 24 hours to remove excess moisture. In addition, the biosorbents were allowed to chill at room temperature and crushed into a fine powder using a grinder (IKA, Grinder MF 10B, Germany, available at the laboratory). The fine powder of SD and CH was then sieved through a small particle size of 4 mesh size. Finally, these raw samples were stored in dry plastic jars separately, and these samples were ready for further batch-scale processing.

Preparation of chemically modified biosorbents

The fine powder of SD and CH was chemically treated with H2SO4 (1M), NaOH (0.5M), and detergent powder (0.5 g per gram of biosorbent). A 500 mL of an acid (H2SO4), base (NaOH), and detergent were assorted with 10 g of biosorbents in a 1,000 mL beaker. After that, the beakers were subjected to a hot plate (DAIHAN Scientific Co., Ltd, Hotplate Stirrer, MSH-20D, Korea, available at the laboratory) attached with a magnetic stirrer at a mixing speed of 300 rpm for 2 hours at 25 °C. The solution was then filtered and washed several times with deionized water. Further, it was oven-dried (Oven Memmert UM200, Germany, available at the laboratory) at 50 °C for 16 hours and stored for further analysis.

Preparation of stock solutions

All the chemical materials used in this experiment were analytical grade. CrCl3, high concentrated H2SO4, pure NaOH, HNO3, HCl, and NaNO3, and detergent powder (ionic) were used for the preparation of the stock solution. The stock chromium (III) solution of 1,000 ppm was produced by mixing 3.1 g of chromium chloride (CrCl3) in 1,000 mL distilled water. Further, stock solutions were placed in a plastic bottle and labeled, and placed in the fridge for further use.

Experimental procedure

Batch-scale experiments were performed to remove Cr (III) from the aqueous solution by using raw SD (RSD), acid-treated SD (ATSD), base-treated SD (BTSD), detergent-treated SD (DTSD), raw CH (RCH), acid-treated CH (ATCH), base-treated CH (BTCH), and detergent-treated CH (DTCH). In an Erlenmeyer flask, a 100 ml solution containing Cr (III) ions with an initial concentration of 513.7230 mg/L was assorted with 1 g of biosorbents. Then, these flasks were allowed to mix thoroughly in an incubator shaker at 300 rpm. After 2 hours, the process was stopped, and the sample was filtered in a vacuum filtration with the help of Whatman's filter paper grade 1. To estimate the final concentration of Cr (III) in water suspension atomic absorption spectroscopy (AAS) (Thermo Scientific, iCE 3000 Series AA spectrometer, USA) was used. This whole experiment was repeated in triplicates.

Characterization of the biosorbents

The raw and chemically modified biosorbents were described by using scanning electron microscope (SEM) and Fourier transform infrared (FTIR). The SEM (JSM-6480, Tokyo, Japan, available at Center for Advanced Studies in Physics (CASP) of Government College University, Lahore) analysis was carried out to analyze the surface morphology (topography and composition) of raw and chemically modified SD and CH at 200 μm, 100 μm, and 50 μm. FTIR (JASCO, FT/IR-6600, USA, available at the laboratory) analysis was carried out in a spectral sequence of 4,000–400 cm−1 to characterize the attached surface functional groups of SD and CH.

Point of zero charge

The pH, at which the net charge on the surface of biosorbent becomes zero, is termed as the point of zero charge (PZC). In the biosorption of Cr (III) ions by using raw and chemically modified SD and CH, the PZC was determined by the mass titration method. Three solutions having a certain range of initial pH were formed by utilizing HNO3 and NaOH. However, the background electrolyte was NaNO3. A 25 ml of solution, having each initial pH value, was poured into six conical flasks and a distinct amount of biosorbents were added to it. A solid friction of 0.05, 0.1, 0.5, 1, 5 and 10% were prepared and placed in shaker for 24 hours. The final pH value was noted and a graph was plotted between the amount of biosorbents and final pH value. The PZC was noted and the surface of biosorbents has positive charge below this pH value and negative charge above this pH value (Chand et al. 2009).

Effects of operating parameters

A batch-scale experiment was performed to examine the impacts of various operational factors on the process of biosorption. Some of these parameters, such as pH, the time of contact, a dose of adsorbent, and temperature, were analyzed for the biosorption of Cr (III) ions present in an aqueous sample by utilizing RSD and RCH, and the optimum biosorption parameters were evaluated.

Effect of pH on biosorption

The impact of pH on biosorption of HM ions by RSD and RCH was estimated by adjusting the initial pH ranging from 0 to 8. The 100 mL aqueous solution having a 513.7230 mg/L of initial concentration of Cr (III) ions was used. The pH of the solutions was maintained by using 1.0 M HCl and 1.0 M NaOH. In a 500 ml flask, 1 g of biosorbents was added in 100 ml of the metal-containing aqueous solution. The solution was then placed in a magnetic stirrer at 300 rpm for 2 hours at 25 °C. Then, the solution was strained, and the concentration of unadsorbed Cr (III) ions in the filtrate residual were estimated using an Atomic Absorption Spectrophotometer (Thermo Scientific, iCE 3000 Series AA spectrometer, USA).

Effect of contact time on biosorption

The impact of contact time on the biosorption of HM ions by RSD and RCH was estimated by studying the biosorption at different intervals of time (30 min, 60 min, 90 min, 120 min, 150 min, and 180 min). The 100 mL aqueous solution having a 513.7230 mg/L of initial concentration of Cr (III) ions was used. In a 500 ml flask, 1 g of biosorbent was added in 100 ml of the metal-containing aqueous solution. The mixture was then placed in a magnetic stirrer at 300 rpm for different intervals of time (30 min, 60 min, 90 min, 120 min, 150 min, and 180 min) at room temperature and optimum pH. Then, the solution was strained and the concentration of unadsorbed Cr (III) ions in the filtrate residual were estimated by using an atomic absorption spectrophotometer (Thermo Scientific, iCE 3000 Series AA spectrometer, USA).

Effect of adsorbent dose on biosorption

The impact of different amounts of RSD and RCH on the adsorption of HM ions was estimated by changing the amount of RSD and RCH. Different biosorbents dosages ranging from 0.25 g to 1.25 g were used in 100 ml of metal-containing solution. The 100 mL aqueous solution having a 513.7230 mg/L concentration of Cr (III) ions was used. In a 500 ml flask, a different dose of adsorbent was added in 100 ml of the metal-containing aqueous solution. The mixture was then placed in a magnetic stirrer at 300 rpm for 2 hours at room temperature and optimum pH. Then, the solution was strained and the concentration of unadsorbed chromium (III) ions in the filtrate residual were estimated by using atomic absorption spectrophotometer (Thermo Scientific, iCE 3000 Series AA spectrometer, USA).

Effect of temperature on biosorption

The impact of temperature on biosorption of HM ions by RSD and RCH was estimated by changing the temperature from 15° to 45 °C. The 100 mL aqueous solution having a 513.7230 mg/ L of initial concentration of Cr (III) ions was used. In a 500 ml flask, 1 g of biosorbents was added in 100 ml of the metal-containing aqueous solution. The mixture was then placed in a magnetic stirrer at 300 rpm at different temperatures for 2 hours and optimum pH. Then, the solution was strained and the concentration of unadsorbed metal ions in the filtrate residual was estimated by using atomic absorption spectrophotometer (Thermo Scientific, iCE 3000 Series AA spectrometer, USA).

Mathematical models for the process of biosorption

Mathematical manipulations of the results were manifested by estimating the % RE, adsorption isotherm models, kinetics models, and thermodynamic parameters for the biosorption of raw and chemically treated SD and CH.

Removal efficiency

The concentration of Cr (III) absorbed per gram of biosorbent at equilibrium was expressed by utilizing the following Equation (1):
formula
(1)
where Co is the initial amount of Cr (III) ions; Ce is the final amount in mg/L; V is the total volume used in liters; m is the amount of biosorbent in grams. The high value of biosorption intensity manifested an efficient process as it indicated that a lower concentration of biosorbent was used to remove metal ions from wastewater (Rafatullah et al. 2009).
The removal efficiency of Cr (III) was measured by using the following Equation (2):
formula
(2)
where Ci and Ce manifested the initial and final concentrations of metal ions in solution (mg/L) (Rafatullah et al. 2009).

Adsorption isotherm models

The initial concentrations of HM ions were used for modeling adsorption isotherms. The Langmuir isotherm model and the Freundlich isotherm model were commonly used, and data were fitted to these models. In addition, the amount of Cr (III) adsorbed on biosorbent was fitted graphically for comparing the biosorption capacity and percentage removal efficiency (% RE) of raw and chemically modified SD and CH.

The linear configuration of the Langmuir equation is presented in Equation (3):
formula
(3)
where, KL is Langmuir equilibrium constant measured in (L.mg−1); qmax is monolayer adsorption capacity expressed in mg.kg−1 (Rafatullah et al. 2009).
The linear configuration of the Freundlich isotherm equation is presented in Equation (4):
formula
(4)
where, KF is the Freundlich isotherm constant which manifested multilayer biosorption intensity; 1/n is the intensity of biosorption. The value of 1/n is an expression of acceptability when 0.1 < 1/n < 1 (Rafatullah et al. 2009).

Adsorption kinetic models

The adsorption kinetics modeling enabled studying the rate of biosorption with time (Kapur & Mondal 2014). To present the kinetics equation highlighting the biosorption of Cr (III) on raw and chemically modified SD and CH, two main kinetics models were employed to evaluate the experimental data. These models are pseudo-first-order kinetics (PFO) and pseudo-second-order kinetics (PSO).

PFO
PFO or the Lagergren-first-order equation is one of the most widely used kinetics model. The expression for the PFO kinetics model is represented in Equation (5):
formula
(5)
This kinetics equation was proposed by Lagergren. By administering the boundary conditions qt = 0 at t = 0 and qe = qt at t = t, the linearized expression of this equation is as shown by Equation (6):
formula
(6)

where the value of qt represents time of adsorption (min)(mg/g); k1 represents the first-order rate constant (1/min); qe represents the adsorption amount at equilibrium (mg/g) (Li et al. 2007; Kapur & Mondal 2014).

PSO
The expression of the PSO kinetics is described in Equation (7):
formula
(7)
After specific integration by administering the conditions qt = 0 at t = 0 and qt = qt at t = t, the linearized expression of this equation is shown in Equation (8):
formula
(8)

where the value of qt represents time of adsorption (min)(mg/g); k2 represents the second-order rate constant (g/mg/min); qe represents the adsorption amount at equilibrium (mg/g) (Li et al. 2007; Kapur & Mondal 2014).

Thermodynamics parameters

The thermodynamic parameters for the biosorption of Cr (III) ions by using raw and chemically treated SD and CH can be estimated by employing following equations:

The free energy change (ΔG°) was measured by Equation (9):
formula
(9)
Similarly, the change in enthalpy (ΔH°) was calculated by Equation (10):
formula
(10)
Further, the entropy (ΔS) was obtained from Equation (11):
formula
(11)
where k represents the equilibrium constant at temperature T; R represents the general gas constant (8.314 × 10−3 kJ/mol K) (Ajmal et al. 1998; Kapur & Mondal 2014).

Statistical analysis

Standard deviation

Standard deviation is a term used in statistics, which quantifies the dispersion of data concerning its mean value. It is measured as the square root of the variance by evaluating the difference in values from their mean value. If there is a greater difference in data points relative to their mean value, the data will be more spread out, and higher will be the standard deviation. The following Equation (12) calculates the standard deviation:
formula
(12)
where σ is standard deviation of sample values; N is the size of the dataset; xi represents each data value from the sample; μ is the mean of the dataset.

The characterization of SD and CH

SEM analysis of sawdust and corn husk

SEM analysis was employed to examine the surface morphology of RSD and pre-treated SD with acid, base, and detergent. The estimation of surface morphology helped to determine the areas of adsorption on SD. The greater the surface area, the higher will be the contact efficiency between pollutants to be removed and SD (Ateş & Özcan 2018).

Figure 1 represented SEM images RSD (a, b, c), ATSD (d, e, f), BTSD (g, h, i), and DTSD (j, k, l) at 200 μm, 100 μm, and 50 μm (magnification ranging from 80X to approximately 650X). At first sight, it was observed, the difference in the morphological structure of fibers before and after chemical treatments (Benyoucef et al. 2020). It was evident that RSD exhibited canal structure (Borhan et al. 2014) and was composed of bundles of various longitudinal sheets (Sewwandi et al. 2012). SD had an irregular and heterogeneous external structure with oriented sheets in the longitudinal direction. Further, the presence of tracheid on the surface and poorly developed tunnel-shaped (Gao et al. 2018) micropores was also observed in the images (M'hamdi et al. 2016).
Figure 1

SEM images of RSD (a, b, c), ATSD (d, e, f), BTSD (g, h, i) and DTSD (j, k, l) at 200, 100, and 50 μm.

Figure 1

SEM images of RSD (a, b, c), ATSD (d, e, f), BTSD (g, h, i) and DTSD (j, k, l) at 200, 100, and 50 μm.

Close modal

After modifications with an acid, base, and detergent, the surface of SD becomes smooth and shiny with irregular cavities. The pores, i.e. micropores and mesopores, appeared out onto the surface of the SD. The visibility of pores was due to the degradation of cellulose, lignin, and hemi-cellulosic material, and extermination of extractible material from SD, after chemical modification of the cellulosic structure. The mechanical properties of SD were enhanced after treatment with acid, base and detergent, and pores provide surface area for attachment of metal contaminants onto the surface of SD (Ahmad et al. 2016; M'hamdi et al. 2016; Benyoucef et al. 2020).

Figure 2 represents SEM images of RCH (a, b, c), ATCH (d, e, f), BTCH (g, h, i) and DTCH (j, k, l) at 200 μm, 100 μm, and 50 μm (magnification ranging from 80X to approximately 650X). At first sight, the fibers of CH showed remarkable differences in their morphological structures before and after chemical treatments (Benyoucef et al. 2020). It was observed from the figure that the surface of RCH fibers was irregular with grooves. The fibers exhibited an uneven, distorted, and honeycomb nest-like structure. The roughness and irregularity in the structure were due to waxes and were non-cellulosic on the outer surface, protecting the inner cellulosic material (Kambli et al. 2016; Herlina et al. 2018; Duru et al. 2021).
Figure 2

SEM images of RCH (a, b, c), ATCH (d, e, f), BTCH (g, h, i) and DTCH (j, k, l) at 200, 100 and 50 μm.

Figure 2

SEM images of RCH (a, b, c), ATCH (d, e, f), BTCH (g, h, i) and DTCH (j, k, l) at 200, 100 and 50 μm.

Close modal

After chemical treatment, the heterogeneity of the surface increased due to the removal of the non-cellulosic and small quantity of lignocellulosic material from the surface. The appearance of irregular cavities on the surface was due to excessive delignification of the fibers by NaOH. The fibers showed a smooth surface and hollow-shaped pores (i.e., both macropores and micropores) become visible on the surface. The appearance of both macropores and micropores enhanced the adhesive properties of CH fiber due to the removal of impurities of non-cellulosic material. Thus, the acid, base, and detergent treatment enhanced the adsorption ability of CH by increasing its surface area for attachment of contaminants present in the wastewater (Herlina et al. 2018; Duru et al. 2021).

FTIR analysis of SD and CH

FTIR spectroscopy assisted in identifying surface functional groups present on the exterior of the biosorbent (Zou et al. 2013). Figure 3(a)–3(d) represents the FTIR spectrum graph of RSD, ATSD, BTSD, and DTSD, respectively. The spectrum exhibited various absorption peaks that indicated SD's complex nature (Zou et al. 2013). The biosorption intensity of sawdust decreased after base and detergent treatment. However, the shape of the spectrum of SD before and after chemical treatment remained almost the same (M'hamdi et al. 2016). The adsorption peak is broad in the region located at 3,350 cm−1, which is attributed to the stretching vibration of bonded OH- groups of lignin and cellulose along with water molecules adsorbed on the surface of the SD (Zou et al. 2013; M'hamdi et al. 2016).
Figure 3

FTIR spectrum of (a) RSD (b) ATSD (c) BTSD and (d) DTSD.

Figure 3

FTIR spectrum of (a) RSD (b) ATSD (c) BTSD and (d) DTSD.

Close modal

The spectral peaks in the region around 2,910 cm−1 and 1,363 cm−1 represented the stretching and bending vibrations of the sp3 C-H bond of methyl groups of cellulose, respectively (Zou et al. 2013). A small spectral peak in the region of 2,360 cm−1 and 2,160 cm−1 represented the availability of C ≡ C and C ≡ N groups. Similarly, the spectral band that existed in 1,725 cm−1, characterized the stretching vibration of carboxylic acid (C = O) and esters. The presence of these groups is the indication of xylan, which is present in hemicellulose (M'hamdi et al. 2016). Further, the small spectral band present at 1,725 cm−1 and 1,610 cm−1 highlighted stretching vibrations of carbonyl groups, i.e., aldehyde and ketones (Zou et al. 2013). The band located at 1,507 cm−1 indicated the aromatic ring structure (C = O) deformation in lignin (M'hamdi et al. 2016). Further, the band spectrum from 1,545 cm−1 to 1,454 cm−1 highlighted the deformations of CH3 and CH2. A small spectral band at 1,425 cm−1 was the consequence of in-plane bending vibrations of HCH and COH functional groups. At 1,365 cm−1, the adsorption peaks characterized the deformation vibrations of C-H and O-H of alcoholic and phenolic groups linked with cellulose (Sewwandi et al. 2012). The spectral band saw at 1,316 cm−1, and 1,267 cm−1 are associated with vibrations of methoxy groups of lignin. The spectrum variation was due to the delignification of wood during activation with chemicals (M'hamdi et al. 2016). The spectrum's peak at region 1,026 cm−1 corresponded to the C-O stretching vibration modes of primary and secondary alcohols and C-O-C bonding groups. A small peak at 899 cm−1 highlighted the out-plane deformation of the C-H bond (M'hamdi et al. 2016).

The FTIR spectrum of RCH, ATCH, BTCH, and DTCH has been shown by Figure 4(a)–4(d), respectively. Figure 4(a)–4(d) represents all characteristics peaks attributed to cellulose, lignin, and hemicellulose present in corn husk sample (Singh et al. 2020). The FTIR analysis of RCH and ATCH exhibited roughly the same trends. Similarly, BTCH and DTCH samples had roughly the same trends in the spectrum. The differences in the spectrum were associated with the variances in chemical composition, specifically cellulose, hemicellulose, lignin, and other crude contents of proteins. A broad spectral peak was present in the area between 3,400 cm−1 and 3,270 cm−1, indicating the stretching vibration of hydrogen-bonded hydroxyl group and α-cellulose. This band also confirms the presence of water (Singh et al. 2020). However, the peak was not sharp in the case of DTCH due to an increase in α-cellulose content. The spectral band located at 2,926 cm−1 and 2,906 cm−1, represented the sp3 stretching vibration of alkyl C-H (Yeasmin & Mondal 2015).
Figure 4

FTIR spectrum of (a) RCH (b) ATCH (c) BTCH and (d) DTCH.

Figure 4

FTIR spectrum of (a) RCH (b) ATCH (c) BTCH and (d) DTCH.

Close modal

The spectral peak at 2,358 cm−1 highlighted the availability of C ≡ C and C ≡ N groups. This peak was more intense in DTCH samples because of unsaturation. Another spectral band near 1,734 cm−1 highlighted the presence of carbonyl stretching vibrations of carboxylic acid (C = O) and esters in hemicellulose (Singh et al. 2020). This spectral peak was less defined in acid and base-treated CH samples because of the removal of hemicellulose after chemical treatment (Singh et al. 2020). A small peak at 1,642 cm−1 represented the (C = O) stretching vibration of the conjugated carbonyl group of lignin (Singh et al. 2020). Finally, a sharp peak at 1,460 cm−1 indicated the stretching vibrations of the C = C group of aromatic skeletal lignin and deformation of the C-H group of lignin and hemicellulose (Kambli et al. 2017; Singh et al. 2020). This peak was more intense in ATCH but less in base-treated and detergent-treated CH samples. Similarly, at 1,423 cm−1, a spectral band corresponding to symmetric bending of CH2 group (Yeasmin & Mondal 2015; Kambli et al. 2016). However, this peak was much prominent in raw and acid-treated CH samples because of more cellulose content in both.

At 1,364 cm−1, the peak highlighted the in-plane bending of OH groups in cellulose and elongation of aliphatic C-H groups of cellulose and hemicellulose hemicellulose (Yeasmin & Mondal 2015; Singh et al. 2020). The spectral peak located in region 1,158 cm−1 described the asymmetrical stretching of C-O-C bonds in cellulose (Singh et al. 2020). A sharp spectral peak at 1,030 cm−1 corresponded to the C = O symmetrical elongation in alcohols present in lignin (Yeasmin & Mondal 2015; Herlina et al. 2018). The intensity of this peak was less in base-treated, and detergent-treated CH samples due to the removal of lignin. The spectral band existing at at 897 cm−1 highlighted the presence of C-O-C elongation vibrations of β (1–4) glycoside linkage of the cellulose (Yeasmin & Mondal 2015; Kambli et al. 2016).

PZC

The PZC value of SD and CH for biosorption of Cr (III) ions was estimated by mass titration method. The PZC of biosorbents were determined to be 3.75 and 4 for SD and CH, respectively. At pH equal to PZC the net charge on the surface of biosorbents is zero and there is no interaction between biosorbent and biosorbate. At lower pH than PZC, the surface of biosorbents is positively charged. While, at higher pH than PZC, the surface of biosorbents is negatively charged (Chand et al. 2009).

Effects of operating parameters

Effect of pH

The initial pH of the solution is regarded as a controlling factor in the biosorption of metal ions onto biosorbents surface. The main fact behind this is that protons present in solution behave as strong competing biosorbate ions. Moreover, pH also causes ionization of the biosorbents' surface functional groups and chemical speciation of the metallic ions present in the aqueous solution (Božić et al. 2009).

Through identifying the influence of pH on the biosorption of Cr (III) ions on raw and chemically modified SD and CH, biosorption batch-scale experiments were performed by utilizing raw biosorbents and also by changing initial pH of aqueous solution ranging from 0 to 8. The results obtained showed that the biosorption of Cr (III) ions onto RSD and RCH was pH-dependent. Figure 5 highlighted the biosorption of Cr (III) ions onto the exterior of RSD and RCH, and it was revealed that chromium (III) ions biosorption increases as the pH increases from 0 to 2 and then gradually decreased with an additional increase in pH. The maximum biosorption of Cr (III) was achieved at pH of 2.45 and 2.7 for SD and CH, respectively. Therefore, optimum pH for Cr (III) biosorption is 2.5 for SD and CH, respectively.
Figure 5

Effect of pH on the biosorption of Cr (III) using RSD and RCH.

Figure 5

Effect of pH on the biosorption of Cr (III) using RSD and RCH.

Close modal

This optimum pH is less than PZC and the surface of biosorbents becomes positively charged. In this regard, the biosorption behavior of HMs onto the biosorbent surface revealed that biosorption of HMs was based on the principle of ions exchange and hydrogen bonding (Rafatullah et al. 2009). At lower pH, the biosorption of HMs is low because of conflict between the metal ions and protons for the attachment on available sites. As the pH rises, the surface of SD and CH becomes less positively charged. So, the attractive electrostatic forces between the HMs and biosorbent surface start increasing and the biosorption capacity also increases (Kapur & Mondal 2014; Gunatilake 2016).

It can be observed from Figure 5 that pH has a strong influence on the biosorption capacity of SD and CH. The biosorption of Cr (III) ion exhibited a similar trend, that is, it increased as the pH increased By observing the results, it is confirmed that biosorption of metal ions obeyed the principle of ion exchange. At lower pH, the biosorption capacity of biosorbents is low because of higher amount of H+ ions present in the aqueous solution, which behave as a competitor for the HM ions and cause hindrance in the biosorption of HM ions onto the surface of biosorbents. At pH ≤ 1, the biosorption capacity of biosorbents is zero because of the presence of relatively high concentration of protons, which may lead to the occupation of the active sites and limit the biosorption of HM ions. At higher value of pH, the concentration of H+ in the aqueous solution starts decreasing, and the available active binding sites onto the biosorbents surface shifted towards their dissociation condition and adsorb HM ions from the aqueous solution. This behavior is similar to the findings of Schiewer & Volesky (1997).

However, the increase in pH increases the biosorption capacity but up to certain pH, above that pH, the increase in pH causes a decrease in biosorption capacity. This is due to formation of metal hydroxides precipitates into the aqueous solution, which settles down and biosorption capacity of HMs decreases. So, in case of Cr (III), the chromium hydroxide precipitates formed at pH greater than 4. Therefore, the optimum pH is less than pH at which precipitates of metal hydroxides formed (Rafatullah et al. 2009).

Effect of contact time

Figure 6 reveals the biosorption intensity of RSD and RCH for biosorption of Cr (III) ions concerning contact time. It is observed from Figure 6 that maximum removal efficiency was achieved in 120 min and then, no significant change in Cr (III) removal was noticed. The biosorption plot manifests that the biosorption capacity for Cr (III) ion removal is initially high because of the availability of excessive active sites on the exterior of RSD and RCH for the attachment of these ions. As the rate of adsorption increases with time, the exhaustion of the available sites and repulsion of solute molecules and the bulk phase start slowed down the process of biosorption (Schiewer & Volesky 1997; Shukla et al. 2005; Božić et al. 2009; Rafatullah et al. 2009; Muhaisen 2017). So, the equilibrium time of 120 min is maintained for the biosorption of chromium (III) ions by using RSD and RCH as an adsorbent.
Figure 6

Effect of contact time on the biosorption of Cr (III) using RSD and RCH.

Figure 6

Effect of contact time on the biosorption of Cr (III) using RSD and RCH.

Close modal

Effect of adsorbent dose

The biosorption of chromium (III) ions onto the exterior of RSD and RCH also depends on the amount of biosorbent. Therefore, a batch-scale experiment was done at room temperature by changing the quantity of RSD and RCH from 0.25 g to 1.25 g, and the pH of the solution was kept at 2.5. Figure 7 described the adsorption capacity of RSD and RCH at varying quantities, respectively. The maximum biosorption of chromium (III) is achieved at a dose of 1 g/100 ml of a solution containing Cr (III) ions. The obtained results depicted that the biosorption of chromium (III) ions on RSD and RCH increased with the rise in biosorbent dose. This is because, at high biosorbent concentration, the availability of active sites for the biosorption of chromium (III) ions was also increased (Shukla et al. 2005; Rafatullah et al. 2009; Muhaisen 2017).
Figure 7

Effect of dose of adsorbents on the biosorption of Cr (III) using RSD and RCH.

Figure 7

Effect of dose of adsorbents on the biosorption of Cr (III) using RSD and RCH.

Close modal

Effect of temperature

Temperature is a significant factor for the biosorption of HMs onto biosorbent surfaces. For investigation of the impact of temperature on the biosorption of Cr (III) ions, a batch scale experiment was conducted at a temperature varying from 15 °C to 45 °C. Figure 8 revealed the effects of temperature on the biosorption of Cr (III) ions onto the exterior of RSD and RCH, respectively. It was observed from the figures that the biosorption of Cr (III) enhanced with the rise in temperature (Rafatullah et al. 2009). Therefore, the optimum temperature of 25 °C was noted for biosorption of Cr (III) ions onto the surface of RSD and RCH.
Figure 8

Effect of temperature on the adsorption of Cr (III) using RSD and RCH.

Figure 8

Effect of temperature on the adsorption of Cr (III) using RSD and RCH.

Close modal

Removal efficiency

The biosorption intensity of raw and chemically modified SD and CH for elimination of chromium (III) ions was calculated by Equation (2). Figure 9 represents the removal efficiency of a raw sample, acid-treated, base-treated, and detergent-treated samples of SD and CH for Cr (III) ions removal. It was evident from Figure 9 that the raw sample of SD and CH has removal efficiency of 23.68 ± 1.52 (%) and 35.52 ± 4.74 (%), respectively. Similarly, the ATSD and ATCH samples have removal efficiency of 77.87 ± 1.64 (%) and 81.96 ± 0.34 (%), respectively. The removal efficiency of BTSD and BTCH sample was 95.53 ± 0.18 (%) and 92.43 ± 0.22 (%), respectively. The DTSD and DTCH samples showed % RE of 99.27 ± 0.15 (%) and 99.16 ± 0.08 (%), respectively. Thus, the results showed that both base-treated and detergent-treated SD and CH have higher adsorption capacity for Cr (III) ions. Table 1 explains a summary of the literature on using SD and CH to eliminate chromium ions from wastewater.
Table 1

A brief summary of literature on using sawdust and corn husk for removal of Cr (III) ions from wastewater

BiosorbentspHContact timeDose of adsorbentInitial conc. of metal ionsBest fitted modelsCharacterization of adsorbentsqmaxRemoval efficiency (%)Adsorption kineticsReferences
Sawdust (beech) 20 min 1 g 20–200 mg/L Langmuir – 41.86 mg/g – PSO  
Sawdust 1 h 0.2 g 5–50 mg/L Langmuir, Freundlich FTIR 5.52 mg/g – PSO Li et al. (2007)  
Sawdust (biochar) 5–7 1 h 1 g 20 mg/L Langmuir, Freundlich FTIR 43.48 mg/L 99% PSO Gunatilake (2016)  
Sawdust 20 min 200 mg – Freundlich, D-R sorption, Langmuir – 5.8 μmole/g 94% PFO Ahmad (2005)  
Sawdust 120 min 0.5 g – Langmuir SEM, FTIR 37.878 mg/g – PSO Rafatullah et al. (2009)  
Corn cob 120 min 0.1 g 100 mg/L Langmuir, Freundlich SEM, EXD 166.6 mg/g 79.8% PSO Manzoor et al. (2019)  
Sawdust (treated) 2.5 120 min 1 g 513.7230 mg/L Langmuir, Freundlich SEM, FTIR 4.69 mg/kg 99.27% PSO Current study 
Corn husk (treated) 2.5 120 min 1 g 513.7230 mg/L Langmuir, Freundlich SEM, FTIR 4.703 mg/kg 99.16% PSO Current study 
BiosorbentspHContact timeDose of adsorbentInitial conc. of metal ionsBest fitted modelsCharacterization of adsorbentsqmaxRemoval efficiency (%)Adsorption kineticsReferences
Sawdust (beech) 20 min 1 g 20–200 mg/L Langmuir – 41.86 mg/g – PSO  
Sawdust 1 h 0.2 g 5–50 mg/L Langmuir, Freundlich FTIR 5.52 mg/g – PSO Li et al. (2007)  
Sawdust (biochar) 5–7 1 h 1 g 20 mg/L Langmuir, Freundlich FTIR 43.48 mg/L 99% PSO Gunatilake (2016)  
Sawdust 20 min 200 mg – Freundlich, D-R sorption, Langmuir – 5.8 μmole/g 94% PFO Ahmad (2005)  
Sawdust 120 min 0.5 g – Langmuir SEM, FTIR 37.878 mg/g – PSO Rafatullah et al. (2009)  
Corn cob 120 min 0.1 g 100 mg/L Langmuir, Freundlich SEM, EXD 166.6 mg/g 79.8% PSO Manzoor et al. (2019)  
Sawdust (treated) 2.5 120 min 1 g 513.7230 mg/L Langmuir, Freundlich SEM, FTIR 4.69 mg/kg 99.27% PSO Current study 
Corn husk (treated) 2.5 120 min 1 g 513.7230 mg/L Langmuir, Freundlich SEM, FTIR 4.703 mg/kg 99.16% PSO Current study 
Figure 9

Percentage removal efficiency of Cr (III) biosorption by using raw and chemically modified SD and CH.

Figure 9

Percentage removal efficiency of Cr (III) biosorption by using raw and chemically modified SD and CH.

Close modal

The modifications of SD and CH by sulfuric acid, sodium hydroxide, and detergent powder can either increase or decrease % RE of Cr (III) ions compared with biosorption of raw biosorbents. The obtained results showed that the biosorption of Cr (III) ions was increased by using chemically modified SD and CH in comparison with untreated sawdust and corn husk. However, the H2SO4 treatment resulted in low adsorption capacity as compared to NaOH and detergent powder activation but high in comparison with raw biosorbents.

The acidic modification of biosorbents enhanced the acidic properties and hydrophilic character of biosorbents (Rehman et al. 2019; Abegunde et al. 2020). Gao et al. (2017) concluded that acid modification affects the lipophilic and hydrophobic character, the charge on the surface, porosity, and the oxygen-containing functional groups, which resulted in alternation in adsorption potential of biosorbents than raw biosorbents. Similarly, the treatment with an alkali or base involved the removal of the surface impurities such as waxes or natural fat. These surface impurities hide the adhesion ability of the biosorbents. The roughness on the surface also increases after removal of impurities by NaOH treatment which thus, exposes more surface –OH and other functional groups on the surface, which increases the adhesion properties (Wong et al. 2003; Ndazi et al. 2007; Abegunde et al. 2020).

Therefore, the % RE of acid treatment is less than that of base and detergent treatment because of the increase in positive charge on biosorbent surface and the excessive concentration of H+ ions in the solution, which cause hindrance in the biosorption of metal ions. Similarly, the base treatment has high % RE than acid treatment because of exposure of excessive active sites for metal binding but less than detergent treatment. The detergent treatment enhanced the adsorption capacity of biosorbents than acid and base treatment, due to the increase in surface porosity and exposure of more functional groups for the attachment of metal ions. Thus, the overall percentage removal efficiency of the biosorbents are: RSD and RCH < ATSD and ATCH < BTSD and BTCH < DTSD and DTCH.

Adsorption isotherms

The association between the amount of a specific adsorbate and its adsorption potential onto the exterior of a biosorbent can be evaluated by using adsorption isotherms (Rafatullah et al. 2009). The biosorption of HMs onto the biosorbent can be described by isotherms. These isotherms specify the scattering of HMs between two phases: liquid phase and solid phase (Desta 2013). To identify the biosorption potential of SD and CH for the biosorption of Cr (III) ions present in an aqueous sample, the Langmuir isotherm model and the Freundlich isotherm model were used. The Langmuir model specified the homogeneity of the surface of the biosorbents, while the Freundlich isotherm model specified the heterogeneity of the surface of the biosorbents.

The Langmuir isotherm model

This model considered that one molecular layer of adsorbate was developed onto the adsorbent surface, and it remained constant even the concentration of adsorbate was high (Gunatilake 2016). The Langmuir isotherm model assumed that a single biosorption film of moleculeswas formed on the exterior of the biosorbent and restricted active sites were present for biosorption (Desta 2013). Further, the biosorbent's surface was homogenous and had the same affinity for all the metal ions (Kapur & Mondal 2014). Therefore, the adsorption of metal ions at a site did not affect the adjacent site, and once a site was filled with metal ions, no further metal ion can attach to that site. In this way, a saturation point was reached where maximum adsorption of metal ions was achieved (Desta 2013).

The linear expression of the Langmuir equation was presented in Equation (3). In this equation, qmax is the single layer adsorption potential of the adsorbent. qmax is the maximum concentration of adsorbate which is absorbed on the exterior of the biosorbent. KL is the Langmuir constant linked with the adsorption energy. Ce is the equilibrium amount of HM ions in the aqueous sample and qe is the equilibrium amount of HM ions on the biosorbent (Gunatilake 2016).

The Langmuir isotherm modelwas tested by plotting a graph between Ce/qe vs Ce. The value of qmax and KL were obtained from the slope of the graph and the intercept, respectively. The value of R2, the Langmuir constant, and qmax were summarized in Table 2 for Cr (III) ions biosorption. The R2 value for Cr (III) ions biosorption on raw and chemically modified SD and CH ranged from 0.9999 to 1 (Table 2), which depicted that both adsorbents had the efficient potential for Cr (III) ions biosorption. The literature had more or less similar R2 values. The value of correlation coefficient R2 indicated that both raw and chemically modified SD and CH had good adsorption potential for Cr (III) and the Langmuir model was best fitted for both adsorbents.

Table 2

Adsorption isotherm model parameters and correlation coefficients for adsorption of Cr (III) ions on raw and chemically modified SD and CH

BiosorbentsLangmuir isotherm model
Freundlich isotherm model
qmaxKLR21/nKFR2
RSD 3.026634 0.009855 0.9999 3.2673 2.5303 0.9996 
ATSD 4.48430 0.34719 0.2794 1.6440 0.9994 
BTSD 4.670715 10.00467 0.047 1.5591 0.9999 
DTSD 4.694836 68.70968 0.0177 1.5517 0.9994 
RCH 3.557453 0.018424 0.9999 1.8887 2.1991 0.9978 
ATCH 4.531038 0.531423 0.2204 1.6209 
BTCH 4.640371 3.367188 0.0824 1.5701 
DTCH 4.703669 303.7143 0.0086 1.5502 0.9998 
BiosorbentsLangmuir isotherm model
Freundlich isotherm model
qmaxKLR21/nKFR2
RSD 3.026634 0.009855 0.9999 3.2673 2.5303 0.9996 
ATSD 4.48430 0.34719 0.2794 1.6440 0.9994 
BTSD 4.670715 10.00467 0.047 1.5591 0.9999 
DTSD 4.694836 68.70968 0.0177 1.5517 0.9994 
RCH 3.557453 0.018424 0.9999 1.8887 2.1991 0.9978 
ATCH 4.531038 0.531423 0.2204 1.6209 
BTCH 4.640371 3.367188 0.0824 1.5701 
DTCH 4.703669 303.7143 0.0086 1.5502 0.9998 

The Freundlich isotherm model

This model was applicable for non-ideal biosorption of metal ions onto the heterogeneous exterior of the biosorbent and multilayer biosorption (Rafatullah et al. 2009). The Freundlich isotherm model was an empirical expression incorporating the heterogeneous exterior of the biosorbent (Desta 2013) and also described the adsorption potential of the biosorbent for biosorbate (Gunatilake 2016).

The linear expression of the Freundlich isotherm model was presented by Equation (4). In this equation, KF was the Freundlich constant, which described the biosorption potential of biosorbent and 1/n is biosorption intensity. The value of 1/n and KF were dependent on the interaction of adsorbent and adsorbate and were calculated from the slope and the intercept of the graph. qe specified the concentration of adsorbate, which was absorbed in equilibrium per amount of biosorbent. Ce highlighted the equilibrium amount of HM ions present in an aqueous sample.

The Freundlich isotherm model was described by plotting a graph between log Ce vs log qe. The value of 1/n, KF, and linear regression coefficient (R2) were summarized in Table 2 for Cr (III) ions adsorption. The n specifies the degree of linearity between aqueous sample concentration and biosorption (Desta 2013; Muhaisen 2017). If n = 1, then the process of biosorption was linear. If n < 1, then the phenomena of biosorption were chemical. If n > 1, then the phenomena of biosorption were physical (Desta 2013; Gunatilake 2016). The value of n, which was obtained from the Freundlich equation for Cr (III) ions adsorption, ranges from 0.0177 to 3.2673 for raw and chemically treated sawdust and from 0.0086 to 1.8887 for raw and chemically treated corn husk (Table 2). The value of n was less than 1 for ATSD, BTSD, DTSD, ATCH, BTCH, and DTCH, which indicated that the phenomenon of adsorption was chemical. Only the value of n for biosorption of Cr (III) ions by using RSD and RCH was greater than 1 (Table 2), which showed that the phenomenon of adsorption was physical.

The value of correlation coefficients (R2) for biosorption of Cr (III) ranged from 0.9978 to 1 (Table 2), which highlighted that both biosorbents were suitable for Cr (III) biosorption. Therefore, for the value of R2, it is estimated that the Freundlich isotherm model was also suited for biosorption of Cr (III) by using raw and chemically modified SD and CH. Nearly literature had been reported with similar results. Therefore, both SD and CH had a good potential for biosorption of HM ions from aqueous samples.

Kinetics studies

The adsorption kinetics models were studied to investigate the change in metal adsorption with time and time required to reach the equilibrium. To present the kinetics equation highlighting the biosorption of Cr (III) on raw and chemically modified SD and CH, PFO and PSO models were used.

PFO kinetics model

The Lagergren-first-order kinetics equation was firstly given by Lagergren in 1898 (Rafatullah et al. 2009) and presented in Equation (5). When it was applied for boundary conditions qt = 0 at t = 0 and qt = qt at t = t and then it rearranged into linearized expression as shown in Equation (6). In this equation, qe described the amount of metal ions adsorbed per unit mass of biosorbent (mg/g), q describes the amount of metal ions adsorbed at any time (mg/g), and k1 is the rate constant of PFO kinetics model. The values of the adsorption kinetic parameters were obtained from the slope and intercept of the graph (Kapur & Mondal 2014).

The PFO kinetics model was described by plotting a graph between log (qeqt) vs t. The value of qe, k1 and R2 were summarized in Table 3 for the biosorption of Cr (III) ions by using raw and chemically treated SD and CH. In case of Cr (III) biosorption, the value of R2 lies between 0.7162 and 0.9983, and it depicted that PFO models have deficient or poor correlation coefficient for the data to best fit. Therefore, it can be summarized that PFO model had no satisfactory values and was not suitable for biosorption of Cr (III) ions using raw and chemically treated SD and CH.

Table 3

Adsorption kinetics model parameters and correlation coefficients for adsorption of Cr (III) ions on raw and chemically modified SD and CH

BiosorbentsPFO kinetics
PSO kinetics
qe (mg/g)K1 (min−1)R2qe (mg/g)K2 (g/mgmin)R2
RSD 13,601.84 −0.00013 0.7795 0.1704 1,7644.51 0.9793 
ATSD 17,744.84 −0.00021 0.9297 0.4285 3.9002 0.9997 
BTSD 705.6928 −8.1666 0.9892 0.2437 0.4186 
DTSD 515.2643 −8.4666 0.9295 0.5069 0.5733 
RCH 27,074.84 −0.00016 0.9983 0.2783 7919.721 0.9970 
ATCH 4,419.006 −0.00016 0.7162 0.4308 1.2247 0.9992 
BTCH 1,135.4 −0.00013 0.9501 0.4784 3.3339 
DTCH 261.8102 −0.00011 0.9199 0.5104 1.2228 
BiosorbentsPFO kinetics
PSO kinetics
qe (mg/g)K1 (min−1)R2qe (mg/g)K2 (g/mgmin)R2
RSD 13,601.84 −0.00013 0.7795 0.1704 1,7644.51 0.9793 
ATSD 17,744.84 −0.00021 0.9297 0.4285 3.9002 0.9997 
BTSD 705.6928 −8.1666 0.9892 0.2437 0.4186 
DTSD 515.2643 −8.4666 0.9295 0.5069 0.5733 
RCH 27,074.84 −0.00016 0.9983 0.2783 7919.721 0.9970 
ATCH 4,419.006 −0.00016 0.7162 0.4308 1.2247 0.9992 
BTCH 1,135.4 −0.00013 0.9501 0.4784 3.3339 
DTCH 261.8102 −0.00011 0.9199 0.5104 1.2228 

PSO kinetics model

The Lagergren-first-order equation was modified by Ho and McKay (Rafatullah et al. 2009) and the proposed form of represented in Equation (7). After specific integration by administering the conditions qt = 0 at t = 0 and qt = qt at t = t, the linearized expression of this equation was showed in Equation (8). In this equation, the value of qt represented time of adsorption (min)(mg/g), k2 represented the second-order rate constant (g/mg/min), qe represented the adsorption amount at equilibrium (mg/g). The values of kinetic parameters can be calculated by plotting a graph between t/q and t. the values of k2 and qe can be obtained from the slope and intercept of the plotted graph. The values of qe, k2, and R2 are summarized in Table 3 for the biosorption of Cr (III) by using raw and chemically treated SD and CH (Li et al. 2007; Rafatullah et al. 2009). The values of correlation coefficient R2 for PSO model was very high. The minimum value of R2 was 0.9793 and maximum was 1 which was ideal values for a perfect model. Therefore, PSO was the best-fitted kinetic model, and the biosorption of Cr (III) onto the surface of SD and CH perfectly obeyed PSO kinetic model.

Thermodynamic parameters

The values of thermodynamic parameters for the biosorption of Cr (III) by using raw and chemically treated SD and CH were calculated by Equations (9)–(11) and summarized in Table 4. The value of Gibb's free energy (ΔG°) was negative which was an indication of the feasibility and spontaneity of the biosorption process. The positive value of ΔH° is indication of the endothermic nature of reaction. While the positive value of entropy (ΔS°) indicated the increased randomness at the solid-solution interface during the biosorption of metal ions onto the active sites present onto the surface of biosorbents. It also showed that both biosorbents had high biosorption potential for Cr (III) ions (Ajmal et al. 1998; Rafatullah et al. 2009).

Table 4

Thermodynamics parameters for adsorption of Cr (III) ions on raw and chemically modified SD and CH

BiosorbentsΔG (KJ/mol)
ΔH (kJ/mol)ΔS (J/Kmol)
15 °C25 °C35 °C45 °C
RSD −7.9987 −8.55307 −8.98299 −9.66882 7.6327 54.2363 
ATSD −13.8388 −14.4766 −15.2956 −15.9782 7.0116 72.3068 
BTSD −18.2529 −19.0292 −19.7336 −20.5317 3.4542 75.3830 
DTSD −20.4776 −21.4733 −22.3712 −23.2414 6.0039 92.0609 
RCH −9.0316 −10.0604 −10.708 −11.2968 12.4569 75.0197 
ATCH −14.5941 −15.1678 −15.7263 −16.424 2.7994 60.3214 
BTCH −16.9353 −17.6419 −18.2679 −19.2891 5.1361 76.4671 
DTCH −22.2122 −23.3441 −24.2136 −25.5718 9.2518 109.1961 
BiosorbentsΔG (KJ/mol)
ΔH (kJ/mol)ΔS (J/Kmol)
15 °C25 °C35 °C45 °C
RSD −7.9987 −8.55307 −8.98299 −9.66882 7.6327 54.2363 
ATSD −13.8388 −14.4766 −15.2956 −15.9782 7.0116 72.3068 
BTSD −18.2529 −19.0292 −19.7336 −20.5317 3.4542 75.3830 
DTSD −20.4776 −21.4733 −22.3712 −23.2414 6.0039 92.0609 
RCH −9.0316 −10.0604 −10.708 −11.2968 12.4569 75.0197 
ATCH −14.5941 −15.1678 −15.7263 −16.424 2.7994 60.3214 
BTCH −16.9353 −17.6419 −18.2679 −19.2891 5.1361 76.4671 
DTCH −22.2122 −23.3441 −24.2136 −25.5718 9.2518 109.1961 

Practical implications of this study

SD and CH are useful biosorbents for the removal of HMs as they effectively remove Cr (III) ions from aqueous samples up to 99%. The process of biosorbtion by SD and CH is reliable, feasible and easy to process. So it will be helpful in designing a wastewater treatment plant on small and large scale for treatment of heavy metals for treatment of industrial and municipal wastewater. The chemical modifications of biosorbents are a successful approach and it increases the biosorption potential for HMs removal. This study also provides an efficient approach to industrialist, government, and private organizations to treat industrial and municipal wastewater with higher removal efficiency by using a variety of biosorbents with chemical modifications and also encourage the practical applications of biosorption techniques over other conventional methods. Furthermore, there is a need for continue research work in order to better understand the mechanism of biosorption, effects of different operating parameters, % RE and adsorption capacity of different biosorbents after modifications by a variety of chemicals. Extensive research work is required in regeneration of biosorbents and utilization of composites and mixtures of different biosorbents and estimation of their adsorption potential. Government and private organizations should abandon the conventional modes of treatments techniques and should encourage such approaches like biosorption in which country's own agricultural waste is utilized for the treatment of other waste, which in terms save many resources and investments, and benefit for the economy of the country.

SD and CH were used as biosorbents to eliminate Cr (III) from the aqueous sample. These biosorbents were chemically treated with H2SO4 (ATSD and ATCH), sodium hydroxide (BTSD and BTCH), and detergent powder (DTSD and DTCH). These raw and chemically treated adsorbents were employed to eliminate chromium (III). The results showed that among all biosorbents, detergents-treated biosorbents (DTSD and DTCH) have efficient removal efficiency up to 99%. The SEM examination of the surface revealed that the surface porosity has increased after chemical treatment. The FTIR analysis highlighted the presence of functional groups and showed that hydroxyl (OH-) and carbonyl groups were highly active functional groups present on the surface of adsorbents which had increased the efficiency for Cr (III) removal. The adsorption isotherm data expresses that the Langmuir and Freundlich isotherm models were best suited to Cr (III) ions biosorption by using raw and chemically modified SD and CH. The biosorption obeyed pseudo-second-order kinetic models and thermodynamic parameters indicated the spontaneity and feasibility of the process. The biosorption process is endothermic and both biosorbents had a greater affinity for biosorption of Cr (III) ions. The order of removal efficiency of raw and chemically modified SD and CH was DTSD and DTCH > BTSD and BTCH > ATSD and ATCH > RSD and RCH. Thus, DTSD and DTCH have a greater potential for the adsorption of HMs. It has dual benefits by removing the HMs from the aqueous solution in a reliable, affordable, and eco-friendly way. Further, the biosorption process utilized agricultural waste products, so it reduced the handling and disposal issues of solid waste management and achieve sustainability.

All relevant data are included in the paper or its Supplementary Information.

The authors declare there is no conflict.

Abegunde
S. M.
,
Idowu
K. S.
,
Adejuwon
O. M.
&
Adeyemi-Adejolu
T.
2020
A review on the influence of chemical modification on the performance of adsorbents
.
Res. Environ. Sustainability
.
Ahmad
Z. S.
,
Munaim
M. S. A.
&
Said
F. M.
2016
Characterization of meranti wood sawdust and removal of lignin content using pre-treatment process
. In:
Proceedings of the National Conference for Postgraduate Research (NCON-PGR 2016)
. pp.
598
606
.
Ahmed
A. S.
,
Sultana
S.
,
Habib
A.
,
Ullah
H.
,
Musa
N.
,
Rahman
M. M.
&
Sarker
M. S. I.
2019
Bioaccumulation of heavy metals in commercially important fish species from the tropical river estuary suggests higher potential child health risk than adults. bioRxiv, p. 681478.
Ajmal
M.
,
Khan
A. H.
,
Ahmad
S.
&
Ahmad
A.
1998
Role of sawdust in the removal of copper (II) from industrial wastes
.
Water Res.
32
(
10
),
3085
3091
.
Al-Qodah
Z.
,
Yahya
M. A.
&
Al-Shannag
M.
2017
On the performance of bioadsorption processes for heavy metal ions removal by low-cost agricultural and natural by-products bioadsorbent: a review
.
DSLNAH Water Treat.
85
,
339
357
.
Arunkumar
C.
,
Perumal
R.
,
Lakshmi
N.
&
Arunkumar
J.
2014
Use of corn cob as low cost adsorbent for the removal of nickel (II) from aqueous solution
.
Int. J. Adv. Biotechnol. Res.
5
(
3
),
325
330
.
Ateş
F.
&
Özcan
Ö.
2018
Preparation and characterization of activated carbon from poplar sawdust by chemical activation: comparison of different activating agents and carbonization temperature
.
Eur. J. Eng. Technol. Res.
3
(
11
),
6
11
.
Benyoucef
S.
,
Harrache
D.
,
Djaroud
S.
,
Sail
K.
,
Gallart-Mateu
D.
&
de la Guardia
M.
2020
Preparation and characterization of novel microstructure cellulosic sawdust material: application as potential adsorbent for wastewater treatment
.
Cellulose
27
(
14
),
8169
8180
.
Borhan
A.
,
Taha
M. F.
&
Hamzah
A. A.
2014
“Characterization of activated carbon from wood sawdust prepared via chemical activation using potassium hydroxide.”
. In:
Advanced Materials Research
vol. 832, pp. 132–137. Trans Tech Publications Ltd, 2014
.
Božić
D.
,
Stanković
V.
,
Gorgievski
M.
,
Bogdanović
G.
&
Kovačević
R.
2009
Adsorption of heavy metal ions by sawdust of deciduous trees
.
J. Hazard. Mater.
171
(
1–3
),
684
692
.
Chand
R.
,
Watari
T.
,
Inoue
K.
,
Luitel
H. N.
,
Torikai
T.
&
Yada
M.
2009
Chemical modification of carbonized wheat and barley straw using HNO3 and the adsorption of Cr (III)
.
J. Hazard. Mater.
167
(
1–3
),
319
324
.
Duru
C.
,
Nnabuchi
M.
&
Duru
I.
2019
Adsorption of Cu onto maize husk lignocellulose in single and binary Cu-Zn solution systems: equilibrium, isotherm, kinetic, thermodynamic and mechanistic studies
.
Egypt. J. Chem.
62
(
7
),
1295
1305
.
Duru
C. E.
,
Enedoh
M. C.
&
Duru
I. A.
2021
Surface modification of powdered maize husk with sodium hydroxide for enhanced adsorption of Pb (II) ions from aqueous solution
.
J. Environ. Treat. Tech.
9
(
1
),
95
104
.
El-Saied
F. A.
,
Abo-Elenan
S. A.
&
El-Shinawy
F. H.
2017
Removal of lead and copper ions from polluted aqueous solutions using nano-sawdust particles
.
Int. J. Waste Res.
7
(
217
),
305
.
Joo
J. H.
,
Hassan
S. H.
&
Oh
S. E.
2010
Comparative study of biosorption of Zn2+ by Pseudomonas aeruginosa and Bacillus cereus
.
Int. Biodeterior.
64
(
8
),
734
741
.
Kambli
N.
,
Basak
S.
,
Samanta
K. K.
&
Deshmukh
R. R.
2016
Extraction of natural cellulosic fibers from cornhusk and its physico-chemical properties
.
Fibers Polym.
17
(
5
),
687
694
.
Kambli
N. D.
,
Mageshwaran
V.
,
Patil
P. G.
,
Saxena
S.
&
Deshmukh
R. R.
2017
Synthesis and characterization of microcrystalline cellulose powder from corn husk fibres using bio-chemical route
.
Cellulose
24
(
12
),
5355
5369
.
Kanamarlapudi
S. L. R. K.
,
Chintalpudi
V. K.
&
Muddada
S.
2018
Application of biosorption for removal of heavy metals from wastewater
.
Biosorption
18
,
69
.
Kimbrough
D. E.
,
Cohen
Y.
,
Winer
A. M.
,
Creelman
L.
&
Mabuni
C.
1999
A critical assessment of chromium in the environment
.
Crit. Rev. Environ. Sci. Technol.
29
(
1
),
1
46
.
Kou
Y.
,
Zhao
Q.
,
Cheng
Y.
,
Wu
Y.
,
Dou
W.
&
Ren
X.
2020
Removal of heavy metals in sludge via joint EDTA-acid treatment: effects on seed germination
.
Sci. Total Environ.
707
,
135866
.
Lim
A. P.
&
Aris
A. Z.
2014
A review on economically adsorbents on heavy metals removal in water and wastewater
.
Rev. Environ. Sci. BioTechnol.
13
(
2
),
163
181
.
Lin
G.
,
Hu
T.
,
Wang
S.
,
Xie
T.
,
Zhang
L.
,
Cheng
S.
,
Fu
L.
&
Xiong
C.
2019
Selective removal behavior and mechanism of trace Hg (II) using modified corn husk leaves
.
Chemosphere
225
,
65
72
.
Manzoor
Q.
,
Sajid
A.
,
Hussain
T.
,
Iqbal
M.
,
Abbas
M.
&
Nisar
J.
2019
Efficiency of immobilized Zea mays biomass for the adsorption of chromium from simulated media and tannery wastewater
.
J. Mater. Res. Technol.
8
(
1
),
75
86
.
M'hamdi
A. I.
,
Kandri
N. I.
,
Zerouale
A.
,
Blumberga
D.
&
Gusca
J.
2016
Treatment and physicochemical characterisation of red wood sawdust
.
Energy Procedia
95
,
546
550
.
Muhaisen
L. F.
2017
Nickel ions removal from aqueous solutions using sawdust as adsorbent: equilibrium, kinetic and thermodynamic studies
.
J. Eng. Sustain. Dev.
21
(
03
),
60
71
.
Ndazi
B. S.
,
Karlsson
S.
,
Tesha
J. V.
&
Nyahumwa
C. W.
2007
Chemical and physical modifications of rice husks for use as composite panels
.
Composites Part A: Appl. Sci. Manufact.
38
(
3
),
925
935
.
Pothan
L. A.
,
Bellman
C.
,
Kailas
L.
&
Thomas
S.
2002
Influence of chemical treatments on the electrokinetic properties of cellulose fibres
.
J. Adhes. Sci. Technol.
16
(
2
),
157
178
.
Rafatullah
M.
,
Sulaiman
O.
,
Hashim
R.
&
Ahmad
A.
2009
Adsorption of copper (II), chromium (III), nickel (II) and lead (II) ions from aqueous solutions by meranti sawdust
.
J. Hazard. Mater.
170
(
2–3
),
969
977
.
Raza
M.
,
Hussain
F.
,
Lee
J. Y.
,
Shakoor
M. B.
&
Kwon
K. D.
2017
Groundwater status in Pakistan: a review of contamination, health risks, and potential needs
.
Crit. Rev. Environ. Sci. Technol.
47
(
18
),
1713
1762
.
Sewwandi
B. G. N.
,
Vithanage
M.
,
Wijesekara
S. S. R. M. D. H. R.
,
Rajapaksha
A. U.
,
Jayarathna
D. G. L. M.
&
Mowjood
M. I. M.
2012
Characterization of aqueous Pb (II) and Cd (II) biosorption on native and chemically modified Alstonia macrophylla saw dust
.
Bioremedia. J.
16
(
2
),
113
124
.
Shamim
S.
2018
Biosorption of heavy metals
.
Biosorption
2
,
21
49
.
Shukla
S. S.
,
Yu
L. J.
,
Dorris
K. L.
&
Shukla
A.
2005
Removal of nickel from aqueous solutions by sawdust
.
J. Hazard. Mater.
121
(
1–3
),
243
246
.
Singh
R.
,
Gautam
N.
,
Mishra
A.
&
Gupta
R.
2011
Heavy metals and living systems: an overview
.
Indian J. Pharmacol.
43
(
3
),
246
.
Singh
G.
,
Jose
S.
,
Kaur
D.
&
Soun
B.
2020
Extraction and characterization of corn leaf fiber
.
J. Nat. Fibers
19
(
5
),
1
11
.
Volesky
B.
2003
Biosorption process simulation tools
.
Hydrometallurgy
71
(
1–2
),
179
190
.
Wong
K. K.
,
Lee
C. K.
,
Low
K. S.
&
Haron
M. J.
2003
Removal of Cu and Pb by tartaric acid modified rice husk from aqueous solutions
.
Chemosphere
50
(
1
),
23
28
.
This is an Open Access article distributed under the terms of the Creative Commons Attribution Licence (CC BY 4.0), which permits copying, adaptation and redistribution, provided the original work is properly cited (http://creativecommons.org/licenses/by/4.0/).