You are on page 1of 35

Accepted Manuscript

Title: Residual biomass of chia seeds (Salvia hispanica) oil


extraction as low cost and eco-friendly biosorbent for effective
reactive yellow B2R textile dye removal: Characterization,
kinetic, thermodynamic and isotherm studies

Authors: Deborah Cristina Crominski da Silva, Juliana


Martins Teixeira de Abreu Pietrobelli

PII: S2213-3437(19)30131-9
DOI: https://doi.org/10.1016/j.jece.2019.103008
Article Number: 103008

Reference: JECE 103008

To appear in:

Received date: 23 October 2018


Revised date: 1 February 2019
Accepted date: 5 March 2019

Please cite this article as: da Silva DCC, de Abreu Pietrobelli JMT, Residual biomass of
chia seeds (Salvia hispanica) oil extraction as low cost and eco-friendly biosorbent
for effective reactive yellow B2R textile dye removal: Characterization, kinetic,
thermodynamic and isotherm studies, Journal of Environmental Chemical Engineering
(2019), https://doi.org/10.1016/j.jece.2019.103008

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Residual biomass of chia seeds (Salvia hispanica) oil extraction as low cost and eco-

friendly biosorbent for effective Reactive Yellow B2R textile dye removal:

Characterization, kinetic, thermodynamic and isotherm studies

Deborah Cristina Crominski da Silvaa,*, Juliana Martins Teixeira de Abreu Pietrobellia


a
Department of Chemical Engineering, Federal University of Technology – Paraná (UTFPR),

T
Av. Monteiro Lobato s/n- Km 4, Zip Code 84016-210 Ponta Grossa – PR – Brazil.

IP
*
Corresponding author. Email address: deborahcrominski@gmail.com.

R
SC
Highlights
U
N
 Residual chia-seed-oil-extraction biomass does not have current application.
A
 At 303 K, 40 and 74% of dye removal at 5 and 60 min, respectively.
M

 Biosorption process followed pseudo-second order model and Toth isotherm.

 Unloaded and dye-loaded biosorbent was characterized by FTIR and SEM.


ED

 Studies indicated possibility of application at room temperatures of 303 K.


E PT

Abstract: Residual chia-seed-oil-extraction biomass was studied as a biosorbent for removal


CC

of Reactive Yellow B2R textile dye from aqueous solutions in batch system to suggest an

appropriate and eco-friendly application, other than incineration or landfill. This residue does
A

not require previous treatment, biosorption process responded very well at temperature of 303

K and time up to 60 min, characterizing a highly practical application in textile wastewater

treatments. Biosorption process was concluded effective by kinetic, thermodynamic and

equilibrium studies. Efficiency was maximum at pH 2 and 150 rpm of agitation speed. Kinetic

studies presented equilibrium time of 60 hours and approximately 92% of dye removal from

1
aqueous solution, and pseudo-second order was the best fit. Toth adsorption isotherm

indicated the maximum biosorption capacity of 70.95 mg g-1 and the heterogeneity parameter

of 1.2785, both suggesting good pollutant removal ability by biosorbent and a slight degree of

heterogeneity of biosorbent surface, respectively, indicated also by the evaluation of the SEM

images of the unloaded and dye-loaded biosorbent surface. Biosorption process occurs

T
favorably, spontaneously and possibly through physical adsorption, indicating possible dye

IP
recovery and reuse of the biosorbent, as concluded by thermodynamic studies.

R
SC
Keywords: Biosorption; Residual chia-seed-oil-extraction biomass; Salvia hispanica;

Reactive Yellow B2R dye; Batch studies.

U
N
A
M

Nomenclature

Absi and Absf initial and final absorbance


ED

𝐪𝐭 adsorption capacity (mg g-1) at contact time 𝒕


𝒒𝒆𝒒 adsorbed amounts of dye (mg g-1) at equilibrium
𝑪𝒊 and 𝑪𝒕 dye concentration (mg L-1) at initial and 𝒕 time.
PT

𝒕 contact time (min)


𝑽 volume of dye solution (L)
E

𝒎 mass of biosorbent (g)


𝒌𝟏 pseudo-first order rate (min-1)
CC

𝒌𝟐 pseudo-second order rate (g mg-1 min-1)


𝛂 the initial adsorption rate (mg g-1 min-1)
𝛃 relationship between the degree of surface coverage and the activation
A

𝒌𝒊 intraparticle diffusion rate constant (mg g-1 min 0,5


)
energy involved in the chemosorption (g mg-1)
𝒙 value of the intersection of the line with the axis 𝒒𝒕 (mg g-1)
𝜟𝑯𝟎 enthalpy
𝜟𝑺𝟎 entropy
𝜟𝑮𝟎 Gibbs free energy
𝒌𝒅 sorption equilibrium constant
𝑨 Arrhenius constant

2
𝑬𝒂 activation energy
𝒒𝒎𝒂𝒙 monolayer biosorption capacity of the biosorbent (mg g-1)
𝒃 Langmuir equilibrium constant (L mg-1)
𝒌𝑭 Freundlich biosorption capacity (mg g-1)
𝒏𝑭 Freundlich biosorption intensity
𝒌𝒕 Temkin constant (L min-1)
𝑩 Temkin constant related to related to the heat of adsorption and the total
𝜺 Polany potencial
number of sites

T
𝜷𝑹𝑫 activity coefficient related to the average sorption energy (mol2 kJ-2)
𝒒𝒎á𝒙𝑹𝑫 maximum biosorption capacity (mg g-1)

IP
𝒏𝑻 heterogeneity parameter of Toth
𝒃𝑻 affinity constant of Toth (L min-1)

R
SC
1 Introduction

Chia seeds (Salvia hispanica) are composed of approximately 30% of fats and 19-27% of

U
proteins (Capitani et al., 2012). Due to the number of fats present in chia seeds, oil extraction
N
is performed for encapsulation and human consumption (Timilsena et al., 2016). According to
A

Capitani et al. (2012), the residual biomass generated in this process could be applied in
M

formulations for the food industry because of the high levels of fibers and proteins. However,
ED

no published studies have been found nor is there any knowledge about industries that use the

residual biomass. Thus, as the residue is usually discarded as a common industrial waste and
PT

the characteristics of the residual biomass have not yet been exploited, it does not have eco-

friendly final disposal.


E
CC

On the other hand, the application of dyes by the textile industry is responsible for

consumption of approximately 60% of the total production of synthetic dyes, in which 10 to


A

15% are discarded in wastewater after dyeing stage (Daneshvar et al., 2012). Therefore,

inefficient treatment of textile dyes in wastewater and posterior discharge in watercourses

causes imbalance in the biological cycle, given that the presence of coloration decreases the

reception of sunlight and affects photosynthesis of aquatic plants. Furthermore, the specific

occurrence of azo dye adds toxic properties, as it is carcinogenic, mutagenic, and hazardous to

3
living organisms (Verma et al., 2012). According to the manufacturer of the dye under study,

Reactive Yellow B2R textile dye is harmful to aquatic organisms, causing long-term adverse

effects on the aquatic environment. Thus, the disposal recommended by the AGS Chemicals

is incineration or landfill disposal.

Conventional effluent treatment processes may not apply to the removal of the synthetic

T
dyes as they are chemically stable. Synthetic dyes are efficiently removed using chemical

IP
processes, microbiological and enzymatic decomposition, and sorption processes (Forgacs et

R
al., 2004). Chemical processes as oxidation, ozonation, Fenton reaction, and UV irradiation,

SC
and microbiological processes as aerobic and anaerobic digestion, mixture of fungus and

bacteria, and enzyme degradation are often reported in the literature as a highly effective

U
method for dye removal, given that these techniques can be straightforward to apply, suitable
N
for several types of dyes and toxins, and reusable. However, they can be costly, require a
A
catalyst, produce toxic by-products, be useful for a specific type of dye only, among other
M

disadvantages (Katheresan et al., 2018). On the other hand, the sorption processes have been

presented as an alternative method for removal of a wide range of dyes (Vikrant et al., 2018),
ED

and they can be applied to purify wastewater and drinking water as well. The key
PT

disadvantage of the sorption processes is that the cost can be high, mainly due to the cost of

the sorbent itself (Katheresan et al., 2018). However, there is a possibility of applying a low-
E

cost sorbent, characterized by a material that requires little processing to be used, such as by-
CC

product or residue from industries (Bailey et al., 1999; Katheresan et al., 2018). In addition,

several sorbents have been used in the removal of dyes, such as activated carbon, residues
A

from agriculture, industry and natural materials (Gisi et al., 2016). In this sense, the

application of organic sorbents has taken an important step towards the cost-effective removal

of dyes over the years (Shahawy and Heikal, 2018). Consequently, the use of organic wastes

as biosorbent for dye removal has been severely reported in the literature, for instance malt

4
bagasse (Da Silva et al., 2019; Junchen et al., 2018; Fontana et al., 2016), banana, cucumber

and potato peels (Stavrinou et al., 2018), grape pomace waste (Oliveira et al., 2018), and

sugarcane bagasse (Scheufele et al., 2016).

Noticing the lack of proper application for the residual biomass of chia seed oil

extraction, the main objective of this study was to support the valorization of industrial waste

T
and encourage a circular economy by suggesting an application to residual biomass of chia

IP
seed oil extraction instead of just disposal. The residual biomass was applied as biosorbent for

R
the removal of Reactive Yellow B2R textile dye from aqueous solutions to promote an

SC
effective treatment of textile industrial effluents with minimal cost and reduce the

environmental impacts generated by waste disposal and industrial effluents. To achieve this

U
objective, biosorbent before and after biosorption was characterized by scanning electron
N
microscopy and Fourier transform infrared spectrometry, and adsorption kinetics, equilibrium
A
isotherms and thermodynamics studies were carried out in a batch system.
M

2 Materials and methods


ED

2.1 Instrumentation
PT

The residual chia-oil-extraction biomass was dried in an oven with forced air circulation

(SL 102, SOLAB), and sieved using the electromagnetic stirrer with granulometric sieves
E

(BERTEL, AAKER). The Fourier transform infrared spectroscopy (FTIR) absorption spectra
CC

were on the spectral range of 400 – 4000 cm-1 (VERTEX 70v, BRUKER). Biosorbent

morphology was evaluated using a scanning electron microscopy (VEGA 3 LMU, TESCAN).
A

The pH and weight measurements were performed using a pH meter (MPA- 210, AAKER)

and analytical balance (MARK, BEL ENGINEERING). Batch experiments were performed

using a metabolic bath (TE 420, TECNAL). Measurements of dye concentration in aqueous

solutions were achieved in a molecular absorption spectrometer UV-Vis (800XI, FEMTO) at

5
the wavelength of 439, 428, 418, and 411 nm for dye solution pH of 2.0, 3.0, 4.0 and above

5.0, respectively. Separation of the residual chia-seed-oil-extraction biomass from dye

aqueous solution was performed in a centrifuge (EXCELSA BABY II 206-R, FANEM).

2.2 Biosorbent and dye solution preparation

Residual chia-seed-oil-extraction biomass without previous treatment or processing, i.e.

T
as presented at the end of the industrial process, was obtained from H. Larocca Carbonar M.E.

IP
company, located in Castro city, Paraná state, Brazil. Biosorbent was dried at temperature of

R
303 K until constant mass. Reactive Yellow B2R textile dye solutions were prepared using

SC
purified water (18MΩcm-1 resistivity) obtained from a system (MASTER ALL, GEHAKA)

and pH adjustments were performed using HCl and NaOH at least of analytical grades at 0.1

mol L-1 of concentration.


U
N
2.3 Batch biosorption trials
A
Preliminary study was carried out to obtain the optimal biosorption operating parameters.
M

The influence of dye solution pH (1.0 – 12.0), agitation speed of metabolic bath (0 – 150 rpm)

and biosorbent particle diameter (0.065 – 0.925 mm and original residue without sieving step)
ED

were evaluated by means of final removal percentage of Reactive Yellow B2R dye from

aqueous solution, presented by Equation 1, in which Absi and Absf indicate the initial and
PT

final absorbance, respectively.


E

Absi -Absf
%Removal = 100 (1)
CC

Absi

Batch experiments were performed with 50 mL of 75 mg L-1 dye solution and 0.3 g of
A

residual chia-seed-oil-extraction biomass, applied as referenced in other works related to the

use of biomass as biosorbents (Da Silva et al., 2019; Juchen et al., 2018; Scheufele et al.,

2016; Honorio et al., 2015). Samples were placed in metabolic bath at temperature of 303 K

and contact time of 22 h. After contact time, the samples were centrifuged at 3000 rpm for 10

min and the initial (stock solution) and final (remaining dye solution) absorbance readings

6
were determined in molecular absorption spectrometer UV-Vis. All the experiments were

performed in triplicate.

2.4 Kinetic and thermodynamic studies

Experiments were performed using 50 mL of 75 mg L-1 dye solution at pH 2.0 and 0.3 g

of residual chia-seed-oil-extraction biomass. Samples were placed in a metabolic bath at 150

T
rpm at temperatures of 303, 313 and 323 K, collected at different times from 0 to 5760 min

IP
(96 h), centrifuged, and both stock and remaining dye solution absorbances were measured.

R
All the experiments were performed in triplicate. Equation 2 was used to obtain the adsorption

SC
capacity q t (mg g-1) at a contact time 𝑡 (min) from dye concentration (mg L-1) at initial 𝐶𝑖 and

final contact time 𝐶𝑡 , volume of dye solution 𝑉 (L) and mass of biosorbent 𝑚 (g).

q t = (𝐶𝑖 − 𝐶𝑡 ) 𝑉 ⁄𝑚 (2)
U
N
Pseudo-first order (Eq. 3) (Lagergren, 1898), pseudo-second order (Eq. 4) (Ho and
A
McKay, 1998), Elovich (Eq. 5) (Turner, 1974) and intraparticle diffusion (Eq. 6) (Weber and
M

Morris, 1962) kinetic linear equations were adjusted to the experimental data and the kinetic

parameters and correlation coefficients (R2) were obtained to verify which model represents
ED

the process.
PT

K1
log(q eq − q t ) = log(q eq ) − (2,303) t (3)

t 1 1
=K + q t (4)
E

qt 2
2 qeq eq
CC

1 1
q t = β ln(αβ) + β ln(t) (5)

q t = k i t 0,5 + x (6)
A

where 𝑞𝑡 were obtained from Eq. 2, 𝑞𝑒𝑞 is the adsorbed amounts of dye (mg g-1) at

equilibrium, 𝑘1 is the pseudo-first order rate (min-1), 𝑘2 is the pseudo-second order rate (g mg-
1
min-1), α and β stands for the initial adsorption rate (mg g-1 min-1) and the relationship

between the degree of surface coverage and the activation energy involved in the chemical

7
sorption (g mg-1), respectively, 𝑘𝑖 is the intraparticle diffusion rate constant (mg g-1 min0,5)

and 𝑥 is the value of the intersection of the line with the axis 𝑞𝑡 (mg g-1).

Adsorption thermodynamic parameters of enthalpy (𝛥𝐻 0 ), entropy (𝛥𝑆 0 ) and Gibbs free

energy (𝛥𝐺 0) allow the thermodynamic understanding of biosorption process and were

calculated from Gibbs energy equation as follows:

T
Δ𝐺 0 = Δ𝐻 0 − 𝑇Δ𝑆 0 (7)

IP
In addition, the Gibbs free energy variation of the adsorption process can be related to the

R
equilibrium constant, 𝑘𝑑 (Fontana et al., 2016), as shown in the following equations:

SC
Δ𝐺 0 = −𝑅 𝑇 ln 𝑘𝑑 (8)
𝑞𝑒𝑞
𝑘𝑑 = 𝐶 (9)
𝑒𝑞

U
The equilibrium constant and the entropy and enthalpy are related through the Van't Hoff
N
equation:
A

Δ𝐻 0 Δ𝑆 0
M

ln 𝑘𝑑 = − + (10)
𝑅𝑇 𝑅

Physical or chemical nature of the adsorption can be predicted from determination of the
ED

activation energy (Wu, 2007), described by Eq. 11.


𝐸𝑎
ln(𝑘) = ln(𝐴) −
PT

(11)
𝑅𝑇

where 𝑘 is the adsorption rate constant according to the fitted kinetic model, 𝐴 is the
E

Arrhenius constant and 𝐸𝑎 is the activation energy (J mol-1).


CC

2.5 Equilibrium studies


A

Equilibrium experiments were performed using 50 mL of dye solution at concentrations

from 50 to 675 mg L-1 at pH 2.0, pH in which the biosorption was optimal, and 0.3 g of

residual chia-seed-oil-extraction biomass. Samples were placed in a metabolic bath at 150

rpm and 303 K, collected at 60 h of contact time, centrifuged, and both stock and remaining

8
dye solution absorbances were measured. Equation 2 was used to obtain the adsorption

capacity (q t ).

The most applied adsorption isotherms in these circumstances are the Langmuir and

Freundlich isotherms, and both have a convex shape considered as favorable to the

biosorption process (Geankoplis, 2003). However, other unconventional models for

T
biosorption are also considered favorable and can be applied, such as Temkin, Dubinin-

IP
Radushkevich, and Toth isotherms, and each model has its own individualities. Adsorption

R
isotherm equations were applied to experimental data to obtain the equilibrium parameters

SC
and correlation coefficients (R2), and evaluate the biosorption process behavior.

The Langmuir model assumes that the biosorption process occurs only in purely specific

U
sites (Langmuir, 1918) and Eq. (12) describes the linear form of the Langmuir isotherm:
N
𝐶𝑒𝑞 1 1
= 𝑏𝑞 +𝑞 𝐶𝑒𝑞 (12)
𝑞𝑒𝑞
A
𝑚𝑎𝑥 𝑚𝑎𝑥

where 𝑞𝑚𝑎𝑥 is the monolayer biosorption capacity of the biosorbent (mg g-1) and 𝑏 is the
M

Langmuir equilibrium constant (L mg-1).


ED

The Freundlich model is an empirical equation that assumes that the adsorption process

takes place in multilayers and can be applied in non-ideal sorption systems with homogeneous
PT

sites (Freundlich, 1906). The linear form of the Freundlich isotherm is described by the

following equation:
E

log 𝑞𝑒𝑞 = log 𝑘𝐹 + 𝑛𝐹 log 𝐶𝑒𝑞 (13)


CC

where 𝑘𝐹 is the Freundlich biosorption capacity (mg g-1) and 𝑛𝐹 is the Freundlich biosorption
A

intensity.

In contrast with Langmuir and Freundlich models, the Temkin model considers the

effects of indirect interactions between sorbate and sorbent molecules, as well as relates the

adsorption energy of sorbate molecules and sorbent surface (Daneshvar et al., 2012). The

Temkin model is represented by the linear form:

9
𝑞𝑒𝑞 = 𝐵 ln 𝑘𝑡 + 𝐵 ln 𝐶𝑒𝑞 (14)

where 𝑘𝑡 (L min-1) and 𝐵 are Temkin constants, and the second parameter is related to the heat

of adsorption and the total number of sites.

The model developed by Dubinin and Radushkevich is based on the adsorption potential

previously proposed by Polany, which considers a heterogeneous surface with adsorption

T
potentials and an adsorption equilibrium that can be expressed independently of temperature

IP
(Dubinin, 1959). The Dubinin-Radushkevich model is described by the linear equation:

R
ln 𝑞𝑒𝑞 = ln 𝑞𝑚á𝑥𝑅𝐷 − 𝛽𝑅𝐷 𝜀 2 (15)

SC
where 𝜀 is the Polany potencial described by equation 10, 𝛽𝑅𝐷 is the activity coefficient

related to the average sorption energy (mol2 kJ-2) and 𝑞𝑚á𝑥𝑅𝐷 is the maximum soption

capacity (mg g-1). U


N
1
𝜀 = 𝑅 𝑇 ln (1 + 𝐶 ) (16)
A
𝑒𝑞
M

where 𝑅 is the ideal gas constant (8,314 J K-1 mol-1) e 𝑇 is the temperature (K).

The Toth isotherm was created in 1962 and recognized as representing the adsorption of
ED

gas mixtures on heterogeneous surfaces as well as liquid-solid adsorption (Toth, 1980). It

aims to reduce the errors between the experimental values and those predicted by the model,
PT

and also considers the adsorption on a heterogeneous surface from a modification of the
E

Langmuir model (Rangabhashiyam et al., 2014). Eq. (17) presents the linear form of the Toth
CC

model:
𝑛𝑇
𝐶𝑒𝑞 1 1 𝑛
(𝑞 ) = (𝑞 𝑛 + (𝑞 𝑛 𝐶𝑒𝑞𝑇 (17)
𝑚𝑎𝑥 𝑏𝑇 ) 𝑇 𝑚𝑎𝑥 ) 𝑇
A

𝑒𝑞

where 𝑛𝑇 and 𝑏𝑇 represent the heterogeneity parameter (dimensionless) and the affinity

constant of Toth (L min-1), respectively.

3 Results and discussion

10
3.1 Characterization of the biosorbent by FTIR and SEM

Fourier transform infrared spectroscopy (FTIR) absorption spectra of unloaded and dye-

loaded residual chia-seed-oil-extraction biomass is presented at Fig. 1. In the FTIR spectra

analysis, the intensity of the peaks after the interaction of the biosorbent with Reactive Yellow

B2R textile dye is primarily noticed. A broad band was observed for both unloaded and dye-

T
loaded biomass at 3300 cm-1, indicating the presence of -OH and maybe -NH groups,

IP
characteristic of this wavenumber. –CH2 e –CH3 groups were indicated by the peaks at 2927 e

R
2856 cm-1. The group of C=O stretching vibrations is usually represented by a very strong

SC
band between 1650 and 1800 cm-1, in which the peak at 1654 cm-1 fits in both spectras, being

more intense for unloaded biomass. The observed peak at 1070 cm-1 can be related to the

U
presence the alkoxy group C-O. Therefore, it was concluded that all groups were involved in
N
the process as all peaks showed increased transmittance after biosorption.
A
M

1 1
ED

0,8 0,8
Transmitance (%)

PT

0,6 0,6

3307 2927
0,4 0,4
E

1647 1045
CC

0,2 0,2
Unloaded
Reactive-Yellow-B2R-Loaded
0 0
A

4000 3600 3200 2800 2400 2000 1600 1200 800 400
Wavenumber (cm-1)

Fig. 1

FTIR spectra for residual chia-seed-oil-extraction biomass.

11
The unloaded biosorbent SEM image, Fig. 2 (a,b), indicated that the biosorbent surface

has a well-structured layer on which cavities without a well-defined format were observed. In

addition, irregular granules are observed on the surface, indicating a heterogeneous surface.

The dye-loaded biosorbent SEM image, Fig. 2 (c,d), suggests a coarse surface morphology,

possibly as a result of the interaction between dye and biosorbent. Additionally, the deposits

T
observed on the surface could be characteristic to the dye adhered by sorption irregularly on

IP
the surface of the biosorbent.

R
SC
U
N
A
M
ED
E PT
CC
A

12
a) b)

T
R IP
SC
c) d) U
N
A
M
ED
E PT
CC

Fig. 2
A

SEM images for: (a)(b): unloaded biosorbent; (c)(d): Reactive-Yellow-B2R-loaded

biosorbent.

13
3.2 Batch biosorption trials

3.2.1 Effect of pH

The pollutant solution pH is responsible for the surface charge of the biosorbent, the

degree of ionization of the material in solution, and the dissociation of functional groups of

biosorbent active sites (Crini et al., 2007). The maximum removal percentage obtained in the

T
experimental conditions of 100 rpm agitation speed and 22 h contact time was 83% at pH 2.0.

IP
From this pH, the percentage of removal decreased to 82 and 11% at pH 3.0 and 4.0

R
respectively, and zero as pH increased. pH 2.0 was determined as the optimal condition.

SC
Likewise, the maximum efficiency in acidic pH and low efficiency in basic pH was

observed in the studies of the biosorption of 134% Yellow Reafix B2R (Da Silva et al., 2019)

U
and Orange Solimax TGL 182% (Fontana et al., 2016) onto malt bagasse, and biosorption of
N
Orange G onto banana, cucumber and potato peels (Stavrinou et al., 2018).
A
According to the analysis of pH effect in the studied biosorption process, the higher the
M

dye solution pH increases the lower the removal efficiency. As reported by the authors

Fiorentin et al. (2010), the decrease in pH leads to an increase in available protons in dye
ED

solution and a decrease in the number of active sites negatively charged on the dye surface.
PT

Consequently, the surface of biosorbent becomes more positively charged due to the

conduction of protons from dye solution to biosorbent and the electrostatic repulsion between
E

azo dye and biosorbent decreases due to the anionic characteristic of dye. As a result, the
CC

adsorption process tends toward higher efficiency at low pH.


A

3.2.2 Effect of metabolic bath agitation speed

Even though studies on dye biosorption do not often assess the relationship between the

agitation speed of the metabolic bath and the final percentage of dye removal, the reports of

the application of biosorption to remove pollutants from water (Honorio et al., 2015;

14
Rangabhashiyam and Selvaraju, 2015; Dirbaz and Roosta, 2017; Shahawy and Heikal, 2018)

have shown that the evaluation of the agitation speed is crucial in biosorption processes.

Agitation speed can influence the dispersion of solute among solution and the reduction of the

boundary layer formed by a film of solute in the surroundings of biosorbent, reducing the

resistance to mass transference (Geethakarthi, 2011). It was obtained 69, 82 and 89% of dye

T
removal at 0, 75 and 150 rpm respectively and 22 h contact time, concluding that the agitation

IP
speeds selected for the process effects on the dye removal percentage from aqueous solutions.

R
Accordingly, 150 rpm of agitation speed was determined as optimal condition.

SC
The agitation speed of 150 rpm was also used in the application of sugarcane bagasse for

removal of Reactive Blue 5G dye (Scheufele et al., 2016) and agricultural wastes for removal

U
of Methylene Blue, Malachite Green and Congo Red (Singh et al., 2017).
N
A
3.2.3 Effect of biosorbent particle diameter
M

As stated by the authors Klimaviciute et al. (2010), biosorbent particle diameter can have

great influence on sorption properties, since the increase of the surface area could be
ED

associated to smaller particle diameter (Jain and Shrivastava, 2008). However, the
PT

experimental data analysis concluded the particle diameter of the biosorbent has little

influence on the final percentage of Reactive Yellow B2R dye removal, considering that the
E

original residual biomass without sieving step presented dye removal of 87% and the smallest
CC

diameter 90%.

In this study, the use of sieving resources to obtain the specific diameter and disposal of
A

the remaining biomass due to higher efficiency was not justified, and the whole stock of

residual chia-seed-oil-extraction biomass obtained directly from the industry was applied to

the biosorption of Reactive Yellow B2R dye. Therefore, the biomass under study has

advantages over other raw materials, such as peels of vegetables (Stavrinou et al., 2018),

15
given that the residual chia-seed-oil-extraction biomass does not require crushing and sieving

processes to be used as biosorbent. Similarly, malt bagasse has been reported as an industrial

waste that does not require those processes to be effective as biosorbent (Fontana et al., 2016;

Da Silva et al., 2019).

T
3.3 Biosorption kinetics

IP
The kinetic studies aim to determine the optimal contact time between the solution and

R
the biosorbent (Daneshvar et al., 2012), that is, the contact time required to biosorbent and

SC
Reactive Yellow B2R dye solution to be at equilibrium state.

The kinetic profile (Fig. 3) showed a high rate of sorption in the initial time for all

U
evaluated temperatures, presenting dye removal rate of 40, 46, and 55% at the first 5 min of
N
process for 303, 313, and 323 K respectively, then slowly decreasing to equilibrium time of
A
60 h, showing approximately 92% of dye removal rate at the three temperatures and final dye
M

solution concentration of 6.05, 6.24, and 6.23 mg L-1 at 303, 313, and 323 K respectively.

This behavior characterizes a feasible real application as 1 h process provided 74, 83, and
ED

84% of dye removal at 303, 313 and 323 K respectively.


PT

80
Reactive Yellow B2R concentration

70
E

60
CC

50
(mg/L)

40
30
A

20
10
0
0 10 20 30 40 50 60
a) Contact time (h)

16
80

Reactive Yellow B2R concentration 70


60
50
(mg/L)

40
30

T
20

IP
10
0

R
0 10 20 30 40 50 60
b) Contact time (min)

SC
Fig. 3

U
Biosorption kinetic profile of Reactive Yellow B2R dye onto residual chia-seed-oil-extraction
N
biomass at: (a) time contact from 0 to 96 h; (b): initial sorption phase from 0 to 60 min.
A
M

According to the authors Ahmad et al. (2009), the behavior presented by the kinetic

profile (Fig. 3b) has three distinct phases: the first phase, in which the high rate of initial
ED

sorption is due to the strong electrostatic attractions between dye and external sorbent surface;

the second phase characterized by the slow progression of the sorption process due to the
PT

occupation of sorbent active sites by dye molecules, creating repulsion forces; and the third
E

and last phase presented with the reach of adsorption equilibrium, when all the active sites of
CC

the sorbent were already occupied.

Similar behavior regarding to the high rate of sorption in the first instants was observed
A

in other studies of industrial residual biomass as biosorbents in dye removal, for example in

the study of Orange Solimax TGL 182% biosorption onto malt bagasse by Fontana et al.

(2016) and Reactive Blue 5G onto orange bagasse by Fiorentin et al. (2010).

17
3.3.1 Adsorption kinetics

Kinetic models are very useful in determining the adsorption mechanisms and evaluating

the efficiency of pollutant removal by the biosorbent used in the process (Michalak et al.,

2013). Lagergren (pseudo-first order), Ho and McKay (pseudo-second order), Elovich, and

Weber and Morris (intraparticle diffusion) kinetic models were applied to the experimental

T
data.

IP
The kinetic parameters and correlation coefficients (R2) for each kinetic model at the

R
temperatures of 303, 313, and 323 K are presented in Table 1. The experimental data was

SC
better described by the pseudo-second order (R2>0.999), and the amount of dye adsorbed at

equilibrium time (qeq) predicted by this model was the closest to the qeq obtained

U
experimentally. As stated by Ho and McKay (1998), the pseudo-second order model suggests
N
that probably “the rate-limiting step may be chemical sorption, which involves valency forces
A
through sharing or exchange of electrons between sorbent and sorbate”.
M
ED
E PT
CC
A

18
Temperature (K)

Kinetic model Parameters 303 313 323

qeq (mg g-1) 1.4709 2.4575 1.2853

Pseudo-first order K1 (min) 0.0004 0.0004 0.0004

R2 0.2498 0.5434 0.5627

T
qeq (mg g-1) 10.9649 10.9770 11.1358
Pseudo-second

IP
K2 (g mg-1 min-1) 0.0032 0.0043 0.0075
order

R
R2 0.9991 0.9994 0.9998

SC
𝛼 (g mg-1 min-1) 131.8518 123.1448 6079.1840

Elovich 𝛽 (g mg-1) 1.1896 1.1397 1.4601

R2 0.7449 U 0.7811 0.5332


N
𝑘𝑖 (mg g-1 min0,5) 0.0502 0.0596 0.0307
A
Intraparticle
𝑥 (mg g-1) 7.9772 8.0470 9.6293
M

diffusion
R2 0.4904 0.6933 0.4385
ED

Experimental
qeq (mg g-1) 11.1107 11.1619 11.1408
(time of 60 h)
PT

Table 1

Kinetics model parameters for the biosorption of Reactive Yellow B2R dye onto residual
E

chia-seed-oil-extraction biomass at different temperatures.


CC
A

Analogously, the pseudo-second order model was the best fit in biosorption studies of

Ooi et al. (2017) in azo dye removal using fish scales waste, Zhao and Zhou (2016) in

methylene blue dye removal using waste dreg of Salvia miltiorrhiza, and Oguntimein (2015)

in reactive and direct dyes removal from industrial effluent using dried sunflower seed hull.

19
3.4 Adsorption isotherms

The study of adsorption equilibrium is crucial to determine the amount of sorbate

retained by the biosorbent at a constant temperature, that is, the biosorbent ability to remove

the sorbate from a solution (Brouers and Al-Musawi, 2015). Therefore, the application of the

adsorption isotherms appropriately to predict with confidence the behavior and adsorption

T
parameters is indispensable (Foo and Hameed, 2010).

IP
The adsorption isotherms were evaluated through the equilibrium parameters and

R
correlation coefficient obtained from the adjustment of the linear equations to experimental

SC
data. Table 2 presents the results obtained for each model. As presented by insufficient

correlation coefficients, Freundlich, Temkin and Dubinin-Radushdevich isotherms were not

U
effective. The isotherm that better fitted experimental data was the Toth isotherm, showing a
N
correlation coefficient of 0.9927. In addition, the Langmuir isotherm presented a correlation
A
coefficient of 0.9903.
M

Adsorption Isotherm Parameters (R2)


ED

𝑞𝑚𝑎𝑥 (mg g-1) 76.3359


Langmuir 0.9903
PT

𝑏 (L mg )-1
0.0368

𝑘𝐹 (mg g-1) 5.6157


E

Freundlich 0.8767
𝑛𝐹 0.5041
CC

𝑘𝑡 (L min-1) 0.4255
Temkin 0.9464
𝐵 15.7780
A

𝑞𝑚𝑎𝑥𝑅𝐷 (mg g-1) 52.0081


Dubinin-Radushkevich 0.8334
𝛽𝑅𝐷 (mol kJ )
2 -2
8 ∙ 10 -6

𝑞𝑚𝑎𝑥 (mg g-1) 70.9495


Toth 0.9927
𝑏𝑇 (L mg ) -1
0.0335

20
Table 2

Isotherm parameters for Reactive Yellow B2R dye biosorption equilibrium onto residual

chia-seed-oil-extraction biomass.

One of the parameters defined by both Toth and Langmuir isotherms is the maximum

T
dye biosorption capacity by the biosorbent (𝑞𝑚𝑎𝑥 ), which indicates the number of available

IP
sites in biosorbent, that is, saturation of biosorbent surface. According to the adjustment of

R
Toth adsorption isotherm to the experimental data, the maximum biosorption capacity of B2R

SC
Reactive Yellow dye by the residual chia-seed-oil-extraction biomass was 70.95 mg g-1 and

for the Langmuir isotherm, the maximum biosorption capacity was 76.34 mg g-1. The second

U
parameter defined by means of both isotherms is the affinity coefficient between the
N
adsorbent and the adsorbate (𝑏) and the affinity increases with the increasing in 𝑏
A
(Sathishkumar et al., 2008). 𝑏 coefficient obtained for the present study was 0.0335 e 0.0368
M

L mg-1 for Toth and Langmuir isotherms respectively. The proximity between the values of

𝑞𝑚𝑎𝑥 e 𝑏 found for isotherms is explained as Toth isotherm is an adaptation of Langmuir


ED

isotherm with the addition of the 𝑛𝑇 parameter, which implies the system hetetogeneity.
PT

When the 𝑛𝑇 parameter is equivalent to 1, the equation is summarized in the Langmuir

isotherm and the more the parameter o 𝑛𝑇 deviates from the unit, the more heterogeneous is
E

the surface (Terzyk et al., 2003). 𝑛𝑇 parameter determined for this study was 1.2785 and
CC

suggests a slight degree of heterogeneity of the biosorbent surface because its closeness to the

unit. For this reason, the Langmuir isotherm did not present better values for correlation
A

coefficient (R2), as this isotherm considers that the biosorbent surface is homogeneous.

This particular finding was observed in the study of methylene blue onto activated

carbon RDFK by Wu et al. (2013), in which it was obtained similar results between Langmuir

and Toth isotherms parameters: 𝑞𝑚𝑎𝑥 of 409 and 414 g kg-1, 𝑏 of 0.241 and 0.205 m3 g-1, and

21
R2 of 0.9997 and 0.9999, respectively, and Toth heterogeneity parameter (𝑛𝑇 ) was 0.94.

Likewise, Vijayaraghavan et al. (2008) observed in their study of methylene blue onto

polysulfone-immobilized Corynebacterium glutamicum that the Toth isotherm was the best fit

and the parameter of Toth heterogeneity (𝑛𝑇 ) was between 1.423 and 1.767.

The study of adsorption isotherms also aims to estimate the behavior of sorption process

T
under different operating conditions to determine whether the process is favorable or not

IP
(Brouers and Al-Musawi, 2015). From the characteristics of Toth isotherm, the biosorption

R
process of Reactive Yellow B2R dye onto residual chia-seed-oil-extraction biomass is

SC
strongly promising.

The biomass under study is a novelty. In this sense, no available literature regarding the

U
use of this particular biomass as biosorbent was available to compare with the findings of this
N
work. However, Table 3 provides a comparison of the measured maximum adsorption
A
capacity of residual chia-seed-oil-extraction biomass onto Reactive Yellow B2R with the
M

recent studies of other types of biomass applied as biosorbent in dye biosorption. According

to the data presented in Table 3, in general the maximum biosorption capacity of dye onto
ED

residual chia-seed-oil-extraction biomass is one of the highest among biomass, except for the
PT

dead leaves of Prunus Dulcis, which presented the maximum biosorption capacity of 97.09

mg g-1 and is also a treated and modified biosorbent (Jain and Gogate, 2017). An important
E

advantage of the residual biomass of chia seeds oil extraction over the dead leaves of Prunus
CC

Dulcis is the characteristic of being an unmodified biosorbent and still having a high

maximum biosorption capacity.


A

22
Biomass Dye Adsorption 𝒒𝒎𝒂𝒙 (mg g-1) Reference

Isotherm

Banana peel Orange G Langmuir 20.90 Stavrinou et al., 2018

Carica papaya wood Methylene Blue Langmuir 32.25 Rangabhashiyam et al.,

2018

Carica papaya wood Malachite green Langmuir 52.63 Rangabhashiyam et al.,

T
2018

IP
Citrus Limetta peel Methylene Blue Dubinin- 8.54 Singh et al., 2017

Radushkevich

R
Cucumber peel Orange G Langmuir 23.60 Stavrinou et al., 2018

SC
Dead leaves of Prunus Acid Blue 113 Langmuir 97.09 Jain and Gogate, 2017

Dulcis

Malt bagasse Reactive Blue BF- Langmuir


U 42.58 Junchen et al., 2018
N
5G
A
Malt bagasse 134% Yellow Langmuir 68.74 Da Silva et al., 2019
M

Reafix B2R

Potato peel Orange G Langmuir 23.60 Stavrinou et al., 2018


ED

Residual biomass Reactive Yellow Toth 70.95 This study

of chia seeds oil B2R

extraction
PT

Zea Mays cob Malachite green Dubinin- 12.84 Singh et al., 2017

Radushkevich
E
CC

Table 3
A

Comparison of maximum biosorption capacities of dyes onto several biomass.

23
3.5 Thermodynamic study

The study of the adsorption thermodynamic parameters is important in determination of

the spontaneity of reaction, as well as the exothermic or endothermic nature of the process

(Daneshvar et al., 2012).

As stated by Alothman et al. (2014), the variation of enthalpy (ΔH0) determines the

T
endothermic or exothermic characteristics of the biosorption process, this is, if the value of

IP
ΔH0 is positive, the process is endothermic, and if negative, exothermic. ΔH0 for the

R
biosorption process studied at the temperatures of 303, 313, and 323 K was 1.5269 kJ mol-1,

SC
characterizing an endothermic process. The enthalpy could also indicate that the biosorption

efficiency is directly proportional to the process temperature.

U
The determination of the mobility of the sorbate molecules after removal is suggested by
N
the order of the system after the adsorption process, which is the variation of entropy (ΔS0).
A
Positive values of ΔS0 points toward an increase in this mobility, while negative values
M

indicate lower mobility of the molecules after the process (Alothman et al., 2014). ΔS0 for the

biosorption process studied was positive at 0.0096 kJ mol-1 K-1, indicating that there is
ED

increased mobility and consequently disorganization of the molecules after the process.
PT

Finally, the variation of the Gibbs free energy represents whether the sorption process

occurs spontaneously and is favorable. Therefore, when ΔG0 decreases throughout the process
E

and is negative, the process is assumed to be spontaneous, and if the Gibbs free energy
CC

variation increases during the process and is positive, the process is not spontaneous, therefore

not favorable in the temperatures of study (Chan et al., 2016). ΔG0 for this biosorption process
A

evaluated at temperatures of 303, 313, and 323 K was -1.1395, -1.5038, and -1.5885 kJ mol-1

respectively. Thus, ΔG0 is negative and inversely proportional to the temperature. In

conclusion, the biosorption process of the Reactive Yellow B2R dye onto residual chia-seed-

oil-extraction biomass occurs spontaneously and favorably at the temperatures under study.

24
Similar thermodynamic behavior was observed in the dye biosorption process using

waste pomegranate peel by the authors Gündüz and Bayrak (2017), using natural clay by

Bentahar et al. (2017), using malt bagasse by Fontana et al. (2016), and using walnut shell by

Cao et al. (2014).

From the pseudo-second order rate constants found in adsorption kinetics at temperatures

T
303–323 K, the activation energy was determined as 33,83 kJ mol-1, obtained from the slope

IP
and intercept of the linear equation. The activation energy represents the minimum energy

R
required for the biosorption process to occur and suggests the physical or chemical nature of

SC
the adsorption, in which values between 5 and 40 kJ mol-1 indicates physical sorption and

values between 40 and 800 kJ mol-1 chemical sorption (Banerjee et al., 1997). Therefore, the

U
biosorption process of Reactive Yellow B2R dye onto residual chia-seed-oil-extraction
N
biomass indicates that this process occurs under physical sorption. Even though the kinetic
A
model of Ho and McKay (pseudo-second order) could indicate chemical sorption, the authors
M

Ho and McKay (1998) already stated that the model does not always fit exclusively for

chemical sorption. According to Ruthven (1938), the physical sorption involves weak
ED

intermolecular forces such as Van der Waals and electrostatic interactions including
PT

polarization, has low activation energy, can be monolayer or multilayer adsorption and makes

the adsorption process reversible.


E
CC

4 Conclusions

Batch biosorption of Reactive Yellow B2R dye onto residual chia-seed-oil-extraction biomass
A

was primarily assessed through the evaluation of dye solution pH, metabolic bath agitation

speed, and biosorbent particle diameter, in which the highest percentage of removal occurred

at pH 2 and 150 rpm of agitation speed using original residual biomass (without sieving step).

The equilibrium time of the batch system performed at 303, 313 and 323 K was reached at 60

25
h with 92% of dye removal. However, the biosorption was marked by a high rate of sorption

in the first hour of process with 74, 83 and 84% of dye removal at 303, 313 and 323 K,

respectively. The experimental data was better adjusted by the pseudo-second order kinetic

model and the biosorbent showed good removal ability and affinity with Reactive Yellow

B2R dye and the biosorption thermodynamic studies performed at 303, 313 and 323 K, in

T
which enthalpy, entropy, variation of the Gibbs free energy, and activation energy were

IP
evaluated, concluded the biosorption process is endothermic, occurs spontaneously and

R
favorably probably through physical sorption. The Toth equilibrium isotherm model was

SC
established as the best fit for the biosorption of Reactive Yellow B2R dye onto residual chia-

seed-oil-extraction biomass, and the maximum biosorption capacity was 70.95 mg g-1.

U
Additionally, the Toth heterogeneity parameter indicated that surface of the biosorbent has a
N
slight degree of heterogeneity, which was also suggested by SEM image of the unloaded and
A
dye-loaded biosorbent. In conclusion, the treatment of wastewater containing Reactive
M

Yellow B2R textile dye could be effective by applying the residual chia-seed-oil-extraction

biomass as biosorbent, as it is characterized as low-cost, given that it does not require


ED

previous treatment or crushing and sieving processes. Moreover, the process responded very
PT

well at temperature of 303 K, which indicated great potential of application in industries

located in regions that have 303 K as room temperature.


E

Declarations of interest: none


CC

Fundind Sources

This research did not receive any specific grant from funding agencies in the public,
A

commercial, or not-for-profit sectors.

References

26
[1] Ahmad, A., Rafatullah, M., Sulaiman, O., Ibrahim, M. H., Hashim, R., 2009.

Scavenging behaviour of meranti sawdust in the removal of methylene blue from

aqueous solution. J. Hazard. Mater. 170, 357-365.

https://doi.org/10.1016/j.jhazmat.2009.04.087

[2] Alothman, Z. A., Habila, M. A., Ali, R., Ghafar, A. A., Hassouna, M. S. E., 2014.

T
Valorization of two waste streams into activated carbon and studying its adsorption

IP
kinetics, equilibrium isotherms and thermodynamics for methylene blue removal.

R
Arab. J. Chem. 7, 1148-1158. https://doi.org/10.1016/j.arabjc.2013.05.007

SC
[3] Bailey, S. E., Olin, T. J., Bricka, M., Adrian, D. D., 1999. A review of potentially low-

cost sorbents for heavy metals. Water Resour. 33, 2469-2479.

https://doi.org/10.1016/S0043-1354(98)00475-8
U
N
[4] Banerjee, K., Cheremisinoff, P. N., Cheng, S. L., 1997. Adsorption kinetics of o-
A
xylene by flyash. Water Resour. 31, 249-261. https://doi.org/10.1016/S0043-
M

1354(96)00003-6

[5] Bentahar, S., Dbik, A., Khomri, M. E., Messaoudi, N. E., Lacherai, A., 2017.
ED

Adsorption of methylene blue, crystal violet and congo red from binary and ternary
PT

systems with natural clay: Kinetic, isotherm, and thermodynamic. J. Environ. Chem.

Eng. 5, 5921-5932. https://doi.org/10.1016/j.jece.2017.11.003


E

[6] Brouers, F., Al-musawi, T. J., 2015. On the optimal use of isotherm models for the
CC

characterization of biosorption of lead onto algae. J. Mol. Liq. 212, 46-51.

https://doi.org/10.1016/j.molliq.2015.08.054
A

[7] Cao, J., Lin, J., Fang, F., Zhang, M., Hu, Z., 2014. A new absorbent by modifying

walnut shell for the removal of anionic dye: Kinetic and thermodynamic studies.

Bioresour. Technol. 163, 199-205. https://doi.org/10.1016/j.biortech.2014.04.046

27
[8] Capitani, M. I., Spotorno, V., Nolasco, S. M., Tomás, M. C., 2012. Physicochemical

and functional characterization of by-products from chia (Salvia hispanica L.) seeds of

Argentina. LWT - Food Sci. and Technol. 45, 94-102.

https://doi.org/10.1016/j.lwt.2011.07.012

[9] Chan, S., Tan, Y. P., Abdullah, A. H., Ong, S., 2016. Equilibrium, kinetic and

T
thermodynamic studies of a new potential biosorbent for the removal of Basic Blue 3

IP
and Congo Red dyes: Pineapple (Ananas comosus) plant stem. J. Taiwan Inst. Chem.

R
Eng. 61, 306-315. https://doi.org/10.1016/j.jtice.2016.01.010

SC
[10] Crini, G., Peindy, H. N., Gimbert, F., Robert, C., 2007. Removal of C.I. Basic Green

4 (Malachite Green) from aqueous solutions by adsorption using cyclodextrin-based

U
adsorbent: Kinetic and equilibrium studies. Sep. Purif. Technol. 53, 97-110.
N
https://doi.org/10.1016/j.seppur.2006.06.018
A
[11] Daneshvar, E., Kousha, M., Sohrabi, M. S., Khataee, A., Converti, A., 2012.
M

Biosorption of three acid dyes by the brown macroalga Stoechospermum marginatum:

isotherm, kinetic and thermodynamic studies. Chem. Eng. J. 195, 297-306.


ED

https://doi.org/10.1016/j.cej.2012.04.074
PT

[12] Da Silva, B. C., Zanutto, A., Pietrobelli, J. M. T. A., 2019. Biosorption of reactive

yellow dye by malt bagasse. Adsorp. Sci. & Tec. 0, 1-24.


E

https://doi.org/10.1177/0263617418823995
CC

[13] Dirbaz, M., Roosta, A., 2018. Adsorption, kinetic and thermodynamic studies for the

biosorption of cadmium onto microalgae Parachlorella sp., J. Environ. Chem. Eng. 6


A

(2), 2302-2309. https://doi.org/10.1016/j.jece.2018.03.039

[14] Dubinin, M. M., 1959. The potential theory of adsorption of gases and vapors for

adsorbents with energetically nonuniform surfaces. Institute of Physical Chemistry,

Academy of Science, Moscow, USSR. https://doi.org/10.1021/cr60204a006

28
[15] Fiorentin, L. D., Trigueros, D. E. G., Módenes, A. N., Espinoza-Quiñones, F. R.,

Pereira, N. C., Barros, S. T. D., Santos, O. A. A., 2010. Biosorption of reactive blue

5G dye onto drying orange bagasse in batch system: Kinetic and equilibrium

modeling. Chem. Eng. J. 163, 68-77. https://doi.org/10.1016/j.cej.2010.07.043

[16] Fontana, K. B.; Chaves, E. S.; Sanchez, J. D. S.; Watanabe, E. L. R. L.; Pietrobelli,

T
J. M. T. A.; Lenzi, G. G., 2016. Textile dye removal from aqueous solutions by malt

IP
bagasse: Isotherm, kinetic and thermodynamic studies. Ecotox. Environ. Saf. 124,

R
329-336. https://doi.org/10.1016/j.ecoenv.2015.11.012

SC
[17] Foo, K. Y., Hameed, B. H., 2010. Insights into the modeling of adsorption isotherm

systems. Chem. Eng. J. 156, 2-10. https://doi.org/10.1016/j.cej.2009.09.013

U
[18] Forgacs, E., Cserháti, T., Oros, G., 2004. Removal of synthetic dyes from
N
wastewaters: a review. Environ. Int. 30, 953-971.
A
https://doi.org/10.1016/j.envint.2004.02.001
M

[19] Freundlich, H. M. F., 1906. Uber die adsorption in loungen. Zeitschrift für

Physikalische Chemie 57A, 385-470.


ED

[20] Geankoplis, C. J., 2003. Transport process and separation process principles.
PT

Prentice Hall Professional Technical reference, 4, New Jersey.

[21] Geethakarthi, A., Phanikumar, B. R., 2011. Adsorption of reactive dyes from
E

aqueous solutions by tannery sludge developed activated carbon: kinetic and


CC

equilibrium studies. Int. J. Environ. Sci. Tec. 8, 561-570.

https://doi.org/10.1007/BF03326242
A

[22] Gisi, S. D., Lofrano, G., Grassi, M., Notarnicola, M., 2016. Characteristics and

adsorption capacities of low-cost sorbents for wastewater treatment: A review.

Sustain. Mat. Technol. 9, 10-40. https://doi.org/10.1016/j.susmat.2016.06.002

29
[23] Gündüz, F., Bayrak, B., 2017. Biosorption of malachite green from an aqueous

solution using pomegranate peel: Equilibrium modelling, kinetic and thermodynamic

studies. J. Mol. Liq. 243, 790-798. https://doi.org/10.1016/j.susmat.2016.06.002

[24] Ho, Y. S., Mckay, G., 1998. Pseudo-second order model for sorption processes.

Process Biochem. 34, 451-465. https://doi.org/10.1016/S0032-9592(98)00112-5

T
[25] Honorio, J. F., Veit, M. T., Gonçalves, G. C., Campos, E. A., Fagundes-Klen, M. R.,

IP
2015. Adsorption of reactive blue BF-5G dye by soybean hulls: kinetics, equilibrium

R
and influencing factors,Water Sci. Tech. 73 (5), 1166-1174.

SC
https://doi.org/10.2166/wst.2015.589

[26] Jain, R., Shrivastava, M., 2008. Adsorptive studies of hazardous dye Tropaeoline

U
000 from an aqueous phase on to coconut-husk. J. Hazard. Mat. 158, 549-556.
N
https://doi.org/10.1016/j.jhazmat.2008.01.101
A
[27] Jain, S. N., Gogate, P. R., 2017. Acid Blue 113 removal from aqueous solution using
M

novel biosorbent based on NaOH treated and surfactant modified fallen leaves of

Prunus Dulcis, J. Environ. Chem. Eng. 6 (4), 3384-3394.


ED

https://doi.org/10.1016/j.jece.2017.06.047
PT

[28] Juchen, P. T., Piffer, H. H., Veit, M. T., Gonçalves, G. C., Palácio, S. M., Zanette, J.

C., 2018. Biosorption of reactive blue BF-5G dye by malt bagasse: kinetic and
E

equilibrium studies, J. Envir. Chem. Eng. 6 (6), 7111-7118.


CC

https://doi.org/10.1016/j.jece.2018.11.009

[29] Katheresan, V., Kansedo, J.; Lau, S. Y., 2018. Efficiency of various recent
A

wastewater dye removal methods: A review. J. Environ. Chem. Eng. 6 (4), 4676-4697.

https://doi.org/10.1016/j.jece.2018.06.060

30
[30] Klimaviciute, R., Bendoraitiene, J., Rutkaite, R., Zemaitaitis, A., 2010. Adsorption

of hexavalent chromium on cationic cross-linked starches of different botanic origins.

J. Hazar. Mat. 181, 624–632. https://doi.org/10.1016/j.jhazmat.2010.05.058

[31] Langmuir, I., (1918). The adsorption of gases on plane surfaces of glass, mica and

platinum. J. Am. Chem. Soc. 40, 1361-1403. https://doi.org/10.1021/ja02242a004

T
[32] Lagergren, S., (1898) About the theory of so-called adsorption of soluble

IP
substances. Bihang till K. Svenska vet. Akad Handlingar, 24, 1-39.

R
[33] Michalak, I., Chojnacka, K., Witek-Krowiak, A., 2013. State of the Art for the

SC
Biosorption Process—a Review. Appl. Biochem. Biotec. 170, 1389-1416.

https://doi.org/10.1007/s12010-013-0269-0

U
[34] Oguntimein, G. B., 2015. Biosorption of dye from textile wastewater effluent onto
N
alkali treated dried sunflower seed hull and design of a batch adsorber. J. Environ.
A
Chem. Eng. 3, 2647-2661. https://doi.org/10.1016/j.jece.2015.09.028
M

[35] Oliveira, A. P., Módenes, A. N., Bragião, M. E., Hinterholz, C. L., Trigueros, D. E.

G., Bezerra, I. G. O., 2018. Use of grape pomace as a biosorbent for the removal of the
ED

Brown KROM KGT dye, Bioresour Technol. Rep. 2, 92-99.


PT

https://doi.org/10.1016/j.biteb.2018.05.001

[36] Ooi, J., Lee, L. Y., Hiew, B. Y. Z., Thangalazhy-Gopakumar, S., Lim, S. S., Gan, S.,
E

2017. Assessment of fish scales waste as a low cost and eco-friendly adsorbent for
CC

removal of an azo dye: Equilibrium, kinetic and thermodynamic studies. Bioresour.

Technol. 245-A, 656-664. https://doi.org/10.1016/j.biortech.2017.08.153


A

[37] Rangabhashiyam, S., Anu, N., Giri Nandagopal, M. S., Selvaraju, N., 2014.

Relevance of isotherm models in biosorption of pollutants by agricultural byproducts.

J. Environ. Chem. Eng. 2, 398-414. https://doi.org/10.1016/j.jece.2014.01.014

31
[38] Rangabhashiyam, S., Selvaraju, N. 2015.Evaluation of the biosorption potential of a

novel Caryota urens inflorescence waste biomass for the removal of hexavalent

chromium from aqueous solutions, J. Taiwan Inst. Chem. Eng. 47, 59-70.

https://doi.org/10.1016/j.jtice.2014.09.034

[39] Ruthven, D. M., 1939. Principles of adsorption and adsorption processes. A Wiley-

T
Interscience publication.

IP
[40] Sathishkumar, M., Vijayaraghavan, K., Binupriya, A. R., stephan, A. M., Choi, J.G.,

R
Yun, S. E., 2008. Porogen effect on characteristics of banana pith carbon and the

SC
sorption of dichlorophenols. J. Colloid Interf. Sci. 320, 22-29.

https://doi.org/10.1016/j.jcis.2007.12.011

U
[41] Scheufele, F. B., Módenes, A. N., Borba, C. E., Ribeiro, C., Espinoza-Quiñones, F.
N
R., Bergamasco, R., Pereira, N. C. Monolayer–multilayer adsorption
A
phenomenological model: Kinetics, equilibrium and thermodynamics, Chem. Eng. J.
M

284, 1328-1341. https://doi.org/10.1016/j.cej.2015.09.085

[42] Shahawy, A. E., Heikal, G., 2018. Organic pollutants removal from oily wastewater
ED

using clean technology economically, friendly biosorbent (Phragmites australis), Ecol.


PT

Eng. 122, 207-218. https://doi.org/10.1016/j.ecoleng.2018.08.004

[43] Singh, H., Chauhan, G., Jain, A. K., Sharma, S. K., 2017. Adsorptive potential of
E

agricultural wastes for removal of dyes from aqueous solutions, J. Environ. Chem.
CC

Eng. 5 (1), 122-135. https://doi.org/10.1016/j.jece.2016.11.030

[44] Stavrinou, A., Aggelopoulos, C. A., Tsakiroglou, C.D., 2018. Exploring the
A

adsorption mechanisms of cationic and anionic dyes onto agricultural waste peels of

banana, cucumber and potato: Adsorption kinetics and equilibrium isotherms as a tool,

J. Environ. Chem. Eng. 6 (6), 6958-6970. https://doi.org/10.1016/j.jece.2018.10.063

32
[45] Tabaraki, R., Nateghi, A., Ahmady-Asbchin, S., 2014. Biosorption of lead (II) ions

on Sargassum ilicifolium: Application of response surface methodology. Int. Biodeter.

Biodegr. 93, 145-152. https://doi.org/10.1016/j.ibiod.2014.03.022

[46] Terzyk, A. P., Chatłas, J., Gauden, P. A., Rychlicki, G., Kowalczyk, P., 2003.

Developing the solution analogue of the Toth adsorption isotherm equation. J. Colloid

T
Interf. Sci. 266, 473-476. https://doi.org/10.1016/S0021-9797(03)00569-1

IP
[47] Timilsena, Y. P., Adhikari, R., Barrow, C. J., Adhikari, B., 2016.

R
Microencapsulation of chia seed oil using chia seed protein isolate‐chia seed gum

SC
complex coacervates. Int. J. Biol. Macrom. 91, 347-357.

https://doi.org/10.1016/j.ijbiomac.2016.05.058

U
[48] Tóth, J., 1981. A uniform interpretation of gas/solid adsorption. J. Colloid Interf.
N
Sci. 79, 85-95. https://doi.org/10.1016/0021-9797(81)90050-3
A
[49] Turner, N. H., 1975. Kinetics of chemisorption: An examination of the Elovich
M

equation. J. Catal., 36, 3. https://doi.org/10.1016/0021-9517(75)90035-4

[50] Uribe, J. A. R., Perez, J. I. N., Kauil, H. C., Rubio, G. R., Alcocer, C. G., 2011.
ED

Extraction of oil from chia seeds with supercritical CO2. J. Supercrit. Fluids 56, 174-
PT

178. https://doi.org/10.1016/j.supflu.2010.12.007

[51] Verma, A. K., Dash, R. R., Bhunia, P. A., 2012. Review on chemical
E

coagulation/flocculation technologies for removal of colour from textile wastewaters.


CC

J. Environ. Manag. 93, 154-168. https://doi.org/10.1016/j.jenvman.2011.09.012

[52] Vijayaraghavan, K., Mao, J., Yun, Y., 2008. Biosorption of methylene blue from
A

aqueous solution using free and polysulfone-immobilized Corynebacterium

glutamicum: Batch and column studies. Bioresour. Technol. 99, 2864-2871.

https://doi.org/10.1016/j.biortech.2007.06.008

33
[53] Vikrant, K., Giri, B. S., Raza, N., Roy, K., Kim, K., Rai, B. N., Singh, R. S., 2018.

Recent advancements in bioremediation of dye: Current status and challenges,

Bioresour. Technol. 253, 355-367. https://doi.org/10.1016/j.biortech.2018.01.029

[54] Weber W. J., Morris, J. C., 1962. Advances in water pollution research: removal of

biologically resistant pollutants from waste waters by adsorption. PICWPS, 2, 231-

T
266.

IP
[55] Wu, C., 2007. Adsorption of reactive dye onto carbon nanotubes: Equilibrium,

R
kinetics and thermodynamics. J. Hazard. Mat. 144, 93-100.

SC
https://doi.org/10.1016/j.jhazmat.2006.09.083

[56] Wu, K.; Wu, P.; Wu, F.; Jreng, R.; Juang, R., 2013. A novel approach to

U
characterizing liquid-phase adsorption on highly porous activated carbons using the
N
Toth equation. Chem. Eng. J. 221, 373-381. https://doi.org/10.1016/j.cej.2013.02.012
A
[57] Zhao, S., Zhou, T., 2016. Biosorption of methylene blue from wastewater by an
M

extraction residue of Salvia miltiorrhiza Bge. Bioresour. Technol. 219, 330-337.

https://doi.org/10.1016/j.biortech.2016.07.121
ED
E PT
CC
A

34

You might also like