You are on page 1of 11

Bioresource Technology Reports 3 (2018) 127–137

Contents lists available at ScienceDirect

Bioresource Technology Reports


journal homepage: www.journals.elsevier.com/bioresource-technology-reports

Use of banana trunk waste as activated carbon in scavenging methylene blue T


dye: Kinetic, thermodynamic, and isotherm studies

Mohammed Danisha, , Tanweer Ahmadb, Shahnaz Majeedc, Mehraj Ahmadd, Lou Ziyange,
Zhou Pine, S.M. Shakeel Iqubalf
a
Bioengineering Technology Section, Universiti Kuala Lumpur Malaysian Institute of Chemical and Bioengineering Technology (MICET), Lot 1988, Kawasan Perindustrian
Bandar Vendor, Taboh Naning, 78000 Alor Gajah, Melaka, Malaysia
b
Department of Chemistry, College of Natural and Computational Science, Madda Walabu University, Bale Robe, Ethiopia
c
Faculty of Pharmacy and Health Sciences, Universiti Kuala Lumpur, Royal College of Medicine, Perak 30450, Malaysia
d
School of Food Science & Bioengineering, Zhejiang Gongshang University, 18 Xuezheng Street Xiasha, Hangzhou, Zhejiang 310018, China
e
School of Environmental Science and Engineering, 800 Dongchuan Road, Shanghai Jiao Tong University, Shanghai 200240, China
f
Department of Basic Science (Chemistry), Ibn Sina National College, Jeddah, Saudi Arabia

A R T I C LE I N FO A B S T R A C T

Keywords: Activated carbon (BTAC) derived from the banana trunk was utilized in this experiment to scavenge the me-
Activated carbon thylene blue (MB) dye from synthetic dye solution. The optimized large surface area (1173.16 m2/g) BTAC was
Banana trunk prepared with phosphoric acid (6.96 mol/L) chemical activation. The maximum adsorption capacity (qmax) of
Isotherm BTAC was experimentally observed to be 166.51 mg/g at 25 °C, and adsorbent dose of 1.5 g/L for 250 ppm dye
Kinetics
concentration. The MB dye adsorption onto BTAC reached equilibrium within 20 min of contact time. The ki-
Methylene blue
Thermodynamics
netic data shown equilibrium achieved quickly (20 min) and followed a linear equation of the pseudo-second-
order kinetic model. The adsorption of dye molecules stacking in three layers on the surface of the BTAC, after
each layer formation the rate of adsorption slow down, as evident from the intraparticle diffusion study. The
thermodynamic study shown that adsorption was spontaneous and exothermic.

1. Introduction and other industries like cosmetic, pharmaceutical, rubber, plastic,


food, and paper industry (Rafatullah et al. 2010). Artificial dyes are also
Water soluble dye molecules are considered to be a hazardous or- toxic to various microorganisms which can either cause inhibition or
ganic compound for the aquatic environment. They are used in textile destruction to their catalytic and functional capabilities. The presence

Abbreviations: qmax, the maximum uptake of MB dye corresponding to the sites saturation through Langmuir isotherm model (mg/g); KL, Constant used in Langmuir
model, it corresponds to the energy of adsorption (L/g).; RL, Dimensionless separation factor; KF, Constant used in the Freundlich isotherm model, it corresponds to
adsorption capacity.; N, Constant used Freundlich isotherm model, it corresponds to adsorption intensity; KT, Temkin isotherm constant, it expressed the equilibrium
binding constant (L/mg); B, Temkin constant for the heat of adsorption (~RT/b); B, Heat of adsorption constant; T, Temperature in Kelvin scale (K); R, Universal gas
constant (8.314 J/K/mol); A, Interaction parameter which can be positive or negative values; KFM, Frumkin constant related to adsorption equilibrium.; ΔG, Gibbs
free energy change; KD-R, Dubinin and Radushkevich constant correspond to the adsorption energy.; EA, Mean free energy of adsorption; Ε, Polanyi potential ~ RTln
(1 + 1/Ce); Ɵ, Fractional occupation calculated by the ratio of adsorption capacity at equilibrium and theoretical monolayer saturation capacity.; qm, Maximum
monolayer adsorption capacity determined through D-R isotherm; HA & HB, Harkins–Jura isotherm constant for multilayer adsorption; W & Wb, Constant parameters
for Smith isotherm equation; U, Fractional attainment of equilibrium; Ce, the equilibrium concentration of MB dye (mg/L).; qe, the amount of MB dye uptake at
equilibrium (mg/g).; kn, Net first order reversible rate constant(min−1); k1, Rate of adsorption of the first-order reversible kinetic model (min−1).; k−1, Rate of
desorption of first-order reversible kinetic model (min−1).; Kc, Equilibrium constant; k2, Rate of adsorption for the pseudo-second-order kinetic model (min−1).; h,
Initial sorption rate for pseudo-second order kinetic model (mg/g/min); qt1, Weber–Morris adsorption capacity in the first phase of adsorption (mg/g); Kd1,
Weber–Morris adsorption rate constant in the first phase of adsorption (mg/g/min.1/2); C1, Boundary layer in the first phase of adsorption in the first phase; qt2,
Weber–Morris adsorption capacity in the second phase of adsorption (mg/g); Kd2, Weber–Morris adsorption rate constant in the second phase of adsorption (mg/g/
min.1/2); C2, Boundary layer in the first phase of adsorption in the second phase; qt3, Weber–Morris adsorption capacity in the third phase of adsorption (mg/g); Kd3,
Weber–Morris adsorption rate constant in the third phase of adsorption (mg/g/min.1/2); C3, Boundary layer in the first phase of adsorption in the third phase

Corresponding author at: Bioengineering Technology Section, Universiti Kuala Lumpur Malaysian Institute of Chemical and Bioengineering Technology (MICET),
Universiti Kuala Lumpur, Lot 1988, Kawasan Perindustrian Bandar Vendor, Taboh Naning, 78000 Alor Gajah, Melaka, Malaysia.
E-mail address: mdanish@unikl.edu.my (M. Danish).

https://doi.org/10.1016/j.biteb.2018.07.007
Received 12 June 2018; Received in revised form 28 July 2018; Accepted 28 July 2018
Available online 11 August 2018
2589-014X/ © 2018 Elsevier Ltd. All rights reserved.
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

of various dye molecules in aquifer reduces light penetration which from banana waste will benefit the farmers in two possible ways: it will
directly affects the photosynthesis process of aquatic flora. The waste- increase the revenue as well as reduce the volume of the waste sig-
waters discharged after dyeing processes showed high biological nificantly. At the same time, it could support the environment by fixing
oxygen demand (BOD), high chemical oxygen demand (COD), visible the carbon as a solid framework. Otherwise, these waste dumping
pollutants, and large amounts of dissolved solids. Some dyes are mu- disposal in water and humid conditions could lead to methane and
tagenic, carcinogenic and teratogenic (Bhattacharyya and Sharma carbon monoxide gas formation or their burning could produce a large
2005; Kaur et al. 2015). Water streams contaminated with artificial quantity of CO2 gas; all these gases are responsible for global warming.
dyes can cause dermatitis, allergy, irritation of the skin, which in turn The current study was designed to assess the scavenging behavior of
can cause the mutation in the nucleotides of deoxyribonucleic acid banana trunk activated carbon (BTAC) against aqueous methylene blue
(DNA), that leads to the growth of cancer cells in humans and animals (MB) dye. For the first time kinetic, thermodynamic, and isotherm
(Veliev et al. 2006). The presence of dye makes the water objectionable study of MB dye adsorption onto the novel prepared banana trunk
for drinking. Therefore, it is in dire need to scavenge dye from waste- waste activated carbon (BTAC) was studied. The effect of adsorbent
water to prevent the continuous environmental pollution and to protect dosage, initial MB dye concentration, contact time, and solution tem-
the water quality for flora and fauna. perature was evaluated for the scavenging of MB dye through BTAC
Methylene blue (MB) has a wide range of applications especially as adsorbent. The thermodynamic parameters of the scavenging behavior
a coloring agent for hair, paper, cotton, wools, and coating paper. of BTAC and MB dye were calculated, and the kinetic rate of MB ad-
Chemical studies showed that MB dye molecule is a heterocyclic aro- sorption onto BTAC was established through various models. The me-
matic compound. The chemical properties of the MB dye molecules chanism of scavenging behavior and isotherms has also been in-
have been intensively studied due to its strong binding with solids, and vestigated.
quite often it serves as a model compound for the separation of colored
bodies and colorless organic compounds from various contaminated 2. Materials and methods
mixtures (Rafatullah et al. 2010). Many advanced techniques and
treatment methods were employed for the separation/mineralization of 2.1. Preparation of BTAC
dye molecules from wastewater such as photo-catalytic (Houas et al.
2001), Fenton process and electrochemical combined treatments The banana trunk was collected from a banana plantation farm at
(Lahkimi et al. 2007), adsorption process (Danish et al. 2013), elec- Tampin, Negeri Sembilan, Malaysia. The banana trunk was cut into
trochemical degradation (Panizza et al. 2007), flocculation (Han et al. small pieces (dimension, 3 cm × 3 cm × 4 cm) and washed thoroughly
2002), chemical coagulation and ozonation (Sarasa et al. 1998), nano- several times with water to remove sand and soil residue from the
filtration and ultrafiltration (Fersi and Dhahbi 2008), cloud point ex- surface. In a subsequent step, the washed pieces of the banana trunk
traction (Bezerra et al. 2005), and reverse osmosis (Al-Bastaki 2004; were set for drying at 105 °C for 24 h in a forced draft oven. The
Nataraj et al. 2009). Adsorption is a surface phenomenon which occurs moisture content of the representative sample of the banana trunk was
at the surface, and the adsorbate is remained stuck at the surface of the taken in triplicate. Nearly 10 g of banana trunk pieces were taken in
adsorbent without any change in chemical composition or structure. So, glass Petri dishes and kept for drying at 105 °C till constant weight was
it has the advantage to be recovered the adsorbate species in its original obtained. After drying, the dried banana trunk pieces were grounded
form through desorption. The adsorption methods can be widely ap- and sieved into a uniform particle size of 0.5 μm.
plied to recover the specific chemical species from wastewater or For activated carbon preparation, 100 g of dried banana trunk
aqueous solutions if we use highly selective adsorbents. The low cost, powder was mixed with 5.09 mol/L of phosphoric acid (H3PO4) (opti-
sustainable, selective, and eco-friendly adsorbents could be proved to mized activating agents for methylene blue removal study, Danish et al.
be an effective and advanced technique to recover dyes from waste- 2018) and kept for complete soaking of the activating agent into the
water. Contemporary researchers are showing great interest in devel- banana trunk powder. The soaking was performed at room tempera-
oping new adsorbent materials with excellent properties, diverse com- ture. Then, the mixture was placed in the muffle furnace at 845 °C for
positions, and selective functionalities. The biomass waste derived 50 min in a loosely covered silica crucible. After the activation process,
activated carbons (AC) are the most promising materials which can the muffle furnace was left for 12 h to bring the temperature down. The
serve as excellent adsorbent materials. Although commercial activated washed BTAC product was dried at 105 °C for 24 h inside a hot air oven.
carbon (mostly derived from coal mining dust) is a widely used ad- Finally, the dried BTAC product was kept in tight lid glass bottles for
sorbent to remove various dye molecules, it is too expensive to be used further analysis and application studies.
in small-scale industries. Literature survey shows that in last decades so
many low-cost alternative adsorbents are proposed that includes co- 2.2. Methylene blue dye stock solution
conut shells (Dwivedi et al. 2008), Acacia mangium wood (Danish et al.
2011), rice husk (Sahu et al. 2009), pomegranate shell (Ghaedi et al. Methylene blue dye was procured from Scharlau's chemicals,
2012), bamboo (Hameed et al. 2007), banana waste (Ahmad and 1.000 ± 0.005 g of dye was dissolved in 1 L standard flask to prepare
Danish 2018), oil palm waste (Rafatullah et al., 2013) and date stone 1000 ppm stock solution. Later this stock solution was used to prepare a
(Danish et al. 2014). Methylene blue dye removal through biochar from desired concentration dye solution for further studies.
oak wood (Babaei et al. 2016), pine wood, pig manure, and cardboard
(Lonappan et al. 2016) were also tried. 2.3. Adsorbent dosage optimization experiment
Banana is the second most commonly cultivated fruit in Malaysia.
According to the statistical report of the department of agriculture Initially, the experiments were performed to optimize the adsorbent
Malaysia, the banana tree plantation cover around 30,320 ha and pro- dosage (g/L) for maximum removal of MB dye through BTAC ad-
duced 321,810 metric tonnes of fresh banana fruits in the year 2016. sorbent. Randomly selected adsorbent dosage of 0.3, 0.5, 0.7, 1.0, 1.5,
Whereas, the world production of banana in 2016 was 144 mil- 1.7, 2.0 g/L were used for finding the optimum dosage for maximum
lion metric tonnes (Ahmad and Danish 2018). Banana plant can pro- removal percentage against 50 ppm MB dye solution. The experiments
duce fruit once in a lifetime, so after fruit harvesting, the whole plant is required three conical flasks for each selected dosage, to find the de-
discarded as waste. It was estimated that near 100 kg of waste gener- viation in the reproducibility of the dye adsorption. The BTAC ad-
ated for every tonne of banana fruit production. If this much huge waste sorbent was weighed on butter papers using analytical balance
can be transformed into useful material, will be a great help to banana (Sartorius Quas) that was capable of weighing accurately in milligrams.
grower farmers. Keeping these in view, the activated carbon formation Butter paper was preferred during the activated carbon weight

128
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

measurement due to its non-sticky surface that minimizes the mass loss (Ci − Ct ) V
qt =
and prevents the powder from being scattered around. For 30 mL of m (3)
50 ppm MB dye, the weight of the BTAC adsorbent was accurately
The percentage MB removal has been calculated by the following
measured using the following relationship (Eq. (1)).
Eq. (4):
dosage (g /L) × MB dye vol. (mL) (Ci − Ce )
Adsorbent weight = %MB removal = × 100
1000 (1)
Ci (4)
The precise weight of BTAC adsorbent was transferred into conical where,
flasks, containing 30 mL of 100 ppm MB solution. The mixture was Ci, Ce and Ct (mg/L), = Initial, equilibrium, and after time-t con-
thoroughly mixed and set for 1 h continuous agitation. After that, BTAC centration of MB dye, respectively.
was filtered through the Whatman filter paper No. 42, and residual MB V = Solution volume of MB dye (units in L).
dye concentration was measured through UV-VIS spectrophotometer m = Amount of adsorbent (units in g).
(Perkin Elmer, Lambda 35 model). The UV-VIS spectrophotometer was
calibrated before the actual measurement was conducted. The standard
3. Results and discussion
solution of MB dye of various concentrations such as 0.01 ppm,
0.03 ppm, 0.05 ppm, 0.07 ppm, and 0.10 ppm was prepared carefully.
3.1. Effect of adsorbent dosage and contact time on MB adsorption
The calibration curve was prepared between the standard solution
concentration and their corresponding absorbance as shown in
The adsorption percentage of MB onto different dosage of BTAC
Supplementary Fig. S1. The regression coefficient (R2) of the calibration
adsorbent with 100 ppm of initial dye concentration (as shown in
curve was calculated to be 0.995 with the intercept at 0.00. This high
Supplementary Fig. S2). The adsorbent dosage increased from 0.3 g/L
correlation coefficient (R2) value gave the highly precise results of re-
to 1.5 g/L, the adsorption percentage increased with regular increment
sidual MB dye concentration.
in the dosage and reached 99.98%. Further increase in BTAC dosage up
to 2.0 g/L didn't change the adsorption percentage anymore. So BTAC
2.4. Kinetic measurements dosage of 1.5 g/L was chosen as maximum removal for 100 ppm MB dye
solution for further studies.
Thirty-nine conical flasks were prepared with a fixed dosage of At an optimum dosage of BTAC adsorbent, the equilibrium contact
1.5 g/L (45 mg of BTAC for 30 mL of 100 ppm MB dye solution) BTAC time was established through the plot of percent adsorption vs. time
adsorbent that was determined through dosage experiment. Each con- curve (as shown in Supplementary Fig. S3). The curve showed that MB
ical flask after mixing the MB dye solution was agitated for a selected dye adsorption was rapid and 99.53% dye was scavenged in the initial
time interval of 1, 2, 3, 4, 5, 6, 7, 8, 9, 10, 15, 20, and 30 min. The 10 min of BTAC and MB dye interaction. At equilibrium, the MB dye
filtered residual MB dye solution was measured through the UV-VIS removal percentage was found to be 99.87%.
spectrophotometer as stated in the previous section. Each time interval
was repeated three times to find the deviation in the measured value. 3.2. Kinetic studies

2.5. Thermodynamic measurements The adsorption of MB dye vs. time data was tested against various
available kinetic models. The aim of the kinetic studies was to explore
Thermodynamic parameters of MB dye adsorption onto BTAC ad- the possible mechanism of sorption of MB dye onto BTAC adsorbent.
sorbent were measured for three dye concentrations viz., 50, 100, and The well established kinetic models such as reversible first order reac-
150 ppm at 30, 40, 50, and 60 °C. For each concentration of MB dye tion, pseudo-first-order, and pseudo-second-order models were tested in
solution, 1.5 g/L adsorbent dosage was used. The temperature of the this study. The conformity between predicted model values and ex-
dye and adsorbent mixture inside a conical flask was maintained by perimental data was expressed by the regression coefficient (R2). The
immersing the conical flask inside the water bath shaker for a fixed time higher regression coefficient value with the available kinetic experi-
of 1 h. After the fixed time of agitation inside the water bath at a mental data is considered as an applicable model.
constant selected temperature, the residual MB dye solution was fil- A first-order reversible kinetic model was used to explore the pos-
tered and kept in a vial at room temperature (25 °C). Later residual sibility of the applicability of the MB dye adsorption data. The simpli-
concentration of the MB dye was measured through the UV-VIS spec- fied reversible kinetic model is given below:
trophotometer as mentioned in section 2.3.
ln(1 − U ) = −kn t (5)

2.6. Adsorption isotherm study where, U represents the fractional attainment of equilibrium; kn is net
rate constant, and t is a time in min.
Isotherm study of MB dye adsorption onto BTAC was conducted The plot between ln(1-U) vs. t, should give a straight line that must
with five different concentrations of dye solutions viz., 50, 100, 150, pass through the origin for a perfect first order reversible kinetic re-
200, and 250 ppm using a fixed dosage of 1.5 g/L adsorbent. The action between adsorbent and adsorbate. We applied MB dye adsorp-
maximum deviation in the dilution of stock solution (1000 ppm) to the tion data to calculate U (as shown in the supplementary data file), then
selected concentration was ± 10 ppm. These deviations in the dye plot ln(1-U) vs. t. The representative of the plot is shown in Fig. 1(a). It
concentration were considered during the calculations to get the correct was observed that the line was not a straight line and the regression
values of isotherm model constants and adsorption capacity. Isotherm coefficient (R2) was found to be 0.856, which suggested that MB dye
study was conducted at room temperature only. The amount of MB adsorption was not followed the first-order reversible kinetic model.
adsorbed per unit mass of adsorbents (q) at equilibrium was calculated Constants kn, k1, and k−1 were calculated and reported in Table 1.
using the following equation: Pseudo-first order kinetic model was applied to the MB dye ad-
sorption data; the following equation was used for model calculation.
(Ci − Ce ) V
qe = k′
m (2)
log(qe − qt ) = log qe − ⎛ 1 ⎞ × t
⎜ ⎟

⎝ 2.303 ⎠ (6)
The amount of MB adsorbed per unit mass of adsorbents at the time
t, was calculated through the following equation: where, qt and qe (mg/g) = adsorption capacity at time, t and

129
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

Fig. 1. (a) First order reversible kinetic model plot, (b) Pseudo-first order kinetic model plot, (c) Pseudo-second-order kinetic model, (d)Weber–Morris diffusion
model plot, for MB dye adsorption onto BTAC.

equilibrium, respectively. BTAC follows the pseudo-second-order kinetic model perfectly.


k1’ (1/min) = rate constant for pseudo-first-order adsorption. Intraparticle diffusion of the MB dye molecules was calculated
The representative of the plot is shown in Fig. 1(b). The regression through Weber–Morris model (the model is represented in Eq. (8)).
coefficient, R2 for this model is 0.961 and not close to the ideal corre-
lation coefficient value. Constant k value was calculated and mentioned qt = kd × t 1/2 + c (8)
in Table 1.
The pseudo-second-order kinetics model (Ho and McKay, 1999) can where, qt (mg/g) = The adsorption capacities at time t; kd (mg/g/min1/
2
be expressed as: ) = diffusion rate constant, t = time, and c (mg/g) = boundary
thickness.
t 1 1
= + t It can be seen from Fig. 1(d), the plot shows three different slopes.
qt h qe (7) Initial slope was sharp increasing with high diffusion rate constant
where; value (kd = 14.694 mg/g/min1/2), when the BTAC adsorbent surface
qe and qt (mg/g) = The adsorption capacities at equilibrium and at and pores were freely available for the MB dye molecules, so thickness
time t, respectively. boundary (35.35 mg/g) was also less. In the second stage, the diffusion
k2 (g/(mg·min)) = Rate constant of the pseudo-second-order reac- rate decreased to 0.6082 mg/g/min1/2, and boundary thickness in-
tions. creased to 62.69 mg/g. In the third stage, the diffusion rate further
h = Initial sorption rate as qt/t → 0 h = k2 × qe2. decreased to 0.1358 mg/g/min1/2, and boundary thickness increased to
The plot of t/qt vs. t of Eq. (7) is shown in Fig. 1(c) and give a linear 65.90 mg/g. It is clear from the results that as the adsorption of MB dye
relationship, from which qe, k2, and h can be calculated. The values for progresses with time the diffusion rate decreased due to pores and
qe, k2, and h were calculated and are summarized in Table 1. The re- surface spaces were occupied the dye molecules. Consequently, an in-
gression coefficient (R2) value for this pseudo-second-order model is crease in boundary thickness at the surface of BTAC adsorbent was
0.999 which is very close to 1.00, the ideal value of the regression observed. The Weber–Morris model parameters with correlation coef-
coefficient. Therefore, it can be concluded that the sorption of MB onto ficients (R2) are reported in Table 1.

130
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

Table 1
Kinetic models parameters for the adsorption of MB onto BTAC.
Kinetic parameters Values

qe (mg/g) (for 100 ppm initial conc of MB dye) 66.57


First-order-reversible reaction kinetics model
kn (min−1) 0.319
k1(min−1) 0.047
k−1(min−1) 0.271
Kc 0.175
R2 0.856

Pseudo-first order reaction kinetics model


qe,cal (mg/g) 31.19
k1(min−1) 0.067
R2 0.961

Pseudo-second order reaction kinetics model


qe,cal (mg/g) 67.57
k2(min−1) 0.056
h (mg/g/min) 256.4
R2 0.999

Weber–Morris model
qt1(mg/g) 101.068
Kd1(mg/g/min.1/2) 14.694
C1 35.354
R12 0.9856
qt2(mg/g) 65.414
Kd2(mg/g/min.1/2) 0.6082
C2 62.694
R22 0.8237
qt3(mg/g) 66.517
Kd3 (mg/g/min.1/2) 0.1358
C3 65.91
R32 0.7019

3.3. Effect of temperature on MB dye sorption

Adsorbate solution temperature is an extremely significant factor Fig. 2. (a) Effect of temperature on MB dye adsorption onto BTAC, (b) Vont
Hoff plot for MB dye adsorption onto BTAC.
during adsorption processes, so the experiments were conducted to
observe it for MB dye adsorption onto BTAC. The experiments were
designed at selected temperatures such as 303.15 K, 313.15 K, ΔS ° ΔH ° 1
ln K c = − ⎛ ⎞
323.15 K, and 333.15 K, for different initial concentrations (50 ppm, R R ⎝T ⎠ (9)
100 ppm, and 150 ppm) of MB dye. The percentage sorption of MB dye
ΔGo = ΔH o − T ΔS o (10)
onto BTAC adsorbent was calculated against various temperature and
initial concentrations to explore the nature of the adsorbate-adsorbent where, Kc is the thermodynamic equilibrium constant (calculated
interaction. With the help of Fig. 2(a), the effect of temperatures on through the ratio of adsorbed concentration to residue MB dye con-
three different initial concentrations of MB dye was represented. It was centration, at equilibrium), T is the temperature in Kelvin scale (K), and
observed that with the rise in temperature from 303.15 K to 333.15 K, R represents the universal molar gas constant (8.314 J/mol/K).
the adsorption percentage was marginally decreased for each higher Fig. 2(b) indicated that the rate of sorption decreased marginally
initial concentrations of MB dye. At a higher initial concentration of MB along with the rise in temperature from 303.15 K to 333.15 K. The
dye and higher temperature among the selected one, the effect of linearity of the Vont Hoff plots were observed to be 0.9892, 0.993, and
temperature was observed to be more. This behavior can be explained 0.9872 for 50 ppm, 100 ppm, and 150 ppm initial concentration of MB
based on the kinetic movements of the adsorbate (MB dye) molecules dye, respectively. The thermodynamic parameters for MB adsorption
and adsorbent particles (BTAC). Since at higher concentration of MB onto BTAC is reported in Table 2. The result showed that the value of
dye molecules, the more molecules adsorb and desorb at a higher ΔG° for MB dye sorption onto BTAC at different temperature and initial
temperature, so net adsorption was observed to be lower. Further, this concentrations were varied between −18.313 kJ/mol
thermal interaction was used to calculate the thermodynamic para- and − 19.719 kJ/mol. The ΔG° < 0 indicated that MB dye sorption
meters. onto BTAC was thermodynamically favorable and spontaneous. Since
the ΔG° values fall within the range of 0 and − 20 kJ/mol, that in-
dicated the electronic interaction between adsorption sites in the ad-
3.4. Thermodynamic studies sorbent (BTAC) and the adsorbate molecules (MB dye) (Al-Othman
et al. 2013). Meanwhile, ΔG° values ranging from −80 and − 400 kJ/
The thermodynamic calculations for MB dye sorption onto BTAC mol revealed that the sorption involving sharing or transferring of
was conducted using Von't Hoff relation and standard Gibbs free energy charge from the surface of the adsorbent to the adsorbate to form a
change (ΔG°), standard enthalpy change (ΔH°), and standard entropy coordinate bond.
change (ΔS°) relationship (as given in Eqs. (9) and (10)). The ΔH° and At a fixed temperature, when we compared the ΔG° values of dif-
ΔS° were evaluated through the slope and intercept of Fig. 2(b) plots, ferent initial concentrations. It was observed increasing trends, which
respectively. The ΔG° was estimated by using ΔH° and ΔS° (calculated indicated the decrease in the spontaneity of sorption. The increase in
from the ln (Kc) vs. 1/T plot) at different temperatures in Eq. (10). ΔG° values with the rise of temperature indicated that the better sorp-
tion of MB dye onto BTAC could be found at lower temperatures. The

131
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

Table 2 higher initial concentration (150 ppm). Because, at low initial MB dye
Thermodynamics parameters for MB dye adsorption onto BTAC. concentrations more heat evolved (for 50 ppm ΔH° = −24.571 kJ/mol
Concentration Temperature ΔG° ΔH° ΔS° and for 100 ppm ΔH° = −22.039 kJ/mol) compared to higher initial
(ppm) (°C) (kJ/mol) (kJ/mol) (kJ/mol K−1) MB dye concentration (150 ppm, ΔH° = −15.600 kJ/mol). Negative
standard entropy change (ΔS°) can be explained as the decrease in
50 30 −19.719 −24.571 −0.0157
randomness, and positive ΔS° values indicated the increase in ran-
40 −19.639
50 −19.482
domness in the system. We found negative ΔS° for 50 ppm and
60 −19.324 100 ppm MB dye solution but positive ΔS° values for 150 ppm solution.
100 30 −19.358 −22.039 −0.0088 The reason for the increase in randomness in the system for higher
40 −19.270 concentration was the replacement of some of the small molecules from
50 −19.182
the surface of adsorbent due to a large crowd of adsorbate molecules
60 −19.093
150 30 −18.313 −15.600 0.0089 approaches to the surface. Some authors considered even displacement
40 −18.403 of water molecules from the surface of the adsorbent added more
50 −18.492 translational entropy than it lost due to the settling of adsorbate mo-
60 −18.582
lecules on the surface of adsorbent; this phenomenon also caused an
increase in randomness in the system (Malkoc and Nuhoglu 2007).
overall ΔG° obtained in this study was indicating for physisorption of
MB dye molecules onto the BTAC surface because of ΔG° lies between 3.5. Adsorption isotherm models
−20 kJ/mol to 0.00 kJ/mol in all cases studied here (Mahmoodi 2011).
The standard enthalpy and entropy changes in the selected range of The adsorption isotherm identifies the distribution of concentration
temperature (303.15 K to 333.15 K~30–60 °C) were found to be variation of adsorbate molecules between solid (adsorbent) and liquid
−24.571 kJ/mol and − 0.0157 kJ/ mol/K, respectively for 50 ppm in- phases at a constant temperature. The isotherms successfully represent
itial MB dye concentration. For 100 ppm MB dye solution the value for the dynamic adsorption behavior of adsorbate. At equilibrium, it can be
ΔH° is −22.039 kJ/mol while the value for ΔS° obtain was expressed as the amount of adsorbate molecules adjusted at the surface
−0.0088 kJ/mol/K. Lastly, for 150 ppm MB dye solution, the standard of per unit mass adsorbent. All available adsorption isotherm models
enthalpy and entropy change in the range of 30–60 °C temperature were are composed of certain constant values which assigned to specific
found to be −15.600 kJ/mol and 0.0089 kJ/mol/K. The negative va- surface behavior and used to compare the adsorption properties of the
lues of the ΔH° indicated that the adsorption was exothermic adsorbents against pollutants (Dursun et al. 2005). To identify the
(ΔH° < 0) (Mortimer 2008). It can be inferred from the obtained data, sorption mechanism of the MB dye onto the BTAC, different initial
at a fixed initial concentration the increased in temperature decreased concentrations of MB dye was keep in contact with the BTAC adsorbent
the spontaneity of the MB dye adsorption onto BTAC surface except at (1.5 g/L) and their respective equilibrium concentrations values were

Table 3
MB dye adsorption onto BTAC isotherm model constants and their correlation coefficients.
Isotherms Models constants Units Values

Langmuir isotherm
Ce
=
1
+
Ce qmax mg/g 227.272
qe qmax × KL qmax
KL L/g 11.000
RL 18.15–3.64 × 10−4
R2 0.9635

Freundlich isotherm
1 Kf 377.851
ln(qe ) = ln(Kf ) + × ln(Ce )
N
N 1.765
R2 0.9926

Dubinin-Raduskevich isotherm
ln(qe) = ln (qm) − KD−R × ε2 KD-R mol2/kJ2 2.00 × 10−11
qm mg/g 199.877
EA kJ/mol 5.00
R2 0.9863

Temkin isotherm
qe = B × ln (KT) + B × ln (Ce) KT L/g 116.531
B 46.207
B J/mol 54.546
R2 0.9608

Frumkin isotherm
A −0.905
ln ⎡
⎣ ( ) θ
1−θ
1

Ce ⎦
= ln(KFM ) + 2 × a × θ
ln(KFM) 4.3344
ΔG kJ/mol −10.924
R2 0.6385

Harkins–Jura isotherm
1 HB 1 HA 1428.571
= −⎡ × log(Ce ) ⎤
qe2 HA ⎣ HA ⎦ HB −0.71429
R2 0.8190

Smith isotherm
qe = Wb − W × ln (1 − Ce) Wb 40.378
W 487.94
R2 0.9565

132
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

Fig. 3. (a) Langmuir plot, (b) Freundlich plot, (c) D-R isotherm plot, (d) Temkin isotherm plot, (e) Frumkin isotherm plot, (f) Harkins–Jura isotherm plot, (g) Smith
isotherm plot, for MB dye adsorption on BTAC (h) FESEM image of BTAC.

applied to the linearized form of Langmuir, Freundlich, Dubinin-Ra- as summarized in Table 3.


duskevich, Temkin, Frumkin, Harkins–Jura, and Smith isotherms
models. The adsorption isotherm constants and regression coefficient
(R2) were calculated through respective linearized isotherm equations 3.5.1. Langmuir isotherm model for monolayer adsorption
The Langmuir isotherm is based on the assumption that the

133
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

adsorbent surface has uniformly distributed homogenous surface ad- The Dubinin-Radushkevich isotherm constant (KD-R) can be used to
sorption sites available for adsorbate molecules. It also assumes that the mean free energy of adsorption (EA) per molecule of adsorbate
during adsorption the energy is uniformly distributed between the ad- when the adsorbate molecules move from infinity to the surface of the
sorbate surface and adsorbent molecules for monolayer adsorption at a adsorbent. It can be calculated by the given below equation (Eq. (13)).
fixed temperature. 1
From Fig. 3(a), the scattered data points of Ce/qe vs. Ce plot has been EA =
2KD − R (13)
shown. The best fit line with their equation is also represented there.
Five gradual increasing initial concentrations (50, 100, 150, 200, and The Dubinin-Radushkevich isotherm constants and its related mean
250 ppm) of MB dye were considered for this plot. The slope and in- free energy of adsorption for MB dye adsorption onto BTAC are re-
tercept of the best fit line were employed to determine the Langmuir ported in Table 3. The linearized D-R plot is shown in Fig. 3(c), the
model parameters such as qmax and KL. The values of isotherm constants correlation coefficient (R2) for the linear plot was found to be 0.9863,
qmax and KL were found to be 227.27 mg/g and 11.00 L/g, respectively. which is in very close agreement with the perfect linear (R2 = 1). After
The nature of the MB dye adsorption on BTAC was also tested through Freundlich isotherm, this isotherm model was more closely followed by
dimensionless constant commonly known as equilibrium parameter or the MB dye adsorption data. According to D-R isotherm model pre-
separation factor, RL. The separation factor was calculated by the fol- dicted maximum adsorption capacity (qm = 199.87 mg/g) of the BTAC
lowing relation (Eq. (11)): adsorbent was also very close to the experimentally calculated ad-
1 sorption capacity (qe = 166.50 mg/g).
RL =
1 + KL Co (11)
3.5.4. Temkin adsorption isotherm model
where C0 is the initial MB dye highest concentration (mg/L), KL is the Temkin isotherm is modeled by assuming that the decrease in the
Langmuir isotherm model constant corresponds to the adsorption en- heat of adsorption followed the linear trend rather than a logarithmic
ergy (L/g). The RL value is an essential feature of the Langmuir iso- curve. The non-linearized form of Temkin isotherm can be mathema-
therm to explore whether the adsorption is “favorable (0 ˂ RL ˂ 1)”, tically expressed by the following equation (Eq. (14)).
“unfavorable (RL ˃ 1)”, “linear (RL = 1)” or “irreversible (RL = 0)”. For
RT
MB dye adsorption on BTAC, the RL values were found to be 0.001815, qe = ln(KT × Ce )
b (14)
0.000908, 0.000606, 0.000454, and 0.000364 while initial concentra-
tions of MB dye were 50, 100, 150, 200, and 250 mg/L at 303.15 K, RT
respectively. These RL values indicated that adsorption of MB dye on B=
b (15)
BTAC was favorable. Table 3 contains all the calculated parameters of
the Langmuir isotherm model. The regression coefficient (R2) of this where, B is constant dealing with the heat involved during adsorption.
model was found to be 0.9635, this regression coefficient shows the The heat of adsorption constant (b) can be calculated through the slope
isotherm data deviated from the linearity. Therefore, adsorption of MB of qe vs. ln Ce plot as shown in Fig. 3(d). The calculated values of the
dye onto BTAC did not seem to be a monolayer. Temkin isotherm constants and its related parameters are summarized
in Table 3. The correlation coefficient of the model for MB dye ad-
sorption data was found to be 0.9608, so the calculated values are near
3.5.2. Freundlich model to the actual heat involved in the adsorption.
This isotherm assumes that the energy is not uniformly distributed
at the surface site of the adsorbent and the adsorbate molecules, rather
3.5.5. Frumkin adsorption isotherm model
energy is exponentially varied in adsorbent surface and the adsorbate
The Frumkin isotherm assumes the homogeneous surface of the
molecules. The Freundlich isotherm model was developed by assuming
adsorbent with uniform interaction with adsorbent molecules. This
the non-uniform distribution of adsorption energy at the heterogeneous
isotherm used a constant term ‘a’ which stands for the interaction of
surface of the adsorbent. The experimental data was applied to the plot
adsorbate molecules with the adsorbed layer. The positive value of ‘a’
of ln(Ce) vs. ln(qe), and the extent of linear relationship was established
indicated that during adsorption the energy increased due to the lateral
through the regression coefficient (R2). The linearized Freundlich
interaction of the adsorbate molecule with the surface of the adsorbent.
model is shown in Table 3. The Freundlich isotherm helps to explore the
The negative value of ‘a’ pointed that the repulsive force was operated
possibility of multilayer adsorption characteristics of the MB dye onto
between the incoming adsorbate molecules and the molecules settled
BTAC.
on the surface of the adsorbent. The linear form of Frumkin isotherm
The experimental data plot of ln(Ce) vs. ln(qe) is shown in Fig. 3(b),
with its parameter values are shown in Table 3. The fractional occu-
the experimental data points are in good agreement with the linearity of
pation (θ) of adsorbate at the surface of adsorbent was calculated
the plot. The regression coefficient (R2) was found to be 0.9926, which
through the ratio of experimental adsorption capacity to the monolayer
seems to be the best fit for the linearity test of the Freundlich isotherm
adsorption capacity (calculated through D-R isotherm) for each initial
model. The sorption of MB dye onto the BTAC was a multilayer. The
concentration of MB dye. From the experimental data of MB dye ad-
Freundlich constant related to adsorption (Kf) for MB dye has very high
sorption on BTAC, it was observed that Frumkin isotherm model gave
value 377.85, and Freundlich constant related to the intensity of ad-
poor linearity with a correlation coefficient (R2) 0.6385 as shown in
sorption was also high (1.765). This model better explains the ad-
Fig. 3(e). The isotherm constant parameter ‘a’ was having negative
sorption characteristics of the MB dye onto BTAC.
values that indicated that there was repulsion between the incoming
MB dye molecules with the molecules settled on the surface. The re-
3.5.3. Dubinin-Radushkevich (D-R) adsorption isotherm model pulsion between the bounded and incoming adsorbate molecule is not
Dubinin-Radushkevich isotherm also considering the concept of true because the experimental data verified multilayer adsorption
heterogeneity of the adsorbent surface as in Freundlich isotherm and (Freundlich isotherm) characteristics. Therefore, the other parameters
assume that the sorption sites are not identical (Misra 1969). The lin- of the isotherm model could not be explained because experimental
earized equation of isotherm is represented in Table 3. The Polyani data was not following this model.
potential (Ɛ) was calculated using the following relation (Eq. (12)).

1⎞ 3.5.6. Harkins–Jura adsorption isotherm model


ε = RT ln ⎛1 +
⎜ ⎟
This isotherm explains the multilayer adsorption on a hetero-
⎝ Ce⎠ (12)
geneous surface with a range of pore distribution. The linearized

134
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

CH2 H2C OH CN-

OH

O
S-
BTAC adsorbent surface functional groups
COOH
+

CH3
H3C
N S N Cl
CH3
MB dye molecules CH3

Third layer of molecules Second layer of molecules


N
Cl
H3C CH3 H3C
N CH3
H3C N CH3
N S N Cl CH3
MB dye molecules CH3 H3C N
SH CH3
SH
N
H3C
CH3 N N
N
S
H3C First layer of molecules N
H3C
CH2 H2C OH CN-

OH
H3C N
CH3
O Cl
S-

BTAC adsorbent surface functional groups


COO H
+

N CH3

S CH3

H3C N

CH3

Fig. 4. Schematic representation of mechanism of MB dye molecules adsorption on the macromolecular surface of BTAC. Different layers of dye molecules are also
represented here.

Harkins–Jura adsorption isotherm in terms of 1/qe2 and log(Ce) was parameters with their values are reported in Table 3. The correlation
used to apply on the adsorption data of MB dye on BTAC adsorbent. The coefficient (R2) of the model was found to be 0.8190, which indicated
plot is shown in Fig. 3(f). The HA represents for multilayer adsorption that the obtained MB dye adsorption data do not follow this model.
isotherm constant and explains the existence of heterogeneous pores,
and HB represents the isotherm constant. The relevant isotherm

135
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

3.5.7. Smith adsorption isotherm model wood biochar through Langmuir isotherm was reported to be
This isotherm model assumes the adsorbent has a heterogeneous 97.55 mg/g at 50 °C (Babaei et al. 2016). On comparing adsorption
surface, and the adsorbed molecules at the surface of adsorbent divided capacities of these biomass-derived adsorbents, BTAC was found to be
into bounded and condense fractions. The bounded fraction of the ad- more efficient with maximum adsorption capacity 199.87 mg/g calcu-
sorbate at the inner or outer surface of the adsorbent facilitates the lated through Dubinin-Radushkevich (D-R) isotherm model at 25 °C and
condensation of the incoming adsorbate molecules (Smith 1947). The adsorbent dose 1.5 g/L.
condensed fraction of the adsorbate molecules forms in layers, each
condense fractions have more than one layers. The linearized form of 4. Conclusions
the Smith isotherm with their constants values are reported in Table 3.
The plot between qe and ln(1-Ce) is shown in Fig. 3(g), the correlation The present study is dedicated to kinetic, thermodynamics, and
coefficient value was found to be 0.9565, which again confirms the isotherm behavior of MB dye adsorption onto BTAC adsorbent. The
heterogeneous surface of BTAC adsorbent and adsorption of MB dye following conclusion was drawn from the study:
was a multilayer. The heterogeneous surface of the BTAC adsorbent can
be in field emission scanning electron microscopic (FESEM) image as in (1) Percentage MB removal increased with an increase in time and
Fig. 3(h). The various size of the diameter of the pores varies between achieves equilibrium after 20 min of contact time.
5.71 μm to 20.22 μm. (2) The kinetics data of MB dye adsorption on BTAC was found to
follow the pseudo-second-order kinetic model with a correlation
3.6. Mechanism and comparison of MB dye adsorption coefficient (R2 > 0.99) close to absolute linearity (R2 = 1.00).
(3) The Weber–Morris intraparticle diffusion model was also applied to
Possible functional groups in the BTAC and MB dye, and changes in the kinetic data to estimate the diffusion rate constants and
the surface of BTAC after MB dye adsorbed was identified through FTIR boundary thickness after each layer of adsorption. The results re-
spectra (as shown in Supplementary Fig. S4). Based on the FTIR spec- vealed that three layers of adsorbate formed on the surface of the
trum information the functional groups on the surface of BTAC was adsorbent for each subsequent layers the diffusion rate was ob-
observed at wavenumbers 3443 cm−1 (-OH groups, a broad peak), served to be decreased, and boundary thickness increased.
2366 cm−1 (-CN groups, weak peak), 1637 cm−1 (-C=C-COO– groups, (4) The thermodynamic parameters are shown adsorption was spon-
medium peak), 1592 cm−1 (-C=C-O-S groups, medium peaks), taneous and decrease with increasing temperature (exothermic).
1415 cm−1 (heterocyclic ring skeleton, weak peak). In MB dye the The thermodynamic parameter ΔG°, ΔH°, and ΔS° indicated that the
peaks were observed at wavenumbers 3433 cm−1 (-OH groups) due to adsorption process was spontaneous and exothermic.
presence of moisture in the MB dye, and at wavenumber 2357 cm−1 (5) The isotherm data were tested against the Langmuir, Freundlich,
(-C=N-C groups, sharp and strong peak),1645 cm−1 (-C=N- group, Dubinin-Radushkevich, Temkin, Frumkin, Harkins–Jura, and Smith
weak peak), and 1550 cm−1 (-S-C-C-N-, heterocyclic ring, medium isotherm models to verify the isotherm characteristics. The iso-
peak) was due to presence of functional groups in the MB dye mole- therm data followed the Dubinin-Radushkevich model closely, and
cules. After MB dye adsorption on to the surface of the BTAC, the BTAC- the model predicted qmax was 199.87 mg/g; whereas, the maximum
MB dye complex was also analyzed through the FTIR, and spectrum adsorption capacity of BTAC calculated through experimental va-
shown some peaks of similar functional groups get stronger, like peak at lues with 250 ppm initial MB dye concentration was 166.51 mg/g.
wavenumber 3443 cm−1 getting strong and broad, in BTAC the weak
peak was observed at 2366 cm−1, but after MB dye adsorption this peak In view of the experimental findings, this study recommended that
shifted 2363 cm−1 with medium strength. Similarly, the peaks in the the banana trunk chemically activated carbon (BTAC) can be used ef-
BTAC surface at wavenumber 1637 cm−1, shifted to 1644 cm−1, with fectively in the scavenging of MB dye from aqueous solution.
increased peak strengths. These changes in the functional groups peak
position and strength in the BTAC-MB dye complex gave the evidence of Acknowledgment
MB dye adsorption onto the BTAC surface. The interaction sites in the
BTAC and MB dye macromolecule was identified and drawn in Fig. 4. This research did not receive any specific grant from funding
The negatively charged functional groups such as -CN-, –O-S-, and agencies in the public, commercial, or not-for-profit sectors. However,
–COO– at the surface of BTAC interact with positively charged the authors acknowledge the Universiti Kuala Lumpur Malaysian
site = N+(CH3)2 of the MB dye molecules. The adsorption of MB dye Institute of Chemical and Bioengineering Technology (MICET) for
on BTAC was a multilayer as established through the isotherm and providing research facilities for this research.
intraparticle diffusion models. The possible BTAC surface macro-
molecule interaction with dye molecules are presented in Fig. 4. The Appendix A. Supplementary data
first layer of adsorption was quick as indicated by the first diffusion rate
(Kd1 = 14.694 mg/g/min1/2) due to negatively charged functional Supplementary data to this article can be found online at https://
groups (-COO-) on the BTAC and positively charged site (=N+(CH3)2) doi.org/10.1016/j.biteb.2018.07.007.
of MB dye molecules. The second layer of adsorption was due to in-
duced dipole formed due to the approach of incoming dye molecules References
but the diffusion rate of stacking of molecules reduced to 0.6082 mg/g/
min1/2, and in the third layer it further reduced and reached Ahmad, T., Danish, M., 2018. Prospects of banana waste utilization in wastewater
0.1358 mg/g/min1/2. As the layers increased on the surface of BTAC the treatment: a review. J. Environ. Manag. 206, 330–348. https://doi.org/10.1016/j.
jenvman.2017.10.061.
induced dipole effect decreased, that was observed through the diffu- Al-Bastaki, N., 2004. Removal of methyl orange dye and Na2SO4 salt from synthetic
sion rate experiments. waste water using reverse osmosis. Chem. Eng. Process. Process Intensif. 43 (12),
The adsorption of the methylene blue onto various adsorbents were 1561–1567. https://doi.org/10.1016/j.cep.2004.03.001.
Al-Othman, Z.A., Habila, M.A., Ali, R., Hassouna, M.S.E., 2013. Kinetic and thermo-
tried. The biochars were prepared from pine wood, pig manure, and dynamic studies for methylene blue adsorption using activated carbon prepared from
paper waste for MB dye removal. The maximum adsorption capacity agricultural and municipal solid wastes. Asian J. Chem. 25, 8301–8306.
was reported to be 3.99 mg/g (pine wood biochar), 16.30 mg/g (pig Babaei, A.A., Alavi, S.N., Akbarifar, M., Ahmadi, K., Ramazanpour Esfahani, A.,
Kakavandi, B., 2016. Experimental and modeling study on adsorption of cationic
manure biochar), and 1.66 mg/g (paper waste biochar) (Lonappan et al. methylene blue dye onto mesoporous biochars prepared from agrowaste. Desalin.
2016). Oak wood was activated to prepare the biochar for MB dye re- Water Treat. 57 (56), 27199–27212. https://doi.org/10.1080/19443994.2016.
moval from aqueous solution. The maximum adsorption capacity of oak 1163736.

136
M. Danish et al. Bioresource Technology Reports 3 (2018) 127–137

Bezerra, M.d.A., Arruda, M.A.Z., Ferreira, S.L.C., 2005. Cloud point extraction as a pro- cost carbon produced from Eichhornia plant: kinetic, equilibrium, and thermo-
cedure of separation and pre-concentration for metal determination using spectro- dynamic studies. Desalin. Water Treat. 53 (2), 543–556. https://doi.org/10.1080/
analytical techniques: a review. Appl. Spectrosc. Rev. 40 (4), 269–299. https://doi. 19443994.2013.841109.
org/10.1080/05704920500230880. Lahkimi, A., Oturan, M.A., Oturan, N., Chaouch, M., 2007. Removal of textile dyes from
Bhattacharyya, K.G., Sharma, A., 2005. Kinetics and thermodynamics of methylene blue water by the electro-Fenton process. Environ. Chem. Lett. 5 (1), 35–39. https://doi.
adsorption on Neem (Azadirachta indica) leaf powder. Dyes Pigments 65 (1), 51–59. org/10.1007/s10311-006-0058-x.
https://doi.org/10.1016/j.dyepig.2004.06.016. Lonappan, L., Rouissi, T., Das, R.K., Brar, S.K., Ramirez, A.A., Verma, M., ... Valero, J.R.,
Danish, M., Hashim, R., Ibrahim, M.N.M., Rafatullah, M., Ahmad, T., Sulaiman, O., 2011. 2016. Adsorption of methylene blue on biochar microparticles derived from different
Characterization of Acacia mangium wood based activated carbons prepared in the waste materials. Waste Manag. 49, 537–544. https://doi.org/10.1016/j.wasman.
presence of basic activating agents. Bioresources 6 (3), 3019–3033. 2016.01.015.
Danish, M., Hashim, R., Ibrahim, M.N.M., Sulaiman, O., 2013. Characterization of phy- Mahmoodi, N.M., 2011. Equilibrium, kinetics, and thermodynamics of dye removal using
sically activated Acacia mangium wood-based carbon for the removal of methyl or- alginate in binary systems. J. Chem. Eng. Data 56 (6), 2802–2811. https://doi.org/
ange dye. Bioresources 8 (3), 4323–4339. 10.1021/je101276x.
Danish, M., Hashim, R., Ibrahim, M.N.M., Sulaiman, O., 2014. Optimized preparation for Malkoc, E., Nuhoglu, Y., 2007. Determination of kinetic and equilibrium parameters of
large surface area activated carbon from date (Phoenix dactylifera L.) stone biomass. the batch adsorption of Cr(VI) onto waste acorn of Quercus ithaburensis. Chem. Eng.
Biomass Bioenergy 61 (0), 167–178. https://doi.org/10.1016/j.biombioe.2013.12. Process. Process Intensif. 46 (10), 1020–1029. https://doi.org/10.1016/j.cep.2007.
008. 05.007.
Danish, M., Ahmad, T., Nadhari, W.N.A.W., Ahmad, M., Khanday, W.A., Ziyang, L., Pin, Misra, D.N., 1969. Adsorption on heterogeneous surfaces: a dubinin-radushkevich equa-
Z., 2018. Optimization of banana trunk-activated carbon production for methylene tion. Surf. Sci. 18 (2), 367–372. https://doi.org/10.1016/0039-6028(69)90179-4.
blue-contaminated water treatment. Appl Water Sci 8, 9. https://doi.org/10.1007/ Mortimer, R.G., 2008. Physical Chemistry, 3rd edition. Elsevier Academic Press.
s13201-018-0644-7. Nataraj, S.K., Hosamani, K.M., Aminabhavi, T.M., 2009. Nanofiltration and reverse os-
Dursun, G., Çiçek, H., Dursun, A.Y., 2005. Adsorption of phenol from aqueous solution by mosis thin film composite membrane module for the removal of dye and salts from
using carbonised beet pulp. J. Hazard. Mater. 125 (1), 175–182. https://doi.org/10. the simulated mixtures. Desalination 249 (1), 12–17. https://doi.org/10.1016/j.
1016/j.jhazmat.2005.05.023. desal.2009.06.008.
Dwivedi, C.P., Sahu, J.N., Mohanty, C.R., Mohan, B.R., Meikap, B.C., 2008. Column Panizza, M., Barbucci, A., Ricotti, R., Cerisola, G., 2007. Electrochemical degradation of
performance of granular activated carbon packed bed for Pb(II) removal. J. Hazard. methylene blue. Sep. Purif. Technol. 54 (3), 382–387. https://doi.org/10.1016/j.
Mater. 156 (1), 596–603. https://doi.org/10.1016/j.jhazmat.2007.12.097. seppur.2006.10.010.
Fersi, C., Dhahbi, M., 2008. Treatment of textile plant effluent by ultrafiltration and/or Rafatullah, M., Sulaiman, O., Hashim, R., Ahmad, A., 2010. Adsorption of methylene blue
nanofiltration for water reuse. Desalination 222 (1), 263–271. https://doi.org/10. on low-cost adsorbents: a review. J. Hazard. Mater. 177 (1), 70–80. https://doi.org/
1016/j.desal.2007.01.171. 10.1016/j.jhazmat.2009.12.047.
Ghaedi, M., Tavallali, H., Sharifi, M., Kokhdan, S.N., Asghari, A., 2012. Preparation of low Rafatullah, M., Ahmad, T., Ghazali, A., Sulaiman, O., Danish, M., Hashim, R., 2013. Oil
cost activated carbon from Myrtus communis and pomegranate and their efficient palm biomass as a precursor of activated carbons: a review. Crit. Rev. Environ. Sci.
application for removal of Congo red from aqueous solution. Spectrochim. Acta A Technol. 43 (11), 1117–1161.
Mol. Biomol. Spectrosc. 86, 107–114. https://doi.org/10.1016/j.saa.2011.10.012. Sahu, J.N., Agarwal, S., Meikap, B.C., Biswas, M.N., 2009. Performance of a modified
Hameed, B.H., Din, A.T.M., Ahmad, A.L., 2007. Adsorption of methylene blue onto multi-stage bubble column reactor for lead(II) and biological oxygen demand re-
bamboo-based activated carbon: kinetics and equilibrium studies. J. Hazard. Mater. moval from wastewater using activated rice husk. J. Hazard. Mater. 161 (1),
141 (3), 819–825. https://doi.org/10.1016/j.jhazmat.2006.07.049. 317–324. https://doi.org/10.1016/j.jhazmat.2008.03.094.
Han, B., Runnells, T., Zimbron, J., Wickramasinghe, R., 2002. Arsenic removal from Sarasa, J., Roche, M.P., Ormad, M.P., Gimeno, E., Puig, A., Ovelleiro, J.L., 1998.
drinking water by flocculation and microfiltration. Desalination 145 (1), 293–298. Treatment of a wastewater resulting from dyes manufacturing with ozone and che-
https://doi.org/10.1016/S0011-9164(02)00425-3. mical coagulation. Water Res. 32 (9), 2721–2727. https://doi.org/10.1016/S0043-
Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Process 1354(98)00030-X.
Biochem. 34 (5), 451–465. https://doi.org/10.1016/s0032-9592(98)00112-5. Smith, S.E., 1947. The sorption of water vapor by high polymers. J. Am. Chem. Soc. 69
Houas, A., Lachheb, H., Ksibi, M., Elaloui, E., Guillard, C., Herrmann, J.-M., 2001. (3), 646–651. https://doi.org/10.1021/ja01195a053.
Photocatalytic degradation pathway of methylene blue in water. Appl. Catal. B Veliev, E.V., Öztürk, T., Veli, S., Fatullayev, A.G., 2006. Application of diffusion model for
Environ. 31 (2), 145–157. https://doi.org/10.1016/S0926-3373(00)00276-9. adsorption of azo reactive dye on pumice. Pol. J. Environ. Stud. 15 (02), 347–353.
Kaur, S., Rani, S., Mahajan, R.K., 2015. Adsorptive removal of dye crystal violet onto low-

137

You might also like