Sunteți pe pagina 1din 9

Ecotoxicology and Environmental Safety 128 (2016) 109–117

Contents lists available at ScienceDirect

Ecotoxicology and Environmental Safety


journal homepage: www.elsevier.com/locate/ecoenv

Adsorption of methyl orange from aqueous solution by aminated


pumpkin seed powder: Kinetics, isotherms, and thermodynamic
studies
Munagapati Venkata Subbaiah, Dong-Su Kim n
Department of Environmental Science and Engineering, Ewha Womans University, 11-1 Daehyun-Dong, Seodaemun-Gu, Seoul 120-750, Korea

art ic l e i nf o a b s t r a c t

Article history: Present research discussed the utilization of aminated pumpkin seed powder (APSP) as an adsorbent for
Received 19 October 2015 methyl orange (MO) removal from aqueous solution. Batch sorption experiments were carried to eval-
Received in revised form uate the influence of pH, initial dye concentration, contact time, and temperature. The APSP was char-
1 February 2016
acterized by using Fourier Transform Infrared Spectroscopy (FTIR) and Scanning Electron Microscopy
Accepted 2 February 2016
(SEM). The experimental equilibrium adsorption data were fitted using two two-parameter models
(Langmuir and Freundlich) and two three-parameter models (Sips and Toth). Langmuir and Sips iso-
Keywords: therms provided the best model for MO adsorption data. The maximum monolayer sorption capacity was
Adsorption found to be 200.3 mg/g based on the Langmuir isotherm model. The pseudo-first-order and pseudo-
Methyl orange
second-order model equations were used to analyze the kinetic data of the adsorption process and the
Kinetics
data was fitted well with the pseudo-second-order kinetic model (R2 40.97). The calculated thermo-
Isotherms
Thermodynamics dynamic parameters such as ΔG0, ΔH0 and ΔS0 from experimental data showed that the sorption of MO
onto APSP was feasible, spontaneous and endothermic in the temperature range 298–318 K. The FTIR
results revealed that amine and carboxyl functional groups present on the surface of APSP. The SEM
results show that APSP has an irregular and porous surface which is adequate morphology for dye ad-
sorption. Desorption experiments were carried to explore the feasibility of adsorbent regeneration and
the adsorbed MO from APSP was desorbed using 0.1 M NaOH with an efficiency of 93.5%. Findings of the
present study indicated that APSP can be successfully used for removal of MO from aqueous solution.
& 2016 Elsevier Inc. All rights reserved.

1. Introduction Over 7  105 tons and approximately 10,000 different types of


dyes and pigments are produced worldwide annually. It is esti-
Water pollution due to the release of various toxic chemicals mated that 10–15% of the dye is lost in the effluent during the
from industrialization and urbanization is a global problem (Gupta dyeing process (Gupta et al., 2011a, 2011b, 2011c). In general, dyes
et al., 2004). Among the various notorious toxic chemicals, dyes, are classified as anionic (direct, acid and reactive dyes), cationic
organics and pharmaceuticals are of highly concerned (Gupta and (basic dyes) and non-ionic (disperse dyes and vat dyes) (Zheng
Nayak, 2012; Gupta et al., 1998, 2011a, 2011b, 2011c, 2012a, 2012b, et al., 2015). Among them, azo dyes (anionic) with the existence of
2005, 2015a, 2015b). However, industrial effluents released from nitrogen-nitrogen double bonds, are considered to be the largest
textile, paint, paper, varnishes, ink, plastics, pulp, cosmetics, tan- and most versatile class of organic dyes (Netpradit et al., 2004;
nery, and plastic are one of the key causes of water pollution, since Gao et al., 2013). But their degradation is difficult due to their
these runoffs comprises highly colored substances. Their discharge complicated aromatic structure and poor biodegradability (Fang
et al., 2010). Methyl orange (MO), a typical water-soluble anionic
into the hydrosphere possesses a significant source of pollution
azo dye, with IUPAC name of Sodium 4-[[4-(dimethyl amino)
due to their visibility even at very low concentrations. This is due
phenyl] diazenyl] benzene sulfonate. MO is commonly present in
to their recalcitrance nature, giving undesirable color to the water,
effluent discharges from textile, food, pharmaceutical, printing and
reducing sunlight penetration, resisting photochemical and bio-
paper manufacturing industries (Cheah et al., 2013). Due to the
logical attack, and their degradation products being toxic or even
toxicity and persistence, these discharges can cause a serious
mutagenic, and carcinogenic (Gupta et al., 2013a, 2013b). threat to physic-chemical properties of fresh water and to aquatic
life. Therefore, it is necessary to provide suitable technology for
n
Corresponding author. the wastewater treatment.
E-mail address: dongsu@ewha.ac.kr (D.-S. Kim). The conventional methods for the removal of dyes from

http://dx.doi.org/10.1016/j.ecoenv.2016.02.016
0147-6513/& 2016 Elsevier Inc. All rights reserved.
110 M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117

wastewaters include coagulation and flocculation (Panswad and 2. Materials and methods
Wongchaisuwan, 1986), oxidation or ozonation (Malik and Saha,
2003; Koch et al., 2002), membrane separation (Ciardelli et al., 2.1. Adsorbent preparation and chemical modification
2000). However, most of these techniques suffer from high oper-
ating cost and sludge generation. Adsorption is one of the most Pumpkin seeds were used as the raw biomass for the pre-
paration of adsorbent, and the seeds were collected from a local
successful technique for color removal from wastewater among all
market in South Korea. The collected seeds were washed several
the techniques so far developed for dye removal (Ali and Gupta,
times with boiled water and finally with deionized water to re-
2007; Gupta et al., 2009). The major advantages of adsorption over
move any adhering dirt. It was then oven dried at 70 °C for 24 h to
conventional treatment methods include; low cost, ease of op- a constant weight. The oven dried seeds were ground well to a fine
eration, high efficiency, minimization of chemical or biological powder, sieved to 20–30 mesh fractions. This pumpkin seed
sludge, no additional nutrient requirement, regeneration of bio- powder was named as PSP.
sorbent and possibility of sorbate recovery (Saleh et al., 2011; The aminated biomass was prepared by employing a reported
Khani et al., 2010; Gupta et al., 2011a, 2011b, 2011c, 2012a, 2012b). method (Brady and Duncan, 1994) of esterification. For the pre-
The agricultural and forestry products have a great potential to be paration of aminated pumpkin seed powder (APSP), 5 g of the well
used as biosorbents. Some of the reported low cost biosorbents washed raw PSP was suspended in 100 mL of ethanolamine, and
include palm ash (Ahmad et al., 2007), de-oiled soya (Mittal et al., 20.8 mL of concentrated hydrochloric (HCl) acid was added to the
2005), broad bean peels (Hameed and EI-Khaiary, 2008), durian suspension. The reaction mixture was agitated on a rotary shaker
peel (Hameed and Hakimi, 2008), rice husk (McKay et al., 1999), at 150 rpm for 6 h. This treatment was expected to result in the
almond shell (Senturk et al., 2010) and neem sawdust (Khattri and increase of amine groups on the biomass by the ester link to the
carboxyl groups via the general equation:
Singh, 2000), etc. These types of biosorbents contain poly-
HCl, room temp
saccharides and proteins having various functional groups such as Biomass − COOH + NH2 CH2 CH2 OH → Biomas
6h
carboxyl, hydroxyl, and phosphates (Bharathi and Ramesh, 2013;
Asgher, 2012). The biosorption of dye molecules by these materials s − COOCH2 CH2 NH2 + H2 O
might be associated with these functional groups. The chemically treated suspension was then centrifuged and
Pumpkin, a gourd-like squash of Cucurbitaceae’s family, is one sequentially washed with distilled water. The treated biomass was
of the widely cultivated plant species for its fruit. The fruit of dried at 60 °C inside a convection oven for 24 h. This treated bio-
pumpkin is one the most important vegetables in traditional mass was named as APSP. The resultant dried APSP was stored in a
agricultural systems in the world. The fruit represents rich sources desiccator, and subsequently used as an adsorbent in the sorption
of pectin-type dietary fiber, antioxidants (carotenoids), vitamins experiments.
(C, E, B6, K, thiamine, and riboflavin), and minerals (potassium,
phosphorus, magnesium, iron, zinc and selenium) (Celekli et al., 2.2. Chemicals and equipment
2014). Pumpkin seeds have been valued as a special source of the
mineral zinc, and the World Health Organization recommends All chemicals used in this work were of analytical reagent grade
their consumption as a good way of obtaining this nutrient. and used without purification. Double deionized water (Milli-Q
Pumpkin seeds, pumpkin seed extracts, and pumpkin seed oil have Millipore 18.2 MΩ cm  1 conductivity) was used for all dilutions.
long been valued for their anti-microbial benefits, including their MO dye [formula weight: C14H14N3NaO3S, molecular weight:
327.34, λmax: 464 nm] was purchased from Junsei Chemicals, Ja-
anti-fungal and anti-viral properties. Thus, in the present study we
pan. Working standards were prepared by progressive dilution of
have chosen pumpkin seeds as biomass source for manufacturing
the stock MO solution. HCl, NaOH and buffer solutions (Dae Jung
efficient adsorbent.
Chemicals, Korea) were used to adjust the solution pH. A pH meter
Although adsorbents have been regarded as promising mate-
(pH-240L, NEOMET, Korea) was used for pH measurements. The
rials for pollutant removal, raw adsorbents (without modification) pH meter was calibrated using standard buffer solutions of pH 4.0,
suffer from their low sorption efficiencies. In recent years, in- 7.0 and 10.0. Infrared spectra of the PSP and APSP sample were
creased interest has been focused on enhancing the sorption ca- obtained using Fourier Transform Infrared Spectrometer (BIO-RAD,
pacity of biomass by introducing various chemical functional FTS-135, USA). For IR spectral studies, 10 mg of sample was mixed
groups (Reddy et al., 2012; Sajab et al., 2013; Kumar et al., 2014). and ground with 100 mg of KBr and made into a pellet. The
As sorption mainly takes place on the biomass surface, enhancing/ background absorbance was measured by busing a pure KBr pellet.
activating the binding sites on the surface would be an effective The morphology of PSP and APSP was analyzed by Scanning
approach for improving the adsorption capacity. Considering the Electron Microscopy (JEOL, JSM-7600F, Japan). The samples were
aforesaid reasons in the present study we have chosen pumpkin first sputter-coated with homogeneous gold layer and then loaded
seeds a raw biomass sources and the surface of biomass was fur- onto a copper substrate. The dye concentrations in the samples
ther functionalized to enhance the sorption capacity. were determined using UV/Vis spectrophotometer (Optizen Pop,
The main objective of the present work is to investigate the Korea) at maximum wavelengths of 464 nm.
potential of APSP as an adsorbent material for the removal of MO
from aqueous solutions. The different parameters such as effect of 2.3. Batch adsorption procedure
the pH, initial dye concentration, contact time, and temperature
The adsorption of MO on the APSP was investigated as an effect
that influences the sorption processes of MO onto APSP were
of pH, initial dye concentration, contact time, and temperature.
evaluated. The isotherm, kinetics as well as thermodynamic
1000 mg/L of MO stock solution was prepared and was used fur-
parameters for the adsorption of MO onto the APSP were calcu- ther to obtain a standard solution by appropriate dilution of stock
lated. Characterization of adsorbent was carried using FTIR and solution. In sorption experiments, 0.05 g of the sorbent was
SEM analysis. brought into contact with 30 mL of MO solution in a 50 mL falcon
tube. The pH values of the solutions were adjusted using small
M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117 111

volumes (1–2 drops) of 0.1 M HCl and 0.1 M NaOH solutions to CiVi − Cf Vf
reach the desired pH. The total added volume to get the desired pH q=
M (1)
was also considered for calculation of initial concentration. All
tubes were agitated in an electrical thermostatic reciprocating where q is the uptake of MO (mg/g), Ci and Cf are the initial and
shaker at 160 rpm. Sorption kinetic experiments were conducted final MO concentrations in the solution (mg/L), Vi and Vf are the
with an initial concentration of 300 mg/L at 298 K. Samples were initial and final (initial plus added HCl or NaOH solutions) solution
collected at various shaking time intervals until the concentration volumes, and M is the mass of sorbent (g) used in MO sorption.
of MO in the dilute phase became constant. Sorption isotherm
experiments were also carried out with different initial dye con- 2.4. Regeneration and reusability of adsorbent
centrations varying from 100 to 1000 mg/L at different tempera-
tures (298, 308 and 318 K) for 4 h. At the end of adsorption, 1 mL The recovery and reusability of adsorbent material is an im-
sample was centrifuged at 3000 rpm for 10 min on a centrifuge. portant parameter related to the application potential of adsorp-
The filtrate was diluted in polythene tubes before analysis. The tion technology. In this work consecutive batch adsorption and
concentration of remaining MO in the adsorption medium was desorption experiments were performed using the same biosor-
determined by using a UV/Vis spectrophotometer. Each experi- bent in order to test for its ability to be reutilized. For this purpose,
ment was repeated three times, and the mean values are used in 0.05 g of adsorbent was contacted with 30 mL MO solution
this study. (300 mg/L) at pH 3.0. APSP saturated with MO was removed from
The MO uptake (q) was calculated from the difference between solution and transferred into 50 mL falcon tubes, to this 30 mL of
the concentration of MO before and after sorption using the fol- 0.1–1 M NaOH solution was added and the tubes were shaken for
lowing equation: 4 h. Then the mixer solution was centrifuged and the

Fig. 1. FTIR spectrum of (A) PSP, and (B) APSP.


112 M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117

concentration of the MO released into eluent solution was ana-


lyzed by using UV/Vis spectrophotometer. Adsorbent was washed
repeatedly with deionized water to remove any residual desorbing
solution. The sorbent was placed into dye solution for the suc-
ceeding adsorption cycle after drying to a constant weight. The
cycles were repeated five times using the same adsorbent. Deso-
rption efficiency was calculated by using the following equation:
Amount of MO desorbed
Desorption efficiency= × 100
Amount of MO adsorbed (2)

3. Results and discussion

3.1. Characterization of APSP

FTIR spectroscopy was used to identify the functional groups


that are responsible for adsorption and also for confirming the
successful modification of PSP. FTIR spectral data of PSP and APSP
was obtained in the range of 4000–400 cm  1 and the data was
shown in Fig. 1. A depicts the IR spectra of the PSP was shown in
Fig. 1A and the spectral analysis results were discussed as follows:
The band observed at around 2919 cm  1 was attributed to the C–
H stretching. The peak at 1732 cm  1 refers to the characteristics of
the C¼ O stretches of the carbonyls mainly those of ketones and
esters. The peak at 1445 cm  1 is due to C–H bending vibrations of
the alkyl group. The band at about 1326 cm  1 corresponds to the
C–N stretching. The peaks at 1162 cm  1 attributed to the
stretching vibrations of C–O group. Fig. 1B depicts the FTIR spec-
trum of the APSP. After modification, a new peak was observed at
3415 cm  1 (assigned to the N–H stretching vibration mode of the
amine functional groups) for APSP. Some shifts in wave numbers
from 2919 to 2925 cm  1, 1732 to 1746 cm  1, 1445 to 1465 cm  1,
1326 to 1365 cm  1 and 1162 to 1166 cm  1 were noticed in the
spectra of PSP and APSP. These significant changes in the wave
numbers of specific peaks suggested that amino and carboxyl
groups could be presented in the APSP. This observation is an
agreement with similar previously reported studies (Saleh et al., Fig. 2. SEM micrographs of (A) PSP, and (B) APSP.

2014; Saleh and Gupta, 2012; Mahmoodian et al., 2015; Gupta


et al., 2015a, 2015b; Saleh, 2015). molecule exists as negatively charged species until pH 3.46 after-
To evaluate the textual structure of adsorbent surface, SEM wards MO exists as neutral and in high pH conditions as slightly
micrographs of PSP and APSP were shown in Fig. 2. The surface positive charged species. Therefore at studied pH 3.0 the ad-
morphology of PSP was differed from that of APSP. From the sur- sorbent surface has maximum positive charge and also the MO
face morphology of PSP it was observed that powder was an as- molecule has maximum negative charge which will contribute to
semblage of fine particles, which did not have regular, fixed shape the high adsorption capacity in an expected electrostatic way. As
and size. The particles were of various dimensions and all of them the pH of the system increases, the number of positively ionized
contained a large number of steps and kinks on the external sur- charged amine functional groups decreases and negatively
face, with broken edges. The size of the voids in the PSP was in- charged sites increases on the adsorbent on the other hand ne-
creased after modification and some distortion of the shape could gative charge on the dye molecule also decreases. The decrease of
be seen in the SEM of APSP. MO adsorption can also be as at higher pH conditions repulsion
occurs due to the competition between the presence of the
3.2. Effect of pH abundant OH  ions in basic solution and MO anions. A negatively
charged site on the adsorbent surface does not favor the adsorp-
pH of the solution is one of the vital factor controlling the tion of anionic dyes due to the electrostatic repulsion (Arami et al.,
adsorption capacity, pH of the solution influences not only the 2005; Fouad and Omar el Farouk, 2015).
surface charge of the adsorbent but also the solubility of dyes Adsorption mechanism may be as follow:
(Reddy and Lee, 2013). Considering this in present study we have First MO is dissolved in an aqueous solution after which the
studied effect of pH on MO removal and effect of initial solution sulfonate groups of MO dye (R-SO3Na) become dissociated and
pH on MO removal was illustrated Fig. 3(A). From these results it converted into anionic dye ions.
was observed that the amount of adsorption of solute decreases (i) R − SO3Na + H2O → R − SO−3 + Na+
from 126.1 to 4.9 mg/g as the pH increases from 3 to 11. The Second the adsorption process is performed due to electrostatic
maximum adsorption of MO on the APSP was observed at pH 3.0. attractions between the adsorbent surface and the MO anions
At this pH a considerable high electrostatic attraction exists be- (Reddy and Lee, 2013).
tween the positively charged surface of the adsorbent and anionic (ii) RNH+3 + RSO−3 → RNH+3 O3SR
dye, due to the ionization of functional groups (amines) of ad- Since the adsorbent shows high adsorption capacities at pH 3.0,
sorbent and anionic dye molecules. pKa of MO was 3.46, the MO all the further studies were carried out at this pH. A similar type of
M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117 113

140 3.5. Adsorption kinetics


(A)
120 In order to investigate the mechanism of the adsorption pro-
cess pseudo-first and second-order kinetic models were used to fit
100
the kinetic experimental data. The pseudo-first-order equation
80 assumes the adsorption is of one adsorbate molecule onto active
site on the adsorbent surface while in pseudo-second-order model
qe [mg/g]

60 one adsorbate molecule is adsorbed onto two active sites. The


non-linear form of the pseudo-first-order (Lagergren, 1898) and
40 pseudo-second-order (Ho and McKay, 1999) kinetic rate equations
are given as (Eqs. (3) and (4)):
20
Pseudo-first-order model
0 qt = q1(1 − exp( − k1t )) (3)

Pseudo-second-order model
2 4 6 8 10 12
q22k2t
Final pH qt =
1 + q2k2t (4)
140
(B)
where q1 and q2 are the amount of dye sorbed at equilibrium (mg/
120 g), qt is the amount of dye sorbed at time t (mg/g), k1 is the
pseudo-first-order equilibrium rate constant (1/min), and k2 is the
100 pseudo-second-order equilibrium rate constant (g/mg min).
The estimated kinetic parameters and coefficient of determi-
80 nation (R2) from the pseudo-first-order and pseudo-second-order
q [mg/g]

models are summarized in Table 1. The low R2 value and the dif-
60 ference between experimental qe and theoretical q1 indicate that
the pseudo-first-order model was not well suited to describe the
40 adsorption of MO by APSP. On the other hand, the R2 value (0.9741)
for the pseudo-second-order model was relatively higher than that
20 of the pseudo-first-order model. Moreover, the q2 value calculated
by the pseudo-second-order model was close to the experimental
0 qe value. Thus, these results suggest that the pseudo-second-order
0 50 100 150 200 250 300 350 model provided a good correlation for the adsorption of MO onto
Time [min] APSP.
Fig. 3. Effect of (A) pH, and (B) contact time on the adsorption of MO onto APSP.
3.6. Adsorption isotherms

behavior was also reported for the adsorption of the MO on dif- Adsorption isotherm is a functional expression that correlates
ferent adsorbents (Asuha et al., 2010; Tian et al., 2015). the amount of solute adsorbed per unit weight of the adsorbent
and the concentration of an adsorbate in bulk solution at a given
temperature under equilibrium conditions. It is important to es-
3.3. Effect of contact time tablish the most appropriate correlations for the batch equilibrium
data using empirical or theoretical equations as it plays a func-
As seen in Fig. 3(B), it is evident that time has significant in- tional role in predictive modeling procedures for analysis and
fluence on the adsorption of dye. It can be observed from the design of adsorption systems. The four most common isotherms
figure that adsorption of MO was quite rapid in the first 60 min, for describing solid-liquid sorption systems are the Langmuir,
then gradually increased with the prolongation of contact time. Freundlich (two-parameter isotherms), Sips and Toth (three-
After 110 min of contact time equilibrium was observed and with parameter isotherms). Therefore, the experimental data were fit-
further increase in contact time there is no additional enhance- ted to these equilibrium models.
ment in sorption capacity. Based on these results, 110 min was
taken as the equilibrium time in batch adsorption experiments. 3.6.1. Langmuir isotherm
Langmuir sorption model serves to estimate the maximum
uptake values where they cannot be reached in experiments.
3.4. Effect of temperature
Langmuir’s model does not take into account the variation in ad-
sorption energy, but it is the simplest description of the adsorption
The adsorption experiments were carried out at three different
temperatures (298, 308 and 313 K) with the different concentra- Table 1
tions to study the effect of temperature on the adsorption of MO Pseudo-first-order and pseudo-second-order model constants for the adsorption of
onto APSP. The adsorption capacity of APSP was increased from MO on APSP.
143.7 to 200.3 mg/g as the temperature was increased from 298 to
Sorbent qe (mg/ Pseudo-first-order Pseudo-second-order
313 K. The increase is due to increase in dye mobility that may g)
occur at high temperature between the adsorbent and the ad- q1 (mg/ k1 (L/ R2 q2 (mg/ k2 (g/mg R2
sorbate (Ahmad et al., 2014). Adsorption of MO onto APSP is an g) min) g) min)
endothermic process; the adsorption capacity increased with in-
APSP 114 108 0.155 0.8553 113 0.002 0.9741
crease in solution temperature.
114 M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117

process. It is based on the physical hypothesis that the maximum The 1/n values between 0 and 1 indicate that the adsorption of MO
biosorption capacity consists of a monolayer adsorption, that there onto APSP is favorable under the conditions studied.
are no interactions between adsorbed molecules, and that the
adsorption energy is distributed homogeneously over the entire 3.6.3. Sips isotherm
coverage surface. Sips isotherm is a combination of the Langmuir and Freundlich
The Langmuir sorption isotherm has been successfully applied isotherm type models and expected to describe heterogeneous
to many pollutant sorption processes and has been the most surfaces much better. At low sorbate concentrations it effectively
widely used to describe the sorption of a solute from a liquid so- reduces to the Freundlich isotherm and thus does not obey Henry’s
lution. A basic assumption of the Langmuir theory is that sorption law. At high sorbate concentrations, it achieves a monolayer
takes place at specific homogeneous sites on the surface of the sorption capacity characteristic of the Langmuir isotherm. The
sorbent. It is then assumed that once a sorbate molecule occupies model can be written as (Sips, 1948).
a site, no further sorption can take place at that site. The rate of
sorption to the surface should be proportional to a driving force KSC fβS
q=
and area. The driving force is the concentration in the solution, 1 + aS C fβS (7)
and the area is the amount of bare surface. The Langmuir equation
is (Langmuir, 1918): where KS is the Sips model isotherm constant (L/g), aS the Sips model
constant (L/mg) and βS the Sips model exponent. The exponent va-
qmKLCe
qe = lues were close to unity meaning that MO sorption data obtained in
1 + KLCe (5) this study is more of Langmuir form rather than that of Freundlich.
where qe is metal concentration on sorbent (mg/g) at equilibrium,
Ce is the equilibrium metal concentration in solution (mg/L), qm is 3.6.4. Toth isotherm
monolayer adsorption capacity of sorbent (mg/g), and KL is the Toth isotherm, derived from potential theory, has proven useful
Langmuir constant (L/mg) related with the sorption free energy. in describing sorption in heterogeneous systems such as phenolic
The qmax value was increased from 143.7 to 200.3 mg/g (Table 2) compounds on carbon. It assumes an asymmetrical quasi-Gaussian
with an increase in the temperature from 298 to 318 K. This in- energy distribution with a widened left-hand side, i.e., most sites
crease in sorption capacity may be attributed to a rise in kinetic have sorption energy less than the mean value. It can be re-
energy of the sorbent particles due to the increase in temperature. presented as (Toth, 1971):
This rise in kinetic energy increases the frequency of collisions qmaxbT Cf
between the adsorbent and the sorbate, resulting in enhanced q=
sorption on to the surface of the sorbent. [1 + (bT Cf )1/ nT ]nT (8)

where bT is the Toth model constant and nT the Toth model ex-
3.6.2. Freundlich isotherm
ponent. It is obvious that for nT ¼ 1 this isotherm reduces to the
Freundlich developed an empirical equation to describe the
Langmuir sorption isotherm equation.
adsorption process. The Freundlich isotherm was based on the
The correlation coefficients, R2 and the Chi-square (χ2) test
assumption that the adsorbent had a heterogeneous surface
were also carried out to find the best fit among the adsorption
composed from different classes of adsorption sites, with adsorp-
isotherm models that are used. The equation for evaluating the
tion on each class of sites following the Langmuir isotherm.
best fit model is to be evolved as
Freundlich demonstrated that the ratio of the amount of solute
adsorbed onto a given mass of an adsorbent to the concentration ⎛ (q − q )2 ⎞
of the solute in the solution was not constant at different solution χ2 = ∑ ⎜⎜ e e, m ⎟

⎝ qe, m ⎠
concentrations. This isotherm does not predict any saturation of
the sorbent by the sorbate; thus infinite surface coverage is pre- where qe,m equilibrium capacity obtained by calculating from a model
dicted mathematically, indicating multilayer sorption of the sur- (mg/g) and qe experimental data of equilibrium capacity (mg/g).
face. The Freundlich model (Freundlich, 1906) is: The different isotherm parameters along with R2 and χ2 values
qe = Kf Ce1/ n are given, respectively, in Table 2. High R2 values and low χ2 values
(6)
for the Langmuir isotherm model indicate that the adsorption of
where Kf (mg/g) is a constant relating the adsorption capacity and MO onto APSP follows the Langmuir isotherm model. The ex-
1/n is an empirical parameter relating the adsorption intensity, cellent fit of the Langmuir isotherm to the experimental adsorp-
which varies with the heterogeneity of the material. The values of tion data confirms that the adsorption is monolayer; adsorption of
Kf increased from 31.8 to 50.3 mg/g (Table 2) with an increase in each molecule has equal activation energy and that sorbate-sor-
temperature of the solution from 298 to 318 K. As the Kf is a bate interaction is negligible. On the basis of correlation coeffi-
measurement of adsorption capacity, the increase in the value cients and chi-square test, the experimental data better described
again confirms that the adsorption process of MO onto APSP is an by the Sips isotherm model followed by Toth. Besides, the Sips
endothermic process. The 1/n values are found in the range of constant confirms the fact that the isotherm is more approaching
0.23–0.22, when the temperature was altered from 298 to 318 K. the Langmuir than the Freundlich isotherms.

Table 2
Langmuir, Freundlich, Sips and Toth isotherm constants for the adsorption of MO on APSP.

T Langmuir Freundlich Sips Toth

qmax b R2 χ2 Kf n R2 χ2 KS aS βS R2 χ2 qmax bT nT R2 χ2

298 143.7 0.021 0.9934 4.2 31.8 4.385 0.9342 13.2 0.7147 0.0053 1.391 0.9992 1.58 132.9 0.015 0.637 0.9985 2.2
308 183.2 0.031 0.9955 4.5 46.3 4.659 0.9195 19.3 2.3200 0.0133 1.275 0.9996 1.51 172.9 0.023 0.705 0.9994 1.8
318 200.3 0.035 0.9987 2.7 50.3 4.559 0.9321 19.6 4.9265 0.0252 1.119 0.9996 1.73 194.7 0.031 0.856 0.9994 1.9
M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117 115

Table 3 100
Comparison of adsorption capacity of APSP with different adsorbents. (A)
Biosorbent qmax pH References
80
Bottom ash 3.618 3.0 Mittal et al. (2007)

Desorption rate %
De-Oiled Soya 16.664 3.0 Mittal et al. (2007)
Banana peel 21.0 6.0-7.0 Annadurai et al. 60
(2002)
Orange peel 20.5 4 7.0 Annadurai et al.
(2002)
40
AC/ferrospinel composite 95.8 – Ai and Jiang (2010)
Modified sporopollenin 5.23 – Ayar et al. (2007)
Multiwalled Carbon nanotubes 52.86 2.3 Yao et al. (2011)
Hypercrosslinked polymeric 76.9 – Huang et al. (2008) 20
adsorbent
m-CS/Ƴ-Fe2O3/MWCNTs 66.09 – Zhu et al. (2010)
Activated clay 16.78 7.0 Ma et al. (2013) 0
Modified ultrafine coal powder 18.52 – Liu et al. (2009)
0.0 0.2 0.4 0.6 0.8 1.0 1.2
Modified wheat straw 50.4 3.0 Su et al. (2014)
Modified HNTs 94.34 6.7 Liu et al. (2012) NaOH Concentration (M)
Acrylic acid grafted Ficus carica fiber 51.55 4.0 Gupta et al. (2013b)
Maghemite/chitosan nanocomposite 29.41 3.0 Jiang et al. (2012) 120
films (B)
Lapindo volcanic mud 333.3 3.0 Jalil et al. (2010) Adsorption
Magnetic chitosan beads 779 4.0 Obeid et al. (2013) 100 Desorption
Phragmites australis activated 217 3.0 Chen et al. (2010)

Adsorption/desorption %
carbon
APSP 200.3 3.0 This study 80

60
3.7. Comparison of APSP with other biosorbents

Adsorption capacities of various adsorbents towards MO as 40


reported in literature were presented in Table 3. A comparison
between this work and other reported data from the literature
20
shows that APSP is a better adsorbent for MO compared to other
adsorbents. Therefore, it could be safely concluded that the APSP
adsorbent has a considerable potential for the removal of MO from 0
1 2 3 4 5
an aqueous solution.
Cycle number
3.8. Thermodynamic studies Fig. 4. Desorption studies of MO: (A) desorption time, and (B) number of adsorp-
tion–desorption cycles.
The determination of thermodynamics parameters has a great
importance to evaluate spontaneity and heat change for the ad- values of ΔG0 with an increase in temperature showed an ad-
sorption reactions (Tanyildizi, 2011). The free energy change (ΔG0) sorption process is more favorable at higher temperatures. The
of the adsorption reaction is given by the following equation. entropy (ΔS0) value indicates whether a particular reaction pro-
ceeds faster or slower than another individual reaction (Hameed
ΔG 0 = − RT ln KL (9) et al., 2007; Ozcan and Ozcan, 2004). The positive value of ΔS0
3 (0.123 kJ/mol K) resulted from the increased randomness due to
where R is the universal gas constant (8.314  10 kJ/mol K), T is
the adsorption of MO, which suggested good affinity of MO to-
an absolute temperature (K) and KL is the Langmuir constant at
wards the APSP and increased randomness at the solid-solution
temperature T.
interface during the fixation of MO on the active site of the APSP.
The enthalpy (ΔH0) and entropy (ΔS0) change values were
The positive value of ΔH0 (33.37 kJ/mol) for the process indicates
estimated from the following equations:
that the adsorption of APSP for MO was an endothermic process.
ΔG 0 = ΔH 0 − T ΔS 0 (10)
3.9. Desorption of MO and reuse of adsorbent
ΔH 0 ΔS 0
ln KL = − + To reduce the adsorption processing cost and to open the
RT R (11)
possibility of recovering the dyes extracted from the liquid phase,
The values of ΔG0 were calculated from Eq. (9). The values of it is desirable to regenerate the adsorbent material. Desorption
ΔH0 and ΔS0 can be calculated from the slope and the intercept of studies were performed with different NaOH concentrations and
the plot of ln KL versus 1/T, respectively (figure not shown). The the results are shown in Fig. 4(A). From the results of this study,
change of free energy for physisorption is generally between  20 with the increasing of NaOH concentration, the desorption rate
and 0 kJ/mol, the physisorption together with chemisorption is at decreased. The maximum percentage recovery of MO was 93.5%
the range of  20 to  80 kJ/mol and chemisorption is at a range of with 0.1 M NaOH solution. In this study 0.1 M NaOH was used as a
80 to  400 kJ/mol (Baseri et al., 2012). Gibbs free energy change regenerating agent. The regenerated adsorbent was reused for up
(ΔGo) was also calculated to be  2.761,  4.473 and  5.201 kJ/ to five adsorption-desorption cycles and the results are illustrated
mol for 298, 308 and 318 K, respectively. The negative values of in Fig. 4(B). An efficiency of 93.5% recovery of MO was obtained
overall free energy changes during the adsorption process indicate with 0.1 M NaOH in the first cycle and is therefore suitable for
the spontaneous nature of the adsorption process. The decrease in regeneration of adsorbent. There is a gradual decrease in MO
116 M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117

adsorption with an increase in the number of cycles. After five Ayar, A., Gezici, O., Kucukosmanoglu, M., 2007. Adsorptive removal of methylene
cycles, the adsorption capacity decreased from 77.1% to 65.3% blue and methyl orange from aqueous media by carboxylated diaminoethane
sporopollenin: on the usability of an aminocarboxilic acid functionality-bearing
while the recovery of MO ions decreased from 93.5% in the first solid-stationary phase in column techniques. J. Hazard. Mater. 146, 186–193.
cycle to 81.3% in the fifth cycle. There was slight decrease in both Baseri, J.R., Palanisamy, P.N., Kumar, P.S., 2012. Adsorption of basic dyes from syn-
the percentage of MO adsorbed and the percentage of MO des- thetic textile effluent by activated carbon prepared from Thevetiaperuviana.
Indian J. Chem. Technol. 19, 311–321.
orbed from the first to fifth cycle. This small decrease may be due Bharathi, K.S., Ramesh, S.T., 2013. Removal of dyes using agricultural waste as low-
to the MO bound to the APSP through electrostatic interactions are cost adsorbents: a review. Appl. Water Sci. 3, 773–790.
not recoverable fully in subsequent cycles. The results showed that Brady, D., Duncan, J.R., 1994. Binding of heavy metals by the cell walls of Sacchar-
omyces cerevisiae. Enzyme Microb. Technol. 16, 633–638.
the APSP could be used repeatedly in MO adsorption studies with Celekli, A., Celekli, F., Cicek, E., Bozkurt, H., 2014. Predictive modeling of sorption
a small loss in the total adsorption capacity. and desorption of a reactive azo dye by pumpkin husk. Environ. Sci. Pollut. Res.
21, 5086–5097.
Cheah, W., Hosseini, S., Khan, M.A., Chuah, T.G., Choong, T.S.Y., 2013. Acid modified
carbon coated monolith for methyl orange adsorption. Chem. Eng. J. 215–216,
4. Conclusions 747–754.
Chen, S., Zhang, J., Zhang, C., Yue, Q., Li, Y., Li, C., 2010. Equilibrium and kinetic
studies of methyl orange and methyl violet adsorption on activated carbon
This study revealed that APSP was an effective adsorbent for
derived from Phragmites australis. Desalination 252, 149–156.
MO removal from aqueous solutions. The batch adsorption para- Ciardelli, G., Corsi, L., Marcucci, M., 2000. Membrane separation for wastewater
meters such as pH of solution, initial dye concentration, contact reuse in the textile industry. Resour. Conserv. Recycl. 31, 189–197.
time, and temperature were effective on the adsorption process. Fang, R., Cheng, X., Xu, X., 2010. Synthesis of lignin-base cationic flocculant and its
application in removing anionic azo-dyes from simulated wastewater. Bior-
The equilibrium sorption phenomena were found to be well de- esour. Technol. 101, 7323–7329.
scribed by both Langmuir and Sips isotherm models. The max- Fouad, K., Omar el Farouk, B., 2015. Removal of methyl orange from aqueous so-
imum monolayer adsorption capacity was found to increase from lution via adsorption on cork as a natural and low-coast adsorbent: equili-
brium, kinetic and thermodynamic study of removal process. Desalin. Water
143.7 to 200.3 mg/g with increase in temperature from 298 to Treat 53, 3711–3723.
318 K. The Langmuir constants were used for calculating different Freundlich, H.M.F., 1906. Uber die adsorption in lasugen. J. Phys. Chem. 57, 385–470.
thermodynamic parameters such as ΔG0, ΔH0, and ΔS0. The ne- Gao, H., Zhao, S., Cheng, X., Wang, X., Zheng, L., 2013. Removal of anionic azo dyes
from aqueous solution using magnetic polymer multi-wall carbon nanotube
gative values of ΔG0 indicate the spontaneity of the process, nanocomposite as adsorbent. Chem. Eng. J. 223, 84–90.
whereas the positive values of ΔH0 and ΔS0 indicate the en- Gupta, V.K., Ali, I., Saini, V.K., Gerven, T.V., Van der Bruggen, B., Vandecasteele, C.,
dothermic nature and increase in randomness of the process, re- 2005. Removal of dyes from wastewater using bottom ash. Ind. Eng. Chem. Res.
44, 3655–3664.
spectively. The experimental results indicated that the adsorption Gupta, V.K., Ali, I., Saleh, T.A., Nayak, A., Agarwal, S., 2012a. Chemical treatment
behavior was well represented by a pseudo-second-order kinetic technologies for waste-water recycling-an overview. RSC Adv. 2, 6380–6388.
model. Five adsorption/desorption cycles were carried out with Gupta, V.K., Carrott, P.J.M., Ribeiro Carrott, M.M.L., Suhas, 2009. Low-cost ad-
sorbents: Growing approach to wastewater treatment a review. Crit. Rev. En-
0.1 M NaOH as the desorbing agent without any loss of adsorbent
viron. Sci. Technol. 39, 783–842.
or appreciable reduction in adsorption capacity. The results in- Gupta, V.K., Gupta, B., Rastogi, A., Agarwal, S., Nayak, A., 2011a. Pesticides removal
dicated that APSP could be a potential alternative adsorbent for from waste water by activated carbon prepared from waste rubber tire. Water
Res. 45, 4047–4055.
the removal of MO from aqueous solution.
Gupta, V.K., Jain, R., Nayak, A., Agarwal, S., Shrivastava, M., 2011b. Removal of the
hazardous dye-Tartrazine by photodegradation on titanium dioxide surface.
Mater. Sci. Eng. C 31, 1062–1067.
Acknowledgments Gupta, V.K., Jain, R., Saleh, T.A., Nayak, A., Malathi, S., Agarwal, S., 2011c. Equilibirum
and thermodynamic studies on the removal and recovery of Safranine-T Dye
from industrial effluents. Sep. Sci. Technol. 46, 839–846.
This research was supported by the R&D Program for Society of Gupta, V.K., Kumar, R., Nayak, A., Saleh, T.A., Barakat, M.A., 2013a. Adsorptive re-
moval of dyes from aqueous solution onto carbon nanotubes: a review. Adv.
the National Research Foundation (NRF) funded by the Ministry of
Colloid Interface Sci. 193–194, 24–34.
Science, ICT & Future Planning (Grant no.: NRF- Gupta, V.K., Mittal, A., Jhare, D., Mittal, J., 2012b. Batch and bulk removal of ha-
2013M3C8A3078596). zardous colouring agent Rose Bengal by adsorption techniques using bottom
This study was supported by the R&D Center for Valuable Re- ash as adsorbent. RSC Adv. 2, 8381–8389.
Gupta, V.K., Nayak, A., 2012. Cadmium removal and recovery from aqueous solu-
cycling (Global-Top Environmental Technology Development Pro- tions by novel adsorbents prepared from orange peel and Fe2O3 nanoparticles.
gram) funded by the Ministry of Environment (Project no.: GT-11- Chem. Eng. J. 180, 81–90.
C-01-070-0). Gupta, V.K., Nayak, A., Agarwal, S., 2015a. Bioadsorbents for remediation of heavy
metals: current status and their future prospects. Environ. Eng. Res. 20 (1),
001–018.
Gupta, V.K., Pathania, D., Sharma, S., Agarwal, S., Singh, P., 2013b. Remediation and
References recovery of methyl orange from aqueous solution onto acrylic acid grafted Ficus
carica fiber: isotherms, kinetics and thermodynamics. J. Mol. Liq. 177, 325–334.
Gupta, V.K., Saleh, T.A., Pathania, D., Rathore, B.S., Sharma, G., 2015b. A cellulose
Ahmad, A.A., Hameed, B.H., Aziz, N., 2007. Adsorption of direct dye on palm ash: acetate based nanocomposite for photocatalytic degradation of methylene blue
kinetic and equilibrium modeling. J. Hazard. Mater. 14 (1), 70–76. dye under solar light. Ionics 21, 1787–1793.
Ahmad, M.A., Puad, N.A.A., Bello, O.S., 2014. Kinetic, equilibrium and thermo- Gupta, V.K., Sharma, S., Yadav, I.S., Mohan, D., 1998. Utilization of bagasse fly ash
dynamic studies of synthetic dye removal using pomegranate peel activated generated in the sugar industry for the removal and recovery of phenol and p-
carbon prepared by microwave-induced KOH activation. Water Resour. Ind. 6, nitrophenol from wastewater. J. Chem. Technol. Biotechnol. 71, 180–186.
18–35. Gupta, V.K., Singh, P., Rahman, N., 2004. Adsorption behavior of Hg(II), Pb(II), and
Ai, L., Jiang, J., 2010. Fast removal of organic dyes from aqueous solutions by AC/ Cd(II) from aqueous solution on Duolite C-433: a synthetic resin. J. Colloid In-
ferrospinel composite. Desalination 262, 134–140. terface Sci. 275, 398–402.
Ali, I., Gupta, V.K., 2007. Advances in water treatment by adsorption technology. Hameed, B.H., Ahmad, A.A., Aziz, N., 2007. Isotherm, kinetics and thermodynamics
Nat. Protoc. 1, 2661–2667. of acid dye adsorption on activated palm ash. Chem. Eng. J. 133, 195–203.
Annadurai, G., Juang, R.S., Lee, D.J., 2002. Use of cellulost-based wastes for ad- Hameed, B.H., EI-Khaiary, M.I., 2008. Sorption kinetics and isotherms studies of a
sorption of dyes from aqueous solutions. J. Hazard. Mater. B92, 263–274. cationic dye using agricultural waste: broad bean peels. J. Hazard. Mater. 154
Arami, M., Limaee, N.Y., Mahmoodi, N.M., Tabrizi, N.S., 2005. Removal of dyes from (1–3), 639–648.
colored textile wastewater by orange peel adsorbent: equilibrium and kinetics Hameed, B.H., Hakimi, H., 2008. Utilization of durian (Duzio zibethinus Murray)
studies. J. Colloid Interface Sci. 288, 371–376. peel as low cost sorbent for the removal of acid dye from aqueous solutions.
Asgher, M., 2012. Biosorpiton of reactive dyes: a review. Water Air Soil Pollut. 223, Biochem. Eng. J. 39 (2), 338–343.
2417–2435. Ho, Y.S., McKay, G., 1999. Pseudo-second order model for sorption processes. Pro-
Asuha, S., Zhou, X.G., Zhao, S., 2010. Adsorption of methyl orange and Cr(VI) on cess Biochem. 34 (5), 451–465.
mesoporous TiO2 prepared by hydrothermal method. J. Hazard. Mater. 181, Huang, J.H., Huang, K.L., Liu, S.Q., Wang, A.T., Yan, C., 2008. Adsorption of Rhoda-
204–210. mine B and methyl orange on a hypercrosslinked polymeric adsorbent in
M.V. Subbaiah, D.-S. Kim / Ecotoxicology and Environmental Safety 128 (2016) 109–117 117

aqueous solution. Colloids Surf. A Physicochem. Eng. Asp. 330, 55–61. adsorption of methyl orange. J. Colloid Interface Sci. 410, 52–58.
Jalil, A.A., Triwahyono, S., Adam, S.H., Diana Rahim, N., Aziz, M.A.A., Hairom, N.H.H., Ozcan, A.S., Ozcan, A., 2004. Adsorption of acid dyes from aqueous solutions onto
Razali, N.A.M., Abidin, M.A.Z., Mohamadiah, M.K.A., 2010. Adsorption of methyl acid activated bentonite. J. Colloid Interface Sci. 276, 39–46.
orange from aqueous solution onto calcined Lapindo volcanic mud. J. Hazard. Panswad, T., Wongchaisuwan, S., 1986. Mechanisms of dye wastewater colour re-
Mater. 181, 755–762. moval by magnesium carbonate-hydrated basic. Water Sci. Technol. 18,
Jiang, R., Fu, Y.Q., Zhu, H.Y., Yao, J., Xiao, L., 2012. Removal of methyl orange from 139–144.
aqueous solutions by magnetic maghemite/chitosan nanocomposite films: Reddy, D.H.K., Lee, S.M., 2013. Application of magnetic chitosan composites for the
adsorption kinetics and equilibrium. J. Appl. Polym. Sci. 125, E540–E549. removal toxic metal and dyes from aqueous solutions. Adv. Colloid Interface Sci.
Khani, H., Rofouei, M.K., Arab, P., Gupta, V.K., Vafaei, Z., 2010. Multi-walled carbon 201-202, 68–93.
nanotubes-ionic liquid-carbon paste electrode as a super selectivity sensor: Reddy, D.H.K., Seshaiah, K., Reddy, A.V.R., Lee, S.M., 2012. Optimization of Cd(II), Cu
application to potentiometric monitoring of mercury ion(II). J. Hazard. Mater. (II) and Ni(II) biosorption by chemically modified Moringa oleifera leaves
183, 402–409. powder. Carbohydr. Polym. 88, 1077–1086.
Khattri, S.D., Singh, M.K., 2000. Color removal from synthetic dye wastewater using Sajab, M.S., Chia, C.H., Zakaria, S., Khiew, P.S., 2013. Cationic and anionic mod-
a biosorbent. Water Air Soil Pollut. 120 (3–4), 283–294. ifications of oil palm empty fruit bunch fibers for the removal of dyes from
Koch, M., Yediler, A., Lienert, D., Insel, G., Kettrup, A., 2002. Ozonation of hydrolyzed aqueous solutions. Bioresour. Technol. 128, 571–577.
azo dye reactive yellow 84 (CI). Chemosphere 46, 109–113. Saleh, T.A., 2015. Isotherms, kinetic, and thermodynamic studies on Hg(II) ad-
Kumar, P.S., Sivaranjanee, R., Vinothini, U., Raghavi, M., Rajasekar, K., Ramakrishnan, sorption from aqueous solution by silica-multiwall carbon nanotubes. Environ.
K., 2014. Adsorption of dye onto raw and surface modified tamarind seeds: Sci. Pollut. Res. Int. 22, 16721–16731.
isotherms, process design, kinetics and mechanism. Desalin. Water Treat. 52, Saleh, T.A., Agarwal, S., Gupta, V.K., 2011. Synthesis of MWCNT/MnO2 and their
2620–2633. application for simultaneous oxidation of arsenite and sorption of arsenate.
Lagergren, S., 1898. About the theory of so-called adsorption of soluble substances. Appl. Catal. B Environ. 106, 46–53.
K. Sven. Vetenskapsakad. Handl. 24, 1–39. Saleh, T.A., Al-Saadi, A.A., Gupta, V.K., 2014. Carbonaceous adsorbent prepared from
Langmuir, I., 1918. The adsorption of gases on plane surfaces of glass, mica and waste tires: experimental and computational evaluations of organic dye methyl
platinum. J. Am. Chem. Soc. 40, 1361–1403. orange. J. Mol. Liq. 191, 85–91.
Liu, R., Fu, K., Zhang, B., Mei, D., Zhang, H., Liu, J., 2012. Removal of methyl orange by Saleh, T.A., Gupta, V.K., 2012. Photo-catalyzed degradation of hazardous dye methyl
modified halloysite nanotubes. J. Disper. Sci. Technol. 33, 711–718. orange by use of a composite catalyst consisting of multi-walled carbon na-
Liu, Z., Zhou, A., Wang, G., Zhao, X., 2009. Adsorption behavior of methyl orange notubes and titanium dioxide. J. Colloid Interface Sci. 371, 101–106.
onto modified ultrafine coal powder. Chin. J. Chem. Eng. 17 (6), 942–948. Senturk, H.B., Ozdes, D., Duran, C., 2010. Biosorption of Rhodamine 6G from aqu-
Ma, Q., Shen, F., Lu, X., Bao, W., Ma, H., 2013. Studies on the adsorption behavior of eous solutions onto almond shell (Prunus dulcis) as a low cost biosorbent. De-
methyl orange from dye wastewater onto activated clay. Desalin. Water Treat. salination 252, 81–87.
51, 3700–3709. Sips, R., 1948. On the structure of a catalyst surface. J. Chem. Phys. 16 (5), 490–495.
Mahmoodian, H., Moradi, O., Shariatzadeha, B., Salehf, T.A., Tyagi, I., Maity, A., Asif, Su., Y., Jiao, Y., Dou, C., Han, R., 2014. Biosorption of methyl orange from aqueous
M., Gupta, V.K., 2015. Enhanced removal of methyl orange from aqueous so- solutions using cationic surfactant modified wheat straw in batch mode. De-
lutions by poly HEMA-chitosan-MWCNT nano-composite. J. Mol. Liq. 202, salin. Water Treat. 52, 6145–6155.
189–198. Tanyildizi, M.S., 2011. Modeling of adsorption isotherms and kinetics of reactive dye
Malik, P.K., Saha, S.K., 2003. Oxidation of direct dyes with hydrogen peroxide using from aqueous solution by peanut hull. Chem. Eng. J. 168, 1234–1240.
ferrous ion as catalyst. Sep. Purif. Technol. 31, 241–250. Tian, Y., Liu, Y., Sun, Z., Li, H., Cui, G., Yan, S., 2015. Fibrous porous silica micro-
McKay, G., Porter, J.F., Prasad, G.R., 1999. The removal of basic dyes aqueous solu- spheres decorated with Mn3O4 for effective removal of methyl orange from
tion by adsorption on low-cost materials. Water Air Soil Pollut. 114 (3–4), aqueous solution. RSC Adv. 5, 106068–106076.
423–438. Toth, J., 1971. State equations of the solid gas interface layer. Acta Chem. Acad.
Mittal, A., Krishnan, L., Gupta, V.K., 2005. Removal and recovery of malachite green Hung. 63, 311–317.
from wastewater using an agricultural waste material, de-oiled soys. Sep. Purif. Yao, Y., He, B., Xu, F., Chen, X., 2011. Equilibrium and kinetic studies of methyl or-
Technol. 43 (2), 125–133. ange adsorption on multiwalled carbon nanotubes. Chem. Eng. J. 170, 82–89.
Mittal, A., Malviya, A., Kaur, D., Mittal, J., Kurup, L., 2007. Studies on the adsorption Zheng, L., Wang, C., Shu, Y., Yan, X., Li, L., 2015. Utilization of diatomite/chitosan-Fe
kinetics and isotherms for the removal and recovery of Methyl Orange from (III) composite for the removal of anionic azo dyes from wastewater: equili-
wastewaters using waste materials. J. Hazard. Mater. 148, 229–240. brium, kinetics and thermodynamics. Colloids Surf. A Physicochem. Eng. Asp.
Netpradit, S., Thiravetyan, P., Towprayoon, S., 2004. Adsorption of three azo reactive 468, 129–139.
dyes by metal hydroxide sludge: effect of temperature, pH, and electrolytes. J. Zhu, H.Y., Jiang, R., Xiao, L., Zeng, G.M., 2010. Preparation, characterization, ad-
Colloid Interface Sci. 270 (2), 255–261. sorption kinetics and thermodynamics of novel magnetic chitosan enwrapping
Obeid, L., Bee, A., Talbot, D., Jaafar, S.B., Dupuis, V., Abramson, S., Cabuil, V., nanosized Ƴ-Fe2O3 and multi-walled carbon nanotubes with enhanced ad-
Welschbillig, W., 2013. Chitosan/maghemite composite: A magsorbent for the sorption properties for methyl orange. Bioresour. Technol. 101, 5063–5069.

S-ar putea să vă placă și