You are on page 1of 12

Chemical Engineering Journal 209 (2012) 577–588

Contents lists available at SciVerse ScienceDirect

Chemical Engineering Journal


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

Application of Mangifera indica (mango) seeds as a biosorbent for removal of


Victazol Orange 3R dye from aqueous solution and study of the biosorption
mechanism
Wagner S. Alencar a, Elie Acayanka b, Eder C. Lima b,⇑, Betina Royer b, Felipe E. de Souza b,
Jerônimo Lameira a, Cláudio N. Alves a
a
Institute of Exact and Natural Sciences, Federal University of Pará (UFPA), R. Augusto Corrêa S/N, 66075-900 Belém, PA, Brazil
b
Institute of Chemistry, Federal University of Rio Grande do Sul (UFRGS), Av. Bento Gonçalves 9500, Postal Box 15003, 91501-970 Porto Alegre, RS, Brazil

h i g h l i g h t s g r a p h i c a l a b s t r a c t

" Mango seed used in its natural and


protonated form as biosorbents.
" Characterization of biosorbents by
FTIR, SEM, BET and BJH techniques.
" A new general order kinetic equation
developed to fit the experimental
results.
" Mechanism of adsorption evaluated
using computational simulation.
" Water plays important role in the
interaction between dye and
biosorbent.

a r t i c l e i n f o a b s t r a c t

Article history: Mango seed, an abundant residue of the food industry, was used in its natural form (MS) and protonated
Received 3 June 2012 form (AMS) as a biosorbent for the removal of Victazol Orange 3R (VO-3R) dye from its aqueous solutions.
Received in revised form 19 August 2012 These biosorbents were characterized by infrared spectroscopy (FTIR), scanning electron microscopy
Accepted 20 August 2012
(SEM), and by nitrogen adsorption/desorption curves. Optimization of the effects of the initial pH of
Available online 27 August 2012
the dye solution, biosorbent dosage and contact time between the dye and the biosorbents on the
biosorption capacities of the biosorbents was studied. Based on an error function (Ferror), the general order
Keywords:
kinetic model provided the best fit to the experimental data when compared to the pseudo-first order and
General order kinetic equation
Mango seed
pseudo-second order kinetic biosorption models. The equilibrium data were fitted to Langmuir, Freund-
Biosorption lich and Liu isotherm models. For both biosorbents, the equilibrium data were best fitted to the Liu
Nonlinear isotherm fitting isotherm model. Finally, the mechanism of biosorption involving VO-3R and lignin–cellulose was evalu-
Thermodynamics ated using the hybrid quantum mechanical/molecular mechanical (QM/MM) approach and molecular
Computational simulation dynamic (MD) simulations. The results suggest that water molecules play a key role in the biosorption
process.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction worldwide [1]. Approximately 10–60% of the reactive dyes are lost
during the manufacturing process, producing large quantities of
Many industries use dyes to color their final products. It has colored wastewater [2]. The dye-containing wastewater dis-
been estimated that about 10,000 different synthetic dyes and charged from industry can adversely affect the aquatic environ-
pigments exist and that over 7  105 tonnes are produced annually ment by impeding light penetration and, as a consequence,
precluding photosynthesis of aquatic flora [3,4]. Moreover, most
⇑ Corresponding author. Tel.: +55 51 3308 7175; fax: +55 51 3308 7304. of the dyes can cause allergy, dermatitis, skin irritation and can
E-mail addresses: eder.lima@ufrgs.br, profederlima@gmail.com (E.C. Lima). also provoke cancer and cell mutation in humans [5,6]. Therefore,

1385-8947/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cej.2012.08.053
578 W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588

effluents containing dyes require treatment before being released 617.54 g mol1, kmax = 489 nm, see Supplementary Fig. 1) at 50%
into the environment [1–4]. purity, was furnished by Dynasty Colourants Co., Ltd. (China) (see
It is rather difficult to treat dye effluent because of its complex Supplementary Fig. 1). The dye was used without further purifica-
aromatic molecular structure that can impart a bright and strong tion. VO-3R has one sulphatoethylsulphone group and one sulfo-
color to other materials. However, the molecular structure of dyes nate group. These groups present negative charges even in highly
make them more stable and non-biodegradable [7–9]. One of the acidic solutions due to their pKa values being lower than zero
methods most often employed for the removal of synthetic dyes [25]. The stock solution was prepared by dissolving the dye in dis-
from aqueous effluents is the adsorption procedure [10,11], due tilled water to a concentration of 5.00 g L1. Working solutions
to its simplicity and high efficiency, as well as the availability of were obtained by diluting the dye stock solutions to the required
a wide range of adsorbents [8–11]. This process transfers the dyes concentrations. To adjust the pH solutions, 0.50 mol L1 sodium
from the aqueous effluent to a solid phase, significantly decreasing hydroxide and hydrochloric acid solutions were used. The pH of
bioavailability of the dye to living organisms [8–14]. The decon- the solutions was measured using a Schott Lab 850 set pH meter.
taminated effluent can then be released to the environment, or
the water can be reutilized in the industrial process [15]. Subse-
quently, the adsorbent can be regenerated or stored in a dry place 2.2. Biosorbent preparation and characterization
with no direct contact with the environment [15].
Activated carbon is one of the most employed adsorbents for The mango seeds (MS) utilized in this work belong to the family
dye removal from aqueous solution because of its excellent adsorp- Mangifera indica L., Tommy Atkins variety. MS was provided by a
tion properties [11,12]. However, the extensive use of activated juice producer in Ubá-MG, Brazil. MS is mainly composed of cellu-
carbon for dye removal from industrial effluents is expensive lose, lignin and hemicelluloses [24,26]. MS was washed with tap
[16,17], due to its high initial and regeneration costs [10], thus water to remove dust, and then with de-ionized water. It was then
limiting more extensive application in wastewater treatment. dried at 70 °C in an air-supplied oven for 8 h [14]. Following this,
There is therefore a growing interest in finding alternative low cost the MS was ground in a disk-mill and subsequently sieved. That
adsorbents for removal of dyes from aqueous solution. Among part of the biosorbent which presented particles of diame-
these alternative adsorbents can be cited: aqai stalk [14]; jackfruit ter 6 106 lm was used. This mango seed was assigned as MS.
leaf powder [15]; cupuassu shell [16]; Brazilian pine-fruit shell In order to increase the amount of VO-3R dye adsorbed by the
[17]; pineapple leaf [18]; rice husk [19]; babassu coconut [20]; MS biosorbent, the biomaterial was protonated as described below
macauba palm [21]; and mango seeds [22,23]. [14]. An amount of 5.0 g of MS was added to 200.0 mL of 3 mol L1
Mango seed is an abundant residue discarded by the mango of HCl, and the slurry was magnetically stirred for 24 h at 70 °C.
juice industry, and is increasing due to the expansion of fruit pro- Subsequently, the slurry was filtered in a sintered glass funnel,
duction [24]. Brazil produces 1.3 million metric tonnes of mangoes and its solid phase was thoroughly washed with water, until the
per year, making it the third largest producer of mangoes in the filtrate reached the pH of distilled water. Subsequently, the mate-
world, most of which are as juice. The seeds correspond to rial was dried at 70 °C in an air-supplied oven for 8 h, yielding the
30–45% of each mango’s weight, depending on the variety, and re- acidified mango seed (AMS).
main as a residue which is usually burned or discarded [24]. The The MS and AMS biosorbents were characterized by vibrational
disposal of large amounts of mango seed (MS) directly into the soil spectroscopy in the infrared region with Fourier Transform (FTIR)
can contaminate the environment in an uncontrolled way, because using a Varian spectrometer (Australia, model 640-IR). The spectra
decomposition of this waste material generates various chemical were obtained with a resolution of 4 cm1 using 100 cumulative
compounds and microorganisms. In this context, using MS as a bio- scans.
sorbent for the removal of dyes from industrial effluents imparts a The surface analyses and porosity were carried out with a volu-
good economic and environmental advantage to developing coun- metric adsorption analyzer, Nova 1000 (Quantachrome Instru-
tries such as Brazil, as it also reduces costs of commercial adsor- ments, United States of America), at 77 K (the boiling point of
bents. Mango seed as a biosorbent for dye removal has been only nitrogen). The samples were pre-treated at 473 K for 24 h under
reported for methylene blue removal [22], and also as a carbon a nitrogen atmosphere in order to eliminate the moisture adsorbed
precursor for the preparation of activated carbon used for the on the surface of the solid sample. The samples were then submit-
adsorption of acid dyes [23]. Mango seeds have not however been ted to conditions of 298 K in a vacuum, reaching the residual pres-
used to biosorb reactive dyes. sure of 104 Pa. For area and pore calculations, the multi-point BET
In this work, use of mango seed in both its natural (MS) and (Brunauer, Emmett and Teller) [27] and BJH (Barret, Joyner and
acidified (AMS) form is being proposed for the first time as poten- Halenda) [28] methods were used.
tial biosorbents for removal of Victazol Orange 3R (VO-3R) reactive The biosorbent samples were also analyzed with scanning elec-
dye from aqueous solutions. This dye is largely used for textile dy- tron microscopy (SEM; Jeol microscope, model JSM 6060) using an
ing in the Brazilian cloth industries. A new kinetic biosorption acceleration voltage of 10 kV and magnification ranging from 200
model was developed for studying the biosorption of the dye by to 5000 [29].
the MS and AMS biosorbents. Additionally, a theoretical investiga- The point of zero charge (pHpzc) of the biosorbent was deter-
tion of the molecular mechanism of biosorption was performed mined by adding 20.00 mL of 0.050 mol L1 NaCl with a previously
through a hybrid quantum mechanical/molecular mechanical adjusted initial pH (the initial pH (pHi) values of the solutions were
(QM/MM) approach and molecular dynamic (MD) simulations. adjusted from 2.0 to 10.0 by adding 0.10 mol L1 of HCl and NaOH)
to several 50.0 mL cylindrical high-density polystyrene flasks
2. Materials and methods (height 117 mm and diameter 30 mm) containing 50.0 mg of the
biosorbent, which were immediately securely capped. The suspen-
2.1. Solutions and reagents sions were shaken in an acclimatized shaker at 298 K and allowed
to equilibrate for 48 h. The suspensions were then centrifuged at
De-ionized water was used throughout the experiments for 14,000 rpm for 10 min to separate the biosorbent from the aqueous
solution preparations. solution. The pHi of the solutions were accurately measured using
The textile dye Victazol Orange 3R, also known as Reactive Or- the solutions that had no contact with the solid biosorbent; the
ange 16 (VO-3R; C.I. 17757; CAS 20262-58-2; C20H17N3O11S3Na2, final pH (pHf) values of the supernatant after contact with the solid
W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588 579

were recorded. The value of pHpzc is the point where the curve of of the software Microcal Origin 7.0. In addition, the models were
DpH (pHf  pHi) versus pHi crosses a line equal to zero [30]. evaluated by a determination coefficient (R2), an adjusted determi-
nation coefficient (R2adj ), as well as by an error function (Ferror) [10],
2.3. Batch biosorption studies which measured the differences in the amount of dye taken up by
the biosorbent as predicted by the models and the actual q mea-
The batch biosorption studies for evaluation of the ability of MS sured experimentally. R2, R2adj and Ferror are given below, in Eqs.
and AMS biosorbents to remove VO-3R dye from aqueous solutions (3)–(5) respectively:
were carried out in triplicate, using the batch contact biosorption Pn   P   !
i;exp 2  ni qi;exp  qi;model 2
qi;exp  q
method. For these experiments, 50.0 mg of biosorbent were placed 2
R ¼ i
ð3Þ
Pn  2
in 50 mL cylindrical polypropylene flasks (117 mm height and 
i qi;exp  qi;exp
30 mm diameter) containing 20.0 mL of dye solution (20.00–
300.0 mg L1), which were agitated for an appropriate time   n  1
(0.0833–24.00 h) using an acclimatized shaker at temperatures R2adj ¼ 1  1  R2  ð4Þ
np
ranging from 298 to 323 K. The pH of the dye solutions ranged
from 2.0 to 10.0. Subsequently, in order to separate the biosorbent
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
u  X n
u 1  2
from the aqueous solutions, the flasks were centrifuged at F error ¼t  qi;exp  qi;model ð5Þ
14,000 rpm for 5 min using a Unicen M Herolab centrifuge (Stutt- np i
gart, Germany), and aliquots of 1–10 mL of the supernatant were
properly diluted with an aqueous solution fixed at pH 2.0. where qi,model is each value of q predicted by the fitted model, qi,exp is
The final concentrations of the dyes remaining in the solution each value of q measured experimentally, q exp is the average of q
were determined by visible spectrophotometry using a T90 + UV– experimentally measured, n is the number of experiments per-
VIS spectrophotometer (PG Instruments, London, England) fitted formed, and p is the number of parameters of the fitted model [10].
with quartz optical cells. Absorbance measurements were made
at 489 nm: the maximum wavelength of VO-3R dye. 2.5. Kinetic biosorption models
The amount of dye taken up and the percentage of the dye re-
moved by the biosorbents were calculated by applying Eqs. (1) See Supplementary material.
and (2), respectively:
2.6. Equilibrium models
ðC o  C f Þ
q¼ V ð1Þ
m See Supplementary material.

ðC o  C f Þ
%Remov al ¼ 100  ð2Þ 2.7. Computational chemistry
Co
in which q is the amount of dye adsorbed by the biosorbent In this study, we undertook a theoretical investigation of possi-
(mg g1), Co is the initial dye concentration placed in contact with ble interaction mechanisms between VO-3R and lignin–cellulose
the biosorbent (mg L1), Cf is the dye concentration (mg L1) after (LC) using the hybrid quantum mechanical/molecular mechanical
the batch biosorption procedure, m is the mass of biosorbent (g) (QM/MM) approach and molecular dynamic (MD) simulations.
and V is the volume of dye solution (L). The initial structure of VO-3R was obtained from pubchem [31],
while the LC structure was constructed using Molden [32]. In order
2.4. Quality assurance and statistical evaluation of the kinetic and to study the protonation state of LC based on the experimental
isotherm parameters study of pH (ranging from 2.0 to 10.0), we considered two models:
(A) the deprotonated LC; and (B) the LC protonating the guaiacyl
To establish the accuracy, reliability and reproducibility of the and siringyl groups. The molecules were pre-optimized in vacuum
collected data, all the batch biosorption measurements were per- and were then placed in a cubic box of pre-equilibrated water
formed in triplicate and the relative standard deviation of all mea- (20 Å side). Finally, 300 ps of hybrid QM/MM Langevin–Verlet
surements was lower than 5% [16]. Blank tests were run in parallel molecular dynamics (NVT) at 300 K were used to equilibrate the
and were corrected when necessary. models. During MD, atoms of VO-3R and LC were selected to be
All dye solutions were stored in glass flasks, which were cleaned treated by QM, using a semi-empirical AM1 Hamiltonian [33,34],
by soaking in 1.4 mol L1 HNO3 for 24 h [16], rinsing five times while the water molecules were described using a TIP3P force field
with de-ionized water, drying and storing them in a flow hood. [35,36] implemented in f-Dynamo. Lastly, the molecular electro-
For analytical calibration, standard solutions with concentra- static potential (MEP) surface was generated from the B3LYP/6-
tions ranging from 5.00 to 150.0 mg L1 of the dyes were used, in 31G(d,p) using the PC Spartan PRO molecular package [37]. This
parallel with a blank solution of water adjusted to pH 2.0. The surface corresponds to an isodensity value of 0.002 au, and it is
linear analytical calibration of the curve was furnished by the UV- interesting to highlight potential non-covalent interactions that
Win software of the T90 + PG Instruments spectrophotometer. The can occur at the molecular level.
detection limit of the method, obtained with a signal/noise ratio of
3 [16], was 0.12 mg L1 of VO-3R. All the analytical measurements 3. Results and discussion
were performed in triplicate, and the precision of the standards
was better than 3% (n = 3). In order to verify the accuracy of the 3.1. Characterization of biosorbents
VO-3R dye sample solutions during spectrophotometric measure-
ments, standards containing dyes at 50.0 mg L1 were employed FTIR technique was used to examine the surface groups of MS
as a quality control every five determinations [16]. and AMS biosorbents and to identify the groups responsible for
The kinetic and equilibrium models were fitted by employing a biosorption of the dye. Infrared spectra of the biosorbents were re-
nonlinear method, with successive interactions calculated by the corded in the range 4000–400 cm1. Supplementary Fig. 2 presents
Levenberg–Marquardt method; interactions were also calculated the FTIR spectra of MS and AMS biosorbents before biosorption
using the Simplex method, based on the nonlinear fitting facilities (MS: Supplementary Fig. 2A; AMS: Supplementary Fig. 2C) and
580 W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588

when loaded with the VO-3R dye (AS + VO-3R: Supplementary and 11.5 m2 g1; average pore diameter (BJH), 4.31 and 5.78 nm;
Fig. 2B; AMS + VO-3R: Supplementary Fig. 2D). It was observed and total pore volume, 0.00918 and 0.0121 cm3 g1, for MS and
that the vibrational bands for AS, AS + VO-3R, AMS and AMS + AMS, respectively. These textural properties obtained for MS and
VO-3R appeared at practically the same wavenumber (differ- AMS are consistent with values of lignin–cellulosic materials re-
ences 6 4 cm1, within the error of the equipment) for the four ported in the literature [16,39,40]. The maximum diagonal length
situations described above. As previously observed for lignin– of VO-3R is 1.68 nm (see Supplementary Fig. 1; the dimensions
cellulosic materials [14], protonation of the biomass did not signif- of the chemical molecule were calculated using ChemBio 3D Ultra
icantly alter the FTIR bands of the biomaterials. It was additionally version 11.0.). The ratios of average pore diameter of the biosor-
observed that the vibrational bands did not alter significantly bents to the maximum diagonal length of dye were 2.56 (£MS/
when compared before and after the biosorption process, indicat- £VO-3R) and 3.44 (£AMS/£VO-3R). Therefore, the mesopores
ing that the strength of the interaction between the dye and the of MS could accommodate up to two molecules of VO-3R dye,
biosorbent is a physical interaction and not a chemical bond. and the mesopores of AMS could accommodate up to three mole-
The intense absorption bands at 3384 and 3385 cm1 were as- cules of VO-3R dye.
signed to OAH bond stretching for AS and AMS, respectively SEM images of the MS and AMS biosorbents are shown in Fig. 1,
[14,16,38]. The CH2 stretching band observed at 2919 cm1 was as- revealing that MS biosorbent is a more compact fibrous material
signed to asymmetric stretching of CH2 groups [38]. Bands at 1722 (see Fig. 1A and B), with only a low number of macropores (pore
and 1718 cm1, for MS and AMS, respectively, were assigned to with £ > 50 nm). AMS biosorbent presents a greater number of
carbonyl groups of carboxylic acid [14,16]. FTIR vibrational bands cavities in the fibrous materials (see Fig. 1C and D) when compared
at 1603 and 1601 cm1 were assigned to stretching of carboxylate to unmodified MS biosorbent.
[38]. Several small bands at 1517, 1457, 1379 and 1325 cm1 for From analysis of the textural properties of MS and AMS biosor-
MS, and 1517, 1457, 1376 and 1379 cm1 for AMS, were assigned bents, it could be inferred that these biosorbent materials predom-
to ring modes of the aromatic rings [38]. The bands at 1160 and inantly contain a mixture of mesopores (pores with diameter
1162 cm1 for MS and AMS, respectively, were assigned to a ranging from 2 to 50 nm, see textural results described above)
CAOAC asymmetric stretch of ether groups of lignin [14,16,38]. and macropores (pore with diameter >50 nm).
The intense FTIR band at 1029 and 1028 cm1 for MS and AMS,
respectively, was assigned to a CAO stretch of primary alcohol of 3.2. Effects of pH of dye solution on biosorption
lignin [38]. Based on these FTIR results, it was expected that the
interaction of VO-3R dye with the MS and AMS biosorbents should One of the most important factors influencing the biosorption of
occur with the OH, C@O, COOH and at the aromatic present in the dye on a biosorbent is the pH of the adsorbate solution [3,12]. Dif-
lignin–cellulosic materials, as previously reported in the literature ferent dyes present different ranges of suitable pH depending on
[14,16,38]. which biosorbent is used. The effects of initial pH on percentage
The textural properties of MS and AMS obtained by nitrogen of removal of VO-3R dye solution (40 mg L1) using MS and AMS
adsorption/desorption curves were: superficial area (SBET), 8.52 biosorbents were evaluated within a pH range between 2 and 10

Fig. 1. SEM of (A) MS with magnification 1000; (B) MS with magnification 1500; (C) AMS with magnification 1000; and (D) AMS with magnification 1500. Accelerating
voltage 10 kV.
W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588 581

Fig. 2. Effect of pH on the biosorption of: VO-3R dye, using (A) MS biosorbent; (B) AMS biosorbent. Conditions: Co = 40.0 mg L1 of dye solution; the temperature was fixed at
298 K; mass of biosorbent 50.0 mg.

(Fig. 2). A considerable difference was observed for the effect of pH 3.4. Kinetic studies
solution on the dye using the biosorbents. For MS biosorbent, the
percentage of dye removed decreased markedly from 48.1% of Nonlinear pseudo-first order, pseudo-second order and general
dye removed at pH 2.0 to less than 0.4% removed at pH 7.0. For order kinetic biosorption models were used to evaluate the kinetics
AMS biosorbent, the percentage of dye removed also decreased of biosorption of VO-3R dye by using the MS and AMS biosorbents,
from 78.5% at pH 2.0 to 13.5% at pH 7.0. as shown in Figs. 3 and 4. The kinetic parameters for the three
The point of zero charge (pHPZC) of MS and AMS biosorbents are kinetic models are listed in Table 1. Taking into account that the
5.45 and 2.15, respectively. For pH values lower than pHPZC, the experimental data were fitted to nonlinear kinetic models, an error
biosorbent presents a positive surface charge [11,30]. The function (Ferror) was used to evaluate the fit of the experimental
dissolved VO-3R dye is negatively charged in water solutions, be- data. The lower the Ferror, the lower the difference in the q calcu-
cause it presents one sulfonate and one sulphatoethylsulphone lated by the model from the experimentally measured q
group [25]. The biosorption of this dye occurs when the biosorbent [3,4,7,8,10–12,14,16,17] (see Eq. (5)). It should be pointed out that
presents a positive surface charge [30]. For MS, electrostatic inter- the Ferror utilized in this work takes into account the number of fit-
action occurs at pH < 5.45, whereas for AMS this interaction occurs ted parameters (p term of Eq. (5)), since it is reported in the liter-
at pH < 2.15. Based on these results, it is clear that the best pH for ature [41,42] that the best fit of the results depends on the number
VO-3R dye biosorption would be pH 2.0 for both biosorbents. of parameters a nonlinear equation presents. For this reason, the
number of fitted parameters should be considered in the calcula-
tion of Ferror.
3.3. Effect of the mass of biosorbent The pseudo-first order kinetic model presented Ferror values of
at least 609% and 592% higher than the values obtained for the gen-
The study of biosorbent masses in the removal of VO-3R dye eral order biosorption model, using MS and AMS biosorbents,
from aqueous solution was carried out using MS and AMS masses respectively. Additionally, the pseudo-second order kinetic model
ranging from 20.0 to 200.0 mg, with the initial dye concentration presented Ferror values that were at least 108% and 92% higher than
fixed at 50.0 mg L1. The highest amount of dye removal was at- the values obtained for the general order biosorption model, using
tained for biosorbent masses of at least 50 mg, corresponding to MS and AMS biosorbents, respectively. These results clearly indi-
biosorbent dosages of 2.5 g L1 (see Supplementary Fig. 3). For cate that the experimental kinetic data were not suitably fitted
biosorbent dosages higher than this value, the percentage of dye by the pseudo-first-order and pseudo-second order kinetic models
removed remained almost constant. Increases in the percentage [43]. On the other hand, the general order kinetic model better
of the dye removed with adsorbent masses up to 50.0 mg explains the biosorption process of VO-3R dye using the MS and
(2.5 g L1) could be attributed to increases in the adsorbent surface AMS biosorbents.
area, augmenting the number of adsorption sites available for Taking into account that the general order kinetic equation pre-
adsorption, as has already been reported in several papers sents different orders (n) when the concentration of the adsorbate
[3,7,8,11,12,30]. Conversely, the increase in the adsorbent masses is changed (see Table 1), it is difficult to compare the kinetic
promoted a remarkable decrease in the amount of dye uptake parameters of the model. Therefore, it is useful to use the initial
per gramme of adsorbent (q), (Supplementary Fig. 3). This could sorption rate h0 [44] to evaluate the kinetics of a given model,
be due to two factors. Firstly, the increase in adsorbent mass at a using the following equation:
fixed dye concentration and volume could lead to unsaturation of
adsorption sites through the adsorption process; and, secondly, h0 ¼ kn  qne ð6Þ
the reduction in adsorbent capacity may have been due to particle
aggregation, resulting from a high adsorbent mass. Such an aggre- in which, h0 is the initial sorption rate (mg g1 h1), kn is the rate
gation would lead to a decrease in the total surface area of the constant [h1 (g mg1)n  1] [45], qe is the amount adsorbed at the
adsorbent and an increase in diffusional path length [28]. There- equilibrium (mg g1), and n is the order of the kinetic model. It
fore, the MS and AMS biosorbent masses must be fixed at should be stressed that when n = 2, this equation is the same initial
50.0 mg, these being the biosorbent masses that corresponded to sorption rate as first introduced by Ho and Mckay [44]. It was ob-
the minimum amount of biosorbent that facilitated consistent served that by increasing the initial dye concentration, the initial
dye removal. sorption rate was increased for all kinetic models, as expected,
582 W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588

Fig. 3. Kinetic biosorption curves of VO-3R dye using MS biosorbent. (A) Co 50 mg L1; (B) Co 100.0; (C) Co 50.0 mg L1; (D) Co 100.0 mg L1. Conditions: pH was fixed at 2.0;
the MS biosorbent mass was fixed at 50.0 mg; and the temperature was fixed at 298 K.

indicating that there is coherence with the experimental data [45]. These results imply that the biosorption processes involved more
Taking into account that the kinetic data were better fitted by the than one sorption rate [45]. For both biosorbents, the biosorption
general order kinetic model, meaning that the order of a biosorption process exhibited three stages, which can be attributed to each
process should follow the same logic as a chemical reaction and linear section as shown in Figs. 3C and D and 4C and D. The first
where the order is experimentally measured [45,46], instead of linear section was attributed to the process of the dye diffusing
being previously stipulated by a given model, the more confident to the biosorbent surface [45]; hence, it was the fastest sorption
initial sorption rates (h0) were obtained by the general order kinetic stage. The second section, ascribed to intra-particle diffusion, was
model. a delayed process [45]. The third stage may be regarded as diffu-
It was observed that the kinetics of biosorption of VO-3R dye on sion through smaller pores, followed by the establishment of equi-
AMS biosorbent were faster than those obtained using MS biosor- librium [45]. The kid values for biosorption of VO-3R on AMS
bent. Considering the initial sorption rate (h0) of VO-3R dye taken biosorbent were 48.9% (50.00 mg L1) and 42.2% (100.00 mg L1)
up by MS and AMS biosorbents (see Table 1), as obtained by the higher than the value for kid obtained for the biosorption of VO-
general order kinetic model, it was observed that h on AMS biosor- 3R dye on MS biosorbent, given the same initial dye concentration.
bent was increased by 120–150% when compared with h0 obtained This difference in kid values is related to differences in the textural
for MS biosorbent. It can be concluded that acid treatment of man- properties of the biosorbents after treatment with acid, as already
go seed promoted an increase in the macropore structure on the reported in the literature [14,45], and also in agreement with the
biomaterial (see Fig. 1), as already reported for other biosorbents results discussed above in the characterization of the biosorbents.
[14,45], contributing to a faster diffusion of the VO-3R dye through The increase in the macropore structure of AMS biosorbent should
the pores of the biosorbent, and also allowing higher amounts of facilitate the diffusion of VO-3R dye molecules through the mac-
the VO-3R dye to be adsorbed by the AMS biosorbent. ropores [27,28]. As the adsorbate diffuses through the pores of
The intra-particle diffusion model [45] was also used to verify the biosorbent, the dye could be adsorbed at the internal sites of
the influence of mass transfer resistance on the binding of VO-3R the AMS biomaterial; therefore both a fast kinetics of biosorption
dye to the biosorbents (Table 1 and Figs. 3C and D and 4C and (increasing the intra-particle diffusion) and higher sorption capac-
D). The intra-particle diffusion constant, kid (mg g1 h0.5), can be ity of AMS biosorbent were expected in relation to MS biosorbent.
obtained from the slope of the plot of qt versus the square root of On the other hand, for the MS biosorbent, with a lower number of
the time. These figures give plots of qt versus t1/2, with three linear macropores, the biosorption is limited to the external surface of
sections for the VO-3R dye using the MS and AMS biosorbents. the biosorbent, decreasing the total amount adsorbed [14,45],
W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588 583

Fig. 4. Kinetic biosorption curves of VO-3R dye using AMS biosorbent. (A) Co 50 mg L1; (B) Co 100.0; (C) Co 50.0 mg L1; (D) Co 100.0 mg L1. Conditions: pH was fixed at 2.0;
the AMS biosorbent mass was fixed at 50.0 mg; and the temperature was fixed at 298 K.

Table 1
Kinetic parameters for VO-3R removal using MS and AMS biosorbent. Conditions: temperature of 298 K; pH 2.0; mass of biosorbent 50.0 mg.

MS AMS
Pseudo-first order 50.00 mg L1 100.00 mg L1 50.00 mg L1 100.00 mg L1
k1 (h1) 2.57 2.57 2.55 2.71
qe (mg g1) 9.46 18.4 16.5 30.8
h0 (mg g1 h1) 24.3 47.2 42.0 83.2
R2adj 0.9874 0.9900 0.9693 0.9648
Ferror 0.3218 0.5594 0.8737 1.712
Pseudo-second order
k2 (g mg1 h1) 0.386 0.199 0.215 0.125
qe (mg g1) 9.89 19.2 17.3 32.1
h0 (mg g1 h1) 37.7 73.2 63.9 129.3
R2adj 0.9993 0.9985 0.9976 0.9970
Ferror 0.0767 0.2167 0.2420 0.4977
General order
kN [h1 (g mg1)n1] 0.586 0.452 0.0628 0.0181
qe (mg g1) 9.75 18.8 17.9 33.7
n 1.79 1.69 2.46 2.58
h0 (mg g1 h1) 34.7 64.5 75.2 160.1
R2adj 0.9998 0.9998 0.9994 0.9998
Ferror 0.0369 0.0789 0.1262 0.1424
Intra-particle
kid (mg g1 h0.5)a 1.39 2.63 2.07 3.74
R2 0.9879 0.9854 0.9835 0.9868
a
Second stage.
584 W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588

Table 2
Isotherm parameters for VO-3R biosorption, using MS and AMS biosorbents. Conditions: biosorbent mass 50.0 mg; pH fixed at 2.0, contact time 6 h.

MS AMS
298 K 303 K 308 K 313 K 318 K 323 K 298 K 303 K 308 K 313 K 318 K 323 K
Langmuir
Qmax (mg g1) 44.8 47.4 51.7 45.9 47.9 51.2 71.6 64.9 62.2 64.9 66.9 68.7
KL (L mg1) 0.0131 0.0135 0.0134 0.0178 0.0184 0.0208 0.0499 0.0626 0.0849 0.0901 0.100 0.116
R2adj 0.9872 0.9797 0.9593 0.9823 0.9776 0.9241 0.9067 0.9761 0.9848 0.9766 0.9755 0.9892
Ferror 1.121 1.515 2.398 1.441 1.881 3.595 5.359 2.697 2.043 2.970 3.2420 2.121
Freudlich
KF (mg g1 (mg L1)1/nF) 2.05 2.27 2.39 3.21 3.10 4.43 10.8 12.0 18.0 17.3 18.7 19.7
nF 1.86 1.88 1.86 2.11 2.04 2.28 2.58 2.87 4.04 3.76 3.90 3.81
R2adj 0.9466 0.9354 0.9041 0.9329 0.9213 0.8396 0.9991 0.8923 0.8982 0.8587 0.8601 0.8967
Ferror 2.292 2.701 3.683 2.809 3.529 5.225 7.330 5.725 5.282 7.295 7.745 6.565
Liu
Qmax (mg g1) 33.4 34.1 34.8 35.5 36.3 36.9 53.3 54.6 56.3 58.3 60.6 63.3
Kg (L mg1) 0.0226 0.0245 0.0264 0.0283 0.0307 0.0333 0.0741 0.0830 0.0931 0.105 0.1169 0.130
nL 1.54 1.73 2.12 1.64 1.77 2.69 2.94 1.69 1.53 1.63 1.676 1.44
R2adj 0.9998 0.9999 0.9998 0.9997 0.9998 0.9998 0.9991 0.9996 0.9997 0.9998 0.9995 0.9997
Ferror 0.1176 0.1279 0.1486 0.1766 0.1610 0.1909 0.5200 0.3444 0.2827 0.2545 0.4602 0.3683

and also leading to slower biosorption kinetics of VO-3R dye. The


cavities generated in the AMS biomaterial can be attributed to its
treatment with 3.0 mol L1 HCl solution.
Figs. 3 and 4 reveal that the minimum contact time of VO-3R
dye with the MS and AMS biosorbents needed to reach equilibrium
was about 5 h for both biosorbents. In order to continue this work,
the contact time between the MS and AMS biosorbents with VO-3R
dye was fixed at 6 h, using the MS and AMS biosorbents. The in-
creased contact time utilized in this work guaranteed that equilib-
rium would be attained for the VO-3R dye even at higher adsorbate
concentrations [14].

3.5. Equilibrium studies

A biosorption isotherm describes the relationship between the


amount of adsorbate adsorbed by the biosorbent (qe) and the
adsorbate concentration remaining in solution after the system
attained equilibrium (Ce), maintaining the process at a constant
temperature. The biosorption parameters of the equilibrium mod-
els often provide some insight into the biosorption mechanism,
surface properties and affinity of the biosorbent with the adsor-
bate. We tested the Langmuir, Freundlich and Liu isotherm models
[10].
The isotherms of biosorption were carried out at 298 to 323 K
with VO-3R dye on the two biosorbents (MS and AMS), and were
performed using the optimum experimental conditions previously
described above (see Table 2, and Fig. 5). Fig. 5 shows the biosorp-
tion isotherms of VO-3R dye using MS and AMS biosorbents at
298 K. Based on the Ferror (see Table 2), the Liu model was found
to be the best isotherm model for both biosorbents at all six pf
the temperatures studied. The Liu model showed (Table 2) the low-
est Ferror values, meaning that the q fit by the isotherm model was
close to the q measured experimentally. The Freundlich isotherm
model was not suitably fitted, presenting Ferror values ranging from
1490% to 2640% and from 1310% to 2770% higher than the Ferror
values obtained for the Liu isotherm model, using MS and AMS Fig. 5. Isotherms of biosorption of VO-3R dye at 298 K. (A) MS; (B) AMS. Conditions:
as biosorbents, respectively. The Langmuir isotherm model was pH was fixed at 2.0; the AMS biosorbent mass was fixed at 50.0 mg; time of contact
6.0 h.
also not suitably fitted, presenting values of Ferror ranging from
850% to 1780%, and from 470% to 1070% higher than the Ferror
values obtained for the Liu isotherm model, using MS and AMS bio-
sorbents, respectively. well-defined sites; each site can only hold one adsorbate species;
The Langmuir isotherm model is based on the following princi- all sites are energetically equivalent; and there are no interactions
ples [45]: adsorbates are chemically adsorbed at a fixed number of between the adsorbate species. The Freundlich isotherm model
W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588 585

Table 3
Thermodynamic parameters of the biosorption of VO-3R dye on MS and AMS biosorbents. Conditions: mass of biosorbent 50.0 mg, pH fixed at 2.0; contact time of 6 h.

Temperature (K)
298 303 308 313 318 323
MS
Kg (L mol1) 1.36  104 1.47  104 1.59  104 1.70  104 1.85  104 2.00  104
DG (kJ mol1) 23.58 24.18 24.76 25.36 25.97 26.60
DH° (kJ mol1) 12.3 – – – – –
DS° (J K1 mol1) 120.4 – – – – –
R2 0.9985 – – – – –
AMS
Kg (L mol1) 4.46  104 5.00  104 5.60  104 6.29  104 7.03  104 7.79  104
DG (kJ mol1) 26.52 27.25 28.00 28.75 29.51 30.25
DH° (kJ mol1) 18.0 – – – – –
DS° (J K1 mol1) 149.4 – – – – –
R2 0.9998 – – – – –

assumes that the concentration of adsorbate on the adsorbent sur-


Table 4
face increases with adsorbate concentration. Theoretically, using Comparison of maxima adsorption capacities for Victazol Orange 3R (VO-3R) dye, also
this expression, an infinite amount of adsorption can occur [45]. known as Reactive Orange 16. The values were obtained at the best experimental
The Liu isotherm model is a combination of the Langmuir and Fre- conditions of each work.
undlich isotherm models, and therefore the monolayer of the Lang- Adsorbent Qmax (mg g1) Ref.
muir model is excluded, as is the infinite adsorption of the
Aqai stalk 61.3 [14]
Freundlich model. The Liu model [45] predicts that the active sites Acidified Aqai Stalk 156 [14]
of the adsorbent could not present the same energy; therefore, the Sewage sludge 47.0 [47]
adsorbent may present active sites that are preferentially occupied Chitosan cross-linked (beads) 30 [48]
Chitosan cross-linked (beads) 5.6 [48]
by the adsorbate molecules [45]. However, saturation of the active
Chitosan 30.4 [49]
sites should occur, unlike the Freundlich isotherm model. Humin immobilized on silica 10.2 [50]
The maximum amounts of VO-3R biosorbed were 36.9 and Mango seed (MS) 36.9 This work
63.3 mg g1, respectively, using MS and AMS as the biosorbent, Acidified mango seed (MAS) 63.3 This work
respectively, at 323 K. It should be highlighted that the maximum
amount (Qmax) of the VO-3R dye adsorbed by the AMS biosorbent
was 60–72% higher when compared with the MS biosorbent. These fident. In addition, the magnitude of enthalpy was consistent with
values indicate that MS and AMS are fairly good biosorbents for the a physical sorption for both biosorbents [55]. The type of interac-
removal of this dye from aqueous solutions (see Table 3) [14,47– tion can be classified, to a certain extent, by the magnitude of
50]. Of eight different adsorbents, AMS presents an adsorption enthalpy change. Physical sorption, such as hydrogen bonding, is
capacity higher than seven. On the other hand, of eight different usually lower than 25 kJ mol1 [55]. Enthalpy changes (DH°) indi-
adsorbents, MS presents an adsorption capacity higher than four. cate that the biosorption followed endothermic processes. Nega-
tive values of DG indicate that the VO-3R dye biosorption by MS
3.6. Thermodynamics of biosorption and AMS biosorbents was a spontaneous and favorable process at
all the studied temperatures. The positive values of DS° confirmed
Thermodynamic parameters related to the biosorption process, a high preference of VO-3R molecules for the surface of MS and
i.e., Gibb’s free energy change (DG°, kJ mol1), enthalpy change AMS, and also suggested the possibility of some structural changes
(DH°, kJ mol1) and entropy change (DS°, J mol1 K1) were deter- or readjustments in the dye–carbon biosorption complex [56].
mined by the following equations:
DG ¼ DH  T DS ð7Þ 3.7. Mechanism of biosorption based on computational calculations

DG ¼ RTLnðKÞ ð8Þ
Hybrid QM/MM MD simulations of 300 ps for VO-3R and lig-
The combination of Eqs. (7) and (8) gives: nin–cellulose (LC) were performed. Representative snapshots of
the averaged structures of VO-3R and lignin–cellulose obtained
DS DH 1
LnðKÞ ¼  x ð9Þ from the 300 ps MD simulations are shown in Fig. 6. As can be
R R T observed in this figure, the water molecules play a key role in
where R is the universal gas constant (8.314 J K1 mol1), T is the dye–lignin–cellulose interactions in both the A and B models,
absolute temperature (Kelvin) and K represents the equilibrium bio- where VO-3R dye interacts with LC through chains of H-bonds
sorption constants of the isotherm fits. It has been reported in the formed by water molecules (Fig. 6).
literature that different biosorption equilibrium constants (K) were In model A, the H-bond linkages are formed in two ways: 1, 2, 3
obtained from different isotherm models [10–12,30,41,45,51–54]. and 4, 5, 6, where the distances correspond to 2.93, 3.72, 3.57 and
Thermodynamic parameters of biosorption can be estimated from 1.88, 4.66, 2.22 Å, respectively. In model B, H-bond linkages are
Kg (Liu equilibrium constant), as already reported in the literature also formed in two ways: 1, 2, 3, 4 and 5, 6, 7, where the distances
[10,41]. observed correspond to 1.66, 1.90, 2.92, 2.92 and 1.87, 1.85, 3.16 Å,
The DH° and DS° values can be calculated from the slope and respectively. The H-bond distances obtained from QM/MM MD are
intercept of the linear plot of Ln(K) versus 1/T. in agreement with the literature [57], which suggests that VO-3R
The thermodynamic results are depicted in Table 4. The R2 dye may be interacting with the biosorbent through chains of
values of the linear fit were at least 0.99, indicating that the values water molecules. As the water molecule chain is located between
of enthalpy and entropy calculated for both biosorbents were con- the siringyl group (see Supplementary Fig. 4) of LC and sulfonic
586 W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588

Fig. 6. Representative snapshots of the averaged structures of VO-3R dye and lignin–cellulose obtained from the 300 ps of QM/MM MD simulations.

groups of VO-3R, we can conclude that the process of diffusion of These surfaces correspond to an isodensity value of 0.002 au.
the siringyl group is better than in the guaiacyl group. Curiously, The most nucleophilic regions (negative electronic potential)
Fourier transform infrared spectroscopy (FTIR) spectra for the are shown in red, while the most electrophilic regions (positive
absorbent before and after the contact with VO-3R dye were sim- electrostatic potential) are shown in blue. In Fig. 8 we can
ilar (see Supplementary Fig. 2), suggesting the existence of a phys- observe a much more intense region of negative electrostatic
ical rather than a chemical reaction [14]. potential around the sulfonate group of the VO-3R (Fig. 8A),
The water molecules form H-bonding links between the VO-3R and a more intense region of positive electrostatic potential
dye and lignin–cellulose (Fig. 7), in agreement with the FTIR spec- around the siringyl group of lignin–cellulose material (Fig. 8B
tra of the biosorbent obtained before and after biosorption of the and C). When the lignin–cellulose material was protonated, the
dye, in which no chemical reaction is observed (Supplementary interactions with the siringyl group were more intense, yielding
Fig. 2). Therefore, water molecules play an important role in the a better biosorption.
biosorption process. This new proposed interaction mechanism The computer simulation results reveal that model B better ex-
(Fig. 7) of lignin–cellulose and dye might improve understanding plains the experimental observations than model A. In fact, we can
of the previous mechanism proposed for the biosorption of dyes observe that the siringyl group and the sulfonate group are close to
using lignin–cellulosic biosorbents [14,16,45]. each other, as demonstrated by the average distance between then,
Fig. 8 shows the molecular electrostatic potential (MEP) for which is 8.50 Å and 7.90 Å for model A and B, respectively. This
the VO-3R (Fig. 8A), LC (Fig. 8B) and protonated LC (Fig. 8C). favorable interaction is reflected in Fig. 8 (MEP) and may explain
MEP surfaces were generated using the Spartan program [37]. the experimental results. Thus, model B is in agreement with

Fig. 7. Mechanism of biosorption process.


W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588 587

general order kinetic model. However, the intra-particle diffusion


model gave multiple linear regions, which suggested that biosorp-
tion may also be followed by multiple sorption rates. The
minimum equilibration time of the dye was obtained after 5 h of
contact between VO-3R dye and the MS and AMS biosorbents.
The equilibrium isotherm of these dyes was obtained, which was
best fitted to the Liu isotherm model. The maximum amounts of
VO-3R adsorbed were 51.2 (323 K) and 71.6 (298 K) mg g1, using
MS and AMS as biosorbents, respectively. The thermodynamic
parameters of biosorption (DH°; DS° and DG) were calculated. Fi-
nally, we investigated the interaction mechanism of biosorbent
and VO-3R using QM/MM MD simulation. The results suggest that
water molecules play a key role in interactions between the biosor-
bent and VO-3R, where a chain of H-bonds formed by water mol-
ecules may be a link to the siringyl group (see Supplementary
Fig. 4) of lignin–cellulosic material to sulfonic groups of VO-3R
dye. Herein, we have proposed a new mechanism of biosorption
of VO-3R on the surface of biosorbents, combining experimental
data (FTIR spectra and pH studies) and theoretical calculations,
that might improve our understanding of biosorption processes.

Acknowledgements

We thank to Professor Guimes Rodrigues-Filho from Institute of


Chemistry, Federal University of Uberlandia for donating the man-
go seeds. The authors are grateful to the National Council for Scien-
tific and Technological Development (CNPq, Brazil), and to the
Foundation for Research Support in the State of Rio Grande do
Sul (FAPERGS, Brazil) for financial support and fellowships. We
are also grateful to Center of Electron Microscopy (CME-UFRGS)
for the use of the SEM microscope. We thank to Herolab (Wiesloch,
Fig. 8. Maps of molecular electrostatic potential (MEP) surfaces derived from Germany) for donating the sealing o-ring of the rotor of the Unicem
B3LYP/6-31G(d,p)calculations for (A) VO-3R dye; (B) siringyl group of LC and (C) M centrifuge.
siringyl group of LC protonated. The increase of negative charges goes from the blue
(positive) to red (negative). (For interpretation of the references to color in this
figure legend, the reader is referred to the web version of this article.)
Appendix A. Supplementary material

Supplementary data associated with this article can be found, in


experimental observations suggesting that the acid pH is a critical
the online version, at http://dx.doi.org/10.1016/j.cej.2012.08.053.
factor in the biosorption process.
Finally, the experimental and theoretical results reported in this
work reveal that water molecules have a key role in the biosorp- References
tion process: H-bonds are formed between the VO-3R and
ligno-cellulose (LC) through water molecules which are essential [1] N. Koprivanac, H. Kusic, Hazardous Organic Pollutants in Colored Wastewaters,
New Science Publishers, New York, 2008.
for biosorption of dye on the biosorbent. In accordance with FTIR [2] C. Hessel, C. Allegre, M. Maisseu, F. Charbit, P. Moulin, Guidelines and
spectra, there is no chemical reaction (Supplementary Fig. 2). In legislation for dye house effluents, J. Environ. Manage. 83 (2007) 171–180.
addition, the magnitude of enthalpy of biosorption obtained in this [3] L.G. da Silva, R. Ruggiero, P.M. Gontijo, R.B. Pinto, B. Royer, E.C. Lima, T.H.M.
Fernandes, T. Calvete, Adsorption of Brilliant Red 2BE dye from water solutions
work (see Table 3) is consistent with the energy of a hydrogen by a chemically modified sugarcane bagasse lignin, Chem. Eng. J. 168 (2011)
bond formation [55]. The formation of acidic sites in the substrate 620–628.
have not been reported in previous works [14,16,45]. On the other [4] E.W. de Menezes, E.C. Lima, B. Royer, F.E. de Souza, B.D. dos Santos, J.R.
Gregório, T.M.H. Costa, Y. Gushikem, E.V. Benvenutti, Ionic silica based hybrid
hand, our results suggest that the formation of acidic sites in the material containing the pyridinium group used as adsorbent for textile dye, J.
substrate is crucial to biosorption processes involving water mole- Colloid Interface Sci. 378 (2012) 10–20.
cule chains. [5] D.S. Brookstein, Factors associated with textile pattern dermatitis caused by
contact allergy to dyes, finishes, foams, and preservatives, Dermatol. Clin. 27
(2009) 309–322.
[6] P.A. Carneiro, G.A. Umbuzeiro, D.P. Oliveira, M.V.B. Zanoni, Assessment of
4. Conclusion water contamination caused by a mutagenic textile effluent/dyehouse effluent
bearing disperse dyes, J. Hazard. Mater. 174 (2010) 694–699.
Mango (M. indica L.) seeds in natural form (MS), as well as a pro- [7] B. Royer, N.F. Cardoso, E.C. Lima, V.S.O. Ruiz, T.R. Macedo, C. Airoldi,
Organofunctionalized kenyaite for dye removal from aqueous solution, J.
tonated form (AMS), are good alternative biosorbents for removal Colloid Interface Sci. 336 (2009) 398–405.
of the textile dye Victazol Orange 3R (VO-3R) from aqueous solu- [8] B. Royer, N.F. Cardoso, E.C. Lima, T.R. Macedo, C. Airoldi, A useful
tions. MS and AMS were characterized by FTIR spectroscopy, SEM organofunctionalized layered silicate for textile dye removal, J. Hazard.
Mater. 181 (2010) 366–374.
and nitrogen adsorption/desorption curves. The VO-3R dye
[9] G.L. Dotto, E.C. Lima, L.A.A. Pinto, Biosorption of food dyes onto Spirulina
interacts with the biosorbents at the solid/liquid interface when platensis nanoparticles: equilibrium isotherm and thermodynamic analysis,
suspended in water. The best conditions were established, with re- Bioresour. Technol. 103 (2012) 123–130.
spect to both pH and contact time, to saturate the available sites [10] F.M. Machado, C.P. Bergmann, T.H.M. Fernandes, E.C. Lima, B. Royer, T. Calvete,
S.B. Fagan, Adsorption of Reactive Red M-2BE dye from water solutions by
located on the biosorbent surface. Four kinetic models were used multi-walled carbon nanotubes and activated carbon, J. Hazard. Mater. 192
to adjust the biosorption, and the best fit was obtained by the (2011) 1122–1131.
588 W.S. Alencar et al. / Chemical Engineering Journal 209 (2012) 577–588

[11] T. Calvete, E.C. Lima, N.F. Cardoso, J.C.P. Vaghetti, S.L.P. Dias, F.A. Pavan, [33] M. Elstner, K.J. Jalkanen, M. Knapp-Mohammady, T. Frauenhem, S. Suhai,
Application of carbon adsorbents prepared from Brazilian-pine fruit shell for Energetics and structure of glycine and alanine based model peptides:
the removal of reactive orange 16 from aqueous solution: kinetic, equilibrium, approximate SCC-DFTB, AM1 and PM3 methods in comparison with DFT, HF
and thermodynamic studies, J. Environ. Manage. 91 (2010) 1695–1706. and MP2 calculations, Chem. Phys. 263 (2001) 203–219.
[12] N.F. Cardoso, R.B. Pinto, E.C. Lima, T. Calvete, C.V. Amavisca, B. Royer, M.L. [34] H. Zhou, E. Tajkhorshid, T. Frauenheim, S. Suhai, M. Elstner, Performance of the
Cunha, T.H.M. Fernandes, I.S. Pinto, Removal of remazol black B textile dye AM1, PM3, and SCC-DFTB methods in the study of conjugated Schiff base
from aqueous solution by adsorption, Desalination 269 (2011) 92–103. molecules, Chem. Phys. 277 (2002) 91–103.
[13] S. Chowdhury, P. Saha, Sea shell powder as a new adsorbent to remove Basic [35] M.J. Field, M. Albe, C. Bret, The Dynamo library for molecular simulations using
Green 4 (Malachite Green) from aqueous solutions: equilibrium, kinetic and hybrid quantum mechanical and molecular mechanical potentials, J. Comput.
thermodynamic studies, Chem. Eng. J. 164 (2010) 168–177. Chem. 21 (2000) 1088–1100.
[14] N.F. Cardoso, E.C. Lima, T. Calvete, I.S. Pinto, C.V. Amavisca, T.H.M. Fernandes, [36] W.L. Jorgensen, J. Chandrasekhar, J.D. Madura, R.W. Impey, M.L.J. Klein,
R.B. Pinto, W.S. Alencar, Application of aqai stalks as biosorbents for the Comparison of simple potential functions for simulating liquid water, J.
removal of the dyes Reactive Black 5 and Reactive Orange 16 from aqueous Chem. Phys. 79 (1983) 926–935.
solution, J. Chem. Eng. Data 56 (2011) 1857–1868. [37] Spartan’08, Spartan 08 Graphical User Interface, Irvine, CA, 2009.
[15] P.D. Saha, S. Chakraborty, S. Chowdhury, Batch and continuous (fixed-bed [38] B. Smith, Infrared Spectral Interpretation. A Systematic Approach, CRC Press,
column) biosorption of crystal violet by Artocarpus heterophyllus (Jackfruit) leaf Boca Raton, 1999.
powder, Colloid Surf. B 92 (2012) 262–270. [39] W.S.W. Ngah, M.A.K.M. Hanafiah, Biosorption of copper ions from dilute
[16] N.F. Cardoso, E.C. Lima, I.S. Pinto, C.V. Amavisca, B. Royer, R.B. Pinto, W.S. aqueous solutions on base treated rubber (Hevea brasiliensis) leaves powder:
Alencar, S.F.P. Pereira, Application of cupuassu shell as biosorbent for the kinetics, isotherm, and biosorption mechanisms, J. Environ. Sci. 20 (2008)
removal of textile dyes from aqueous solution, J. Environ. Manage. 92 (2011) 1168–1176.
1237–1247. [40] J.C. Zheng, H.M. Feng, M.H.W. Lam, P.K.S. Lam, Y.W. Ding, H.Q. Yu, Removal of
[17] B. Royer, N.F. Cardoso, E.C. Lima, J.C.P. Vaghetti, N.M. Simon, T. Calvete, R.C. Cu(II) in aqueous media by biosorption using water hyacinth roots as a
Veses, Applications of Brazilian-pine fruit shell in natural and carbonized biosorbent material, J. Hazard. Mater. 171 (2009) 780–785.
forms as adsorbents to removal of methylene blue from aqueous solutions – [41] F.M. Machado, C.P. Bergmann, E.C. Lima, B. Royer, F.E. de Souza, I.M. Jauris, T.
kinetic and equilibrium study, J. Hazard. Mater. 164 (2009) 1213–1222. Calvete, S.B. Fagan, Adsorption of Reactive Blue 4 dye from water solutions by
[18] S. Chowdhury, S. Chakraborty, P. Saha, Biosorption of Basic Green 4 from carbon nanotubes: experiment and theory, Phys. Chem. Chem. Phys. 14 (2012)
aqueous solution by Ananas comosus (pineapple) leaf powder, Colloid Surf. B 11139–11153.
84 (2011) 520–527. [42] M.I. El-Khaiary, G.F. Malash, Y.S. Ho, On the use of linearized pseudo-second-
[19] S. Chakraborty, S. Chowdhury, P.D. Saha, Adsorption of Crystal Violet from order kinetic equations for modeling adsorption systems, Desalination 257
aqueous solution onto NaOH-modified rice husk, Carbohydr. Polym. 84 (2011) (2010) 93–101.
1533–1541. [43] M.I. El-Khaiary, G.F. Malash, Common data analysis errors in batch adsorption
[20] A.P. Vieira, S.A.A. Santana, C.W.B. Bezerra, H.A.S. Silva, J.A.P. Chaves, J.C.P. de studies, Hydrometallurgy 105 (2011) 314–320.
Melo, E.C. da Silva-Filho, C. Airoldi, Kinetics and thermodynamics of textile dye [44] Y.S. Ho, G. Mckay, Sorption of dye from aqueous solution by peat, Chem. Eng. J.
adsorption from aqueous solutions using babassu coconut mesocarp, J. Hazard. 70 (1988) 115–124.
Mater. 166 (2009) 1272–1278. [45] W.S. Alencar, E.C. Lima, B. Royer, B.D. dos Santos, T. Calvete, E.A. da Silva, C.N.
[21] S.S. Vieira, Z.M. Magriotis, N.A.V. Santos, M.D. Cardoso, A.A. Saczk, Macauba Alves, Application of aqai stalks as biosorbents for the removal of the dye
palm (Acrocomia aculeata) cake from biodiesel processing: an efficient and low Procion Blue MX-R from aqueous solution, Sep. Sci. Technol. 47 (2012) 513–
cost substrate for the adsorption of dyes, Chem. Eng. J. 183 (2012) 152–161. 526.
[22] K.V. Kumar, A. Kumaran, Removal of methylene blue by mango seed kernel [46] R. Chang, General Chemistry: The Essential Concepts, sixth ed., McGraw-Hill,
powder, Biochem. Eng. J. 27 (2005) 83–93. New York, 2011.
[23] M.M. Davila-Jimenez, M.P. Elizalde-Gonzalez, V. Hernandez-Montoya, [47] S.W. Won, S.B. Choi, Y.S. Yun, Performance and mechanism in binding of
Performance of mango seed adsorbents in the adsorption of anthraquinone Reactive Orange 16 to various types of sludge, Biochem. Eng. J. 28 (2006) 208–
and azo acid dyes in single and binary aqueous solutions, Bioresour. Technol. 214.
100 (2009) 6199–6206. [48] I.Y. Kimura, M.C.M. Laranjeira, V.T. Fávere, L. Furlan, The interaction between
[24] C.S. Meireles, G. Rodrigues-Filho, M.F. Ferreira Jr., D.A. Cerqueira, R.M.N. reactive dye containing vinylsulfone group and chitosan microspheres, Int. J.
Assunção, E.A.M. Ribeiro, P. Poletto, M. Zeni, Characterization of asymmetric Polym. Mater. 51 (2002) 759–768.
membranes of cellulose acetate from biomass: newspaper and mango seed, [49] G. Crini, P.M. Badot, Application of chitosan, a natural aminopolysaccharide,
Carbohydr. Polym. 80 (2010) 954–961. for dye removal from aqueous solutions by adsorption processes using batch
[25] J.D. Roberts, M.C. Caserio, Basic Principles of Organic Chemistry. W.A. studies: a review of recent literature, Prog. Polym. Sci. 33 (2008) 399–447.
Benjamin Incorporation, second ed., London, 1977. [50] A.M.D. Jesus, L.P.C. Romão, B.R. Araújo, A.S. Costa, J.J. Marques, Use of humin as
[26] M.F. Ferreira Jr., E.A.R. Mundim, G. Rodrigues-Filho, C.S. Meireles, D.A. an alternative material for adsorption/desorption of reactive dyes,
Cerqueira, R.M.N. de Assunção, M. Marcolin, M. Zeni, SEM study of the Desalination 274 (2011) 13–21.
morphology of asymmetric cellulose acetate membranes produced from [51] P. Leechart, W. Nakbanpote, P. Thiravetyan, Application of ‘waste’ wood-
recycled agroindustrial residues: sugarcane bagasse and mango seeds, shaving bottom ash for adsorption of azo reactive dye, J. Environ. Manage. 90
Polym. Bull. 66 (2011) 377–389. (2009) 912–920.
[27] R.A. Jacques, R. Bernardi, M. Caovila, E.C. Lima, F.A. Pavan, J.C.P. Vaghetti, C. [52] M.S. Bilgili, Adsorption of 4-chlorophenol from aqueous solutions by xad-4
Airoldi, Removal of Cu(II), Fe(III) and Cr(III) from aqueous solution by aniline resin: isotherm, kinetic, and thermodynamic analysis, J. Hazard. Mater. 137
grafted silica gel, Sep. Sci. Technol. 42 (2007) 591–609. (2006) 157–164.
[28] J.C.P. Vaghetti, M. Zat, K.R.S. Bentes, L.S. Ferreira, E.V. Benvenutti, E.C. Lima, 4- [53] S. Nethajia, A. Sivasamya, G. Thennarasu, S. Saravanan, Adsorption of
Phenylenediaminepropylsilica xerogel as a sorbent for copper determination Malachite Green dye onto activated carbon derived from Borassus aethiopum
in waters by slurry-sampling ETAAS, J. Anal. Atom. Spectrom. 18 (2003) 376– flower biomass, J. Hazard. Mater. 181 (2010) 271–280.
380. [54] V.K. Gupta, R. Jain, S. Malathi, A. Nayak, Adsorption–desorption studies of
[29] R.C. Jacques, E.C. Lima, S.L.P. Dias, A.C. Mazzocato, F.A. Pavan, Yellow passion- indigocarmine from industrial effluents by using deoiled mustard and its
fruit shell as biosorbent to remove Cr(III) and Pb(II) from aqueous solution, comparison with charcoal, J. Colloid Interface Sci. 348 (2010) 628–633.
Sep. Purif. Technol. 57 (2007) 193–198. [55] C.L. Sun, C.S. Wang, Estimation on the intramolecular hydrogen-bonding
[30] T. Calvete, E.C. Lima, N.F. Cardoso, S.L.P. Dias, F.A. Pavan, Application of carbon energies in proteins and peptides by the analytic potential energy function, J.
adsorbents prepared from the Brazilian-pine fruit shell for removal of Procion Mol. Struct. 956 (2010) 38–43.
Red MX 3B from aqueous solution – kinetic, equilibrium, and thermodynamic [56] D.D. Asouhidou, K.S. Triantafyllidis, N.K. Lazaridis, K.A. Matis, S.S. Kim, T.J.
studies, Chem. Eng. J. 155 (2009) 627–636. Pinnavaia, Sorption of reactive dyes from aqueous solutions by ordered
[31] Pub Chem compound, 20262-58-2 – Compound Summary (CID 5744432). hexagonal and disordered mesoporous carbons, Microporous Mesoporous
<http://pubchem.ncbi.nlm.nih.gov/summary/summary.cgi?cid=5744432&loc= Mater. 117 (2009) 257–267.
ec_rcs#x281> (accessed 01.06.12). [57] E. Meyer, Internal water molecules and H-bonding in biological
[32] G. Schaftenaar, J.H. Noordik, Molden: a pre- and pós-processing program for macromolecules: a review of structural features with functional
molecular and electronic structures, J. Comput.-Aided Mol. Des. 14 (2000) implications, Protein Sci. I (1992) 1543–1562.
123–134.

You might also like