Sunteți pe pagina 1din 10

Article

pubs.acs.org/jced

Dye Adsorption by Leather Waste: Mechanism Diusion, Nature


Studies, and Thermodynamic Data
Jeferson S. Piccin,* Liliana A. Feris, Mrian Cooper, and Mariliz Gutterres
Chemical Engineering Post Graduation Program, Laboratory for Leather and Environmental Studies (LACOURO), Federal
University of Rio Grande do Sul, Luiz Englert str., s/no., 90.040-040, Porto Alegre, RS, Brazil
ABSTRACT: Tannery solid waste (leather) is a possible adsorbent of dye
contaminants in wastewater. In this paper the nature and mechanisms of dyes
adsorption by chromium-tanned leather waste (CTLW) are proposed and
discussed on the basis of isotherms, adsorption kinetics, and thermodynamics
of three dyes: Red 357, Black 210, and Yellow 194 in aqueous solutions.
Langmuir, BrunauerEmmettTeller (BET) and Henry isotherm models
were used to t the adsorption equilibrium data, respectively. The kinetic data
were evaluated using boundary layer mass transfer and intraparticle diusion
models. The boundary layer mass transfer coecient was in the order of 106
and 105 mmin1 for the three studied dyes. However, the intraparticle
diusion were of the order of 108 and 1011 m2min1 for the Red 357 and
Black 210 dyes, respectively, demonstrating that intraparticle diusion is the
predominant mass transfer mechanism of these dyes. The values of H, S,
G, and Ea suggest that adsorption is spontaneous, exothermic, and chemical in nature. The chemical nature is also conrmed by
Fourier transform infrared (FT-IR) analysis.

1. INTRODUCTION
Many industries, such as leather, food, cosmetics, plastics, and
textiles, use dyes to confer, intensify, or restore the color of
their products. The world consumption of dyes is estimated to
be around 700 000 tonnes per year.1,2 In these industries,
during the processes involved in production usually dyes are
dissolved in water. Despite being a contaminant that
signicantly contributes to the elevation of chemical oxygen
demand of wastewaters, at low concentrations they may
signicantly change the color of the water, causing aesthetic
problems in water bodies polluted with industrial euents.
Moreover, the biological treatment systems may not be ecient
for the removal of color, when the objective is the wastewater
reuse, especially in tanneries that perform only the leather
nishing, because in these cases the wastewater has higher dye
concentrations.
Adsorption is one of the most eective methods used for
removal of dyes and other soluble substances from wastewater.
Activated carbon is the most widely used adsorbent in various
industrial sectors. However, especially due to its production and
regeneration cost, the use of adsorption in industrial processes,
as tanneries, is limited. Therefore, the use of industrial waste as
alternative adsorbents has been considered. In this context,
solid waste generated in the leather processing operation has
been used as alternative adsorbents for the removal of dyes,
metals, oils, and surfactants.36 Although a nal cleanup of
water-based euents using an adsorbent like activated carbon is
not common practice on the industrial scale in the leather
industry, the use of a solid waste of the process may be a
recycling alternative to reduce the cost of disposal this waste
and to reuse the treated water. The ongoing development on
2013 American Chemical Society

advanced treatment systems enables wastewater regeneration


requirements for segregated or integrated water recycling or
reuse and the removal of dicult contaminants to meet
environmental regulatory standards for discharge.7
Because of the design of xed bed column adsorption
systems, the adsorption capacity and mass transfer rate of
solution for the adsorbent particle are the most important
parameters.8 The adsorption capacity is described by solid
liquid equilibrium curves and represented by isotherm models.
However, the mass transfer mechanisms for the solution to
solid sorbent can be obtained from the adsorption kinetics
using kinetic models described in the literature. In addition,
isotherms and kinetics provide important thermodynamic data
for the characterization of the nature of adsorption.
In this paper the eect of temperature on the adsorption of
dye, adsorption isotherm, and adsorption kinetics of leather
wastes were investigated. Isotherm models were proposed, and
adsorption mechanisms were evaluated using boundary layer
mass transfer and intraparticle diusion models (HSDM). The
thermodynamic parameters were calculated, and the adsorption
nature was evaluated using Fourier transform infrared (FT-IR)
analysis.

2. EXPERIMENTAL SECTION
2.1. Adsorbate. Three commercial dyes used were supplied
by the Business Leather Unit of Lanxess Company. Information
Received: September 4, 2012
Accepted: February 12, 2013
Published: February 21, 2013
873

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

on the C.I. Acid Red 357 (Red 357, an azo-dissulphonated Crorganic-complex dye, CAS No. 57674-14-3, and purity of 55
%), C.I. Acid Black 310 (Black 210, an amine dissulfonated
triazo organic-dye, CAS No. 99576-15-5, and purity of 70 %),
and C.I. Acid Yellow 194 (Yellow 194, an azo-dissulphonated
Co-organic-complex dye, CAS No. 85959-73-5, and purity of
50 %) were obtained from the American Association of Textile
Chemists and Colorists (AATCC), the United States Environmental Protection Agency (EPA), and previous work.9 Figure 1
shows their optimized chemical structure using ChemBio 3D
11.0.1 software and others characteristics. Solutions of
commercial dyes with approximately 800 mgL1 were
produced for the experiments, and these correspond to an
initial concentration of 375 mgL1, 465 mgL1, and 555
mgL1 of Yellow 194, Red 357, and Black 210 dyes,
respectively, calculated based on the purity of the dye products.
2.2. Adsorbent Preparation. Leather waste samples from
chromium-tanned leather shaving operation were obtained
from a local tannery (Portao/RS, Brazil). The so-called
adsorbent chromium-tanned leather waste (CLTW) was
dried, ground, and sieved according to Piccin et al.9 Table 1
shows the physical-chemical characteristics of the adsorbent
used.
2.3. Sorption Experiments. Equilibrium adsorption and
kinetics studies were carried out by batch conditions at dierent
temperatures [(15 to 45) C] according Piccin et al.9 The dye
concentrations were determined by UVvis spectrophotometry
using standard curves obtained considering the purity degree of
the dyes. All adsorption experiments were performed at pH 2.5
and in duplicate, and data were considered satisfactory with a
coecient of variation of less than 2.5 %. The Giles et al.10
classication was adopted to evaluate the behavior of
equilibrium data, and Langmuir, BrunauerEmmettTeller
(BET), and Henry isotherm models were proposed to correlate
experimental data.9 From kinetic data, lm uid in boundary
layer mass transfer and intraparticle diusion models were
proposed, and the mechanisms of adsorption were checked.
2.4. FT-IR Analysis. The adsorbent samples were dried to
constant weight at 105 C, before and after the adsorption
process. Afterward, the samples were analyzed by infrared
spectroscopy (Perkin-Elmer, Spectrum 1000, USA), in the
range of (4000 to 400) cm1, using transmittance spectrum
with a potassium bromide disc (1 part of adsorbent for 20 parts
of KBr).11,12

3. RESULTS AND DISCUSSION


3.1. Adsorption Isotherms. Figure 2 shows the eect of
temperature on adsorption equilibrium of leather dyes in
CTLW. According Giles et al.,10 it is shown that Red 357 dye
adsorption isotherm is apparently of the H2 type, traditional
shape observed in studies of alternative adsorbents. In this case,
the H2 type isotherm was represented by the Langmuir
model with satisfactory adjustment to the experimental data. It
was the same with adsorption of dyes on chromium containing
leather wastes,3,6 thermo-chemical modied leather wastes,12
chitosan beads,13 and Spirulina platensis.14
Nonconventional equilibrium data are observed for Black
210 and Yellow 194, corresponding to H3 and C1 isotherm
types, respectively. Black 210 adsorption by CTLW are of H3
and L3 types, corresponding to temperatures less than 25 C
and more than 35 C, respectively. However, Yellow 194 dye
adsorption shows C1 isotherm type. In literature, H3 and
L3 types isotherms have been satisfactorily represented by the

Figure 1. Optimized three-dimensional chemical structures of the


leather dyes Red 357 (a), Black 210 (b), and Yellow 194 (C).

type of BET model, and it was the same with pentaclorophenol


adsorption by carbonized pine bark,15 dyes in groundwater by
clay,16 and biosorption for color removal of pulp mill euent
by fungal biomass.17 Moreover, linear isotherms of C1 type
874

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

Table 1. Characterization of Chromium-Tanned Leather


Waste (CTLW)
parameter

parameter

determined
valuea

method

moisture (%)
ashes (%, D.B.)
total carbon (%, D.B.)

7.8 0.8
9.0 0.4
37.1 2.5

total chromium
(%, D.B.)
particle diameter (mm)
density (kgm3)

2.5 0.1

ASTM D3790-79
ASTM D2617-06
instrumental (Shimadzu
SSM-5000A)
ABNT NBR 11054

0.98 0.22
1450.2 37.0

screening
picnometry

Table 2. Adsorption Isotherm Parameters of Tannery Dye


Adsorption by CTLW

KL (Lmg1)
QM (mgg1)
R2
ARE (%)
K1 (Lmg1)
K2 (Lmg1)103
QBET (mgL1)
R2
ARE (%)

Mean standard deviation, n = 3.

KH (Lg1)
R2
ARE (%)

were represented by the Henry model, as observed for


phosphate adsorption by loess modied with zinc,18 the
adsorption of Reactive Black and Reactive Yellow hydrolyzed
dyes by diatomaceous earth19 and the adsorption of Methylene
Blue, Reactive Black, and Reactive Yellow dyes by calcined
diatomite.20 A previous study reported by Piccin et al.9
presented the Langmuir (eq 1), BET (eq 2), and Henry (eq
3) isotherm models to predict the equilibrium data of Red 357,
Black 210, and Yellow 194 dyes adsorption by CTLW,
respectively. Table 2 shows the observed parameters, the
correlation coecient (R2), and average relative errors (AREs)
for the adsorption of tannery dyes by CLTW, according to the
selected models for each one of them.

qe =

qe =

25 C

Red 357
0.218
0.142
218.8
232.0
0.980
0.975
3.7
5.7
Black 210
2.907
0.522
2.10
1.70
43.2
108.8
0.994
0.998
4.857
2.494
Yellow 194
2.172
1.909
0.986
0.985
5.2
3.6

35 C

45 C

0.112
239.9
0.953
5.4

0.090
250.6
0.973
6.0

0.042
1.50
138.1
0.996
3.711

ND
ND
ND
ND
ND

1.100
0.988
4.7

0.993
0.997
1.4

qBETk1Ce
(1 k 2Ce)(1 k 2Ce + k1Ce)

qe = kHCe

(2)
(3)

High coecients of determination (R2 > 0.95) exhibited in


Table 2 indicate that more than 95 % of the variability of qe as a
function of Ce increase can be explained by the proposed
models. Moreover, low AREs demonstrate that the fraction
error of qe distributed across the entire concentration range is
lower than 10 %. Therefore, the isotherms models proposed for
each dye has a good theoretical correlation with experimental

qmkLCe
1 + kLCe

15 C

(1)

Figure 2. Adsorption isotherms of Red 357 (a), Black 210 (b), and Yellow 194 (c) in dierent temperature conditions.
875

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

Figure 3. Adsorption kinetics of Red 357 (a), Black 210 (b), and Yellow 194 (c) dyes in dierent temperature conditions.

constant (kH). In this case, the number of sites is very superior


to the number of adsorbate molecules. Henrys law applies to
the adsorption on a uniform surface at suciently low
concentrations so that all molecules are isolated from their
nearest neighbors. This suggests that occurs a hydrophobic
interaction between the adsorbent and adsorbate,10 causing a
linear increase in adsorption capacity. The linear relationship
between the uid phase and adsorbed phase equilibrium
concentrations, with a constant of proportionality, which is
equal to the adsorption equilibrium constant known as Henrys
constant (kH). An increase in temperature from (15 to 45) C
causes a decrease in kH values of 2.172 Lg1 for 0.993 Lg1.
This indicates that adsorption capacities are approximately 2.2
times superior with reduction in temperature from (45 to 15)
C. At high temperatures, weak physical chemical interaction
takes place between the dye and the adsorbent due to reduction
in hydrogen bonds and van der Waals interactions, reducing the
adsorption capacity.14,22
For the Black 210 equilibrium data, Figure 2b shows that the
BET isotherm model predict satisfactorily experimental data.
The BET isotherm is an extension of the Langmuir theory for
monolayer adsorption to multilayer adsorption, when the
increase of adsorbateadsorbate interaction and multilayer
formation occurs due to secondary adsorption at a given
site.12,23 The multilayer formation may occur due to a change in
organizational form of dye molecules arranged on the surface of
the adsorbent, in horizontal to vertical alignment, or due to
solubility reduction caused by supercial hydrophobic interactions between the adsorbate and the adsorbent.10,15,17,23
These phenomena can be complemented by chemical
interactions between adsorbate in solution and one that has
been adsorbed, since the Black 210 dye has R-NH2 group,

data of adsorption equilibrium and, thus, can be used to


represent the equilibrium curves. The solid lines traced in
Figure 2 represent the correlation of the proposed models to
equilibrium experimental data.
For the Red 357 adsorption isotherm (Figure 2a), the
Langmuir theory (eq 1) is observed when the adsorbent surface
and the adsorptions sites are energetically homogeneous and
only one adsorption occurs by site.13 This behavior indicates
that high adsorption capacities are observed in the solid phase
with low concentrations of solute in the liquid phase, due to the
high anity of the dye molecules for the leather surface.
Therefore, Table 2 shows an increase in qm values and decrease
in kL values with temperature increase from (15 to 45) C for
Red 357 adsorption. The behavior of these parameters indicates
that higher temperatures increase the maximum adsorption
capacity, which can only be achieved with higher liquid
concentrations. The observed values of qm in this study are
higher than those reported in other research for the adsorption
of dyes for alternative and conventional adsorbents, as the case
of Reactive Red (C.I. 18286) by leather wastes (56 mgg1 to
163 mgg1) and activated carbon (48 mgg1),6 FD&C 40 by
Spirulina platensis (225.2 mgg1 at 25 C and pH = 4),14 and
methylene blue and crystal violet by clay (50.8 and 57.8,
respectively).16 An increase in the maximum adsorption
capacity and decrease in kL constant with the increase in
temperature was observed with dye adsorption on FD&C Red
40 and C.I. Reactive Black 13 dyes by chitosan.13,21
Concerning the linear isotherm shape, as it is observed for
isotherm data from Yellow 194 dye in Figure 2c, the adsorption
capacity increases linearly with an increase in equilibrium
concentration. Henrys law applies when the relationship of
equilibrium between concentrations on the uid phase and
solid phase are linear (C1 type) and proportional to Henrys
876

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

Figure 4. Plotting of the liquid lm diusion model: (a) Red 357; (b) Black 210; (c) Yellow 194.

adsorption. The Cs in the adsorption of crystal violet dye in


groundwater by clay was 116 mgg1, whereas the presence of
NH4Cl resulted in reduction of Cs from 137 mgg1 to 116
mgg 1. 16 Besides, the C s value of pentachlorophenol
adsorption by carbonized bark was 5.43 mgg1.15,23 A decrease
in k1 and k2 values indicates that adsorption of Black 210 is
favored by low temperatures.
3.2. Adsorption Kinetics. Figure 3 shows the eect of
temperature on the adsorption kinetics of dyes by chromium
tanned leather waste. The adsorption capacity of Yellow 194
increases rapidly during the rst hour of adsorption and then
remains nearly constant after it, suggesting that equilibrium was
quickly reached. For temperatures of (35 and 25) C, although
the adsorption capacity of Red 357 is initially lower than that of
Yellow 194 dye, after (30 to 60) min, the adsorption capacity of
Red 357 dye is higher than that of the Yellow 194. Higher
temperatures are associated with a greater adsorption capacity
of Red 357 dye compared to Yellow 194. Experiments have
shown a clear trend to increased absorption capacity of the Red
357 and Black 210 dyes after 120 min, meaning that
equilibrium adsorption capacity has not been reached.
Adsorption rates involve the following mechanisms: (i) liquid
bulk diusion; (ii) convection in mass boundary layer
surrounding the particle, or liquid lm diusion; (iii)
intraparticle diusion (surface or pore diusion); (iv) available
adsorption sites. However, bulk diusion and adsorption
reaction are typically instantaneous, and the mass transfer is
controlled only by convection in the boundary layer or
intraparticle diusion separately or simultaneously.24,25
In the case of liquid mass transfer, or external convection, the
model assumes for short periods of time, that the diusion step
does not aect the adsorption rate. The model can be described
by the liquid linear driving force (LDF), according eq 4:24,25

which under acidic conditions can be protonated, resulting in


additional adsorption of dye in solution.
Therefore, for the Black 210 dye adsorption, there is a strong
inuence of temperature on the equilibrium sorption. The
increase of temperature favor the formation of monolayers,
causing an increase in the qBET value of 42.3 mgg1 to 138.1
mgg1, which were superior than the values observed for qBET
of methylene blue and crystal violet dyes present in
groundwater by clay (44 mgg1 and 47 mgg1, respectively)16
and pentachlorophenol by carbonized bark (0.84 mgg1).15,23
As the temperature increase occurs, an increase of free volume
and a decrease of the interaction between solvents and solid
surfaces are observed, exposing a higher number of adsorption
sites, which are favorable to adsorption.
However, the values of k1 and k2 decrease with increasing
temperature. Like kL in the Langmuir model, k1 represents the
inverse of the equilibrium concentration in the liquid when the
adsorption capacity reaches half the capacity of monolayer
adsorption (or f(1/k1) = 0.5qBET, where f is the function of the
BET isotherm). In other words, the k1 increase indicates that
smaller equilibrium concentrations are necessary to saturate the
monolayer. This implies a more rectangular isotherm curve to
the achievement of monolayer saturation, as it is clearly seen in
Figure 2b.
The k2 values represent the inverse of the concentration
value when the isotherm becomes a vertical line (Cs = 1/k2)
and are associated with supercial solubility of dye.16 In general,
the solubility of the dyes is also increased with higher
temperatures. Therefore, adsorbateadsorbate interactions are
reduced. Thus, a temperature reduction from (45 to 15) C
reduces Cs values from 666.6 mgL1 to 476.2 mgL1,
indicating that high adsorption capacity by multilayer formation
is obtained with lower equilibrium concentrations, favoring
877

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

Figure 5. Plotting of intraparticle diusion model: (a) Red 357; (b) Black 210; (c) Yellow 194.

qt

dq
= k f a(C Ce)
dt

(4)

qe

= A

Dst
R p2

(8)

where q is the dye concentration on the solid phase (mgg ), kf


is the mass transfer coecient on the boundary layer
(mmin1), a is the specic surface area (m2g1), and C and
Ce are the dye concentrations in the liquid phase (mgm3).
Using appropriate initial conditions, the solution of eq 4 is
given according to eq 5.
C Ce
w
ln
= k f ap t
V
C0 Ce

where Rp is the mean radius of the particles and A equal to 3.38


and 4.24 for the linear isotherm and for the rectangular
isotherm, respectively. Plotting qt versus t1/2 (Webber and
Morris plot, Figure 5) it shows multilinearity portions; an initial
and curved behavior indicates the convective resistance, and the
nal linear portions, relative to solid diusion eects.2,2931
Figures 4 and 5 shows that Red 357 and Yellow 194 dye have
linear portions starting of the origin throughout the experiment
for both propose the mass transfer mechanisms. This suggests
that the two proposed models are able to explain the kinetics of
adsorption and the adsorption rate is controlled simultaneously
by uid-lm mass transfer and intraparticular diusion.
However, by adsorption of Black 210, the sharp curvature in
the Figure 4b and linear portion starting throughout the
experiment of the Weber and Morris plot (Figure 5b) suggest
that the adsorption rate mechanisms is of intraparticular
diusion. The resistance mechanism, boundary layer convection, or intraparticle diusion are evaluated by nondimensional Biot number (NBi), calculated from eq 9.32,33

(5)

Plotting the left side of eq 5 versus the time will provide


linearized data (Figure 4), with angular coecient equal to kfa,
for the range where convection limits the rate of adsorption.
When the internal resistance is predominant in the mass
transfer process, the intraparticle diusion can be represented
by the HSDM model, according to eq 6.24,26,27
2Q
q
2 Q
= Ds 2 +

t
R R
R

(6)

NBi =

Using appropriate boundary and initial conditions, the


solution of the model is given according to eq 7.
qt
qe

=1

n=1

Dt
1
n 2 2 s
exp
n2
R p2

k f R pC0
Dsp qo

(9)

where Rp is the particle radius (m), p is the adsorbent density


(kg3m3), and q0 is the adsorption capacity (mgg1)
considering Ce = C0. Table 3 shows the values obtained for
the convection coecient, the eective diusion inside the solid
particle, calculated according to eqs 5 and 7, respectively, and
dimensionless Biot number.

(7)

Suziki28 and Qiu et al.24 reports that, for short periods of


time, when qt/qe is less than 0.3, eq 7 approximates to eq 8:
878

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

external mass transfer control predominates in the adsorption


process of dye Yellow 194, while both mechanisms are active in
the adsorption of Red 357 dye and the intraparticular diusion
controls the mass transfer dye Black 210, due the low values of
solid diusion coecients.
3.3. Adsorption Thermodynamics. Adsorption thermodynamics was determined using the thermodynamic equilibrium coecients obtained at dierent temperatures for the
three dyes, to verify possible adsorption mechanisms.
The Gibbs free energy change (G) is associated to
spontaneity of the process, and it is calculated according to
eq 10.

Table 3. Convective and Intraparticle Diusion Coecients


and Biot Number of Tannery Dye Adsorption by CTLW
dye/temperature

kf

mmin1

15
25
35
45

4.41106
3.98106
4.41106
4.95106

15
25
35

7.11106
4.46106
4.77106

15
25
35
45

1.33105
1.39105
1.58105
2.19105

Ds
R2
Red 357
0.888
0.849
0.963
0.979
Black 210
0.852
0.812
0.811
Yellow 194
0.921
0.935
0.929
0.977

m2min1

R2

Bi

1.36109
2.87109
4.89109
1.01108

0.983
0.994
0.989
0.994

6.78
2.75
1.73
0.91

5.621012
3.251011
6.151011

0.951
0.956
0.981

147.5
90.8
50.3

1.48108
1.63108
2.16108
3.56108

0.937
0.985
0.968
0.981

0.40
0.44
0.65
0.60

G = RT ln kD

(10)

where kD is the thermodynamic equilibrium constant, R is the


universal gas constant (8.314 Jmol1K1), and T is the
temperature (K). kD was obtained from the slope of initial
linear portion of qe vs Ce,38 and converted to a dimensionless
constant according Milonjic.39
Gibbs free energy is the dierence between adsorption
enthalpy (H) and adsorption entropy (S), at constant
temperature. Thus, by applying this concept to eq 12, vant
Hos (eq 11) can be used to determine the thermochemical
parameters through the plot of ln kD vs 1/T obtaining a angular
coecient equal to H/R and linear coecient equal to S/R.

The data in the Table 3 show that, in magnitude, the solid


diusion of the dye in the adsorbent is Yellow 194 > Red 357
Black 210, while the convection coecient is on the order of
105 mmin1 to 106 mmin1 for the three dyes. The high
diusivity values of the Yellow 194 dye makes that the
equilibrium is reached within a few minutes, while the Red 357
dye equilibrium is reached in a few hours and the Black 210 dye
takes days or even weeks, as can be evaluated in Figure 3. The
kf values observed were inferior that for the adsorption
herbicides and pentaclorophenol by activated charcoal (104
mmin1 and 103 mmin1, respectively)32,34 and pesticide by
acid treated oil shale ash (102 mmin1).35 However, the Ds
values of the Yellow 194 and Red 357 dyes were superior to
other alternative adsorbents, such as FD&C Red 40 dye by
chitosan (1010 m2min1) and pesticide by acid treated oil
shale ash (109 m2min1). This may be because the stirring of
the system and the high concentration of dye used in this study,
which causes an increase in solution viscosity. These dierences
can be explained, because while the rst is aected principally
by the stirring of the system, that is not identical in the dierent
studies, both could also have aected by initial dye
concentration, which was higher in this present work, reducing
viscosity and causing a decrease in external and internal mass
transfer coecients.32,34
In relation to the eect of temperature on the diusion
coecient, Table 3 shows that for the three dyes the increase in
temperature favored intraparticle diusion due the increase in
Ds values. Similarly, for the dyes Red 357 and Yellow 194,
increase in temperature led to increase in external convection
coecient (kf), reducing resistance to mass transfer in
boundary layer. In general, the convection and intraparticle
diusion coecients increase due the increase of adsorbate
molecular diusion in the solvent at greater temperatures.8,32,34,36 In addition, increasing temperature reduces the
viscosity of a solution, facilitating penetration of the adsorbate
in the solid sorbent.35
The decrease in Biot number indicates that as temperature
rises the thickness of the boundary layer surrounding the
adsorbent and the mass transport resistance of the adsorbate in
the boundary layer is reduced. According Cooney,37 for NBi <
0.5, adsorption mass transfer is complete dominance of the
external resistance, while for NBi > 30, there is reasonably
complete dominance of intraparticle resistance. Therefore, the

ln kD =

H
S
+
RT
R

(11)

Moreover, the diusion coecients are represented as an


exponential function of temperature, according to the
Arrhenius equation.27,28,40
E
Ds = D0 exp a
RT

(12)

where Ea is the activation energy (kJ mol1) and D0 is the


diusion at a temperature of 0 K. Then, Ea and D0 are
determined by regression of the linearized form of the
Arrhenius equation, by plotting ln(Ds) vs 1/T, obtaining a
linear coecient equal to ln(D0) and a slope equal to Ea/R.
Figures 6 and 7 shown vant Hos and Arrhenius plots for Red
357, Black 210, and Yellow 194 dyes adsorption. Table 4 shows
Gibbs free energy change (G), enthalpy (H), and entropy
(S), activation energy (Ea) obtained through eqs 10, 11, and
linearized form of eq 12 and respective R2 of the plots.

Figure 6. vant Ho plot of tannery dye adsorption by CTLW.


879

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

Figure 7. Arrhenius plot of tannery dye adsorption by CTLW.


Figure 8. FT-IR spectra of: (a) CLTW; (b) CLTW + Red 357; (c)
CLTW + Black 210; (d) CLTW + Yellow 194.

Table 4. Thermodynamics Parameters of Tannery Dye


Adsorption by CTLW
parameters

T/C

Red 357

Black 210

Isotherm Thermodynamic Data


15
9.25
11.57
25
8.66
10.01
35
8.43
4.49
45
8.24
ND
R2
0.952
0.920
0.920
H (kJmol1)
18.6
112.9
24.9
S (kJmol1K1)
0.033
0.35
0.07
Kinetics Thermodynamic Data
R2
0.996
0.940
Ea (kJmol1)
23.6
89.4
G (kJmol1K1)

1033) cm1 was observed due the vibration of SO3+ group.


Additionally, visible changes are observed in peaks of (1239 and
1127) cm1. Such behaviors are caused by changes in the
molecular composition of the groups possibly by dye
electrostatic interaction with the amino group of the leather
sorbent, proving the chemical nature of adsorption.

Yellow 194
1.86
1.60
0.24
0.02

4. CONCLUSIONS
Chromium-tanned leather waste was used as alternative
adsorbents of Red 357, Black 210, and Yellow 194 dyes in
aqueous solutions at dierent temperatures. The maximum
monolayer adsorption capacity of Red 357 and Black 210 dyes
were 250.6 mgg1 at 45 C and 138.1 mgg1 at 35 C,
respectively, and the temperature reduction led to a decrease of
monolayer saturation. The observed values for respective
isotherms were superior to adsorption of others dyes by
conventional or alternative adsorbents. Moreover, for the Black
210 dye adsorption, a nonconventional isotherm, the reduction
in temperature favored the formation of multilayers at lower
equilibrium concentrations, and the supercial solubility at 15
C was of 476.2 mgL1. Yellow 194 adsorption isotherms did
not provide formation of mono- or multilayers, and the
adsorption capacity at equilibrium for a liquid concentration of
100 mgL1 decreased from 217.2 mgg1 to 99.3 mgg1 with a
temperature increase of (15 to 45) C.
The solid diusion coecients (Ds) were in the order of 108
2
m min1, 109 m2min1, and 1011 m2min1 for Red 357,
Black 210, and Yellow 194 dyes, and the temperature reduction
lead to an increase of diusion coecients. Convective
coecients (kf) were in the order of 105 to 106 mmin1
for all dyes and conditions studied. A Biot number (NBi) greater
than 0.5 shows that both mechanisms of mass transfer
(convection and diusion) control the adsorption process of
Red 357. However, for Black 210 with NBi > 30 and for Yellow
194 with NBi < 0.5 it was found that solid diusion and
boundary layer mass transfer have complete dominance over
the mass transfer, respectively.
Negative values of G, H, and S suggest a spontaneous,
exothermic, and favorable adsorption process for three studied
dyes. Moreover, thermodynamic studies of activation energy
(Ea) and FT-IR analysis suggest that dye adsorption has a
chemical nature.
The results observed for the adsorption capacity and mass
transfer rates in the order of magnitude of other conventional
and nonconventional adsorbents clearly demonstrated the

0.927
22

Negative G values indicate that the adsorption process is


more favorable and spontaneous at low temperatures, not
requiring external energy from outside of the system.14
However, for Black 210 dye adsorption, a signicant increase
in G values as temperature increases leads to a decrease in the
adsorption driving force, reducing adsorption.13,30 Negative H
values indicate that the adsorption process is exothermic
because there is a decrease in system enthalpy. A decrease in
enthalpy indicates that the Black 210 dye adsorption is most
aected by temperature. Negative S values indicated that dyes
are arranged in ordered manner in the solid phase. Therefore,
adsorption causes a reduction in system disorder. For Red 357
and Black 210 dyes, although internal mass transfer resistance
was important, activation energy (Ea) values were higher than
50 kJmol1. These values suggest that the adsorption process
has a chemical nature.29 However, for Yellow 194 dye,
intraparticle diusion can be neglected, for it is not inuenced
by temperature. Thus, a low Ea value is observed.
3.4. IR Spectroscopy. Figure 8 shows FT-IR analysis of
leather waste adsorbent before and after adsorption of dyes.
Before adsorption, the peaks near (3500 to 3300) cm1
represent the symmetric stretching vibration of the OH
group that overlaps the vibration of the NH group. Absorption
peaks for the sample in 1960 cm1 may result from the
stretching vibration of CH. The structure of the protein
material is demonstrated by the signals in 1659 cm1 in relation
to the carbonyl group (CO) and in 1547 cm1 to the NH
group. Moreover, the peak at 1239 cm1 is attributed to a
combination of NH bending and CN stretching and the
peak of 1127 cm1 to CO stretching.11,12 After the
adsorption process, the presence of a peak near (1042 to
880

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

(15) Edgehill, R. U.; (Max) Lu, G. Q. Adsorption characteristics of


carbonized bark for phenol and pentachlorophenol. J. Chem. Technol.
Biotechnol. 1998, 71, 2734.
(16) Al-Futaisi, A.; Jamrah, A.; Al-Hanai, R. Aspects of cationic dye
molecule adsorption to palygorskite. Desalination 2007, 214, 327342.
(17) Grainger, S.; Fu, G. Y.; Hall, E. R. Aspects of cationic dye
molecule adsorption to palygorskite. Water, Air, Soil Pollut. 2011, 217,
233244.
(18) Li, Z.; Tang, X.; Chen, Y.; Wang, Y. Behaviour and mechanism
of enhanced phosphate sorption on loess modified with metals:
Equilibrium study. J. Chem. Technol. Biotechnol. 2009, 84, 595603.
(19) Al-Ghouti, M. A.; Khraisheh, M. A. M.; Allen, S. J.; Ahmad, M.
N. The removal of dyes from textile wastewater: A study of the
physical characteristics and adsorption mechanisms of diatomaceous
earth. J. Environ. Manage. 2003, 69, 229238.
(20) Khraisheh, M. A. M.; Al-Ghouti, M. A.; Allen, S. J.; Ahmad, M.
N. Effect of OH and silanol groups in the removal of dyes from
aqueous solution using diatomite. Water Res. 2005, 39, 922932.
(21) Annadurai, G.; Ling, L. Y.; Lee, J. F. Adsorption of reactive dye
from an aqueous solution by chitosan: isotherm, kinetic and
thermodynamic analysis. J. Colloid Interface Sci. 2008, 286, 3642.
(22) Anjos, F. S. C.; Vieira, E. F. S; Cestari, A. R. Interaction of
Indigo Carmine dye with chitosan evaluated by adsorption and
thermochemical data. J. Colloid Interface Sci. 2002, 253, 243246.
(23) Ebadi, A.; Soltan Mohammadzadeh, J. S.; Khudiev, A. What is
the correct form of BET isotherm for modeling liquid phase
adsorption? Adsorption 2009, 15, 6573.
(24) Qiu, H.; Lv, L.; Pan, B.-C.; Zhang, Q.-J.; Zhang, W.-M.; Zhang,
Q.-X. Critical review in adsorption kinetic models. J. Zhejiang Univ.:
Sci. A 2009, 10, 716724.
(25) Cheung, W. H.; Szeto, Y. S.; McKay, G. Intraparticle diffusion
processes during acid dye adsorption onto chitosan. Bioresour. Technol.
2007, 98, 28972904.
(26) Cooney, D. O. Adsorption design for wastewater treatment; Lewis
Publisher: New York, 1999.
(27) Cotoruelo, L. M.; Marques, M. D.; Rodrguez-Mirasol, J.;
Rodrguez, J. J.; Cordero, T. Lignin-based activated carbons for
adsorption of sodium dodecylbenzene sulfonate: Equilibrium and
kinetic studies. J. Colloid Interface Sci. 2009, 332, 3945.
(28) Suzuki, M. Adsorption Engineering; Elsevier Science Publishers B.
V.: Amsterdam, The Netherlands, 1990.
(29) Baccar, R.; Blanquez, P.; Bouzid, J.; Feki, M.; Sarra, M.
Equilibrium, thermodynamic and kinetic studies on adsorption of
commercial dye by activated carbon derived from olive-waste cakes.
Chem. Eng. J. 2010, 165, 457464.
(30) Senthil Kumar, P.; Ramalingam, S.; Senthamarai, C.; Niranjanaa,
M.; Vijayalakshmi, P.; Sivanesan, S. Adsorption of dye from aqueous
solution by cashew nut shell: Studies on equilibrium isotherm, kinetics
and thermodynamics of interactions. Desalination 2010, 261, 5260.
(31) Dotto, G. L.; Pinto, L. A. A. Adsorption of food dyes onto
chitosan: Optimization process and kinetic. Carbohydr. Polym. 2011,
187, 164170.
(32) Kim, T.-Y.; Park, S.-S.; Kim, S.-J.; Cho, S.-Y. Separation
characteristics of some phenoxy herbicides from aqueous solution.
Adsorption 2008, 14, 611619.
(33) Piccin, J. S.; Dotto, G. L.; Vieira, M. L. G.; Pinto, L. A. A.
Kinetics and mechanism of the food dye FD&C Red 40 adsorption
onto chitosan. J. Chem. Eng. Data 2011, 56, 37593765.
(34) Leyva-Ramos, R.; Bernal-Jacome, L. A.; Mendoza-Barron, J.;
Hernandez-Orta, M. M. G. Kinetic modeling of pentachlorophenol
adsorption onto granular activated carbon. J. Taiwan Inst. Chem. Eng.
2009, 40, 622629.
(35) Al-Qodah, Z.; Shawaqfeh, A. T.; Lafi, W. K. Two-resistance
mass transfer model for the adsorption of the pesticide deltamethrin
using acid treated oil shale ash. Adsorption 2007, 13, 7382.
(36) Wilke, C. R.; Chang, P. Correlation of diffusion coefficients in
dilute solution. AIChE J. 1957, 264270.
(37) Cooney, D. O. Comparison of simple adsorber breakthrough
curve method with exact solution. AIChE J. 1993, 30, 355358.

eciency of leather waste to dyes removal present in tannery


wastewater, allowing survival of an industrial byproduct before
the disposal in landlls of hazardous waste or by other
traditional method.

AUTHOR INFORMATION

Corresponding Author

*Tel.: +55 51 3308 3954; Fax: +55 51 3308 3277. E-mail


address: jspiccin@enq.ufrgs.br (J. S. Piccin), mariliz@enq.ufrgs.
br (M. Gutterres).
Funding

The authors would like to thank CNPq (National Council of


Science and Technological Development) for the nancial
support by Edict CTAGRO No. 40/2008 and the Business
Leather Unit of Lanxess Company for the technical support.
Notes

The authors declare no competing nancial interest.

REFERENCES

(1) Maljaei, A.; Arami, M.; Mahmoodi, N. M. Decolorization and


aromatic ring degradation of colored textile wastewater using indirect
electrochemical oxidation method. Desalination 2009, 249, 1074
1078.
(2) Noroozi, B.; Sorial, G. A.; Bahrami, H.; Arami, M. Equilibrium
and kinetic adsorption studies of a cationic dye by a natural adsorbentSilkworm pupa. J. Hazard. Mater. 2007, 139, 167174.
(3) Zhang, M.; Shi, B. I. Adsorption of dyes from aqueous solution by
chromium-containing leather waste. J. Soc. Leather Technol. Chem.
2004, 88, 236241.
(4) Gammoun, A.; Tahiri, S.; Albizane, A.; Azzi, M.; Moros, J.;
Garrigues, S.; de la Guardia, M. J. Separation of motor oils, oily wastes
and hydrocarbons from contaminated water by sorption on chrome
shavings. J. Hazard. Mater. 2007, 145, 148153.
(5) Oliveira, D. Q. L.; Goncalves, M.; Oliveira, L. C. A.; Guilherme,
L. R. G. Removal of As(V) and Cr(VI) from aqueous solutions using
solid waste from leather industry. J. Hazard. Mater. 2008, 151, 280
284.
(6) Oliveira, L. C. A.; Goncalves, M.; Oliveira, D. Q. L.; Guerreiro,
M. C.; Guilherme, L. R. G.; Dallago, R. M. Solid waste from leather
industry as adsorbent of organic dyes in aqueous-medium. J. Hazard.
Mater. 2007, 141, 344347.
(7) Cooper, M.; Gutterres, M.; Marclio, N. R. Environmental
Developments and Researches in Brazilian Leather Sector. J. Soc.
Leather Technol. Chem.s 2011, 95, 24349.
(8) Fujiki, J.; Shinomiya, T.; Kawakita, T.; Ishibashi, S.; Furuya, E.
Experimental determination of fluid-film mass transfer coefficient from
adsorption uptake curve. Chem. Eng. J. 2011, 173, 4954.
(9) Piccin, J. S.; Gomes, C. S.; Feris, L. A.; Gutterres, M. Kinetics and
isotherms of leather dyes adsorption by tannery solid waste. Chem.
Eng. J. 2012, 183, 3038.
(10) Giles, C. H.; MacEwan, T. H.; Nakhwa, S. N.; Smith, D. Studies
in adsorption. Part XI. A system of classification of solution adsorption
isotherms, and its use in diagnosis of adsorption mechanisms and in
measurement of specific surface areas of solids. J. Chem. Soc. 1960,
39733993.
(11) Mohamed, O. A.; Kassem, N. F. Utilization of waste leather
shavings as filler in paper making. J. Appl. Polym. Sci. 2010, 118, 1713
1719.
(12) Anandkumar, J.; Mandal, B. Adsorption of chromium(VI) and
Rhodamine B by surface modified tannery waste: Kinetic, mechanistic
and thermodynamic studies. J. Hazard. Mater. 2011, 186, 10881096.
(13) Piccin, J. S.; Dotto, G. L.; Pinto, L. A. A. Adsorption isotherms
and thermochemical data of FD&C Red no. 40 binding by chitosan.
Braz. J. Chem. Eng. 2011, 28, 195302.
(14) Dotto, G. L.; Lima, E. C.; Pinto, L. A. A. Biosorption of food
dyes onto Spirulina platensis nanoparticles: Equilibrium isotherm and
thermodynamic analysis. Biores. Technol. 2012, 103, 123130.
881

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

Journal of Chemical & Engineering Data

Article

(38) Liu, Y. Is the free energy change of adsorption correctly


calculated? J. Chem. Eng. Data 2009, 54, 19811985.
(39) Milonjic, S. K. A consideration of the correct calculation of
thermodynamic parameters of adsorption. J. Serb. Chem. Soc. 2007, 72,
13631367.
(40) Liang, H.; Gao, H.; Kong, Q.; Chen, Z. Adsorption of
tetrahydrofuran + water solution mixtures by zeolite 4A in a fixed bed.
J. Chem. Eng. Data 2007, 52, 695698.

882

dx.doi.org/10.1021/je301076n | J. Chem. Eng. Data 2013, 58, 873882

S-ar putea să vă placă și