Вы находитесь на странице: 1из 8

Desalination 268 (2011) 150–157

Contents lists available at ScienceDirect

Desalination
j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / d e s a l

Biosorption of Ni(II) from aqueous phase by Moringa oleifera bark, a low


cost biosorbent
D. Harikishore Kumar Reddy a, D.K.V. Ramana a, K. Seshaiah a,⁎, A.V.R. Reddy b
a
Analytical & Environmental Chemistry Division, Department of Chemistry, Sri Venkateswara University, Tirupati 517 502, India
b
Analytical Chemistry Division, Bhabha Atomic Research Centre, Trombay, Mumbai 400 085, India

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

Article history: Moringa oleifera bark (MOB), an agricultural solid waste by-product has been developed into an effective and
Received 23 July 2010 efficient biosorbent for the removal of Ni(II) from aqua solutions. The biosorbent was characterized by X-ray
Received in revised form 27 September 2010 diffraction (XRD), scanning electron microscopy (SEM), elemental analysis and FTIR analyses. The
Accepted 2 October 2010
experimental equilibrium adsorption data were analyzed by four widely used two-parameter equations –
Available online 30 October 2010
Langmuir, Freundlich, Dubinin–Radushkevich (D–R), and Temkin isotherms. Among the four isotherm
Keywords:
models Langmuir model provided a better fit with the experimental data than others as revealed by high
Moringa oleifera bark correlation coefficients, low chi-square values. The kinetics data fitted well into the pseudo-second-order
Biosorption model with correlation coefficient greater than 0.99. Desorption experiments were carried out to explore the
Nickel feasibility of regenerating the biosorbent and the biosorbed Ni(II) from MOB was desorbed using 0.2 M HCl
Kinetics with an efficiency of 98.02% recovery. The thermodynamic parameters (ΔH, ΔS and ΔG) of the nickel ion
Isotherms uptake onto MOB indicated that, the process is endothermic and proceeds spontaneously. The findings of the
Thermodynamics present study indicates that MOB can be successfully used for separation of Ni(II) from aqueous solutions.
© 2010 Elsevier B.V. All rights reserved.

1. Introduction cally viable sources of biosorbents; this makes biosorption an


inexpensive alternative treatment method. Several workers have
Industrial effluents are the major sources for contamination of reported on the potential use of agricultural by-products as good
water resources by heavy metals. Mining and metallurgy of nickel, substances for the removal of metal ions from aqueous solutions and
stainless steel, aircraft industries, nickel electroplating, battery and wastewaters [4–11].
manufacturing, pigments and ceramic industries wastewaters contain Moringa oleifera belongs to the family Moringaceae, which is native
high amounts of nickel ions [1]. Nickel is known as one of the most to the sub-himalayan tracts of India. M. oleifera bark is abundantly
common toxic metals, and the US Environmental Protection Agency available in India with no commercial use. Since its fruits and leaves
(EPA) has established 0.5 mg L−1 as the permissible concentration for are edible and consumed in all the seasons, these trees are widely
nickel in drinking water [2]. In view of this, it is necessary to remove cultivated species in India. Due to this, massive amount of bark is
nickel from wastewater before they are discharged into natural water produced, which is being disposed off as waste. The plant contains
bodies. The current technologies for the removal of nickel include various amino acids, fatty acids, vitamins and nutrients [12]
chemical precipitation, ion exchange, or electrochemical processes. glucosinolates and phenolics (flavonoids, anthocyanins, proanthocya-
Many of these methods are neither effective nor economical, nides and cinnamates) which are the functional groups that are
especially when used for the reduction of heavy metal ions to low capable of anchoring metals. The bark tissue of M. oleifera contains 4-
concentrations. Hence, there is a need to develop cost-effective (α-L-rhamnopyranosyloxy)-benzylglucosinolate [13]. Every glucosi-
technologies for treating the wastewater that reduce metal ion nolate contains a central carbon atom which is bonded to the
concentrations to environmentally acceptable levels. thioglucose group (making a sulfated ketoxime) via a sulfur atom and
Biosorption has the potential to contribute to the achievement of to a sulfate group via a nitrogen atom. These functional groups
this goal. Biosorption is a process that uses inexpensive biomaterials containing sulfur and nitrogen are good metal sequesters from the
to sequester metals from aqueous solutions [3] and the biomaterials aqueous solution. Because of the composition of M. oleifera bark and
used in this process are termed as biosorbents. The by-products from its availability, it was chosen as a biosorbent for the removal of Ni(II)
agricultural, food and pharmaceutical industries provide economi- from aqueous solutions.
In the present study, a simple and economic preparation of the
biosorbent was performed and biosorption experiments have been
⁎ Corresponding author. Tel.: +91 877 2289303. conducted to determine the metal biosorption mechanism. The main
E-mail address: seshaiahsvu@yahoo.co.in (K. Seshaiah). objectives of the present study include (1) characterizing the

0011-9164/$ – see front matter © 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.desal.2010.10.011
D.H.K. Reddy et al. / Desalination 268 (2011) 150–157 151

biosorbent through FTIR, XRD, SEM and Elemental analysis to know The amount of metal ion sorbed onto the MOB, Qe, was computed
the physico-chemical properties of the biosorbent, (2) to study the by the following equation:
sorption potential of MOB for Ni(II) removal from water under
equilibrium conditions and (3) to understand the kinetic mechanism. v
The data from the experiments were fitted with different kinetic Qe = ðC −Ce Þ ð1Þ
m 0
models to identify the adsorption mechanism.

Where, C0 and Ce are the initial and equilibrium concentrations of


2. Materials and methods Ni(II) in solution whereas v and m are solution volume and mass of
biosorbent, respectively.
2.1. Preparation of the biosorbent
3. Result and discussion
Samples of M. oleifera bark were collected from Punganur (Andhra
Pradesh, India) in December 2007, air dried and ground in a mill to get 3.1. Characterization of the biosorbent
fine powder. The powder was washed twice with deionized water and
dried at 60 °C for 24 h, then boiled in double distilled water by Characterization of biosorbent surface and structure hold keys to
changing the water repeatedly until water becomes colorless, which understanding the metal binding mechanism onto biomass. The
indicates the removal of water soluble color compounds. The washed proximate and ultimate analysis of MOB is presented in Table 1.
and boiled bark was oven dried at 80 °C for 24 h and stored in Percent ash content, moisture content, fixed carbon, and bulk density
desiccator to prevent moisture adsorption before its use. This treated are also given in Table 1. Elemental analysis results showed that MOB
M. oleifera bark powder is called as MOB. is composed of 44.8 ± 1.50% carbon, 5.9 ± 0.18% hydrogen, 0.8 ± 0.01%
nitrogen, 0.9 ± 0.01% sulphur and 47.6 ± 1.82% oxygen. Sulphur
groups, which are soft bases, have chemical affinity towards metal
2.2. Chemicals and equipment ions and presence of sulphur in MOB qualifies it as a potential
biosorbent.
All reagents used were of AR grade. Deionized double distilled
water was used throughout the experimental studies. Stock solution 3.2. Scanning electron microscopy
(1000 mg L−1) was prepared by dissolving NiSO4·6H2O. This was
further diluted to obtain the desired concentration for practical use. Scanning electron microscope has been widely used to study the
ACS reagent grade HCl, NaOH and buffer solutions (E. Merck) were morphological features of the biosorbent. Study of the SEM micro-
used to adjust the solution pH. An Elico (LI-129) pH meter was used graphs of MOB showed in Fig. 1(a) and (b) indicated the presence of
for pH measurements. The pH meter was calibrated using buffer asymmetric pores and open pore structure, which may provide high
standard solutions of pH 4.0, 7.0 and 9.2. Fourier transform infrared internal surface area and a rough structure on the surface of MOB,
spectrophotometer (Thermo-Nicolet FT-IR, Nicolet IR-200, USA) was which is favorable for biosorption of Ni(II) from aqua solutions.
used to analyze the organic functional groups of the biosorbent. Vario
EL, Elementar, Germany was used for elemental analysis of the MOB. 3.3. X-ray diffraction
The metal concentrations in the samples were determined using
atomic absorption spectrophotometer (AAS, Varian spectra AA, 55 XRD pattern of the MOB shown in Fig. 2 illustrates the presence of
Australia). Wide angle X-ray diffraction (WAXD) patterns of powder a significant amount of amorphous material due to lignin and tannin
MOB sample was recorded on an X-ray diffractometer (XRD-6000, in the sample.
Shimadzu), by using Cu Kα radiation (λ = 1.5406A) at 40 kV and
30 mA with a scan rate of 5°/min in the range of 2θ from 8 to 50°. 3.4. Infrared spectroscopy
Scanning Electron Microscopy (Model Evo15, Carl Zeiss, England) was
used to study the surface morphology of the biosorbent. The FTIR spectra of MOB (Fig. 3a), showed the presence of many
functional groups, indicating the complex nature of MOB biosorbent.
A strong band at 3418 cm−1 indicated the presence of hydroxyl
2.3. Biosorption experiments groups. A peak at 2920 cm−1 is due to the C–H stretching frequency
and the peak at 1644 cm−1 is due to C=O stretching mode of the
Batch biosorption studies were performed by taking 50 mL of Ni primary and secondary amides (NH2CO). The peaks at 1510 cm−1 and
(II) solution with known pH in a 250 mL flask and by adding 200 mg 1375 cm−1 are indicative of the N–H stretching of the primary and
of MOB. The sealed flask was placed in a shaking incubator at a speed secondary amides, and the presence of amide (III) or sulfamide band,
of 300 rpm and temperature of 25 °C. After 3 h, biosorbent was respectively. Bands at 1320 and 1246 cm−1 indicate presence of
separated using an ashless filter paper and the concentration of metal carboxylic acids [14].Weak bands at 1462 and 1510 cm−1 are
ion in solution was analyzed by AAS. attributed to aromatic C=C and two sharp peaks at 1733 and
The influence of pH on Ni(II) biosorption was determined by 1644 cm−1 characteristic of carbonyl group stretching were also
equilibrating the suspensions in solutions of different pH ranging observed. The strong C–O band at 1054 cm−1 confirms the lignin
from 2.0 to 8.0. The effect of contact time on batch experiments was structure of the MOB [15]. The asymmetrical stretching vibration at
examined by varying the contact time of suspensions from 5 to 3418 cm− 1 was shifted to 3342 cm−1 after the biosorption of Ni(II).
100 min. For the adsorption isotherm studies, the initial metal ion The change in hydroxyl group after biosorption shown in Fig. 3b,
concentration was varied over a range of 20 to 200 mg L−1. The indicated that it had been changed from multimer to monopolymer or
concentration of MOB was varied between 100 and 800 mg to even dissociative state [16] which showed that the degree of hydroxyl
determine the ratio required for optimum biosorption. The interfer- polymerization in lignocellulose was decreased by binding of Ni(II).
ence caused by the presence of other cations that are normally The spectrum after interaction with Ni(II) showed increased intensity
present in water was checked by the addition of monovalent Na+ and and shifted to asymmetrical stretching bands at 1732, 1644 and
K+, and divalent Mg2+ and Ca2+ metal solution ranging from 0 to 1054 cm− 1 when compared with those observed in the spectra of
300 μg mL−1. MOB. These results indicated that the biosorption of Ni(II) occurs at
152 D.H.K. Reddy et al. / Desalination 268 (2011) 150–157

Table 1
Characteristics of Moringa oliefera bark (MOB).

Biosorbent Bulk Moisture Ash Fixed PZC Elemental analysis (%)


density content content carbon (%)
Carbon Hydrogen Nitrogen Sulphur Oxygena
(g/cm3) (%) (%)

MOB 0.5 ± 0.03 2.5 ± 0.08 11.1 ± 0.32 20.1 ± 0.51 4.2 ± 0.67 44.8 ± 1.50 5.9 ± 0.18 0.8 ± 0.01 0.9 ± 0.01 47.6 ± 1.82
a
Estimated by difference.

hydroxyl, carboxyl, and carbonyl functional groups present on the (PZC) of biosorbent was determined according to the procedure
surface of MOB. described by Lu et al. [18]. At pH below 4.2 ± 0.67 (PZC), the surface of
MOB would positively charged due to protonation. This protonation
3.5. Influence of pH effect was more pronounced at lower pH values due to the presence of
higher concentration of H+ ions in the solution which resulted in
It is well known that pH of the medium is most important factors more unfavorable Ni(II) biosorption. At the optimum pH value (pH
that influence the biosorption process [17]. The pH level affects the 6.0) the surface of the MOB is negatively charged and favorable to the
network of negative charge on the surface of the biosorbing cell walls, biosorption of Ni(II). Decreased biosorption at higher pH (pH N 6) was
as well as physicochemsitry and hydrolysis of the metal. Therefore, due to the formation of soluble hydroxylated complexes of the metal
preliminary experiments have been performed to find out the ions and their competition with the active sites, and as a consequence,
optimum pH for maximizing the metal removal. Percentage removal the retention had been decreased again. Therefore further experi-
of the metal ion as a function of pH is shown in Fig. 4. It has been ments were carried out with initial pH value of 6. Previous studies also
observed that under highly acidic conditions (pH ≈ 2.0) the amount of reported that the maximum biosorption efficiency for Ni(II) metal ion
Ni(II) removal was very small, while the sorption had been increased on biomass was observed at pH 6 [19–21].
with the increase in pH from 3.0 to 6.0 and then decreased in the
range 7.0 and 8.0. The lower removal efficiency at low pH is 3.6. Effect of biosorbent dose
apparently due to the presence of higher concentration of H+ in the
solution which compete Ni(II) ions for the adsorption sites of the Biosorption of Ni(II) onto MOB was studied by changing the
MOB. With increase in pH, the H+ concentration decreases leading to quantity of sorbent from 0.1 to 0.8 g in the test solution while
increased Ni(II) uptake. maintaining the initial concentration 50 mg mL−1, pH 6 and contact
On considering the point of zero charge of MOB another time 2 h constant. Biosorption of Ni(II) as a function of biomass shown
explanation for Ni(II) has been provided. The point of zero charge in Fig. 5, indicates the effect of sorbent dose on the Ni(II) biosorption
by MOB. Obviously, the biosorption efficiency increased as the sorbent
dose increased, but it remained almost constant when the sorbent
dose reached 0.4 g. This may be explained by the following analysis.
When sorbent ratio is small, the active sites for binding metal ions on
the adsorbent surface is less, so the adsorption efficiency is low; when
biosorbent dose increased more metal ions were adsorbed. Thus it
results in the increment of adsorption efficiency until saturation.

3.7. Effect of contact time and biosorption kinetics

For practical applications, the process design and operation


control, the sorption kinetics were very important. Sorption kinetics
in wastewater treatment was significant, as it provides valuable
insights into the reaction pathways and the mechanism of the
Intensity (CPS)

10 15 20 25 30 35 40 45 50
2θ (deg)

Fig. 1. SEM micrographs of MOB surface: (a) low magnification, (b) high magnification. Fig. 2. XRD spectrum of MOB.
D.H.K. Reddy et al. / Desalination 268 (2011) 150–157 153

100
(b)

90
Transmittance (%)

% Removal
80

(a)
70

60
Ni(II)

4000 3500 3000 2500 2000 1500 1000 500 50


100 200 300 400 500 600 700 800
Wavenumber (cm-1)
Amount of biosorbent (mg)
Fig. 3. FT-IR spectra of : (a) MOB ; (b) Ni-loaded MOB.
Fig. 5. Effect of sorbent dose on the sorption of Ni(II) by MOB. Error bars represent ± S.D.

sorption reactions. Biosorption of Ni(II) onto MOB at various initial The pseudo-second-order model proposed by Ho and McKay [23]
concentrations had been carried out at different time intervals (5– was based on the assumption that the adsorption follows second-
100 min) and shown in Fig. 6. order chemisorption. The linear form can be written as follows:

3.8. Biosorption kinetics t 1 t


= + ð3Þ
q k2 q2e qe
In order to further investigate the biosorption mechanism of Ni(II)
onto MOB and rate-controlling steps, a kinetic investigation was where, k2 (g mg−1 min−1) is the rate constant of adsorption. By
conducted, Pseudo-first, pseudo-second-order and intra-particle plotting a curve of t/qt against t, qe and k2 can be evaluated. The initial
diffusion kinetic models have been used for testing experimental data. adsorption rate, h0 (mg g−1 min−1) is defined as
The pseudo-first-order kinetic model was proposed by Lagergren
[22]. The integral form of the model generally expressed as follows: 2
h0 = k2 qe ð4Þ
k1
logðqe −qÞ = log qe  t ð2Þ
2:303 The values of qe, k2, h0 and R2 are listed in Table 2. The dependence
of t/qt vs t gives a linear relation for all the experimental concentra-
Where, qe (mg g−1) and qt (mg g−1) are the adsorption amount at tions (Fig. 7.), and all the R2 values are closer to 1 (Table 2), confirming
equilibrium and time t (min), respectively. k1 (min−1) is the rate the applicability of pseudo-second-order equation. In addition, there
constant in the pseudo-first-order adsorption process. The constants is only a little difference between qe,exp and qe,cal (Table 2), reinforcing
were determined experimentally by plotting log(qe − qt) vs t. The the applicability of this model. From Table 2, it can be observed that,
results are given in Table 2. The theoretical values (qe,cal) were far with an increase in initial metal concentration, the initial sorption rate
lower than those experimental data, qe,exp and low correlation (h0) also increased. In accordance with the pseudo-second reaction
coefficient values obtained for the pseudo first-order model indicate mechanism, the overall rate of Ni(II) sorption processes appears to be
that sorption is not occurring exclusively onto one site per ion. controlled by the chemical processes, through sharing of electrons

15
100
50 mg L-1
25 mg L-1
80
10 mg L-1
10
% Removal

qe (mg g-1)

60
Ni(II)

40
5

20

0 0
0 2 4 6 8 0 20 40 60 80 100
Initial pH Contact time (min)

Fig. 4. Effect of pH on the sorption of Ni(II) by MOB. Error bars represent ± S.D. Fig. 6. Effect of contact time on the sorption of Ni(II) by MOB. Error bars represent ± S.D.
154 D.H.K. Reddy et al. / Desalination 268 (2011) 150–157

Table 2
The pseudo-first-order and pseudo-second-order and Weber and Morris constants.

Initial Ni(II) Experimental Pseudo-first-order Pseduo-second-order Weber and Morris


concentration value qe,exp
qe,cal k1 × 10−2 R2 qe,cal k2 × 10−2 h0 R2 qe,cal Kid R2
(mg L−1) (mg g−1)
(mg g−1) (min–1) (mg g−1) (g mg–1 min–1) (mg g−1 min−1) (mg g−1) (mg g−1 min−1/2)

50 9.70 8.00 7.02 0.9888 10.29 1.91 2.03 0.9971 1.5 0.20 0.9353
25 6.74 4.72 5.61 0.9344 7.14 2.70 1.38 0.9964 2.88 0.44 0.9213
10 3.27 2.16 5.73 0.9233 3.43 12.74 1.05 0.996 4.03 0.67 0.9243

between biosorbent and sorbate, or covalent forces, through the where, qe,m is equilibrium capacity obtained by calculating from
exchange of electrons between the particles involved. model (mg g−1), qe is the experimental data of equilibrium capacity
An empirically found functional relationship, common to the most (mg g−1).
adsorption processes, is that the uptake varies almost proportionally The isotherm constants, correlation coefficient (R2) of these
with t1/2, as per the Weber–Morris plot, rather than with the contact models for sorption of Ni(II) onto MOB for three different tempera-
time, t [24] tures are presented in Tables 3 and 4.
The Langmuir isotherm theory assumes monolayer coverage of
1=2 adsorbate over a homogeneous adsorbent surface [26].
q = Ki t ð5Þ
qm KL ce
qe = ð7Þ
where q (mg g−1) is the adsorbed metal amount, Ki intraparticle 1 + KL ce
diffusion rate constant (mg g−1 min−1/2). According to this model, the
plot of uptake (q) vs the square root of time should be linear if In Eq. (7) qm is the monolayer biosorption capacity of the
intraparticle diffusion is involved in the adsorption process and if biosorbent (mg g−1), qe is the equilibrium metal ion concentration
these lines pass through the origin then intraparticle diffusion is the on the biosorbent (mg g−1), Ce is the equilibrium metal ion
rate-controlling step. The low correlation coefficient values obtained concentration in the solution (mg L−1), and KL is the Langmuir
(Table 2) for the intraparticle diffusion model indicated that biosorption constant (L mg−1) related to the free energy of biosorp-
adsorption is not occurring in the pores of biosorbent accordance tion. The constants qm and KL increased with increase in temperature.
with surface adsorption. The R2 and χ2 (Table 4) values for MOB sorbent indicated that
Langmuir theory describes the sorption phenomena more favorably.
3.9. Biosorption isotherm The essential features of the Langmuir biosorption isotherm can be
expressed in terms of a dimensionless constant separation factor (RL),
The successful representation of the dynamic adsorptive separa- which is defined in Eq. (8).
tion of solute from solution onto an adsorbent depends upon an
appropriate description of the equilibrium separation between two 1
RL = ð8Þ
phases [25]. In order to determine the mechanism of Ni(II) ð1 + KL C0 Þ
biosorption onto MOB and to evaluate the relationship between
biosorption temperatures, the experimental data was applied to the where KL is the Langmuir constant (L mg−1) and C0 is the initial
two-parameter non-linear isotherm models, i.e. Langmuir, Freundlich, adsorbate concentration (mg L−1). The value of RL indicates the type
Dubinin–Radushkevich (D–R) and Temkin. The Chi-square (χ2) test of isotherm [27,28]. The values of RL are all in the range of 0–1, which
was also carried out to find out the best fit adsorption isotherm model. indicate the favorable biosorption of Ni(II) onto MOB.
The equation for evaluating the best fit model is The Freundlich model assumes a heterogeneous sorption surface
[29].
 2
1=n
2
qe −qe;m qe = KF Ce ð9Þ
χ =∑ ð6Þ
qe;m
In Eq. (9) Kf is a constant relating the biosorption capacity and 1/n
35 is an empirical parameter relating the biosorption intensity, which
varies with the heterogeneity of the material. From the graphs, Kf
30 values were found to be 7.62 ± 0.78, 8.59 ± 0.90 and 10.41 ± 1.00 and
50 mg L-1 the n values were found as 3.01 ± 0.30, 3.07 ± 0.33 and 3.41 ± 0.38.
25
25 mg L-1 The 1/n values were between 0 and 1 indicating that the biosorption
t/qt (min g mg-1)

10 mg L-1 of Ni(II) onto the MOB was favorable at studied conditions. The low
value of R2 and high χ2 values indicated that Freundlich model was
20
not altogether properly describing the relationship between the
amounts of sorbed metal ions and their equilibrium concentrations in
15
the solution.
The Dubinin–Radushkevich (D–R) model, which does not assume
10
a homogeneous surface or a constant biosorption potential as the
Langmuir model, was also used to test the experimental data [30].
5
    2 
1
qe = Qm exp −K RT ln 1 + ð10Þ
0 Ce
0 20 40 60 80 100
t (min)
 
2
qe = Qm exp −kε ð11Þ
Fig. 7. Pseudo-second-order kinetics plots for the sorption of Ni(II) onto the MOB.
D.H.K. Reddy et al. / Desalination 268 (2011) 150–157 155

Table 3
Langmuir, Freundlich, Temkin and Dubinin–Radushkevich constants for Ni(II) biosorption by MOB.

Metal Temp. Langmuir Freundlich Dubinin–Radushkevich Temkin


ion (K)
qm (mg/g) KL (L/mg) Kf (mg/g) n Qm K b KT

Ni(II) 303 26.84 ± 0.91 0.26 ± 0.02 7.62 ± 0.78 3.01 ± 0.30 21.82 ± 2.02 0.06 ± 0.01 7.76 ± 0.56 5.26 ± 1.24
313 29.46 ± 0.99 0.27 ± 0.03 8.59 ± 0.90 3.07 ± 0.33 24.12 ± 2.20 0.05 ± 0.01 6.33 ± 0.48 5.56 ± 1.42
323 30.38 ± 0.82 0.31 ± 0.03 10.41 ± 1.00 3.41 ± 0.38 25.13 ± 1.86 0.02 ± 0.01 6.09 ± 0.35 7.43 ± 1.51

Where Qm is the maximum amount of the metal ion that could be Eq. (14) can be written as
sorbed onto unit weight of sorbent (mg g−1), ε is the Polanyi potential
which is equal to RT ln (1 + 1/Ce), where R and T are the universal gas
constant (kJ mol−1 K−1) and the absolute temperature (K), respec- ΔSB ΔHB
ln K = − ð16Þ
tively. The D–R isotherm parameters for three different temperatures R RT
have been presented in Table 3. The values of correlation coefficient
were lower than that of other three isotherm values. In all the cases,
The values of ΔG° were calculated from Eq. (14). Reciprocal of
the D–R equation represents the least fit to experimental data than
temperature (1/T) was plotted against ln k and a straight line was
the other isotherms equations.
obtained as shown in Fig. 8. The values of ΔH° and ΔS° calculated from
The Temkin isotherm [31] assumes that the heat of adsorption of
the slope and intercept of the straight line were summarized in
all the molecules in the layer decreases linearly with coverage due to
Table 5. The negative value of ΔG° indicates the feasibility of the
adsorbent–adsorbate interactions [32], and that the adsorption is
process and indicated the spontaneous nature of the adsorption. The
characterized by a uniform distribution of the binding energies, up to
values of ΔG° were found to increase as temperature increased,
some maximum binding energy.
indicating more driving force and hence resulting in higher biosorp-
RT tion capacity. The value of ΔH° was positive, indicating the
qe = lnðK T Ce Þ ð12Þ endothermic nature of the biosorption of Ni(II) onto MOB. In addition,
b
the positive values of ΔS° shows an affinity of biosorbent and the
qe = B1 lnðK T Ce Þ ð13Þ increasing randomness at the solid-solution interface during the
biosorption of Ni(II) on the bark [34].
Where constant B1 = RT/b, which is related to the heat of
adsorption, R is the universal gas constant (J mol− 1 K−1), T is the 3.11. Biosorption of Ni(II) in presence of other cations
temperature (K), b is the variation of adsorption energy (J mol−1) and
KT is the equilibrium binding constant (L mg−1) corresponding to the For the determination of interference caused by the presence of
maximum binding energy. Temkin isotherm parameters are listed in other cations, Na+ K+, Ca2+ and Mg2+ were added to the Ni(II)
Table 3. solution. These elements are major constituents of saline and hard
By comparing the results, the values of χ2 and the correlation waters and are likely to be encountered in most industrial effluents
coefficients (Tables 3 and 4), it has been found that the Langmuir [10] from which toxic metals are intended to be removed by
isotherm best represented the equilibrium adsorption of Ni(II) onto biosorption. The presence of Na+ and K+ had no antagonistic effect
MOB. on sorption of Ni(II) by MOB biomass as shown in Fig. 9. The effect of
Ca2+ and Mg2+ was only marginal. These observations pointed to a
3.10. Biosorption thermodynamics significant advantage in favor of MOB in an industrial water treatment
plant over the ion exchanger resin system since the binding of Ca2+
Different thermodynamic parameters, such as the standard Gibbs and Mg2+ with these limits their application for the sorption of other
free energy ΔG° (kJ mol−1), the standard enthalpy (ΔH°), and the metals, since particularly both these divalent cations are present in
standard entropy (ΔS°) for biosorption of Ni(II) onto MOB were high concentrations in wastewaters.
calculated using the following equations

ΔGB = −RT ln K ð14Þ 2.4

−1
where K (L g ) an equilibrium constant obtained by multiplying the
Langmuir constants qm and KL [33], R is the universal gas constant 2.2
(8.314 × 10−3 kJ mol−1 K−1) and T is the absolute temperature (K).
The enthalpy (ΔH°) and entropy (ΔS°) parameters were estimated
ln K

from the following equation: 2.0

ΔGB = ΔH B−TΔSB ð15Þ


1.8

Table 4
Chi-square and correlation coefficient for isotherms at different temperatures.
1.6
T Langmuir Freundlich Dubinin– Temkin
(K) Raduskovich
0.00310 0.00315 0.00320 0.00325 0.00330
R2 χ2 R2 χ2 R2 χ2 R2 χ2
1/T (K-1)
303 0.9972 0.9843 0.9507 3.8117 0.9312 13.0593 0.9650 2.7029
313 0.9970 1.1920 0.9453 5.0504 0.9308 15.6236 0.9607 3.6256
Fig. 8. Plot of ln K vs 1/T for the estimation of thermodynamic parameters for biosorption of
323 0.9994 2.0626 0.9395 6.0889 0.9655 16.5379 0.9775 4.2588
Ni(II) onto MOB.
156 D.H.K. Reddy et al. / Desalination 268 (2011) 150–157

Table 5
Thermodynamic parameters of the Ni(II) biosorption on the MOB at different 100
temperatures.

T (K) −ΔG° ΔS° ΔH° R2


(KJ mol−1) (J mol−1 K−1)a (KJ mol−1)a 90

desorption rate %
303 4.89
313 5.39 56.27 12.15 0.9916
323 6.02 80
a
Measured between 303 and 323 K.

70

3.12. Desorption and regeneration studies


60
In order to investigate desorption of metal ion from metal loaded Ni(II)
MOB, the metal-loaded biosorbent was treated with HCl, which has
been reported as an efficient metal desorbent [35–37]. Desorption 50
studies were performed with different hydrochloric acid concentra- 0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35
tions and the results are shown in Fig. 10. From the results of this HCl Concentration (M)
study, with the increasing of hydrochloric acid concentration, the
desorption rate also increased initially, and then become almost Fig. 10. Effect of HCl concentration on desorption of Ni(II) from the MOB. Error bars
represent ± S.D.
stable. The maximum percentage recovery of Ni(II) was 98.02% with
0.2 M HCl solution.
In biosorption process, to keep the processing cost down and to adsorption capacity. The maximum Ni(II) biosorption by MOB at
open the possibility of recovering the metal(s) extracted from the optimum pH 6.0 in this study (30.38 mg g−1) was comparable and
liquid phase, it is desirable to regenerate the biosorbent material. In found to be higher than many other biosorbents. The economical
this study 0.2 M HCl was used as a regenerating agent (or eluent). The success of any biosorption process depends largely on the cost of
regenerated biosorbent was reused for up to six biosorption– biosorbent used. The biosorbent that is used in the present
desorption cycles and the results are illustrated in Fig. 11. An investigation is easily available waste material; hence the cost of the
efficiency of 98.02% recovery of Ni(II) was obtained with 0.2 M HCl biosorbent would be very less compared to ion exchange resins. The
in the first cycle and is therefore suitable for regeneration of easy availability and cost effectiveness of MOB biomass are additional
biosorbent. There is a gradual decrease in Ni(II) biosorption with an advantages, which make it better biosorbent for treatment of nickel
increase in the number of cycles. After a sequence of six cycles, the Ni waste waters.
(II) uptake capacity of the biosorbent was reduced from 98.02% to
92.02%. The lost in the biosorption capacity of the biomass for metal
ions was found to be b6%. This might be due to the ignorable amount 4. Conclusion
of biomass lost during the biosorption–desorption process. These
results indicate that the MOB could be used repeatedly in Ni(II) The experimental investigation concluded that MOB could be used
biosorption studies without any detectable loss in the total biosorp- as potential sorbent, for the removal of Ni(II) from aqueous solution.
tion capacity. The batch study parameters; pH of solution, biomass concentration,
contact time, and temperature were found to be effective on the
3.13. Comparison of nickel removal with different biosorbents biosorption processes. The kinetic studies revealed that the biosorp-
tion process followed the pseudo-second-order kinetic model. The
The biosorption capacity of the biosorbent MOB for the removal of maximum biosorption capacity of Ni(II) was 30. 38 mg g−1 at an
nickel has been compared with those of other biosorbents reported in optimum pH 6.0. The calculated thermodynamic parameters showed
literature [38–48] and the values of adsorption capacities (qm) have the feasibility, endothermic and spontaneous nature of the biosorp-
been presented in Table 6. The values are based on Langmuir tion of Ni(II) onto MOB biomass. The reusability of the biosorbent was

100
Adsorption Desorption
Na+ 100
K+
Adsorption/Desorption (%)

95 Ca2+
Mg2+
Ni(II) Removal (%)

80

60
90

40

85
20

80 0
0 50 100 150 200 250 300 1 2 3 4 5 6
Concentration (μg mL-1) Cycle number

Fig. 9. Effect of other metal ions on the sorption of Ni(II) by MOB. Error bars represent±S.D. Fig. 11. Six cycles of Ni(II) adsorption–desorption with 0.2 M HCl. Error bars represent±S.D.
D.H.K. Reddy et al. / Desalination 268 (2011) 150–157 157

Table 6 [15] C.P. Neto, J. Rocha, A. Gil, N. Cordeiro, A.P. Esculcas, S. Rocha, I. Delgadillo, J.D.
The maximum biosorption capacities of various biosorbents. Pedrosa de Jesus, A.J.F. Correia, 13 C solid-state nuclear magnetic resonance and
Fourier transform infrared studies of the thermal decomposition of cork, Solid
Biosorbent Biosorption pH Reference State Nucl. Magn. Reson. 4 (1995) 143–151.
capacity (mg/g) [16] R. Kellner, J.M. Mermet, M. Otto, Analytical Chemsitry, WILEY-VHC Verlag Gmbh
Press, New York, 1998, p. 824.
Dye groundnut shells 7.49 5.16 38 [17] R.A. Anayurt, A. Sari, M. Tuzen, Equilibrium, thermodynamic and kinetic
Dye loaded sawdust 9.87 5.16 38 studies on biosorption of Pb(II) and Cd(II) from aqueous solution by
Tea factory waste 18.42 4.0 39 macrofungus (Lactarius scrobiculatus) biomass, Chem. Eng. J. 151 (2009)
Cone biomass 12.42 4.0 40 255–261.
of T. orientalis [18] D. Lua, Q. Caob, X. Caoa, F. Luoa, Removal of Pb(II) using the modified lawny grass:
Baker's yeast 11.40 6.75 41 mechanism, kinetics, equilibrium and thermodynamic studies, J. Hazard. Mater.
Barley straw 35.8 4.85 42 166 (2009) 239–247.
untreated [19] H. Pahlavanzadeh, A.R. Keshtkar, J. Safdari, Z. Abadi, Biosorption of nickel(II) from
aqueous solution by brown algae: equilibrium, dynamic and thermodynamic
Bagasse fly ash 6.49 6.0 43
studies, J. Hazard. Mater. 175 (2010) 304–310.
Irish peat moss 14.5 4.7 44
[20] N. Fiol, I. Villascusa, M. Martinez, N. Mirralles, J. Poch, J. Seralos, Sorption of Pb(II),
Coir pith 9.5 3.92 45
Ni(II), Cu(II) and Cd(II) from aqueous solution by olive stone waste, Sep. Purif.
Modified coir pith 38.9 3.92 45 Technol. 50 (2006) 132–140.
Biomatrix from 5.4 6.0 46 [21] X. Li, Y. Tang, Z. Xuan, Y. Liu, F. Luo, Study on the preparation of orange peel
rice husk cellulose adsorbents and biosorption of Cd2+ from aqueous solution, Sep. Purif.
Black carrot residues 6.51 5.25 47 Technol. 55 (2007) 69–75.
Waste pomace of olive 14.80 4.0 48 [22] S. Lagergren, About the theory of so-called adsorption of soluble substances, K.
oil factory (WPOOF) Sven. Vetenskapsakad. Handl. 24 (1898) 1–39.
MOB 30.38 6.0 This study [23] Y.S. Ho, G. McKay, Pseudo-second order model for sorption processes, Process
Biochem. 34 (1999) 451–465.
[24] W.J. Weber, J.C. Morris, Kinetics of adsorption on carbon from solution, J. Sanit.
Eng. Div. A. S. C. E. 89 (1963) 31–59.
[25] J. Febrianto, A.N. Kosasih, J. Sunarso, Y. Jua, N. Indraswati, S. Ismadji, Equilibrium
good after six consecutive biosorption–desorption cycles without any and kinetic studies in adsorption of heavy metals using biosorbent: a summary of
considerable loss in biosorption capacity based on all results, the MOB recent studies, J. Hazard. Mater. 162 (2009) 616–645.
[26] I. Langmuir, The adsorption of gases on plane surfaces of glass, mica and platinum,
biomass can be used as alternative biosorbent for treatment of waste J. Am. Chem. Soc. 40 (1918) 1361–1403.
waters containing Ni(II) ions. [27] E. Malkoc, Y. Nuhoglu, Potential of tea factory waste for chromium(VI) removal
from aqueous solutions: thermodynamic and kinetic studies, Sep. Purif. Tehnol. 54
(2007) 291–298.
Acknowledgments [28] A. Ozer, D. Ozer, A. Ozer, The adsorption of copper(II) ions on to dehydrated wheat
bran (DWB): determination of the equilibrium and thermodynamic parameters,
This work was financially supported by the Board of Research in Process Biochem. 39 (2004) 2183–2191.
[29] H. Freundlich, Adsorption in solution, W.J. Helle, J. Am. Chem. Soc. 61 (1939) 2–28.
Nuclear Sciences, Mumbai under Project BRNS No: 2007/37/48/BRNS/ [30] M.M. Dubinin, L.V. Radushkevich, Equation of the characteristic curve of activated
2919. One of the authors, D. H. K. Reddy thanks BRNS for his Junior charcoal, Proc. Acad. Sci. USSR 55 (1947) 331–333.
Research Fellowship. [31] M.I. Temkin, V. Pyzhev, Kinetics of ammonia synthesis on promoted iron catalysts,
Acta Physiochim. URSS 12 (1940) 327–356.
[32] M. Hosseine, S.F.L. Mertens, M. Ghorbani, M.R. Arshadi, Asymmetrical Schiff bases
References as inhibitors of mild steel corrosion in sulphuric acid media, Mater. Chem. Phys. 78
(2003) 800–807.
[1] A. Ozer, G. Gurbuz, A. Calimli, B.K. Korbahti, Investigation of nickel(II) biosorption [33] Z.-Y. Yao, J.-H. Qi, L.-H. Wang, Equilibrium, kinetic and thermodynamic studies on the
on Enteromorpha prolifera: optimization using response surface analysis, J. Hazard. biosorption of Cu(II) onto chestnut shell, J. Hazard. Mater. 174 (2010) 137–143.
Mater. 152 (2008) 778–788. [34] Q. Li, L. Chai, Z. Yang, Q. Wang, Kinetics and thermodynamics of Pb(II) adsorption onto
[2] V.K. Gupta, A. Rastogi, A. Nayak, Biosorption of nickel onto treated alga modified spent grain from aqueous solutions, Appl. Surf. Sci. 255 (2009) 4298–4303.
(Oedogonium hatei): application of isotherm and kinetic models, J. Colloid [35] Z.R. Holan, B. Volesky, I. Prasetyo, Biosorption of cadmium by biomass of marine
Interface Sci. 342 (2010) 533–539. algae, Biotechnol. Bioeng. 41 (1993) 819–825.
[3] B. Volesky, Biosorption and me, Water Res. 41 (2007) 4017–4029. [36] A. Saeed, M. Iqbal, Bioremoval of cadmium from aqueous solution by black gram
[4] P. King, K. Anuradha, S.B. Lahari, Y.P. Kumar, V.S.R.K. Prasad, Biosorption of zinc husk (Cicer arientinum), Water Res. 37 (2003) 3472–3480.
from aqueous solution using Azadirachta indica bark: equilibrium and kinetic [37] P.R. Puranik, K.M. Paknikar, Biosorption of lead and zinc from solutions using
studies, J. Hazard. Mater. 152 (2008) 324–329. Streptoverticillium cinnamoneum waste biomass, J. Biotechnol. 55 (1997) 113–124.
[5] E.M. Saad, R.A. Mansour, A. El-Asmy, M.S. El-Shahawi, Sorption profile and [38] S.R. Shukla, R.S. Pai, Adsorption of Cu(II), Ni(II) and Zn(II) on dye loaded
chromatographic separation of uranium (VI) ions from aqueous solutions onto groundnut shells and sawdust, Sep. Purif. Technol. 43 (2005) 1–8.
date pits solid sorbent, Talanta 76 (2008) 1041–1046. [39] E. Malkoc, Y. Nuhoglu, Investigations of nickel(II) removal from aqueous solutions
[6] I. Ghodbane, O. Hamdaour, Removal of mercury(II) from aqueous media using using tea factory waste, J. Hazard. Mater. 127 (2005) 120–128.
eucalyptus bark: kinetic and equilibrium studies, J. Hazard. Mater. 160 (2008) [40] E. Malkoc, Ni(II) removal from aqueous solutions using cone biomass of Thuja
301–309. orientalis, J. Hazard. Mater. 137 (2006) 899–908.
[7] S. Schiewer, S.B. Patil, Pectin-rich fruit wastes as biosorbents for heavy metal [41] V. Padmavathy, P. Vasudevan, S.C. Dhingra, Biosorption of nickel(II) ions on
removal: equilibrium and kinetics, Biores. Technol. 99 (2008) 1896–1903. Baker's yeast, Process Biochem. 38 (2003) 1389–1395.
[8] E. Pehlivan, B.H. Yank, G. Ahmetli, M. Pehlivan, Equilibrium isotherm studies for [42] A. Thevannan, R. Mungroo, C.H. Niu, Biosorption of nickel with barley straw,
the uptake of cadmium and lead ions onto sugar beet pulp, Biores. Technol. 99 Biores. Technol. 101 (2010) 1776–1780.
(2008) 3520–3527. [43] V.C. Srivastava, I.D. Mall, I.M. Mishra, Equilibrium modelling of single and binary
[9] A. Bhatnagar, A.K. Minocha, Biosorption optimization of nickel removal from adsorption of cadmium and nickel onto bagasse fly ash, Chem. Eng. J. 117 (2006) 79–91.
water using Punica granatum peel waste, Colloids Surf. B Biointerfaces 76 (2010) [44] B.S. Gupta, M. Curran, S. Hasan, T.K. Ghosh, Adsorption characteristics of Cu and Ni
544–548. on Irish peat moss, J. Environ. Manage. 90 (2009) 954–960.
[10] A. Gundogdu, D. Ozdes, C. Duran, V.N. Bulut, M. Soylak, H.B. Senturk, Biosorption [45] A. Ewecharoen, P. Thiravetyan, W. Nakbanpote, Comparison of nickel adsorption
of Pb(II) ions from aqueous solution by pine bark (Pinus brutia Ten.), Chem. Eng. J. from electroplating rinse water by coir pith and modified coir pith, Chem. Eng. J.
153 (2009) 62–69. 137 (2008) 181–188.
[11] O.S. Lawl, A.R. Sanni, I.A. Ajayi, O.O. Rabiu, Equilibrium, thermodynamic and [46] K.K. Krishnani, X. Mengb, C. Christodoulatos, V.M. Boddu, Biosorption mechanism
kinetic studies for the biosorption of aqueous lead(II) ions onto the seed husk of of nine different heavy metals onto biomatrix from rice husk, J. Hazard. Mater. 153
Calophyllum inophyllum, J. Hazard. Mater. 177 (2010) 829–835. (2008) 1222–1234.
[12] P.-H. Chuang, C.-W. Lee, J.-Y. Chou, M. Murugan, B.-J. Shieh, H.-M. Chen, Anti- [47] F. Guzel, H. Yakut, G. Topal, Determination of kinetic and equilibrium parameters
fungal activity of crude extracts and essential oil of Moringa oleifera Lam, Biores. of the batch adsorption of Mn(II), Co(II), Ni(II) and Cu(II) from aqueous solution
Technol. 98 (2007) 232–236. by black carrot (Daucus carota L.) residues, J. Hazard. Mater. 153 (2008)
[13] R.N. Bennett, F.A. Mellon, N. Foidl, J.H. Pratt, M.S. Dupont, L. Perkins, P.A. Kroon, 1275–1287.
Profiling Glucosinolates and phenolics in vegetative and reproductive tissues of [48] Y. Nuhoglu, E. Malkoc, Thermodynamic and kinetic studies for environmentaly
the multi-purpose trees Moringa oleifera L. (Horseradish Tree) and Moringa friendly Ni(II) biosorption using waste pomace of olive oil factory, Biores. Technol.
stenopetala L. J. Agric. Food Chem. 51 (2003) 3546–3553. 100 (2009) 2375–2380.
[14] D.L. Pavia, G.M. Lampman, G.K. Kriz Jr., Introduction to spectroscopy, Saunders
Golden Sunburst Series. 1979.

Вам также может понравиться