You are on page 1of 9

This article appeared in a journal published by Elsevier.

The attached
copy is furnished to the author for internal non-commercial research
and education use, including for instruction at the authors institution
and sharing with colleagues.
Other uses, including reproduction and distribution, or selling or
licensing copies, or posting to personal, institutional or third party
websites are prohibited.
In most cases authors are permitted to post their version of the
article (e.g. in Word or Tex form) to their personal website or
institutional repository. Authors requiring further information
regarding Elsevier’s archiving and manuscript policies are
encouraged to visit:
http://www.elsevier.com/copyright
Author's personal copy

Journal of Hazardous Materials 173 (2010) 750–757

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

Reduction of COD in refinery wastewater through adsorption on date-pit


activated carbon
Muftah H. El-Naas ∗ , Sulaiman Al-Zuhair, Manal Abu Alhaija
Chemical and Petroleum Engineering Department, United Arab Emirates University, 17555 Al-Ain, United Arab Emirates

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

Article history: Experiments were carried out to evaluate the batch adsorption of COD from petroleum refinery wastewa-
Received 6 May 2009 ter on a locally prepared date-pit activated carbon (DP-AC), and its adsorption effectiveness was compared
Received in revised form 31 August 2009 to that of commercially available BDH activated carbon (BDH-AC). Adsorption equilibrium and kinetic
Accepted 1 September 2009
data were determined for both adsorbents and fitted to several adsorption isotherm and kinetics models,
Available online 4 September 2009
respectively. The Langmuir monolayer isotherm fitted well the equilibrium data of COD on both adsor-
bents; whereas, the kinetics data were best fitted by the pseudo-second order model. Modeling of the
Keywords:
controlling mechanisms indicated that both intrinsic kinetics and mass transfer contributed to controlling
COD reduction
Petroleum refinery
the adsorption process. Mass transfer seemed to be the dominant mechanism at low COD content, while
Date-pit intrinsic kinetics dominates at high concentrations. In general, the adsorption effectiveness of locally
Controlling mechanism prepared DP-AC was proven to be comparable to that of BDH-AC. Therefore, DP-AC can be utilized as an
Adsorption effective and less expensive adsorbent for the reduction of COD in refinery wastewater.
© 2009 Elsevier B.V. All rights reserved.

1. Introduction industrial wastewater. Mohan et al. [4] investigated the adsorption


effectiveness of AC from different raw materials for COD removal,
Various treatment technologies have been used for the reduc- and reported reduction percentages of up to 74% using AC from
tion of COD, which is a major contaminant in petroleum refinery coconut shell fiber and rice husk. Recently, date-pits (DP) have
and industrial wastewater. These techniques include filtration, ion received considerable attention as a lignin origin for preparing
exchange, coagulation/flocculation, reverse osmosis and electro- AC. In addition to producing AC from inexpensive raw material,
dialysis. Adsorption provides an attractive alternative treatment, using DP is considered a waste minimization process. DP consti-
especially if the adsorbent is inexpensive and readily available. tute approximately 10% of the total weight of dates [5–6], making
Adsorption of COD has been previously studied to evaluate the them the largest agricultural by-product in palm growing countries,
overall adsorption behavior in wastewaters [1]. Considerable effort including the UAE. Therefore, finding ways to use this agricultural
has been devoted to the production of low cost adsorbents using by-product profitably will benefit date farmers substantially and
less expensive and readily available materials [2–4]. The use of acti- offers a useful alternative for its disposal or current use as animal
vated carbon (AC), as an absorbent, has proven to be effective in a feed [6].
wide range of applications, including the removal of both organic Several studies have examined different activation processes
and inorganic pollutants from wastewater [5]. The versatility of of DP and the possibility of using the produced AC in wastewa-
high surface area, porous structure and surface adsorption capac- ter adsorption treatment. DP-AC was prepared by physical [7–8]
ity, which can be appropriately modified by physical and chemical and chemical activation [9]. In addition, physical activation of DP
treatments, are among the reasons for the use of such adsorbent. has been carried out in different reactor configurations including
However, the cost of commercially available AC is relatively high fluidized bed [10]. The adsorption effectiveness of DP-AC of alu-
[2], and therefore great attention has been focused on the pro- minum was evaluated and compared to that of commercial BDH-AC
duction of low-cost AC of properties comparable to those of the [11]. Generally, the effectiveness of DP-AC was found to be compa-
commercially available. In order to achieve this target, agricultural rable to that of BDH-AC, and was even preferable under certain
waste was considered as a promising raw material for generat- conditions, such as low Al concentrations and low pH.
ing AC capable of removing various pollutants including COD from The objective of the present work is to explore the viability
of using DP-AC, locally prepared by physical activation, as a low
cost adsorbent for the reduction of COD from petroleum refinery
∗ Corresponding author. Tel.: +971 3 713 3637; fax: +971 3 762 4262. wastewater. Equilibrium and kinetics studies were carried out to
E-mail address: muftah@uaeu.ac.ae (M.H. El-Naas). determine the mechanisms of the adsorption process.

0304-3894/$ – see front matter © 2009 Elsevier B.V. All rights reserved.
doi:10.1016/j.jhazmat.2009.09.002
Author's personal copy

M.H. El-Naas et al. / Journal of Hazardous Materials 173 (2010) 750–757 751

1.1. Adsorption isotherms where qD is the D–R isotherm constant related to the degree of
sorbate sorption by the sorbent surface (mg/g) and BD is related to
Adsorption is a process that results in the removal of a solute the free energy of sorption per mole of sorbate as it migrates to
from a solution and concentrating it at the surface of the adsorbent, the surface of the adsorbent from infinite distance in the solution
until the amount of the solute remaining in the solution is in equi- (mol2 /kJ2 ). This means free energy can be estimated from the value
librium with that at the surface. This equilibrium is described by of BD using the following equation:
expressing the amount of solute adsorbed per unit mass of adsor-
1
bent, qe , as a function of the concentration of solute remaining E=  (7)
in solution, Ce . An expression of this type is termed adsorption 2BD
isotherm.
Many theoretical and empirical models have been developed The above mentioned isotherms are characterized by parame-
to represent the various types of adsorption isotherms. At present, ters that express the surface properties and affinity of the adsorbent
there is no single model that satisfactory describes all mechanisms towards different pollutants. Different isotherm models can be
and shapes. Langmuir and Freundlich equations are examples used to fit the equilibrium data. The shapes of various model
of such models that are commonly used to describe adsorption isotherms depend on the type of adsorbate/adsorbent and the inter-
isotherms in water and wastewater treatment applications. The molecular interactions between the fluid and the surface [1]. The
Langmuir isotherm assumes uniform and constant binding of the model that fits the experimental data most accurately can then be
sorbate on the surface of the adsorbent, which is usually described used to describe the system and predict the adsorption behavior
by: for practical process design.

qm bCe
qe = (1) 1.2. Adsorption kinetics
1 + bCe
where qe is the equilibrium amount of solute adsorbed (mg/g of The adsorption kinetics can be described by diffusion through
solid), Ce is the equilibrium concentration of solute in solution the adsorbent, such as film, pore and surface diffusions, and pore-
(mg/L), and qm (mg/g) and b (mg/L)−1 are constants, representing surface adsorption or any combination of these four steps. In order
the maximum adsorption capacity for the solid phase loading and to examine the controlling mechanism of an adsorption process
the energy constant related to the heat of adsorption, respectively. and to determine the minimum necessary time to achieve equi-
The parameters qm and b are temperature dependent parameters. A librium, several kinetics models such as the pseudo-first order,
convenient way of expressing this temperature dependence takes Elovich’s, pseudo-second order and intraparticle diffusion models
advantage of Van’t Hoff’s relationship between the equilibrium can be used. Brief descriptions of these models are given in the
constant and the standard enthalpy change associated with the following sections.
process under consideration, which leads to the following relations:

qm = ˛(1 + ε T ) (2) 1.2.1. Lagergren pseudo-first order model


 −H  The Lagergren pseudo-first order model is most commonly used
ads to describe the adsorption of solute from a liquid solution. The
b = ˇ exp (3)
RT linearized form of this model is given by:
where Hads is the standard enthalpy change accompanying the
kt
reversible adsorption (kJ mol−1 ), R is the universal gas constant, T ln (qe − q) = ln (qe ) − (8)
2.303
is the absolute temperature and ˛, ˇ and ε are constants. The con-
stants found in Eqs. (2) and (3) can be determined experimentally where k is the kinetics constant of pseudo-first order adsorption
by measuring the equilibrium adsorption at different temperatures. (min−1 ), qe and q (mg/g) are the amounts adsorbed at equilibrium
Unlike the Langmuir isotherm model, the Freundlich isotherm and at time t (min), respectively.
(Eq. (4)), which has been widely used in correlating equilibrium
data, does not have any thermodynamic basis and does not offer 1.2.2. Elovich’s model
much physical interpretation of the adsorption data [12]. This In recent years, Elovich’s model has been successfully used to
model is not bound by a maximum uptake, and it does not approach describe the adsorption of pollutants from aqueous solutions. The
Henry’s law at low concentrations. linearized form of this model is given by Eq. (9):
The Freundlich isotherm can be written as:
1/b 1 1
qe = aCe (4) q= ln (ab) + ln (t) (9)
b b
where, a and b are constants where, a is the initial adsorption rate (mg/g/min), and 1/b (mg/g) is
The Sips isotherm combines the Langmuir and Freundlich a parameter related to the number of sites available for adsorption.
isotherms into one equation. At low sorbent concentration, the
Sips isotherm approaches the Freundlich isotherm, whereas it
approaches the Langmuir isotherm at a high concentration. This 1.2.3. Pseudo-second order model
model is presented in the Langmuir–Freundlich form as follows: In this model, the rate limiting step is the surface adsorption
that involves chemisorption, where the adsorbate removal from
KLF CenLF a solution is due to physicochemical interactions between the two
qe = (5)
1 + (aLF Ce)nLF phases [13]. The model is usually represented by its linearized form
where, KLF , nLF and aLF are constants. as follows:
Another isotherm that has seen considerable applications in t 1 1
adsorption processes is the Dubinin–Radushevich: = + t (10)
q K2 q2t qe
   2 
1 where, K2 (g/mg/min) is the pseudo-second order rate constant of
qe = qD exp −BD RT ln 1+ (6)
Ce adsorption.
Author's personal copy

752 M.H. El-Naas et al. / Journal of Hazardous Materials 173 (2010) 750–757

Table 1 concentrations as follows:


Surface characteristics of BDH and date-pit (DP) activated carbons.
(Ci − Cf )V
Characteristics BDH DP q= (12)
M
Particle size (␮m) 850–1700 125–212
BET surface area (m2 /g) 1220 490
where, q is the uptake (mg adsorbate/g adsorbent), Ci is the initial
BET surface area in the 1173 423 COD concentration, Cf is the final COD (mg/L), M is the adsorbent
micropore region (m2 /g) amount (g) and V is the solution volume (l).
Total pore volume (cm3 /kg) 534 229
Total pore volume in the 479 165
micropore region (cm3 /kg) 2.3. Regeneration

Regeneration is an important aspect in evaluating the capac-


ity and practicality of any adsorbent. The regeneration of activated
1.2.4. Intraparticle diffusion model carbon has been meticulously examined in the literature and many
The intraparticle diffusion model describes adsorption pro- researchers have successfully regenerated activated carbon using
cesses, where the rate of adsorption depends on the speed at which different methods. These included the utilization of pyrolysis [14],
adsorbate diffuses towards adsorbent (i.e., the process is diffusion- steam [15], water under sub-critical conditions [16], surfactants
controlled), which is presented by Eq. (11). [17], direct ozonation [18], microwave [19], electrochemical meth-
√ ods [20,21] wet peroxide oxidization [22] bio-regeneration [23],
q = kd t +  (11) and ultrasound [24]. Although the results depended on the applica-
tion and the type of wastewater treated, in most cases the activated
where, kd is the rate constant of the intraparticle transport carbon was fully regenerated and reused for many cycles. These
(mg/g/min1/2 ) and  (mg/g) is a constant related to the thickness studies have confirmed that activated carbon can be easily regener-
of the boundary layer. High values of  indicate greater boundary ated and that the cost associated with the regeneration process can
layer effect. be substantially reduced. In the present study, the regeneration of
activated carbon was not addressed and will be thoroughly exam-
2. Experimental methods ined in a future study that evaluates the continuous adsorption of
COD in a packed bed column.
2.1. Preparation of DP-AC
3. Results and discussion
DP-AC was prepared from raw DP granules, obtained from
Al-Saad Date Processing Factory, Al-Ain, UAE. The granules were 3.1. Effect of adsorbent dose
washed, dried, grinded and screened. The collected granules were
carbonated and activated to produce DP-AC. The carbonization is The amount of adsorbent, which is usually referred to as the
performed in a tube furnace (Thermolyene, USA) which has been adsorbent dose (grams of adsorbent per 1 L of solution) plays an
initially purged with a flow of nitrogen for 10 min. After that, the important role in the adsorption process. The effect of the adsorbent
furnace was heated at a rate of 10 ◦ C/min up to 600 ◦ C and then kept dose on COD removal for the two adsorbents (DP-AC and BDH-
at this temperature for 4 h. After cooling to room temperature, the AC) was evaluated by varying the dose from 1–80 g/L. The results
material is considered carbonized, but still inactive. After weighing are presented in Fig. 1 and indicate that higher percent removal
the inactive carbon, it was activated in the same tube furnace at is achieved for higher adsorbent dose, but the uptake decreases
a temperature of 900 ◦ C using a flow of carbon dioxide instead of with increasing the amount of adsorbent. This may be attributed
nitrogen. The resulting AC was then degassed under vacuum (Shel to many factors such as availability of solute, interference between
Lab, USA) for about 2 h before use. The main surface characteristics binding sites, electrostatic interactions, and reduced mixing due to
of the prepared DP-AC and commercially available DBH-AC were high adsorbent concentration in the solution [25–26]. Many of the
determined using a surface area analyzer (Micromeritics, Model adsorption sites, therefore, remain unsaturated due to the limited
ASAP-2010) and presented in Table 1.

2.2. COD batch adsorption

Batch adsorption equilibrium experiments were carried out by


contacting a specified amount of adsorbent with 50 ml wastew-
ater sample, of a known initial COD concentration, in a sealed
glass bottle. The bottle was kept on a shaker (WSB-30, Korea) at
different temperatures for 24 h to reach equilibrium. At regular
intervals, samples were withdrawn and the COD was measured
using UV Spectrophotometer (DR-5000, Germany). The COD uptake
was determined as a function of time, using 20 g/L DP-AC and BDH-
AC, contacted with refinery wastewater samples obtained from a
local petroleum refinery. Wastewater samples of three different ini-
tial concentrations, namely, 3490, 1662 and 950 (±0.5) mg/L were
tested. In addition, the experiment was carried at different values
of temperature and pH to determine their effects.
After 24 h, the samples were filtered and the final COD was
determined. Adsorption efficiencies of the prepared DP-AC were
compared to those of commercial BDH-AC. The uptake, q, was then Fig. 1. Effect of adsorbent dose on COD uptake at initial COD concentration of
calculated from the difference between the initial and the final COD 3490 mg/L and 25 ◦ C.
Author's personal copy

M.H. El-Naas et al. / Journal of Hazardous Materials 173 (2010) 750–757 753

Table 2
Fitted kinetic parameters for the adsorption of COD onto DP-AC and BDH-AC at 25 ◦ C.

Co (mg/L) Pseudo-first order Elovich’s equation Pseudo-second order


−1 2 2
qe (mg/g) k (min ) R a (mg/g/min) 1/b (mg/g) R K (g/mg.min) qe (mg/g) R2

BDH 3490 31.37 0.0122 0.9497 7.7 × 104 10.53 0.93 0.0016 151.52 1.00
DP 12.38 0.0299 0.8658 7.2 × 1016 3.40 0.87 0.0069 147.06 1.00

BDH 1662 19.33 0.0188 0.7716 8.9 × 104 5.03 0.94 0.0051 74.63 1.00
DP 27.54 0.0258 0.8975 294.4 8.68 0.91 0.0029 73.53 1.00

BDH 950 18.24 0.0295 0.9088 15.3 × 106 5.18 0.87 0.0054 42.74 1.00
DP 7.81 0.0208 0.7426 151.3 2.05 0.91 0.0145 42.55 1.00

availability of the sorbate, which in turn lowers the uptake and the Elovich’s model to describe the adsorption mechanism. The best
adsorption efficiency [27]. The figure also shows that increasing straight lines that pass through the data points were used to deter-
the dose beyond 20 g/L had little effect on the COD reduction and mine the kinetics parameters of both models, which are shown in
hence this value was considered as the optimum dose and used for Table 2. The deviation of the values of R2 from unity is a measure
the rest of the experimental work. of the incompatibility of each respective case.
On the other hand, when t/q was plotted against t, according to
3.2. Adsorption kinetics the linearized pseudo-second order model (Eq. (10)), a clear lin-
ear relationship was observed for all cases, as shown in Fig. 5. The
The COD adsorption rate was determined for the two adsorbents data were fitted perfectly well by straight lines, with R2 values of
by contacting refinery wastewater samples with different initial 1.00. This proves that the adsorption kinetics is more accurately
COD contents (3490, 1662 and 950 mg/L) using an adsorbent dose
of 20 g/L. The results indicate that most of the COD reduction takes
place during the first 30 min for both absorbents as shown in Fig. 2.
After that the COD contents remained almost unchanged, which is
assumed to be the equilibrium amount. The adsorption rate data
were fitted to the different kinetic models described in Section
1.2, namely: Lagergren’s pseudo-first order (Eq. (8)), Elovich’s (Eq.
(9)), pseudo-second order (Eq. (10)) and intraparticle diffusion (Eq.
(11)). Figs. 3–6 present the applicability of the four models at dif-
ferent initial COD, respectively.
Fig. 3 shows the relationship between ln(qe − q) and t, according
to the linearized Lagergren’s pseudo-first order equation (Eq. (8)).
It is observed that the results deviate significantly from a straight
line, which indicates that the pore diffusion is not the sole rate con-
trolling step, as the Lagergren’s pseudo-first order model suggests.
The adsorption data shows a curvature in the initial period, which
is attributed to the intraparticle diffusion or external mass transfer
effects. Similar results were found by plotting q vs. ln(t) accord-
ing to the linearized Elovich’s equation (Eq. (9)), shown in Fig. 4
for both adsorbents, which also proves the incompatibility of the
Fig. 3. ln (qe − q) vs. t according to pseudo-first order model. (䊉) C0 = 3490 mg/L, ()
C0 = 1662 mg/L, () C0 = 950 mg/L, (black) BDH-AC and (white) DP-AC.

Fig. 2. Kinetics of COD uptake using 20 g/L adsorbents at 25 ◦ C and different ini-
tial COD contents. (䊉) C0 = 3490 mg/L, () C0 = 1662 mg/L, () C0 = 950 mg/L, (black) Fig. 4. q vs. ln(t) according to Elovich’s model. (䊉) C0 = 3490 mg/L, ()
BDH-AC and (white) DP-AC. C0 = 1662 mg/L, () C0 = 950 mg/L, (black) BDH-AC and (white) DP-AC.
Author's personal copy

754 M.H. El-Naas et al. / Journal of Hazardous Materials 173 (2010) 750–757

Fig. 5. t/q vs. t according to the pseudo-second order model. (䊉) C0 = 3490 mg/L, ()
C0 = 1662 mg/L, () C0 = 950 mg/L, (black) BDH-AC and (white) DP-AC.
Fig. 7. Langmuir fitting for the equilibrium isotherm data of the COD adsorption on
described by pseudo-second order model. The kinetics parameters DP-AC at different temperatures.

of the model are shown in Table 2. √


To further confirm this conclusion, q was plotted against t, the experimental data were best fit by the Langmuir model with R2
which would, according to Eq. (11), give a straight line if intra- values of almost 1.0.
particle diffusion was the limiting process. However, the results
in Fig. 6 show that this is not the case. The early stage sharper 3.4. Effect of temperature and pH
portion is assumed to be due to the mass transfer external resis-
tance; whereas the later linear portion is an indication of some As shown in Table 3, the adsorption isotherms of COD from
intraparticle diffusion control. wastewater were determined at different temperatures of 25,
40 and 60 ◦ C. The results show that increasing the temperature
3.3. Adsorption isotherms resulted in increasing adsorption capacity. Increasing the temper-
ature from 25 to 60 ◦ C, resulted in improving the uptake capacity
The equilibrium adsorption isotherms of COD were determined by 30% and 53% for BDH-AC and DP-AC, respectively. This is due to
at different temperatures of 25, 40 and 60 ◦ C, and the results are the increase in the kinetic energy of the adsorbate with tempera-
shown in Figs. 7 and 8 for DP-AC and BDH-AC, respectively. The ture, which enhances the adsorbate availability at the active sites of
experimental data were fitted to the Langmuir, Freundlich, Sips the adsorbent. In addition, raising the solution temperature results
and Dubinin–Radushevich isotherm models using Sigma Plot non- in expansion of the pores within the adsorbent particles, which in
linear regression, which uses the Marquardt–Levenberg algorithm turn enhances the adsorption capacity.
to find the parameters that give the best fit between a set of data The value of standard enthalpy change accompanying the
and a proposed non-linear equation. Values for the determined reversible adsorption, Hads and the constants ˛, ˇ and ε found
parameters of each isotherm are shown in Table 3, together with in Eqs. (2) and (3) were determined by fitting the values of qm and
the respective R2 value for each regression. The results show that b at different temperatures. The values of Hads for BDH-AC and
DP-AC were found to be 5.48 and 7.36 kJ/mol which confirms the


Fig. 6. q vs. t according to the intraparticle diffusion model. (䊉) C0 = 3490 mg/L, Fig. 8. Langmuir fitting for the equilibrium isotherm data of the COD adsorption on
() C0 = 1662 mg/L, () C0 = 950 mg/L, (black) BDH-AC and (white) DP-AC. BDH-AC at different temperatures.
Author's personal copy

M.H. El-Naas et al. / Journal of Hazardous Materials 173 (2010) 750–757 755

Table 3
Isotherm model parameters for the adsorption of COD on DP-AC and BDH-AC at different temperatures.

Isotherm Parameter 25 ◦ C 40 ◦ C 60 ◦ C

BDH DP BDH DP BDH DP

Langmuir b (L/mg) 1.62 × 10−3 2.07 × 10−3 1.89 × 10−3 2.4 × 10−3 2.05 × 10−3 2.85 × 10−3
qm (mg/g) 252.81 191.58 293.65 217.82 331.19 241.45
R2 0.99 1.00 0.93 0.99 0.99 0.99

Freundlich aF 1.518 2.651 1.304 1.79 0.989 1.59


bF 0.716 0.6022 0.763 0.665 1.006 0.699
R2 0.98 0.99 0.92 0.97 1.00 0.98

Sips (L–F) nLF 1.489 0.902 2.605 1.228 1.312 1.187


aLF 0.005 0.002 0.009 0.0036 0.0041 0.0035
KLF 0.0623 0.833 5.8 × 10−4 0.169 0.147 0.226
R2 0.93 0.94 0.93 0.93 0.93 0.96

D–R qD (mg/g) 128.49 105.3 131.01 102.8 124.66 113.94


BD 0.0021 1.706 × 10−3 0.0011 1.605 × 10−3 0.001 1.1 × 10−3
E (J/mol) 15.43 17.12 21.32 17.65 22.36 21.32
R2 0.76 0.81 0.76 0.85 0.76 0.83

endothermic nature of the adsorption. The enthalpies of adsorption on the characteristics of the sorbent as well as the sorbate. Ion
on both adsorbent were in the same order of magnitude, which is exchange is expected to be less important with increasing pH, since
the reason for the comparable adsorption effectiveness. However, most functional groups become dissociated if the pH is above a cer-
the higher absolute value of DP-AC indicates that it has a higher tain limit [34]. The ion exchange mechanisms can be significant
adsorption affinity than BDH-AC which is reflected in the higher if the bonding energy is in the range of 8-16 kJ/mol [35–36]. The
improvement of the uptake capacity. The values for Hads are con- estimated bonding energy (E) for the adsorption of COD on DP-AC
sistent with physical adsorption, where the heat of adsorption is was found to be 0.017 kJ/mol, which suggests that ion exchange
known to be of the order of 10 kJ/mol [28]. does not play an important role and that physical adsorption or
The effect of initial pH on the adsorption of COD was also evalu- chemisorptions is the major adsorption mechanism.
ated at 25 ◦ C for an initial concentration of 3490 mg/L and different Mathematical modeling of the adsorption process is an essen-
initial pH values in the range of 2–10. The typical pH of the refinery tial step in understanding the relative contributions of the different
wastewater was about 8, and it was adjusted to the desired value mechanisms. Mao et al. [37] modeled protein adsorption with
by the addition of few drops of 0.1 M HCl or 0.1 M NaOH. Fig. 9 porous and non-porous particles to determine the relative impor-
shows equilibrium uptake as a function of pH for both adsorbents. tance of the main controlling mechanism: surface reaction and
Although there was a slight enhancement in COD adsorption for pH mass transfer. The model was also applied to the biosorption of
values ranging from 7.5 to 8, the initial pH of the wastewater did lead [12] and copper [38] by C. vulgaris. It is assumed that the trans-
not seem to have any significant effect on the adsorption process. port of the adsorbate from the bulk solution to the surface of the
Similar results for the optimum range of pH were reported for the adsorbent can be described by film resistance mechanism:
adsorption of COD [29], COD and BOD [30–31] and dyes [32–33].
dp
= aK × 10−3 (C − Ci ) (13)
3.5. Modeling and adsorption mechanisms dt

where q and C are the adsorbate concentrations in the adsorbent


Adsorption of contaminants on any solid surface can take place
(mg/g) and in the solution (mg/L), respectively; a is the surface area
through physical adsorption, chemical adsorption or ion exchange.
of the adsorbent per unit mass (cm2 /g); K is the liquid film mass
It is rather difficult to predict the exact mechanism, since it depends
transfer coefficient (cm/min); and Ci is the concentration of the
adsorbate in the solution at the internal solid/liquid interface. The
adsorbate concentration in the adsorbent, q, is related to that in the
solution, C, through mass balance on the adsorbate:

C0 − C
q= (14)
m

where, C0 is the adsorbate concentration in the bulk solution (mg/L)


and m is the adsorbent dose (g/L).
The interaction between the adsorbate and the active sites on
the surface of the adsorbent can be described by the following
second-order reversible equation:

dp
= k1 [(qm − q)Ci − Kd q] (15)
dt

where, k1 is the second-order forward rate constant (L/mg/min); qm


is the maximum adsorption capacity of the adsorbent (mg/g); Kd is
the adsorption equilibrium constant (mg/L), which is the reciprocal
of the constant, b, in Langmuir isotherm.
Combining Eqs. (13)–(15) to eliminate Ci , q and dq/dt, will give
Fig. 9. Equilibrium isotherm data for the adsorption of COD at different pH values. the rate of change in the adsorbate concentration in the solution
Author's personal copy

756 M.H. El-Naas et al. / Journal of Hazardous Materials 173 (2010) 750–757

Table 4
Model parameters for single and dual control mechanisms.

Mechanism k1 (L/mg/min) K (cm/min)

Intrinsic kinetics 0.00002 –


Mass transfer – 6.3 × 10−8
Combined 0.000049 1.2 × 10−7

[34]:
dC
− = (C − x1 )(C − x2 ) (16)
dt
 
mqm −C0 +C 1
where,  = −3 + k1
; x1 and x2 are the roots of the
maK×10
quadratic equation,

C 2 − ˛C − Kd C0 = 0 (17)
  
1
where, ˛ = C0 − mqm − Kd and x1 , x2 = 2
˛± ˛2 + 4Kd C0
Fig. 11. Effect of the initial COD concentration on the contribution ratio. Adsorbent
There are two limiting cases that may be considered here: (1) dose = 20 g/L.
intrinsic kinetics control and (2) mass transfer control. In the first
case, the contribution of mass transfer is assumed to be negligible
(i.e. K → ∞) or of the two mechanisms. It is defined as the ratio of the mass transfer
1 contribution (Eq. (19)) to that of intrinsic kinetics (Eq. (18)):
= (18)  
k1 mqm − C0 + Ceq
= k1 (20)
Eq. (16) is then solved with the new value of  and the solution is maK × 10−3
fitted to the experimental data to estimate the kinetics parameter
k1 . In the second case, the surface reaction is assumed to have very where, Ceq is the adsorbate concentration when the system reaches
fast kinetics (i.e.k1 → ∞) or equilibrium, which is equal to the positive root (x1 ) of Eq. (17). The
  parameter  was calculated for three initial COD contents (950,
mqm − C0 + C 1662 and 3490 mg/L) and found to decrease with increasing the ini-
= (19)
maK × 10−3 tial concentration C0 as shown in Fig. 11. This suggests that for low
initial concentrations (less than 1300 mg/L), the mass transfer is the
Again, Eq. (16) is solved and the solution is fitted to the experi- more dominant mechanism, while intrinsic kinetics dominates for
mental data to estimate the mass transfer parameter K. In all cases high initial concentrations. For the refinery wastewater evaluated
the model differential equations were solved numerically using in this study, the COD content is usually higher than 1300 mg/L and
EZ-Solve software. Values for the estimated model parameters for hence surface reaction will be the dominate mechanism.
kinetics control, mass transfer control and dual control are shown
in Table 4; the surface area a (490 m2 /g) was estimated by surface
area analysis (Micromeritics, Model ASAP-2010). A comparison of 4. Conclusions
the model parameters for single and dual control reveals that there
is no single control and that both mass transfer and intrinsic kinetics The kinetics and equilibrium of the adsorption of COD were
contribute to the adsorption process. examined using two types of adsorbents, namely commercial BDH
The model predictions for the reduction of COD are compared and locally prepared date-pit (DP) activated carbons. The adsorp-
with the experimental data in Fig. 10. El-Naas et al. [12] developed tion effectiveness of DP-AC was found to be comparable to that of
a dimensionless parameter, , to assess the relative contributions the commercial BDH-AC and the equilibrium data for both were
almost equivalent. For both adsorbents, kinetic data were best fit-
ted by the pseudo-second order model, and the equilibrium data
followed the Langmuir isotherm. Modeling the adsorption mech-
anisms indicated that mass transfer was dominant for low COD
content; whereas for high concentrations the intrinsic reaction
kinetics is the dominant mechanism. Overall, the study revealed
that DP-AC is a promising alternative for the highly expensive acti-
vated carbon currently available commercially. In addition, the
utilization of DP-AC can provide an excellent disposal option for
the date palm industry.

Acknowledgements

The authors would like to acknowledge the financial support


provided by the Japan Cooperation Center, Petroleum (JCCP) and
the technical support of the Nippon Oil Research Institute Co., Ltd
(NORI). They would also like to thank Abu Dhabi Oil Refining Com-
pany (TAKREER) and the Research Affairs at the UAE University for
Fig. 10. Predictions of the dual-resistance model compared with experimental data.
their support. Special thanks are also due to Sami Abdulla for his
Initial COD concentration, C0 = 1662 mg/L, adsorbent dose = 20 g/L. help with the experimental work.
Author's personal copy

M.H. El-Naas et al. / Journal of Hazardous Materials 173 (2010) 750–757 757

References [20] H. Zhang, Regeneration of exhausted activated carbon by electrochemical


method, Chem. Eng. J. 85 (2002) 81–85.
[1] D.D. Duong, Adsorption Analysis: Equilibria and Kinetics, Imperial College [21] C.-H. Weng, M.-C. Hsu, Regeneration of granular activated carbon by an elec-
Press, London, 1998. trochemical process, Sep. Purif. Technol. 64 (2008) 227–236.
[2] S. Babel, T.A. Kurniawan, Low-cost adsorbents for heavy metals uptake from [22] K. Okawa, K. Suzuki, T. Takeshita, K. Nakano, Regeneration of granular activated
contaminated water: a review, J. Hazard. Mater. 97 (2003) 219–243. carbon with adsorbed trichloroethylene using wet peroxide oxidation, Water
[3] A. Baran, E. Blcak, S.H. Baysal, S. Onal, Comparative studies on the adsorption Res. 41 (2007) 1045–1051.
of Cr (VI) ions on to various sorbents, Bioresour. Technol. 98 (2007) 661–665. [23] Ö. Aktaş, F. Çeçen, Bioregeneration of activated carbon: A review, Intern. Biod.
[4] D. Mohan, K. Singh, V. Singh, Wastewater treatment using low cost activated Biodegr. 59 (2007) 257–272.
carbons derived from agricultural by-products: a case study, J. Hazard. Mater. [24] J.-L. Lim, M. Okada, Regeneration of granular activated carbon using ultrasound,
152 (2008) 1045–1053. Ultra Sonochem. 12 (2005) 277–282.
[5] E. El-Sharkawy, A. Soliman, K. Al-Amer, Comparative study for the removal [25] E. Fourest, J.C. Roux, Heavy metal biosorption by fungal mycelial byprod-
of methylene blue via adsorption and photocatalytic degradation, J. Colloid ucts: mechanisms and influence of pH, Appl. Microbiol. Biotechnol. 37 (1992)
Interface Sci. 310 (2007) 498–508. 399–403.
[6] J. Hamada, I. Hashim, F. Sharif, Preliminary analysis and potential uses of date [26] A.J. Meikle, G.M. Gadd, R.H. Reed, Manipulation of yeast for transport studies:
pits in foods, Food Chem. 76 (2002) 135–137. critical assessment of cultural and experimental procedure, Enzyme Microb.
[7] F. Banat, S. Al-Asheh, D. Al-Rousan, A comparative study of copper and zinc ion Technol. 12 (1990) 865–872.
adsorption on to activated and non-activated date-pits, Adsorpt. Sci. Technol. [27] F. Abu Al-Rub, M.H. El-Naas, I. Ashour, M. Al-Marzouqi, Biosorption of copper
20 (2002) 319–335. on Chlorella vulgaris from single, binary and ternary metal aqueous solutions,
[8] F. Banat, S. Al-Asheh, L. Al-Makhadmeh, Kinetics and equilibrium study of cad- Process Biochem. 41 (2006) 457–464.
mium ion sorption onto date pits: an agricultural waste, Adsorpt. Sci. Technol. [28] J.m. Smith, Chemical engineering kinetics, 3rd Edition, McGraw-Hill, New York,
21 (2003) 245–260. 1981, pp. 310-322.
[9] F. Banat, S. Al-Asheh, L. Makhadmeh, Preparation and examination of activated [29] A.K. Parande, A. Sivashanmugam, H. Beulah, N. Palaniswamy, Performance eval-
carbons from date pits impergnated with potassium hydroxide for the removal uation of low cost adsorbents in reduction of COD in sugar industrial effluent,
of methylene blue from aqueous solutions, Adsorpt. Sci. Technol. 21 (2003) J. Hazard. Mater. 168 (2009) 800–805.
597–606. [30] R. Devi, R.P. Dahiya, COD and BOD removal from domestic wastewa-
[10] F. Banat, S. Al-Asheh, L. Makhadmeh, Utilization of raw and activated date ter generated in decentralised sectors, Bioresour. Technol. 99 (2008) 344–
pits for the removal of phenol from aqueous solutions, Chem. Eng. Technol. 349.
27 (2004) 80–86. [31] R. Devi, V. Singh, A. Kumar, COD and BOD reduction from coffee process-
[11] S.A. Al-Muhtaseb, M.H. El-Naas, S. Abdullah, Removal of aluminum from aque- ing wastewater using Avacado peel carbon, Bioresour. Technol. 99 (2008)
ous solutions by adsorption on date-pit and BDH activated carbons, J. Hazard. 1853–1860.
Mater. 158 (2008) 300–307. [32] A. Mittal, V. Gajbe, J. Mittal, Removal and recovery of hazardous triph-
[12] M.H. El-Naas, F. Abu Al-Rub, I. Ashour, M. Al Marzouqi, Effect of competitive enylmethane dye, Methyl Violet through adsorption over granulated waste
interference on the biosorption of Lead(II) by C. vulgaris, Chem. Eng. Process. materials, J. Hazard. Mater. 150 (2008) 364–375.
46 (2007) 1391–1399. [33] R. Jain, S. Sikarwar, Adsorptive removal of Erythrosine dye onto activated low
[13] Y.S. Ho, Removal of copper ions from aqueous solution by Tree Fern, Water Res. cost de-oiled mustard, J. Hazard. Mater. 164 (2009) 627–633.
37 (2003) 2323–2330. [34] A. Delgado, A.M. Anselmo, J.M. Novais, Heavy metal biosorption by dried
[14] E. Sabio, E. González, J.F. González, C.M. González-García, A. Ramiro, J. Gañan, powdered mycelium of Fusarium Flocciferum, Water Environ. Res. 70 (1998)
Thermal regeneration of activated carbon saturated with p-nitrophenol, Carbon 370–375.
42 (2004) 2285–2293. [35] Y.S. Ho, J.F. Porter, G. McKay, Equilibrium isotherm studies for the sorption
[15] G. San Miguel, S.D. Lambert, N.J.D. Graham, The regeneration of field-spent of divalent metal ions onto peat: copper, nickel and lead single component
granular-activated carbons, Water Res. 35 (2001) 2740–2748. systems, Water Air Soil Pollut. 141 (2002) 1–33.
[16] F. Salvador, C. Sánchez Jiménez, A new method for regenerating activated car- [36] A. Özcan, T. Sibel, A. Tamer, K. Ismail, Determination of the equilibrium, kinetic
bon by thermal desorption with liquid water under subcritical conditions, and thermodynamic parameters of adsorption of copper(II) ions onto seeds of
Carbon 34 (1996) 511–516. Capsicum annuum, J. Hazard. Mater. 124 (2005) 200–208.
[17] M.K. Purkait, A. Maiti, S. DasGupta, S. De, Removal of congo red using activated [37] Q.M. Mao, R. Stockmann, I.G. Prince, M.T.W. Hearn, Modeling of protein adsorp-
carbon and its regeneration, J. Hazard. Mater. 145 (2007) 287–295. tion with non-porous and porous particles in a finite bath, J. Chromatogr. 646
[18] P.M. Álvarez, F.J. Beltrán, V. Gómez-Serrano, J. Jaramillo, E.M. Rodríguez, Com- (1993) 67–80.
parison between thermal and ozone regenerations of spent activated carbon [38] M.A. Hashim, K.H. Chu, Modeling of the batch adsorption of copper by the
exhausted with phenol, Water Res. 38 (2004) 2155–2165. Miroalga Chlorella vulgaris, in: Proceedings of the Sixth World Congress of
[19] C.O. Ania, J.A. Menéndez, J.B. Parra, J.J. Pis, Microwave-induced regeneration Chemical Engineering, Melbourne, Australia, 2001.
of activated carbons polluted with phenol. A comparison with conventional
thermal regeneration, Carbon 42 (2004) 1383–1387.

You might also like