You are on page 1of 16

Colloids and Surfaces A: Physicochem. Eng.

Aspects 272 (2006) 89–104

Adsorptive removal of phenol by bagasse fly ash and activated carbon:


Equilibrium, kinetics and thermodynamics
Vimal C. Srivastava, Mahadeva M. Swamy 1, Indra D. Mall ∗, Basheswar Prasad, Indra M. Mishra

Department of Chemical Engineering, Indian Institute of Technology Roorkee, Roorkee, Uttaranchal 247667, India

Received 3 April 2005; received in revised form 9 July 2005; accepted 14 July 2005
Available online 26 September 2005

Abstract

Present study deals with the adsorption of phenol on carbon rich bagasse fly ash (BFA) and activated carbon-commercial grade (ACC)
and laboratory grade (ACL). BFA is a solid waste obtained from the particulate collection equipment attached to the flue gas line of the
bagasse-fired boilers of cane sugar mills. Batch studies were performed to evaluate the influences of various experimental parameters like
initial pH (pH0 ), contact time, adsorbent dose and initial concentration (C0 ) on the removal of phenol. C0 varied from 75 to 300 mg/l for the
adsorption isotherm studies and the effect of temperature on adsorption. Optimum conditions for phenol removal were found to be pH0 ≈ 6.5,
adsorbent dose ≈10 g/l of solution and equilibrium time ≈5 h. Adsorption of phenol followed pseudo-second order kinetics with the initial
sorption rate for adsorption on ACL being the highest followed by those on BFA and ACC. The effective diffusion coefficient of phenol is of
the order of 10−10 m2 /s. Equilibrium isotherms for the adsorption of phenol on BFA, ACC and ACL were analysed by Freundlich, Langmuir,
Temkin, Redlich–Peterson, Radke–Prausnitz and Toth isotherm models using non-linear regression technique. Redlich–Peterson isotherm
was found to best represent the data for phenol adsorption on all the adsorbents. The change in entropy (S◦ ) and heat of adsorption (H◦ )
for phenol adsorption on BFA were estimated as 1.8 MJ/kg K and 0.5 MJ/kg, respectively. The high negative value of change in Gibbs free
energy (G◦ ) indicates the feasible and spontaneous adsorption of phenol on BFA. The values of isosteric heat of adsorption varied with the
surface loading of phenol.
© 2005 Elsevier B.V. All rights reserved.

Keywords: Adsorption; Phenol removal; Bagasse fly ash (BFA); Activated carbon; Adsorption thermodynamics; Temperature; Kinetics; Isotherms

1. Introduction and its melting point is 181 ◦ C. It is a weak acid dissociating


slightly in aqueous solution: for this reason it is also known
Phenol, a derivative of benzene, is an important raw mate- as carbolic acid. The Ministry of Environment and Forests
rial and/or product of chemical and allied industries (e.g. (MOEF), Government of India and EPA, USA, have listed
petrochemicals, oil refineries, plastics, leather, paint, phar- phenol and phenolic compounds on the priority-pollutants
maceutical, steel industries and pesticides) [1,2]. It is highly list. Chronic toxic effects due to phenols reported in humans
soluble in water and is very toxic in nature. It is a colour- include vomiting, difficulty in swallowing, anorexia, liver and
less, hygroscopic and crystalline substance, which turns pink kidney damage, headache, fainting and other mental distur-
in air owing to its oxidation. Its solubility in water is 98 g/l bances [3]. That phenol is highly toxic and difficult to degrade
biologically have led to setting up of rigid limits on the accept-
able level of phenol in the environment. While the MOEF has
∗ Corresponding author. Tel.: +91 1332 285319 (O)/285106 (R);
set a maximum concentration level of 1.0 mg/l of phenol in
fax: + 91 1332 276535/273560.
the industrial effluents for safe discharge into surface waters,
E-mail address: id mall2000@yahoo.co.in (I.D. Mall).
1 Present address: Department of Environmental Engineering, Sri Jay- the WHO recommends the permissible phenolic concentra-
achamarajendra College of Engineering, Mysore, Karnataka, India. tion of 0.001 mg/l in potable waters [4].

0927-7757/$ – see front matter © 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.colsurfa.2005.07.016
90 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

Nomenclature
qe,exp experimental value of solid phase concentra-
1/n heterogeneity factor, dimensionless tion of adsorbate at equilibrium (mg/g)
aR constant of Redlich–Peterson isotherm (l/mg) qe∞ monolayer adsorption capacity parameter in
A adsorbate Toth isotherm (mg/g)
ACC activated carbon-commercial grade qm maximum adsorption capacity of adsorbent
ACL activated carbon-laboratory grade (mg/g)
A.S activated complex qt amount of adsorbate adsorbed by adsorbent at
BFA bagasse fly ash time t (mg/g)
C constant that gives idea about the thickness of R universal gas constant (8.314 J/K mol)
boundary layer (mg/g) Ra radius of the adsorbent particle assumed to be
C0 initial concentration of adsorbate in solution spherical (m)
(mg/l) RL separation factor, dimensionless
Ce equilibrium liquid phase concentration (mg/l) S active site on the adsorbent
CS adsorbent concentration in the solution t time (min)
De effective diffusion coefficient of adsorbate in T absolute temperature (K)
the adsorbent phase (m2 /s) Th constant in Toth isotherm, dimensionless
f exchange of number of moles of water per V volume of the solution (l)
mole of adsorbate at the adsorption site w mass of the adsorbent (g)
F(t) fractional uptake of adsorbate on adsorbent, XAe fraction of the adsorbate adsorbed on the adsor-
0 < F(t) < 1 bent under equilibrium
h initial sorption rate (mg/g min) z an integer
HYBRID hybrid fractional error function G◦ Gibbs free energy of adsorption (KJ/mol)
k0 constant in Bangham equation H◦ enthalpy of adsorption (KJ/mol)
kf rate constant of pseudo-first-order adsorption Hw heat of adsorption of water (normally assumed
model (min−1 ) to be zero)
kid intra-particle diffusion rate constant Hsol heat of solution
(mg/g min1/2 ) Hst,0 isosteric heat of adsorption with zero coverage
kA adsorption rate constant for the adsorption Hst,a apparent isosteric heat of adsorption
equilibrium Hst,net net isosteric heat of adsorption
kD desorption rate constant for the adsorption S◦ entropy of adsorption (J/K mol)
equilibrium
kRP constant in Radke–Prausnitz isotherm Greek symbols
((mg/g)/(mg/l)1/P ) α Bangham constant (<1)
kS rate constant of pseudo-second-order adsorp- β constant of Redlich–Peterson isotherm
tion model (g/mg min) (0 < β < 1)
KA equilibrium adsorption constant λvap heat of vapourisation of phenol
KF constant of Freundlich isotherm
((mg/g)/(l/mg)1/n )
KL constant of Langmuir isotherm (l/mg)
KR constant of Redlich–Peterson isotherm (l/g)
Several methods for the treatment of phenolic waste water
KRP constant in Radke–Prausnitz isotherm (l/g)
have been proposed in the literature. These include physico-
KTh constant in Toth isotherm ((mg/l)Th )
chemical treatment processes, chemical oxidation and bio-
m mass of adsorbent per liter of solution (g/l)
logical degradation. The physico-chemical processes include
n number of data points
adsorption and ion exchange. Various oxidizing agents (oxy-
N number of data points
gen, hydrogen peroxide, ozone, etc.) have been used for
p number of parameters
wet oxidation of phenolic waste waters. For high strength
P constant in Radke–Prausnitz isotherm, dimen-
and low volume of phenolic waste waters, phenol removal
sionless
by adsorption using granular/powdered activated carbon has
MPSD Marquardt’s percent standard deviation
been widely used [5–12]. However, high costs of activated
qe equilibrium solid phase concentration (mg/g)
carbon and 10–15% loss during regeneration and the dif-
qe,cal calculated value of solid phase concentration
ficulties faced in the recovery/disposal of phenol make the
of adsorbate at equilibrium (mg/g)
utilization of activated carbon prohibitive in the developing
countries [8]. This has led to search for cheaper carbonaceous
substitutes to activated carbon.
V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104 91

Most of the studies on phenol removal dealt with the Bulk densities were determined by using MAC bulk den-
adsorptive equilibrium and kinetics, both in batch and contin- sity meter whereas particle size analysis was done using
uous mode. The equilibrium relationships were represented standard sieves. The specific surface area and the pore
by various isotherm models, viz. two parameter models diameter of the samples were measured by N2 adsorption
of Freundlich [8–21] and Langmuir [9,11–13,17–22]; three isotherm using an ASAP 2010 Micromeritics instrument
parameter models of Redlich–Peterson [18]. The details of and by Brunauer–Emmett–Teller (BET) method, using the
these models and their characteristics have been extensively software of Micromeritics. Nitrogen was used as cold bath
dealt with by Swamy et al. [8]. Most of the authors have used (77.15 K).
Vermeulen’s [23] approximations to the analytical solution of X-ray diffraction analyses of the adsorbents were carried
a differential equation to represent the fractional approach to out using Phillips diffraction unit (Model PW 1140/90), using
equilibrium. Mechanistic studies to understand sorption char- copper as the target with nickel as filter media, and K radiation
acteristics have been reported by many investigators [24,25]. maintained at 1.542 Å. Goniometer speed was maintained
Temperature is known to have pronounced effect on adsorp- at 1◦ /min. Scanning electron microscopic (SEM) analyses
tion phenomenon and intraparticle diffusional process. of the adsorbents were carried out by using a scanning
In the present investigation, sugarcane bagasse fly ash electron microscope (Model SEM-501, Phillips, Holland)
(BFA) collected from the particulate collection device [27].
attached to the flue gas line of a bagasse-fired boiler stack
of a nearby canesugar mill, and activated carbon-commercial 2.2. Adsorbate
grade (ACC) and laboratory grade (ACL) have been used
as adsorbents for the removal of phenol from aqueous solu- Phenol (C6 H5 OH) of analytical reagent (AR) grade sup-
tion. The aim of the present work is to explore the possibility plied by Ranbaxi Laboratories Ltd., India, was used for the
of BFA, which is a carbonaceous waste, being utilized as preparation of the synthetic adsorbate solutions of various C0
an adsorbent for the removal of phenol from wastewater. in the range of 75–300 mg/l. The required quantity of phenol
BFA has fairly good amount of carbon and silica in it. ACL was accurately weighed and dissolved in a small amount of
has been used as the standard adsorbent for the comparison distilled water and subsequently made-up to 1 l in a measur-
of adsorptive capacities of BFA and ACC. Effects of such ing flask. Fresh stock solution as required was prepared every
parameters as initial pH (pH0 ), adsorbent dose (m), initial day and was stored in a brown colour glass reservoir of 5 l
concentration (C0 ), and contact time (t) on the sorption of capacity to prevent photo-oxidation. The C0 was ascertained
phenol have been investigated. The kinetics of adsorption before the start of each experimental run.
of phenol on the adsorbents has been studied using differ-
ent models. The sorption capacity of the adsorbents has been 2.3. Analytical measurements
studied using the adsorption isotherm technique. Experimen-
tal data were fitted to various isotherm equations to determine The concentration of phenol was determined by finding
the best isotherm to correlate the experimental data. The out the absorbance of the solution at 270 nm wavelength
effect of temperature for the adsorption of phenol onto BFA using UV/vis spectrophotometer (model UV 210 A; Schi-
has also been investigated. Thermodynamics of adsorption madzu, Japan). The calibration plot of absorbance versus
process have been studied and the change in Gibbs free energy concentration for phenol showed a linear variation up to
and the enthalpy, and isosteric heat of adsorption have also 40 mg/l concentration. Therefore, the samples with higher
been determined. concentration of phenol (>40 mg/l) were diluted with dis-
tilled water, whenever necessary, to make the concentration
less than 40 mg/l, for the accurate determination of the phe-
2. Material and methods nol concentration with the help of the linear portion of the
calibration curve. The pH of the aqueous solution of phenol
2.1. Adsorbents and their characterization varied from 7.04 to 7.59 for the phenol concentration varying
from 10 to 200 mg/l.
BFA was obtained from U.P. State Sugar Corporation
Ltd., Doiwala Unit, Dehradun, India. The commercial grade 2.4. Batch experimental programme
activated carbon (ACC) manufactured by Rajasthan Brew-
eries Ltd. was procured from the open market (Kharibaoli, For each experiment, 50 ml of the phenol solution of
Delhi) and the laboratory grade activated carbon (ACL) man- known C0 , pH0 and a known amount of the adsorbents were
ufactured by GSE Chemical Testing Laboratory and Allied taken in a 100 ml stoppered conical flask. This mixture was
Industry, New Delhi, was procured from local suppliers. The agitated in a temperature-controlled shaking water bath at a
adsorbents were used as procured. constant speed of 145 rpm. Small amount of the sample was
The physico-chemical characteristics of the adsorbents withdrawn after 5 h and was filtered through Whatman filter
were determined using standard procedures. Proximate anal- paper no. 42 (pore size ca. 2.5 ␮m) and analysed for phenol
ysis was carried out using the standard procedure [26]. concentration. The sample was again transferred to the con-
92 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

ical flask. The pH0 of the adsorbate solutions was adjusted rium adsorption data. For adsorption isotherms, experiments
using 1 M aqueous solution of either HCI or NaOH. were carried out at pH0 of 6.5 by contacting fixed amount of
The percentage removal of phenol and equilibrium adsorp- adsorbent with 50 ml of phenol solution at the C0 varying over
tion uptake, qe (mg/g), were calculated using the following the range of 75–300 mg/l. Adsorbents were separated from
relationships: the solution after 5 and 24 h, and the phenol concentration in
the solution and in the adsorbent was estimated.
100(C0 − Ce )
% removal = , (1) In order to investigate the kinetics of adsorption of phe-
C0 nol on the adsorbents, various kinetic models, like pseudo-
first-order, pseudo-second-order, Bangham and intraparticle
(C0 − Ce )V diffusion models were used. Two parameters model, viz.,
amount adsorbed qe =
w Langmuir, Freundlich and Temkin, and three-parameter mod-
(mg of adsorbate/g of adsorbent), (2) els, viz., Redlich–Peterson, Radke–Prausnitz and Toth have
been used to describe the equilibrium nature of adsorption of
where C0 is the initial sorbate concentration (mg/l), Ce the phenol in the present study.
equilibrium sorbate concentration (mg/l), V the volume of the Two different error functions of non-linear regression
solution (l) and w is the mass of the adsorbent (g). basin were employed in this study to find out the most
suitable kinetic and isotherm models to represent the experi-
2.5. Effect of initial pH (pH0 ) mental data respectively. The hybrid fractional error function
(HYBRID) [28] and the Marquardt’s percent standard devi-
The sorption of phenol by the adsorbents was studied over ation (MPSD) error function [29] have been used previously
a pH0 range of 3–10 at 30 ◦ C and the studies were carried out by a number of researchers in the field [30,31]. These error
for 5 h. C0 was 100 mg/l and the adsorbent dose was kept at functions are given as:
10 g/l for ACL and BFA and 12 g/l for ACC.
n  
100  qe,meas − qe,calc
HYBRID = (3)
2.6. Effect of temperature and estimation of n−p qe,meas
i=1 i
thermodynamic parameters

 n  
 1  qe,meas − qe,calc 2
The effect of temperature on the sorption characteristics MPSD = 100 (4)
was investigated by determining the adsorption isotherms at n−p qe,meas i
i=1
25, 30, 35, 40 and 45 ◦ C. Apparent and net isosteric heats of
adsorption at various surface coverages have been determined HYBRID was developed to improve the fit of the square of
using classical thermodynamic equations. C0 was varied from errors function at low concentration values. The MPSD is
75 to 300 mg/l but the pH0 of the solutions was maintained similar in some respects to a geometric mean error distribu-
at 6.5. tion modified according to the number of degrees of freedom
of the system.
2.7. Batch kinetic and isotherm study and error analysis

To determine the t necessary for adsorption, 50 ml of the 3. Results and discussion


aqueous solution containing 100 mg/l of phenol was taken in
a series of conical flasks. Preweighed amounts of the adsor- 3.1. Characterisation of adsorbents
bents were added to different flasks. The flasks were kept in
a temperature-controlled shaking water bath and the aque- Physico-chemical characteristics of the three adsorbents
ous solution–adsorbent mixtures were stirred at 145 rpm. At are presented in Table 1 and particle size analysis is given
the end of the predetermined t, the flasks were withdrawn, in Table 2. The ash obtained by incineration of BFA con-
their contents were filtered, and the filtrate analyzed for phe- sists mainly of carbon, silica, alumina, calcium oxide, etc.
nol. The amount of phenol adsorbed onto the filter paper From Table 1, it is observed that the bulk density of BFA is
during the filtration was found to be less than 1% for all lower than that of ACC or ACL. The specific surface areas
the C0 between 20 and 60 mg/l. Above 75 mg/l, the phenol of the adsorbents used in the present investigations are given
adsorption onto the filter paper was found to be less than in Table 1. From geometrical considerations of the particle
0.7%. Hence, the effect of filter paper on adsorption by the shape, it is expected that the surface area per unit mass of
adsorbents has been neglected during the studies. Adsorption BFA will be much lower than that of either ACC or ACL.
kinetics was followed for 24 h and it was observed that after The morphologies of BFA, ACC and ACL were exam-
1 h, there was gradual but very slow removal of adsorbate ined under scanning electron microscope. The SEMs of BFA,
from the solution. ACC and ACL as obtained are shown in Fig. 1. The BFA has
To optimize the design of an adsorption system for the linear type of fibres with holes in it and at other places and
removal of adsorbate, it is important to obtain the equilib- has skeletal structure. Further, the number of pores in BFA
V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104 93

Table 1 Table 1. Fig. 2 shows the pore size distribution of BFA and
Characteristics of bagasse fly ash and activated carbons ACC. While ACC has a narrower pore size distribution, BFA
Characteristics ACLa ACC BFA has a much wider size distribution. The average pore size dis-
Proximate analysis tribution of BFA is also much larger. Although, the detailed
Moisture (%) 0.31 9.00 2.51 pore size analysis of the ACL is not available, it is expected
Ash (%) 2.65 8.78 30.98 that the pore size distribution shall be in between that of BFA
Volatile matter (%) 7.74 24.08 23.48
Fixed carbon (%) 89.30 58.14 43.03
and ACC. It has both macro- and micro-pores [32].
Bulk density (kg/m3 ) 750.82 610.00 270.50 X-ray diffraction patterns for BFA, ACC and ACL are
shown in Fig. 3. Major components identified in BFA are
Chemical analysis of ash
SiO2 (%) 8.00 54.93 51.05 crystalline quartz, alumina, cristobalite, and calcium orthosil-
Al2 O3 (%) ND 6.84 10.75 icate, whereas, tridymite and silicate carbon are the major
CaO (%) ND 11.15 6.04 components of ACC and ACL [33]. Diffraction peaks cor-
Fe2 O3 (%) ND 1.26 3.45 responding to crystalline carbon were not observed in BFA
MgO (%) ND 2.36 1.10
and activated carbons. Nowackiite was also observed in ACL.
Rest others – – –
The broad peaks in all the three samples indicate the pres-
Surface area of pores (m2 /g) ence of amorphous form of silica. From the micrographs and
(i) BET 492.00 336.60 168.39
(ii) BJH
X-ray diffractograms, it is seen that the activated carbons and
(a) adsorption cumulative – 64.43 70.90 the BFA have heterogeneous surface. Proximate analysis of
(b) desorption cumulative – 34.03 45.30 adsorbents showed 43.03, 89.03 and 58.14% fixed carbon in
BJH cumulative pore volume (cm3 /g) BFA, ACL and ACC, respectively. Thus, BFA has about one-
(i) Single point total – 0.1622b 0.1067c half of the carbon present in ACL and about three-fourth of
(ii) BJH adsorption – 0.0425 0.0844 the carbon present in ACC.
(iii) BJH desorption – 0.0224 0.0622
Average pore diameter (Å)
(i) BET – 19.72 25.54
3.2. Effect of adsorbent dosage (m)
(ii) BJH adsorption – 26.62 49.85
(iii) BJH desorption – 26.34 58.44 The effect of m on the uptake of phenol on BFA, ACC
a Detailed analyses of pore surface/volume not performed. and ACL was studied and is shown in Fig. 4. This figure
b Pores less than 2239 Å. reveals that the removal of phenol increases with increase in
c Pores less than 2194 Å. adsorbent dosage. The removal of phenol at adsorbent dosage
larger than 10 g/l remains almost unchanged. An increase in
are found to be less than that of ACL and these are also rela- the adsorption with the adsorbent dosage can be attributed to
tively larger in size, but the numbers of pores are more than greater surface area and the availability of more adsorption
that present in ACC. Activated carbon is generally described sites. At m < 5 g/l, the adsorbent surface become saturated
as an amorphous form of graphite with a random structure of with phenol and the residual concentration in the solution is
graphite plates; having highly porous structure with a range large. With increase in the m, the phenol removal increases
of cracks and crevices reaching molecular dimensions. The due to increased phenol uptake by the increases amount
summary of porous structure of BFA and ACC are shown in of adsorbent. At m > 5 g/l, the incremental phenol removal

Table 2
Particle size analysis of different adsorbents
ACL ACC BFA

Size (mm) Weight (%) Size Weight (%) Size (␮m) Weight (%)
>4.00 1.42 >1.19 mm 6.72 >1000 0.73
−4.00 + 3.33 3.23 −1.19 + 1.18 mm 17.21 −1000 + 850 0.80
−3.33 + 2.80 11.70 −1.18 + 1.14 mm 14.92 −850 + 710 3.58
−2.80 + 2.36 45.30 −1.14 + 1.0 mm 21.80 −710 + 600 1.42
−2.36 + 2.00 22.85 −1.0 mm + 850 ␮m 32.07 −600 + 500 0.84
−2.00 + 1.70 9.51 −850 + 710 ␮m 5.00 −500 + 355 3.84
−1.70 + 1.40 4.89 −710 + 425 ␮m 2.23 −355 + 300 2.78
<1.40 1.10 <425 ␮m 0.35 −300 + 150 19.20
−150 + 125 13.84
−125 + 106 22.83
−106 + 90 0.42
−90 + 75 14.81
−75 + 63 9.45
−63 + 45 1.78
<45 3.68
94 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

Fig. 2. Pore size distribution of adsorbents (a) BFA and (b) ACC.

becomes very low, as the surface phenol concentration and the


solution phenol concentration come to equilibrium with each
other. At about m = 10 g/l, the removal efficiency becomes
almost constant. There is very little difference between the
phenol uptake by ACL and BFA.

3.3. Effect of initial pH (pH0 )

The pH of the solution affects the surface charge of the


adsorbents as well as the degree of ionization and specia-
tion of different pollutants [34]. Change in pH affects the
adsorptive process through dissociation of functional groups
on the adsorbent surface active sites. This subsequently leads
to a shift in reaction kinetics and equilibrium characteris-
tics of adsorption process. Adsorption of various anionic and
cationic species on such adsorbents has been explained on the
basis of the competitive adsorption of H+ and OH− ions with
the adsorbates [35]. It is a common observation that the sur-
face adsorbs anions favourably at lower pH due to presence
of H+ ions, whereas, the surface is active for the adsorption
of cations at higher pH due to the deposition of OH− ions
[36].
Fig. 1. Scanning electron micrograph of adsorbents at 150× magnification
The influence of the pH0 of phenolic solution on the extent
(1 cm = 100 ␮m).
of adsorption of phenol is shown in Fig. 5. Adsorption of phe-
nol decreases with increase in pH0 . Up to pH0 6.5 the decrease
in adsorption is gradual, which, however, drops drastically
for pH0 > 6.5. The presence of oxides of aluminum, calcium
and silicon on the adsorbents (BFA, ACC and ACL) leads to
V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104 95

Fig. 5. Effect of pH0 on the removal of phenol. T: 30 ◦ C; t: 5 h; C0 : 100 mg/l;


ACL dosage: 10 g/l; BFA dosage: 10 g/l; ACC dosage: 12 g/l.

the development of charge when adsorbent is in contact with


water according to the pH of the solution as follows:
M OH + H+ → M OH2 +

M OH + OH− → M O− + H2 O
where M stands for Al, Ca and Si. Except silica, all other
oxides will possess positive charge for a pH range of inter-
est because the zero point charge of SiO2 , Al2 O3 and CaO
are 2.2, 8.3 and 11.0, respectively [37]. For pH0 below 6.5, a
significantly high electrostatic attraction exists between the
positively charged surface of the adsorbent and the phenolate
ion, C6 H5 O− ion. Phenol being a weak acid is adsorbed to
a lesser extent at higher pH values as the negatively charged
surfaces of the adsorbent did not favour the adsorption of
C6 H5 O− ion due to electrostatic repulsion. Such observa-
tions were also reported by other workers [38]. From Fig. 5,
it is observed that the adsorption of phenol by ACL is slightly
higher than that by BFA for phenol while the ACC adsorption
capacity is much lower. The phenol adsorption is limited by
the micropore volume and the acid–basic characteristics of
the adsorbents [39]. Since the diameter of phenol molecule
Fig. 3. X-ray diffraction patterns for BFA, ACC and ACL. is about 6 Å [40], the mesoporous adsorbents with sufficient
pore surface area may show better adsorption characteristics.
This is clearly discerned from the pore size characteristics
shown in Table 1 and Fig. 2. Due to higher surface area of
mesoporous structure of BFA than that of ACC, and pos-
sibly favourable surface functionalities, BFA accounts for
higher phenol uptake than ACC. Similar reasons influence
the adsorption of phenol onto ACL as well. Since the change
from acidic pH to alkaline pH reduces the adsorption capac-
ity of the adsorbents drastically, the rest of the experiments
were performed at slightly acidic pH (∼6.5).

3.4. Effect of initial phenol concentration (C0 )

The effect of C0 on the extent of adsorption on BFA as a


function of time is shown in Fig. 6. At any time the amount
Fig. 4. Effect of adsorbent dosage on the removal of phenol. pH0 : 6.5; T: of phenol adsorbed per unit weight of adsorbent increased
30 ◦ C; t: 5 h; C0 : 100 mg/l. with increasing C0 . The C0 provides the necessary driving
96 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

of phenol in solution versus contact time for the adsorbents at


an C0 of 100 mg/l. The rate of phenol removal is found to be
very rapid during the initial 30 min and thereafter the rate of
phenol removal decreases. No significant change in phenol
removal is observed after about 120 min. It is also found that
the removal of phenol by BFA is only slightly less than that
by ACL at any contact time and that the removal of phenol
by ACC is around 15% less than that by BFA. Similar trends
were observed for different C0 , viz. 50, 150 and 200 mg/l. It
was found that the adsorptive removal of the phenol ceases
after 300 min of contacting with the three adsorbents.

Fig. 6. Effect of initial phenol concentrations on removal of phenol by BFA. 3.6. Adsorption kinetic study
pH0 : 6.5; T: 30 ◦ C; t: 5 h; BFA dosage: 10 g/l.

force to overcome the resistances to the mass transfer of phe- Four kinetic models, viz. pseudo-first-order, pseudo-
nol between the aqueous and the solid phases. The increase second-order, Bangham and intraparticle diffusion models,
in C0 also enhances the interaction between phenol and the were used to investigate the adsorption process of phenol on
BFA. Therefore, an increase in C0 of phenol enhances the BFA, ACC and ACL.
adsorption uptake of phenol.
3.6.1. Pseudo-first-order model
3.5. Effect of contact time The sorption of organic molecules from a liquid phase to
a solid phase can be considered as a reversible process with
Available adsorption results reveal that the uptake of equilibrium being established between the solution and the
adsorbate species is fast at the initial stages of the contact solid phase. Assuming a non-dissociating molecular adsorp-
period, and thereafter, it becomes slower near the equilib- tion of phenols on BFA particles, the sorption phenomenon
rium. In between these two stages of the uptake, the rate of can be described as the diffusion controlled process [41].
adsorption is found to be nearly constant. This is obvious kA
from the fact that a large number of vacant surface sites are A + SAS (5)
kD
available for adsorption during the initial stage, and after a
lapse of time, the remaining vacant surface sites are difficult where A is adsorbate and S is the active site on the adsor-
to be occupied due to repulsive forces between the solute bent and AS is the activated complex. kA and kD are the
molecules on the solid and bulk phases. adsorption and desorption rate constants, respectively. Using
Aqueous phenol solutions with different C0 were kept in first-order kinetics it can be shown that with no adsorbate
contact with the adsorbents for 24 h. The values of the resid- initially present on the adsorbent (i.e. CAS0 = 0 at t = 0) the
ual concentrations at 5 h contact time were found to be higher fractional uptake of the adsorbate by the adsorbent can be
by a maximum of ∼2% than those obtained after 24 h contact expressed as:
 
time. Therefore, after 5 h contact time, a steady state approx- XA kA
imation was assumed and a quasi-equilibrium situation was = 1 − exp kA CS + t (6)
XAe KS
accepted. Accordingly all the batch experiments were con-
ducted with a contact time of 5 h under vigorous shaking where XAe is the fraction of the adsorbate adsorbed on the
conditions. Fig. 7 presents the plot of residual concentration adsorbent under equilibrium condition, KS = kA /kD and CS
is the adsorbent concentration in the solution Eq. (6) can be
transformed as:
 
qe − q −kf
log10 = t (7)
qe 2.303
where
 
kA
kf = kA CS + (8)
kS
q = XA and qe = XAe
This equation is the so-called Lagergren equation [42]. Eq.
(7) when plotted (not shown here) as log10 ((qe − q)/qe ) versus
t for the adsorption of phenol CA0 = 100 mg/l at 30 ◦ C and at
Fig. 7. Effect of contact time on removal of phenol. pH0 : 6.5; T: 30 ◦ C; C0 : a pH0 6.5, showed a straight line passing through the origin.
100 mg/l; ACL and BFA dosage: 10 g/l; ACC dosage: 12 g/l. However, after 15 min of contact time, the data is represented
V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104 97

by the two distinct straight lines. The adsorption data do not surface diffusion and adsorption on the pore surface, or a
fit this equation, as the R2 values for BFA, ACC and ACL are combination of more than one step. In a rapidly stirred batch
found to be 0.8931, 0.4996 and 0.9509, respectively. adsorption, the diffusive mass transfer can be related by an
apparent diffusion coefficient, which will fit the experimental
3.7. Pseudo-second-order model sorption-rate data. Generally, a process is diffusion controlled
if its rate is dependent upon the rate at which components
The pseudo-second-order model can be represented in the diffuse towards one another. The possibility of intra-particle
following form [43]: diffusion was explored by using the intra-particle diffusion
model [25,44–46].
tKS qe2
qt = (9)
1 + tKS qe qt = kid t 1/2 + C (11)
the initial sorption rate, h (mg/g min), as t → 0 can be defined where, kid is the intra-particle diffusion rate constant
as: (mg/g min1/2 ) and C (mg/g) is a constant that gives idea
h = kS qe2 (10) about the thickness of the boundary layer, i.e., larger the
value of C the greater is the boundary layer effect [47]. If
The equilibrium adsorption capacity (qe ) and the initial the Weber–Morris [25] plot of qt versus t0.5 gives a straight
sorption rate (h) alongwith the pseudo-second-order con- line, then the sorption process is controlled by intra-particle
stant kS were determined from the non-linear fitting of the diffusion only. However, if the data exhibit multi-linear plots,
data. Kinetic parameters alongwith correlation coefficient then two or more steps influence the sorption process. The
for pseudo-second order kinetic model are listed in Table 3. mathematical dependence of fractional uptake of adsorbate
The calculated correlation coefficient is closer to unity for on t1/2 is obtained if the sorption process is considered to be
pseudo-second-order kinetic model than the pseudo-first- influenced by diffusion in the cylindrical (or spherical) and
order kinetic model. Therefore, the sorption reaction can be convective diffusion in the adsorbate solution. It is assumed
approximated more favourably by the pseudo-second-order that the external resistance to mass transfer surrounding the
kinetic model for all the three adsorbents. Error functions particles is significant only in the early stages of adsorption.
as shown in Table 3 are also considerably less for pseudo- This is represented by first sharper portion. The second linear
second-order kinetic model reinforcing the applicability of portion is the gradual adsorption stage with intra-particle
the pseudo-second-order kinetic model. diffusion dominating.
Fig. 8 presents the plots of mass of phenol adsorbed per
3.8. Intra-particle diffusion study unit mass of adsorbent versus t1/2 for all the adsorbents. In the
figure the data points are related by two straight lines—the
The adsorbate transport from the solution phase to the first straight portion depicting macropore diffusion and the
surface of the adsorbent particles occurs in several steps. The second representing micro-pore diffusion [46]. These show
overall adsorption process may be controlled either by one only the pore diffusion data. Extrapolation of the linear por-
or more steps, e.g. film or external diffusion, pore diffusion, tions of the plots back to the y-axis gives the intercepts, which

Table 3
Kinetic parameters for the removal of phenol by different adsorbents
Adba h (mg/g min) qe (mg/g) kS (g/mg min) R2 HYBRID MPSD
Pseudo-second-order constants
ACL 3.6784 9.6436 0.0396 0.9999 −5.4414 11.6985
ACC 1.4591 7.0689 0.0292 0.9998 −3.8630 6.1933
BFA 3.3379 9.4989 0.0370 0.9999 −5.5468 12.0807

Adba kid,1 b R2 HYBRID MPSD kid,2 b R2 HYBRID MPSD


Weber–Morris constants
ACL 0.48742 0.9527 −19.8824 28.1277 0.0126 0.9591 −0.00021 0.1446
ACC 0.32786 0.9820 −22.9889 32.4988 0.0286 0.9490 −0.0059 0.7653
BFA 0.45137 0.9902 −23.3487 33.0060 0.0137 0.8967 −0.00033 0.1797

Adba k0 (l/(g/l)) α R2 HYBRID MPSD


Bangham constants
ACL 5.0262 0.3566 0.9326 −26.2841 41.6086
ACC 3.6370 0.3091 0.9515 −95.9866 103.581
BFA 4.8033 0.3468 0.9305 −30.3758 46.67921
a Adsorbent.
b Unit (mg/g min1/2 ).
98 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

where bn ’s are roots of J0 (bnp R) = 0.


Vermeulen’s approximation [23] of the Eq. (12) fits the
whole range 0 < F(t) < 1, for adsorption on spherical particles.
  1/2
−π2 De t
F (t) = 1 − exp (14)
R2a
This equation could further be simplified to cover most of the
data points for calculating effective particle diffusivity.
 
1 π 2 De t
ln = (15)
1 − F (t)
2 R2a

Fig. 8. Weber and Morris intra-particle diffusion plots for removal of phenol. Thus the slope of the plot of ln[1/(1 − F2 (t))] versus t would
pH0 : 6.5; T: 30 ◦ C; C0 : 100 mg/l; ACL and BFA dosage: 10 g/l; ACC dosage: give De . Table 4 also presents the value of effective diffu-
12 g/l. sion coefficient (De ) as calculated from Eq. (15). Value of
provide the measure of the boundary layer thickness. The De is 1.800, 1.052 and 0.388 × 10−10 m2 /s, respectively, for
deviation of straight lines from the origin (Fig. 8) may be due phenol adsorption on ACL, BFA and ACC. This shows that
to difference in rate of mass transfer in the initial and final ACL has highest overall pore diffusion rate. Table 4 shows
stages of adsorption. Further, such deviation of straight line the values of the effective pore diffusivities and the homoge-
from the origin indicates that the pore diffusion is not the neous surface diffusion model (HSDM) surface diffusivities
sole rate-controlling step. The adsorption data for q versus for the adsorption of phenol on activated carbon and other
t1/2 for the initial period show curvature, usually attributed adsorbents. Wide variation is seen in the data with minimum
to boundary layer diffusion effects or external mass transfer value of surface diffusivity being 3.5 × 10−13 m2 /s and the
effects [45,48,49]. The slope of the Weber and Morris plots – maximum value being that of De for wood derived activated
q versus t1/2 – are defined as a rate parameter, characteristic carbon, De = 3.80 × 10−8 m2 /s at 21 ◦ C. Our values are for a
of the rate of adsorption in the region where intra-particle dif- temperature of 30 ◦ C, but fall in between those for Ds and
fusion is rate controlling. The values of rate parameters (kid,1 De .
and kid,2 ) as given in Table 3 show that ACL has highest value
of the kid,1 followed by BFA and ACC in that order but the 3.8.2. Bangham’s equation
trend is just opposite for kid,2 . The values of error functions Kinetic data can further be used to check whether pore-
are given in Table 3. They show that the Weber–Morris model diffusion is the only rate-controlling step or not in the adsorp-
shows better representation of the data than pseudo-first order tion system using Bangham’s equation [52].
kinetic model.    
C0 k0 m
log log = log + α log(t) (16)
3.8.1. Determination of diffusivity C0 − q t m 2.303V
Kinetic data could be treated by models given by Boyd et where α (<1) and k0 are constants. If the experimental data
al. [50], which are valid under the experimental conditions are represented by Eq. (16), then it is an indication that the
used. With diffusion rate controlling in the adsorption on adsorption kinetics is limited by the pore diffusion. However,
particles of spherical shape, the solution of the simultaneous Eq. (16) does not give a good fit of the experimental data,
set of differential and algebric equations leads to: indicating thereby that the diffusion of adsorbate into pores
∞  2 2  of the sorbent is not the only rate-controlling step [53]. With
61 −z π De t
F (t) = 1 − 2 exp (12) increase in the contact time, the effect of diffusion process
π z2 R2a on overall sorption could be ignored. Values of Bangham
z=1
parameters, correlation coefficient and error function values
where F(t) = qt /qe is the fractional attainment of equilibrium are given in Table 3. MPSD values are greater than 41.60
at time t, De the effective diffusion coefficient of adsorbates exposing poorer fit of the model.
in the adsorbent phase (m2 /s), Ra the radius of the adsorbent
particle assumed to be spherical (m), and z is an integer. The
3.9. Adsorption equilibrium study
BFA particles are not spherical; they may be taken as cylin-
drical particles. If one assumes that the diffusion takes place
To optimize the design of an adsorption system for
radially with diffusion in the angular and axial direction to
the removal of adsorption of adsorbates, it is important to
be negligible, one gets the solution given by Skelland [51],
establish the most appropriate correlation for the equilibrium
which after rearrangement is:
curves. Various isotherm equations have been used to

4 1 describe the equilibrium nature of adsorption. Large num-
F (t) = 1 − exp[−De bn2 t] (13) bers of researchers in the field of environmental engineering
π2 bn2
z=1 have used Freundlich and Langmuir isotherm equations
V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104 99

Table 4
Comparison of effective pore diffusivity and HSDM surface diffusivities for adsorption of phenol on activated carbon
Adsorbent Conditions De × 1010 (m2 /s) Reference
AC (coal; cecarbon GAC 40) T = 21 ◦ C 141.0 [10]
AC (coconut shell; 208C) T = 21 ◦ C 276.0 [10]
AC (wood; Pica 103) T = 21 ◦ C 380.0 [10]
AC (Coal; SP 207A) T = 21 ◦ C 368.0 [10]
AC (straw type 5; 800/200/12) T = 21 ◦ C 221.0 [10]
AC (tyre type 6; 900/100/18) T = 21 ◦ C 263.0 [10]
Granular AC pH = 3, T = 21 ◦ C 0.012a [65]
Granular AC pH = 7, T = 21 ◦ C 0.0076a [65]
Granular AC pH = 11, T = 21 ◦ C 0.0035a [65]
AC – 0.0138a [66]
BFA pH = 6.5, T = 30 ◦ C 1.052 Present work
ACC pH = 6.5, T = 30 ◦ C 0.388 Present work
ACL pH = 6.5, T = 30 ◦ C 1.800 Present work
AC: activated carbon.
a HSDM surface diffusivity.

to represent equilibrium adsorption data using activated 5.0 for Windows to minimize the deviation between calcu-
carbon–organic contaminants systems. This, despite the fact lated and experimental values.
that these equations have serious limitations on their usage,
the most popular Freundlich isotherm is suitable for highly 3.9.1. Choosing best isotherm model
heterogeneous surfaces, however, it is valid for adsorption Since each of the error functions produce a different
data over a restricted range of concentrations. For highly set of isotherm parameters, an overall optimum parame-
heterogeneous surfaces and extremely low concentrations, ter set is difficult to identify directly. Thus, a normalisa-
Henry’s law is valid. However, Freundlich equation [54] does tion of each parameter is employed in order to have a bet-
not approach Henry’s law at vanishing concentrations. The ter comparison between the parameter sets for the single
Langmuir equation [55], although follows Henery’s law at isotherm model [31]. In the normalisation processes first
vanishing concentrations, is valid for homogeneous surfaces. each error function was selected in turn and the results for
Thus, both these isotherm equations may not be suitable each parameter set were determined. Secondly, the errors
for phenol adsorption on activated carbons and BFA for the determined for a given error function were divided by the
whole range of concentrations used in the study. Temkin maximum to obtain the normalised errors for each parameter
isotherm contains a factor that explicitly takes into account set. Lastly, the normalised errors for each parameter set were
the interactions between adsorbing species and the adsorbate. summed up. Parameters as calculated for different isotherms
This isotherm assumes that (i) the heat of adsorption of all the for all phenol–adsorbent systems are tabulated in Table 5.
molecules in the layer decreases linearly with coverage due to By comparing the results of the values for the error func-
adsorbate–adsorbate interactions, and (ii) adsorption is char- tion and correlation coefficients (Table 5), similar ‘best-fit’
acterized by a uniform distribution of binding energies, up to results for phenol adsorption on ACL, ACC and BFA are
some maximum binding energy [56,57]. Radke and Prausnitz obtained. According to them, Redlich–Peterson isotherm best
[58] presented a simple equation, based on thermodynamic fitted the equilibrium data for all phenol–adsorbent systems.
ideal solution concept, for dilute solutions. The Redlich Fig. 9 presents how well the six equations fit the data for
and Peterson equation [59] and the Toth equation [60] are
three parameter-equations, often used to represent solute
adsorption data on heterogeneous surfaces. These equations
reduce to Henry’s equation at very low concentrations.
Due to experimental limitations, we could not obtain
adsorption equilibrium data at very low concentrations.
None the less, we tried to use the six isotherm equations
given by, Freundlich, Langmuir, Temkin, Redlich–Peterson,
Radke–Prausnitz and Toth to fit the experimental data for
phenol on ACL, BFA and ACC. For each system and the
isotherm equation all the experimental data were used. Using
the least-squares data reduction method, the isotherm param-
eters (two for Freundlich, Langmuir and Temkin; and three
for Redlich–Peterson, Radke–Prausnitz and Toth isotherms)
were obtained using a statistical software Statistica version Fig. 9. Equilibrium isotherms for the removal of phenol by BFA at 30 ◦ C.
100 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

Table 5 Several authors have reported Freundlich and Langmuir


Isotherm parameters for different phenol–adsorbent systems at 30 ◦ C constants for adsorption of phenol on various adsorbents. The
Adsorbent KF [(mg/g)/ 1/n R2 HYBRID MPSD Freundlich and Langmuir constants values obtained in some
(mg/l)1/n ] of these works, although under different environmental con-
Freundlicha ditions, are compared with the values obtained in the present
BFA 4.5782 0.3678 0.9945 1.6465 6.5314 work in Table 6. It may be seen that the isotherm parameters
ACC 2.4143 0.5031 0.9774 −0.1389 1.6692
ACL 7.3771 0.3146 0.99451 −0.4508 6.4218
differ widely in their values for activated carbons of differ-
ent origins. Hence one should be cautious while using these
Adsorbent qm (mg/g) KL (l/mg) R2 HYBRID MPSD values in design of adsorption systems.
Langmuirb
BFA 23.832 0.0884 0.9897 −1.3836 6.7864 3.10. Effect of temperature
ACC 30.2187 0.0291 0.9916 2.74078 9.1324
ACL 24.6458 0.2391 0.9543 3.7889 16.0369 Temperature has a pronounced effect on the adsorption
Adsorbent KT (l/mg) B1 R2 HYBRID MPSD
capacity of the adsorbents. Fig. 10 shows the plots of adsorp-
tion isotherms, qe versus Ce for phenol–BFA system at dif-
Temkinc
BFA 5.0303 0.9317 0.9965 0.1702 3.2803
ferent temperatures ranging from 25 to 45 ◦ C. It shows that
ACC 0.3023 6.5015 0.9864 0.6562 10.2645 with the increase in temperature the adsorptivity of phenol
ACL 3.4117 4.6443 0.9844 0.2722 8.2107 increases. This figure also shows that at lower adsorbate
concentrations, qe rises sharply and thereafter the increase
Adsorbent KR (l/g) aR (l/mg) β R2 HYBRID MPSD is gradual with solute concentration in the solution. Since
Redlich–Petersond sorption is an exothermic process, it would be expected that
BFA 4.4167 0.5233 0.7643 0.9976 −0.3101 3.1300 an increase in temperature of the adsorbate–adsorbent sys-
ACC 7.3820 2.5652 0.4384 0.9994 0.0242 1.5865
ACL 57.3165 7.1016 0.7078 0.9945 −0.3903 6.3051
tem would result in decreased sorption capacity. However, if
the adsorption process is controlled by the diffusion process
Adsorbent kRP [(mg/g)/ KRP (l/g) P R2 HYBRID MPSD (intraparticle transport-pore diffusion), the sorption capacity
(mg/l)(1/P) ] will show an increase with an increase in temperatures. This
Radke–Prausnitze is basically due to the fact that the diffusion process is an
BFA 4.1629 13013.63 2.414 0.9955 −6.6353 8.9522 endothermic process [24]. With an increase in temperature,
ACC 1.8300 9882.36 1.654 0.6806 −8.7421 12.2868 the mobility of the phenolate ions increases and the retarding
ACL 6.9933 9953.78 2.823 0.9913 −4.8782 9.4707
forces acting on the diffusing ions decrease, thereby increas-
Adsorbent qe∞ (mg/g) KTh Th R2 HYBRID MPSD ing the sorptive capacity of adsorbent. As has been shown
[(mg/l)Th ] earlier, the diffusion of adsorbate into pores of the sorbent is
Tothf not the only rate-controlling step, and the diffusion process
BFA 53.0560 1.6612 0.3501 0.9971 −0.3729 3.5018 could be ignored with adequate contact time. Therefore, the
ACC 47.7308 7.9121 0.5878 0.9968 1.4123 5.6561 increase in sorption capacity with an increase in temperature
ACL 80.6903 0.7791 0.2309 0.9914 0.4199 6.5198 may be attributed to chemisorption. The increase in phenol
1/n
a qe = K F Ce . sorption capacity of the carbonaceous adsorbents with the
qm K L C e
b qe = 1+K L Ce
. increase in temperature has also been reported by other inves-
c qe = B1 ln KT + B1 ln Ce . tigators [38,61]. These investigators used activated carbon
d qe = KR Ce β .
1+aR Ce and activated date pits, respectively, for phenol adsorption.
e 1
qe = KRP Ce + k C1/P
1 1
. Isotherm parameters for six isotherms at different temper-
RP e

qe Ce
f qe = . atures for phenol–BFA system are given in Table 7 along with
1/Th
[KTh +CeTh ]

phenol–BFA system. Following equilibrium relationships are


recommended for the three adsorbate–adsorbent systems.

57.3165Ce
qe = (phenol–ACL system) (17)
1 + 7.1016Ce0.7078

7.3820Ce
qe = , (phenol–ACC system) (18)
1 + 2.5652Ce0.4348

4.41670Ce
qe = , (phenol–BFA system) (19) Fig. 10. Equilibrium adsorption isotherms at different temperature for
1 + 0.5233Ce0.7643 phenol–BFA system. pH0 : 6.5; BFA dosage: 10 g/l.
V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104 101

Table 6
Freundlich and Langmuir constants for adsorption of phenol on various adsorbents
Adsorbent KF [(mg/g)/(mg/l)1/n ] 1/n qm (mg/g) KL (l/mg) Reference
AC – – 216.43 0.0690 [9]
Na–Y zeolites – – 75.28 0.0096 [9]
Ni/Na–Y zeolites – – 84.69 0.00106 [9]
AC (coal; cecarbon GAC 40) 36.902 0.293 – – [10]
AC (coconut shell; 208C) 32.083 0.278 – – [10]
AC (wood; pica 103) 9.9906 0.436 – – [10]
AC (coal; SP 207A) 29.300 0.245 – – [10]
AC (straw type 5; 800/200/12) 60.871 0.103 – - [10]
AC (tyre type 6; 900/100/18) 26.436 0.272 – – [10]
AC 39.97 0.188 1.1989 3.4443 [11]
AC (pinewood-based) 23.33 0.256 240.6 0.034 [12]
Dried activated sludge 15.10 0.45 236.8 0.0146 [13]
Amnerlite XAD-4 resin 0.259 1.639 – – [14]
NJ-8 0.808 2.356 – – [14]
Modified-bentonite 3.64 0.464 – – [15]
A-pillared bentonite 1.81 0.605 – – [15]
CTAB-bentonite 1.48 0.642 – – [15]
Thermal bentonite 1.30 0.652 – – [15]
TiO2 surface 0.676 0.919 – – [16]
Bentonite 0.2013 4.840 – – [17]
Peat 0.0362 1.570 – – [17]
Fly ash 19.8 0.26 [17]
Dried aerobic activated sludge 1.28 0.820 194.2 0.004 [18]
Bentonite 0.100 0.473 1.712 0.0141 [19]
Amnerlite XAD-16 resin 0.0748 0.8720 1.5029 0.0511 [20]
AC 37.0 0.17 309.7 0.0533 [21]
Filtrasorb (F-400) 36.3 0.60 205.1 0.0420 [21]
HiSiv 1000 (zeolite-Y structure) 0.047 0.70 319.0 0.00055 [21]
Activated carbon cloth (ACC) – – 402.74 0.0253 [22]
Pd (1%)/ACC – – 400.86 0.0176 [22]
Pd (9.2%)/ACC – – 376.40 0.0156 [22]
BFA 4.5782 0.3678 23.832 0.0884 Present work
ACC 2.4143 0.5031 30.2187 0.0291 Present work
ACL 7.3771 0.3146 24.6458 0.2391 Present work
AC: activated carbon; A-pillared: aluminum-hydroxypolycation as a pillaring agent; CTAB: cetyltrimethyl ammonium bromide.

the values of the correlation coefficient and the error func- isotherm, from which Hst,0 = 503.88 kJ/kg and
tions values. On increasing the temperature from 25 to 45 ◦ C, S◦ = 1798.14 kJ/kg K have been obtained. For signif-
the value of the maximum adsorption capacity, qm increases icant adsorption to occur, the free energy changes of
from 23.33 to 26.09 mg/g confirming endothermic nature adsorption, G◦ , must be negative. The thermodynamics
of overall-sorption process. Table 7 clearly reinforces ear- relation between G◦ , H◦ and S◦ suggests that either
lier observation at 30 ◦ C that the Redlich–Peterson isotherm (i) H◦ is positive and S◦ is positive and that the value
shows slightly better fit than all other isotherms. However,
these values have been used to define and determine the
isosteric heat of adsorption Hst,0 that corresponds to zero
surface coverage (qe = 0).

3.11. Estimation of thermodynamic parameters

S◦ can be determined by the slope of the linear Van’t


Hoff plot i.e. as ln K versus (1/T), using equation:
 
◦ d ln K
H = R (20)
d(1/T )
H◦ obtained here corresponds to isosteric heat of adsorp-
tion (Hst,0 ) with zero surface coverage (i.e. qe = 0) [62]. Fig. 11. Van’t Hoff plot of adsorption equilibrium constant, K using
Fig. 11 shows the Van’t Hoff’s plot for Redlich–Peterson Redlich–Peterson isotherm.
102 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

Table 7
Isotherm parameters for phenol–BFA system at different temperatures
Temperature (◦ C) KF [(mg/g)/(mg/l)1/n ] 1/n R2 HYBRID MPSD
Freundlich
25 3.6268 0.3993 0.9976 1.8853 8.2170
30 4.5782 0.3678 0.9945 1.6465 6.5314
35 5.2582 0.3528 0.9955 2.3981 7.9805
40 5.9182 0.3457 0.9951 2.9751 9.2871
45 7.1621 0.3146 0.9958 3.8357 12.3222

Temperature (◦ C) qm (mg/g) KL (l/mg) R2 HYBRID MPSD


Langmuir
25 23.331 0.0586 0.9846 −0.2849 3.6270
30 23.832 0.0884 0.9897 −1.3836 6.7864
35 24.645 0.1052 0.9901 −1.5441 6.7622
40 25.578 0.1282 0.9873 −1.7707 7.2070
45 26.088 0.1706 0.9811 −1.5823 6.9270

Tempeature (◦ C) KT (l/mg) B1 R2 HYBRID MPSD


Temkin
25 5.1695 0.5964 0.9912 0.3360 5.5258
30 5.0303 0.9317 0.9965 0.1702 3.2803
35 5.0602 1.1969 0.9979 0.3570 3.4119
40 5.1588 1.5278 0.9973 0.4322 4.0064
45 4.9700 2.3986 0.9963 0.5558 4.9311

Tempeature (◦ C) KR (l/g) aR (l/mg) β R2 HYBRID MPSD


Redlich–Peterson
25 3.8696 0.6336 0.7083 0.9991 −0.0712 2.7842
30 4.4167 0.5233 0.7643 0.9976 −0.3101 3.1300
35 5.5747 0.6003 0.7734 0.9995 −0.1202 1.4252
40 7.5220 0.7681 0.7697 0.9992 −0.1371 1.6193
45 13.438 1.3349 0.7661 0.9993 −0.0815 1.2575

Temperature (◦ C) kRP [(mg/g)/(mg/l)(1/P) ] KRP (l/g) P R2 HYBRID MPSD


Radke–Prausnitz
25 3.0492 9938.99 2.1168 0.9945 −8.9098 11.4630
30 4.1629 13013.63 2.4143 0.9955 −6.6353 8.9522
35 4.6378 9210.27 2.3956 0.9948 −8.3398 10.7482
40 5.4111 9460.33 2.5049 0.9948 −5.4963 7.6947
45 6.6894 12097.61 2.7991 0.9913 −5.8189 8.2302

Temperature (◦ C) qe∞ (mg/g) KTh [(mg/l)Th ] Th R2 HYBRID MPSD


Toth
25 63.677 1.916 0.3288 0.9961 0.1991 3.1622
30 53.056 1.6612 0.3501 0.9971 −0.3729 3.5018
35 50.353 1.5528 0.3626 0.9992 −0.1845 1.8621
40 53.566 1.3541 0.3470 0.9989 −0.3067 2.0571
45 69.532 0.9161 0.2660 0.9990 −0.2052 1.8865

of TS is much larger than H◦ , or (ii) H◦ is negative water molecule for their adsorption and this results in the
and S◦ is positive or that the value of H◦ is more than endothermicity of the adsorption process. Therefore, the
TS. Phenol adsorption is endothermic in nature, giving a Hst,0 will be positive. The positive value of S◦ suggests
positive value of H◦ . Hence, S◦ has to be positive and increased randomness at the solid/solution interface with
that the positive value of TS has to be larger than H◦ . some structural changes in the adsorbate and adsorbent and
The positive Hst,0 value confirms the endothermic nature an affinity of the BFA towards phenol. Also, positive S◦
of the overall-sorption process. The adsorption process in value corresponds to an increase in the degree of freedom
the solid–liquid system is a combination of two processes: of the adsorbed species [63]. G◦ values were negative
(a) the desorption of the molecules of solvent (water) indicating that the sorption process led to a decrease in
previously adsorbed, and (b) the adsorption of adsorbate Gibbs free energy. Negative G◦ indicates the feasibility
species. The phenolate ions have to displace more than one and spontaneity of the adsorption process.
V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104 103

to be zero and Hsol of the adsorbate in the solvent may be


calculated from heats of formation data. Since Hsol of phe-
nol is 116.49 kJ/kg and Hw should be negligible, Hst,net as
calculated from Eq. (23) is equal to 890.41, 713.59, 610.15,
536.77 and 479.84 kJ/kg for qe = 5, 10, 15, 20 and 25 mg/g,
respectively. Since heat of vapourisation of phenol is given
as λvap = 478.48 kJ/kg, therefore, Hst,net lies between 1.00
and 1.86 times λvap .

4. Conclusions

The present study shows that the bagasse fly ash (BFA)
Fig. 12. Adsorption isosters for determining isosteric heat of adsorption. is an effective adsorbent for the removal of phenol from
aqueous solution. Higher percentage of phenol removal by
3.11.1. Isosteric heat of adsorption BFA, ACC and ACL was possible provided that the C0 in the
Apparent isosteric heat of adsorption (Hst,a ) at constant solution was low. Optimum conditions for phenol removal
surface coverage (qe = 5, 10, 15, 20, 25 mg/g) is calculated were found to be pH0 ≈ 6.5, adsorbent dose ≈10 g/l of solu-
using Clausius–Clapeyron equation [64]. tion for phenol concentration up to 50 mg/l. The equilibrium
d ln Ce −Hst,a between the adsorbate in the solution and on the adsorbent
= (21) surface was practically achieved in 5 h. Adsorption kinetics
dT RT 2
was found to follow second-order rate expression with
d ln Ce
Hst,a =R (22) initial sorption rate being highest for adsorption on BFA.
d(1/T ) qe Equilibrium adsorption data for phenol on all the adsorbents,
viz., ACL, ACC and BFA were best represented by the
For this purpose, the equilibrium concentration (Ce ) at con-
Redlich–Peterson isotherm. Adsorption of phenol on BFA is
stant amount of adsorbed phenol is obtained from the adsorp-
favourably influenced by an increase in the temperature of
tion isotherm data at different temperatures. Hst,a is cal-
the operation. This shows that the overall sorption process is
culated from the slope of the ln Ce versus (1/T) for differ-
controlled by intraparticle diffusion of phenol. The negative
ent amount of phenol adsorption on BFA (Fig. 12). The
value of G◦ads indicate spontaneous adsorption of phenol on
Hst,a varies with the surface loading (Fig. 13), indicating
BFA. The values of isosteric heat of adsorption varied with
that the BFA has an energetically heterogeneous surface.
surface coverage implying an energetically heterogeneous
The variation in Hst,a with surface loading can be also
surface of BFA. Overall, BFA showed excellent adsorptive
be attributed to the possibility of having lateral interactions
characteristics for the removal of phenol from wastewater
between adsorbed phenolate ions.
samples and could be used as a very good adsorbent due
The apparent isosteric heat of adsorption in aqueous
to its high carbon content and the presence of silica and
adsorption Hst,a , is a resultant of net isosteric heat of adsorp-
alumina.
tion Hst,net , heat of solution Hsol , and heat of adsorption
of water, Hw , i.e.:
References
Hst,a = Hst,net − Hsol − fHw (23)
where, f is the exchange of number of moles of water per mole [1] W. Kujawski, A. Warszawski, W. Ratajczak, T. Porebski, W. Capaa,
I. Ostrowska, Sep. Purif. Technol. 40 (2004) 123.
of adsorbate at the adsorption site. Hw is normally assumed [2] H. Li, M. Xu, Z. Shi, B. He, J. Colloid Interf. Sci. 271 (2004)
47.
[3] J.K. Fawell, S. Hunt, Environmental Toxicology: Organic Pollutants,
Halsted Press, John Wiley & Sons, NY, 1988, 398.
[4] World Health Organization. International Standards for Drinking
Water, Geneva, 1963, p. 40.
[5] J.S. Zogorski, S.D. Faust, Equilibria of adsorption of phenols
by granular activated carbon, in: A.J. Rubin (Ed.), Chemistry of
Wastewater Technology, Ann Arbor Science Publishers, Ann Arbor,
Michigan, 1978 (Chapter 9).
[6] C.T. Hsieh, H. Teng, J. Colloid Interf. Sci. 230 (2000) 171.
[7] J.T. Paprowicz, Environ. Technol. 11 (1990) 71.
[8] M.M. Swamy, I.D. Mall, B. Prasad, I.M. Mishra, Pollut. Series 16
(1997) 170.
[9] B. Okolo, C. Park, M.A. Keane, J. Colloid Interf. Sci. 226 (2000)
308.
Fig. 13. Variation of Hst,a with respect to surface loading. [10] M. Streat, J.W. Patrick, M.J.C. Perez, Water Res. 29 (1995) 467.
104 V.C. Srivastava et al. / Colloids and Surfaces A: Physicochem. Eng. Aspects 272 (2006) 89–104

[11] A.R. Khan, T.A. Al-Bahri, A. Al-Haddad, Water Res. 31 (1997) [38] F. Banat, A.A. Sameer, A.M. Leema, Chem. Eng. Technol. 27 (2004)
2102. 80.
[12] R.L. Tseng, F.C. Wu, R.S. Juang, Carbon 41 (2003) 487. [39] A.P. Terzyk, J. Colloid Interf. Sci. 268 (2003) 301.
[13] Z. Aksu, J. Yener, Process Biochem. 33 (1998) 649. [40] D.A. Mooney, F. Muller-Plathe, K. Kreneg, Chem. Phys. Lett. 294
[14] A. Li, Q. Zhang, G. Zhang, J. Chen, Z. Fei, F. Liu, Chemosphere (1998) 135.
47 (2002) 981. [41] H.S. Fogler, Elements of Chemical Reaction Engineering, 3rd ed.,
[15] S. Al-Asheh, F. Banat, L. Abu-Aitah, Sep. Purif. Technol. 33 (2003) Prentice-Hall PTR, 1998.
1. [42] S. Lagergren, Ksver. Veterskapsakad. Handl. 24 (1898) 1.
[16] D. Robert, S. Parra, C. Pulgarin, A. Krzton, J.V. Weber, App. Surface [43] Y.S. Ho, G. McKay, Process Biochem. 34 (1999) 451.
Sci. 167 (2000) 51. [44] V.J.P. Poots, G. McKay, J.J. Healy, J. Water Pollut. Control Fed. 50
[17] T. Viraraghavan, F.D.M. Alfaro, J. Hazard Mater. 57 (1998) 59. (1978) 926.
[18] Z. Aksu, D. Akpkinar, Sep. Purif. Technol. 21 (2000) 87. [45] G. McKay, M.S. Otterburn, A.G. Sweeney, Water Res. 14 (1980)
[19] F.A. Banat, B. Al-Bashir, S. Al-Asheh, O. Hayajneh, Environ. Pollut. 15.
107 (2000) 391. [46] S.J. Allen, G. Mckay, K.Y.H. Khader, Environ. Pollut. 56 (1989) 39.
[20] K. Abburi, J. Hazard Mater. B105 (2003) 143. [47] K. Kannan, M.M. Sundaram, Dyes Pigments 51 (2001) 25.
[21] N. Roostaei, F.H. Tezel, J. Environ. Manage. 70 (2004) 157. [48] J. Crank, The Mathematics of Diffusion, vol. 84, 1st ed., Oxford
[22] G.E. Shter, Y. Shindler, Y. Matatov-Meytal, G.S. Grader, M. Shein- Clarendon Press, London, 1965.
tuch, Carbon 40 (2002) 2547. [49] H.M. Asfour, M.M. Nassar, O.A. Fadali, M.S. El-Geundi, J. Chem.
[23] T. Vermeulen, Ind. Eng. Chem. 45 (1953) 1664. Technol. Biotechnol. A 35 (1985) 28.
[24] W.J. Weber Jr., Physicochemical Processes for Water Quality Con- [50] G.E. Boyd, A.W. Adamson, L.S. Meyers, J. Am. Chem. Soc. 69
trol, Wiley Interscience, New York, 1972, p. 206. (1947) 2836.
[25] W.J. Weber Jr., J.C. Morris, J. Sanitary Eng. Div. ASCE 89 (SA2) [51] A.H.P. Skelland, Diffusional Mass Transfer, Wiley, NY, 1974.
(1963) 31. [52] C. Aharoni, S. Sideman, E. Hoffer, J. Chem. Technol. Biotechnol.
[26] IS 1350 (Part I), Methods of Test for Coal and Coke Proximate 29 (1979) 404.
Analysis, Bureau of Indian Standards, Manak Bhawan, New Delhi, [53] E. Tutem, R. Apak, C.F. Unal, Water Res. 32 (1998) 2315.
India, 1984. [54] H.M.F. Freundlich, J. Phys. Chem. A 57 (1906) 385.
[27] V.C. Srivastava, M.M. Swamy, I.D. Mall, B. Prasad, I.M. Mishra, J. [55] I. Langmuir, J. Am. Chem. Soc. 40 (1918) 1361.
Chem. Technol. Biotechnol. (2005), Communicated. [56] M.I. Temkin, V. Pyzhev, Acta Physiochim. URSS 12 (1940) 327.
[28] J.F. Porter, G. McKay, K.H. Choy, Chem. Eng. Sci. 54 (1999) 5863. [57] Y. Kim, C. Kim, I. Choi, S. Rengraj, J. Yi, Environ. Sci. Technol.
[29] D.W. Marquardt, J. Soc. Ind. Appl. Math. 11 (1963) 431. 38 (2004) 924.
[30] I.D. Mall, V.C. Srivastava, N.K. Agarwal, I.M. Mishra, Colloids Surf. [58] C.J. Radke, J.M. Prausnitz, AIChE J. 18 (1972) 761.
A: Physicochem. Eng. Aspects 264 (2005) 17. [59] O. Redlich, D.L. Peterson, J. Phys. Chem. 63 (1959) 1024.
[31] Y.C. Wong, Y.S. Szeto, W.H. Cheung, G. McKay, Process Biochem. [60] J. Toth, Acta Chem. Acad. Hung. 69 (1971) 311.
39 (2004) 693. [61] P.R. Vijayalakshmi, V.J. Raksh, J. Rodriguez, J. Chem. Technol.
[32] H.M. Stenzel, Chem. Eng. Prog. 89 (1993) 36. Biotechnol. 71 (1998) 173.
[33] Powder Diffraction File (PDF), JCPDS International Centre for [62] M. Suzuki, T. Fujii, AIChE J. 28 (1982) 380.
Diffraction Data, Swarthmore, PA, USA, 1979. [63] C. Raymon, Chemistry: Thermodynamic, McGraw-Hill, Boston,
[34] H.A. Elliott, C.P. Huang, Water Res. 15 (1981) 849. 1998, p. 737.
[35] P. Khanna, S.K. Malhotra, Indian J. Environ. Health 19 (1977) 224. [64] D.M. Young, A.D. Crowell, Physical Adsorption of Gases, Butter-
[36] C. Huang, C.T. Stumm, J. Colloid Interf. Sci. 43 (1973) 409. worths, London, 1962, p. 426.
[37] K.K. Panday, G. Prasad, V.N. Singh, Environ. Technol. Lett. 50 (7) [65] N.S. Abuzaid, G.F. Nakhla, J. Hazard Mater. 49 (1996) 217.
(1986) 547. [66] V.K.C. Lee, G. McKay, Chem. Eng. J. 98 (2004) 255.

You might also like