Sei sulla pagina 1di 16

791

Breakthrough Curve Analysis for Fixed-Bed Adsorption of Azo Dyes Using


Novel Pine ConeDerived Active Carbon
Mohammad R. Samarghandi1, Mahdi Hadi2 and Gordon McKay3,4,* (1) Department of
Environmental Health Engineering, School of Public Health, Center for Health Research, Hamadan University of Medical
Sciences, P.O. Box No. 4171, Hamadan, Iran. (2) Center for Water Quality Research (CWQR), Institute for Environmental
Research (IER), Tehran University of Medical Sciences, Tehran, Iran. (3) Division of Sustainable Development, College
of Science, Engineering and Technology, Hamad Bin Khalifa University, Qatar Foundation, Doha, Qatar. (4) Department
of Chemical and Biomolecular Engineering, Hong Kong University of Science and Technology, Clearwater Bay, New
Territory, Hong Kong, SAR.

(Received date: 20 July 2014; Accepted date: 12 October 2014)

ABSTRACT: A novel activated carbon has been produced from pine cone, a
sustainable resource, by phosphoric activation at 900 C. The BET surface area
of the activated carbon was 869 m2/g and the methylene blue isotherm area was
734 m2/g. The adsorption of two single-component acid azo dyes, Acid Blue 113
(AB113) and Acid Black 1 (AB1), onto this activated carbon was studied using
fixed-bed adsorption. Breakthrough curves and equilibrium isotherms were
obtained for the adsorption of these two dyes onto the prepared active carbon.
The highest adsorption capacities achieved for AB113 and AB1 were 485 and
286 mg dye/g carbon, respectively. Two-column breakthrough curve models
were applied to correlate the experimental adsorption breakthrough curves for
each dye: the Thomas model, based on solid-phase internal diffusion, has been
applied with considerable success; and the saturation isotherm model approach
of Michaels was used with a reasonable degree of success.

1. INTRODUCTION

Increasing government legislative requirements for effluent discharges emphasize the need for
effective treatment methods. In general, coloured organic compounds constitute only a small
fraction of the total organic load in wastewaters; however, their high degree of colour is easily
detectable and affects the aesthetic value of rivers and receiving waters. In addition, the presence
of colour reduces the photosynthetic process and certain dyes, particularly some azo dyes, are
carcinogenic and toxic (Gottlieb et al. 2003; Mathur et al. 2005). Large volumes of aqueous
effluent are discharged by the clothing industries, including textiles, leather and dyeing, and 70%
of commercial dyes produced are azo dyes (Carliell et al. 1998).
Adsorption is gaining prominence as an effective and cost-effective method for the removal of
pollutants from wastewaters (McKay 1995). The mechanism of adsorption is based on the ability of
porous solids (i.e. adsorbents) to concentrate and hold solute molecules (e.g. dyestuff) on their
surfaces. In the past thirty years, and particularly in the last ten years, extensive research has been
carried out to identify suitable adsorbents for dyes. These include adsorption of dyes on carbon
(Davies et al. 1973; McKay 1982a, b), peat (Ho and McKay 1999; Poots et al. 1976), waste tyre
carbon (Mui et al. 2010), wood (Ho and McKay 1998; McKay and Ho 1998) and Fullers earth
(McKay et al. 1985). A number of pith-based materials have also been tested as adsorbents, including
*Author to whom all correspondence should be addressed. E-mail: kemckayg@ust.hk (G. McKay).
792 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

orange peel (Namasivayam et al. 1996) and bagasse (McKay et al. 1997a, b). Activated carbon is the
most commonly used adsorbent in the treatment of water, municipal wastewater and organic industrial
wastewaters, because of its ability to adsorb a wide variety of organic compounds. However, a number
of studies are carried out to identify cheaper adsorbents (McKay 1995; Snoeyink 1990).
A wide variety of activated carbons have been prepared from agricultural and wood wastes,
such as bamboo (Wu et al. 1999), cassava peel (Rajeshwarisivaraj et al. 2001), bagasse (Valix
et al. 2004), coir pith (Namasivayam and Kavitha 2002), banana pith (Kadirvelu et al. 2003), date
pits (Banat et al. 2003), chickpea (Nadeem et al. 2006), pinewood (Oktem et al. 2012) and pecan
nut shell (Johns et al. 1999).
In this study, an activated carbon has been produced from pine cone (pine cone activated carbon
or PCAC) as pine trees are widely available throughout the world in large numbers in national
parks. Thus, ground pine cone may be abundantly available as an unused natural waste resource,
and it would be worthwhile to develop a low-cost adsorbent from this waste. The activation agent
used in this research was phosphoric acid because it has already been applied very successfully to
similar materials, such as coals and hardwoods (Toles et al. 2000). Acid-activated carbons from
almond shells had improved physical, chemical and adsorptive properties with reduced cost of
production (Jagtoyen and Derbyshire 1998; Molina-Sabio et al. 1995).
However, the major problem in all these previous studies is that they are either equilibrium or
batch-contact systems. Only limited studies are available on fixed-bed adsorption using column
design models, and none of these has studied adsorption using activated carbon synthesized from
modified agricultural waste materials. Furthermore, it is often difficult to apply batch
experimental data to fixed-bed adsorbents because isotherms cannot provide accurate scale-up
data in fixed-bed systems. This is attributed to the following reasons: a flow column is not at
equilibrium, uneven flow patterns occur in fixed beds and the problems of recycling and
regeneration cannot be studied. Therefore, it is necessary to carry out flow tests using columns
before obtaining design models. The problem in designing columns is that it is difficult to predict
how long they will last before regeneration or replacement becomes necessary.
The most widely applied simplified design model for this purpose is the bed-depth service time
(BDST) model (Bohart and Adams 1920; Hutchins 1973; McKay and Bino 1990; Walker and
Weatherley 1997).
The BDST model is based on a surface chemical reaction and not diffusion, which is the
predominant mechanism in dye adsorption. No predictive equations were developed or applied to
analyze the fixed-bed applications and few simplified diffusion models are available to test the
performance of dye-adsorption columns. A fixed-bed column study is important to predict the
column breakthrough or the shape of the wavefront, which determines the operation lifespan of
the bed, and it is essential to study the behaviour of the continuous flow system. Adsorption
isotherms from batch studies may not provide accurate scale-up information in the column
operation systems (McKay 1995).
The main aim of this study was to produce and evaluate the adsorption efficiency of granular
local ground PCAC under continuous flow condition to remove two azo dyes [Acid Blue 113
(AB113) and Acid Black 1 (AB1)] from aqueous solutions.
The adsorption characteristics of AB113 and AB1 dyes in the batch system and fixed-bed
columns of granular PCAC were investigated. The breakthrough curves were measured and
characterized for the two dyes. The Michaels breakthrough curve model was applied to the fixed-
bed results (Michaels 1952). Furthermore, the Thomas (Lee et al. 2004; Thomas 1944) diffusion
model, which offers a relatively simple approach and rapid prediction of adsorbent performance,
was applied for modelling the adsorption of dyes on the activated carbon columns.
Fixed-Bed Adsorption of Azo Dyes Using Novel Pine ConeDerived Active Carbon 793

The effects of influent dye concentration, flow rate and bed depth on the adsorption of azo dyes
by the PCAC bed column were investigated. Based on the data obtained from the batch isotherm
studies, the theoretical breakthrough curves were predicted and compared with experimental
breakthrough curves.

2. MATERIALS AND METHODS


2.1. Equipment and Dye Concentration Measurement

Materials were weighed using an analytical balance with precision of 0.0001 g (Sartorius
ED124S). The materials were dried in an electric oven (PARS TEB) and carbonized in a muffle
furnace (Exiton). The pH values of the solutions were measured using a digital pH meter (model
Sartorius Professional Meter PP-50). The dye solutions were stirred using an inductive stirring
system (Oxitop IS 12) within a WTW-TS606/2-i incubator. The samples were centrifuged (301
Sigma Centrifuge). The dye concentrations in the samples were measured spectrophotometrically
using UV-1700 PharmaspecShimadzo spectrophotometer.

2.2. Raw Materials

Dried pine cone was used as the raw material to produce the adsorbent. The pine cones were
collected from the Mardom Park in front of Hamadan University of Medical Sciences (Hamadan,
Iran). AB113 [1-naphthalenesulophonic acid, 8-(phenylamino)-5-4-(3-sulphophenyl)], a disazo-type
dye containing two azo groups, and AB1 [(4-amino-5-hydroxy-3-(p-nitrophenylazo)-6-(phenylazo)-
2,7-naphthalenedisulphonic], a sulphonated azo dye having a naphthyl radical containing a hydroxyl
functional group adjacent to the azo bond and a phenyl radical adjacent (unsubstituted) to the azo
bond were used in this study. The dyes were obtained from Alvansabet Dyestuff and Textile
Auxiliary Manufacturer Company in the west of Iran. Their structures are presented in Figure 1.

N N N N N
H

NaO3S
SO3Na

AB113

NH2 OH
N N N N

NO2 NaO3S SO3Na

AB1
Figure 1. Structure of AB113 and AB1.
794 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

2.3. Adsorbent Preparation

Pine cone was the source material for the adsorbent used in this study (i.e. PCAC). The PCAC
was produced by exposing the raw pine cones to a thermalchemical process. First, the pine
cones were crushed and washed with hot water and then dried at 100 C in an oven overnight.
The crushed sample (50 g) was mixed with a pre-determined volume of phosphoric acid
(concentration, 95% wt/wt in the mass ratio of 1:10). This mixture was transferred to a stainless
steel tube (50-mm diameter and 250-mm long). The tube was place in a muffle furnace, which
was programmed to gradually reach up to 900 C within 3 hours. This temperature was
maintained for 1 hour, and then the tube and carbon were gradually cooled down to room
temperature. The end product was repeatedly washed with hot distilled water until achieving a
pH of approximately 6.9. Then, the washed sample was again dried at 120 C in an oven
overnight. The final sample was then ground in a household-type blender and passed through a
series of sieves (20, 30, 40 and 50 U.S. Standard mesh). A mixture of the residuals on the sieves
30, 40 and 50 was collected and stored in an air-tight bottle. This is used as the adsorbent.
The average adsorbent particle size was 0.5 mm.

2.4. Adsorbent Properties

The specific surface area of the produced PCAC was obtained by determining the optimal
concentration of methylene blue dye adsorbed onto the PCAC adsorbent at a constant
temperature of 20 C. The test for methylene blue adsorption on the prepared activated
carbon was performed using a batch process by mixing 0.30 g of adsorbent in stoppered
conical flasks with 100 ml of methylene blue solutions of concentration ranging from 100 to
1000 mg l1 at a pH of 7.5 0.2. For all the concentration ranges studied (1001000 mg l1),
the mixture was magnetically stirred at a constant revolution for 3 days, which is more than
sufficient to reach equilibrium. Absorbance measurements were performed on methylene
blue solutions at 660 nm to determine the equilibrium concentration and the calculated
surface area was 734 m2 g1.
The capacity of the adsorbent for the adsorption of dye can be evaluated through iodine
adsorption from aqueous solutions using test conditions referred to as the iodine number
determination. This indicates the relative activation level of the adsorbent and the surface area
available for micropores. Usually adsorbents with a high iodine number have a high surface
area and are suitable for adsorbing small compounds. The iodine number was measured
according to the standard procedure [ASTM D4607-94 (ASTM 1999)] using 0.1N
standardized iodine solution. Sample volumes of 100 mL of the iodine solution were treated
with 0.6, 0.9 and 1.2 g of the different samples. After equilibrium, the remaining iodine in the
supernatants was titrated with 0.1N sodium thiosulphate solution. The iodine number was
reported as the amount of iodine adsorbed per gram of adsorbent at a residual iodine
concentration of 0.02N. The calculated iodine number value was 483.5 mg g1.
The apparent density was calculated by filling a calibrated cylinder with a given weight of
activated carbon until a minimum volume is achieved. This density was referred as tapping or
bulk density of adsorbent (as the minimum volume is calculated by tapping the cylinder).
To obtain real density values, a pycnometer method was used. In this method, the pycnometer
was filled with the prepared activated carbon and a solvent (methanol) was added to fill the
void. The weight at each step was then determined. The apparent and real density values were
equal to 0.954 and 1.599 g cm3, respectively.
Fixed-Bed Adsorption of Azo Dyes Using Novel Pine ConeDerived Active Carbon 795

The pore volume and the porosity were determined using a volumetric method. In this
method, the calibrated cylinder was filled with a volume (V1) of activated carbon (mass m1)
and a solvent (methanol) until volume V2 (total mass m2) is reached. Knowing the density of
the solvent, total porosity volume (0.407 cm3 g1) and the porosity (40.3%) of the adsorbent
were easily calculated.
The nitrogen BET surface area was measured as 734 m2/g.

2.5. Dye Concentration Measurement and Equilibrium Experiment

An accurately weighed quantity of dye was dissolved in distilled water to prepare a stock solution
(500 mg l1). Experimental solutions of the desired concentrations were obtained by successive
dilutions. Spectrophotometric scanning of a dilute dye solution was performed and the maximum
wavelengths of adsorption (lmax) for AB113 and AB1 were identified as 574 and 622 nm,
respectively. The calibration curves for AB113 and AB1 were linear from 0.125 to 100 mg l1
(R2 = 0.999) and 0.062 to 100 mg l1 (R2 = 0.999), respectively. The adsorption experiments were
carried out in a batch process. Adsorption processes were studied in 250-ml Erlenmeyer flasks
within an incubator container. Solutions of synthetic dyes were prepared by dissolving dyes in
distilled water to produce a solution of 150 mg l1 for each dye. To determine the equilibrium time
of adsorption of AB113 and AB1 onto PCAC, accurate amounts of PCAC (0.12 g) were added to
two Erlenmeyer flasks (volume = 250 cm3) containing 250 ml of each dye solution (150 mg l1).
The contents of the flasks were mixed using a magnetic stirrer and 1-ml samples were obtained at
regular intervals. The equilibrium times for AB113 and AB1 were determined as 250 and 167
hours, respectively. For each dye solution, accurately weighed amounts of PCAC adsorbent
(0.020.20 g in 0.02-g intervals) were added to each flask containing 250 ml of the respective dye
solution (150 mg l1) at pH 7.4 0.2 for AB113 and 7.0 0.2 for AB1 dyes. The contents of the
Erlenmeyer flasks were thoroughly mixed for 250 and 167 hours for AB113 and AB1, respectively
at 20.5 1 C using magnetic stirrers at constant revolution. A 5-ml sample was taken out after the
equilibrium time and centrifuged at 3800 rpm for 5 minutes. The dye concentration in the samples
was then measured spectrophotometrically.
The amount of dye adsorbed onto the adsorbent was calculated using equation (1) as follows:
(C0 Ce )
qe = (1)
m
All experiments were performed in triplicate, and the mean values are reported.

2.6. Experimental Studies

Continuous-flow adsorption experiments were conducted in a glass column (2-cm internal


diameter and 20-cm length). To avoid entrapping of air bubbles inside the PCAC particles, the
particles were soaked for 24 hours before packing into the column. The column was packed with
PCAC (particle size, 297595 mm) between two supporting layers of glass wool to prevent the
floating of adsorbent particles. The schematic diagram of the column is presented in Figure 2. The
column was washed with distilled water for 30 minutes at a constant flow rate before loading
the dye solution. The column studies were carried out to evaluate the effects of initial dye
concentration, the flow rate and carbon-bed depth.
The dye solution was pumped into the column in an upflow mode using a peristaltic pump with
the volumetric flow rate of 8.07 ml/minute for studying the effects of different bed depths (10, 15
796 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

Glass wool

5cm
Sampling points
GAC

5cm
fixed-bed
column

10cm
Glass wool

Peristaltic pump Effluent wastewater

Influent wastewater

Figure 2. Schematic diagram of the column.

and 20 cm). The initial concentration of the dye was 150 mg/l. The column was also loaded with
higher (200 mg/l) and lower (100 mg/l) dye concentrations. The bed depth and the flow rate in
this case were 10 cm and 8.07 ml/minute, respectively. The effect of dye-solution flow rate was
also studied using an influent dye concentration of 150 mg/l and bed depth of 10 cm at flow rates
of 10 and 6.1 ml/minute.
The mass of PCAC in the column was 9, 14.5 and 19 g at depths of 10, 15 and 20 cm,
respectively. Samples were collected at regular intervals from sampling points and centrifuged at
3800 rpm for 5 minutes. The dye concentrations in the samples were measured
spectrophotometrically. The pH value of dye solution in all experiments was adjusted by adding
0.1 mol/l nitric acid or NaOH. The pH value was maintained at 7.5 and the temperature was
maintained at 20 2 C.

3. RESULTS AND DISCUSSION


3.1. Adsorption Column Behaviour

The results of this study show that the profile of breakthrough curves varies with bed depth,
solution flow rate and initial dye concentration. The breakthrough curves do not follow the
characteristic S-shaped profile produced in ideal adsorption systems, a behaviour which is
associated with adsorbates of smaller molecular diameter and more simple structure (McKay,
1995).
The velocity at which the active adsorption zone moves through the column is assumed to be
constant except during the period at which it is formed. This velocity defines the adsorption zone
height d, as shown in equation (2)
t L
= U t = (2)
tx tf
Fixed-Bed Adsorption of Azo Dyes Using Novel Pine ConeDerived Active Carbon 797

where L is the column length (cm), tf is the active adsorption zone formation time (minutes),
Ud is the active adsorption zone velocity (cm minute1) and td is the time required to cross the
adsorption zone height (minutes).
The length of the adsorption zone was found to be higher than total bed depth for both dyes,
suggesting that beds of an increased height may be required for dye adsorption. This higher
length of the adsorption zone is attributed to the large molecular structure of the dyes as well
as their higher resistance to internal diffusion. Because the dye molecules do not have
sufficient contact time to diffuse from the surface of the particle to the adsorption sites, they
are pumped further up the column towards the fresh activated carbon in which the rate of dye
uptake is higher.
The formation time (tf) at the active adsorption zone can be estimated from equations (3) and (4)

tf = (1 f)t (3)

PS
f= (4)
Ptc

where f is the fractional capacity of the active adsorption zone, Ps is the quantity of solute adsorbed
in the active adsorption zone from breakthrough to exhaustion (mg) and Ptc is the total capacity of
carbon in the active adsorption zone (mg). The values of Ps and Ptc are obtained as follows
[equations (5) and (6)]:

Vx

PS = (C
Vb
0 C) dV (5)

Ptc = (Vx Vb)C0 (6)

The fractional capacity is then evaluated from equations (7) and (8)

Vx

PS Vb
(C0 C) dV (7)
f= =
Ptc (Vx Vb )C

1
C d(V Vb )
f = 1 (8)
0
C0 (Vx Vb )

where C is the dye concentration (mg l1), C0 is influent dye concentration (mg l1), Cx is the
effluent dye concentration at exhaustion (mg l1), V is the volume of liquid passing through the
column (l) and dV is the differential volume.
As f approaches 1, tf approaches 0, and the ideal plug-flow conditions are approximated. When
channelling or mass-transfer limitations prevail, the breakthrough curve rises very rapidly until it
approaches Cx, which is the exhaustion effluent solvent concentration. Under these conditions,
798 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

Ps is small and the fractional capacity, f, is approaching 0. The depth of the active adsorption zone,
d, can then be expressed in terms of fractional capacity, f, as follows:

L ( Vx Vb )
= (9)
Vb + f ( Vx Vb )

The degree of saturation at the breakthrough point can be evaluated using equation (10)

M
= y ( f ) y (10)
MS

where M is the total amount of dye adsorbed at the breakthrough point, MS is the total amount of
dye adsorbed at exhaustion point and y is the total column bed depth.
When column breakthrough occurs, the only portion of the adsorption column bed, which has
not been saturated, exhausted or in equilibrium with the influent solute concentration C is the final
portion of depth d in the column. The total adsorption capacity (Sb) at equilibrium (breakthrough
curve) can then be defined using equation (11) as

x
Sb = p ( L ) + ( f ) (11)
m C0

where Sb is the mass of solute adsorbed at equilibrium per unit of cross-sectional area of
adsorption bed, (x/m)C0 is the mass of organic material adsorbed per unit mass of carbon at the
equilibrium concentration C0, rp is the apparent packed density of the carbon adsorbent and L is
the bed depth.
The effective adsorption capacity of a carbon column depends on the fractional capacity f. As
the fractional capacity decreases, adsorption zone height (d) increases in size and mass of the
solute absorbed per cross-sectional area (Sb) decreases in magnitude. Therefore, it is very
important to maximize the fractional capacity (f) of the active adsorption zone.
The values of important parameters of adsorption column behaviour for the 20-cm bed depth
from the breakthrough point to the exhaustion point were calculated and are presented in Table 1.

TABLE 1. Column Behaviour Parameters for 20-cm Bed Depth for the Adsorption of AB1 and AB113 Dyes
Dye AB1 AB113

Parameter value
Ptc (mg) 4720 809
Ps (mg) 1740 219
F 0.369 0.271
kTh (l mg1 min1) 0.0308 0.0172
td (min) 3910 666
tf (min) 2470 485
tx (min) 4040 684.0
d (cm) 49.7 67.0
M/Ms 0.0807 0.0905
Ud (cm min1) 0.0127 0.101
Fixed-Bed Adsorption of Azo Dyes Using Novel Pine ConeDerived Active Carbon 799

3.2. Thomas Model

The expression of Thomas model (Thomas 1944) for an adsorption column is given by the
following equation:

Ct 1
=
C0 m
1 + exp k Th q 0 k Th C0 t (12)
Q

where kTh is the Thomas adsorption rate constant (l mg1 minute1), q0 is the equilibrium dye
uptake per gram of adsorbent (mg g1), M is the mass of PCAC in the column (g), Q is the
volumetric flow rate (ml minute1), Ct is the effluent concentration at time t (mg l1) and t is time
(minutes).
Thomas model parameters for several column conditions are presented in Table 2. The effect of
bed height is shown in Figure 3 for AB113, which shows a comparison between experimental
breakthrough curves (solid lines) at three bed heights, namely, 10, 15 and 20 cm, and the Thomas
model theoretical curves (dashed lines).

3.3. Effect of Flow Rate

The column was run with flow rates of 10 and 6.1 ml/minute, whereas the original flow rate was
8.07 ml/minute. The bed depth and column diameter were 10 and 2 cm, respectively. The
concentration of the influent dyes was kept constant (150 mg/l). The breakthrough curves obtained

TABLE 2. Thomas Model Parameters for Several Column Conditions


Q x C0 Qt q0 q0 kTh Reduced Adjusted
(ml min1) (cm) (mg l1) (mg l1) (mg/g) (l mg1 min1) Chi2 R2

AB1
10.00 10.00 148.4 23,080 57.26 0.0000637 0.01500 0.8335
8.070 10.00 149.6 37,490 93.04 0.0000254 0.01489 0.8288
6.100 10.00 152.0 59,340 147.2 0.0000164 0.01460 0.8475
8.070 10.00 149.6 37,490 93.03 0.0000254 0.01489 0.8288
8.070 15.00 149.6 55,950 138.8 0.0000148 0.01204 0.8780
8.070 20.00 149.6 90,880 225.5 0.0000093 0.00664 0.9355
8.070 10.00 198.2 26,770 66.42 0.0000388 0.00863 0.8953
8.070 10.00 149.6 37,490 93.04 0.0000254 0.01489 0.8288
8.070 10.00 98.00 35,330 87.68 0.0000318 0.01190 0.8583
AB113
10.00 10.00 151.4 6835 16.96 0.000225 0.00884 0.8971
8.070 10.00 150.5 7183 17.86 0.000252 0.01051 0.8954
6.100 10.00 149.2 8891 22.06 0.000129 0.01568 0.8038
8.070 10.00 150.5 7183 17.82 0.000252 0.01051 0.8954
8.070 15.00 150.5 6603 16.39 0.000165 0.01507 0.8508
8.070 20.00 150.5 9084 22.54 0.000086 0.01263 0.8826
8.070 10.00 201.9 4608 11.43 0.000218 0.01784 0.7019
8.070 10.00 150.5 7183 17.82 0.000252 0.01051 0.8954
8.070 10.00 99.69 8408 20.86 0.000227 0.01296 0.8838
800 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

1.0

0.8

0.6
C/C0

0.4

20 cm
15 cm
0.2
10 cm
Thomas model

0.0
0 150 300 450 600 750
Time (min)

Figure 3. Breakthrough curves obtained for AB113 dye removal by PCAC fixed-bed column using different bed depths.

1.0

0.8

0.6
C/C0

0.4

10 ml min1
8.07 ml min1
0.2
6.1 ml min1
Thomas model

0.0
0 500 1000 1500 2000 2500 3000
Time (min)

Figure 4. Breakthrough curves obtained for AB1 dye removal by PCAC fixed-bed column using different flow rates.

with different flow rates for AB1 dyes are shown in Figure 4. The results show that the
breakthrough curve was obtained quickly with higher flow rates. The time required to reach
saturation increased significantly with a decrease in flow rate.
Results in Figure 4 show that an increase in solution flow rate decreases the volume of dye
solution treated to achieve the breakthrough point; consequently, this decreases the service time
of the adsorbent bed.
Fixed-Bed Adsorption of Azo Dyes Using Novel Pine ConeDerived Active Carbon 801

TABLE 3. Empty Bed Contact Times at Different Flow Rates for the Adsorption of AB1 and AB113
Flow (ml min1) Bed volume (ml) EBCT (min)

10 31.4 3.14
8.07 31.4 3.89
6.1 31.4 5.15

This is owing to the decreased contact time between the dye and the carbon at higher flow rates.
Thus, at higher flow rates, the rate of mass transfer increased, that is, the amount of dye adsorbed
onto the unit bed height (mass transfer zone) is increased with increasing flow rate, which leads
to faster saturation at higher flow rates (Ko et al. 1999; Lee et al. 2004).
The flow rate represents the empty bed contact time (EBCT) in the column, which affects the
volume required to achieve breakthrough and the shape of the breakthrough curve. The EBCT is
determined from the equation (13)

Vc H
EBCT = =
Q v (13)

where Vc is the adsorption bed volume (ml), Q is the volumetric influent flow rate (ml minute1),
H is the bed depth (cm) and v is the linear flow rate (cm minute1).
The EBCTs at different flow rates for the adsorption of AB1 and AB113 are presented in Table 3.

3.4. Effect of Initial Dye Concentration

The column was run with initial dye concentrations of 200 and 100 mg/l, whereas the original
two dye concentrations were 150 mg/l. The bed depth and column diameter were 10 and 2 cm,
respectively. The influent flow was kept constant at 8.07 ml minute1. Figure 5 shows the
breakthrough curves of AB113 removal using the PCAC fixed-bed column for the
aforementioned initial dye concentrations. An increase in the inlet adsorbate concentration
increased the slope of the breakthrough curve and reduced the volume of adsorbate treated.
However, an increased inlet dye concentration at a constant flow rate decreases the throughput
until breakthrough is achieved. This may be due to the high adsorbate concentrations saturating
the activated carbon more quickly, subsequently decreasing the breakthrough time. In addition,
the effect of increasing influent concentration results in sharper breakthrough curves, which
demonstrates that the change in concentration gradient affects both the saturation rate and the
breakthrough times.

3.5. Theoretical Breakthrough Curve

Using the data obtained from the batch isotherm studies, theoretical breakthrough curves are
predicted, which can be well compared with the experimental curve. The theoretical breakthrough
curves were generated using the initial concentration of 150 mg/l, in accordance with the
procedure detailed by Michaels (1952).
The equilibrium curves for AB1 and AB113 dyes were obtained using Langmuir adsorption
isotherm (Choy et al. 2000; Langmuir 1918), which are presented as equations (14) and (15)
802 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

1.0

0.8

0.6
C/C0

0.4

201.86 mg I1
150.54 mg I1
0.2
99.69 mg mg I1
Thomas model

0.0
0 100 200 300 400 500 600

Time (min)

Figure 5. Breakthrough curves obtained for AB113 dye removal by PCAC fixed-bed column using different initial
concentrations.

Equilibrium equation for AB1 dye

356.81Ce
qe = (14)
1 + ( 0.787Ce )

Equilibrium equation for AB113 dye

37.055Ce
qe = (15)
1 + ( 0.124Ce )

where qe is the amount of dye (mg) adsorbed per unit mass of PCAC (g) and Ce is the equilibrium
concentration of dye remaining in the solution (mg/l).The equilibrium curve and operating lines
are shown for AB1 in Figure 6.
An operating line was drawn passing through the origin and through the point given by C0 and
the corresponding qe. The qe values corresponding to 150 mg/l concentration of AB1 and
AB113 dyes were 458.1 and 286.0 mg/g, respectively. The significance of this operating line
is that the data of a continuously mixed batch reactor and the data of fixed-bed reactor are
identical at these two points: first at the initiation and the other at the exhaustion of the
reaction (Choy et al. 2001).
The rate of transfer of the adsorbate from the solution over a differential depth of column dy is
given using the mass-transfer model shown in equation (16)
dC
Fm = = K a ( C = Ce )
dy (16)
Fixed-Bed Adsorption of Azo Dyes Using Novel Pine ConeDerived Active Carbon 803

500

400

300
qe , mg g1

200

100 Equilibrium curve

Operating line

0
0 20 20 60 80 100 120 140 160
Ce , mg L1

Figure 6. Plot of equilibrium and operating lines for the determination of theoretical breakthrough curve of AB1 dye.

where Fm is the flow rate, Ka is the overall Michaels mass transfer coefficient, which includes the
resistances offered by the film and internal diffusion, Ce is the equilibrium concentration for the
amount of dye adsorbed, y is the adsorbent bed length and C is the concentration of the adsorbate
at a given time t. The internal diffusion models are well-established, particularly for batch systems
(Choy et al. 2001, 2004), and two resistance column models are available for dye adsorption (Ko
et al. 2000; McKay et al. 1984). However, the authors of these studies point out that the external
film resistance is only influential in the first 510% of the adsorption, following which a single
mass transfer coefficient is sufficient to describe the system, thus simplifying the model solution.
The term (C Ce) represents the difference in concentration between the actual and the
equilibrium value of any point in the adsorbent and is the driving force for adsorption, which is
equal to the difference between the operating line and equilibrium curve at any given qe value.
Integrating equation (16) and solving for the height of the adsorption zone gives the following
equation:

Cx
Fm dC
=
Ka (C C )
Cb e
(17)

where Cb and Cx are concentrations of the adsorbate in effluent at breakthrough and exhaustion,
respectively. The plot of (C Ce)1 versus C for AB113 dye is shown in Figure 7.
The area under the curves represents the value of the integration in equation (17). For any value
of y < d, corresponding to a concentration C between Cb and Cx, equation (17) can be rewritten
as follows:
C
Fm dC
y=
Ka (C C )
Cb e
(18)
804 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

0.35

0.35

(c-ce)1, L mg1 0.25

0.20

0.15

0.10

0.05

0.00
0 10 20 30 40 50 60
ce,mg L1

Cx

Figure 7. Curve to evaluate dC/(C C ). for determining the theoretical breakthrough curve of AB113 dye.
e
Cb

Combining equations (17) and (18) yields the following relationship:

C
dC
y
(C C )
Cb e V Vb
= Cx = (19)
dC Vx Vb
C (C Ce )
b

where Vb and Vx are total volume of solution treated until breakthrough and up to exhaust point,
respectively, and V is the volume of solution treated within Vx for the effluent concentration C
within Cx. C

The values of (V Vb)/(Vx Vb) were calculated by dividing the values of dC / (C Ce ) by the
Cb
Cx

value of dC/(C C ). Finally the theoretical reakthrough curves that plotting (VVb)/(VE Vb)
e
Cb

versus (C/C0) are shown in Figure 8 for AB1 dye. Although the experimental and theoretical
breakthrough curves followed the same trend, there were differences in the results obtained.
There is an excellent correlation between the experimental and the theoretical data in Figure 8
for the adsorption of AB1. However, for the adsorption of AB113, the correlation is not as good
as that of AB1, particularly in the region of dimensionless liquid-phase concentration from 0.4 to
0.6. This discrepancy can be attributed to the breaking and slight lifting of the bed at the value
corresponding to C/C0 = 0.20 up to 0.38. At this point, the liquid phase becomes stationary and
does not move further. Consequently, the experimental breakthrough curve requires to be moved
to the right by an amount (VVb)/(VEVb) = 0.1 equivalent to the loss in adsorption capacity due
to the empty section of the bed at the point where the bed lifting occurred. The two breakthrough
curves for AB1 show very good agreement.
Fixed-Bed Adsorption of Azo Dyes Using Novel Pine ConeDerived Active Carbon 805

1.0

0.8

0.6
C/C0

0.4

Experimental breakthrough curve


for 20 cm bed depth
0.2
Theoretical breakthrough curve
for 20 cm bed depth

0.0
0.0 0.2 0.4 0.6 0.8 1.0 1.2
(V - VB)/(VE - VB)

Figure 8. Experimental and theoretical breakthrough curves of AB1 dye adsorption by PCAC fixed-bed column.

4. CONCLUSION

The adsorption of two acid dyes, AB1 and AB113, onto an activated carbon derived from pine
cones, PCAC, has been studied using fixed-bed column adsorption. Two models, Thomas and
Michaels models, were used to predict the breakthrough curves. The Michaels model provided a
better correlation for the breakthrough curves.

REFERENCES

ASTM (1999) Standard test method for determination of iodine number of activated carbon. ASTM D4607-
94, West Conshohocken, PA.
Banat, F., Al-Asheh, S. and Al-Makhadmeh, L. (2003) Process Biochem. 39, 193.
Bohart, G.S. and Adams, E.Q. (1920) J. Chem. Soc. 42, 523.
Carliell, C.M., Agridiotis, V. and Forstyer, C.F. (1998) Environ. Technol. 19, 1133.
Choy, K.K.H., Porter, J.F. and McKay, G. (2000) Chem. Eng. J. 45, 575.
Choy, K.K.H., Porter, J.F. and McKay, G. (2001) Chem. Eng. J. 81, 213.
Choy, K.K.H., Porter, J.F. and McKay, G. (2004) Chem. Eng. J. 103, 133.
Davies, R.A., Kaempf, H.J. and Clemens, M.M. (1973) Chem. Ind. 27, 827.
Gottlieb, A., Shaw, C., Smith, A., Wheatley, A. and Forsythe, S. (2003) J. Biotechnol. 101, 49.
Ho, Y.S. and McKay, G. (1998) Process Biochem. 34, 451.
Ho, Y.S. and McKay, G. (1999) Chem. Eng. J. 34, 735.
Hutchins, R.A. (1973) Chem. Eng. 80, 133.
Jagtoyen, M. and Derbyshire, F. (1998) Carbon. 36, 1085.
Johns, M.M., Marshall, W.E. and Toles, C.A. (1999) J. Chem. Technol. Biotechnol. 74, 1037.
Kadirvelu, K., Kavipriya, M., Karthika, C., Radhika, M., Vennilamani, N. and Pattabhi, S. (2003) Bioresour.
Technol. 87, 87.
Ko, D.C.K., Porter, J.F. and McKay, G. (1999) Ind. Eng. Chem. Res. 38, 4868.
Ko, D.C.K., Porter, J.F. and McKay, G. (2000) Chem. Eng. Sci. 55, 5819.
Langmuir, I. (1918) J. Am. Chem. Soc. 40, 1361.
806 M.R. Samarghandi et al./Adsorption Science & Technology Vol. 32 No. 10 2014

Lee, V.K.C., Porter, J.F. and McKay, G. (2004) Chem. Eng. J. 98, 255.
Mathur, N., Bhatnagar, P. and Bakre, P. (2005) Ecotoxicol. Environ. Safety. 61, 105.
McKay, G. (1982a) J. Chem. Technol. Biotechnol. 40, 759.
McKay, G. (1982b) J. Chem. Technol. Biotechnol. 32, 773.
McKay, G. (1995) Use of Adsorbents for the Removal of Pollutants from Wastewaters, CRC, Boca Raton,
New York, London, Tokyo.
McKay, G. and Bino, M.J. (1990) Water Air Soil Pollut. 51, 33.
McKay, G., Blair, H.S. and Gardner, J.R. (1984) J. Appl. Polym. Sci. 29, 1499.
McKay, G., El Geundi, M. and Nassar, M.M. (1997a) Adsorpt. Sci. Technol. 15, 753.
McKay, G., El Geundi, M. and Nassar, M.M. (1997b) Adsorpt. Sci. Technol. 15, 186.
McKay, G. and Ho, Y.S. (1998) Trans. I ChemE. 70, 115.
McKay, G., Otterburn, M.S. and Sweeney, A.G. (1985) Water Air Soil Pollut. 24, 307.
Michaels, A.S. (1952) Ind. Eng. Chem. 44, 1922.
Molina-Sabio, M., RodRguez-Reinoso, F., Caturla, F. and Sells, M.J. (1995) Carbon. 33, 1105.
Mui, E.L.K., Cheung, W.H., Valix, M. and McKay, G. (2010) Microporous Mesoporous Mater. 130, 287.
Nadeem, M., Tan, I.B., Haq, M.R.U., Shahid, S.A., Shah, S.S. and McKay, G. (2006) Adsorpt. Sci. Technol.
24, 269.
Namasivayam, C. and Kavitha, D. (2002) Dyes Pigm. 54, 47.
Namasivayam, C., Muniasamy, N., Gayatri, K., Rani, M. and Ranganathan, K. (1996) Bioresour. Technol.
57, 37.
Oktem, Y.A., Soylu, S.G.P. and Aytan, N. (2012) J. Sci. Ind. Res. 71, 817.
Poots, V.J.P., McKay, G. and Healy, J.J. (1976) Water Res. 10, 1061.
Rajeshwarisivaraj, S.S., Senthilkumar, P. and Subburam, V. (2001) Bioresour. Technol, 80, 233.
Snoeyink, V.L. (1990) Water Quality and Treatment, 4th Ed. McGraw-Hill, New York.
Thomas, H.C. (1944) J. Am. Chem. Soc. 66, 1466.
Toles, C.A., Marshall, W.E., Johns, M.M., Wartelle, L.H. and McAloon, A. (2000) Bioresour. Technol. 71,
87.
Valix, M., Cheung, W.H. and McKay, G. (2004) Chemosphere. 56, 493.
Walker, G.M. and Weatherley, L.R. (1997) Water Res. 31, 2093.
Wu, F. C., Tseng, R. L. and Juang, R.S. (1999) J. Environ. Sci. Health. A34, 1753.

Potrebbero piacerti anche