Sie sind auf Seite 1von 9

Journal of Hazardous Materials 177 (2010) 842850

Contents lists available at ScienceDirect

Journal of Hazardous Materials


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

Removal and adsorption characteristics of polyvinyl alcohol from aqueous


solutions using electrocoagulation
Wei-Lung Chou
Department of Safety, Health and Environmental Engineering, Hungkuang University, Sha-Lu, Taichung 433, Taiwan

a r t i c l e

i n f o

Article history:
Received 6 July 2009
Received in revised form
20 December 2009
Accepted 26 December 2009
Available online 6 January 2010
Keywords:
Electrocoagulation
Polyvinyl alcohol (PVA)
Adsorption kinetics
Adsorption isotherms
Thermodynamics

a b s t r a c t
The study was to investigate the performance of electrocoagulation (EC) for the efcient removal of
polyvinyl alcohol (PVA) from aqueous solutions. Several parameters were evaluated to characterize the
PVA removal efciency, such as various electrode pairs, current densities, supporting electrolytes, temperatures, and initial electrolyte concentrations. The effects of the current density, supporting electrolyte,
and temperature on the electrical energy consumption were also investigated. The experimental results
indicate that a Fe/Al electrode pair is the optimum choice out of four different electrode pair combinations.
The optimum current density, supporting electrolyte concentration, and temperature were found to be
5 mA cm2 , 0.008N NaCl, and 298 K, respectively. The PVA removal efciency decreased with increasing in the initial concentrations. The kinetic studies indicated that the EC process was best described
using pseudo-second-order kinetics. The experimental data were also compared to different adsorption
isotherm models in order to describe the EC process. The adsorption of PVA was best tted by the Langmuir adsorption isotherm model. Thermodynamic parameters such as the Gibbs free energy, enthalpy,
and entropy indicated that the adsorption of PVA on metal hydroxides was feasible, spontaneous and
endothermic in the temperature range of 288318 K.
2010 Elsevier B.V. All rights reserved.

1. Introduction
Polyvinyl alcohol (PVA) is a well-known water-soluble and
biodegradable synthetic polymer obtained by hydrolysis of
polyvinyl acetate (PVAc). PVA is frequently used in the textile industry as a size for nylon and as a raw material for the production of PVA
bers [1]. In addition, PVA is used as an ophthalmic lubricant in the
pharmaceutical industry and is also widely used in the adhesives,
emulsion paints, paper coating, and detergent-based industries. It
is generally recognized that PVA is a high-k polymer with interesting properties such as good surface alignment effects, compatibility
with water, low cost, and inexpensive processing [2]. The large
amount of discharged PVA from industrial efuents is harmful to
human health and to the environment [3]. It is often difcult to
clean wastewaters containing PVA and to generate harmless endproducts such as water and carbon dioxide; thus, PVA adversely
affects the ecosystem and accumulates in the human body via the
food chain [4]. PVA also creates environmental issues due to its
ability to prevent the sedimentation of heavy metals in lakes and
streams [5]. Conventional biological technologies do not offer an
effective treatment of PVA because the degrading capacity of most
microorganisms is extremely restricted and specic for PVA [6].

Tel.: +886 4 26318652x4005; fax: +886 4 26319175.


E-mail address: wlchou@sunrise.hk.edu.tw.
0304-3894/$ see front matter 2010 Elsevier B.V. All rights reserved.
doi:10.1016/j.jhazmat.2009.12.110

Using mixed cultures acclimatized to PVA solutions, a past study


has shown that only about 40% of the PVA was mineralized after 48
days of incubation [7]. The formation of foam in biological equipment for the treatment of wastewater containing PVA makes stable
operations and the achievement of acceptable results very difcult
[8]. Therefore, there is a growing interest in the development of
new methods for the removal of PVA.
A large number of scientic studies on the removal of PVA
have been carried out. Most of these studies have focused on
photochemically initiated degradation processes [911]. Other
physico-chemical studies on the removal of PVA focus on methods
such as ultrasonic techniques [12], direct oxidation by KMnO4 [13],
radiation-induced degradation [14], adsorption by various materials, and extraction resins [1518]. However, little attention has
been paid so far to the development of an electrochemical treatment for the removal of PVA.
Electrochemical technologies contribute in many ways to a
cleaner environment and cover a very broad range of techniques
[19]. There has been a growing interest in the signicant use of
environmental electrochemistry. Environmental electrochemistry
[2023] involves electrochemical techniques in order to remove
impurities and to prevent environmental pollution. Among these
techniques, electrocoagulation (EC) is an electrolytic process that
generates metallic hydroxide ocks in situ via electro-dissolution of
the soluble sacricial anode immersed in the wastewater. The generation rate of ocks can be controlled by an applied current. The

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

electrochemically generated metallic ions can be hydrolyzed next


to the anode and generate a series of metal hydroxides that are able
to destabilize the dispersed particles present in the wastewater to
be treated. The destabilized particles are believed to be responsible for the aggregation and precipitation of the suspended particles
and for the adsorption of the dissolved pollutants. The pollutants
can then be removed by sedimentation and turned into slurry. This
technique also presents several benets compared to conventional
methods: higher energy efciency, easy operations, simple equipment, less sludge production, and no requirements for the use of
chemicals. EC is an efcient method for the treatment of drinking
water [24], metal laden wastewater [25], restaurant wastewater
[26], colored water [27], chemical mechanical polishing wastewater [2830], and wastewater containing phosphates [31], uoride
[32], arsenic [33], and indium [34,35]. To our knowledge, the effects
of the EC parameters on the removal of PVA in aqueous solutions
have not been studied to date.
The electrochemical removal of PVA under various parameters
such as the current density, supporting electrolyte, temperature,
and initial concentration are investigated in this study. In general,
an effective process must be economically advantageous regarding
the electrical energy consumption (EEC), and it must be applicable
to many environmental problems. The effects of various operational parameters (current density, supporting electrolyte, and
temperature) on EEC using optimum conditions were evaluated.
The adsorption kinetics of the electro-coagulants was analyzed
using pseudo-rst- and second-order kinetic models. The equilibrium adsorption behavior was analyzed and compared to the
model predictions of the Langmuir and Freundlich isotherms. The
effect of temperature on the adsorption isotherms was determined,
and the thermodynamic parameters during the adsorption such as
the free energy, enthalpy and entropy were calculated for various
temperatures.

2. Experimental
2.1. Materials and apparatus
The polyvinyl alcohol (PVA, with a molecular weight in the
range of 13,00023,000 g mol1 ) was obtained from SigmaAldrich
(Saint Louis, MO 63103, USA) with a hydrolysis degree ranging
from 98 to 99%. An aqueous solution containing polyvinyl alcohol
was prepared in deionized water at 363 K under stirring. The concentration of the supporting electrolyte was adjusted by adding
NaCl (Tedia Company, USA). Iodine (I) was obtained from Toyobo
Co. Ltd. (Osaka, Japan), and potassium iodide (KI) was purchased
from Union Chemical Work Ltd. (Hsin-Chu, Taiwan). The boric acid
(H3 BO3 ) was purchased from Merck (Darmstadt, Germany). All of
the chemicals used in this work were reagent grade. The chemical
reagents were prepared by dilution with deionized water to obtain
the desired concentrations. Fig. 1 is a schematic diagram of the
experimental apparatus and the electrode assembly for the electrocoagulation system used in this work. The electrolytic cell with
an inner radius of 5 cm and a depth of 24 cm had a volume of about
2.0 L and was a Pyrex glass reactor equipped with a water jacket
and a magnetic stirrer. The electrolytic cell was thermostated by
means of its water jacket using a refrigerated water bath circulator
(Model BL-720, Taiwan). A magnetic stirrer bar (Suntex, SH-301,
Taiwan) was spun at the center of the bottom of the reactor. Cast
iron (Fe) and aluminum (Al) plates (12 cm 6 cm 0.2 cm) were
used for the four different combinations of anode and cathode pairs.
The electrode pair was dipped into an aqueous solution of polyvinyl
alcohol at a depth of 6.5 cm, and the two electrodes were approximately 2 cm apart. The effective area of the immersed electrode pair
was 40 cm2 . The assembly was connected to a direct current power

843

Fig. 1. Schematic diagram of the electrocoagulation equipment.

source (2400 Series Sourcemeter, Cleveland, OH, USA). Thus, the EC


cell can be operated in a galvanostatic mode with a constant current supply and a maximum source power of 22 W. The polyvinyl
alcohol aqueous solutions were characterized using a pH meter
(Sartorius, Professional Meter PP-20, Germany) and by measuring
their conductivity (Euteoh, CyberScan 510, Singapore).
2.2. Experimental procedure and analysis
Before each experiment, the electrodes were polished with
sandpaper to remove the scale build-up and then dipped in a 3N HCl
solution at a depth of 8 cm for 10 min. Finally, they were cleaned
with deionized water. During each test run, a circular container
with 1 L of aqueous PVA solution was used as the reactor. The magnetic stirrer was turned on and set at 300 rpm; this stirrer speed
was fast enough to provide a good mixing in the electrolytic cell
and slow enough at the same time to prevent the breakup of the
ocks formed during the treatment process. A xed amount of
NaCl was added to the aqueous solution to increase the solution
conductivity and thus facilitate the EC process. A constant temperature was maintained by circulating refrigerated water through the
water jacket. Each EC test lasted 120 min for all experiments. After
EC, particulates of colloidal ferric hydroxide were produced, which
led to the yellow-brown color of the aqueous solution. The samples were periodically taken from the reactor and allowed to settle
down for 6 h in a 15 mL Pyrex glass column. After the EC treatment,
the conductivity and pH of the polyvinyl alcohol aqueous solutions
were measured with a multi-meter and a pH meter, respectively.
The PVA concentration in the aqueous solutions was determined
using a HACH Model DR2800 spectrophotometer (HACH Company,
USA) after addition of boric acid and iodine solutions according
to the procedure described by Finley [36]. A calibration curve was
obtained by plotting the absorbance value at 680 nm as a function
of the polyvinyl alcohol concentration. The PVA removal efciency
after EC was calculated according to:
RE (%) =

C0 V0 Ct Vt
100,
C0 V0

(1)

where C0 is the initial concentration in mg L1 , Ct is the concentration at time t in mg L1 , V0 is the initial volume of the treated
wastewater in L, and Vt is the volume of the treated wastewater at
time t in L.

844

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

Fig. 2. Effect of different electrode pairs on the removal efciency of PVA


(PVA = 100 mg L1 , current density = 2.5 mA cm2 , t = 120 min, T = 298 K, d = 2 cm,
NaCl = 0.002N, agitation speed = 300 rpm).

3.2. Effect of the current density

3. Results and discussion


3.1. Type of electrode pair
In electrochemical processes, the electrode material has signicant effects on the treatment efciency. The type of electrode pair is
regarded as an important factor controlling the performance of the
EC process [37]. Therefore, an appropriate selection of the electrode
pair is of paramount importance. Iron and aluminum were chosen
as electrode materials because they were non-toxic and readily
available. Four combinations of iron and aluminum plates were
investigated in this study in order to determine the optimum electrode pair. In Fig. 2, we show the effects of the electrode pair on the
PVA removal efciency. Also shown in this gure are the results for
the iron anode and aluminum cathode that had the highest removal
efciency after 120 min of electrolysis. The PVA removal efciencies were approximately 94.9%, 68.7%, 47.3%, and 30.1% for a Fe/Al,
Fe/Fe, Al/Al, and Al/Fe electrode pair, respectively. In addition, the
Fe/Al electrode pair gave the best PVA removal efciency, followed
by the Fe/Fe, Al/Al, and Al/Fe pairs. The PVA removal efciencies for
Fe/Al and Fe/Fe pairs using Fe as the anode were greater than those
for Al/Al and Al/Fe pairs using Al as the anode. This can be explained
by the chemical reactions that take place at the aluminum anode
and the iron anode.
For the Al anode:
Al Al3+ 3e

(2)

For the Fe anode:


Fe Fe2+ + 2e

Fig. 3. Effect of the current density on the removal efciency of PVA


(PVA = 100 mg L1 , t = 120 min, T = 298 K, d = 2 cm, NaCl = 0.012N, agitation
speed = 300 rpm).

(3)

The nascent aluminum and iron ions are very efcient coagulants for the occulation of particulates. From the two reactions
(2) and (3), we can calculate the electrochemical equivalent mass
for Al and Fe. The electrochemical equivalent mole for aluminum is
12.43 mmol (Ah)1 , and it is 18.59 mmol (Ah)1 for iron. Also, some
researchers have reported that the adsorption of dye molecules by
hydrous aluminum oxides is much lower than those using hydrous
ferric oxides [38]. Therefore, more coagulants are theoretically produced by iron anodes when the same electrical charge is applied.
In the following, all EC experiments were conducted using a Fe/Al
electrode combination.

One of the most signicant parameters affecting the EC performance is the current density. The current density was determined
by dividing each current by the corresponding electrode area. The
effect of the current density on the efciency of the PVA removal
from aqueous solutions was studied at 1.25, 2.5, 5, and 7.5 mA cm2 .
Fig. 3 shows the effect of the current density on the PVA removal
efciency for various electrolysis times. As the duration of the electrolysis was increased, comparable increases in the PVA removal
efciency were observed for all current densities. After 120 min of
electrolysis, we observed that 71.2%, 81.5%, 92.1%, and 93.4% of the
original PVA was removed at current densities of 1.25, 2.5, 5, and
7.5 mA cm2 , respectively. The current density strongly determines
the coagulant dosage rate. As the current density was increased, the
PVA removal efciency also increased. At high current densities, the
anodic dissolution of iron increases, which leads to an increase in
the amount of metal hydroxides and results in an improved PVA
removal. As time progresses, the amount of dissolved iron hydroxides increases and results in a larger amount of precipitate and an
improved PVA removal efciency. This can be explained by the fact
that the PVA adsorption increases with increasing iron hydroxides
concentrations. The treatment times required to reach the removal
of more than 70% of the PVA were 28, 30, 62, and 108 min for current densities of 7.5, 5, 2.5, and 1.25 mA cm2 , respectively. As the
current density increased, the required times for the EC process
decreased. When there is enough current in the solution, the metal
ions generated by the dissolution of the sacricial electrode are
hydrolyzed and form a series of metal hydroxides. These hydroxides neutralize the electrostatic charges on the dispersed particles,
reducing the electrostatic repulsion between the particles to the
point where van der Waals attractions become predominant and
thus facilitate agglomeration [39]. However, a slight improvement
in the removal efciency was observed when the current density
was increased from 5 to 7.5 mA cm2 . The performance of the PVA
removal for a given energy consumption was evaluated at a constant current density during EC in order to determine the optimum
current density. The corresponding results are given in the following section.
3.2.1. Effect of the current density on removal efciency and
electrical energy consumption
It is essential to evaluate the electrical energy consumption
for the treatment of wastewater to determine whether EC is a
nancially viable method for the removal of PVA from aqueous
solutions. Once the required currents and corresponding voltages

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

Fig. 4. Effect of the current density on the specic energy consumption and PVA
removal efciency (PVA = 100 mg L1 , t = 120 min, T = 298 K, d = 2 cm, NaCl = 0.012N,
agitation speed = 300 rpm).

are obtained from EC experiments, the amount of consumed energy


can be estimated. In the case of EC with a constant applied current, the EEC (kWh m3 ) was calculated as a function of time for
the removal of 1 m3 of aqueous solution containing PVA using the
following equation:
EEC =

U dt

V 60

(4)

where U, I, and t are, respectively, the applied voltage (V), current (A), and electrolysis time (min), and V is the volume in m3 of
the treated aqueous solution containing PVA. Reasonable removal
efciency and relatively low EC are determined below.
After 120 min of electrolysis, the PVA solutions were treated by
EC at current densities in the range of 1.257.5 mA cm2 in order
to determine the optimum removal efciency and EEC. The effects
of the current density on the removal efciency and on the energy
consumption are shown in Fig. 4. The results show that an increase
in the current density from 1.25 to 7.5 mA cm2 led to a signicant increase from 71.2% to 93.4% in the PVA removal efciency.
When the current density was increased from 1.25 to 5 mA cm2 ,
the PVA removal efciency strongly increased from 71.2% to 92.1%,
whereas the corresponding EEC only increased slightly. However,
when the current density was increased from 5 to 7.5 mA cm2 , the
PVA removal efciency was slightly improved from 92.1% to 93.4%,
whereas the corresponding EEC signicantly increased from 0.079
to 0.211 kWh m3 . Consequently, a current density of 5 mA cm2
provides the optimum performance for the present study and
results in reasonably good removal efciency and relatively low
EEC.

845

Fig. 5. Effect of the supporting electrolyte concentration on the removal efciency of


PVA (PVA = 100 mg L1 , current density = 5 mA cm2 , t = 120 min, T = 298 K, d = 2 cm,
agitation speed = 300 rpm).

0.012, and 0.016N, the removal efciency increased from 85.1% to


91.1%, 92.1%, and 94.1%, respectively. This is most certainly due to
the Cl anions that destroy the passivation layer of the electrode
and catalyze its dissolution material through a pitting corrosion
phenomenon, which is a type of localized corrosion caused by a high
chloride concentration in the solution [40]. Therefore, we thought
that the EC in the presence of NaCl could improve the PVA removal
efciency by increasing the availability of metal coagulants in the
solution and by leading to a reduction of the oxide layer and an
enhancement of the anodic dissolution of the electrode material.
Consequently, the problem of electrode passivation was partially
solved when NaCl was used as a supporting electrolyte. The treatment times required to reach a PVA removal efciency of 80% were
approximately 42, 50, 60, and 80 min for 0.016, 0.012, 0.008, and
0.004N, respectively. However, no signicant improvement in the
PVA removal efciency was observed after 120 min of electrolysis
when the supporting electrolyte concentration was above 0.008N.
3.3.1. Effect of the supporting electrolyte on removal efciency
and electrical energy consumption
The supporting electrolyte concentration was adjusted by
adding a suitable amount of NaCl to the solution. Fig. 6 shows the
effect of the supporting electrolyte on the removal efciency and
the EEC during the EC. As shown in the gure, the PVA removal
efciency increased from 85.1% to 94.1% after 120 min of electrolysis, whereas the supporting electrolyte concentration increased
from 0.004 to 0.016N and the corresponding EEC was considerably

3.3. Effect of the supporting electrolyte concentration


The solution conductivity increases with increasing supporting electrolyte concentrations, so that the current passing through
the circuit increases for the potentiostatic mode [40]. Any electrochemical treatment requires the presence of salts as a supporting
electrolyte to make the solution more conductive. In the present
study, NaCl was used as a supporting electrolyte to increase the
solution conductivity and thus reduce the EEC. The effect of various
supporting electrolyte concentrations on the PVA removal efciency is shown in Fig. 5. A signicant increase in the PVA removal
efciency was observed with increasing electrolysis times, regardless of the supporting electrolyte concentration. This strongly
suggests that the presence of the supporting electrolyte improves
the PVA removal efciency. As the concentration of the supporting electrolyte was increased, the PVA removal efciency was
improved. After 120 min of electrolysis and when the concentration
of the supporting electrolyte was increased from 0.004 to 0.008,

Fig. 6. Effect of the supporting electrolyte concentration on the specic


energy consumption and PVA removal efciency (PVA = 100 mg L1 , current density = 5 mA cm2 , t = 120 min, T = 298 K, d = 2 cm, agitation speed = 300 rpm).

846

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

Fig. 7. Effect of temperature on the removal efciency of PVA (PVA = 100 mg L1 ,


current density = 5 mA cm2 , t = 120 min, d = 2 cm, NaCl = 0.008N, agitation
speed = 300 rpm).

decreased by almost 85%. This is clear evidence that the solution


conductivity increases when the concentration of the supporting
electrolyte increases. A higher ionic strength will generally lead to
a decrease in the cell voltage with increasing solution conductivity at a constant current density. Accordingly, the required voltage
to reach a certain current density will be diminished, and the consumed electrical energy will decrease. We observed constant EEC
when the concentration of the supporting electrolyte was above
0.008N. When the concentration of the supporting electrolyte was
increased from 0.008 to 0.012N, the removal efciency was slightly
increased from 91.1% to 92.1%, whereas the corresponding EEC
decreased from 0.104 to 0.079 kWh m3 . Consequently, a NaCl
concentration of 0.008N provides the optimum conditions for reasonably good removal efciency at relatively low EEC.
3.4. Effect of temperature
In this study, the effect of the temperature on the PVA removal
efciency was studied at 288, 298, 308, and 318 K, as shown in
Fig. 7. As the electrolysis time was increased, comparable increases
in the PVA removal efciency were observed for the different temperatures. After 120 min of electrolysis, we observed that the PVA
removal efciency reached 82.9%, 91.1%, 92.6%, and 93.8% for a temperature of 288, 298, 308, and 318 K, respectively. The treatment
times required to reach a removal of 80% of the PVA were approximately 44, 48, 60, and 88 min for a temperature of 318, 308, 298,
and 288 K, respectively. Thus, the electrochemical reaction rate, as
the other chemical reaction rates, increased with increasing solution temperatures. The temperature inuence can be attributed
to an improved destruction of the iron oxide lm on the anode
surface and an increase in the rate of all reactions involved in
the process according to the Arrhenius equation [20]. In addition,
higher temperatures promote the generation of iron hydroxides
formed in the EC process, which leads to a greater mobility and
more frequent collisions with the iron hydroxides, resulting in an
increased adsorption rate between the iron hydroxides and the pollutants [41]. However, little improvement was observed for the
PVA removal efciency after 120 min of electrolysis once the solution temperature was above 298 K, as shown in Fig. 7. It can be
explained by the fact that higher temperatures most certainly lead
to an increase in the solubility of the precipitates or to the generation of unsuitable ocks, which represents an adverse effect [42].
According to the current results, it seems that within the temperature range of 288298 K, the benecial effects dominate over the
adverse effects. For temperatures higher than 298 K, the benecial
effects are balanced by the adverse effects.

Fig. 8. Effect of temperature on the specic energy consumption and PVA removal
efciency (PVA = 100 mg L1 , current density = 5 mA cm2 , t = 120 min, d = 2 cm,
NaCl = 0.008N, agitation speed = 300 rpm).

3.4.1. Effect of temperature on removal efciency and electric


energy consumption
To investigate the temperature effects on the EEC and on the
PVA removal efciency, several experiments were performed after
120 min of electrolysis and using a Fe/Al electrode pair, an initial
PVA concentration of 100 mg L1 , 0.008N NaCl, a current density of
5 mA cm2 , and an agitation speed of 300 rpm. The temperature of
the PVA solution was controlled using the refrigerated circulating
bath. Fig. 8 shows the effects on the temperature on the performance of the PVA removal efciency and on the EEC after 120 min
of electrolysis by iron electrocoagulation. The electrical energy consumption signicantly decreased by almost 67% when the solution
temperature was increased from 288 to 298 K, whereas the corresponding PVA removal efciency increased from 82.9% to 91.1%.
However, above a solution temperature of 298 K, the EEC remained
almost constant. When the solution temperature was increased
from 298 to 308 and 318 K, the EEC increased from 0.104 to 0.078
and 0.0619 kWh m3 , respectively, whereas the corresponding PVA
removal efciency only slightly increased from 91.1% to 92.6% and
93.8%, respectively. Consequently, considering both the EEC and
the PVA removal efciency, 298 K offers the best compromise and
provides a reasonably good PVA removal efciency and a relatively
low EEC.
3.5. Effect of the initial concentration
Different aqueous solutions with PVA concentrations in the
range 50150 mg L1 were treated, and the PVA removal efciencies were measured after various electrolysis times during
the EC. Fig. 9 shows the effect of the initial PVA concentration
in the aqueous solution on the removal efciency for various
electrolysis times. We observed that the PVA removal efciency
reached 93.2%, 91.1%, and 84.7% after 120 min of electrolysis for
initial PVA concentrations of 50, 100, and 150 mg L1 , respectively. When the initial concentration was increased, the PVA
removal efciency decreased. Also shown in Fig. 9 are the treatment times required to reach a removal of 80% of the PVA: about
33, 61, and 74 min for concentrations of 50, 100, and 150 mg L1 ,
respectively. According to Faradays law, a constant amount of
metal hydroxides is dissolved from the iron anode and passes to
the solution for the same current density and electrolysis time
for all PVA concentrations. Consequently, the same amount of
metal hydroxides is produced in the aqueous solution. This is
most likely why the amount of hydroxyl and metal ions produced on the electrodes at high PVA concentrations was not
sufcient to absorb all the PVA in the solution at a constant current
density.

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

847

from a solution [44]. The pseudo-rst-order equation is given by


dqt
= k1 (qe qt )
dt

(6)

where qe (mg g1 ) and qt (mg g1 ) are the amounts of PVA adsorbed


on the adsorbent at equilibrium and at any time t, respectively, and
k1 (min1 ) is the rate constant of the pseudo-rst-order model.
After integration of Eq. (6) and after applying the boundary conditions qt = 0 at t = 0 and qt = qt at t = t, we obtain:
ln(qe qt ) = ln qe k1 t

Fig. 9. Effect of the initial concentration on the removal efciency of PVA (current density = 5 mA cm2 , t = 120 min, T = 298 K, d = 2 cm, NaCl = 0.008N, agitation
speed = 300 rpm).

3.6. Adsorption kinetics


Metal hydroxides that are formed during the EC process possess a very high afnity for adsorption. Coagulated particles attract
and absorb different ions and colloidal particles from the wastewater. If iron electrodes are used, the generated Fe ions immediately
undergo further spontaneous reactions to produce corresponding
hydroxides and/or polyhydroxides. Ferric ions are commonly generated by oxidizing ferrous ions produced at the anode during the
dissolution of iron, while OH ions are generated at the cathode.
Mixing the solution produces hydroxide species, which remove
pollutants by adsorption and co-precipitation. In the present study,
the pH values of aqueous solutions containing PVA were in the
range 6.59.5 during the EC process, which indicates that Fe(OH)3
is the dominant metal hydroxide functioning as an adsorbent [43].
The amount of PVA adsorbed by the Fe(OH)3 was calculated from
the difference between the PVA quantity added to the Fe(OH)3 and
the PVA content of the precipitation using the following equation:
qe =

V (C0 Ce )
M

(5)

where qe is the PVA uptake (mg g1 ), C0 and Ce are, respectively,


the initial and nal (or equilibrium) concentrations in the solution
(mg L1 ), V is the volume of the solution (L) and M is the mass of
Fe(OH)3 (s). In addition, qe was calculated on the basis of assuming that the current efciency of the electrochemical cell is unity.
In order to investigate the mechanisms of the PVA adsorption process, two different kinetic models, namely, the pseudo-rst-order
model and pseudo-second-order model were applied to describe
the kinetics of the PVA adsorption onto iron hydroxides. The bestt model was selected according to the linear regression correlation
coefcient values, R2 .

3.6.1. Pseudo-rst-order model


The rst-order rate equation of the Lagergren model is one of the
most widely used expression describing the adsorption of solute

(7)

The pseudo-rst-order model considers that the rate of occupation of the adsorption sites is proportional to the number of
unoccupied sites. The values of k1 and qe can be obtained from
the slope and intercept of the linear plot of ln(qe qt ) versus t,
respectively. The adsorption rate constants determined from the
pseudo-rst-order model are listed in Table 1. We observed that
the correlation coefcients for the rst-order kinetic model were
relatively lower than those obtained for the second-order kinetic
model for various initial concentrations.
3.6.2. Pseudo-second-order model
The pseudo-second-order kinetic model is based on the adsorption equilibrium capacity and can be written as [45]:
dq
= k2 (qe qt )2
dt

(8)

where k2 (g mg1 min1 ) is the rate constant of the pseudo-secondorder equation, qe (mg g1 ) is the maximum adsorption capacity,
and qt (mg g1 ) is the amount of adsorption at time t (min). After
integration and after applying the boundary conditions qt = 0 at t = 0
and qt = qt at t = t, Eq. (8) becomes:
t
1
t
=
+
qt
qe
k2 q2e

(9)

If second-order kinetics is applicable, plotting t/qt as a function


of t results in a straight line, and qe and k2 can be obtained from
its slope and intercept, respectively. The adsorption rate constants
determined from the pseudo-second-order model are also listed in
Table 1.
As discussed above, the validity of the model of Lagergren
and the pseudo-second-order kinetic model can be identied for
each linear plot. The value of the correlation coefcient R2 for the
pseudo-second-order model was greater than 0.991 for all the initial concentrations. The large correlation coefcients (R2 > 0.991 in
all initial concentrations) and the adsorption capacities calculated
from the model are also close to those determined experimentally.
These results indicate that the second-order kinetic model can be
applied to predict the PVA adsorption process onto iron hydroxides.
3.7. Adsorption isotherms
The formed Fe(OH)n(s) (where n = 2 or 3) complexes remain
in the aqueous stream as a gelatinous suspension. These gelatinous charged hydroxyl cationic complexes can effectively remove
pollutants by adsorption and produce charge neutralization by

Table 1
Comparison of the pseudo-rst-order and pseudo-second-order adsorption orate constants and calculated and experimental qe values for different initial concentrations.
Initial concentration (mg L1 )

qe,exp (mg g1 )

Pseudo-rst-order
1

k1 (min
50
100
150

57.4
111.6
152.2

0.044
0.039
0.037

qe (mg g
44.4
101.8
140.2

Pseudo-second-order
1

k2 104 (g mg1 min1 )

qe (mg g1 )

R2

0.948
0.946
0.931

8.622
5.275
3.543

58.8
113.6
153.8

0.995
0.991
0.994

848

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

Table 2
Langmuir and Freundlich isotherm constants for adsorption of PVA on metal hydroxides.
Langmuir
KL (L mg

Frendlich
)

0.0044

aL (mg g

1428

KF ((mg g1 )(L mg1 )1/n )

R2

0.86

5.454

0.917

0.94

complexation, electrostatic attraction, and enmeshment in a precipitate [37]. The Fe(OH)3 complex is the dominant species in the
range of pH 69 according to the predominance-zone diagram for
Fe(III) chemical species in aqueous solution [45]. The electrode
consumption can be estimated according to Faradays law, and
the amount of generated ocks can be stoichiometrically determined. Since the amount of coagulant can be estimated for a given
time, the pollutant removal can be modeled using an adsorption
phenomenon. The two most commonly employed adsorption equilibrium models are the Langmuir and Freundlich equations. In this
study, the adsorption characteristics for the PVA removal from
aqueous solutions were compared to the two models of adsorption
isotherms.
The Langmuir model was originally introduced to describe the
chemisorption at a set of well-dened localized adsorption sites
having the same adsorption energy, irrespective of the surface coverage and without interactions between the adsorbed molecules.
This model assumes a monolayer deposition on a surface having a
nite number of identical sites. It is well known that the Langmuir
equation is valid for a homogeneous surface. The mathematical
expression for the Langmuir isotherm is [46]:
qe =

aL KL Ce
1 + KL Ce

(10)

where aL (mg g1 ) is a constant related to the area occupied by a


monolayer of adsorbate and reecting the maximum adsorption
capacity; Ce (mg L1 ) is the equilibrium liquid-phase concentration; KL (L mg1 ) is a direct measure of the intensity of adsorption;
and qe (mg g1 ) is the amount of adsorbed species at equilibrium.
From the plot of 1/qe versus 1/Ce , KL and aL can be determined from
the slope and intercept of the resulting straight line.
The Freundlich adsorption isotherm usually enables tting
the experimental data for a wide range of concentrations. This
empirical model takes into account the surface heterogeneity and
exponential distribution of the active sites as well as their energies.
The isotherm can be used to describe a reversible adsorption and is
not restricted to the formation of a monolayer. The mathematical
expression for the Freundlich mode is [46]:
1/n

qe = Kf Ce

(11)

where Kf ((mg/g)(L/mg)1/n ) and n (dimensionless) are constants


that account for all factors affecting the adsorption process, such
as the adsorption capacity and intensity.
The Freundlich constants Kf and 1/n can be respectively determined from the intercept and the slope of the linear plot of ln qe
versus ln Ce . The applicability of the isotherm equations was compared based on the correlation coefcients (R2 ) in the present study.
The constants of the Langmuir and Freundlich isotherms are listed
in Table 2. We observed that the Freundlich isotherm model gave a

better t than the Langmuir isotherm model for the adsorption of


PVA onto iron hydroxides. The variable Kf of the Freundlich equation is related to the adsorption capacity of the adsorbent, and the
parameter n gives an indication of the magnitude of the deviation of
the adsorption from linearity. When the value of n is equal to unity,
the adsorption sites are homogeneous in energy, and no interaction
takes place between the adsorbed species. If n < 1, the adsorption
process is largely physical. If n > 1, the adsorption process is chemical [47]. Since the value of n at equilibrium was 0.94 at 298 K in this
study, the physical adsorption was dominant for the PVA removal
from aqueous solutions by adsorption onto iron hydroxides.
3.8. Adsorption thermodynamics
The thermodynamic parameters, such as the Gibbs free energy
change (G ), the enthalpy change (H ), and the entropy change
(S ), were used to determine whether or not the adsorption process was spontaneous. The values of G are calculated using the
following equation:
G = RT ln Kd

(12)
(8.314 J mol1

K1 ),

where R is the universal gas constant


T is the
absolute temperature (K), and Kd is the distribution coefcient. The
Kd value is calculated using the following equation [48]:
Kd =

qe
Ce

(13)

where qe and Ce are the equilibrium concentrations of PVA for the


adsorbent (mg L1 ) and in the solution (mg L1 ), respectively. The
relationship between G , H and S is given by:
G = H T S

(14)

Combining Eqs. (12) and (14) leads to:


ln Kd =

H
S

R
RT

(15)

The effect of the temperature on the PVA adsorption efciency


was evaluated at different temperatures ranging from 288 to 318 K.
At different temperatures, the corresponding Ce values for various
constant qe values were calculated. The thermodynamic parameters H and S were respectively calculated from the slope and
intercept of the plot of ln Kd versus 1/T, while the Gibbs free energies
were calculated from Eq. (12). The values of the thermodynamic
parameters G , H , and S for the PVA adsorption at different temperatures are given in Table 3. For all temperatures, G
was negative, which indicates the spontaneous nature of the PVA
adsorption process onto iron hydroxides. The increase in the absolute magnitude of G with increasing temperatures indicates that
the process is favorable at high temperatures. The positive value of

Table 3
Thermodynamic parameters of the PVA adsorption on the metal hydroxides at different temperatures.
T (K)

Thermodynamic
equilibrium constant (Kd )

288
298
308
318

3.916
6.344
9.288
10.747

G (kJ mol1 )
3.268
4.579
5.708
6.279

S (J mol1 K1 )

H (kJ mol1 )

102.71

26.16

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

H indicates that the adsorption process is endothermic. In addition, the positive value of S suggests the random character of
the PVA adsorption on iron hydroxides at the solidsolution interface. Even if the adsorption process were endothermic, the process
would be spontaneous because of the positive entropy change.
4. Conclusion
EC is an appropriate remediation process for treating aqueous
solutions containing PVA. Various operating parameters such as the
electrode pairs, the current density, the supporting electrolyte, and
the temperature were investigated in this study. The PVA removal
efciencies were approximately 94.9%, 68.7%, 47.3%, and 30.1% for
Fe/Al, Fe/Fe, Al/Al, and Al/Fe electrode pairs, respectively. The Fe/Al
electrode pair constitutes the best choice of electrode out of the
four combinations tested in this study. Considering the removal
efciency and EEC, a supporting electrolyte of 0.008N NaCl, a current density of 5 mA cm2 , and a solution temperature of 298 K
were found be the optimum values for the present study. Increasing the initial PVA concentration from 50 to 150 mg L1 led to a
decrease in the PVA removal efciency. A pseudo-rst-order model
and pseudo-second-order model were used to t the adsorption
kinetics. A very good agreement with the experimental data was
obtained for the pseudo-second-order kinetic model, which best
described the PVA adsorption onto iron hydroxides. The gelatinous
charged metal hydroxides generated during the EC can efciently
remove the PVA by adsorption. The EC was modeled using adsorption isotherm models. The adsorption of PVA was best tted by the
Langmuir adsorption isotherm, which suggests monolayer coverage of the adsorbed molecules. Various thermodynamic parameters
(G , H and S ) were also determined, and their values indicate that the adsorption process was favorable, spontaneous, and
endothermic in nature. As the temperature increased from 288 to
318 K, G became less negative, resulting in an enhanced adsorption capacity at high temperatures. The positive value of H
conrmed the endothermic nature of the process, which means that
the reaction consumed energy. The positive value of S indicated
the random nature of the PVA adsorption onto iron hydroxides
from aqueous solutions. Under these conditions, the maximum PVA
removal efciency was found to be 95%.
Acknowledgement
The authors would like to thank the National Science Council
of Taiwan, ROC for nancially supporting this study under contract
number NSC98-2221-E-241-004-MY2.
References
[1] Y. Tokiwa, G. Kawabata, A. Jarerat, A modied method for isolating poly (vinyl
alcohol)-degrading bacteria and study of their degradation patterns, Biotechnol. Lett. 23 (2001) 19371941.
[2] J. Veres, S. Ogier, G. Lloyd, Gate insulators in organic eld-effect transistors,
Chem. Mater. 16 (2004) 45434555.
[3] J.A. Giroto, R. Guardani, A.C.S.C. Teixeira, C.A.O. Nascimento, Study on the
photo-Fenton degradation of polyvinyl alcohol in aqueous solution, Chem. Eng.
Process. 45 (2006) 523532.
[4] K. Yoo, Sequential biological treatment including ozonation for persistent
organic compounds, PhD Thesis, Korea Advanced Institute of Science and Technology, Republic of Korea, 1999.
[5] H. Schonberger, A. Baumann, W. Keller, Study of microbial degradation of
polyvinyl alcohol (PVA) in wastewater treatment plants, Am. Dyestuff Rep. 86
(1997) 918.
[6] J.G. Lim, D.H. Park, Degradation of polyvinyl alcohol by Brevibacillus laterosporous: metabolic pathway of polyvinyl alcohol to acetate, J. Microbiol.
Biotechnol. 11 (2001) 928933.
[7] R. Solaro, A. Corti, E. Chillini, Biodegradation of poly(vinyl alcohol) with different molecular weights and degree of hydrolysis, Polym. Adv. Technol. 11 (2000)
873878.
[8] H. Yu, G. Gu, L. Song, Degradation of polyvinyl alcohol in sequencing batch
reactors, Environ. Technol. 17 (1996) 12611267.

849

[9] L.C. Lei, X.J. Hu, P.L. Yue, S.H. Bossmann, A. Gd, A.M. Braun, Oxidative degradation of polyvinyl alcohol by the photochemically enhance Fenton reaction, J.
Photochem. Photobiol A: Chem. 116 (1998) 159166.
[10] S.H. Bossmann, E. Oliveros, S. Gob, M. Kantor, A. Goppert, L. Lei, P.L. Yue,
A.M. Braun, Degradation of polyvinyl alcohol (PVA) by homogeneous and heterogeneous photocatalysis applied to the photochemically enhanced Fenton
reaction, Water Sci. Technol. 44 (2001) 257262.
[11] Y.X. Chen, Z.S. Sun, Y. Yang, Q. Ke, Heterogeneous photocatalytic oxidation
of polyvinyl alcohol in water, J. Photochem. Photobiol A: Chem. 142 (2001)
8589.
[12] A. Grnroos, P. Pirkonen, J. Heikkinen, J. Ihalainen, H. Mursunen, H. Sekki, Ultrasonic depolymerization of aqueous polyvinyl alcohol, Ultrason. Sonochem. 8
(2001) 259264.
[13] R.M. Hassan, New coordination polymers. III: oxidation of poly(vinyl alcohol) by
permanganate ion in alkaline solutions. Kinetics and mechanism of formation
of intermediate complex with a spectrophotometric detection of manganate
(VI) transient species, Polym. Int. 30 (1993) 59.
[14] S.J. Zhang, H.Q. Yu, Radiation-induced degradation of polyvinyl alcohol in aqueous solutions, Water Res. 38 (2004) 309316.
[15] Th.F. Tadros, Adsorption of polyvinyl alcohol on silica at various pH values and
its effect on the occulation of the dispersion, J. Colloid Interface Sci. 64 (1978)
3647.
[16] A.K. Bajpai, N. Vishwakarma, Adsorption of polyvinyl alcohol onto Fullers earth
surfaces, Colloids Surf. A: Phys. Eng. Asp. 220 (2003) 117130.
[17] S.G. Bussetti, E.A. Ferreiro, Adsorption of polyvinyl alcohol on Montmorillonite,
Clays Caly Miner. 52 (3) (2004) 334340.
[18] S.K. Behera, J.H. Kim, X. Guo, H.S. Park, Adsorption equilibrium and kinetics
of polyvinyl alcohol from aqueous solution on powdered activated carbon, J.
Hazard. Mater. 153 (2008) 12071214.
[19] E. Brillas, P.L Cabot, J. Casado, in: M. Tarr (Ed.), Chemical Degradation Methods
for Wastes and Pollutants Environmental and Industrial Applications, Marcel
Dekker, New York, 2003.
[20] D. Pletcher, F.C. Walsh, Industrial Electrochemistry, second ed., Blackie Academic & Professional, London, 1993.
[21] C.A.C. Sequeira, Environmentally Oriented Electrochemistry, Elsevier, Amsterdam, 1994.
[22] K. Rajeshwar, J.G. Ibanez, Environmental Electrochemistry: Fundamentals and
Applications in Pollution Abatement, Academic Press, London, 1997.
[23] K. Jttner, U. Galla, H. Schmieder, electrochemical approaches to environmental
problems in the process industry, Electrochim. Acta 45 (2000) 25752594.
[24] P.K. Holt, G.W. Barton, M. Wark, C.A. Mitchell, A quantitative comparison
between chemical dosing and electrocoagulation, Colloids Surf. A: Phys. Eng.
Asp. 211 (2002) 233248.
[25] I. Heidmann, W. Calmano, Removal of Zn(II), Cu(II), Ni(II), Ag(I) and Cr(VI)
present in aqueous solutions by aluminium electrocoagulation, J. Hazard.
Mater. 152 (2008) 934941.
[26] X. Chen, G. Chen, L.Y. Po, Separation of pollutants from restaurant wastewater
by electrocoagulation, Sep. Purif. Technol. 19 (2000) 6576.
[27] J.Q. Jiang, N. Graham, C. Andre, H.K. Geoff, N. Brandon, Laboratory study
of electro-coagulation-otation for water treatment, Water Res. 36 (2002)
40644078.
[28] C.T. Wang, W.L. Chou, L.S. Chen, S.Y. Chang, Silica particles settling characteristics and removal performances of oxide chemical mechanical polishing
wastewater treated by electrocoagulation technology, J. Hazard. Mater. 161
(2009) 344350.
[29] W.L. Chou, C.T. Wang, S.Y. Chang, Study of COD and turbidity removal
from real oxide-CMP wastewater by iron electrocoagulation and the evaluation of specic energy consumption, J. Hazard. Mater. 168 (2009)
12001207.
[30] C.T. Wang, W.L. Chou, Performance of COD removal from oxide chemical
mechanical polishing wastewater using iron electrocoagulation, J. Environ. Sci.
Health, Part A 44 (12) (2009) 12891297.
[31] S. I rdemez, N. Demircioglu, Y.S. Yildiz, Z. Bingl, The effects of current density
and phosphate concentration on phosphate removal from wastewater by electrocoagulation using aluminum and iron plate electrodes, Sep. Purif. Technol.
52 (2006) 218223.
[32] N. Mameri, H. Lounici, D. Belhocine, H. Grib, D.L. Piron, Y. Yahiat, Deuoridation
of Sahara water by small plant electrocoagulation using bipolar aluminium
electrodes, Sep. Purif. Technol. 24 (2001) 113119.
[33] J.R. Parga, D.L. Cocke, J.L. Valenzuela, J.A. Gomes, M. Kesmez, G. Irwin, H. Moreno,
M. Weir, Arsenic removal via electrocoagulation from heavy metal contaminated groundwater in La Comarca Lagunera Mexico, J. Hazard. Mater. B124
(2005) 247254.
[34] W.L. Chou, C.T. Wang, K.Y. Huang, Effect of operating parameters on indium
(III) ion removal by iron electrocoagulation and evaluation of specic energy
consumption, J. Hazard. Mater. 167 (2009) 467474.
[35] W.L. Chou, Yen-Hsiang Huang, Electrochemical removal of indium ions from
aqueous solution using iron electrodes, J. Hazard. Mater. 172 (2009) 4653.
[36] J.H. Finley, Spectrophotometric determination of polyvinyl alcohol in paper
coatings, Anal. Chem. 33 (1961) 19251927.
[37] M.Y.A. Mollah, R. Schennach, J. Parga, D.L. Cocke, Electrocoagulation (EC)science and applications, J. Hazard. Mater. B 84 (2001) 2941.
[38] P.R. Kumar, S. Chaudhari, K.C. Khilar, S.P. Mahajan, Removal of arsenic from
water by electrocoagulation, Chemosphere 55 (2004) 12451252.
[39] A.S. Koparal, .B. gtveren, Removal of nitrate from water by electroreduction
and electrocoagulation, J. Hazard. Mater. B 89 (2002) 8394.

850

W.-L. Chou / Journal of Hazardous Materials 177 (2010) 842850

[40] R.T. Foley, Localized corrosion of aluminum alloysa review, Corrosion 42


(1986) 277288.
[41] E.-S.Z. El-Ashtoukhy, N.K. Amin, O. Abdelwahab, Treatment of paper mill efuents in batch-stirred electrochemical tank reactor, Chem. Eng. J. 146 (2009)
205210.
[42] A.K. Golder, A.N. Samanta, S. Ray, Removal of phosphate from aqueous solutions
using calcined metal hydroxides sludge waste generated from electrocoagulation, Sep. Purif. Technol. 52 (2006) 102109.
[43] I.A. Sengil, M. zacar, B. mrl, Decolorization of C.I. reactive red 124 using
the electrocoagulation method, Chem. Biochem. Eng. Q 18 (2004) 273277.
[44] S. Lagergren, About the theory of so-called adsorption of soluble substance,
Kung Sven. Veten. Hand. 24 (1898) 139.

[45] Y.S. Ho, G. McKay, Kinetic models for the sorption of dye from aqueous solutionby wood, J. Environ. Sci. Health Part B: Process. Saf. Environ. Prot. 76 (1998)
183191.
[46] J. Walter, J. Weber, Physicochemical Processes for Water Quality Control, WileyInterscience, Canada, 1972.
[47] J.Q. Jiang, C. Cooper, S. Quki, Comparison of modied montmorillonite
adsorbents. Part I. Preparation, characterization and phenol adsorption,
Chemosphere 47 (2002) 711716.
[48] E.I. Unuabonah, K.O. Adebowale, B.I. Olu-Owolabi, Kinetic and thermodynamic
studies of the adsorption of lead (II) ions onto phosphate modied kaolinite
clay, J. Hazard. Mater. 144 (2007) 386395.

Das könnte Ihnen auch gefallen