Sei sulla pagina 1di 7

Chemical Engineering Journal 172 (2011) 137143

Contents lists available at ScienceDirect

Chemical Engineering Journal


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

Electrochemical phosphates removal using iron and aluminium electrodes

Engracia Lacasa, Pablo Canizares,


Cristina Sez, Francisco J. Fernndez , Manuel A. Rodrigo
Department of Chemical Engineering, Faculty of Chemical Sciences, University of Castilla-la Mancha, Avda. Camilo Jos Cela, 12, 13071 Ciudad Real, Spain

a r t i c l e

i n f o

Article history:
Received 15 January 2011
Received in revised form 4 May 2011
Accepted 19 May 2011
Keywords:
Electrocoagulation
Phosphates
Iron electrodes
Aluminium electrodes
Current density

a b s t r a c t
In the present work, the removal of phosphates from waters is studied through electrocoagulation
using iron and aluminium electrodes. This technology is an alternative to the conventional method of
coagulation, which leads to the complete removal of phosphates below the detectable limits of ionic
chromatography (0.1 mg dm3 ). The effect of the current density using both electrodes is also studied.
The results show that the pH increases with the current density. In addition, a mechanistic model is proposed for phosphate removal that considers the solubility of iron, aluminium and phosphate species, and
the zeta potential values. In the case of aluminium electrodes, the coexistence of both direct precipitation
and adsorption onto metal is observed, whereas in the case of iron electrodes, the adsorption mechanism
is less effective.
2011 Elsevier B.V. All rights reserved.

1. Introduction
The growth in the human population and the rise in the consumption of resources have increased the burden on aquatic
ecosystems and have affected the global biogeochemical cycles of
carbon, nitrogen, and phosphorous [1]. Moreover, the excessive
application of fertilisers, the intensive exploitation of farms and
the major contribution from industry have increased the nutrient load discharged into receiving waterways [2,3]. One of the
greatest recent problems of water resources is cultural eutrophication, which refers to a dramatic growth of algae in continental
and coastal waters. Cultural eutrophication is caused by the excess
discharge of phosphorous and nitrogen compounds from the efuents of municipal or industrial wastewater treatment plants into
the environment. In most cases, phosphorus is the limiting factor
in the eutrophication process and not nitrogen [4] because nitrogen
xation is naturally performed by diazotrophs. Therefore, most of
the recent nutrient-removal studies have focused on the removal
of phosphorus.
In the environment, the usual forms of phosphorus found
in solutions include orthophosphate, polyphosphate and organic
phosphate [5]. Nevertheless, the principal phosphorus compounds
in wastewater are generally orthophosphate forms together with
smaller amounts of organic phosphate [6,7]. The removal of
phosphorous from wastewater can be performed through physicochemical or biological processes. The most commonly used
physico-chemical processes are the following: chemical precip-

Corresponding author. Tel.: +34 902204100; fax: +34 926295318.


E-mail address: FcoJesus.FMorales@uclm.es (F.J. Fernndez).
1385-8947/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.cej.2011.05.080

itation including strong oxidant such as ferrate [8], adsorption,


ion-exchange, electrodialysis, hybrid systems containing y-ash
adsorption and membrane ltration and electrocoagulation [9].
However, the presence of organic phosphorous in aqueous solution makes necessary a pre-treatment of electrolysis in which these
compounds decrease their toxicity and increase their biodegradability [10,11].
The electrocoagulation process involves the in situ generation
of coagulants through the electro-dissolution of a sacricial anode,
which is usually made of iron or aluminium [12]. The reactions
involved in the electrochemical cell are summarised in Eqs. (1) and
(2). The main electrochemical reactions are the oxidation of the
metallic anode (Fe or Al) and the reduction of water. In the case
of iron electrodes, iron (II) is rapidly oxidised to iron (III), and the
system consequently behaves as if the iron (III) species were the
dosed reagent.
Anode : Al Al3+ + 3e or Fe Fe2+ + 2e

(1)

Cathode : H2 O + e (1/2)H2 + OH

(2)

In addition, the oxidation of water at the anode (Eq. (3)) is a


secondary reaction that competes with the oxidation of aluminium
or iron.
2H2 O O2 + 4H+ + 4e

(3)

The advantages of the electrocoagulation process over the conventional dosing of coagulants have been reported in several
studies [1315]. The main advantages are the simplicity of the
equipment and the ease of automation of the process. The process does not require the addition of chemicals, and therefore, the
coagulant dose can be controlled easily by varying the applied elec-

138

E. Lacasa et al. / Chemical Engineering Journal 172 (2011) 137143

Fig. 2. Layout of the electrochemical coagulation bench-scale plant. Detail of the


electrochemical cell.

supporting electrolyte type and current density. However, information regarding the mechanisms involved in the process and its
technical and economical feasibility is scarce.
The objective of the present work is to study the inuence of the
current density (i.e., the main operational parameter) on the electrocoagulation process using iron and aluminium, and to determine
the conditions in which the removal of phosphate is technically and
economically feasible.
2. Experimental methods
Fig. 1. (a) Solubility diagram of iron phosphate and iron hydroxide according to pH.
(b) Solubility diagram of aluminium phosphate and aluminium hydroxide according
to pH.

trical current. The low current requirement allows the use of green
energy sources such as solar cells, windmills and fuel cells [16].
The key point in the electrocoagulation process is the pH and
metal dosing. Electrocoagulation processes produce iron and aluminium hydroxides as the nal products, which slightly increase
the pH of the treated water. However, in the conventional chemical coagulation using non-electrochemical technology, the acidic
properties of the chloride or the sulphates of iron or aluminium
decrease the pH of the treated water [17]. Usually, the later use
of treated water requires neutralisation through the addition of
an acidic or a basic solution. The neutralisation of treated water
increases its salinity and thus reduces the quality of the treated
water.
However, coagulation and electrocoagulation perform equally
well when identical doses of the metal and pH are used to treat
wastewater using the two technologies [18]. The results depend on
the particular speciation of metals used as electrodes (iron or aluminium) and on the formation of insoluble phosphate precipitates
(AlPO4 or FePO4 ). Fig. 1 shows the solubility diagram of iron, aluminium and phosphate insoluble precipitates as a function of their
concentrations and pH. These diagrams have been widely reported
in literature [14,19].
Several studies on the removal of phosphate through electrocoagulation processes can be found in literature [7,15,2023].
These studies have investigated the inuence of the operational
parameters such as initial pH, supporting electrolyte concentration,

2.1. Experimental procedure


Bench-scale electrocoagulation studies were carried out to
characterize the treatability and the operational costs. The electrocoagulation experiments have been carried out in a bench-scale
plant which is shown in Fig. 2. The coagulant reagent was obtained
through the dissolution of the iron or aluminium electrodes placed
in a single-compartment electrochemical ow cell. Both electrodes
(anode and cathode) were square in shape (100 cm2 ), and the electrode gap was 9 mm. The electrical current was applied using a
DC power supply (FA-376 PROMAX). The current owing through
the cell was measured with a Keithley 2000 digital multimeter.
The range of current density studied was from 0.1 to 5.0 mA cm2
(experiments were carried out galvanostatically), being 7 V the
maximum exert potential monitored for the larger current density.
The synthetic wastewater was stored in a glass tank (5000 cm3 )
stirred by a Heidolph RZR 2041 overhead stainless steel rod stirrer and recirculated through the electrolytic cell using a peristaltic
pump. The synthetic water was composed of sodium phosphate
(27 mg PO4 3 -P dm3 ) as Na3 PO4 H2 O and a supporting electrolyte
to increase its conductivity (500 mg dm3 Na2 SO4 ).
2.2. Analysis procedure
Phosphate ions were determined, according to the standard
methods [24], through ion chromatography using a Shimadzu LC20A system by means of a Shodex IC I-524A column for anionic
separation. The mobile phase was an aqueous solution of 2.5 mM
phthalic acid with a pH of 4.0. The ow rate was 1 ml min1 .

E. Lacasa et al. / Chemical Engineering Journal 172 (2011) 137143

139

Fig. 3. Variation in the phosphorous concentration with time and specic electrical
charge during the electrocoagulation of phosphate solutions using iron electrodes
at () 0.1 mA cm2 , () 1.0 mA cm2 , () 3.0 mA cm2 and () 5.0 mA cm2 (discontinuous operation mode, 155 mg Na3 PO4 H2 O dm3 + 500 mg Na2 SO4 dm3 , pH not
modied).

The total aluminium or iron concentrations were measured offline using an inductively coupled plasma Liberty Sequential Varian
system according to the standard methods [24]. To determine the
total metal concentration, the samples were diluted 50:50 v/v with
4 N HNO3 to ensure the total solubility of the metal.
3. Results and discussion
Figs. 3 and 4 show the changes in the concentration of phosphorus with time when phosphate-loaded water (27 mg PO4 3 -P
dm3 ) was electrocoagulated (in discontinuous operation mode)
with the iron or aluminium electrodes at four different operating current densities. The PO4 3 -P concentration vs. time graphs
indicate the rate of the electrocoagulation process. In these last
graphs, the slope of the curves is directly proportional to the elec-

Fig. 4. Variation in the phosphorous concentration with time and specic electrical
charge during the electrocoagulation of phosphate solutions using aluminium electrodes at () 0.1 mA cm2 , () 1.0 mA cm2 , () 3.0 mA cm2 and () 5.0 mA cm2
(discontinuous operation mode, 155 mg Na3 PO4 H2 O dm3 + 500 mg Na2 SO4 dm3 ,
pH not modied).

Fig. 5. Electrodissolution of metal molar concentration with specic electrical


charge during the electrocoagulation of phosphate solutions at () 0.1 mA cm2 ,
() 1.0 mA cm2 , () 3.0 mA cm2 , () 5.0 mA cm2 and (solid line) Faradays Law.

trocoagulation rate, and consequently, inversely proportional to


the hydraulic residence time required for the completion of the
process. However, the graphs which show the PO4 3 -P concentration vs. the electric charge applied (Q) indicate the efciency of the
processes. The slope of these last curves is directly proportional
to the efciency of the process because the same electric charge
applied may allow to reach the same phosphorus removal using
different current densities.
In each case, the phosphates was removed completely (the
detection limit of the HPLC technique used was 0.1 mg dm3 ).
The current density increased the rate of the iron and aluminium
electrocoagulation [13,22], particularly at lower values of current
density, although the differences observed between the rates of the
process became negligible for current densities over 3.0 mA cm2 .
In terms of efciency, the observed trends for iron electrocoagulation differed signicantly with aluminium one. The results of the
electrocoagulation process with iron electrodes depended strongly
on the current density applied, and the efciency increased with the
use of low current densities. In the case of aluminium electrocoagulation, the current density had a smaller inuence on the results
than iron electrocoagulation, and only a small increase in efciency
was observed at the lowest current density. In addition, the electric charge applied required to completely remove the phosphorus
was lower in the case of aluminium, as it is reported in literature [7,13,15,22]. These aluminium results are comparable with the
lowest current density employed in iron electrocoagulation.
To better understand these results, the current charge-courses
of coagulation reagents (iron or aluminium) and pH were studied.
The two parameters are shown in Figs. 5 and 6, respectively. Surprisingly, the current density had a small effect on the dosage of iron
and aluminium, and only a small improvement could be estimated

140

E. Lacasa et al. / Chemical Engineering Journal 172 (2011) 137143

Fig. 6. Variation in pH with specic electrical charge during the electrocoagulation of phosphate solutions using (a) iron and (b) aluminium electrodes at ()
0.1 mA cm2 , () 1.0 mA cm2 , () 3.0 mA cm2 and () 5.0 mA cm2 .

Fig. 7. Variation in the zeta-potential with specic electrical charge during the electrocoagulation of phosphate solutions using (a) iron and (b) aluminium electrodes
at () 0.1 mA cm2 , () 1.0 mA cm2 , () 3.0 mA cm2 and () 5.0 mA cm2 .

at the higher current densities. However, the current density had


a signicant inuence in the phosphate removal depending on the
metal selected as electrode. The efciency in the production of the
coagulant reagent was clearly higher in the case of aluminium electrocoagulation in which the results overcame the expected values
according to the Faradays Law. This observation can be explained
by the chemical dissolution (corrosion) of the aluminium electrodes
reported in previous work [2527].
Therefore, in an electrocoagulation cell with aluminium electrodes, the aluminium added to the electrolyte comes from both the
chemical and the electrochemical dissolution of the electrodes. It is
worth noting that the anode and cathode materials were identical
in both cases. This is a normal practice in industrial electrocoagulation processes because it allows the inversion of the polarity
in order to reach comparable dissolution of the anode and cathode sheets in long-term processes. This allows to avoid operational
problems, which can be caused by the formation of carbonate lms
on the surface of the cathodes or by the passivation of the anodes.
On the contrary, the results obtained with iron were below this
electrochemical stoichiometric limit and could be explained by the
competition between iron and oxygen production (from water electrolysis) onto the anode surface. However, the production rate of
iron is almost constant during batch electrocoagulation processes
in which current density was maintained constant. Then, the iron
cannot be considered a limiting reagent in its electrodissolution
process because this process takes place from the oxidation of the
iron anode itself, and no from the electrooxidation of an iron solution which could have mass transfer limitations. In addition, the

efciency of iron production was also constant during all of the


experiments.
The pH parameter changed abruptly during the process, especially during an initial phase where the maximum removal of
phosphates is showed. The initial abrupt change was followed by a
gradual change to a nal pH value that appeared to be related to the
current density employed during the electrocoagulation process.
Thus, higher current densities resulted in higher nal pH values. The
changes in the pH during an electrocoagulation process are related
to the electrolysis of water and to the chemistry of the coagulation
process. For this reason, they are complex during the initial phosphorus removal stage and lead to a constant value at higher electric
charge applied. The increase in the nal pH value with the current
density is easily explained by Eqs. (1)(3). Hydroxyl ion production
is the main process at the cathode, whereas iron (or aluminium)
production competes with proton production at the anode. Consequently, a net production of hydroxyl anions is obtained in the
process and becomes larger at higher current densities. The range
of the nal pH values was wider in the case of iron, suggesting that
a different chemical process in the coagulation of phosphates will
take place with the two reagents (iron and aluminium). The differences appeared during the initial phase and were maintained
throughout the process.
Fig. 7 shows the variations in the z-potential during the electrocoagulation processes. The z-potential indicates the supercial
charge of the particles and can be used to estimate a mechanistic pathway for the electrocoagulation process, such as the process
proposed in Fig. 8.

E. Lacasa et al. / Chemical Engineering Journal 172 (2011) 137143

141

Fig. 8. Schematic of the electrocoagulation mechanism of phosphates.

The trend observed in Fig. 7 indicates the formation of negatively charged particles during the initial phase. The increase in
the negative charge during the initial phase lowered the value to
a minimum of approximately 35 mV in each case. Following the
initial lowering, an increase was observed in the voltage up to a
nal value that was close to 0 in the case of iron electrocoagulation, and a negative value that depended on the current density
in the case of aluminium electrocoagulation. The increase in the
negative charge observed during the initial phase in both iron and
aluminium electrocoagulation may be related to the adsorption of
phosphate anions, hydroxyl anions and metal hydroxoanions onto
the surface of the growing precipitate particles (this was true particularly in the case of aluminium because the negative charge was
almost negligible in the case of iron). The electrochemical dosing
of metal ions reversed the charge of the adsorbed particles and
increased the quantity of the precipitates. In the case of iron, the
process exhibited the complete neutralisation of the z-potential.
In the case of aluminium, the z-potential led to a constant value
but always remained negative, and a higher absolute value was
observed for higher current densities. Initially, this last part was
difcult to explain because the hydroxyl ions concentration was
lower than in the case of iron, as may be observed in Fig. 6. However, the observation in zeta potential values during aluminium
electrocoagulation was clearly explained by the adsorption of negatively charged aluminium monomeric anionic species (Al(OH)4 ),
which was negligible in the case of iron due to the low solubility
product of iron hydroxide.
Regarding the solubility diagram shown in Fig. 1 may be guess a
mechanistic pathway for phosphate removal in electrocoagulation
processes, although it may only be used as a rst approach due to
the signicant effect of other inorganic compounds contained in the
bulk liquid. Thus, in the case of aluminium, both aluminium phosphate (AlPO4 ) and aluminium hydroxide (Al(OH3 )) competed over
the range of pH values used [7,22,27]. Then, lower current densities resulted in lower pH values and hence favoured the aluminium
phosphate formation. However, in the case of iron, the solubility of
iron phosphate (FePO4 ) was higher than iron hydroxide (Fe(OH3 )),
except at the lower pHs which were reached at the lowest current

densities. In this last context, iron hydroxide may be afrmed that


was the main coagulation product above all at higher current densities. Therefore, the particles of amorphous metal hydroxide were
formed in both systems while the metal phosphate was formed
only at very low current densities (the lowest pH values) during iron
electrocoagulation, or in aluminium electrocoagulation throughout
all processes as it showed in Fig. 1.
All of these stages are summarised graphically in the mechanistic model shown in Fig. 8. It is very important to point out the
signicant inuence of the pH on the results. Fig. 9 shows the relationship between the nal pH and the metal dose required for the
complete removal of phosphates. Stoichiometry close to one was
maintained for iron and aluminium at the lowest current density
and at the lowest pH values, suggesting that the main mechanism
in this case was the direct precipitation of the metal phosphate.
The required metal dose increased with the pH and hence with the
current density, suggesting that adsorption onto metal hydroxide

Fig. 9. Relationship between the nal pH and the metal dose required for the complete removal of phosphates during the electrocoagulation process using iron ()
and aluminium () electrodes.

142

E. Lacasa et al. / Chemical Engineering Journal 172 (2011) 137143

improved at the lower current densities. The solubility diagram


indicated that in the case of aluminium, both aluminium phosphate and aluminium hydroxide competed in the range of the pH
values used, although the formation of aluminium phosphate was
favoured by the low current densities. Due to the higher solubility of the iron phosphate, iron hydroxide was the main product
of coagulation when iron electrodes were used. In addition, the
pH exhibited a signicant inuence on the results. Hence, the
required metal dose to remove completely phosphates increased
with the pH and subsequently with the current density, indicating
that adsorption onto the metal hydroxide was a signicant pathway
in the phosphate removal process although it was less efcient than
metal phosphate direct precipitation. In the case of aluminium, a
coexistence of both direct precipitation and adsorption onto metal
was observed, whereas in the case of iron the adsorption mechanism was less effective. Then, the optimum conditions to remove
phosphates by electrocoagulation are at the lowest current densities for iron and aluminium electrodes because it is favoured the
direct metal phosphate precipitation, although if current density
increases the required reagent dose and the electric consumption
for aluminium electrocoagulation are lower than using iron electrodes. Finally, it is important to highlight that the nal pH of the
treated water was close to neutral in all cases.
Acknowledgements
This work was supported by the MCT (Ministerio de Ciencia y Tecnologia, Spain) and the EU (European Union) through
projects CTM2007-60472/TECNO, CTM2010-18833/TECNO and
CONSOLIDER-INGENIO 2010 (CSD2006-044).
References
Fig. 10. Requirement of reagent dose (a) and electric consumption (b) for the complete removal of phosphates during the electrocoagulation process using iron ()
and aluminium () electrodes.

was a signicant pathway in the phosphate removal process. The


smaller required dose of aluminium compared to iron indicated the
coexistence of both mechanisms, whereas the exponential decrease
observed for iron indicated the importance of the less effective
adsorption mechanism.
Fig. 10 summarises the electrode and power consumption for
the complete removal of the phosphate under the different current
densities. As can be observed, the low current densities preserved
the raw metal and reduced the energy consumption. It is also
important to note that under these operation conditions, the pH
of the treated water was close to neutral, which is important
because the neutralisation of the treated water is usually required
in coagulation processes, which increases the salinity and therefore
reduces the quality of the treated water. The metal electrode and
energy consumption varied in the range 18.867.4 g metal/m3 and
0.060.73 kWh/m3 for Al electrode and 27.4586.4 g metal/m3 and
0.054.38 kWh/m3 for Fe electrode with increasing current density
from 10 to 50 A/m2 . These ranges observed for electrode consumptions were in the same magnitude order than the found in literature
[22], although in this work the energy consumption values were
lower.
4. Conclusions
The concentration of phosphates in the aqueous solution
decreased to under 0.1 mg dm3 during the electrocoagulation
experiments using iron and aluminium electrodes. The current density was an important parameter in the phosphate removal process
above all when iron electrodes were used, where the efciency

[1] G. Friedl, A. West, Disrupting biogeochemical cyclesconsequences of


damming, Aquat. Sci. 64 (2002) 5556.
[2] P.M. Vitousek, H.A. Mooney, J. Lubchenco, J.M. Melillo, Human domination of
Earths ecosystem, Science 277 (1997) 494499.
[3] C. Sommariva, A. Converti, M. Del Borghi, Increase in phosphate removal from
wastewater by alternating aerobic and anaerobic conditions, Desalination 108
(1996) 255260.

[4] F.J. Fernndez, J. Villasenor,


L. Rodrguez, Effect of the internal recycles on the
phosphorus removal efciency of a WWTP, Ind. Eng. Chem. Res. 46 (2007)
73007307.
[5] G. Tchobanoglous, F.L. Burton, Wastewater Engineering, McGraw-Hill, 1991.
[6] D.G. Grubb, M.S. Guimaraes, R. Valencia, Phosphate immobilization using an
acidic type F y ash, J. Hazard. Mater. 76 (2000) 217236.
[7] S. I rdemez, Y.S. Yildiz, V. Tosunoglu, Optimization of phosphate removal from
wastewater by electrocoagulation with aluminum plate electrodes, Sep. Purif.
Technol. 52 (2006) 394401.
[8] C. Stanford, J.-Q. Jiang, M. Alsheyab, Electrochemical production of ferrate (iron
VI): application to the wastewater treatment on a laboratory scale and comparison with iron (III) coagulant, Water Air Soil Pollut. 209 (2010) 483488.
[9] H.-G. Kim, H.-N. Jang, H.-M. Kim, D.-S. Lee, T.-H. Chung, Effect of an electro
phosphorous removal process on phosphorous removal and membrane permeability in a pilot-scale MBR, Desalination 250 (2010) 629633.
[10] N.A. Salles, F. Fourcade, F. Geneste, D. Floner, A. Amrane, Relevance of an
electrochemical process prior to a biological treatment for the removal of an
organophosphorous pesticide, phosmet, J. Hazard. Mater. 181 (2010) 617623.
[11] F. Kirzhner, Y. Zimmels, Y. Shraiber, Combined treatment of highly contaminated winery wastewater, Sep. Purif. Technol. 63 (2008) 3844.

C. Jimnez, F. Martnez, C. Sez, M.A. Rodrigo, Study of the electro[12] P. Canizares,


coagulation process using aluminum and iron electrodes, Ind. Eng. Chem. Res.
46 (2007) 61896195.
[13] 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 (2) (2006) 218223.
[14] S. I rdemez, N. Demircioglu, Y.S. Yildiz, The effects of pH on phosphate removal
from wastewater by electrocoagulation with iron plate electrodes, J. Hazard.
Mater. 137 (2) (2006) 12311235.
[15] S. Vasudevan, J. Lakshmi, J. Jayaraj, G. Sozhan, Remediation of phosphatecontaminated water by electrocoagulation with aluminium, aluminium alloy
and mild steel anodes, J. Hazard. Mater. 164 (2009) 14801486.
[16] M.Y.A. Mollah, P. Morkovsky, J.A.G. Gomes, M. Kesmez, J. Parga, D.L. Cocke,
Fundamentals, present and future perspectives of electrocoagulation, J. Hazard.
Mater. 114 (2004) 199210.

E. Lacasa et al. / Chemical Engineering Journal 172 (2011) 137143

[17] P. Canizares,
C. Jimnez, F. Martnez, M.A. Rodrigo, C. Sez., The pH as a key
parameter in the choice between coagulation and electrocoagulation for the
treatment of wastewaters, J. Hazard. Mater. 163 (2009) 158164.

[18] P. Canizares,
F. Martnez, C. Jimnez, J. Lobato, M.A. Rodrigo, Coagulation and
electrocoagulation of wastes polluted with colloids, Sep. Sci. Technol. 42 (2007)
21572175.

F. Martinez, J. Lobato, M.A. Rodrigo, Electrochemically assisted


[19] P. Canizares,
coagulation of wastes polluted with Eriochrome Black T, Ind. Eng. Chem. Res.
45 (2006) 34744348.
[20] N. Bektas, H. Akbulut, H. Inan, A. Dimoglo, Removal of phosphate from aqueous
solutions by electro-coagulation, J. Hazard. Mater. 106 (2004) 101105.
[21] S. Vasudevan, G. Sozhan, S. Ravichandran, J. Jayaraj, J. Lakshmi, S.M. Sheela,
Studies on the removal of phosphate from drinking water by electrocoagulation
process, Ind. Eng. Chem. Res. 47 (2008) 20182023.
[22] M. Kobya, E. Demirbas, A. Dedeli, M.T. Sensoy, Treatment of rinse water from
zinc phosphate coating by batch and continuous electrocoagulation processes,
J. Hazard. Mater. 173 (13) (2010) 326334.

143

F. Martinez, M. Carmona, J. Lobato, M.A. Rodrigo, Continuous elec[23] P. Canizares,


trocoagulation of synthetic colloid-polluted wastes, Ind. Eng. Chem. Res. 44
(22) (2005) 81718177.
[24] APHA-AWWA-WPCF, Standard Methods for the Examination of Water and
Wastewater, 17th ed., American Public Health Association, Washington, DC,
1989.

M. Carmona, J. Lobato, F. Martnez, M.A. Rodrigo, Electrodissolu[25] P. Canizares,


tion of aluminum electrodes in electrocoagulation processes, Ind. Eng. Chem.
Res. 44 (2005) 41784185.
[26] X. Chen, G. Chen, P.L. Yue, Separation of pollutants from restaurant wastewater by electrocoagualation, Sep. Purif. Technol. 19 (2000)
6576.
[27] S. Tchamango, C.P. Nanseu-Njiki, E. Ngameni, D. Hadjiev, A. Darchen, Treatment
of dairy efuents by electrocoagulation using aluminium electrodes, Sci. Total
Environ. 408 (2010) 947952.

Potrebbero piacerti anche