Você está na página 1de 11

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 2 7 3 7

Available online at www.sciencedirect.com

ScienceDirect
www.journals.elsevier.com/journal-of-environmental-sciences

San copolymer membranes with ion exchangers for Cu(II)


removal from synthetic wastewater by electrodialysis
Simona Caprarescu 1 , Mihai Cosmin Corobea 2,, Violeta Purcar 2 , Catalin Ilie Spataru 2 ,
Raluca Ianchis 2 , Gabriel Vasilievici 2 , Zina Vuluga 2
1. Politehnica University of Bucharest, Faculty of Applied Chemistry and Materials Science, Department of Inorganic Chemistry, Physical
Chemistry and Electrochemistry, Calea Grivitei, no. 132, 010737, Bucharest, Romania
2. National Research and Development Institute for Chemistry and Petrochemistry ICECHIM, Polymer Department, Splaiul Independentei,
no. 202, 060021, Bucharest, Romania

AR TIC LE I N FO

ABS TR ACT

Article history:

Heterogeneous membranes were obtained by using styrene-acrylonitrile copolymer (SAN)

Received 3 November 2014

blends with low content of ion-exchanger particles (5 wt.%). The membranes obtained by phase

Revised 15 December 2014

inversion were used for the removal of copper ions from synthetic wastewater solutions by

Accepted 13 February 2015

electrodialytic separation. The electrodialysis was conducted in a three cell unit, without

Available online 27 May 2015

electrolyte recirculation. The process, under potentiostatic or galvanostatic control, was followed
by pH and conductivity measurements in the solution. The electrodialytic performance,

Keywords:

evaluated in terms of extraction removal degree (rd) of copper ions, was better under

SAN copolymer

potentiostatic control then by the galvanostatic one and the highest (over 70%) was attained at

Membrane

8 V. The membrane efficiency at small ion-exchanger load was explained by the migration of

Ion exchange

resin particles toward the pores surface during the phase inversion. The prepared membranes

Copper ions

were characterized by various techniques i.e. optical microscopy, Fourier transform infrared

Electrodialysis

spectroscopy, scanning electron microscopy, thermogravimetric analysis and differential

Wastewater treatment

thermal analysis and contact angle measurements.


2015 The Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences.
Published by Elsevier B.V.

Introduction
Water contamination with cooper ions is still proven as a serious
issue for the decontamination processes, in the recent years.
Exceeding concentrations above 2 mg/L, from industrial wastewaters, increase drastically the toxicity, sometimes with irreversible damage on the environment (Barakat, 2011; Fu and
Wang, 2011; Hashim et al., 2011; Peng et al., 2011; Vinodh et al.,
2011).
The electrochemical recovery of heavy metal ions (as the
copper ones), from sea water and industrial wastewaters may

be recovered for recycling. Deionization processes such as


electrodeionization (Arar et al., 2011, 2014a), reverse osmosis
(Arar et al., 2014a) and electrodialysis (Rodrigues et al., 2008;
Tzanetakis et al., 2003) are some major alternatives. The
separation principle, the advantages and disadvantages can
be considered as follows.
The electrodeionization method is a combination between
electrodialysis and ion exchange methods (Arar et al., 2011,
2014b). This method has already been applied for the removal
of Cu2 +, Ni2 + and Cr6 + ions from the dilute solutions (Dzyazko,
2006; Dzyazko et al., 2008; Spoor et al., 2002; Mahmoud et al.,
2007; Feng et al., 2007). But in these cases the process is

Corresponding author. E-mail: mcorobea@yahoo.com (Corobea Mihai Cosmin).

http://dx.doi.org/10.1016/j.jes.2015.02.005
1001-0742 2015 The Research Center for Eco-Environmental Sciences, Chinese Academy of Sciences. Published by Elsevier B.V.

28

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 27 3 7

complicated by the deposition of insoluble compounds on the


ion exchange membrane. Therefore the reported removal of
Cu2 + ions by electrodeionization starts usually from diluted
initial solutions and uses high performance membranes like
Nafion (Mahmoud et al., 2007).
The reverse osmosis process uses pressure to force a solution
through a semi-permeable membrane that retains the solute on
one side and allows the pure solvent to pass to the other side. The
removal of heavy metals from synthetic wastewater (like copper
and cadmium ions) was investigated by reverse osmosis (Qdais
and Moussa, 2004). In this case high removal efficiency is
available (98% and 99% for copper and cadmium, respectively),
but reverse osmosis in contrast with normal osmosis requires
high pressure for the separation. Therefore the major disadvantages of reverse osmosis are the high power consumption needed
for the pressure pump and the restoration of the membranes.
Due to the high pressure requirements (even 200 psig or more
above the osmotic pressure) reverse osmosis is usually not
applicable for concentrated solution (Fu and Wang, 2011).
Electrodialysis is an electrochemical separation process, in
which selective separation of certain ions from water solution
is made. The phenomena consist in the migration of ions,
across the membranes under the electric field influence. The
electrodialytic cell contains at least one or more anionic and
cationic membranes, alternating between two electrodes.
The properties of the membrane used in the process are
very important because these determine the degree of
separation in electrodialysis (Nasef and Gven, 2012;
Strathmann, 2010). The electrodialysis membrane requirements include good thermal and chemical stability, since the
process can be run at elevated temperatures, in solution of
very low or high pH values (Caprarescu et al., 2012).
Electrodialysis has been extensively used in the
electroplating, metal finishing and metallurgy in general, to
remove and recycle metal ions such as copper, lead, zinc,
silver, among other metals from plating bath rinse solutions
or to remove inert electrolyte salts that build-up during
plating (Barakat, 2011; Fu and Wang, 2011; Nasef and Gven,
2012; Sadrzadeh et al., 2008).
Few reports are describing the use of styrene-acrylonitrile
copolymer (SAN), alone as membrane materials. However in
some combinations (with cellulose acetate, or butadiene) SAN
copolymer have been proven as useful membrane material (in
dialysis, ultrafiltration, enzyme-immobilization, molecular
imprinting and pervaporation) (Joshi et al., 2001; Fritzsche,
1986; Radha et al., 2009, 2014; Silva et al., 2008; Murthy et al.,
2012). From the best of our knowledge SAN was not yet
exploited in electrodialysis, despite its optimal profile in
terms of chemical resistance, thermal resistance, high
strength, dimensional stability, processability and durability
(Murthy et al., 2012). These high performances are completed
by a certain versatility of polymer material processing,
availability and relatively low cost (as supply material).
Therefore, such a stable copolymer could bring benefits in
electrodialysis separations in terms of costs and stability over
a large range of pH (since the ion exchange resins alone are
either brittle for the cation resins or too soft in the case of
anion resins). The reduction of the ion exchange resin
consumption in the separation process would be an advantage since SAN copolymer is several times cheaper.

The objective of this work was to investigate the feasibility


of obtaining SAN-ion exchange membranes and their use in
the removal of copper ions from a synthetic wastewater
using low amounts of ion-exchangers (5 wt.%). The proposed
approach emphasizes new cost effective solution in membranes for electrodialysis process for water purification.
Electrodialysis experiments were operated both in
potentiostatic and galvanostatic modes in order to compare
them in view of percent extraction of copper ions. The
membranes have been characterized using Optical Microscopy, Fourier transform infrared spectroscopy, scanning
electron microscopy, contact angle measurements, thermogravimetric analysis and differential thermal analysis.

1. Experimental
1.1. Materials
The chemical reagents used in these experiments were of
analytical grade. In all experiments fresh distilled water was
used for the preparation of aqueous synthetic electroplating
wastewaters.
The styrene-acrylonitrile copolymer (SAN) was a commercial product (Luran 358N, from BASF PLASTICS,
Ludwigshafen, Germany) and N, N-dimetylformamide (DMF)
(Scharlab S.L., Barcelona, Spain; 99.5%). The ion exchange
resins were the strong acidic cation exchanger Puropack
PPC100 (PPC100) and the weak basic anion exchanger Purolite
A100 (A100) both from Purolite, Romania. Their initial ionic
forms were Na+ and Cl respectively (Fig. 1).
Copper sulfate pentahydrate (CuSO45H2O) (Chimopar,
Romania) and sulfuric acid 98% (Merck) were used as received.
A stock solution of Cu(II), containing 1 g Cu(II)/L, was prepared
by dissolving both copper sulfate pentahydrate and sulfuric
acid (molar ratio 1:1) in distilled water.

1.2. Membrane preparation


In order to obtain the heterogeneous membrane a typical
procedure was followed. A mixed matrix was prepared as
follows. 8 wt.% SAN was purred in DMF under mechanical
stirring (400 r/min at 40C) until a clear solution was
obtained. Then 5 wt.% ion exchanger resin, previously
processed by ball milling up to a fine powder grain size of
2.5 m was added to polymer. In order to obtain a uniform
dispersion, the mix was stirred for several hours, then
sonicated for a few minutes and left over night, in order to
achieve the air removing and swelling of the polymer
solution into the pores of the resin particles.
The day after, the dispersion was again stirred for 1 hr
and then used for casting the membranes. The dispersion
was thus spread with a film applicator (0.45 mm) on a glass
panel. Immediately after coating, this panel was immersed
in a coagulation bath filled with distilled water at 22C.
Within a few minutes, the membrane detaches itself from
the glass support, then it immersed again in a fresh water
bath (the procedure of fresh water immersion was repeated 5
times) and finally left after night in water, in order to remove
completely the DMF solvent. The membrane was kept

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 2 7 3 7

Cation exchange site

29

case of a potentiostatic control). A power supply (Protek,


Germany) ensuring up to 30 V and up to 5 A was used.
The solution conductivity was measured using a PIERRON
(France) conductometer and the pH measurements were
performed with a Hanna Instruments HI 8915 pH-meter using
a combined glass electrode. The copper ions concentration was
determined by titration as previously reported (Caprarescu et
al., 2012).

1.4. Methods for characterization of membranes

n
Anion exchange site

Fig. 1 Idealized structure of acidic cation exchanger,


Puropack PPC100 (a) basic anion exchanger, Purolite A100 (b)
used for membrane obtaining.

permanently immersed at constant temperature in water


until the measurements have begun. The thickness of wet
membranes was between 0.012 and 0.023 cm.
Before pouring the grain particles of ion exchange resins into
the membranes, they were brought to the suitable ionic form,
namely H+ for the cation exchanger and OH for the anion one, by
repeated immersion into concentrated aqueous solutions of 23%
HCl respectively 25% NH3.

Fourier transform infrared spectroscopy (FT-IR) was done


using a Bruker-Tensor 37 instrument. Samples were analyzed
in attenuated total reflection (ATR) module with a Golden
Gate diamond unit.
The surface morphology of the SAN anion-exchange
membranes was followed by an optical microscope (MC 5,
IOR, Bucharest, Romania) with color high resolution video
camera and scanning electron microscopy (SEM), using a
S-2600N, HITACHI, Japan, electron microscope. All the
membrane samples were dipped in liquid nitrogen and gold
coated with a thin layer, before SEM analysis. Water contact
angle (CA) was measured at room temperature using a
Contact Angle Tensiometer (CAM 200, KSV Instruments,
Helsinki, Finland). Static contact angle measurements were
done on 2 cm 2 cm specimens (droplet 10 L, 2 CA error).
Thermogravimetrical and differential thermal analysis of the
membranes was performed by using a Du Pont TGA 2000
instrument (heating rate of 10C/min in air).

2. Results and discussions


1.3. Three-compartment electrodialysis cell
2.1. Electrodialytic current and voltage time profiles
Experimental tests were performed in a laboratory device for
electrodialysis of own construction, composed of three detachable cylindrical compartments reinforced by gaskets, made of
clear acrylic, separated by the two types of SAN/ion exchange
heterogeneous membranes (anionic respectively cationic) with
active areas of 23.316 cm2 and two working plan parallel
electrodes (Fig. 2). The electrodes, both of pure copper (99.9%)
and presenting the same effective area of 23.32 cm2 were
disposed at a distance of 3.6 cm as extremities of the electrodialysis cell. Each compartment has an external diameter of
9.75 cm, inside diameter of 5.45 cm and thickness without
silicone gaskets of 0.918 cm. The thickness of each compartment
with silicone gaskets was 1.279 cm. The distance between the
inner faces of the membranes was 1.2 cm.
In the electrodialysis cell, SAN/anion-exchange membrane
separates the central compartment from the anode, and in a
similar manner the SAN/cation-exchange membrane was placed
between the central compartment and the cathode. The liquid
volume was 29.85 cm3 in each of the three compartments. All
experiments were carried out during 90 min, at room temperature (22 1C) and without recirculation. The three compartments were filled with the same solution.
Electrodialysis was carried out either applying an
anode-cathode current of 0.05 A and of 0.1 A (current density
2.14 mA/cm2 respectively 4.29 mA/cm2) under galvanostatic
control, or an anode-cathode voltage of 6 V and of 8 V, (in the

In acid medium, the following reactions occur at the electrodes during the electrodialysis of water containing copper
ions:
Anode : 2H2 OO2 g 4H 4e

Fig. 2 Three compartment electrodialysis laboratory unit


representation.

30

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 27 3 7

Cathode : 2H 2e H2 g

Cu2 2e Cus

Cation exchange membrane : 2 H CER Cu2 aq 2 H aq


Cu2 aq CER

Anion exchange membrane : 2 OH AER SO4 2 aq 2 OH aq


SO4 2 aq CER

Energy consumption also increased with applied voltage as


expected. These results are in good agreement with the ones
reported in literature (Gven and Karabacakolu, 2005; Kabay et
al., 2002).
Although the investigated membranes were not entirely
made of ion exchangers, they behave in the electrodialysis
likewise high-performance membranes. The behavior was
similar like the ones reported in literature for example in the
removal of Ag+ ions with Nafion 424 and Ionac
MA-on-exchange membranes. Later, the current decreased
rapidly with time, while the Ag+ ions are transferred from the
dilute side to the concentrate side. On the other hand, during
the operation, the initial concentration of Ag+ ions in the treated
water decreased with time (Gven and Karabacakolu, 2005).

where the subscripts aq, CER and AER designate respectively


the aqueous solution, the cation exchange resin and the anion
exchange resin.
In order to prevent the fall of conductance by accumulation of the gaseous products H2 and O2, one small hole on the
top of each compartment was made to release gases from the
electrodialysis cell.

2.1.2. Constant current operation


In galvanostatic mode, a constant current was imposed between
electrodes. The current flowing through the cell is kept constant
all over the batch operation; the potential drop across the cell,
E = I R varies with time.
In these studied conditions, when constant current was
used (Fig. 3b), the voltage value gradually increases. The
voltage increase occurs, because during the separation process, the ions migrate toward the electrodes and the conductivity of the water in the central compartment decreases.
Therefore as the concentration of ions decreases, the resistance gradually increases as well.
At higher current intensity at 0.1 A (i.e. under a current
density of 43 A/m2), the cell voltage, was near two times
greater than at 0.05 A (which corresponds to a current
density 21 A/m2). Thus the electric resistance of cell
remains fast independent of the current intensity, but
shows an increase of 2.5 to 3 times during the process.
Again heterogeneous SAN-ion exchange membranes
showed similar results in the electrodialysis process in comparison with high performance membranes found in the
existing literature (Gven and Karabacakolu, 2005; Chen et
al., 2009; Dalla Costa et al., 2002; Tanaka, 2002; Cifuentes et al.,
2009; Parulekar, 1998; Melnyk and Goncharuk, 2009). The
experimental results confirmed that the potentiostatic control
is safer then the galvanostatic control, by avoiding the

2.1.1. Constant voltage operation


Cell voltage E (V) as the driving force of the electrodialysis
determines the migration rate of Cu2+ ions across the
membranes. In potentiostatic operation control, the current I
(A) flowing through the electrodialysis cell vary with the time t
(hr). The potential fall E between the cell electrodes is
proportional to the cell resistance R, according to the
relationship I = E/R.
Fig. 3a shows the progress of the electrodialysis with the
proposed membranes for two values of applied voltage. To note
that in both cases (6 V and 8 V) the current gradually fall with
time, since the concentration of Cu2+ ions in all three aqueous
solutions decreased leading to an increase in the cell resistance.
The explanation for the phenomena could be given from the ion
migration rate (responsible for the high conductivity of the initial
electrolyte solution). In contrast with the later experiment
(Fig. 3b) the migration seems faster at constant voltage, then in
constant current mode. Further the separation degree should be
of higher aspect, which will be clarified later in the removal
degree section. As reported in the literature, the amount of ions
transported through the membrane is directly proportional to the
electrical current or current density. If the electrical potential
difference was increased, the current density will increase.

0.45

10

0.40

0.35
Cell voltage (V)

Current (A)

0.30
0.25
0.20
0.15
0.10

10

20

7
6
5
4
3
2

6V
8V

0.05
0.00

0.05 A
0.10 A

1
30

40 50 60
Time (min)

70

80

90

100

10

20

30

40

50 60
Time (min)

70

80

90

100

Fig. 3 Time decrease of current intensity at constant voltage (a) and time increase of cell voltage at constant current (b).

31

14.26
7.49
2.44
2.16
14.26
6.06
4.68
2.04
14.26
7.65
2.78
1.22
14.26
7.34
2.84
3.94
1.60
2.60
3.15
2.57
1.60
3.33
2.32
2.47

0.1 A
0.05 A
8V
6V
0.1 A

pH

Galvanostatic
control

Table 1 presents the pH values for the different solutions for the
potentiostatic and galvanostatic control, measured after 1.5 hr, at
room temperature. The solutions from Table 1 are acidic all
present low pH values in good agreement with the data reported
in literature, hence H+ compete with Cu2 + at the cathode
surface resulting in hydrogen embitterment in the copper
deposits (Caprarescu et al., 2011; Kabay et al., 2002; Chen et
al., 2009). The pH of the solution is not considered to affect
the removal rate and there is no need to adjust the pH and
hence to add additional reagents before and after treatment
(Dalla Costa et al., 2002; Tanaka, 2002).

0.05 A

2.1.3. Final pH

Potentiostatic
control

Conductivity (mS/cm)

concentration
polarization phenomena
(Gven and
Karabacakolu, 2005; Chen et al., 2009; Dalla Costa et al., 2002;
Tanaka, 2002) and development of high voltages at low final
concentration in the final electrolyte.
Electrodialysis process operated both in potentiostatic
and galvanostatic modes for the treatment of metal
finishing wastewater was described in literature using
Nafion 450 and Selemion AMP. Results indicated that
potentiostatic control could be safer to operate in the sense that
it does not permit the development of high voltages across the
stack when the ion concentration in the diluate is excessively
low, although each species of metallic ions exhibited a different
rate of extraction from the treated solution. This operation mode
avoids the occurrence of concentration polarization and the
problems related to it (Dalla Costa et al., 2002).

Galvanostatic control

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 2 7 3 7

rd

C in;cat C fi;cat
 100%
C in;cat

where, Cin,cat (mol/dm3) and Cfi,cat (mol/dm3) are the initial,


respectively the final concentration of the copper ion in the
cathodic compartment.

1.60
3.31
3.24
5.15
1.60
3.25
3.03
2.17

8V
6V

Potentiostatic
control
Samples

Untreated wastewater
Treated water from anodic compartment
Treated water from central compartment
Treated water from cathodic compartment

The laboratory electrodialysis cell performance was evaluated in terms of extraction removal degree (rd) of copper in
the cathodic compartment, defined by:

0
1.5

2.1.5. Final removal degree and energy consumption

Time (hr)

The electrical conductance of the system is determined by


the ohmic resistances of the ion-exchange membranes and
by the concentrations of different ions in solutions. The
final changes in solution conductivity for anodic, central
and cathodic compartments are shown in Table 1.
In both cases (potentiostatic and galvanostatic control) the
conductivity decreased in all three compartments during the
electrodialysis. The decrease was moderate in the anodic
compartment and more important in the central cell compartment (indicating the removal of Cu2+ and SO2
4 ions and in the
cathodic compartment because of the copper electrodeposition
reaction on the cathode Caprarescu et al., 2012; Dalla Costa et al.,
2002).
In three of the four experiments, the conductivity fall was
the most pronounced for the cathodic compartment and in a
single one (namely, at potentiostatic control under lower
voltage) this decline was the most important for the central
compartment.

Table 1 Values of solution pH for the potentiostatic and galvanostatic control.

2.1.4. Final solution conductivity

32

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 27 3 7

The specific energy consumption (E.C., kWh/m3) necessary


in an electrodialysis process is given by Eq. (7):
E:C:

1

1000  V cat

tt fin
t0

Et  It  dt

where, E (V) is the applied potential, tfin (hr) is the final time
and Vcat (m3) is the volume of the cathodic compartment.

2.2. Optimal conditions for electrodialysis


The cell voltage applied in the electrodialysis is considered to be one of the most important operational parameters, which can affect radically the copper recovery and
energy consumption. As can be seen from Table 2, rd
values for copper ions increase from 64.69% to 70.31%
with the values (6 and 8 V) of the current or voltage used
in the electrochemical setup (Rodrigues et al., 2008;
Caprarescu et al., 2012; Dzyazko, 2006; Spoor et al., 2002;
Dzyazko et al., 2008). It was demonstrated that a high
copper ion concentration (1 g/L Cu2 + ) can be removed in
90 min of the electrodialysis operation performed under
two constant applied voltages at the room temperature.
For the same operating time, the purification efficiency rd
is a bit lower under galvanostatic control 55.63% and
65.00%; but the energy consumption is much better in the
case of galvanostatic control.
During the process study, no precipitations in membranes
were observed; the solutions remain clear after measurements. The membranes were not damaged during the
electrodialysis process and could be reused.
The results obtained with heterogeneous SAN-ion exchange membranes were comparable with the existing
literature on electrodialysis with high performance membranes. For example using Ionac MA3475 as anion, and
respectively Nafion423 as cation exchange membranes, the
reported removal degree was 52.6% by an energy consumption of 5 kWh/m3 at 20 V, in a period of 75 min, for an initial
concentration of 100 mg Cu2+/L water. Our result for 0.05 A
was very similar: a bit better on the removal (55.6%), for a bit
higher energy consumption of 6.5 kWh/m3 (tveren et al.,
1997). In our setup, separations over 70% can be achieved in
90 min in comparison with other commercial products were
120 min are needed (Chang et al., 2010). Moreover the
separation in our setup starts from a highly concentrated
initial solution (twofold more than the one reported for
similar products) (Chang et al., 2010). This shows that the
SAN-ion exchange membrane used in this study does not

Table 2 Final cell performance and energy consumption.


E.C. (kWh/m3)

rd
Potentiostatic control
Galvanostatic control

6V
64.69%
0.05 A
55.63%

8V
70.31%
0.1 A
65.00%

6V
40
0.05 A
6.5

8V
61
0.1 A
25.3

Observations: cathodic compartment, t = 1.5 hr, Vcat = 2985 105 m3.

hinder the transport of counter ions despite the affordable


operating costs.

2.3. Membrane characteristics


2.3.1. Optical microscopy
The first inspection by optical microscopy revealed for
SAN-ion exchangers membranes a highly texturized surface,
in contrast with the apparent flat and homogeneous shell at
macroscale (Fig. 4). Micropores were formed during membrane coagulation with quite near distribution of pore sizes
for each membrane. However in the case of SAN-Purolite A100
the tendency of forming larger pores was higher than for the
SAN-Puropack PPC 100 membrane. This aspect suggests a
more polar character of the Purolite A100 in contrast with
Puropack PPC 100, which favors the non-solvent (water)
adsorption and increases the pores formation rate. Using the
ion exchanger in the membrane formation could be an
advantage for the process, which for highly nonpolar polymers is generally assisted by a surfactant (Voicu et al., 2014).
Next to the pore size the distribution as well seems to be
favorable to SAN-Puropack PPC 100.

2.3.2. FT-IR spectroscopy


The chemical structure of the obtained membranes (initial
membranes and after use membranes, at different operated
conditions) was investigated by FT-IR spectroscopy (Fig. 5).
The samples were examined before and after the exposure to
electrodialysis process (galvanostatic and potentiostatic
mode).
FT-IR spectra showed the aromatic CH stretching vibration
at 3028, 3061 cm1 coming from the styrene units, the phenyl
ring stretching vibration C_C at 1603 cm1, for the out-of-plane
hydrogen bending vibration CH, at the 760 cm1 and 701 cm1
(Xia et al., 2010; Cai et al., 2007). The CH bending modes (two
signals) are indicating the monosubstituted ring specific in
polystyrene units in contrast with para substituted rings from
Purolite A100 (Fig. 5a).
Two peaks were observed at 2925 and 2856 cm 1 corresponding to the CH symmetric and asymmetric stretching
vibration from CH2 aliphatic chain from both SAN copolymer
and Purolite polymer chain. The vibration of C`N groups
appeared at 2237 cm1 indicating C`N stretching vibration
belonging to acrylonitrile units from the SAN copolymer. The
CH bending from polystyrene units at 1451 cm1, respectively 1495 cm1 from acrylonitrile units. The bands absorption
from 2362 cm1 and 2320 cm1 belongs very probable to CO2
absorption band (Choi et al., 2003; Zhang et al., 2009; Yu et al.,
2003). CO2 adsorption occurs during sample manipulation in
air, when used in experiments. Carbonation of the polymer
materials is known especially when it comes for a large
surface area (like a membrane) (Kazarian et al., 1996). The
adsorption of CO2 was more pronounced when Purolite
particles were used, since they have an even higher specific
surface area (in comparison with neat SAN membrane).
Supplementarily, we should consider that Purolite particles
posses a higher polarity in comparison with the copolymer
alone.
The ion exchanger presence in the heterogeneous membrane was evidenced with bands similar to SAN copolymer

33

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 2 7 3 7

SAN-Purolite A100

SAN-Puropack PPC100

Fig. 4 Optical microscope images for the surface morphology of SAN-Purolite A100 (a) and SAN-Puropack PPC100 (b). SAN:
styrene-acrylonitrile copolymer.

0.14
0.12

0.12
ATR units

ATR units

0.10
0.08
0.06

0.02

0.16
0.14

900

400

SAN A100 blank


SAN A100 6 V
SAN A100 8 V

0.12
ATR units

0.08
0.06

2900 2400 1900 1400


Wavenumber (cm-1)

900

400

900

400

PPC100 powder
SAN blank
SAN PPC100 blank

0.08
0.06
0.04

0.04

0.02

0.02
0
3900

0.16

3400

0.10

0.10

0.18

3900

0.14

0.12

3400

2900

2400 1900 1400


Wavenumber (cm-1)

900

0
3900

400

0.16

SAN PPC100 blank


SAN PPC100 0.05 A
SAN PPC100 0.1 A

0.14

0.14

ATR units

0.1
0.08
0.06

0.06

0.02

0.02
2900

2400 1900 1400


Wavenumber (cm-1)

SAN PPC100 blank


SAN PPC100 6 V
SAN PPC100 8 V

0.08

0.04

3400

900

400

2900 2400 1900 1400


Wavenumber (cm-1)

0.10

0.04
0
3900

3400

0.12

0.12
ATR units

0.06

0.02
2900 2400 1900 1400
Wavenumber (cm-1)

SAN A100 blank


SAN A100 0.05 A
SAN A100 0.01 A

0.08

0.04

3400

0.10

0.04

0
3900

ATR units

0.14

A100 powder
SAN blank
SAN A100 blank

0
3900

3400

2900 2400 1900 1400


Wavenumber (cm-1)

900

400

Fig. 5 FT-IR spectra of anion exchange (a) and cation exchange (b) membrane before electrodialysis and of its constituents;
FT-IR spectra of SAN-anion (c) and SAN-cation (d) exchange membranes before and after electrodialysis at given current; FT-IR
spectra of SAN-anion (e) and SAN-cation (f) exchange membranes before and after electrodialysis at given voltage. FT-IR:
Fourier transform infrared spectroscopy; SAN: styrene-acrylonitrile copolymer.

34

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 27 3 7

because of the same polystyrene segments (Fig. 5a, c, and e).


Supplementarily, some specific bands appeared like CN
specific absorptions from the tertiary amine segments
(found at 1256 cm 1) in stretching mode. Other specific
bands for the ion exchanger are indicated by the 854, 813,
760 and 701 cm1 from CH (phenyl ring) in out of plane
bending mode. These absorptions are confirming the coexistence of mono and bisubstituted (ortho, metha, para) ring
from the segments obtained with divinyl benzene for
cross-linking and the ion exchanger groups from the active
sites (Fig. 1) (Cai et al., 2007)
Puropack PPC100 spectra (Fig. 5b, d, and f) shows the specific
bands for polystyrene sulfonate cross-linked with divinyl
benzene. The specific bands are showing small deviations
from the ones observed in Purolite A100 since the functional
groups are different (from tertiary amine, sulfonic acid in this
case). The sulfonic group specific peaks can be found at
1038 cm1, S_O in symmetric stretching mode and 1176 cm1,
S_O in asymmetric stretching mode. The para substituted
position appears also in the 832 cm1 (CH from the aromatic
ring, out of plane bending mode) (Xia et al., 2010; Cai et al., 2007).
The FT-IR spectra in Fig. 5c, and e show aromatic CH
stretching vibrations at 2928 cm1 and aliphatic CH stretching
at 2858 cm1 (cross-linked polystyrene). In addition, the peaks
at 2237 cm1 and 1451 cm1 indicate C`N stretching and CH
bending, respectively (Cai et al., 2007). The bands in all spectra
from 2362 cm1 and 2320 cm1 are very probable from carbon
dioxide absorption band. The peak observed at 1740 cm1 can
be attributed to C_O stretching vibration. A distinct band
appears at 1236 cm1 for SAN-Purolite A100 and 1237 cm1 for
SAN-Puropack PPC100 membranes. We believe that this band
could indicate the separated ion complex perturbing the tertiary
amine CN bonds (initial at 1256 cm1 in stretching mode)
(Purolite A100 powder), respectively the S_O asymmetric
stretching mode (initial 1176 cm1) (Puropack PPC100).
The FT-IR profiles of the membranes obtained from SAN
with Purolite A100 or Puropack PPC100 confirmed the
absence of strong chemical or physical interactions between
membrane partners. This aspect was considered essential for
the ion exchanger nature and sequentially for his ability to
participate to the exchange process. Moreover in terms of
accessibility a favorable placement of the ion exchanger
particles toward membrane pores surface is very probable,

judging after FTIR profiles (measured on surface on ATR


module) and by the ion exchanger polarity which shows
affinity for the water phase. This behavior could be involved
in the membrane casting section, in the particular case of the
coagulation step occurring in the water phase.

2.3.3. TGA DTG


The thermogravimetric analysis and differential thermal
analysis (TGA DTG)thermogravimetrical analysis and differential thermal analysis-profiles (Fig. 6) displayed the
membrane thermal stability over a wide range of operation
temperature (this is essential not only in electrodialysis but
also for other membrane applications) (Xia et al., 2010; Voicu
et al., 2014).
The importance of the thermostability should be considered in electrochemical process since the local heat up in the
membrane exceeds the temperatures from the water phase.
Neat SAN copolymer membrane was found stable up to 263C,
meanwhile the SAN-ion exchanger membranes thermal
stability increased to over 290C. The ion-exchanger resin
not only provided functionality from the separation point of
view, but can also increase the final membrane thermal
stability even at such low amounts (5 wt.%). Another advantage of the SAN-ion exchange heterogeneous membranes was
found in the operational process; since the thermal stability of
the operated membranes (Fig. 6) showed an increase in
thermal stability (the degradation starts over 300C). The
cross-linked structure of the porous ion-exchanger could
provide sites of adsorption of the SAN copolymer phase
before degradation and can act as a certain barrier effect for
heat diffusion and sorption of radicals during decomposition.

2.3.4. SEM
The SEM pictures for the SAN-ion exchange membranes (Fig. 7)
confirmed the operational performance since a large availability of the ion exchangers was found. The ion exchange particles
were found in large amounts toward pore surfaces. This aspect
can be explained by the migration of the polar resin particles
toward water phase in the casting process.
Macro and microspores were observed according to SEM
images (Fig. 7), suggesting a possible involvement of the
membrane morphology in the current vs. time deviation from
an ideal curve. The pores offered a large surface area on

2.5

100

SAN A100 blank


SAN PPC 100 8Va

SAN A100
Deriv (% / C)

Weight (%)

SAN PPC 100

80

SAN PPC100 8V
60

SAN blank
SAN PPC 100 blank

SAN blank

SAN A100 8V

40

SAN A100 8V

1.5

0.5

20

0
0

100

200

300

400

Temperature (C)

500

600

700

100

200

300

400

500

600

700

Temperature (C)

Fig. 6 Thermal stability by TGA DTG profiles (a) and by DTG profiles (b) of the membranes before and after operating. TGA:
thermogravimetric analysis; DTG: differential thermal analysis.

35

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 2 7 3 7

which the ion exchanger presence was observed. This aspect


explains also the good ion mobility and extraction values
obtained on such a low amount of ion exchanger in the
membrane. The ion-exchange particles are acting in a
concerted manner with the solvent phase in relation with
the polymer phase during membrane obtaining. The
ion-exchange resin has a good affinity for the SAN matrix
(based on polystyrene segments), but strongly favors water
adsorption (as the DMF) because it needs the hydration on the
strong polar functional groups. Moreover the ion-exchange
particles are insoluble in the water phase despite the great
affinity and the placement at the interface between water
and SAN phase was again favored. Considering also the ion
exchanger particles superficial embedding in the fast coagulating SAN phase the surface displacement should be again a
favorable event. Ion exchange particle polar character is
favoring besides pore formation in a concerted manner the
general morphological profile and particular the active
surface. SEM analyses confirmed the primary FT-IR data
above, which highlighted the clear absorptions of the resin
functional groups on material surface. These favorable
events were the driven vectors for obtaining the above
emphasized separation dynamics involved in the electrodialysis process at such low ion exchanger loads (5 wt.% to SAN
phase) (Rodrigues et al., 2008; Arthanareeswaran and Starov,
2011).

2.3.5. Contact angle measurements


Contact angle (CA) is a convenient way to assess the
hydrophilic/hydrophobic properties and wettability of the
membrane surface. The CA values (Fig. 8) showed that SAN
Purolite heterogeneous membranes are quite hydrophilic. CA
measurements sustain once again the surface placement of
the ion exchanger resin with polar groups.
Ion exchange particles are increasing the water contact angle
in the analyzed membranes (Fig. 8). The resin particles induced a
texturizing effect on the membrane surface as could be seen also
in the morphology details above. The increase in contact angle
was more pronounced after operating the membranes suggesting
a certain decrease of the membrane reliability in time, involving

40

120

180

2500

the wettability if longer operation times or more cycles are


intended. In the same time this drawback could be interesting in
other applications which should involve lower polarity fluids (or
organic phases). The highest obtained value (92 CA) was exactly
on the edge of the hydrophilic and hydrophobic surfaces, which
still offers a certain versatility of the membrane applications. This
kind of versatility, next to the potential antibacterial effect of the
copper ions trapped on the membrane surface could be the basis
of reusing such membranes in biowaste separation (Rusen et al.,
2014).

3. Conclusion
The removal of copper ions from synthetic electroplating
wastewater can be carried out effectively by applying
electrodialysis with heterogeneous membranes, composed
of styrene-acrylonitrile copolymer and small loads (5 wt.%) of
ion-exchange resin particles: The process parameters evidenced a good current and ion mobility through the membrane, in the electrodialysis, since the transport through the
pores allowed the gradual separation. If the duration of the
electrodialysis is 90 min, one can operate either in
potentiostatic control at 8 V, leading to a high extraction
degree (over 70%); or preferably, under constant current
conditions, as for example at a current density of 20 A/m2,
when the extraction degree is a bit lower (52%) but the
energetic consumption performs 10 times better only
6 kWh/m3. The performances obtained at small loads of ion
exchangers are close to those of the homogeneous ion
exchange membranes (Chang et al., 2010). SAN copolymer
was proven as an efficient matrix with high thermal stability
(over 260C). SAN heterogeneous membranes thermal stability showed even an increase after operation (degradation
above 290C). The operability of the membrane under these
conditions was explained from the structural and morphological characteristic occurred as a consequence of ion
exchanger placements on the pores surfaces. The small
amounts of cooper trapped in the membrane ion exchange
sites, the increase in contact angles after operating, next to a

40

800

1800

2500

Fig. 7 SEM images of the SAN-anion exchange membrane (Puropack PPC100) (a) and SAN-cation exchange membrane
(Purolite A100) (b) before experiments. SEM: scanning electron microscopy; SAN: styrene-acrylonitrile copolymer.

36

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 27 3 7

Blank

6V

8V

0.05 A

0.1 A

Contact angle (C)

120
100
80
60
40
20
0
SAN A100
SAN PPC100
Membrane type
Fig. 8 Surface modifications of the SAN-ion exchange
membranes followed by contact angle. SAN:
styrene-acrylonitrile copolymer.

the thermal stability profile could open new directions in


reusing such membranes in biowaste applications (for
example in antibacterial separations since cooper ions
activity in such a direction is known).

Acknowledgment
Prof. Florin Danes is kindly acknowledged for his valuable
support in the manuscript editing. Special thanks to: Mrs.
Mariana Andrei and Mr. Corneliu Andrei from Politehnica
University of Bucharest for their logistic support.

REFERENCES

Arar, ., Yksel, ., Kabay, N., Yksel, M., 2011. Removal of Cu2+ ions
by a micro-flow electrodeionization (EDI) system. Desalination 277
(1-3), 296300.
Arar, ., Yksel, ., Kabay, N., Yksel, M., 2014a. Various applications
of electrodeionization (EDI) method for water treatmenta short
review. Desalination 342, 1622.
Arar, ., Yksel, ., Kabay, N., Yksel, M., 2014b. Demineralization
of geothermal water reverse osmosis (RO) permeate by
electrodeionization (EDI) with mixed bed configuration.
Desalination 342, 2328.
Arthanareeswaran, G., Starov, M.V., 2011. Effect of solvents on
performance of polyethersulfone ultrafiltration membranes:
investigation of metal ion separations. Desalination 267 (1),
5763.
Barakat, M.A., 2011. New trends in removing heavy metals from
industrial wastewater. Arab. J. Chem. 4 (4), 361377.
Cai, Y.B., Hu, Y., Xiao, J.F., Song, L., Fan, W.C., Deng, H.X., et al.,
2007. Morphology, thermal and mechanical properties of poly
(styrene-acrylonitrile) (SAN)/clay nanocomposites from
organic-modified montmorillonite. Polym.-Plast. Technol.
Eng. 46 (5), 541548.
Caprarescu, S., Purcar, V., Vaireanu, D.I., 2012. Separation of copper
ions from synthetically prepared electroplating
wastewater at different operating conditions using
electrodialysis. Sep. Sci. Technol. 47 (16), 22732280.
Chang, J.H., Elli, S.A.V., Tung, C.H., Huang, W., 2010. Copper cation
transport and scaling of ionic exchange membranes using
electrodialysis under electroconvection conditions. J. Membr. Sci.
361 (1-2), 5662.

Chen, S.S., Li, C.W., Hsu, H.D., Lee, P.C., Chang, Y.M., Yang, C.H.,
2009. Concentration and purification of chromate from
electroplating wastewater by two-stage electrodialysis
processes. J. Hazard. Mater. 161 (2-3), 10751080.
Choi, Y.S., Xu, M., Chung, I.J., 2003. Synthesis of exfoliated
poly(styrene-co-acrylonitrile) copolymer/silicate nanocomposite
by emulsion polymerization; monomer composition effect on
morphology. Polymer 44 (22), 69896994.
Cifuentes, L., Garca, I., Arriagada, P., Casas, J.M., 2009. The use
of electrodialysis for metal separation and water recovery
from CuSO4H2SO4Fe solutions. Sep. Purif. Technol. 68 (1),
105108.
Dalla Costa, R.F., Klein, W.C., Bernardes, A.M., Ferreira, J.Z., 2002.
Evaluation of the electrodialysis process for the treatment of
metal finishing wastewater. J. Braz. Chem. Soc. 13 (4), 540547.
Dzyazko, Y.S., 2006. Purification of a diluted solution containing
nickel using electrodeionization. Desalination 198 (1-3), 4755.
Dzyazko, Y.S., Rozhdestvenskaya, L.M., Vasilyuk, S.L., Belyakov, V.N.,
Kabay, N., Yuksel, M., et al., 2008. Electro-deionization of Cr (VI)containing solution. Part I: chromium transport through
granulated inorganic ion-exchanger. Chem. Eng. Commun. 196
(1-2), 321.
Feng, X., Wu, Z.C., Chen, X.F., 2007. Removal of metal ions from
electroplating effluent by EDI process and recycle of purified
water. Sep. Purif. Technol. 57 (2), 257263.
Fritzsche, A.K., 1986. Effect of ionizing radiation on styrene/
acrylonitrile copolymer hollow fiber membranes. J. Appl.
Polym. Sci. 32 (2), 35413550.
Fu, F.L., Wang, Q., 2011. Removal of heavy metal ions from
wastewaters: a review. J. Environ. Manag. 92 (3), 407418.
Gven, A., Karabacakolu, B., 2005. Use of electrodialysis to remove
silver ions from model solutions and wastewater. Desalination 172
(1), 717.
Hashim, M.A., Mukhopadhyay, S., Sahu, J.N., Sengupta, B., 2011.
Remediation technologies for heavy metal contaminated
groundwater. J. Environ. Manag. 92 (10), 23552388.
Joshi, M., Mukherjee, A.K., Thakur, B.D., 2001. Development of a
new styrene copolymer membrane for recycling of polyester
fibre dyeing effluent. J. Membr. Sci. 189 (1), 2340.
Kabay, N., Demircioglu, M., Ersz, E., Kurucaovali, I., 2002. Removal
of calcium and magnesium hardness by electrodialysis.
Desalination 149 (1-3), 343349.
Kazarian, S.G., Vincent, M.F., Bright, F.V., Liotta, C.L., Eckert, C.A., 1996.
Specific intermolecular interaction of carbon dioxide with polymers.
J. Am. Chem. Soc. 118 (7), 17291736.
Mahmoud, A., Muhr, L., Grvillot, G., Lapicque, F., 2007. Experimental
tests and modelling of an electrodeionization cell for the treatment
of dilute copper solutions. Can. J. Chem. Eng. 85 (2), 171179.
Melnyk, L., Goncharuk, V., 2009. Electrodialysis of solutions
containing Mn (II) ions. Desalination 241 (1-3), 4956.
Murthy, Z.V.P., Mungray, A.A., Singh, J., 2012. Preparation and
characterization of acrylonitrile butadiene styrene and
trifluoroacetylethyl cellulose blend nanofiltration membrane
and performance in the separation of mercury. Proceedings of
the World Congress on Engineering and Computer Science
Vol. II. WCECS, San Francisco, USA.
Nasef, M.M., Gven, O., 2012. Radiation-grafted copolymers for
separation and purification purposes: status, challenges and
future directions. Prog. Polym. Sci. 37 (12), 15971656.
tveren, .B., Koparal, S., zel, E., 1997. Electrodialysis for the
removal of copper ions from wastewater. J. Environ. Sci. Health
Part A: Environ. Sci. Eng. Toxic. 32 (3), 749761.
Parulekar, J.S., 1998. Optimal current and voltage trajectories for
minimum energy consumption in batch electrodialysis. J. Membr.
Sci. 148 (1), 91103.
Peng, C., Liu, Y., Bi, J., Xu, H., Ahmed, A.S., 2011. Recovery of copper and
water from copper-electroplating wastewater by the combination
process of electrolysis and electrodialysis. J. Hazard. Mater. 189 (3),
814820.

J O U RN A L OF E N V I RO N ME N TA L SC IE N CE S 3 5 (2 0 1 5) 2 7 3 7

Qdais, H.A., Moussa, H., 2004. Removal of heavy metals from


wastewater by membrane processes: a comparative study.
Desalination 164 (2), 105110.
Radha, K.S., Shobana, K.H., Mohan, D., Narayana, L.Y., 2009.
Synthesis and characterization of styrene-acrylonitrile
copolymer blend ultrafiltration membranes. Desalin. Water
Treat. 12 (1-3), 114126.
Radha, K.S., Shobana, K.H., Tarun, M., Mohan, D., 2014. Studies on
sulfonated styrene acrylonitrile and cellulose acetate blend
ultrafiltration membranes. Desalin. Water Treat. 52 (1-3), 459469.
Rodrigues, M.A.S., Amado, F.D.R., Bischoff, M.R., Ferreira, C.A.,
Bernardes, A.M., Ferreira, J.Z., 2008. Transport of zinc complexes
through an anion exchange membrane. Desalination 227 (1-3),
241252.
Rusen, E., Mocanu, A., Nistor, L.C., Dinescu, A., Clinescu, I.,
Mustea, G., et al., 2014. Design of antimicrobial membrane
based on polymer colloids/multiwall carbon nanotubes hybrid
material with silver nanoparticles. ACS Appl. Mater. Interfaces
6 (20), 1738417393.
Sadrzadeh, M., Mohammadi, T., Ivakpour, J., Kasiri, N., 2008.
Separation of lead ions from wastewater using electrodialysis:
comparing mathematical and neural network modeling.
Chem. Eng. J. 144 (3), 431441.
Silva, A.L.A., Takase, I., Pereira, R.P., Rocco, A.M., 2008.
Poly(styrene-co-acrylonitrile) based proton conductive
membranes. Eur. Polym. J. 44 (5), 14621474.
Spoor, P.B., Grabovska, L., Koene, L., Janssen, L.J.J., ter Veen, W.R.,
2002. Pilot scale deionisation of a galvanic nickel solution
using a hybrid ion-exchange/electrodialysis system. Chem.
Eng. J. 89 (1-3), 193202.

37

Strathmann, H., 2010. Electrodialysis, a mature technology with a


multitude of new applications. Desalination 264 (3), 268288.
Tanaka, Y., 2002. Current density distribution, limiting current
density and saturation current density in an ion-exchange
membrane electrodialyzer. J. Membr. Sci. 210 (1), 6575.
Tzanetakis, N., Taama, W.M., Scott, K., Jachuck, R.J.J., Slade, R.S.,
Varcoe, J., 2003. Comparative performance of ion exchange
membranes for electrodialysis of nickel and cobalt. Sep. Purif.
Technol. 30 (2), 113127.
Vinodh, R., Padmavathi, R., Sangeetha, D., 2011. Separation of heavy
metals from water samples using anion exchange polymers by
adsorption process. Desalination 267 (2-3), 267276.
Voicu, S.I., Dobrica, A., Sava, S., Ivan, A., Naftanaila, L., 2014.
Cationic surfactants-controlled geometry and dimensions of
polymeric membrane pores. J. Optoelectron. Adv. Mater. 14
(11-12), 923928.
Xia, Q., Zhao, X.J., Chen, S.J., Ma, W.Z., Zhang, J., Wang, X.L., 2010. Effect
of solution-blended poly(styrene-co-acrylonitrile) copolymer on
crystallization of poly(vinylidene fluoride). Express Polym. Lett. 4
(5), 284291.
Yu, J.R., Yi, B.L., Xing, D.M., Liu, F.Q., Shao, Z.G., Fu, Y.Z., et al.,
2003. Degradation mechanism of polystyrene sulfonic acid
membrane and application of its composite membranes in
fuel cells. Phys. Chem. Chem. Phys. 5 (3), 611615.
Zhang, L., Li, A.M., Wang, J.N., Lu, Y.F., Zhou, Y.D., 2009. A novel
aminated polymeric adsorbent for removing refractory dissolved
organic matter from landfill leachate treatment plant. J. Environ.
Sci. 21 (8), 10891095.

Você também pode gostar