You are on page 1of 10

Resources, Conservation and Recycling 26 (1999) 115124

Electrochemical remediation of copper (II) from


an industrial effluent
Part I: monopolar plate electrodes
Carlo Solisio a, Marco Panizza b,*, Piergiorgio Paganelli c,
Giacomo Cerisola b
Institute of Chemical Engineering, Engineering Faculty, Genoa Uni6ersity, 6ia Opera Pia 15,
16145 Genoa, Italy
b
Chemical Institute, Engineering Faculty, Genoa Uni6ersity, J.F.Kennedy 1 sq., 16129 Genoa, Italy
c
Organic Chemicals s.r.l., Piazza dell Vittoria 10, 17056 Cengio, Sa6ona, Italy
a

Accepted 24 November 1998

Abstract
Many studies indicate that flow-through reactors with three-dimensional electrodes can be
successfully used for metal removal from dilute streams, but they present high capital costs.
So the simplest titanium and stainless steel AISI 904L plate were chosen as cathode materials
in tank reactors. A three cell plant was used for cathodic deposition of copper from an
industrial effluent containing copper (II). The influence of the flow rate and the initial copper
concentration on the removal efficiency and current yield was studied. 1999 Elsevier
Science B.V. All rights reserved.
Keywords: Effluent copper removal; Plate electrode; Wastewater

* Corresponding author. Tel.: +39-010-3536033; fax: + 39-010-3536028.


0921-3449/99/$ - see front matter 1999 Elsevier Science B.V. All rights reserved.
PII: S 0 9 2 1 - 3 4 4 9 ( 9 8 ) 0 0 0 7 8 - 0

116

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

1. Introduction
Metal ion solutions are generally treated by chemicalphysical processes such as
chemical precipitation, cementation and ion exchange. The most utilised treatment
is to precipitate heavy metals as hydroxide sludge with milk of lime or caustic or
sodium carbonate.
This process is quite simple to operate and its plant and exercise costs are cheap,
but a large quantity of sludge are formed and they have no real use and involve
other environmental problems.
For these reasons the electrochemical recovery of metal from dilute solution
offers several promising approaches. Among the most attractive features of this
technology are high efficiency, amenability to automation, lack of sludge to dispose
of, possibility of recovering pure metal.
Different types of electrochemical cell design [114] have been described in
literature and applied industrially for metal recovery using porous electrode cells.
Pletcher et al. [7 9] studied mass transport and removal of copper from dilute
solution in acidic sulphate media using a reticulated vitreous carbon cathode,
Bisang [10] investigated the effect of side reaction such as the reduction of oxygen,
on copper deposition on packed-bed electrodes.
Scott [11] describes an experimental study of a moving bed of solid metal
particles which can be used for the electrolytic recovery of copper and other metals.
The flow-through porous electrodes for copper depletion were investigated by
Bennion and Newman [12] both theoretically and experimentally; Trainham and
Newman [13] also estimated the minimum metal concentration attainable with this
device.
Campbell et al. [14] analysed how the use of inert fluidised bed cell and a packed
graphite particle cathode improved mass transfer and the removal of copper.
Even if these studies show that flow reactors with three-dimensional electrodes
can be efficiently used for heavy metal removal from dilute streams, tank cells with
titanium and stainless steel plane electrodes were used in this work because they
have lower capital costs.
In particular, a pilot plant constituted of three cells with vertical monopolar
electrodes was used in order to deplete copper ions from an industrial wastewater.
The wastewater comes from a plant for copper phthalocyanine production in the
chemical factory of ACNA (Cengio, Italy). The efficiency of removal and the
current yield were evaluated as a function of the flow rate and copper
concentration.
2. Experimental
It should be pointed out that the main purpose of the pilot plant was to reduce
the copper concentration to a value suitable for a biological softening plant present
in the factory, in order to maintain the normal activity of bacteria. Because the
maximum copper concentration suitable for a biological plant is about a few mg/l
[15] it was considered to have 5 mg/l as a limit value for decopperizing plant.

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

117

The wastewater with the copper to be removed comes from the plant of copper
phthalocyanine, which is produced by a discontinuous process from phthalic
anhydride, urea and cupric chloride. Even if a catalyst was used to favour the main
reaction, the yield in respect of phthalic anhydride was about 92%, and for this
value the cupric chloride is proportioned. A schematic diagram of the process is
shown in Fig. 1.
The wastewater which was sent to the removal plant was generated during the
filtration which was carried out by a press filter in order to retain the phthalocyanine crystals on the cloth.
The flow rate may change from 2 to 6 m3/h because the production of phthalocyanine was not constant. The copper concentration also presented variation during
time (Fig. 2) especially due to:
different reaction yield caused by non-standard quality of reagents or by a lower
activity of the catalyst which induced less consumption of Cu2Cl2;
variation of quantity of cupric chloride which is loaded as a whole bag of 35 kg;
different effect of dilution of rinse water.
Nevertheless, copper concentration showed an average value of about 340 mg/l
and standard deviation was such to configure 92% of observations within the
acceptability range. The average feature of the wastewater is illustrated in Table 1.
A simple scheme of the copper removal plant is illustrated in Fig. 3. At the
beginning the wastewater was heated to 55C to avoid the precipitation of salt such
as chlorides or sulphates, present in high concentration (Table 1). Then it went into
the first electrochemical cell (A) and afterwards it was equidistributed into the B
cells which worked in parallel. The stream was controlled by valves and flow
meters.

Fig. 1. Schematic diagram of the copper phthalocyanine production process.

118

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

Fig. 2. Variation of copper concentration with time in the effluent.

Parallelepiped cells made in PVC with iron reinforcing bars were used. Every cell
had a volume of 1.10 m3 and it was constituted of 20 titanium anodes and 21
cathodes (Fig. 4).
The anodes were titanium net in order to allow bubbles to escape rapidly, and so
to reduce overpotential resistance. The cathodes were titanium plate, for the A cell
and stainless steel AISI 904L plate for the B cells. Each electrodes had an area of
0.94 m2, a thickness of 0.15 cm and a distance from each other of 2 cm. Electrical
contact was made by a titanium bar.
Vibrant solenoids fed by alternate current were used in order to make the
electrodes vibrate. In this way the metal disjoins from the electrodes and sinks to
the floor of the cells where a temporised valve discharges the deposit into filtering
bags, allowing the recovery of the copper. Electrode vibration also improves the
mass transport in the electrolyte.
It was chosen to work in a continuous and a potentiostatic manner applying a
differential potential of 2.5 V at the electrodes because of the high oxygen evolution
overpotential on the titanium anodes [16].
Table 1
Average effluent features
Cu2+
Cl
SO2
4
COD
pH
Kinematic viscosity
Density
Flow rate

340 (mg/l)
5000 (mg/l)
30 000 (mg/l)
10 000 (mg/l)
B1
6107 (m2/s)
10201050 (kg/m3)
26 (m3/h)

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

119

Fig. 3. Representation of the electrochemical plant for copper depletion showing: (1) effluent reservoir,
from the filtration unit; (2) heat exchanger; (3) flow meters; (4) valves; (5) A cell; (6) B cells; (7) copper
removal; (8) biological plant.

During the experiments metal concentration was measured by atomic absorption


spectroscopy (Perkin-Elmer 3030) before and after each cell. Current flowing in the
cells was also monitored.
As the quantity of wastewater to be treated was not constant but it was
dependent on the production of phthalocyanine, a series of experiments were
performed in order to find the dependence on flow rate condition both for the
copper removal and for current yield.
Then, the influence of metal concentration in the electrolyte was evaluated.

3. Results and discussion


An increase of the flow rate may induce different effects on the removal
efficiency.
If, on the one hand, it improves the mass transport in the cell and so increases
the copper removal, on the other hand it decreases the residence time in the cells,
opposing the deposition of copper.
In order to evaluate which had the main effect on our cells, a first series of
experiments was done at different flow rates, from 2 to 6 m3/h. Copper concentration was 340 mg/l and the cell voltage was 2.5 V.
The results are shown in Fig. 5 for both the cells A, B and for the total plant. It
can be observed that the efficiency decreases when the flow rate increases and so the

120

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

Fig. 4. Simplified view of the electrochemical tank reactor.

effect associated with the decrease in residence time was greater than one associated
with reduced mass-transport resistance. Similar results were found by Matlosz and
Newman [17] during the removal of mercury with carbon electrodes: their effluent
contained more mercury at the higher flow rate.

Fig. 5. Effect of flow rate on the removal efficiency. Inlet copper concentration 340 mg/l, cell voltage 2.5
V. +, data for A cell; , data for B cells; , data for the total plant.

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

121

Fig. 6. Variation of Km with flow rate. + , data for cell A. , data for cells B.

The mass-transport coefficient Km can be evaluated by removal data instead of


by limited current measurements [7]. The electrolytic cells used in this work may be
modelled as CSTR under diffusion control, and so we may write [1]:
Xa =1

C out
=1
C in

1
K A
1+ m
Q%

(1)

where Xa, is the removal efficiency; C, the inlet and outlet copper concentration
(g/m3); Q%, the feed flow rate (m3/s); A, the electrode area (m2); Km, mass-transport
coefficient (m/s)
The trend of Km against flow rate is evaluated using Eq. (1) with the data in Fig.
5 and it is illustrated in Fig. 6. Such a trend was also obtained by Pletcher et al. [7]
with vitreous carbon electrodes.
In order to see if it is convenient to work with a lower flow rate, the influence of
this parameter on the current yield was evaluated. It is defined as:
h=

(C in C out)Q%Fz
100
MI

(2)

where h, is the current yield; C, the inlet and outlet copper concentration (g/m3);
Q%, the flow rate (m3/s); F, the Faraday constant (96485 C); z, the number of
electrons taking part in the overall reaction; M, the molecular weight; I, the current
intensity (A).
The results obtained (Fig. 7) show that the current yield increases with the flow
rate. So at a lower flow rate a better copper removal is obtained, however, the
worsening in the current yield makes the process economically unviable. In fact at
2 m3/h there is a yield of just 13%.

122

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

Fig. 7. Effect of flow rate on the current yield for the total plant. Inlet copper concentration 340 mg/l,
cell voltage 2.5 V.

Other experiments, executed with a constant flow rate of 4 m3/h, analysed the
influence of copper concentration on the removal efficiency and current yield. These
results are shown in Figs. 8 and 9. Both parameters increase with initial concentration, as was found by other authors [14,18]. This could be explained by the fact that
at low copper concentration the removal reaction comes in competition with
hydrogen evolution reaction.

Fig. 8. Removal efficiency of the total plant against copper concentration. Flow rate 4 m3/h, cell voltage
2.5 V.

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

123

Fig. 9. Current yield of the total plant against copper concentration. Flow rate 4 m3/h, cell voltage 2.5
V.

Nevertheless, at the higher metal concentration removal efficiency was not


sufficiently satisfactory, so it was preferred to abandon this device.

4. Conclusion
An electrochemical three cell plant, working continuously, was used to remove
copper from an industrial wastewater. Titanium and stainless steel AISI 904L plate
cathodes rather than porous electrodes were chosen because of their lower capital
costs.
The influence of the flow rate on removal efficiency and current yield was
studied. It was pointed out that this parameter principally influenced residential
time rather than mass-transport. So, when it was low metal depletion increased, but
the current yield became very small and the process economically unviable.
The influence of the copper concentration was also studied and it was shown that
a higher inlet concentration favours metal deposition and increases current yield.
Nevertheless, an outlet copper concentration below 5 mg/l was not reached in any
condition. So it was preferred to abandon the plate electrodes in favour of
three-dimensional electrodes, because they have a high specific surface and they
supply better performances even if they are more expensive. The results of the
porous electrode experiments will be published in the next paper.

124

C. Solisio et al. / Resources, Conser6ation and Recycling 26 (1999) 115124

Acknowledgements
We would like to especially thank the ACNA personnel who made significant
contributions to the project. The financial support of MURST project Elettrocatalisi ed elettrosintesi is acknowledged.

References
[1] Pletcher D, Walsh FC. Water purification, effluent treatment and recycling of industrial process
streams. In: Industrial Electrochemistry, 2nd Edition. London: Chapman and Hall, 1990;331 384.
[2] Pletcher D, Walsh FC, Whyte I. I Chem E Symp Ser 1990;116:195 210.
[3] Ismail MI (Ed). Electrochemical Reactors Their Science and Technology, part A, Amsterdam:
Elsevier, 1990.
[4] El Ghaoui EA, Jansson REW, Moreland C. J Appl Electrochem 1982;12:59.
[5] Ehdaie S, Fleischmann M, Jansson REW. J Appl Electrochem 1982;12:69.
[6] El Ghaoui EA, Jansson REW, Moreland C. J Appl Electrochem 1982;12:75.
[7] Pletcher D, Whyte I, Walsh FC, Millington JP. J Appl Electrochem 1991;21:659.
[8] Pletcher D, Whyte I, Walsh FC, Millington JP. J Appl Electrochem 1991;21:667.
[9] Pletcher D, Whyte I, Walsh FC, Millington JP. J Appl Electrochem 1993;21:82.
[10] Bisang JM. J Appl Electrochem 1966;26:135.
[11] Scott K. J Appl Electrochem 1981;11:339.
[12] Bennion DN, Newman J. J Appl Electrochem 1972;2:113.
[13] Trainham JA, Newman J. J Electrochem Soc 1977;124:1528.
[14] Campbell DA, Dalrympe IM, Sunderland JG, Tilston D. Resour Conserv Recycl 1994;10:25.
[15] Avezzu F, Bissolotti G, Collivignarelli C, Volpi Ghirardini A. Waste Manage and Res 1995;13:103.
[16] Gileadi E. Multi step electrode reactions. In: Electrode Kinetics for Chemists, Chemical Engineers,
and Material Scientists. New Yoek: VCH, 1993, pp 127 154.
[17] Matlosz M, Newman J. J Electrochem Soc 1986;133:1850.
[18] Bartolozzi M, Marconi PF, Braccini G, Magnani G. Resour Conserv Recycl 1991;6:79.

You might also like