You are on page 1of 10

Separation and Purification Technology 96 (2012) 296–305

Contents lists available at SciVerse ScienceDirect

Separation and Purification Technology


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

Bench scale electrocoagulation studies of bio oil-in-water and synthetic


oil-in-water emulsions
M. Karhu a,⇑, V. Kuokkanen b, T. Kuokkanen b, J. Rämö c
a
Chemical Process Engineering Laboratory, Department of Process and Environmental Engineering, University of Oulu, P.O. Box 4300, 90014 Oulu, Finland
b
Department of Chemistry, University of Oulu, P.O. Box 3000, 90014 Oulu, Finland
c
Thule Institute, University of Oulu, P.O. Box 7300, 90014 Oulu, Finland

a r t i c l e i n f o a b s t r a c t

Article history: Electrocoagulation (EC) test runs of oil-in-water (O/W) emulsions (0.6% and 2%) prepared from bio oils
Received 30 November 2011 and synthetic oils with stainless steel (SS) or aluminum (Al) anode were performed in a batch mode with
Received in revised form 18 May 2012 a novel bench scale EC apparatus. The efficiency of EC for breaking O/W emulsions was evaluated versa-
Accepted 3 June 2012
tilely by measuring COD, TOC, total surface charge (TSC) and turbidity. Particle sizes with laser diffraction
Available online 12 June 2012
and BOD were also measured for a number of test runs. The turbidity of the studied emulsions decreased
the most, between 75% and 100% whereas COD, TSC and TOC reductions were in the range of 25–95%, 40–
Keywords:
98% and 20–75%, respectively. The most successful reductions were measured for tall oil and synthetic oil
Aluminum anode
Electrocoagulation
emulsions. BOD values were also found to decrease during test runs. Laser diffraction measurements
Oil-in-water emulsion proved that the emulsification of the oils, as well as the breaking of the emulsions, were both successful.
Oily wastewater There was no significant difference between treatment results when using a SS or Al anode. However the
Stainless steel anode sludge produced using a SS anode was substantially thicker in structure and thus easier to separate. To
summarize, the EC of O/W emulsions studied in this work was found to be cost-effective and feasible
even without specific optimization of the process parameters.
Ó 2012 Elsevier B.V. All rights reserved.

1. Introduction related to oily wastewaters. Naval and commercial vessels generate


oily wastewaters mostly in the form of bilge water and ballast
Environmentally problematic oily wastewaters are generated in water. Produced water is the largest waste stream generated by
large volumes every day. Oil and grease (O&G) for example, are oil and gas industries containing different kinds of organic and inor-
common pollutants found in effluents from a broad range of indus- ganic components [4]. Ferro and Smith [5] wrote in 2007 that global
tries such as petroleum refineries, petrochemical, metal manufac- water production (produced water) associated with oil and gas
ture, machining and finishing, food processors, textile and leather fields was estimated at around 250 million barrels per day com-
[1,2]. A report [3] published in March 2011 by the oil companies’ pared with around 80 million barrels per day of oil. The volume
European association for environment, health and safety in refining of produced water onshore will increase more rapidly until 2011
and distribution (CONCAWE) summarize data gathered by CONC- and after that the increase of volume will become much slower.
AWE in surveys of effluent water quantity, oil content and treat- The volume of produced water offshore is forecasted to increase un-
ment process for refinery locations situated in the EU-27 til 2015. Because most wastewater produced offshore is discharged
countries and those in Norway and Switzerland for the years into the sea, effective cleaning and separation is critical.
2005 and 2008 (data from previous surveys included for compari- Oily wastewaters can be divided into three categories: free-
son). The data of surveys indicated that oil discharged with refined floating oil, unstable oil/water emulsions, and highly stable oil/
aqueous effluent relative to the refining capacity reduced continu- water emulsions. Free floating oil can be readily removed by
ously from 1969 (127 g t1 capacity) to 2008 (1.18 g t1 capacity). mechanical separation whilst unstable oil-in-water (O/W) emul-
In 2008 the quantity of total aqueous effluents from all the emission sions must be mechanically or chemically broken and separated
points from 125 European oil refineries was 1112  106 t year1 gravitationally. Stable emulsions however need more sophisticated
whilst the total oil discharged was 993 t year1. Furthermore bilge treatment to meet today’s effluent standards [6]. The effectiveness
water and produced water are one of the major global problems of conventional treatment methods is limited, therefore advanced
and cost-effective treatment methods must be developed to meet
ever-stricter environmental regulations and to treat heterogeneous
⇑ Corresponding author. Tel.: +358 8 553 2357; fax: +358 8 553 2304. oily wastewaters containing various types of oils, surfactants, met-
E-mail address: mirjam.karhu@oulu.fi (M. Karhu). als, etc.

1383-5866/$ - see front matter Ó 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.seppur.2012.06.003
M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305 297

Chemical methods are the most widely used techniques em-


ployed in the treatment of oily wastewaters. Ferric and aluminum
salts are still the most popular agents used for demulsification
although also other coagulants are under development. The chem-
ical treatment process usually consists of rapid mixing of the coag-
ulant chemicals with the wastewater followed by flocculation and
flotation or settling [7]. Unfortunately, conventional coagulation
involves a number of drawbacks such as the high amount of re-
quired coagulant, corrosion problems with decreasing pH and
problems with produced sludge.
The principle of electrocoagulation (EC) has been common
knowledge for over 100 years with electrochemical methods first
being used for water and wastewater treatment in 1887 [8]. In
1906 Dieterich patented the first electric water purifier which used
aluminum electrodes [9] whilst J.T. Harries received a patent in
1909 for wastewater treatment by electrolysis with sacrificial alu-
minum and iron anodes [8]. The majority of research took place
when EC was first patented, however interest in the technique in
later years faded due to high electricity prices and underdeveloped
EC technology. Since then EC technology has improved, becoming
more economical, and is one of the most promising alternative
for various types of wastewaters such as: dairy [10], dye and textile
[11,12], heavy metals containing wastewaters [13–15] and waste-
waters from the pulp and paper mill industry [16,17].
EC has several advantages over conventional coagulation, one of
them being that EC can efficiently demulsificate stable oil emul-
sions without substantial chemical addition. Therefore EC has been
studied for real wastewaters containing O&G and also for O/W
emulsions such as: oily bilge water [18], synthetic O/W emulsions
[19,20], wastewaters from petroleum refinery [21], biodiesel pro-
duction [22], vegetable oil industries [23,24], restaurants [25,26],
and agro-industry [27,28].
The main purpose of this work was to study the effectiveness of
a novel bench scale EC apparatus for the treatment of both bio and
synthetic O/W emulsions when using stainless steel (SS) or alumi-
num (Al) as the anode material. This study has been performed
with a novel large scale (bench scale) apparatus with continuous
recycling of sample during test runs and with electrode configura-
tion consisting of multiple vertical electrodes and one electrode Fig. 1. (A) Electrocoagulation apparatus and (B) PI scheme of the system used.
(also used as a basin) (Fig. 1). To the best of our knowledge, there
are no publications using this type of electrode configuration as
differences between the cells. The EC apparatus used by Fouad
well as there are no previous comparative studies on electrocoag-
et al. possessed only two vertical anodes (both the anode and the
ulation of both bio oils and synthetic oils. Breaking of the studied
cathode made of aluminum) compared to six vertical electrode in
O/W emulsions with EC was verified by several analyses such as
this study (the anode and the cathode made of different materials)
chemical oxygen demand (COD), total organic carbon (TOC), total
as well as in this study plane electrode was used as basin not only
surface charge (TSC), turbidity, biological oxygen demand (BOD)
as an electrode. Fouad et al. discovered that a cell with horizontal
and particle size analysis with laser diffraction.
anode electrodes was more efficient in separating oil from oily
Yüksel et al. [29] studied the peroxi-electrocoagulation method
wastewater by electrocoagulation than a cell with vertical anodes.
for the removal of sodium dodecyl sulfate (SDS, used for emulsifi-
This was accounted for by a higher mixing efficiency and higher
cation of oils in our study) in synthetic wastewater. SDS in an
floating ability as a result of the uniform distribution of the H2 bub-
aqueous phase was removed effectively by the peroxi-electrocoag-
bles evolving from all over the cell cross-section of the cathode.
ulation. EC studies have also usually been performed using a labo-
ratory scale EC system with a cell volume ranging from 250 ml to
1500 ml and with electrodes of similar shape placed directly oppo- 2. Theory
site each other and submerged fully in the sample water. To help
evaluate the feasibility of the industrial application of the EC pro- In electrocoagulation, the generation of coagulants is performed
by electrically dissolving metal ions from anode electrodes.
cess, research using larger scale equipment should be performed
already in the beginning of the research to identify practical prob- Austenitic steel with molybdenum used in our studies consisted
mainly of iron (67 w-%). The electrochemical reactions with alumi-
lems related to EC. Thus the bench scale EC apparatus used in this
study represents actual industrial applications of EC better than num or iron as the anode material are presented as the following
equations:
the smaller systems.
Fouad et al. [20] studied the oil separation performance of an At the anode :
electrocoagulation cell with horizontally or vertically oriented an- 3þ
AlðsÞ
Al ðaqÞ þ 3e E0 ¼ þ1:66 V ð1Þ
ode electrodes with a horizontal aluminum plate rested on the cell
bottom. This cell with vertical electrodes was quite similar to the
cell that was used in our study. However there were some notable FeðsÞ
Fe2þ ðaqÞ þ 2e E0 ¼ þ0:44 V ð2Þ
298 M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305

2H2 OðlÞ
O2 ðgÞ þ 4Hþ ðaqÞ þ 4e E0 ¼ 1:23 V ð3Þ Zn, Cr, Mo, Al, Fe, Ni, As and Pb) contents of pure RSO, SFO and
2+ 3+
TO were below 10 mg kg1. The Ca, Zn and Mo contents for MCO
Oxidation of Fe to Fe by atmospheric oxygen or anode oxida-
and MO were 1530/1480 mg kg1, 730/1160 mg kg1 and 70/
tion can be considered [12]:
<10 mg kg1, respectively. The contents of other metals for MCO
Fe2þ ðaqÞ
Fe3þ ðaqÞ þ e E0 ¼ 0:77 V ð4Þ and MO were all below 10 mg kg1. Carbon and hydrogen contents
of RSO, SFO and TO were between 77.7–77.9% and 11.9–12.5%,
2Fe2þ ðaqÞ þ 1=2O2 ðgÞ þ H2 OðlÞ
2Fe3þ ðaqÞ þ 2OH respectively. For MCO and MO carbon and hydrogen contents were
higher than those for bio oils, between 84.8–84.9% and 14.4–14.9%,
E0 ¼ 0:37 V ð5Þ respectively.
Since real oily wastewaters usually contain some surfactants
At the cathode : but also for emulsification and homogenization, anionic surfactant
2H2 O þ 2e
H2 ðgÞ þ 2OH E0 ¼ 0:83 V ð6Þ sodium dodecyl sulfate (SDS, CH3(CH2)11OSO3Na) by Sigma-Al-
drichÒ was added to the O/W solutions. Specific amounts of the
The generated Al3+(aq) or Fe3+(aq) ions will immediately under- surfactant for each type of emulsion was determined in prelimin-
go further spontaneous reactions to produce corresponding ary studies and selected according to COD and TSC measurements
hydroxides and/or polyhydroxides [30]. Sarpola [31] studied the as well as visual inspection. Sodium chloride (NaCl, L.T. Baker) was
hydrolysis of aluminum, the polymerization of the hydrolysis added to the oil-in water emulsion to increase conductivity to a va-
products, and how these can be monitored by mass spectrometric lue of 4.0–4.5 mS cm1 as determined for real oily wastewater. The
methods. pHs of simulated O/W emulsions were not adjusted.
The amount of generated metal ions (mmetal, g) can be calculated The O/W emulsions were prepared by adding 7.5 L of tap water,
from Faraday’s law (Eq. (7)). The current density (i, A m2) (Eq. (8)) oil, surfactant and sodium chloride into a plastic bucket (Table 1).
can be calculated as the current in the EC cell (I, A) per effective an- The solutions were then mixed with a whisk-equipped Heidolph
ode area below the water surface (Aeff, m2). RZR 1 overhead stirrer for 20 min at a rate 900 rpm (about
ItM 280 rpm in the test runs 1–4). The prepared O/W emulsions were
mmetal ¼ ð7Þ colored and turbid. The stability of the O/W emulsions was esti-
nF
mated with TCS measurements. After the 2nd test run, the amount
I of oil used in O/W emulsions was increased to correspond better
i¼ ð8Þ with real oily wastewaters. The number of test runs conducted
Aeff
for RSO emulsion was low because the emulsification of oil was
where t is the treatment time of the EC process (s), M the molar poor compared to SFO. The volume of the prepared O/W emulsions
mass of the metal concerned (g mol1), n the number of electrons was greater when there was a reason to treat again the effluent
in the oxidation reaction and F Faraday’s constant (96,500 C mol1). from the previous test run with EC, or a significant part of the
It has been experimentally observed that the generation of metal emulsion was required for other analyses. The repeated EC treat-
ions (especially in the case of aluminum) has been higher than ment of oil emulsion was performed to study reductions of water
the value calculated by Faraday’s law. This phenomenon has been quality parameters after removing sludge from the surface of the
referred as superfaradaic efficiencies. Cãnizares et al. [32] studied settled sample after the first test run. Increased removal of oil by
the chemical dissolution of aluminum sheets in four different pHs. the repeated treatment was hypothesized to occur because the
The alkaline pHs increased the dissolution rate of aluminum by or- sludge removal destabilizes the chemical balance of the whole
ders of magnitude. studied solution. The volume of emulsion in the repeated treat-
The mode of action of the aluminum and iron is generally ex- ment was increased with tap water since a minimum amount of
plained in terms of two distinct mechanisms: charge neutralization emulsion (about 17 L) was required for the pump of the apparatus
of negatively charged colloids by cations and the incorporation of to continue functioning.
impurities in a produced amorphous hydroxide precipitate as
sweep flocculation. The relative importance of charge neutraliza- 3.2. Electrocoagulation set-up
tion and sweep flocculation depends on factors such as pH and
coagulant dosage [33]. Other important factors influencing the effi- Bench scale EC studies were carried out with horizontal flow
ciency of the EC process are: electrode materials, current density electrode configuration in batch mode with water circulation.
and treatment time, composition of wastewater and the type of Fig. 1 presents the EC apparatus and PI scheme of the system used.
pollutants to be removed, conductivity, electrode gap, passivation The EC cell composed of austenitic steel with molybdenum (al-
of the anode, temperature, solution chemistry, water flow rate, etc. loy 316: Fe 67.8 w/w, Cr 17.6 w/w, Ni 9.9 w/w, Mo 2.5 w/w, Mn 1.6
w/w and Si 0.6 w/w) basin (65 cm  27 cm  5 cm) and six rectan-
3. Experimental gular aluminum bars (58 cm  8 cm  1 cm) oriented vertically
above the SS basin. The gap between the Al bars was about
3.1. O/W emulsions and their preparation 2.5 cm and the gap between the Al bars and the SS basin was kept
constant as 7 mm. Onwards from the 3rd test run a stainless steel
O/W emulsions treated with EC were prepared by using bio oils flow barrier was installed at the end of the basin and consequently
(rapeseed oil, sunflower, tall oil-based chain oil) and synthetic oils the height of the water column was raised to 17 mm. Since the 5th
(motor oil and chain oil-based on recycled industrial oils) from the test run the height of the water column was raised to 25 mm. The
following locations. Rapeseed oil (RSO) was cold pressed in Ilmaj- Aeff for the SS basin was 1840 cm2 (without flow barrier) and
oki at Seinäjoki University of Applied Sciences, School of Agricul- 1960 cm2/2100 cm2 (with flow barrier) when the height of water
ture and Forestry [34]. Sun flower oil (SFO) used was a column was 17 mm/25 mm, respectively. For the Al anode, Aeff
commercial product of Rainbow from Finland, tall oil-based chain was 1190 cm2 when the height of water column was 25 mm. The
oil (TO) was a product of Ekopine Oy whilst the synthetic motor volume electrolytic cell was 1.8 L/2.9 L when the height of water
oil (MO) was that of Mobil Super 3000 5 W-40. Finally, the Metsuri column was 17 mm/25 mm, respectively.
chain oil (MCO) was a mineral oil-based chain oil produced from DC power was applied to the apparatus using a GW Instek GPR-
recycled industrial oils by Ekokem Oy Ab, Finland. The metal (Ca, 1810H, 18 V/10A whilst water flow in the system was achieved
M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305 299

Table 1
Properties of the O/W emulsions prepared for EC experiments.

Test run O/W emulsion SDS (g L1) NaCl (g L1) pH Volume (L)
1: RSO 0.6%-SS RSO 0.6% 0.3 2.1 7.3 23.5
2: SFO 0.6%-SS SFO 0.6% 0.3 2.1 7.5 23.5
3: SFO 2% -SS SFO 2% 0.4 2.2 7.3 23.5
4: SFO 2% -SS SFO 2% 0.4 2.2 8.0 23.5
5: MO + MCO 2%-SS MO + MCO 2% 0.7 2.2 9.2 22.5
6: MO + MCO-SSconta MO + MCO – – 7.9 14.0/17.0b
7: MO + MCO 2%-SS MO + MCO 2% 0.7 2.2 9.2 22.5
8: RSO 2%-SS RSO 2% 0.7 2.2 6.9 22.5
9: SFO 2%-SS SFO 2% 0.4 2.2 6.2 31.0
10: SFO 2%-Al SFO 2% 0.4 2.2 6.6 23.5
11: SFO 2%-Al SFO 2% 0.4 2.2 6.0 23.5
12: MO + MCO 2%-Al MO + MCO 2% 0.7 2.2 8.7 23.5
13: MO + MCO 2%-Al MO + MCO 2% 0.7 2.2 9.0 31.0
14: MO + MCO-Alconta MO + MCO – – 8.1 13.3/18.3b
15: TO 2%-Al TO 2% 1.5 2.2 6.6 22.5
16: TO 2%-Al TO 2% 1.5 2.2 6.6 22.5
a
The repeated treatment of the effluent of the previous test run.
b
Volume of treated emulsion per volume of tap water and treated emulsion combined.

with an Iwaki MD15R-230GS water pump. Electrical current was Pure oils were added directly to the BOD OxiTopÒ bottle
measured with an Amrel 3200 Multimeter and the water temper- whereas oil-in-emulsion samples from the test runs had to be di-
ature with a TES-1300 Type K Thermometer. WTW inoLab pH 720 luted before added to the bottle. For each BOD measurement the
was used for measuring the pH. Water flow was measured electro- OxiTopÒ bottle was sealed and BOD measurements were carried
magnetically with a combination of the detector Toshiba LF470 out under a constant temperature (20 ± 0.2 °C). The biodegradation
and converter Toshiba LF420. The conductivity of the O/W emul- degree [%] was calculated from the relation of the BOD [g g1] va-
sions was measured with either a YSI 600 OMS-sonde with YSI lue given by the OxiTopÒ instrument and theoretical oxygen de-
650 MDS or WTW Multiline P4 Universal meter. mand, ThOD [g g1]. ThOD values were calculated through
carbon and hydrogen contents of the studied oil.
3.3. Metal analyses of pure oils
3.6. Particle size measurement with laser diffraction
The metal contents of the pure oils were determined by induc-
tively coupled plasma-optical emission (ICP-OES) except for RSO Droplet sizes of O/W emulsions were determined by a Beckman
which was analyzed by inductively coupled plasma-mass spec- Coulter LS 13 320 Laser Diffraction Particle Size Analyzer since this
trometry (ICP-MS). type of laser diffraction apparatus can measure particle sizes rang-
ing between 0.017 and 2000 lm [37]. The laser diffraction mea-
3.4. Analyses of COD, TOC, TSC and turbidity surements were performed for samples from test runs 11: SFO
2%-Al, 13: MO + MCO 2%-Al and 15: TO 2%-Al to compare the drop-
The chemical oxygen demand (COD) was measured using the let sizes of various O/W emulsions and to study the efficiency of EC
Hach Lange photometric cuvette test in different measuring range. breaking different types of O/W emulsions.
The vantage of the cuvette test is that samples need not be diluted.
The total organic carbon (TOC) was measured by Sievers 900 Por- 3.7. Procedure for the determination of volume and water content of
table TOC Analyzer however; samples had to be hundredfold di- sludge
luted because the limit value of the TOC measurement was
50 ppm. The total surface charge (TSC) of particles in the colloidal The procedure for the determination of volume of sludge pro-
solution was measured by Mütek PCD 03 pH whilst turbidity was duced during EC test runs was developed during experiments;
measured by a Hach Ratio XR Turbidity meter. Samples for COD, however it was later discovered that Drogui et al. [27] had used
TOC, TSC and turbidity analyses were taken for all test runs 1–7 be- a similar procedure in their experiments. After test runs, 4–9 sam-
fore and after each run. In test runs 8–16, samples were also taken ples were taken from the filling vessel and were left to settle for
during each run (in addition to before and after) to evaluate the ef- 24 h. The thickness of the formed sludge layer was then measured
fect of the treatment time on the reductions. and compared to the height of the whole water-sludge column. The
sludge percentage was then multiplied with the height of the
3.5. Analysis of BOD whole water-sludge column in the filling vessel. The volume of
the sludge was then calculated using the thickness of the sludge
High BOD values of oily wastewaters usually cause problems in layer and its surface area whilst water content of sludge was deter-
municipal treatment plant. Thus, in this study, BOD measurements mined according to the SFS-EN 12,880 [38] standard. Volume and
were carried out to determine the effect of EC treatment on BOD water content of the sludge were determined in test runs 8–11
values of O/W emulsions. The measurements of BOD28 for pure oils and 13.
(RSO, SFO, MCO and MO) and BOD7 for samples of two test runs
(11: SFO 2%-Al and 12: MO + MCO 2%-Al) treated with EC were per- 3.8. Electrocoagulation test runs
formed. The apparatus used for biological oxygen demand (BOD)
measurements was the OxiTopÒ Control system (WTW). The the- Test runs 1–9 were conducted using a SS anode and test runs
ory behind BOD measurement and its calculation can be found 10–16 using an Al anode. All experiments were performed without
more detail in [35,36]. BOD measurements were performed for specific optimization of the process parameters whilst water flow
pure oils and test runs 12: MO + MCO 2%-Al and 16: TO 2%-Al. was kept at its maximum value (about 6.1–6.4 L min1). Since
300 M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305

the 3rd test run, a 20 min blank-run without electricity was per- for the test runs 8–16 are presented in Table 2. After the optimal
formed before the actual test run to homogenize the water and treatment time had passed the COD reduction has either leveled
to ensure that no separation occurred without electricity. The cur- up or there was only minor increasing.
rent used in every test run was adjusted to its maximum value The EC of O/W emulsions resulted in repeatable high reductions
(about 8–9 A) and kept there for the duration of the test run by of COD, TOC, TSC and turbidity. The best reductions were measured
increasing or decreasing the voltage if needed. All test runs were for the TO and MO + MCO emulsions. In all the test runs the turbid-
performed at room temperature between 20 and 22 °C. The emul- ity decreased the most, 75–100% whilst the COD and TSC reductions
sions were circulated in the EC apparatus with the power switched were in the range of 25–95% and 40–98%, respectively. The TOC val-
on for the time needed for visible clarification of water (60– ues decreased the least of all the parameters measured, 20–75%.
175 min). After each test run the EC apparatus was carefully The reliability of TOC measurements was low because significant
cleaned. dilutions had to be performed for the oil emulsion samples. It
Conductivity, current, pH, temperature and voltage were mea- should be noted that the EC was effective for TO 2% emulsions that
sured during all the test runs, at the start of the run at 2-min inter- had especially high initial COD values, whose reduction was 95%.
vals and at the end of the run at 5-min interval. Documentation The repeated treatment (test runs 6 and 14) was found to remove
and sampling were emphasised towards the start of the test runs more oil from the once-treated O/W emulsions even though the ini-
since it was well known from literature that the most important tial values of COD, TSC, turbidity and TOC were quite low. The re-
changes in solutions to be treated occur in this period of EC sults show that reductions for SS and Al anodes were almost
treatment. similar; however it seems that the SS anode performed slightly bet-
ter for synthetic oils and the Al anode for bio oils. Analyses of the
samples taken during test runs 8–16 revealed that it is most likely
4. Results and discussion
the optimal treatment times would have been much shorter than
the total treatment times used here, even as short as 20 min for
4.1. Initial values of the parameters of O/W emulsions
SFO and MO + MCO emulsions with Al as the anode material. Over-
all, analyses of the heterogeneous samples were found to be some-
The initial COD, TOC, TSC and turbidity values measured after a
what difficult and caused uncertainty to the results.
20 min blank run were (without the repeated test runs) in the
Fig. 2–4 present the development of COD, TSC and turbidity val-
range of 2210–26,800 mg L1, 220–1510 mg L1, (600) to
ues during EC test runs. The parameters increase slightly at the
(10,470) leq L1 and 110–1760 NTU, respectively. The initial val-
beginning of the test runs possibly due to increasing emulsification
ues of the parameters varied greatly even for the same type of O/W
of oils in water due to the water flow, or free oil on the surface of
emulsions. This can be result of several reasons: differences in
water could have been mixed into the bulk solution. After this the
sample preparation (more or less free oil in the surface layer)
decreasing of the parameters occurred asymptotically with
and heterogeneous samples taken for analysis. The variation of ini-
increasing time. The TO emulsions had initial values higher than
tial values of parameters did not affect the EC process itself; test
those of the other O/W emulsions, which explains why the reduc-
runs for the same oil emulsion type, were repeatable. Similarly,
tion curves of TO differ from the reduction curves other O/W emul-
when dealing with real oily wastewaters, variations between
sions. The correlation between COD values (Fig. 2) and TSC values
parameter values are also expected. The emulsification of oil is a
(Fig. 3) can also be noticed. Thus, the TSC, as an easily performable
complex phenomenon and the success of emulsification depends
measurement, can be considered as a good process adjustment
strongly on the conditions of sample preparation and also on the
parameter, even on-line. The higher the TSC value of the oil-in-
oil type. The largest amount of surfactant was required for the
water emulsion sample is, the more there are oil droplets causing
emulsification of tall oil as it was also the most viscous oil. Sun
increase of the COD.
flower and rapeseed oil were the least viscous oils.

4.2. Results of COD, TOC, TSC and turbidity measurements 4.3. Results of BOD measurements

The reductions of COD, TOC, TSC and turbidity for different O/W Fig. 5 presents biodegradation degrees over 28 days for RSO,
emulsion types for test runs 1–16 and also the estimated optimal over 26 days for MCO and over 22 days for SFO and MO. The curves
treatment time without specific optimization of the conditions are drawn as average curves of replicates that were performed

Table 2
Reductions of COD, TOC, TSC and turbidity for total treatment times of different O/W emulsion types.

Test run Treatment timea (min) Current density (A m2) COD (Re-%) TOC (Re-%) TSC (Re-%) Turbidity (Re-%)
1: RSO 0.6%-SS 60 40 35 45 50 75
8: RSO 2%-SS 80(80) 40 80 55 70 75
2: SFO 0.6%-SS 60 50 25 50 60 80
3: SFO 2% -SS 60 35 55 50 75 80
4: SFO 2% -SS 60 40 95 60 70 90
9: SFO 2%-SS 60(40) 40 65 50 55 75
10: SFO 2%-Al 60(20) 70 80 – 60 98
11: SFO 2%-Al 60(20) 70 70 – 60 95
5: MO + MCO 2%-SS 60 40 90 75 98 95
6: MO + MCO-SScont 60 40 80 – 65 90
7: MO + MCO 2%-SS 60 40 90 70 80 100
12: MO + MCO 2%-Al 60(20) 70 90 50 70 85
13: MO + MCO 2%-Al 80(20) 70 85 20 55 98
14: MO + MCO-Alcont 60(16) 70 25 65 40 80
15: TO 2%-Al 150(120) 70 95 75 85 100
16: TO 2%-Al 175(150) 70 95 – 90 –
a
Estimated optimal treatment time in brackets.
M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305 301

45000 8: RSO 2%-SS 60


9: SFO 2%-SS SFO
40000 10: SFO 2%-Al
11: SFO 2%-Al 50

Biodegradation degree [%]


35000 12: MO+MCO 2%-Al
13: Mo+MCO 2%-Al RSO

30000 14: MO+MCO-Alcont 40


15: TO 2%-Al
COD [mg L-1]

16: TO 2%-Al
25000
30
MO
20000
20
15000

10000 10

5000 MCO
0
0 0 5 10 15 20 25 30
0 50 100 150 200 Time [day]
Time [min]
Fig. 5. Biodegradation degrees for RSO, SFO, MO and MCO.
Fig. 2. Development of COD values during test runs.

19% (0–400 mg L1), respectively. Differences between biodegra-


Time [min] dation degrees of the replicates with different measuring ranges
0
0 50 100 150 200 can be explained by the inhibiting effect of oil. The relative amount
-500 of oil was greater for higher measuring range than for lower mea-
-1000
suring range. The results show that bio oils biodegraded much bet-
ter than synthetic oils therefore continuous treatment of EC sludge
-1500 would be much easier if bio O/W emulsions are treated.
Vähäoja [39] studied the biodegradation of certain forestry
TSC [µeq L-1]

-2000
hydraulic oils (either mineral oils or synthetic bio oils) in ground-
-2500 8: RSO 2%-SS water and in conditions described by OECD 301 F, with an OxiTopÒ
-3000
9: SFO 2%-SS Control system. The results indicated that mineral oils usually bio-
10: SFO 2%-Al
degraded at a lesser degree than bio oils in groundwater. However
-3500 11: SFO 2%-Al
12: MO+MCO 2%-Al the usage of oils had a significant effect on the biodegradation abil-
-4000 13: MO+MCO 2%-Al ity of hydraulic oils. This is especially the case for used bio oils
14: MO+MCO-Alcont which biodegraded significantly slower than corresponding new
-4500 15: TO 2%-Al
ones. All bio oils studied except for one biodegraded significantly
16: TO 2%-Al
-5000 faster in the mineral rich standard conditions described by OECD
301 F than in groundwater.
Fig. 3. Development of TSC values during test runs. Development of the BOD7 values during EC test runs 11: SFO
2%-Al and 12: MO + MCO 2%-Al resembled each other. The BOD va-
lue decreased during EC treatment; the BOD reduction for the
9000 8: RSO 2%-SS
9: SFO 2%-SS whole EC test run being 95% for SFO 2% emulsion and 85% for
8000 10: SFO 2%-Al MO + MCO 2% emulsion. For both O/W emulsions, the decrease of
11: SFO 2%-Al BOD7 values occurred very rapidly in the first 10 min and settled
7000 12: MO+MCO 2%-Al down after that.
Turbidity [NTU]

13: Mo+MCO 2%-Al


6000
15: TO 2%-Al
5000 4.4. Results of laser diffraction measurements

4000
Laser diffraction measurements performed for samples from
3000 test runs 11: SFO 2%-Al, 13: MO + MCO 2%-Al and 15: TO 2%-Al
showed that the average droplet size was 9.8 lm, 5.1 lm and
2000
5.9 lm, respectively. The small droplet sizes as proved that the
1000 emulsification of oils had succeeded. Droplet sizes were in the
0 same range than stabilized oil-in-water emulsion (often with
0 20 40 60 80 100 120 140 160 indigenous anionic surfactant) exists in oily waste water (3–
Time [min] 20 lm) [40]. During the test runs, droplet size and the number of
size groups had increased which proved that the O/W emulsion
Fig. 4. Development of turbidity values during test runs. had started to break and flocs were starting to form. Droplet size
measurements for the samples taken from the end of the each test
run were impossible to perform because these samples did not
with different measuring ranges except for RSO (one measuring contain enough droplets. This proved that the separation of oil
range, 0–2000 mg L1). The biodegradation degrees of replicates had been successful.
for RSO were 43% and 45% where as for SFO the difference between
biodegradation degrees of the replicates was the greatest (48%, 0– 4.5. The behavior of essential process parameters during test runs
2000 mg L1 and 75%, 0–800 mg L1). For MO and MCO the biodeg-
radation degrees of replicates were almost equal 22% (0– The current densities (Eq. (8)) were calculated with average cur-
800 mg L1) and 25% (0–400 mg L1), 21% (0–800 mg L1) and rent values used in EC test runs (Table 2). They were higher for the
302 M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305

Al anode than for SS anode because the Aeff was much smaller how- Since chromium in certain speciations (especially Cr6+) is detri-
ever the current used was similar in magnitude (8A). The current mental for the environment and human health, usage of stainless
densities without a flow barrier were 40–50 A m2 for the SS anode steels as anode materials should be under strict control. On the
compared to 35–40 A m2/40 A m2 when the height of water col- other hand, in the case of stainless steels, anode surface remain
umn was increased to 17 mm/25 mm. For the Al anode, the current even as compared to those with carbon steels. Dissolving trivalent
density was 70 A m2 at a water column height of 25 mm. chromium is also an effective coagulation agent, since according to
The current used in the EC process behaved differently when the Schultze–Hardy rule the critical coagulation concentration is
using the SS or Al anode material. In the case of SS anode the cur- an extremely strong function of the charge of the coagulating ion
rent usually increased during the test run, and as a result the volt- (high valences provide a low critical coagulation concentration).
age had to be decreased by 0.5 V to prevent shortcuts. In the case of Thus, it can be assumed that the majority of chromium departs
the Al anode, the current usually decreased and so the voltage had with the oil. It must be noted, too, that initial heavy metal concen-
to be increased to maintain higher current values. To the best of trations in waste oil emulsions are often high as compared to the
our knowledge, only Yang [7] has observed the similar spontane- amount dissolving from anode. In any case, however, heavy metal
ous decreasing of voltage during the EC process, when iron elec- control must be an important part of the oil removal processes. Re-
trodes were used as anodes. moval of heavy metals is often performed with ion exchangers.
The solution temperature increased during every test run at a
rate of 0.02–0.03 °C min1 which is probably due to the warming
4.6. Volume and water content of EC sludge
effect of the current or centrifugal pump. The increase of the tem-
perature was so insignificant it was concluded that it had no affect
The samples taken during test runs 8: RSO 2%-SS and 10: SFO
on the EC process. Furthermore, the conductivity did not substan-
2%-Al are presented in Fig. 8.
tially change during test runs.
Precipitates produced during test runs either sank to the bot-
Development of pH during test runs is presented in Fig. 6 and 7
tom of the filling vessel or rose to the surface. When SS was used
for both anode materials. Observations showed that the pH in-
the precipitate had a rusty color (Fig. 8a) because of iron and
creased in every test run, as expected, with the greatest increase
was more easily separated than the precipitate containing alumi-
of pH occurring during the first 30 min until a somewhat constant
num, which was very fragile and white in color (Fig. 8b). The alu-
value was achieved. The pH increase was due to the forming of
minum precipitate was difficult to separate, especially when TO
OH ions (Eq. (6)) and increase was more substantial for emulsions
emulsions were treated however the sludge of SS contained more
with acidic and neutral initial pH. For the RSO, SFO and TO emul-
water (87–91%) than the aluminum sludge (24–82%). The SFO
sions the initial pHs were less than that for MO + MCO emulsions,
and RSO sludges contained approximately equal amounts of water.
however the increase of the pHs for bio O/W emulsions was greater
However, MO + MCO sludge contained significantly less water. The
than that for synthetic O/W emulsions with both anode materials.
volumes of produced sludge from EC treatment of different emul-
When comparing the pH increase between the two anode materials,
sion types were somewhat similar, ranging between 89 and
pH seems to increase more when the SS anode was used. This was
115 L m3.
also observed by Zongo et al. [13]. Kobya et al. [41] stated that EC
exhibits some buffering capacity especially at alkaline conditions
which can be seen more clearly in the aluminum case. The final 4.7. Advantages and disadvantages of the novel EC apparatus and
pHs of O/W emulsions were higher when the SS anode was used. electrode configuration studied
The final pH varied between 8.8 and 10.8 for the bio oils treated
with SS anode, and between 10.3 and 10.9 for synthetic oils. For test The EC apparatus used in this study performed very well in
runs 1 and 2 the final pH was lower than any of the other test runs. treating bio and synthetic O/W emulsions without significant dif-
This could be due to the low height of the water column thus the ferences between treatment results when using Al or SS anode.
effective cathode surface area for water reduction (Eq. (6)) was When using SS or Al basin as anode, vast corrosion of the basin
low (less OH ions). The final pH varied between 9.3 and 9.9 for is expected over time. However, in this study the corrosion pro-
the bio O/W emulsions treated with the Al anode, and between gressed seemingly very slowly due to the high surface area of the
9.0 and 9.1 for synthetic O/W emulsions. The greater pH increase SS basin. The usage of SS cathode as basin eliminates the need
for the SS anode than for the Al anode could be explained due to for a separated solution basin. The use of plane cathode as basin
the fact that aluminum complexes more hydroxide ions than iron enables even spreading of hydrogen bubbles for mixing as well
thus lowering the pH value. as for flotation.

1: RSO 0.6%-SS 2: SFO 0.6%-SS 5: MO+MCO 2%-SS


12 3: SFO 2%-SS 4: SFO 2%-SS 12 6: MO+MCO-SScont
8: RSO 2%-SS 9: SFO 2%-SS 7: MO+MCO 2%-SS
11 11

10 10

9 9
pH

pH

8 8

7 7

6 6

5 5
0 10 20 30 40 50 60 0 10 20 30 40 50 60
Time [min] Time [min]
(a) (b)
Fig. 6. Development of pH during test runs for SS anode for (a) bio and (b) synthetic O/W emulsions.
M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305 303

12: MO+MCO 2%-Al


10: SFO 2%-Al 11: SFO 2%-Al 13: MO+MCO 2%-Al
11 10
15: TO 2%-Al 16: TO 2%-Al 14: MO+MCO-Alcont
10
9

9
8

pH
8
pH
7
7

6
6

5 5
0 25 50 75 100 125 150 175 0 20 40 60 80
Time [min] Time [min]
(a) (b)
Fig. 7. Development of pH during test runs for Al anode for (a) bio and (b) synthetic o/w emulsions.

(a)

0 min 10 min 20 min 30 min 40 min 50 min 60 min


(b)

0 min 4 min 6 min 10 min 20 min 30 min 40 min 50 min 60 min


Fig. 8. Samples taken during EC test runs: (a) 8: RSO 2%-SS and (b) 10: SFO 2%-Al.

A lot of foam was built up between the Al electrodes on the sur-


face of the water in all the test runs 1–16. This was more obvious treatment time. The specific energy consumption (SEC in Eq. (9)) is
for the Al anode than the SS anode, which could be explained by defined as the amount of electric energy consumed per unit mass
the lighter nature of the aluminum structure and also because of of removed contaminant [15]:
the different shape of the Al anode bars and SS anode basin. During UIt
test runs 15 and 16 some of the foam was mechanically removed, SEC ¼ ð9Þ
g removed contaminant
leading to a rise in the current value. This value decreased when
the amount of foam increased again. In the case of the SS anode, where U is the cell voltage (V) during test run. As seen from Table 3
the current increase could be due to the decreasing of the resis- the energy consumptions are low even without specific optimiza-
tance according to Ohm’s law (U = I  R) in the EC cell, and in the tion of the process parameters. The energy consumption was calcu-
case of Al the foam increased resistance so the current decreased. lated as kW h (U  I  t) per sample volume (m3). For the O/W
Submerging the electrodes fully into the treated emulsion or add- emulsions, the SECCOD values ranged between 0.1 and 4.2 kW h
ing a mixing unit to the EC cell could solve this problem. COD-kg1 for the SS anode and between 0.3 and 13.8 kW h COD-
The continuous flow of solution through the filling vessel might kg1 for the Al anode. The SECCOD values for test runs 1 and 2 are
break the flocs formed if the flow rate is too high, but this was not higher because water column height was less than in rest of the test
observed in this study. With lower solution flow rates, the delay of runs. The SECCOD values for repeated test runs are somewhat high
solution in the EC cell would be longer thus lengthening the con- because of the already high level of purification of samples to be
tact time of dissolved metal cations and oil droplets, and also the treated. The energy consumption for the O/W emulsions treated
separation of flocs would be easier. varied between 1.6 and 2.8 kW h m3 using a SS anode and between
2.0 and 7.7 kW h m3 using the Al anode. The corresponding energy
costs were 0.17–0.31 € m3 and 0.21–0.84 € m3. The sacrificial
4.8. Preliminary estimations of energy consumption and treatment electrodes dissolved according to Faraday’s law and must be re-
costs of the EC process placed periodically. Therefore, Table 3 involves preliminary calcula-
tions for treatment costs ranging between 2.14 and 3.91 € m3 for a
The energy consumptions calculated for the test runs are pre- SS anode and between 0.41 and 1.46 € m3 for an Al anode. The dis-
sented in Table 3, where all the parameters were calculated for solution of iron (Fe2+) from the SS anode was calculated to be signif-
the total treatment time. Treatment costs of EC process were calcu- icantly higher than the dissolution of Al because of the greater
lated for both the total treatment time and for the optimal effective surface area. Therefore the material costs of SS were
304 M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305

Table 3
Preliminary calculations of energy consumption and treatment costs.

Test run SECCOD (kW h COD-kg1) Energy consump. (kW h m3) Energy costsa(€ m3) Dissolution of anodeb (g m3) Treatment costsa,b,c(€ m3)
1: RSO 0.6%-SS 2.8 2.0 0.22 326 2.42
8: RSO 2%-SS 1.0 2.6 0.29 530 3.86(3.86)
2: SFO 0.6%-SS 4.2 2.5 0.27 394 2.93
3: SFO 2% -SS 0.3 1.6 0.18 303 2.22
4: SFO 2% -SS 0.1 1.8 0.20 368 2.68
9: SFO 2%-SS 1.0 1.6 0.17 293 2.14(1.43)
10: SFO 2%-Al 0.3 2.0 0.21 115 0.41(0.13)
11: SFO 2%-Al 0.4 2.5 0.27 117 0.47(0.14)
5: MO + MCO 2%-SS 0.4 2.3 0.25 399 2.93
6: MO + MCO-SScont 4.0 2.8 0.31 536 3.91
7: MO + MCO 2%-SS 0.3 2.3 0.26 390 2.89
12: MO + MCO 2%-Al 0.3 2.4 0.26 115 0.46(0.13)
13: MO + MCO 2%-Al 0.8 2.5 0.28 119 0.48(0.11)
14: MO + MCO-Alcont 13.8 2.5 0.27 152 0.54(0.14)
15: TO 2%-Al 0.2 5.9 0.65 307 1.18(1.01)
16: TO 2%-Al 0.3 7.7 0.84 358 1.46(1.26)
a
Electricity price for industry (500–1999 MW h year1) in Finland in July 2011 was 0.1096 € kW h1 (including electrical energy, the distribution of electricity, and taxes)
[42].
b
Dissolution of anode (Fe2+, Al3+) based on Faraday’s law (Eq. (7)).
c
Treatment costs (energy costs + material costs based on dissolution of anode and iron content of SS, 67 w-%). The price of SS 316 hot rolled plate was 4514 € t1 [43] and
the price of Al was 1723 € t1 calculated from 2.380 USD kg1 [44]). Treatment costs for the optimal treatment time in brackets.

almost 10–20 times higher that the corresponding costs for Al. Alu- Acknowledgments
minum appeared to be a more effective coagulant, because of lower
amount of Al than Fe from a SS anode was needed for reductions of The authors would like to thank the Graduate School in Chem-
almost similar magnitude. ical Engineering (GSCE), the VALOKATA-project (University of
Oulu, The Finnish Funding Agency for Technology and Innovation
5. Conclusions (Tekes) from the European Regional Development Fund), Maa- ja
vesitekniikan tuki ry, Ekopine Oy, FA Forest Oy and Dr. Mark Jack-
In this study, bench scale EC test runs for bio and synthetic oil- son from Opal Blue for editing the English.
in-water emulsions with SS and Al anodes were performed.

 The final pHs of treated waters of similar types were higher for a References
SS anode than for an Al anode. The pH increased more for the
bio O/W emulsions than for synthetic O/W emulsions regardless [1] M. Cheryan, N. Rajagopalan, Membrane processing of oily streams, wastewater
of the anode material used. treatment and waste reduction, J. Membr. Sci. 151 (1998) 13–28.
[2] D. Mysore, T. Viraraghavan, Y-C. Jin, Treatment of oily waters using
 The precipitate produced by the SS anode was more easily sep- vermiculite, Water Res. 39 (2005) 2643–2653.
arated than the Al precipitate, but contained more bound water. [3] CONCAWE, Trends in oil discharged with aqueous effluents from oil refineries
 The test runs conducted succeeded very well with high reduc- in Europe 2005 and 2008 survey data, Report no. 2/11, March 2011, p. 17.
<http://www.concawe.org> (accessed 05.10.11).
tions of COD, TOC, TSC and turbidity for both anode materials. [4] A. Fakhru’l-Razi, A. Pendashteh, L.C. Abdullad, D.R.A. Biak, S.S. Madaeni, Z.Z.
It appears that the SS anode performed slightly better for syn- Abidin, Review of technologies for oil and gas produced water treatment, J.
thetic O/W emulsions and the Al anode for bio O/W emulsions. Hazard. Mater. 170 (2009) 530–551.
[5] B.D. Ferro, M. Smith, Global onshore and offshore water production, 2007.
 Laser diffraction and BOD analyses conducted for waters of a
<http://www.touchoilandgas.com/global-onshore-offshore-water-a7137-
number of test runs proved the breaking of O/W emulsions. 1.html> (accessed 06.07.11).
 The current behaved differently when using SS or Al as the [6] M. Cheryan, Ultrafiltration and Microfiltration Handbook, Technomic
anode material. It was assumed that the structure of the pro- Publishing Co. Inc., Pennsylvania, 1998. p. 527.
[7] C-L. Yang, Electrochemical coagulation for oily water demulsification, Sep.
duced precipitate and the unusual structure of the EC cell were Purif. Technol. 54 (2007) 388–395.
behind this. [8] E.A. Vik, D.A. Carlson, A.S. Eikum, E.T. Gjessing, Electrocoagulation of potable
 Energy consumption was low for both types of anodes. The dis- water, Water Res. 18 (1984) 1355–1360.
[9] A.E. Dieterich, Patent No. US 823671 A, Electric water purifier, 1906.
solution of Fe2+ from the SS anode was calculated to be signifi- [10] S. Tchamango, C.P. Nanseu-Njiki, E. Ngameni, D. Hadjiev, A. Darchen,
cantly higher than the Al3+ dissolution from the Al anode. Treatment of dairy effluents by electrocoagulation using aluminium
Therefore, the treatment costs using SS as anode material were electrodes, Sci. Total Environ. 408 (2010) 947–952.
[11] B. Merzouk, B. Gourich, A. Sekki, K. Madani, Ch. Vial, B. Barkaoui, Studies on the
almost trebled compared to the treatment costs using Al. decolorization of textile dye wastewater by continuous electrocoagulation
 Analyses of the samples taken during test runs 8–16 revealed process, Chem. Eng. J. 149 (2009) 207–214.
that the most likely optimal treatment times would have been [12] S. Zodi, O. Potier, F. Lapicque, J-L. Leclerc, Treatment of the textile wastewaters
by electrocoagulation: effect of operating parameters on the sludge settling
much shorter than the total treatment times used here, even characteristics, Sep. Purif. Technol. 69 (2009) 29–36.
as short as 20 min for SFO and MO + MCO emulsions with Al [13] I. Zongo, J.-P. Leclerc, H.A. Maïga, J. Wéthé, F. Lapicque, Removal of hexavalent
as the anode material. chromium from industrial wastewater by electrocoagulation: a
comprehensive comparison of aluminium and iron electrodes, Sep. Purif.
Technol. 66 (2009) 159–166.
To summarize, the EC process using SS or Al as the anode mate- [14] F. Akbal, S. Camcı, Copper, chromium and nickel removal from metal plating
rial was found to be a cost-effective and feasible method for break- wastewater by electrocoagulation, Desalination 269 (2011) 214–222.
ing the stable bio and synthetic oil-in-water emulsions studied in [15] I. Kabdasßli, T. Arslan, T. Ölmez-Hancı, I. Arslan-Alaton, O. Tünay, Complexing
agent and heavy metal removals from metal plating effluent by
this paper. This is even without the specific optimization of the electrocoagulation with stainless steel electrodes, J. Hazard. Mater. 165
most important process parameters. (2009) 838–845.
M. Karhu et al. / Separation and Purification Technology 96 (2012) 296–305 305

[16] S. Khansorthong, M. Hunsom, Remediation of wastewater from pulp and paper [31] A. Sarpola, Doctoral thesis, The hydrolysis of aluminium, a mass spectrometric
mill industry by the electrochemical technique, Chem. Eng. J. 151 (2009) 228– study, Faculty of technology, Department of process and environmental
234. engineering, Water resources and environmental engineering laboratory,
[17] M. Vepsäläinen, H. Kivisaari, M. Pulliainen, A. Oikari, M. Sillanpää, Removal of University of Oulu, Finland, 2007, p. 100. <http://herkules.oulu.fi/
toxic pollutants from pulp mill effluents by electrocoagulation, Sep. Purif. isbn9789514285578/index.html?lang=en> (accessed 05.10.11).
Technol. 81 (2011) 141–150. [32] P. Cãnizares, M. Carmona, J. Lobato, F. Martínez, M.A. Rodrigo,
[18] M. Asselin, P. Drogui, S.K. Brar, H. Benmoussa, J-F. Blais, Organics removal in Electrodissolution of aluminum electrodes in electrocoagulation processes,
oily bilgewater by electrocoagulation process, J. Hazard. Mater. 151 (2008) Ind. Eng. Chem. Res. 44 (2005) 4178–4185.
446–455. [33] J. Duan, J. Gregory, Coagulation by hydrolyzing metal salts, Adv. Colloid
[19] P. Cãnizares, F. Martínez, C. Jiménez, C. Sáez, M.A. Rodrigo, Coagulation and Interfac. 100–102 (2003) 475–502.
electrocoagulation of oil-in-water emulsions, J. Hazard. Mater. 151 (2008) 44–51. [34] V. Vauhkonen, R. Lauhanen, S. Ventelä, J. Suojaranta, A. Pasila, T. Kuokkanen, H.
[20] Y.O.A. Fouad, A.H. Konsowa, H.A. Farag, G.H. Sedahmed, Performance of an Prokkola, S. Syväjärvi, The phytotoxic effects and biodegradability of
electrocoagulation cell with horizontally oriented electrodes in oil separation stored rapeseed oil and rapeseed oil methyl ester, Agr. Food Sci. 20 (2011)
compared to a cell with vertical electrodes, Chem. Eng. J. 145 (2009) 436–440. 131–142.
[21] M.H. El-Naas, S. Al-Zuhair, A. Al-Lobaney, S. Makhlouf, Assessment of [35] M. Karhu, J. Kaakinen, T. Kuokkanen, J. Rämö, Biodegradation of light fuel oils
electrocoagulation for the treatment of petroleum refinery wastewater, J. in water and soil as determined by the manometric respirometric method,
Environ. Manage. 91 (2009) 180–185. Water Air Soil Poll. 197 (2009) 3–14.
[22] O. Chavalparit, M. Ongwandee, Optimizing electrocoagulation process for the [36] K. Roppola, Doctoral thesis, Environmental applications of manometric
treatment of biodiesel wastewater using response surface methodology, J. respirometric methods, Faculty of science, Department of the Chemistry,
Environ. Sci. 21 (2009) 1491–1496. University of Oulu, Finland, 2009, p. 82. <http://herkules.oulu.fi/
[23] N. Adhoum, L. Monser, Decolourization and removal of phenolic compounds isbn9789514290794/index.html?lang=en> (accessed 03.10.11).
from olive mill wastewater by electrocoagulation, Che. Eng. Process. 43 (2004) [37] Beckman Coulter Inc., Meeting and exceeding the standard for laser diffraction
1281–1287. particle size analysis, LS 13 320 particle size analyzer, Brea (CA), Beckman
[24] U.T. Un, A.S. Koparal, U.B. Ogutveren, Electrocoagulation of vegetable oil Coulter Inc., 2009.
refinery wastewater using aluminium electrodes, J. Environ. Manage. 90 [38] SFS-EN 12880, Characterization of sludges, Determination of dry residue and
(2009) 428–433. water content, Lietteen karakterisointi, Kuiva-aineen ja vesipitoisuuden
[25] X. Chen, G. Chen, P.L. Yue, Separation of pollutants from restaurant wastewater määrittäminen, Helsinki, Finland, 2000, p. 8.
by electrocoagulation, Sep. Purif. Technol. 19 (2000) 65–75. [39] P. Vähäoja, Doctoral thesis, Oil analysis in machine diagnostics, Faculty of
[26] X. Xu, X. Zhu, Treatment of refectory oily wastewater by electro-coagulation science, Department of the Chemistry, Department of mechanical engineering,
process, Chemosphere 56 (2004) 889–894. University of Oulu, Finland, 2006, p. 76. <http://herkules.oulu.fi/
[27] P. Drogui, M. Asselin, S.K. Brar, H. Benmoussa, J-F. Blais, Electrochemical isbn9514280768/index.html?lang=en> (accessed 05.10.11).
removal of pollutants from agro-industry wastewaters, Sep. Purif. Technol. 61 [40] R. Moosai, R.A. Dawe, Gas attachment of oil droplets for gas flotation for oily
(2008) 301–310. wastewater cleanup, Sep. Purif. Technol. 33 (2003) 304–314.
[28] S. Bayar, Y.S ß . Yıldız, A.E. Yılmaz, S _
ß . Irdemez, The effect of stirring speed and [41] M. Kobya, M. Bayramoglu, M. Eyvaz, Techno-economical evalution of
current density on removal efficiency of poultry slaughterhouse wastewater electrocoagulation for the textile wastewater using different electrode
by electrocoagulation method, Desalination 280 (2011) 103–107. connections, J. Hazard. Mater. 148 (2007) 311–318.
_ Ayhan S
[29] E. Yüksel, I. ß engil, M. Özacar, The removal of sodium dodecyl sulfate in [42] Statistics Finland, Energy prices. <http://www.stat.fi/index_en.html>
synthetic wastewater by peroxi-electrocoagulation method, Chem. Eng. J. 152 (accessed 19.10.11).
(2009) 347–353. [43] MEPS – EU stainless steel product price. <http://www.meps.co.uk/
[30] M.Y.A. Mollah, P. Morkovsky, J.A.G. Gomes, M. Kesmez, J. Parga, D.L. Cocke, Stainless%20Price-eu.htm> (accessed 12.09.11).
Fundamentals, present and future perspectives of electrocoagulation, J. [44] Metal Prices, Current primary and scrap metal prices. <http://
Hazard. Mater. 114 (2004) 199–210. www.metalprices.com/> (accessed 12.09.11).

You might also like