Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Ecotoxicology and Environmental Safety 145 (2017) 184–192

Contents lists available at ScienceDirect

Ecotoxicology and Environmental Safety


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

Treatment of highly toxic cardboard plant wastewater by a combination of MARK


electrocoagulation and electrooxidation processes
Erhan Gengec
University of Kocaeli, Department of Environmental Protection, Kocaeli, Turkey

A R T I C L E I N F O A B S T R A C T

Keywords: The objective of this study was to investigate the removal efficiencies of the electrochemical treatment systems
Cardboard Wastewater as an alternative for the treatment of cardboard plant wastewater (CPW). In accordance with this purpose, CPW
Toxicity was treated by electrocoagulation (EC) with Al electrodes and the effects of current density (CD), operating time
Electrocoagulation (t), and initial pH (pHi) were investigated. The results showed that EC at optimum treatment conditions (CD:
Electrooxidation
7.5 mA/cm2, pHi: 7.0 and t: 60 min) have limited removal efficiencies for total organic carbon (TOC; 17.1%) and
Boron-doped Diamond Electrode
chemical oxygen demand (COD, 14.2%), on the contrary of turbidity (98.7%). Due to the low TOC and COD
removal efficiencies, a secondary treatment was needed and the electrocoagulated effluent was subjected to
electrooxidation (EO) by using a boron doped diamond (BDD) electrode for investigating the effect of CD, t, pHi
and electrolyte concentration (Ce). Higher TOC (83.7%) and COD (82.9%) removal efficiencies were obtained by
EO under the optimum treatment conditions (CD: 100 mA/cm2, pHi: 7.2, Ce: 5.0 g/L Na2SO4 and t: 180 min). In
addition, a toxicity test was carried out to the raw and treated wastewater under the optimum operating con-
ditions. This study demonstrated that the combination of EC and EO have a satisfactory potential for real in-
dustrial wastewater with a high organic content, suspended solids and toxicity.

1. Introduction of large volumes of toxic sludge, and (II) an increase in metal con-
centration in treated water are the main disadvantages of these systems.
The pulp and paper industry use lignocellulogic materials (wood Another alternative treatment system in the paper industry is the bio-
and/or recycled paper etc.) as a raw material, and produce paper, logical method (mainly aerated lagoon), which is commonly used due
cardboard and other cellulose-based products. During production of the to the low cost and minimal operation requirements. The biological
paper and cardboard, wastewater that has high suspended solids, color, methods are also very effective for the removal of soluble and biode-
and organic substance in the range from 75 to 275 m3 per ton of pro- gradable organic pollutants (Pokhrel and Viraraghavan, 2004). How-
duct (Jaafarzadeh et al., 2016) is generated. Moreover, this industrial ever, the existence of bio-refractory components in wastewater, the
sector wastewater contains the chlorinated lignosulphonic acids, resin sensitivity of microorganisms to toxic material, production of high
acids, phenols and hydrocarbons besides dibenzo-p-dioxin and di- amounts of sludge and lengthy time requirements are the disadvantages
benzofuran, highly toxic and recalcitrant compounds (Kumar et al., of the biological methods (Renault et al., 2009). Due to the mentioned
2015). Due to these components, these effluents are extremely ha- disadvantages, a physico-chemical method is usually needed before a
zardous for the ecosystem because they exhibit a strong mutagenic ef- biological method. Furthermore, adsorption, ozonation and membrane
fect, have acute toxicity, cause a decreasing of oxygen, and prevent filtration of chlorinated phenolic compounds and adsorable organic
light penetration. Therefore, various wastewater treatment systems and halides and chemical oxidation of colorful components were carried out
their combinations have been employed to minimize the negative in- effectively in paper industry wastewater (Pokhrel and Viraraghavan,
fluence of pulp and paper industry effluents on the environment until 2004). However, these processes involve a great deal of disadvantages
now (Birjandi et al., 2014). Inside these treatment systems, coagulation such as high cost, selective removal and difficulties in practice. On the
and flocculation processes in which polyaluminium chloride (PAC) and other hand, due to the increasing fresh water cost and the limited dis-
aluminum sulphate are used are very effective for the removal of sus- charge standards of wastewater, the focus of researchers has shifted to
pended solids (SS), colloids and toxicity in wastewater. However, there relatively new alternatives (Mansour and Kesentini, 2008) such as
are two problems involving the potential ecotoxicity of metallic coa- electrochemical treatment systems. Electrocoagulation (EC), electro-
gulants and other toxic components in wastewater: (I) The production oxidation (EO), and electro-fenton (EF) are the most promising

E-mail address: erhan.gengec@kocaeli.edu.tr.

http://dx.doi.org/10.1016/j.ecoenv.2017.07.032
Received 5 June 2017; Received in revised form 10 July 2017; Accepted 15 July 2017
0147-6513/ © 2017 Elsevier Inc. All rights reserved.
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

electrochemical treatment systems. of EC with EO can be a good alternative for treatment of the CPW
Due to a great deal of competitive advantages such as no selectivity (Jaafarzadeh et al., 2016; Zaidi et al., 2016). In addition, EO presents a
towards pollutants, low operating cost, easy operable equipment, short large number of advantages: (I) the oxidation of pollutants up to the
operation time, low requirement of chemicals, and low production final product or to less toxic by-products in contrast with coagulation,
amount of sludge, the EC process is frequently suggested in literature membrane, filtration, and adsorption processes, (II) short oxidation
for the removal of chemical oxygen demand (COD), biological oxygen time requirements, no sludge problem, non-selective to toxic and non-
demand (BOD), total organic carbon (TOC), suspended solids (SS), biodegradable material in contrast with biological oxidation processes,
turbidity, and toxicity (David et al., 2015; Makwana and Ahammed, and (III) addition of chemical reagents not required (except for elec-
2016). Generally, a power supply (alternating current; AC or direct trolytes or catalyst addition at low amounts) in contrast with chemical
current; DC) and electrodes (aluminum or iron) are used during the EC oxidation. EO can be categorized as direct and indirect oxidation. In the
process and anode electrodes are consumed under a potential. Thus, indirect oxidation process, oxidizing reagents such as chlorine, hypo-
three successful stages during this process are realized: (i) formation of chlorous acid, hypochlorite, hydrogen peroxide and ozone are formed
coagulants by electrolytic oxidation of the sacrificial anode electrode, at the electrodes surfaces and then, destroy pollutants in wastewater.
(ii) destabilization of the pollutants, particulate suspension, and The direct oxidation process involves (I) electron transfer to the pol-
breaking of emulsions, (iii) aggregation of the destabilized phases to lutants and (II) the forming of hydroxyl radicals on anode's surface
form flocs. The main reaction at the Al anode is shown as follows (Särkkä et al., 2015; Wu et al., 2014). When direct and indirect oxi-
(Dastyar et al., 2015; Elabbas et al., 2015; Ezechi et al., 2014; Hamdan dations are compared, direct oxidation has some advantages such as
and El-Naas, 2014; Makwana and Ahammed, 2016; Mella et al., 2015; requiring a lower amount of chemical substance and producing no
Vasudevan, 2012; Zaidi et al., 2016): secondary contaminants (Särkkä et al., 2015). The electrocatalytic ac-
tivity and electrochemical stability of the electrode is a key point for
Al → Al3 + + 3e− (1)
direct oxidation. Although several anode materials have been used in
In addition to aluminum dissolution, oxygen evolution is realized the direct oxidation, the complete mineralization of the pollutants has
via a reaction (Eqn. 2) at the anode: been possible by use of high oxygen overvoltage anodes such as SnO2,
PbO2 and boron doped diamond (BDD) (Panizza and Cerisola, 2001).
2H2 O → O2(g) + 4H+ + 4e− (2)
The BDD electrode has been identified as an excellent non-active anode
At cathode, hydrogen evolution is realized by the following reaction for the removal of pollutants in wastewater (Özcan et al., 2008) and the
(Eq. (3)) and the formed hydrogen causes the floatation of pollutants. formation of hydroxyl radical (•OH) on the BDD anode surface has been
Moreover, the produced hydroxyls ions at cathode generally cause an reported as the main removal mechanism. Due to having an inert sur-
increase of pH. face and low adsorption capacity, the BDD surface interacts weakly
with the •OH. During EO with BDD, the complete mineralization of
3H2 O + 3e− → 3/2H2 + 3OH− (3)
pollutants in wastewater is possible due to the physisorbed •OH by the
The hydrolysis constants for aluminum cover a very narrow range, following reactions (Eqs. (7) and (8), R: organic pollutants in waste-
and all of the aluminum deprotonations are ‘squeezed’ into an interval water) (Gao et al., 2004; Marco Antonio Quiroz et al., 2006; Martínez-
of three pH units. Therefore, except for a narrow pH region (approxi- Huitle and Ferro, 2006; Scialdone, 2009; Sires et al., 2014)
mately 5–8), the dominant soluble species are Al3+ at low pH and
M+ H2 O → M(HO•) + H+ + e (7)
Al(OH)−4 at high pH (Gengec et al., 2012; Kobya et al., 2014; Kobya and
Gengec, 2012). The produced positively charged metal ions (Al3+) in M(HO•) + R→ M + mCO2 + nH2 O + H+ + e (8)
the acidic conditions neutralize the negatively charged pollutants. At
about neutral pH (5–8) metal ions hydrolyze to monomeric-polymeric Although scientists recent attention has focused on EO, there is a
aluminum hydroxides and amorphous species, Al(OH)3 (sweep coagu- lack of information in literature regarding EO: (I) removal efficiencies
lation). Monomeric-polymeric species such as Al(OH)2 +, Al(OH)+2 , of EO in different real wastewater (only 20% EO paper in the last 3
Al(OH)−4 , Al2 (OH)24 +, Al 6 (OH)15
3+ 4+
, Al7 (OH)17 4+
, Al8 (OH)20 , Al13O4 (OH)724+, years have dealt with treatment of the real wastewater), (II) comparison
and Al13 (OH)34 transform initially into Al(OH)3(s) and finally poly-
5+ of the electrodes performance, (III) contribution level on removal effi-
merize to Aln(OH)3n (Eqs. (4) and (5)) according to the pH of the ciencies of EO at a combination of different treatment systems, (IV)
wastewater and concentration of Al: removal mechanism, (V) electrical consumption (Sires et al., 2014) and
(VI) possible formation of toxic by-products. To the best of the author's
Al3 + + 3H2 O → Al(OH)3(s) + 3H+ (4) knowledge, the combination of the electrocoagulation (Al electrode)
and electrooxidation processes (BDD electrode) has yet to be studied for
nAl(OH)3 → Aln (OH)3n (5)
the treatment of CPW. Moreover, the influence of the EC and EO
These hydroxides are excellent coagulants and cause adsorption of combination on toxicity of real wastewater has not been investigated
pollutants. On the other hand, the formed negatively charged hydro- sufficiently.
xides such as Al (OH )−4 (Eq. (6)) usually reduce the removal efficiencies Therefore, the purposes of this study were: (I) present the combi-
at high pH values (> 9); nation of EC and EO as a competitive alternative for the treatment of
industrial wastewater with high suspended solids, organic content and
2Al + 6H2 O + 2OH → Al(OH)4 + 3H2(g) (6) toxicity, (II) investigate the effect of process conditions on removal
Until now, EC technology has been reported as an effective treat- efficiencies, and (III) calculate the electrical and electrode consumption
ment system for industrial wastewater such as electroplating (Kobya during these processes.
et al., 2016), distillery (David et al., 2015), tannery (Elabbas et al.,
2015), baker's yeast production (Gengec et al., 2012) and textile dyeing 2. Material and methods
(Kobya et al., 2014). Therefore, removal of turbidity and toxicity in
cardboard plant wastewater (CPW) can be carried out successfully by 2.1. Wastewater
EC. However, the mineralization of organic components in CPW up to
the legal discharge standards of wastewater does not seem possible via The wastewater was taken from the starting point of a cardboard
EC. plant treatment system. The coated cardboard is made out of recycled
EO is an effective method for the mineralization of organic sub- paper and the annual cardboard production capacity of the plant has
stances in wastewater, though not for turbidity. Thus, the combination reached 240,000 t (Table 1).

185
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

Table 1 2.3. The toxicity test


Some physical and chemical specification of raw cardboard wastewater.
The toxicity test was performed according to the International
Parameters Parameters
Standard Method (ISO) 21338: Water Quality-Kinetic determination of
Flow rate (m3/day)a 6000–8000 COD (mg/L) 2358 the inhibitory effects of sediment, other solids and colored samples on
pH 6.9–7.1 TOC (mg/L) 1006 the light emission of Vibrio fischeri (kinetic luminescent bacteria test)
Conductivity (mS/cm) 2.271–2450 NH3-N (mg/L) < 0.5
(International Standard Method 21338, 2010). Luminescence was
Turbidity (NTU) 783 SO4 (mg/L) 241
Total Solids (mg/L) 5000–5850 PO4-P (mg/L) 2.8 measured using a high performance Sirius luminometer and light
Settleable Solids (mL/L) 750–850 EC50b 20.11 output was recorded automatically by FB12 software (Berthold detec-
Total Suspended Solids (mg/L) 3000–5000 EC20b 6.52 tion systems, Germany). Then, the inhibition percentage, EC50 (the
a
concentration that causes 50% reduction of bacteria relative to control)
The information was supplied from the manufacturer. Other specifications were
and EC20 values were calculated (Al-Mutairi, 2006; International
measured during this work.
b
EC50 and EC20: The concentration of wastewater that causes 50 % and 20 % of Standard Method 21338, 2010).
bacteria reduction relative to control sample, respectively. In brief, the toxicity test was realized over six steps: (I) the waste-
water was diluted with 2% NaCl solution to obtain dilution series (1:2,
2.2. The experimental set-up and analytical methods 1:4, 1:8, 1:16, 1:32 and 1:64), (II) the Vibrio fischeri was injected to the
diluted wastewater, (III) luminescence of Vibrio fischeri in diluted
The EC experiments were realized by a batch process using a 1 L wastewater was measured every 0.1 s during the initial five seconds,
capacity of a reactor made from Plexiglas with dimensions of (IV) the maximum value recorded, (V) II, III and IV steps were carried
12×11×11 cm. Aluminum (Al) plates (purity > 99.5%) with dimen- out for all of the diluted wastewater and control samples, (V) lumi-
sions of 5.0 × 8.0 × 0.3 cm were used as the sacrificial electrodes. In nescence of Vibrio fischeri was re-measured after a contact time of
each run, four Al plate electrodes (two anodes and two cathodes) 30 min, (VI) the ratios of the inhibitory effect (Inh %) and effective
spaced by 0.5 cm were placed vertically in the reactor and connected at concentrations (EC20 or EC50) were calculated by the luminescence
the monopolar parallel mode. All the EC runs were performed with results of bacteria in this dilution series.
0.85 L of wastewater at a constant 25 °C and 300 rpm (Heidolp MR
3000D). Before each run, the impurities on the electrode surfaces were 3. Results and discussion
removed by a hydrochloric acid-hexamethylenetetramine aqueous so-
lution. The electrodes were dipped into the wastewater to a depth of As seen in Table 1, CPW has a high organic content, suspended
80 mm yielding a total effective electrode surface area of 120 cm2 solids and toxicity. In this study, the effect of initial pH (pHi), current
(Gengec et al., 2012). Then, the current was arranged at the desired density (CD, mA/cm2), and operating time (t, min) were investigated
value and the experiment was started. The samples taken from the EC during EC with Al electrode. Following this, the electrocoagulated
reactor at different operating times were analyzed. At the end of the EC wastewater was treated by EO processes under different experimental
runs, electrical energy consumption (ENC, kWh/m3, Eqn. 9) and elec- conditions: pHi, CD, t, and electrolyte concentration (Ce) with BDD
trode consumption (ELC, kg/m3, Eq. (10)) were calculated via Faraday's electrode.
Law:
3.1. Effect of processes parameters on EC
U×i×t
ENC =
v (9) It is well known that pHi, CD (A/m2) and t (min) are the major
parameters which affect the removal efficiencies in the EC process
i × t × Mw (Hamdan and El-Naas, 2014; Jaafarzadeh et al., 2016; Kobya et al.,
ELC =
z×F×v (10) 2016; Makwana and Ahammed, 2016). The optimum values of these
parameters are changeable according to the characterization of the
where U is cell voltage (V), I is current (A), t is operating time (s) and v wastewater (pHi, pollutants species, affinity of pollutants etc.) and
is volume (m3) of wastewater, Mw is molecular mass of aluminum species of electrodes (Al, Fe etc.). Therefore, Al electrode was used as a
(26.98 g/mol), zAl is number of electron transferred (zAl = 3) and F is sacrificial electrode in this study and optimum process conditions were
Faraday's constant (96,487 C/mol). evaluated by a series of experiments in order to remove the COD, TOC
EO experiments were conducted in batch mode by using an open, and turbidity.
undivided, glass cell (12 × 9 × 7 cm). The 0.75 L capacity glass cell The effect of pHi on treatment efficiencies was examined at 3.0, 4.0,
was surrounded by a double jacket for the circulation of external 6.0, 7.0, 9.0, and 10.0 pH's over a period of 60 min. It is clear that when
thermostated water, controlled at 25 ± 1 °C. A thin-film of BDD elec- the pH value was altered towards two extremes, the TOC and COD
trode was used in EO experiments and the effective area of the anodes removal efficiencies decreased (Fig. 1a and b). The maximum removal
was 39 cm2. The BDD anode was supplied by DiaCCon GmbH, the efficiencies were obtained for TOC (16.8%) and COD (10.1%) at pH 7,
cathode was a stainless steel sheet and the inter-electrode distance was and for turbidity (97%) at pHi 10. The original pHi value (pH 7) of CPW
about 1.5 cm. All EO experiments were performed with 0.5 L of was- seemed as the optimum one due to the unnecessary pH arrangement
tewater at 300 rpm (Heidolp MR 3000D). Na2SO4 was preferred as an and the obtained high removal efficiencies for TOC (16.8%), COD
electrolyte and added to the wastewater in various amounts. The (10.1%) and turbidity (95%, residual turbidity: 38 NTU).
electrodes were connected to a digital DC power supply (Agilent 6675A During the EC process, three major interaction mechanisms, charge
model; 30 V, 6 A) operated at galvanostatic mode and the current was neutralization, adsorption and sweep flocculation are dominant at dif-
held constant at the desired values for each run. ferent pH ranges and metal concentrations since pH and metal con-
The TOC levels were determined by the combustion of the samples centration affect the solubility of metal species in wastewater (Fig. 2c,
at 680 °C (Shimadzu, TOC-L model). COD measurements were realized only mononuclear Al species were considered). In low pH values,
by Standard Methods (APHA, 1998). The pH of the sample was adjusted charge neutralization is dominant and realized by mono metal species
with 0.1 M H2SO4 or NaOH and measured by a pH meter (WTW Inolab such as Al+3 which are produced at the anodes. These positively
pH 720). The analysis was triplicated and the average data was re- charged metal ions neutralize the anionic charges of pollutants in
ported. The accuracy of these measured values for COD, TOC and tur- wastewater and reduce their solubility. Between the pH range 5.0–8.0,
bidity were calculated below 3.0%. amorphous metal hydroxides which have a large specific surface are

186
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

Fig. 1. Effect of pH (CD: 6.0 mA/cm2, initial TOC: 1006 mg/L, initial COD: 2358 mg/L, and initial turbidity: 783 NTU) on (a) TOC, (b) COD, (c) turbidity, and (d) Activity–pH diagram for
Al(III) species in equilibrium with Al(OH)3 (amorphous) (Ghernaout et al., 2009; Metcalf and Eddy, 2004; Yilmaz et al., 2007).

formed and sweep flocculation of pollutants is possible. Moreover, cm2, 7 of pHi, and 60 min were selected as the optimum operating
polynuclear and polymeric species at higher and lower pH values cause conditions. The EC process was an effective method for the removal of
adsorption (Elabbas et al., 2015; Gengec et al., 2012; Metcalf and Eddy, turbidity during the treatment of CPW. In addition, the quality of the EC
2004). In this study, at 6.0 mA/cm2 and 7 of pH, the produced Al effluent under the optimum conditions (residual TOC: 834 mg/L re-
concentration was calculated by Faraday's laws (Eqn. 10) as 1.76 × 10−3 , sidual COD:2025 mg/L and residual turbidity: 10 NTU) meets the reuse
3.51 × 10−3 , 7.02 × 10−3 , 1.05 × 10−2 , 1.4 × 10−2 , 1.76 × 10−2 , and standards for turbidity in the paper industry, which is 14–56 NTU
2.11 × 10−2 molar for 5, 10, 20, 30, 40, 50, and 60 min, respectively. (Terrazas et al., 2010) during the paper bleaching stage. However, the
According to these results and Fig. 1c, sweep flocculation seemed as the obtained TOC and COD removal efficiencies with the EC were very low
dominant removal mechanism in this study. when compared with other studies which dealt with the treatment of
As mentioned before, the pH and the Al concentration affect the the paper industries wastewater by EC (Jaafarzadeh et al., 2016). The
species of Al which not only influences the treatment efficiency but also various used raw materials (recycled paper, wood etc.) and the pro-
the operating cost of the EC process. According to Faraday Law's (Eq. duction conditions of cardboard may cause this phenomenon. As a re-
(10)), the dissolved amount of coagulant at the anode depends on sult, it was determined that the EC could be used efficiently for the
current density (calculated by dividing of the applied current by the removal of high molecular weight dissolved organic matters (Kamali
active electrode area), and operating time. Moreover, current density and Khodaparast, 2015), turbidity and suspended solids and it is a
and operation time determine the bubble production rate at the elec- competitive alternative to be conducted prior to an oxidation treatment
trode, which causes electroflotation. For the reasons mentioned above, system.
the effects of current density and operating time were investigated
(Fig. 2) in the range of 3.0–12.0 mA/cm2 and 0–60 min, respectively.
The maximum removal efficiencies at 3.0, 6.0, 7.5, 12.0 mA/cm2 3.2. Effect of processes parameters on EO
during 60 min were obtained as 14.2%, 16.8%, 17.1%, and 18.2% for
TOC, 10.2, 11.0, 12.8, 13.8, and 14.1 for COD, and 92.3%, 95.9%, During the EO process, the current density is one of the most im-
98.7%, and 98.7% for turbidity, respectively. As seen in results, the portant driving forces because it not only affects the degradation rate of
increased current density and operating time improved removal effi- the EO but also the reactions rate (bubble production rate, O2 reduc-
ciencies slightly and caused an increasing of electrical (Fig. 2c) and tion/evaluation rates etc.) (Gengec et al., 2012; Kourdali et al., 2014).
electrode consumptions. The electrical (Eq. (9)) and electrode con- Thus, the electrocoagulated CPW at the presence of 2 g/L Na2SO4 as the
sumptions (Eq. (10)) were calculated respectively as 1.85, 5.50, 9.18, electrolyte was treated at different current densities (from 25 mA/cm2
13.90, and 21.30 kWh/m3 and 0.142, 0.284, 0.355, 0.466 and 0.569 kg to 100 mA/cm2). When 25, 50, 75 and 100 mA/cm2 of the current
Al/m3 for 3.0, 6.0, 7.5 and 12.0 mA/cm2 during 60 min. Thus, 7.5 mA/ densities were applied, the removal efficiencies were increased to
47.5%, 66.8%, 77.9%, and 88.2% for TOC and 49.2%, 68.0%, 80.0%,

187
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

Fig. 2. Effect of current density (pHi: 7.0, initial TOC: 1006 mg/L, initial COD: 2358 mg/L, and initial turbidity: 783 NTU) on (a) TOC (b) COD and (c) Turbidity.

and 89.2% for COD during 180. min (Fig. 3a and b), respectively. It is increase of removal efficiencies (Jalife-Jacobo et al., 2015). Most stu-
well known that the rise of the current density causes the increase of dies have shown that the highest removal efficiencies were obtained by

OH generation on the BDD anode. Therefore, higher removal effi- usage of NaCl in comparison with other electrolytes. However, it was
ciencies were possible using an increased current density and these reported that the usage of NaCl caused the production of many orga-
results are well in agreement with literature. For instance, Jalife-Jacobo nochlorinated species as intermediates and final products that can be
et al. (2015) studied the diazo dye Congo Red (3 L of 100 mg/L) de- even more harmful than the raw pollutants (Comninellis and Pulgarin,
gradation using a BDD anode. They reported that when the current 1991; Gotsi et al., 2005; Sires et al., 2014). Therefore, Na2SO4 was used
density rose from 7.5 to 50 mA/cm2 at the presence of Na2SO4, removal between 2.0 g/L and 6.0 g/L of concentrations at 7.2 of pH and
efficiencies were increased from 15% to 50%, respectively. Moreover, 100 mA/cm2 of CD in order to observe the effects of supporting elec-
Jong Young Choi et al. (2010) investigated the effect of the various CD trolyte on electrical consumption and removal efficiencies. The addition
(5, 15, and 25 mA/cm2) on the degradation of 1,4-dioxane, a refractory of 2.0, 3.0, 4.0, 5.0, and 6.0 g/L Na2SO4 (Fig. 4a) decreased the po-
water pollutant with COD = 32 mol O2/m3, by BDD electrode at 5.5 of tentials of the process (28.99, 25.75, 22.46, 20.07, and 19.06 V) and
pHi, and the presence of 5 g/L of Na2SO4. They showed that a higher electrical consumption (678.29, 602.51, 525.64, 469.52 and
current density had a much faster degradation rate. The complete de- 446.07 kWh/m3), respectively. According to these results, the addition
gradation of 1,4-dioxane occurred at about 18 min at 5 mA/cm2 and of 5.0 g/L Na2SO4 seemed as an optimum electrolyte concentration due
7 min at 25 mA/cm2 (the results were seen at graph). As can be seen by to providing balance between electrical and chemical consumption.
these studies, current density in EO systems affect the removal effi- These results indicated that electrical consumption could be decreased
ciencies depending on other conditions such as pollutant type, the up to 34% by adding electrolyte. On the other hand, the effect of adding
presence of electrolyte, electrolyte species, operating time and pH. electrolyte into wastewater was commonly reported as not only causing
It must be noted that the rise current density from 25 mA/cm2 to the decrease of electrical consumption but also the increase of removal
100 mA/cm2 (Fig. 3c) not only increased the removal efficiencies but efficiencies. Generally, significant progresses in removal efficiencies by
also the electrical consumption (Eqn. 9) from 76.03 kWh/m3 to the forming of radicals and stable oxidants were reported by the usage
678.28 kWh/m3 (average volt from 12.93 to 28.99). These higher of Na2SO4 (Eqs. (11) and (12)) (Jalife-Jacobo et al., 2015; Sires et al.,
electrical consumptions should be decreased for economic reasons. 2014):
The best way of lowering electrical consumption was to decrease
SO24− → (SO−4 )• + e− (11)
wastewater resistance via the addition of electrolyte into the waste-
water. It is well known that adding electrolyte such as NaCl, KCl,
(SO−4 )• + (SO−4 )• → S2 O82 − (12)
Na2SO4, HClO4, and H2SO4 decreased electrical consumption in all
electrochemical treatment processes. In addition to the decreasing However, in this study the increasing of electrolyte (2.0, 3.0, 4.0,
electrical consumptions, the inclusion of electrolyte usually causes an 5.0, and 6.0 g/L Na2SO4) caused a slightly decreasing in removal

188
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

Fig. 3. Effect of current density and time on (a) TOC removal, (b) COD removal and (c) electric voltage (pHi: 7.2, Ce: 2 g Na2SO4/L, initial TOC: 1006 mg/L, initial COD: 2358 mg/L, and
initial turbidity: 783 NTU).

efficiencies (88.24%, 87.28%, 84.92%, 83.66%, and 83.65% for TOC the nature of the wastewater.
and 89.2%, 85.8%, 82.6%, 82.86%, and 81.1% for COD) (Fig. 4a and
b). The stated decline in removal efficiencies (only 5% for TOC and 8%
for COD) showed that the radicals were not formed or the formed ra- 3.3. Evaluation of acute toxicity of wastewater during EC and EO
dicals (SO−4 )• had an antagonistic effect on the other components in
wastewater under tested conditions. Similar to our results, a slight de- Generally, the toxicity levels of wastewater are identified by the
cline or no effect on removal efficiencies during an increased amount of EC50 and EC20 values. The EC50 and EC20 values are the concentration
electrolyte was reported in some other studies (Bottesi, 2016; Choi of wastewater that causes 50% and 20% of bacteria reduction relative
et al., 2010). to a control sample, respectively. In this study, these values were de-
The pH of wastewater affects chemical properties of pollutants and termined according to the ISO Standard Method (21338, 2010) by Vi-
the interaction of pollutants with the electrode in EO. Therefore, the brio fischeri, a species of bioluminescent bacterium.
effect of the initial pH was investigated in Na2SO4 medium (TOCi: The EC50 and EC20 values of raw wastewater were calculated as
836 mg/L, CODi: 2025 mg/L, Ce: 5.0 g/L Na2SO4 and CD: 100 mA/ 20.11 and 6.52, respectively. It meant that the wastewater had high
cm2). As seen in Fig. 5, after 180 min of electrolysis, the removal effi- acute toxicity and when wastewater was evenly diluted at the ratio of
ciencies at pH 2.0, 4.0, 6.0, 7.2, 8.0, and 10.0 were recorded as 78.7%, 20.11 and 6.52, it caused a 50% and 20% of bacteria reduction in
79.0%, 82.2%, 83.7%, 84.3% and 85.7% for TOC and 75.0%, 77.1%, 30 min, respectively. As seen in Table 2, the inhibition ratio for treated
79.4%, 82.9%, 83.0% and 84.0% for COD, respectively. The degrada- wastewater by EC and EO were below 50% for the whole dilution series.
tion efficiencies during the EO were slightly higher at > 6 pH than These results indicated that the treated wastewater was not toxic
acidic medium. Therefore, EO results showed that the electro- (Kováts et al., 2012). Moreover, the measured negative inhibition va-
coagulated wastewater could be used in EO without any pH arrange- lues demonstrated that better life conditions for bacteria were formed
ment due to the final pH of electrocoagulated wastewater measured as after treatment by EC and EO. The possible way for negative inhibition
7.2 under the optimum process condition of EC, as stated in Section 3.1. values is explained by the presence of organic substances (non-toxic or
Similar to our results, Lissens et al. (2003), Canizares et al. low concentration of toxic compound) which were used as nutrients by
(2004a,2004b) and Koparal et al. (2007) reported that the oxidation bacteria and thus, the amount of bacteria increased. The obtained
process by BDD electrode is more favourable in neutral and/or alkaline toxicity test results were quite promising when compared with the lit-
medium. On the other hand, Choi et al. (2010) indicated that pH had no erature. In contrast with our findings, some studies which dealt with
influence on removal efficiencies and Scialdone et al. (2008) reported toxic wastewater reported that toxicity during the EO process rose due
higher removal efficiencies at an acidic medium. According to this lit- to the increase of toxic materials such as organochlorinated species
erature, it can be concluded that the effect of pH strongly depends on (Chatzisymeon et al., 2006; Gotsi et al., 2005).

189
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

Fig. 4. Effect of electrolyte concentration and time on (a) electric potential, (b) TOC and (c) COD removal (CD:100 mA/cm2, pHi: 7.2, initial TOC: 1006 mg/L, initial COD: 2358 mg/L,
and initial turbidity: 783 NTU).

4. Conclusions fraction of organic component was removed during EO systems. Current


density and operating time had an important effect on the removal
The EC and EO processes were used to treat cardboard plant was- efficiency for both electrochemical treatment systems. Moreover, both
tewater. The optimum process conditions were determined as 7.5 mA/ processes showed optimum removal efficiencies at a neutral pH value.
cm2, 7.0 of pHi and 60 min for EC and 100 mA/cm2, pH of 7.2, 5.0 g/L Thus, under the optimum processes conditions, a chemical addition for
Na2SO4 and 180 min for EO. Satisfactory removal efficiencies for tur- pH arrangement was not necessary in either system. However, the ad-
bidity (98.7%), TOC (86.4%) and COD (85.3%) were provided via a dition of electrolyte during EO was necessary due to causing a decrease
combination of EC and EO under the optimum process conditions. of electrical consumption.
Treated wastewater by EC and EO had 347 mg/L of COD, 86.4 mg/L of According to toxicity test results, EC with Al electrode has an im-
TOC, 10 NTU of turbidity and 9.2 of pH. It must be noted that EC only portant potential for decreasing acute toxicity in wastewater. In addi-
provided an effective removal in turbidity. On the other hand, a large tion, the forming of toxic species, one of the disadvantages of the EO

Fig. 5. Effect of pH and time on (a) TOC and (b) COD removal (CD: 100 mA/cm2, Ce: 5.0 g Na2SO4/L, initial TOC: 1006 mg/L, initial COD: 2358 mg/L, and initial turbidity: 783 NTU).

190
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

Table 2 Ghernaout, D., Ghernaout, B., Saiba, A., Boucherit, A., Kellil, A., 2009. Removal of humic
The inhibition ratios for raw and treated wastewater at different dilution levels. acids by continuous electromagnetic treatment followed by electrocoagulation in
batch using aluminium electrodes. Desalination 238, 295–308. http://dx.doi.org/10.
1016/j.desal.2008.04.001.
Wastewater Dilution Ratios (%) Inhibition Ratios (Inh %)
Gotsi, M., Kalogerakis, N., Psillakis, E., Samaras, P., Mantzavinos, D., 2005.
Electrochemical oxidation of olive oil mill wastewaters. Water Res. 39, 4177–4187.
Raw EC effluent EO effluent http://dx.doi.org/10.1016/j.watres.2005.07.037.
Hamdan, S.S., El-Naas, M.H., 2014. Characterization of the removal of Chromium(VI)
50.00 66.06 7.33 11.35 from groundwater by electrocoagulation. J. Ind. Eng. Chem. 20, 2775–2781. http://
33.33 63.51 −3.47 −10.03 dx.doi.org/10.1016/j.jiec.2013.11.006.
25.00 62.46 −10.47 −16.90 International Standard Method 21338, 2010. Water quality — Kinetic determination of
16.67 55.01 −13.51 −15.80 the inhibitory effects of sediment, other solids and coloured samples on the light
12.50 47.46 −17.84 −16.06 emission of Vibrio fischeri (kinetic luminescent bacteria test).
8.33 23.22 −9.98 −11.34 Jaafarzadeh, N., Omidinasab, M., Ghanbari, F., 2016. Combined electrocoagulation and
6.25 12.35 −8.69 −10.93 UV-based sulfate radical oxidation processes for treatment of pulp and paper was-
tewater. Process Saf. Environ. Prot. 102, 462–472. http://dx.doi.org/10.1016/j.psep.
2016.04.019.
Jalife-Jacobo, H., Feria-Reyes, R., Serrano-Torres, O., Gutiérrez-Granados, S., Peralta-
process, was not observed or the concentration of the toxic compound Hernández, J.M., 2015. Diazo dye Congo Red degradation using a Boron-doped
was low and the bacteria were using it as nutrient. It meant that the diamond anode: an experimental study on the effect of supporting electrolytes. J.
usage of Na2SO4 was a good choice in terms of toxicity test results. This Hazard. Mater. 319, 78–83. http://dx.doi.org/10.1016/j.jhazmat.2016.02.056.
Kamali, M., Khodaparast, Z., 2015. Review on recent developments on pulp and paper
study demonstrated that the combination of EC and EO processes, mill wastewater treatment. Ecotoxicol. Environ. Saf. 114, 326–342. http://dx.doi.
which have a great deal of advantages such as easy operable equipment, org/10.1016/j.ecoenv.2014.05.005.
short operation time, low requirement of chemicals, is a vigorous al- Kobya, M., Demirbas, E., Ozyonar, F., Sirtbas, G., Gengec, E., 2016. Treatments of alkaline
non-cyanide, alkaline cyanide and acidic zinc electroplating wastewaters by elec-
ternative for the treatment of wastewater with a high turbidity, organic trocoagulation. Process Saf. Environ. Prot. 105, 373–385. http://dx.doi.org/10.
content and toxicity. 1016/j.psep.2016.11.020.
Kobya, M., Gengec, E., 2012. Decolourization of melanoidins by a electrocoagulation
process using aluminium electrodes. Environ. Technol. 1–10. http://dx.doi.org/10.
Conflict of interest 1080/09593330.2012.671371.
Kobya, M., Gengec, E., Sensoy, M.T., Demirbas, E., 2014. Treatment of textile dyeing
The author declares that there are no conflicts of interest and affirm wastewater by electrocoagulation using Fe and Al electrodes: optimisation of oper-
ating parameters using central composite design. Color. Technol. 130, 226–235.
that this paper consists of original and unpublished work.
http://dx.doi.org/10.1111/cote.12090.
Koparal, A.S., Yavuz, Y., Gürel, C., Öǧütveren, Ü.B., 2007. Electrochemical degradation
References and toxicity reduction of C.I. Basic Red 29 solution and textile wastewater by using
diamond anode. J. Hazard. Mater. 145, 100–108. http://dx.doi.org/10.1016/j.
jhazmat.2006.10.090.
Al-Mutairi, N.Z., 2006. Coagulant toxicity and effectiveness in a slaughterhouse waste- Kourdali, S., Badis, A., Boucherit, A., 2014. Degradation of direct yellow 9 by electro-
water treatment plant. Ecotoxicol. Environ. Saf. 65, 74–83. http://dx.doi.org/10. Fenton: process study and optimization and, monitoring of treated water toxicity
1016/j.ecoenv.2005.05.013. using catalase. Ecotoxicol. Environ. Saf. 110, 110–120. http://dx.doi.org/10.1016/j.
APHA, 1998. Standard Methods for the Examination of Water and Wastewater, 18th ed. ecoenv.2014.08.023.
American Public Health Association, Washington, DC. Kováts, N., Refaey, M., Varanka, B., Reich, K., Ferincz, Á., Ács, A., 2012. Comparison of
Birjandi, N., Younesi, H., Bahramifar, N., 2014. Treatment of wastewater effluents from conventional and Vibrio fischeri bioassays for the assessment of municipal waste-
paper-recycling plants by coagulation process and optimization of treatment condi- water toxicity. Environ. Eng. Manag. J. 11, 2073–2076.
tions with response surface methodology. Appl. Water Sci. 1, 1–10. http://dx.doi. Kumar, S., Saha, T., Sharma, S., 2015. Treatment of pulp and paper mill effluents using
org/10.1007/s13201-014-0231-5. novel biodegradable polymeric flocculants based on anionic polysaccharides : a new
Bottesi, L., 2016. Electrochemical oxidation of salicylic acid using BDD as electrode way to treat the waste water. Int. Res. J. Eng. Technol. 1415–1428.
material. Nor. Univ. Sci. Technol. Lissens, G., Pieters, J., Verhaege, M., Pinoy, L., Verstraete, W., 2003. Electrochemical
Canizares, P., Garcia-Gomez, J., Lobato, J., Rodrigo, M.A., 2004a. Modeling of waste- degradation of surfactants by intermediates of water discharge at carbon-based
water electro-oxidation processes Part II. Application to active electrodes. Ind. Eng. electrodes. Electrochim. Acta 48, 1655–1663. http://dx.doi.org/10.1016/S0013-
Chem. Res. 43, 1923–1931. http://dx.doi.org/10.1021/ie0341303. 4686(03)00084-7.
Canizares, P., Sez, C., Lobato, J., Al, 2004b. Electrochemical treatment of 4-nitrophenol Makwana, A.R., Ahammed, M.M., 2016. Continuous electrocoagulation process for the
containing aqueous wastes using boron-doped diamond anodes. Ind. Eng. Chem. Res. post-treatment of anaerobically treated municipal wastewater. Process Saf. Environ.
43, 1944–1951. Prot. 102, 724–733. http://dx.doi.org/10.1016/j.psep.2016.06.005.
Chatzisymeon, E., Xekoukoulotakis, N.P., Coz, A., Kalogerakis, N., Mantzavinos, D., 2006. Mansour, L. Ben, Kesentini, I., 2008. Treatment of effluents from cardboard industry by
Electrochemical treatment of textile dyes and dyehouse effluents. J. Hazard. Mater. coagulation-electroflotation. J. Hazard. Mater. 153, 1067–1070. http://dx.doi.org/
137, 998–1007. http://dx.doi.org/10.1016/j.jhazmat.2006.03.032. 10.1016/j.jhazmat.2007.09.061.
Choi, J.Y., Lee, Y.-J., Shin, J., Yang, J.-W., 2010. Anodic oxidation of 1,4-dioxane on Marco Antonio Quiroz, A., Ferro, S., Martínez-Huitle, C.A., Vong, Y.M., 2006. Boron
boron-doped diamond electrodes for wastewater treatment. J. Hazard. Mater. 179, doped diamond electrode for the wastewater treatment. J. Braz. Chem. Soc. 17,
762–768. http://dx.doi.org/10.1016/j.jhazmat.2010.03.067. 227–236.
Comninellis, C., Pulgarin, C., 1991. Anodic oxidation of phenol for waste water treatment. Martínez-Huitle, C.A., Ferro, S., 2006. Electrochemical oxidation of organic pollutants for
J. Appl. Electrochem. 21, 703–708. http://dx.doi.org/10.1007/BF01034049. the wastewater treatment: direct and indirect processes. Chem. Soc. Rev. 35,
Dastyar, W., Amani, T., Elyasi, S., 2015. Investigation of affecting parameters on treating 1324–1340. http://dx.doi.org/10.1039/b517632h.
high-strength compost leachate in a hybrid EGSB and fixed-bed reactor followed by Mella, B., Glanert, A.C., Gutterres, M., 2015. Removal of chromium from tanning was-
electrocoagulation – flotation process. Process Saf. Environ. Prot. 95, 1–11. http://dx. tewater and its reuse. Process Saf. Environ. Prot. 95, 195–201. http://dx.doi.org/10.
doi.org/10.1016/j.psep.2015.01.012. 1016/j.psep.2015.03.007.
David, C., Arivazhagan, M., Tuvakara, F., 2015. Decolorization of distillery spent wash Metcalf, E., Eddy, H., 2004. Wastewater engineering: treatment and reuse. In:
effluent by electro oxidation (EC and EF) and Fenton processes: a comparative study. Techobanoglous, G., Burton, F.L., Stensel, H.D. (Eds.), Wastewater Engineering,
Ecotoxicol. Environ. Saf. 121, 142–148. http://dx.doi.org/10.1016/j.ecoenv.2015. Treatment, Disposal and Reuse, 4th edition. Tata McGraw-Hill Publishing Company
04.038. Limited. http://dx.doi.org/10.1016/0309-1708(80)90067-6.
Elabbas, S., Ouazzani, N., Mandi, L., Berrekhis, F., Perdicakis, M., Pontvianne, S., Pons, Özcan, A., Şahin, Y., Koparal, A.S., Oturan, M.A., 2008. Propham mineralization in
M.N., Lapicque, F., Leclerc, J.P., 2015. Treatment of highly concentrated tannery aqueous medium by anodic oxidation using boron-doped diamond anode: influence
wastewater using electrocoagulation: influence of the quality of aluminium used for of experimental parameters on degradation kinetics and mineralization efficiency.
the electrode. J. Hazard. Mater. 319, 69–77. http://dx.doi.org/10.1016/j.jhazmat. Water Res. 42, 2889–2898. http://dx.doi.org/10.1016/j.watres.2008.02.027.
2015.12.067. Panizza, M., Cerisola, G., 2001. Removal of organic pollutants from industrial wastewater
Ezechi, E.H., Isa, M.H., Rahman, S., Kutty, M., Yaqub, A., 2014. Boron removal from by electrogenerated Fenton's reagent. Water Res. 35, 3987–3992. http://dx.doi.org/
produced water using electrocoagulation. Process Saf. Environ. Prot. 92, 509–514. 10.1016/S0043-1354(01)00135-X.
http://dx.doi.org/10.1016/j.psep.2014.08.003. Pokhrel, D., Viraraghavan, T., 2004. Treatment of pulp and paper mill wastewater - A
Gao, W., Chen, J., Guan, X., Jin, R., Zhang, F., Guan, N., 2004. Catalytic reduction of review. Sci. Total Environ. 333, 37–58. http://dx.doi.org/10.1016/j.scitotenv.2004.
nitrite ions in drinking water over Pd–Cu/TiO2 bimetallic catalyst. Catal. Today 95, 05.017.
333–339. http://dx.doi.org/10.1016/j.cattod.2004.06.013. Renault, F., Sancey, B., Charles, J., Morin-Crini, N., Badot, P.M., Winterton, P., Crini, G.,
Gengec, E., Kobya, M., Demirbas, E., Akyol, A., Oktor, K., 2012. Optimization of baker's 2009. Chitosan flocculation of cardboard-mill secondary biological wastewater.
yeast wastewater using response surface methodology by electrocoagulation. Chem. Eng. J. 155, 775–783. http://dx.doi.org/10.1016/j.cej.2009.09.023.
Desalination 286, 200–209. http://dx.doi.org/10.1016/j.desal.2011.11.023. Särkkä, H., Bhatnagar, A., Sillanpää, M., 2015. Recent developments of electro-oxidation

191
E. Gengec Ecotoxicology and Environmental Safety 145 (2017) 184–192

in water treatment – a review. J. Electroanal. Chem. 754, 46–56. http://dx.doi.org/ Vasudevan, S., 2012. Effects of alternating current (AC) and direct current (DC) in elec-
10.1016/j.jelechem.2015.06.016. trocoagulation process for the removal of iron from water. Can. J. Chem. Eng. 90,
Scialdone, O., 2009. Electrochemical oxidation of organic pollutants in water at metal 1160–1169. http://dx.doi.org/10.1002/cjce.20625.
oxide electrodes: a simple theoretical model including direct and indirect oxidation Wu, W., Huang, Z.-H., Lim, T.-T., 2014. Recent development of mixed metal oxide anodes
processes at the anodic surface. Electrochim. Acta 54, 6140–6147. http://dx.doi.org/ for electrochemical oxidation of organic pollutants in water. Appl. Catal. A Gen. 480,
10.1016/j.electacta.2009.05.066. 58–78. http://dx.doi.org/10.1016/j.apcata.2014.04.035.
Scialdone, O., Galia, A., Guarisco, C., Randazzo, S., Filardo, G., 2008. Electrochemical Yilmaz, A.E., Boncukcuoǧlu, R., Kocakerim, M.M., 2007. A quantitative comparison be-
incineration of oxalic acid at boron doped diamond anodes: role of operative para- tween electrocoagulation and chemical coagulation for boron removal from boron-
meters. Electrochim. Acta 53, 2095–2108. http://dx.doi.org/10.1016/j.electacta. containing solution. J. Hazard. Mater. 149, 475–481. http://dx.doi.org/10.1016/j.
2007.09.007. jhazmat.2007.04.018.
Sires, I., Brillas, E., Oturan, M.A., Rodrigo, M.A., Panizza, M., 2014. Electrochemical Zaidi, S., Chaabane, T., Sivasankar, V., Darchen, A., Maachi, R., Msagati, T.A.M.,
advanced oxidation processes: today and tomorrow. A Rev. Environ. Sci. Pollut. Res. Prabhakaran, M., 2016. Performance efficiency of electro-coagulation coupled
21, 8336–8367. http://dx.doi.org/10.1007/s11356-014-2783-1. electro-flotation process (EC-EF) versus adsorption process in doxycycline removal
Terrazas, E., Vázquez, A., Briones, R., Lázaro, I., Rodríguez, I., 2010. EC treatment for from. Process Saf. Environ. Prot. 102, 450–461. http://dx.doi.org/10.1016/j.psep.
reuse of tissue paper wastewater: aspects that affect energy consumption. J. Hazard. 2016.04.013.
Mater. 181, 809–816. http://dx.doi.org/10.1016/j.jhazmat.2010.05.086.

192

You might also like