You are on page 1of 27

18

CHAPTER 2

REVIEW OF LITERATURE

The textile effluent shows a large extent of variation in the


characteristics of wastewater from plant-to-plant and sample-to-sample
because of the difference in the process adopted and dyestuffs employed. The
removal of colour from textile industry and dyestuff manufacturing industry
wastewaters represents a major environmental concern since most of the
commercial dyes are poor in biodegradability with high COD/BOD ratio.
Several physical, biological and chemical processes like adsorption,
coagulation, flocculation, membrane filtration, ozonation, electrochemical,
bacterial and advanced oxidation processes is known to decolorize the textile
industrial effluents (Kannan & Sundaram, 2001; Rai et al, 2005; Solmaz et al,
2009; Wojnarovits & Takacs, 2008; Ince & Tezcanh, 1999; Chaudhari et al,
2011). On one hand, physicochemical treatments transfer pollutants in the
effluents from one phase to other (Erswell et al., 1988) whereas the biological
treatment takes much time for degradation and even fails due to the
complexity of the pollutants (Aksu 2005).

Hence a choice of suitable treatment method is required, which can


effectively treat pollutants without expensive chemical addition. Among the
different treatment processes, advanced oxidation processes and
electrocoagulation process are often the preferred choice for the treatment of
textile wastewater due to its adaptability for higher and lower pollution load,
easy handling and cost effectiveness. However, literatures pertaining to
general description of treatment processes with its operational variables have
been presented here with a view to highlight the gap in those researchers and
to reiterate the rationale of the present research.
19

2.1 ADVANCED OXIDATION PROCESS (AOP)

2.1.1 Photo Assisted Chemical Oxidation (PACO)

Advanced Oxidation Process (AOP) has gained much attention due


to its high level of decolourization and mineralization of wide range of dyes
to harmless products. It is an oxidation technique in which hydroxyl radicals
are produced in a different way for the possible degradation compounds of
high chemical stability and low biodegradability. Hydroxyl radicals are
efficient to mineralize the organic compounds completely into carbon dioxide
and water.

The basic mechanism of advanced oxidation process is the


production of OHᵒ radicals which are capable of destroying complex organic
compounds as shown in Table 2.1 (a) and (b) (Al-Kdasi et al. 2004;Vogelpohl
and Kim 2004). The OHᵒ radicals are found to have an oxidizing power of
2.05 V and to oxidize substances more quickly than conventional oxidants
(Gogate & Pandit 2003). The OHᵒ radicals formed are considered as reactive
electrophiles due to their preference to electrons and hence they react rapidly
towards electron rich organic compounds.

2.1.2 Basics of PACO process

Photo Assisted Chemical Oxidation (PACO), a simple method of


production of OHᵒ radical, is the photolysis of hydrogen peroxide by the UV
light source – UV/H2O2. Photolysis of H2O2 yields hydroxyl radicals by a
direct process with yield of two radicals formed per photon absorbed by 254
nm (Baxendable and Willson, 1957). The highest hydroxyl radical yield is
obtained using short-wave ultraviolet radiations between 200–280 nm.

The molar extinction coefficient of H2O2 by 254 nm is 19.6 M-1cm-1


(Lay, 1989) which is very low. Comparatively, the molar extinction
coefficients of ozone, naphthalene and pentachlorophenol (PCP) are 3300,
20

3300 and 10,000 M-1cm-1, respectively. It indicates that to generate high level
of OHᵒ radicals in a solution needs strong photon absorbers and high
concentration of H2O2, but in PACO process, low pressure mercury vapor
lamps with a 254 nm peak emission are used to generate OHᵒ radicals by
direct photolysis.

The organic substances are first attacked by the generated OHᵒ


radicals, followed by hydrogen abstraction and then the electron transfer as
follows (Stasinakis 2008).

Table 2.1 Oxidizing species and oxidizable compounds (Ghaly et al, 2014)

(a) Relative oxidizing power of some oxidizing species and (b) Oxidizable
compounds by OHᵒ radicals

(a) (b)
Oxidizing Oxidizing
Group Components
Species Power
Hydroxy Formic, gluconic, lactic, malic, propionic,
2.05 Acids
radical tartaric
Benzyl, tert-butyl, ethanol, ethylene
Atomic
1.78 Alcohols glycol, glycerol,isopropanol, methanol,
oxygen
propenediol
Acetaldehyde, benzaldehyde,
Ozone 1.52 Aldehydes formaldehyde, glyoxal, isobutryaldehyde,
trichloroacetaldehyde
Benzene, chlorobenzene, chlorophenol,
Hydrogen creosote, dichlorophenol, hydroquinone, p-
1.31 Aromatics
peroxide nitrophenol, phenol, toluene,
trichlorophenol, xylene, trinitrotoluene
Aniline, cyclic amines, diethylamine,
dimethylformamide,
Permanganate 1.24 Amines
ethylenediaminetetraaceticacid,
propanediamine, n-propylamine
Chlorine 1.00 Esters Tetrahydrofuran
21

H2O2 + UV 2OHᵒ (2.1)

R + OHᵒ ROH (2.2)

RH + OHᵒ Rᵒ + H2O (2.3)

Rn + OHᵒ Rn-1 + OHᵒ- (2.4)

R – reactive organic compound

The oxidation rate of PACO process depends upon the


concentration of (a) target pollutant, (b) radicals generated and (c) oxygen
present in the surrounding.

2.1.3 Textile dyes used in PACO process

Hydroxyl radicals formed by the PACO process can react with


many numbers of complex organic dyes and undergo degradation rapidly
(Ray et al. 2006; Poyatos et al. 2010). Initially, OHᵒ radicals easily attack the
chromophoric group of the dye molecules that are ready to undergo oxidation.
The rate of decolourization efficiency of the dye molecules may vary
depending upon the nature of the dye molecules. High intensity colour
producing dyes undergo decolourization very slowly than the low intensity
colour causing dyes.

Table 2.2 shows the treatability of PACO process on different types


of dyes such as acidic, basic, direct and reactive dyes. It is evident from the
table that percentage of the colour removal efficiency of the differnt types of
dyes vary depends upon the intensity of UV radiation and concentration of the
dye solutions. An internal optical density of dye solution is increased due to
the increase in dye concentration and the dye solution becomes more and
more impermeable to UV irradiation. As a consequence the rate of photolysis
22

of hydrogen peroxide is reduced and hence the formation of OHᵒ radicals.


Therefore it can be inferred that for more concentrated dye wastewater as well
as effluents exhibiting higher optical density are less suitable for photo
assisted AOPs.

Table 2.2 PACO used for treatment of different types of dyes

UV % of
S. Dye and
Structure lamp colour Reference
No. Class
power removal
Direct red
2 x 15 Mahmoodi et
1 80 (tetra 100
W al. (2005)
azo)

Acid blue 7W Shu et al.


2 96.66
113 (2005)

Acid black 1400 Shu et al.


3 90
1 W (2004)

Muruganandh
Reactive am and
4 --- 93.47
orange 4 Swaminathan,
(2004)

Basturk and
Reactive 100
5 99 Karatas,
blue 181 mW
(2015)

Beikmohamm
Basic
6 50 W 98 adi et al.
yellow-28
(2016)
23

Table 2.2 (Continued)

UV % of
S. Dye and
Structure lamp colour Reference
No. Class
power removal

Reactive
7 16 W 99%
black 5

Bali et al.
Direct (2004)
8 16 W 98%
yellow 12

Direct red
9 16 W 70%
28

Reactive Zuorro et al.


10 6W 98.99%
green 19 (2013)

Most of the researchers (Mariana et al. 2002; Rosario et al. 2002)


have reported that sufficient reaction time for the removal of dye using PACO
process is subjective. Some of the researchers have documented complete
destruction of reactive and azo dyes in 30 to 90 minutes (Georgiou et al. 2002;
Tanja et al. 2003), while Perkowski & Kos (2003) have recorded 99% of colour
removal efficiency of dye house wastewater after 2 hrs of reaction time.
24

2.1.4 Effect of oxidant concentration in PACO process

Galindo & Kalt (1998) have reported that the effective colour
removal efficiency of dyes solution by PACO process was takes place in an
acid medium (pH≈ 3-4), but the process fails at alkaline pH due to the fact
that decomposition of hydrogen peroxide lead to oxygen and water.

Table 2.3 Effect of oxidant concentration on treatment by PACO process

Optimized Percentage of
S.
Dye concentration of colour AOP Process Reference
No.
oxidant removal
Direct Red Mahmoodi et al.
1 300 mg/l 100% UV/TiO2/H2O2
80 (2005)
Reactive
2 24.5 mmol/l 99.58% UV/H2O2
Yellow 84
Reactive Neamtu et al.
3 24.5 mmol/l 99.53% UV/H2O2
Red120 (2002)
Reactive
4 24.5 mmol/l 99.82% UV/H2O2
Black 5
Acid Blue
5 46.53 mmol/l 96.66% UV/H2O2 Shu et al. (2005)
113
6 Acid Black 1 21.24 mmol/l 90% UV/H2O2 Shu et al. (2004)
Muruganandham &
Reactive
7 10 mmol 93.47% UV/H2O2 Swaminathan
Orange 4
(2004)
Beikmohammadi et
8 Yellow 28 25 mg/l 97.47% UV/H2O2
al. (2016)
Reactive Basturk & Karatas,
9 500 mg/l 99% UV/H2O2
Blue 181 (2015)
Reactive
10 25 mmol 99% UV/H2O2
Black 5
Direct
11 25 mmol 98% UV/H2O2 Bali et al. (2004)
Yellow 12
Direct Red
12 25 mmol 70% UV/H2O2
28
Reactive
13 30 mmol 98.99% UV/H2O2 Zuorro et al. 2013
Green 19
25

Furthermore, the scavenging effects of carbonate at higher pH


values are more sensitive to PACO process. Temperature has been found to
have no significant effect on colour removal efficiency.

Arslan et al. (1999) and Ince (1999) have reported that the colour
removal efficiency of textile dyes by PACO process gets increased on
increasing the effective dose hydrogen peroxide up to a “critical” level as
shown in Table 2.3. It is evident from the Table 2.3 that the critical level for
the different dyes vary depends upon many factors such as complexity of the
dye structure, concentration of the dye solution and intensity of the UV light
and so on. Higher concentration of peroxide acts as a radical scavenger while,
lower concentration of peroxide generate insufficient quantity of hydroxyl
radicals (OHᵒ) consumed by dyes and dye auxiliaries and this leads to slower
rate of oxidation of organic dyes.

Arslan & Isil (2001 & 2002) have shown that the treatment of
textile wastewater is not effective in PACO oxidation unless a preliminary
treatment is introduced to produce sufficient OHᵒ to observe a significant
colour and COD reduction. Many researchers have documented that even at
50 mmol of H2O2 concentration, raw textile effluent does not show any
significant degradation and COD removal is negligible.

2.1.5 Energy Consumption in PACO Process

The selection of methodology for the treatment of industrial


effluent involves many important factors such as inducing economics, effluent
quality, operation like maintenance, control and safety and robustness i.e.
flexibility of the process. Photodegradation of aqueous dye solution is an
electric energy intensive process. The UV lamp is completely submerged into
the reactor, the entire irradiation emitted by the lamp is absorbed by the dye
26

wastewater and thus the electrical energy dose at each treatment level is
calculated as the energy consumption (EE/O) of the lamp. The EE/O is
defined as the number of kWh of electrical energy required to reduce the
concentration of a pollutant in a unit volume of contaminated water. Higher
EE/O value indicates that it is more difficult and more costly to remove the
pollutant relative to those with lower EE/O values.

Table 2.4 Effect of Energy consumption on treatment by PACO process

% Colour
S. Energy
Process Dye Optimized condition removal Reference
No. Consumption
efficiency
For US = 633.79 and
Malachite [Dye] 0 = 5 mgL−1 Behnajady et al.
1 US/UV/H2O2 for US/UV/H2O2 = 100%
Green [H2O2]0=400 mgL−1 (2009)
19.98 kWhm−3
2 Acid Orange 7 1.341 kWhm-3
3 Acid Orange 8 -5
[Dye]0 = 9 x 10 M 1.561 kWhm-3
Daneshvar et al.
UV/H2O2 Acid Orange [H2O2]0 = 4.98 x10-2 >95%
4 M 2.177 kWhm-3 (2005)
52
5 Acid Blue 74 1.165 kWhm-3
[Dye] 0 =10ppm Daneshvar et al.
6 UV/H2O2 Rhodamine B 26.6 kWhm-3 72%
[H2O2]0 =450ppm, (2008)
[Dye]0 =30ppm Aleboyeh, et al.
7 UV/H2O2 Acid Orange 7 2.696 kWhm−3 100%
[H2O2]0 = 285ppm (2008)
[Dye] 0 =20ppm
UV/ZnO/ Acid Orange Daneshvar et al.
8 [ZnO]0=160ppm 172 kWh/m3 90%
H2O2 7 (2007)
[H2O2]0=10 mmol
Reactive black [Dye]0=30mg/L Marandi et al.
9 UV/H2O2 59.25 kWh/m3 84.34%
B [H2O2]0=500mg/L (2013)
[Dye]0 =10ppm
10 Basic Blue 3 7.67 kWh/m3 95.03 %
[H2O2]0 = 1.2 g/l Kasiri and
UV/H2O2
Acid [Dye]0 =10ppm 3
Khataee, (2011)
11 5.76 kWh/m 98.16%
Green 25 [H2O2]0 = 1.2 g/l

The connection between electric energy and overall reaction


kinetics of AOPs is obvious. The electrical energy required to eliminate
specific amount of pollutant is directly proportional to electric power of the
bath used and inversely proportional to the treated volume of water and the
observed overall rate constant of the process. Table 2.2 exhibited the
application of different power of the UV lamp used during the process and
27

hence the energy consumed for the treatment process. Table 2.4 consolidated
the energy consumption of the treatment of the different textile dyes with its
optimal conditions. The energy consumption during the treatment process
may be increased due to the nature of the dye auxiliaries used at the time of
dyeing of fabric materials. Since some of the dye auxiliaries may be act as
radical scavengers and reduced the formation of the oxidizing species.

2.1.6 Industrial application of PACO process

Various types of industries deliver different characteristics of


effluent. Treatment processes availed for the effective removal of pollutants
from the effluent have their own mechanism. Textile industrial effluent is one
among typical pollutant characterized with high colour, COD, BOD and TDS.
This could be mainly due to the presence of bulkier/complex organic dyes in
the wastewater stream. PACO is promising technique that has gained much
attention due to its high level of decolorization and mineralization of wide
range of dyes into harmless products. Some of the advantages of PACO
process are: (a) it does not form any chemical sludge, (b) it effectively
removes phenolic compounds (Stasinakis 2008), chlorinated compounds and
chlorophenols (Munter, 2001), from effluents, (c) UV acts as a disinfectant
and (d) it is effective in the degradation of TOC, 80-82% removal efficiency
is obtained after 1-2 hours of treatment (Alaton et al. 2002, Amin et al. 2008).

The experiments performed by many researchers on different


industrial effluents are reported in Table 2.5 and it clearly shows that the
optimum conditions varied depending upon the nature of the industrial
effluent for the effective treatment of the process.
28

Table 2.5 PACO used for the treatment of different industrial effluents

% of
S. UV Light % of COD
Industrial Effluent colour Reference
No. intensity removal
removal
Industrial
wastewater
Mahvi et al.
1 containing 6.3 mW/cm2 --- 69.7%
(2012)
high amounts of
refractory organics
Oil recovery Dincer et al.
2 12 W --- 55%
industry wastewater (2008)
Peerez et al.
3 Coffee wastewater UV/H2O2/O3 --- 87%
(2007)
Bleach effluent of Subashini &
4 pulp and paper 400 W 52% 59% Kanmani
industry (2013)
Abdel-Aal et
5 Tannery wastewater 150 W --- 85%
al. (2015)
Pharmaceutical Azizi et al.
6 50 W --- 87.5%
wastewater (2016)
Pieczykolan
7 Landfill leachate 15 W --- 74.6%
et al. (2012)
Sahunin et al.
8 Textile wastewater 60 W 52% 90%
(2006)
15 fold diluted Arslan Alaton
9 reactive 25 W 100% --- et al. (2002)
dyebath effluent
Nagel-
Textile industry
10 250 W 73% 77% Hassemer et
effluents
al. (2011)

2.2 ELECTROCOAGULATION

2.2.1 Basics of Electrocoagulation

Electrocoagulation is a simple electrochemical process in which,


sacrificial anode generates a metal ion that follows a different mechanism to
produce hydroxide or polyhydroxide in the bulk of the solution, strongly
adsorbs dispersed compound molecules and results in coagulation (Heidmann
29

et al. 2008). Hydrogen produced in cathodes can help mass particles float and
come out of water (Ghosh et al. 2008a; Yahiaoui et al. 2011). Daneshvar et al.
(2006) have proposed two mechanisms for the production of the metal
hydroxide during electrocoagulation process.

Mechanism 1:

At the anode

4Fe(s) 4 Fe2+(aq) + 8 e- (2.5)

At the bulk of the solution

4Fe2+(aq) + 10H2O(l) + O2(g) 4Fe(OH)3(s) + 8H+(aq) (2.6)

At the cathode

8H+(aq) + 8e- 4H2(g) (2.7)

Overall reaction

4Fe(s) + 10H2O(l) + O2(g) 4Fe(OH)3(s) + 4H2(g) (2.8)

nFe(OH)3 Fen(OH)3n(s) (2.9)

Mechanism 2:

At the anode

Fe(s) 4 Fe2+(aq) + 2 e- (2.10)

At the bulk of the solution

4Fe2+(aq) + 2OH-(aq) Fe(OH)2(s) (2.11)


30

At the cathode

2H2O(aq) + 2e- 2OH-(aq) + H2(g) (2.12)

Overall reaction

Fe(s) + 2H2O(l) Fe(OH)2(s) + H2(g) (2.13)

nFe(OH)2 Fen(OH)2n(s) (2.14)

2.2.2 Textile Dyes used in Electrocoagulation

Decolourization of dye solution by electrocoagulation has become


an attractive method in the recent years. The electronic transition between the
molecular orbitals determines the intensity of the colour of the dye.
Understanding the basic structure and its complexity of the dye is much
important before going into the experimentation. In general, all the dye
molecules contain chromophores, delocalized electron systems with
conjugated double bonds and auxochromes, electron-withdrawing or electron-
donating substituents that cause or intensify the color of the chromophore by
altering the overall energy of the electron system. For example, at least one
nitrogen-nitrogen (N=N) double bond is present in azo dyes, but many
different structures are also possible. Two, three or more azo groups present
in some of the dye molecules are called monoazo, diazo or polyazo dyes.
These azo groups are connected with the side groups such as benzene,
naphthalene or aromatic compounds that are necessary for bringing the colour
and intensity to the dye molecules.

Decolorization efficiency of the dye solutions indicates the


degradation of chromophores of the organic dye molecules and thus decreases
its conjugation in dye structure with fragmentation of the organic dye into
simple molecules that takes place more effectively.
31

Many researchers have registered their findings on the colour


removal efficiency of textile dyes by EC, i.e., disperse and reactive dyes (Kim
et al. 2002), vat dyes (Roessler et al. 2003) acid orange 7 (Fernandes et al.
2004), acid green dye (Ghalwa et al. 2005), reactive black 5 (Sengil and
Ozacar, 2009), reactive red 141 (Zidane et al. 2008), acid blue 9 (Khataee et
al. 2009), acid red 14 (Aleboyeh et al. 2008a), acid Orange II (Xiong et al.
2001), bomaplex red CR-L (Yildiz, 2008), and the mixture of disperse dyes 2-
naphthoic and 2-naphthol (Essadki et al. 2008; Merzouk et al. 2009).

Table 2.6 Electrocoagulation used for treatment of different types of dyes

S. Electrode & % of
Dye and Class Structure Reference
No. Exp.Conditions removal
SS-SS, Fe-Fe
(pH:7, CD: 110
Reactive A/m2 for SS, 130
Yuksel et
1 Orange 84 and A/m2 for Fe, cond. 66, 76
al. (2013)
double azo 3000 lS/cm, initial
dye conc. 100
mg/L)

Al-Al (pH = 5, CD:


40 A/m2, NaCl: 2
Acid Black 52 Pajootan et
2 g/L and 99
and Single azo al. (2012)
initial dye conc.
200 mg/L).)

Al-Al (pH
6; NaCl = 2 g/L;
Direct red 81 Aoudj et
3 initial dye conc. 50 98
and double azo al. (2010)
mg/L; electrolysis
time 60 min.)

Fe-Al, Al-Fe (pH


Crystal violet 9; CD: 28Am−2; Durango-
4 and Na2SO4;1420mgL−1 100 Usugaa et
triarylmethane initial conc. 54 al. (2010)
mgL−1)
32

Table 2.6 (Continued)


S. Electrode & % of
Dye and Class Structure Reference
No. Exp.Conditions removal
Fe-Fe (pH, 6;
CD:10 mA/cm2;
Direct Red 23 Kobya et al.
5 initial conc.: 99
and mono azo (2010)
250 mg/L;
time, 10 min)
Fe-Fe (initial
Acid yellow 36 conc=50 ppm, Kashefialasl
6 83
and mono azo NaCl =10 g/l, et al. (2006)
Time = 10 min,)

Reactive SS-SS (pH 11.5


Kabdasli et
7 yellow 145 and CD: 22mA/cm2 100
al. (2009)
mono azo NaCl 40g/L)

Al-Al (CD=
0.25 mA/cm2
Bomaplex red initial conc = Yildiz
8 99
CR-L 600 mg/L, pH = (2008)
5.0,
CaCl2 and)

Al-Al (CD:
100 A/m2,
initial Conc.:
Levafix orange Kobya et al.
9 250 mg/L, time: 95
and mono azo (2006)
15 min,
conductivity:
750 S/cm)

Fe-Fe (CD =
4.575mAcm−2;
time = 5 min; Sengil and
Reactive Black
10 initial conc. 98 Ozacar
-5 and diazo
=100mgL−1; (2009)
T=298K;
NaCl=3gL−1)

Table 2.6 represents different classes of aqueous dyes solution


treated by EC process using different types of electrode materials with the
removal efficiency. Each dye shows complex structures with different
characteristics and ensures that the adaptability of the treatment process. The
33

dye removal efficiency varies depending on the nature of the dye and the
electrode materials used.

2.2.3 Electrodes used in Electrocoagulation

Table 2.7 shows the performance of different electrodes on the


textile dye solution depending upon the nature of the sacrificial anode availed
for the treatment process, nature of the flocs formed and interaction of the
flocs with the organic dyes reflected in the percentage of dye degradation /
decolourization. The amount of flocs formed is directly proportional to the
dissolution of the sacrificial electrode during the EC process. According to
Faraday’s law, the dissolution of the anode is directly proportional to the
current density / applied voltage. The stabilization of the flocs formed and
interaction of flocs with dyes also depend upon the pH of the medium.

A number of researches have been done on different electrodes


such as iron, aluminium and Stainless Steel (SS) for the effective removal of
aqueous dye solution from the textile industries (Ghalwa et al, 2016). Linares-
Hernández et al. (2009) have reported higher colour removal efficiency with
aluminium electrodes, whereas iron shows effective COD removal efficiency
than aluminium from industrial wastewater. A combination of iron and
aluminium electrodes removes both colour (71%) and COD (69%) with
higher removal efficiency. Similar electrode combination shows better
performance in treating paper mill wastewaters (Katal et al. 2011). Both the
electrodes perform well in treating heavy metal removals like arsenic (Gomes
et al. 2007), copper, chromium and nickel (Akbal and Camci, 2011).
34

Table 2.7 Electrocoagulation with different electrodes for treatment of


textile dyes

% of Colour
S. Current Electrode &
Dye removal Reference
No. density Exp.Conditions
efficiency
Senthilkum
SS (pH 8 and CD =
1 Reactive red 120 50A/m2 98 ar et al.
50 A/m2)
(2010)
SS-SS, Fe-Fe (pH:7,
CD: 110 A/m2 for SS,
Reactive Orange Yuksel et
2 130 A/m2 66 & 76 130 A/m2 for Fe, cond.
84 3000 lS/cm, initial dye al. (2013)
conc. 100 mg/L)
SS-SS (pH 11.5 CD:
Reactive yellow Kabdasli et
3 220 A/m2 100 22mA/cm2 NaCl
145 40g/L) al. (2009)
Fe-Fe (CD = 4.575
mAcm−2; time = 5 Sengil and
4 Reactive Black -5 45.75 A/m2 98 min; initial conc. Ozacar
=100mgL−1; T=298K;
(2009)
NaCl=3gL−1)
Al-Al (pH
6; NaCl = 2 g/L; initial
Aoudj et al.
5 Direct red 81 18.75 98 dye conc. 50 mg/L;
electrolysis time 60 (2010)
min.)

Fe-Fe (initial conc=50 Kashefialas


2
6 Acid yellow 36 127.8 A/m 83 ppm, NaCl =10 g/l, l et al.
Time = 10 min,) (2006)
Al-Al (pH = 5, CD: 40
2 A/m2, NaCl: 2 g/L and Pajootan et
7 Acid Black 52 40 A/m 99 initial dye conc. 200 al. (2012)
mg/L).)
Fe-Al, Al-Fe (pH 9; Durango-
2 CD: 28Am−2;
8 Crystal violet 28 A/m 100 Na2SO4;1420mgL−1
Usugaa et
initial conc. 54 mgL−1) al. (2010)
Al-Al (CD= 0.25
Bomaplex red CR- 2 mA/cm2 initial conc = Yildiz
9 5 A/m 99 600 mg/L, pH = 5.0,
L (2008)
CaCl2 and)
Al-Al (CD:
100 A/m2, initial
Conc.: 250 mg/L, Kobya et
10 Levafix orange 355 A/m2 95
time: 15 min, al. (2006)
conductivity: 750
S/cm)
35

2.2.4 Energy and Electrode Consumption in Electrocoagulation

Coagulant (flocs) concentration produced by electrolysis on anodes


is directly proportional to the electric charge added per volume (coulombs per
litre). However, the total amount of coagulant dissolved also includes
chemical dissolution of the electrodes in low pH and the dissolution of
electrodes.

The increase in current density increases the applied voltage to the


EC reactor and hence electrode and electrical energy consumption also
increase. With the increasing current density, electrode consumption and
bubble generation get increased; consequently the amount of hydroxyl
polymers available for the adsorption of dye molecule is also increased. It is
well known that the amount of current density determines the electrode
consumption / coagulant production rate, and adjusts the rate and size of the
bubble production, and hence affects the growth of flocs (Mollah et al. 2004;
Chen, 2004).

Table 2.8 showcase the collection of literature on electrode and


energy consumption during the EC treatment process. Tezcan & Aytac (2013)
have noted that electrical energy consumption after 90 min of operation is
1.05x10-4, 1.45x10-4 and 2.66x10-4 kWh/mg of COD is removed at the current
densities of 20, 30 and 50 mA/cm2 respectively. The increase in current
density during the process reduces the treatment time and hence the cost of
electrical energy consumption would be reduced to 6.91x10-5 kWh/mg for 60
min EC at 20 mA/cm2, 6.63 x10-5 kWh/mg for 40 min EC at 30mA/cm2 and
9.39 x10-5 kWh/mg for 30 min EC at 50 mA/cm2. Rajkumar and Kim, (2006)
also have suggested that choosing an optimum current density by considering
a shorter reaction time with a maximum pollutant removal is essential for the
36

scaling up of the electrochemical process to reduce the operating costs


(Muthukumar et al. 2004; Rajkumar et al. 2003).

Table 2.8 Energy and Electrode consumption in Electrocoagulation


process during the treatment of different types of dyes

Electrode Energy
S.
Dye consumption consumption Electrode Reference
No.
(kg/m3) (kWh/m3)
Reactive Orange
1 0.060 2.02 SS
84 Yuksel et al.
Reactive Orange (2013)
2 0.190 2.80 Fe
84
Ghosh et al.
3 Crystal Violet 0.497 10.01 Al
(2008)
1.2 kg Fe/ kg 3.3 kWh/ kg Kobya et al.
4 Remazol 3B Fe
dye dye (2010)
Arslan-
Simulated reactive
5 --- 1.8 & 3.6 Al & SS Alaton et al.
dyebath effluent
(2009)
Synthetic Merzouk et
6 0.016 -- Al
wastewater al. (2011)
Egg processing Sridhar et al.
7 --- 16.24 Al
industrial effluent (2014)
Naje et al.
8 Textile wastewater 0.100 8.49 Al
(2015)
Kobya et al.
9 Textile wastewater 0.163 & 0.107 0.63 & 0.70 Fe & Al
(2007)
Textile dye
10 0.05 2.43 SS
wastewater Yuksel et al.
Textile dye (2013)
11 0.28 2.38 Fe
wastewater

2.2.5 Industrial Application of Electrocoagulation

Increased industrialization in the developing countries creates


harmful water pollution due to high level of industrial effects and sewage
effluent discharges. Depending upon the type of industries, characteristics of
the effluent vary significantly in terms of the presence of pollutant,
concentration of the pollutant, treatment and disposal techniques. The
37

industrial effluent treatment technique is selected on the basis of the nature,


concentration, volume and toxicity of the pollutant. Electrocoagulation is
attracted by many researchers due to its adaptability to treat various kinds of
industrial wastewater.

Table 2.9 Electrocoagulation used for the treatment of different


industrial effluents

% of
S. Industrial Current
pH Electrode COD Reference
No. Effluent density
removal
Hospital
12.2 Mahajan et
1 operation 6.75 Fe-Fe 100
mA/cm2 al. (2013)
theatre
Paint Akyol
2 35 A/m2 6.95 Fe-Fe 93
manufacturing (2012)
Landfill 15.9 Top et al.
3 7.00 Al-Al 45
leachate mA/cm2 (2011)
Pulp and 15 Sridhar et al.
4 7.00 Al-Al 90
Paper mA/cm2 (2011)
15 Yavuz et al.
5 Dairy industry 7.00 Fe-Al 79
mA/cm2 (2011)
Tannery 20 Babu et al.
6 6.80 Fe-Al 52
industry mA/cm2 (2007)
Tannery Feng et al.
7 1A 9.80 MS-MS 68
industry (2007)
Potato chips 300 Kobya et al.
8 4-6 Al-Al 76
manufacturing A/m2 (2006a)
Textile (RuO2-Ti) Muthukumar
9 5 A/dm2 12.0 85
industry – Ti et al. (2004)
Textile 200 Bayramoglu
10 11.0 Fe-Fe 76
industry A/m2 et al. (2004)
38

Some of the advantages of the methods are simple equipment, easy


operation, less operating time and decreased amount of sludge generation
which sediments rapidly and retains less water. Electrocoagulation used for
the removal of pollutants from different industrial wastewaters has been
successfully reported by the researchers and some of these are consolidated in
Table 2.9. Many studies have recorded (Khandegar and Saroha 2013;
Ogutveren et al. 1992) for the treatment of dye and textile industry effluent
using electrocoagulation.

2.3 TREATMENT STUDIES REPORTED ON THE SELECTED


DYES

2.3.1 C.I. Reactive Blue 194

Degradation of Reactive Blue 194 (RB 194) using sponge iron in


the presence of UltraSound (US) has been investigated by Liu et al. (2014).
Results indicate that the increased sponge iron dosage increases the
degradation rate of the dye and increased particle size of sponge iron has
inverse effect on the degradation of the dye. The rate constants for the
degradation of the dye are 0.00494, 0.02 and 0.138 min-1 for sponge iron,
ultrasonic irradiation and sponge iron-US respectively.

A microstructured tin doped electrode is synthesized and


investigated for the treatment of Reactive blue 194 in aqueous solution by An,
et al. (2012). The electrode shows good decolourization and mineralization
capability with an initial dye concentration of 30 mg/L and at a current
density of 150 mA/cm2 in neutral pH.

C.I. Reactive Blue 194 and C.I. Reactive Blue 222 have similar
molecular structure as given in Figure 2.1 and they are expected to exhibit
similar behavior during the treatment process. Kiran et al. (2013) have
reported that Reactive Blue 222 dye is subjected to photo-fenton’s oxidation
39

with decolorization of ≈90% which is further increased to 96.88 and 95.23 %


after aerobic treatment using biological processes. Similarly, electrolysis
using boron doped diamond electrode at 20 mA/cm2 is reported by Rangel et
al. (2013) with 100% of mineralization of Reactive Blue 222 dye with an
energy consumption of 11.8 KWh/m3.

Figure 2.1 (a) Reactive Blue 194 Figure 2.1(b) Reactive Blue 222

2.3.2 C.I. Direct Red 7

The structure of C.I. Direct Red 7 resembles C.I. Direct Blue 15 as


given in Figure 2.2 and it is expected to show similar behavior during the
treatment process.

Figure 2.2 (a) C.I. Direct Red 7 Figure 2.2 (b) C.I. Direct Blue 15

The application of Fenton oxidation process for decolorization of


direct blue 15 is investigated by Sun et al. (2009). The effect of initial pH,
initial concentration of H2O2, H2O2/Fe2+, H2O2/dye ratios and reaction
temperature on the decolorization efficiency and kinetic of direct blue 15 is
studied and it is reported that for the treatment of 4.7×10−5 mol/L of direct
40

blue 15 aqueous solution, 100% of decolorization efficiency within 50 min at


rate of 0.1694 min−1 under the optimum conditions is achieved.

Biosorption kinetic study of direct blue 15 on bacterial cellulose is


reported by Ashjaran et al. (2012). The adsorption rate gets decreased at high
temperatures, indicating exothermic nature of the process. The values of ∆H#,
∆G# and Ea suggest that the biosorption of the direct blue 15 onto bacterial
cellulose is spontaneous and physisorption.

Advanced Fenton process coupled with ultrasonic irradiation is


investigated by Weng, et al. (2013) on the decolourization of direct blue 15.
The process with an acoustic power of 140 W/L achieves more than 99% of
decolourization of 4130 ADMI solution within 10 min. Saroj et al. (2014)
have reported the biodegradation of azo dyes such as Acid Red 183, Direct
Blue 15 and Direct Red 75.

2.3.3 C.I. Acid Violet 17

The adsorption of acid violet 17 from aqueous solution by orange


peel as adsorbent is reported by Sivaraj, et al. (2001). A maximum removal of
87% is achieved at pH 2.0, adsorbent dose of 600 mg and 50 ml of 10 mg/L
initial dye concentration. The adsorption capacity Qo has also been calculated
as 19.88 mg/g at initial pH of 6.3. The adsorption increases with the increase
in initial pH.

The adsorption study on the activated carbon from an agricultural


solid waste by-product, sunflower seed hull is carried out for the
decolourization of acid violet 17 dye by Thinakaran et al. (2008). The
adsorption capacity of the activated carbon prepared at different temperatures
273, 353 and 393 K is found to be 65.8, 116 and 65.8 mg/g respectively.
41

Acidic pH is found to favor the adsorption process and thermodynamic


parameters reveal that the process is endothermic and spontaneous in nature.

Electrochemical regeneration of graphite intercalation compound


loaded with acid violet 17 is investigated in a spouted bed reactor (Liu et al.
2016). A good performance of simultaneous adsorption and electrochemical
regeneration is achieved with a liquid flow rate of 7.36 ml/s and at the current
density of 5 mA/cm2. Around 98% of removal efficiency of acid violet is
reported at the initial concentration of 100 ppm, effective surface area of 100
cm2 and 140 g of graphite intercalation compound by simultaneous adsorption
and electrochemical regeneration at the time of 60 minutes, with energy
consumption of 7.32 kWh per kg of acid violet 17. Acid violet 17 and crystal
violet show similarity in their structure as can be seen in Figure 2.3.

A comparative study on the kinetics of photo decolouration of


crystal violet by UV/H2O2 process and Fenton’s reagent is made by Alshamsi
et al. (2007). Crystal violet is found to degrade rapidly and efficiently in
H2O2/UV, but is better degraded by the Fenton process. Decolourization using
UV/H2O2 approach is the most efficient at 1.67 mmol of initial peroxide
concentration. Rate of decolourization of Fenton process is found to be
proportional to the oxidant concentration and Fe2+ concentration.

Figure 2.3 (a) C.I. Acid violet 17 Figure 2.3(b) Crystal violet
42

The photolytic degradation of crystal violet dye by Ag+ doped TiO2


and nanoanatase TiO2 is reported by Samira et al. (2013). The colour removal
efficiency of more than 99.5% is achieved for nanoanatase TiO2 and 75% is
achieved for Ag+ doped TiO2 at 45 min of UV illumination. Varying the
operational parameters show that nanoanatase TiO2 has a higher efficiency
than the Ag+ doped Titanium di oxide as catalyst.

Ghosh., et al. (2008) have noted that 99.0% of COD removal and
99.75% of colour removal of crystal violet by electrocoagulation using Al
electrode at initial dye concentration of 100 mg/L with the current density of
1112.5 A/m2, solution conductivity of 1.61 S/m and initial pH of 8.5 at the
end of 1hr of operation.

Optimization of the adsorption of crystal violet dye by the


adsorbent prepared from mango seed is studied by Sudamalla, et al. (2012).
The maximum removal efficiency of more than 89% at the initial
concentration of 50 mg/L, adsorbent dosage of 0.075 g and temperature of
40°C is obtained.

Adsorption of crystal violet from aqueous solutions on Palm kernel


fiber has been carried out at different initial dye concentrations from 20 to
160 mg/L, contact time, pH from 1.0 to 11.0 and sorbent doses from 0.4 to 8.0
g/L. Maximum adsorption capacity Qo is 78.9 mg/g for crystal violet at an
optimum pH reported by El-Sayed, et al. (2011).

2.4 OBJECTIVE OF THE PRESENT RESEARCH

The present work has been carried out to investigate the treatability
of Photo Assisted Chemical Oxidation (UV/H2O2) and Electrocoagulation
process on three different aqueous textile dye solutions. The performance of
the two processes is validated using simulated dye bath effluent. Finally,
43

feasibility of the two processes is tested for the real time textile industrial
effluents. To achieve the desired result, the present study has been started
with the following objectives.

1. To investigate the performance of PACO process of the three


different classes of aqueous dye solutions such as Reactive
blue 194 (RB194), Direct Red 7 (DR7) and Acid violet 17
(AV17) dyes with various parameters such as initial
concentration of dye, pH and oxidant dose.

2. To investigate the performance of EC on the same three


classes of dyes with varying operational parameters such as
initial concentration of dye, pH, electrolyte concentration,
applied voltage and electrode distance and to characterize the
sludge formed during the process using SEM and EDAX.

3. To prepare simulated dye bath effluent similar to real time


textile effluent.

4. To study the effect of operating parameters such as pollution


load, initial pH, initial oxidant concentration and treatment
time on the percentage of colour and COD removal of
simulated dye bath effluent using PACO process and to
develop a model using Box–Behnken experimental design
with four factors and three levels of optimization.

5. To study the effect of operating parameters such as pollution


load, initial pH and applied voltage on the percentage of
colour and COD removal of simulated dye bath effluent using
EC process and to develop a model for EC process using Box–
Behnken experimental design.
44

6. To check the adaptability of the PACO and EC process for


real time textile industrial effluent and to analyze the
percentage removal of various water quality parameters before
and after the treatment.

7. To compare the efficiency of PACO and EC processes in


terms of economic and environmental impacts.

You might also like