Anda di halaman 1dari 39

Accepted Manuscript

Treatment of oily wastewater from drilling site using electrocoagulation fol-


lowed by microfiltration

M. Changmai, M. Pasawan, M.K. Purkait

PII: S1383-5866(18)32103-8
DOI: https://doi.org/10.1016/j.seppur.2018.08.007
Reference: SEPPUR 14828

To appear in: Separation and Purification Technology

Received Date: 19 June 2018


Revised Date: 31 July 2018
Accepted Date: 5 August 2018

Please cite this article as: M. Changmai, M. Pasawan, M.K. Purkait, Treatment of oily wastewater from drilling site
using electrocoagulation followed by microfiltration, Separation and Purification Technology (2018), doi: https://
doi.org/10.1016/j.seppur.2018.08.007

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
Treatment of oily wastewater from drilling site using electrocoagulation followed by

microfiltration

M. Changmai*, M. Pasawan and M.K. Purkait

Indian Institute of Technology, Guwahati

Department of Chemical Engineering

Assam-781039

*Corresponding author

e-mail id: murchana@iitg.ac.in


1
Abstract

The present work deals with the utilization of electrocoagulation-microfiltration in batch mode

for the treatment of oil wastewater containing oil, grease along with metals like Na, Cr, Cu, Pb

and Ni. Samples were pre-treated using the electrocoagulation by varying working parameters

such as current density (20-80 A/m2), electrode distance (0.005-0.2 m) and initial pH (3.6-8.7).

The concentration of oil and grease was reduced from 35 mg/L to 10.2 mg/L in just 20 min.

Microfiltration was carried out using indigenously prepared ceramic membranes to remove these

flocs generated during the electro-coagulation process at three different pressures of 98,196 and

194 kPa.

Keywords: Oily wastewater; Electrocoagulation; Microfiltration; Membrane; Water quality

2
1. Introduction

Effluents from oil and gas well sites heave a potential impact on the environment. The chemical

and physical properties associated with these fluids influences the environmental impact ability.

Oily wastewater and drill cuts are the two most important wastes generated during the drilling

process. Some of the additives present in these wastes introduce potentially toxic compounds

into the fluids thereby increasing the toxic levels to a great extent [1]. Oil-water emulsions

through discharge from several industrial work in machining, automotive shops, oil-drilling,

refineries, off shore oil explorations, oil transportation and distribution have been one of the

major causes of water pollution. Moreover, emulsions of oil-in-water have also been widely used

in metal processing industry such as in industrial work of cooling and lubrication which are later

replaced and discharged due to substantial degradation of certain components. This oil

discharged into the environment may be of different forms such as free (oil and grease),

dispersed and emulsions [2], [3, 4].

Till date a large number of methods have been proposed by various researchers from all over the

world to treat such forms of oil in water. It is a necessity, in the present times, to identify new

technologies which would be efficient enough in separating oil form such oil-in water emulsions.

Methods such as adsorption [5], coagulation [6-8] anaerobic treatment [9], reverse osmosis [10],

ultrafiltration [11], chemical destabilization by using inorganic salts [12], flocculation [13],

dissolved air flotation [14] and membrane processes [15, 16] have been widely used in the past

to treat such water. Previous research studies have suggested that electrocoagulation process has

a high rate of success in the separation of oil from oily rejects in comparison to other available

processes. Thus, electrocoagulation plays an important role in the treatment of oily wastewaters.

This process in combination with other techniques can be an effective method for a highly

3
efficient process [17, 18]. The main advantages being the non-requirement of any additional

additive chemicals to enhance the process of destabilizing the emulsion, a simple and easy to use

working set-up, lower investment and lower operating costs and most importantly a decreased

amount of sludge production [19-21]. However, owing to the consumption of electrode with time

and its corrosion, it becomes necessary to replace the electrodes for better performance.

The process of electrocoagulation involves the dissolution of the sacrificial anode as a result of

application of potential difference across the electrodes. This dissolution of anode results in the

production of flocs in the form of metallic hydroxides within the sample of water to be treated.

This in-situ process mainly includes three major steps i.e. an electrolytic reaction at the electrode

surface, formation of coagulants and finally the adsorption of the colloidal particles onto the

produced flocs which gets removed either by sedimentation of floatation. It is also of utmost

importance that a proper electrode material is selected for carrying out the process of

electrocoagulation. Materials such as aluminium and iron are frequently used and have been

found to give good results in comparison to other electrode materials [22-24]. Combination of

electrocoagulation (EC) with processes such as electro-floatation, reverse –osmosis and

nanofiltration have already been reported previously. These methods were basically exploited for

the treatment of polluted drinking water, like the removal of fluoride from drinking water,

treatment wastewater from textile industry (electrocoagulation-nanofiltration), treatment of

brackish water (electrocoagulation-reverse osmosis) [25-27]. In order to use the combination of

electrocoagulation along with filtration it is very important to know the particle size of the

generated flocs. The particle size would in turn help in determining the type of hybrid process

viz. determine the type of filtration process that can be clubbed with EC process. In addition to

the synthetically prepared oil water, EC process has been utilized in the treatment of oily

4
wastewater such as refectory oily wastewater, vegetable oil refinery industry wastewater, palm

oil mill effluent, vegetable oil wastewater, crude oil and restaurant oil water [28, 29]. With

increasing threat to the aquatic life due to high toxic levels of the drilling effluents, it has become

an utmost necessity to treat the effluents before being discharged into the environment. Different

standards have been set up to control the water pollution scenario due to such industrial effluents.

As per central pollution board of India (CPCB) the permissible limit for oil and grease stands at

35 mg/l [30-36]. Hence it has become an utmost importance to develop a method to maintain the

permissible limits of drilling effluents before being released into the environment [37-40].

To the best of our knowledge no work has been done till date on treatment of drilling wastewater

using the hybrid technique adopted in this work. In view to this, present work utilized EC

process with a microfiltration (MF) set up so as to remove the flocs produced during the EC

process. Effluent water from the drilling site of Barekuri, Oil India Limited (OIL) was collected

on 29th August, 2017. The obtained sample was from a drilling location TAU having a well

depth of 3364 m. Obtained effluent samples had very high quantity of suspended solids, TDS,

sulphates, chlorides, COD and BOD. The oil and grease content was around 35 mg/L.

Electrocoagulation process was carried out on these effluent samples using Al-electrodes with

varying operating parameters such as current density, electrode distance and initial pH. Al-

electrodes were preferred over Fe-electrodes due to the fact that the use of the later resulted in a

reddish tinge to the treated water. Once electro-coagulated the sample was sent to a

microfiltration set-up assembled with ceramic membrane prepared form thermal power plant

slag. The slag was preferred for membrane preparation due to the presence of high concentration

of alumina and silica. Four different sintering temperatures were chosen for investigating its

effect on the morphological and microfiltration properties of the membrane. The performance of

5
prepared microfiltration membrane was investigated to separate electro-coagulated byproduct. A

cost estimation on the electrocoagulation process was also done to investigate the power

consumption and electrode material utilized during each process. Detailed flux studies were done

to study the performance of the membrane during the microfiltration process.

2. Mechanism of oil removal by electrocoagulation

Electrocoagulation (EC) process utilizes the benefits of flocculation and coagulation for the

removal of oil and grease from water. The process involves the usage of Al electrodes which

when under the influence of a potential difference results in the generation of flocculating

species from the anode. Amorphous Al(OH)3 flocs produced during the EC process have large

surface areas which in turn have high affinity towards colloidal particles [41-45]. The hydrogen

and oxygen gas produced during the EC process further aids in the removal of oil and grease

from aqueous solutions [28, 29]. The reactions occurring are briefly given below

Reactions occurring at the anode


3+ -
Al (s) Al (aq) + 3e (1)

2H2O O2 (g) + 4H+ + 4e- (2)

Reactions occurring at the cathode

2H2O + 2e- H2 (g) + 2OH- (aq) (3)

Reactions occurring at bulk

Al3+ (aq) + 3OH- Al(OH)3 (s) (4)

Al(OH)3 (s) + (Under alkaline conditions) (5)

6
2.1 Experimental

2.1.1 Materials and methods

Effluent water from the drilling site of Barekuri, Oil India Limited (OIL) was collected on 29 th

August, 2017. The obtained sample was from a drilling location TAU having a well depth of

3364 m. Obtained effluent samples had very high quantity of suspended solids, TDS, sulphates,

chlorides, COD and BOD. The oil and grease content was around 35 mg/L. The initial

concentration of all the components of the feed are given in Table 2.

2.1.2 Measurement and analysis

The concentration of oil and grease in the effluent after EC treatment was determined using UV-

Vis spectrophotometer (Make: Perkin Elmer). The quality of the treated effluent after

electrocoagulation and microfiltration were studied in terms of turbidity, conductivity, salinity,

TDS, D.O and pH using a water analyzer kit (Make: VSI electronics pvt. Ltd.). Studies on the

produced flocs during EC process and the membrane used for microfiltration were done using

EDX (Make: Leo), FESEM (Make: Leo) and Delsa Nano (Make: Beckman Coulter). The metals

present in the sample such as Fluorides, Chromium, Copper, Mercury, Nickel, Lead etc. were

analyzed using Atomic Absorption Spectroscopy (Make: Varian). Stock solution were purchased

from Sigma and corresponding standards were prepared for each metal before analysis. Since the

concentration of the metals were more or less near the permissible limit values, a final

concentration after electrocoagulation has been analyzed instead of a detailed study.

2.1.3 Membrane preparation

Thermal power plant slag was mixed with sodium carbonate, boric acid, sodium metasilicate and

alumina using a mortar and pastel. The obtained mixture was mixed with a definite amount of

7
Millipore water to form a paste of suitable consistency for casting on ring mold. The paste was

thus casted on an MS ring to form disk shaped membranes which were then kept under a uniform

weight of 1-2 kg overnight to prevent deformation of the membrane during the drying and

sintering process [27]. This membrane was then removed from the MS ring very carefully and

the placed in a hot air oven at a temperature of 110 oC to remove the water and excess moisture

content for 12 h. The dried membrane was then sintered at three different temperatures of 700,

800, 900 and 1000 oC for 12 h. The obtained membranes were then polishes and washed before

being used for microfiltration studies.

2.1.4 Electrocoagulation bath

The collected effluent from OIL was pre-treated using electrocoagulation before being filtered

using a ceramic membrane. The set up consisted a rectangular box of dimensions 0.23 m × 0.08

m × 0.04 m with a holding capacity of 600 ml. Aluminium sheets of dimension 0.07 m × 0.77 m

× 0.001 m with an inter-electrode distance of 0.005 m were used as the electrode which were

connected to a direct current power source (Fig.1). The EC process was carried out maintain

various operating parameters to study the removal of oil and grease during the electrocoagulation

process. Thus, EC was carried out by varying the parameters such as current density, electrode

distance and initial pH.

Figure 1 here

3. Results and discussion

3.1 Effect of current density, electrode distance and initial pH on removal of oil and grease

Three important parameters i.e. current density (20-80 A/m2), electrode distance (0.005-0.02 m)

and initial pH (3.6-8.7) were considered when studying the removal of oil and grease from

8
effluent water. All the varying parameters were studied experimentally for an operating time

period of 20 min. During the EC process flocs are produced due to anodic oxidation which helps

in the capturing of oil and grease from the aqueous solution. Thus more is the current density,

more flocs will be produced and hence more removal of oil and grease will be observed. As seen

from the Figure 2a, with an increase in current density as 20, 50 and 80 A/m2 the final

concentration of oil in the effluent reduced to 29.3, 27.9 and 26.86 mg/L respectively from an

initial concentration of 35 mg/L. With an increase in applied current density, anodic oxidation

accelerates resulting in the dissolution of the sacrificial anode facilitating the production of

aluminium hydroxide flocs which aid in the removal of oil from effluent [27], [31], [32].

Electrode distance also plays an important role in the removal of oil and grease from effluent

water. Higher is the distance in between the electrodes, more is the current required to produce

aluminium hydroxide responsible for the removal of oil from effluent. Figure 2b shows the

effect of varying electrode distance with a constant current density of 80 A/m2 and a constant

operating time of 20 min. At the same current density, the least flocs would be produced by the

highest electrode distance which is well explained by the Figure. Although the removal of oil is

not much effected as the concentrations were reported to differ by a mere 0.64 mg/L when the

electrode distance was increased from 0.005 to 0.02 m. Inter-electrode distance is another

important parameter controlling the electrocoagulation process [31], [32].

Mainly, two mechanisms were found responsible for removal performance during

electrocoagulation process. These interactions are adsorption and precipitation, each being

governed by a definite pH range. At a low pH range, flocculation as precipitation is dominant

whereas at higher pH range (> 6.5) adsorption is the major mechansim. The dominant aluminum

species when the solution pH is 2-4 are Al3+, Al(OH)22+ and Al(OH)2+ and for pH between 4-9

9
dominant species include species such as Al13O4(OH)247+ are formed and get precipitated as

Al(OH)3(s). For initial pH of 3.6, maximum removal was observed with a final residual

concentration of 10.21 mg/L (Fig. 2c).

Figure 2 here

3.2 Effect of current density, electrode distance and initial pH on removal of chlorides, TSS

and sulphates

Similar studies were done to see the effect of current density, electrode distance and initial pH on

the removal of chlorides, total suspended solids and sulphates present in the effluent water. As

seen from Figure 3 with an increase in current density the chlorides and sulphates concentration

reduced to 1274.6 mg/L and 18218 mg/L respectively. However the amount of suspended solids

increased with an increase in current density due to more production of aluminium hydroxide at

higher current densities. Similar observations were made for changing pH values. At acidic pH

the chlorides concentration was high whereas at basic pH the concentration of chlorides was low.

Similarly TSS was high at acidic pH due to high flocculation by precipitation with a

concentration of 630 mg/L. The sulpahtes concentration dropped to 18189 mg/L at acidic pH

whereas it increased to 18218 at basic pH.

Electrode distance had very less impact on the removal of chlorides, suspended solids and

sulphates from the effluent water. Concentration at varying electrode distance of 0.005-0.02 m

almost remained the same as seen from Figure 3c.

Figure 3 here

10
3.3 Estimation of corrosion, film thickness, energy consumption and operating cost

In order to determine the lifetime of the electrocoagulation process it is necessary to carry out the

corrosion studies on the used electrodes. As electrocoagulation proceeds with time there will be a

weight loss of the electrodes which shall determine the amount of corrosion the electrodes

undergo. As seen from the Figure 4a with an increase in current density from 20 to 80 A/m2 the

corrosion increased from 24.5 to 71.11 mg. This is because of the fact that at higher current

densities more anodic dissolution occurs thereby resulting in the weight loss of the electrodes.

However at increasing electrode distances the amount of corrosion of the electrodes reduced

drastically. Low electrode distance of 0.005 m had corrosion of 71.11mg whereas with an

increase in electrode distance to 0.02 m, corrosion decreased to 37.6 mg.

During the EC process a gelatinous layer formed on the surface of the electrodes which forms an

additional barrier towards the formation of aluminum hydroxide. The film thickness of this

gelatinous layer formed over the electrode surface has already been mentioned in our previous

work. The film thickness increased from 2.43 to 7.05 µm when the current density increased

from 20 to 80 A/m2. A high electrode distance resulted in low thickness of gel layer and a low

electrode distance resulted in a high gel layer thickness. The gel layer thickness were 7.05 and

3.73 µm for 0.005 and 0.02 m electrode distance, respectively (Fig 4b).

Figure 4c shows the variation of operating costs at different current densities at the end of 20

min of EC operation for a working volume of 600 ml effluent. With increased current density

energy cost increased due to higher energy consumptions. Cost estimation was carried out as per

the previous work [27]. Costs associated with consumption of electricity (0.0936 US$/Kwh) and

electrode (2.046 US$/ Kg of Al) were considered as per the year 2017. As seen, from the Fig. 4c

increasing current density resulted in higher electrode cost and energy cost which was due to

11
more anodic oxidations and increased energy consumptions. The total cost as a result increased

from 0.0502 to 0.253 US$/m3 when current density increased from 20 to 80 A/m2.

Figure 4 here

4. Water permeation experiment

4.1. Porosity

The membrane porosity at 700, 800, 900 and 1000 oC were determined using the Archimedes’

principle. The following equation was used

P= (6)

Where, Mw is the mass of the ceramic membrane saturated with water, Md is the dry mass of the

ceramic membrane and Ms is the mass of the suspended membrane when submerged into water.

The membranes were first weighed in air followed by suspending the membranes in water to

measure the suspended weight .The membranes were then soaked in water for 24 h, removed

from water and measured again for the soaked weight.

Figure 5a shows the changes in bulk porosity as the sintering temperature varies from 700-1000
o
C for the ceramic membranes prepared by paste method. As seen from the figure, it is evident

that as sintering temperature increases, ceramic membrane porosity decreased. A decrease from

55 % to 30.13 % was seen when the sintering temperature increased from 700 to 1000 oC. This

decrease in porosity is because of the fusion of the membrane constituents at higher

temperatures. When the sintering temperature is increased more melting of the membrane

constituents shall occur resulting in removal of the existing voids and thereby leading to

structural densification.

12
4.2 Hydraulic permeability

Hydraulic permeability variation at different sintering temperatures were also determined. The

hydraulic permeability was estimated by using the following equation

(7)

Where, J is the flux across the membrane in m3/m2s, ΔP is the transmembrane pressure in kPa, ε

is the porosity of the ceramic membrane (ε=nπr2), µ is the viscosity of water, l is the length of the

pore. As seen from Fig. 5a (top inset) the hydraulic permeability increased with an increase in

the sintering temperature. The hydraulic permeability increased from 5.7×10-9 m/Pa.s to 2.47×10-
8
m/Pa.s when the sintering temperature increased from 700 to 1000 oC. This increase in the

hydraulic permeability was because of the fact that at higher sintering temperature the pore size

of the ceramic membranes increased resulting in a increase in the permeate flux of the ceramic

membrane. The hydraulic permeability was also determined after microfiltration of electro-

coagulated oil samples and it was found that the values decreased from 2.30×10 -9 to 4.00×10-9

m/Pa.s respectively when the sintering temperature of the ceramic membrane increased from 700

to 1000 oC. It was already reported in our previous work, [27] that void spaces were created

more when ceramic membranes were prepared by paste method. Also reports from FESEM

analysis suggested that high sintering temperature, increased the membrane pore size which has

been discussed in details in the section below.

4.3 The average pore size

13
The average pore size of the ceramic membranes were determined using water permeation

experiments and imageJ software for SEM analysis. For the water permeation experiments, the

equation used for the determination of the pore size is as follows

(8)

Where, µ is the viscosity of water, l is the length of the pore, L p is the hydraulic permeability and

ε is the porosity of the membrane. Figure 5a (bottom inset) shows the change in pore sizes with

sintering temperature. As seen the pore size increased with sintering temperature which was

because of large voids created at higher temperatures due to burning off of CaO. The Fig. 5a

(bottom inset) shows that the average pore size for both the methods of pore size determination

agreed with the assumption. Pore size increased from 1.175 µm to 24.76 µm for temperature

changes from 700 to 1000 oC by imageJ software and from 0.72 µm to 1.87 µm for temperature

changes from 700 to 1000 oC for water permeation experiments. Moreover, the figure suggests

that pore sizes obtained by imageJ software were considerably high from the water permeation

experiments. This is because the software mainly determines pore size based on surface pore

which may or may not extend throughout the entire membrane whereas the water permeation

experiment considers the continuous pores that are actually involved in the microfiltration of

permeate. Thus, the actual pore size of 0.72 µm to 1.87 µm were involved in the microfiltration

of the electro coagulated oily wastewater and the pore channels were continuous from the top to

the bottom of the membrane.

4.4 Permeation experiment

Permeation experiments were carried out using the electro-coagulated sample of effluent at 80

A/m2. Each of the membranes sintered at 700, 800, 900 and 1000 oC were used for the

14
permeation experiments in a batch mode. Trans-membrane pressures were maintained at three

different pressures of 98,196 and 194 kPa. Each membrane was used to filter deionized water at

a transmembrane pressure of 392 kPa in order to ensure that the pores were unblocked before

carrying out the effluent permeation experiments. It was observed that initially the membrane

flux was very high and as the microfiltration experiments progressed it gradually reduced and

reached a final steady state value. Steady state flux was achieved within 20 min of operation. The

permeate flux (Fig. 5b) followed an increasing trend for higher sintering temperatures because at

higher sintering temperatures, the pore size increased. As evident, at a pressure of 98 kPa the

permeate flux increased from 264 to 423 L/h.m2 when sintering temperatures increased from 700

to 1000oC. However, with time there was a decline in flux as the suspended flocs deposited on

the membrane surface thereby blocking the active pores of the membranes and resulting in a

decrease of permeate flux. Again, increase in permeate flux with pressure is due to the more

driving force.

Figure 5 here

4.5 Permeate flux profile after MF experiments

In order to determine the usability of the membrane, the pure water flux profiles before and after

microfiltration studies were determined. Figure 6a shows the pure water flux of the ceramic

membrane before and after the experiments respectively. As observed from the figure 6a (inset),

there is a decline in the flux after microfiltration of electro-coagulated samples. As seen, for the

pressure 98 kPa the steady state flux of 264 L/h.m2 was reached at 16 min (Fig. 6b), the decrease

in flux with time was due to the deposition of flocs on the membrane surface. Pure water flux

had a steady state value of 2011.75 L/h.m2 at 20 min, however after microfiltration of EC

sample, the flux was reduced to 811 L/h.m2 at 10 min. This declination in flux profile was

15
evident for all the membranes. The decrease in flux after usage was due to the deposition of the

suspended flocs of the EC samples on the surface of the membrane thereby blocking the active

pores which were available for the filtration process.

Figure 6 here

To quantify the results, flux decline ratio (FDR) and flux recovery ratio (FRR) were determined

[29].

Flux decline ratio (FDR) = (9)

Flux recovery ratio (FRR) = (10)

Where, Ji is the initial permeate flux, Jf is the steady state permeate flux and Jw is the pure water

flux. It can be observed from Table 1, that FDR of 0.659-0.745, 0.6-0.75, 0.778-0.837 and 0.76-

0.854 was observed for the membranes sintered at 700, 800, 900 and 1000 oC respectively at

pressure of 98-294 kPa indicating that 75-85 % of the initial flux was lost during the

microfiltration of the electro-coagulated samples. Similarly, FRR of 0.131-0.041, 0.082-0.040,

0.054-0.049 and 0.048-0.044 was observed for the membranes sintered at 700, 800, 900 and

1000oC respectively suggesting that the cleaning of membrane is not much effective in all

conditions. Thus, all the membranes were washed several times with soap water to retain

sufficient flux values. The organic solutes, responsible for fouling, deposit over the membrane

surface forming a thick cake layer which needs to be scraped out and the membrane surface

needs to be washed rigorously to retain atleast 50% of the former flux properties. The flux could

be further recovered by washing the membrane surface with a good detergent using a brush,

followed by ultrasonication to remove the adhered oil and grease. This process helped in

16
recovering almost 80 % of the flux. The clean membrane, thus obtained could be re-used for

further experiments.

Table 1 here

4.6 Floc characterization and membrane characterization (before and after filtration)

Floc size of the electro-coagulated samples were determined using delsa-nano. The size of the

flocs varied from 2.61 to 5.83 µm when the applied current density for electrocoagulation varied

from 20 to 80 A/m2. As evident from Fig. 7a the size variation curve was bimodal in nature.

Similarly floc size was 124.49 µm at a pH of 3.6 and 11.87 µm at pH of 5.7 flocs. The EDX

analysis of the produced flocs provided a details on the elements present. Presence of metals

such as Cr, Ni, Hg, Cu, Pb, F etc. suggested that electrocoagulation has successfully removed

these metals from the sample oily water (Fig. 7b).

The membranes prepared at different sintering temperatures were analyzed using XRD (Fig. 7c).

As seen, the crystalline nature increased with an increase in sintering temperature form 700 to

1000oC. Phases [111], [200] and [222] were present which corresponded to the calcium oxide,

dicalcium silicates. XRD analysis of the flocs suggested an amorphous nature of the dried flocs.

From the water permeation experiments it was observed that the membrane pore size varied from

0.72 µm to 1.87 µm for temperature changes from 700 to 1000 oC. As discussed above, floc size

obtained in the electrocoagulation experiments were 2.61 to 5.83 µm when the applied current

density varied from 20 to 80 A/m2. Hence, the present membranes were suitable for carrying out

the microfiltration of the electro coagulated oily wastewater. Figure 7d shows the membrane

after microfiltration experiment. As evident from the image, a layer of filtered flocs covers the

17
membrane top layer which can be removed easily by washing. Figure 7e shows the flocs

retained by ceramic membrane during the microfiltration process. Figure 7f shows the FESEM

image of the dried flocs suggesting that the flocs do not inherit any specific shape.

Figure 7 here

4.8 Performance of hybrid process

The quality of the feed and electro-coagulated water were determined and compared along with

the available central pollution control board (CPCB) permissible limits Table 2 shows the

obtained results. As can be seen from the table TDS values decreased from 3230 to 2780 mg/L.

Conductivity of the feed increased from 2.65 to 6.19 which was due to the addition of NaCl,

salinity remained more or less the same. pH was within permissible limits for all the sets of

experiments. Turbidity increased from 28.1 to 71.6 because of large amounts of flocs produced

during the electrocoagulation process. Metals like sodium, chromium, copper, lead, nickel were

brought down within their permissible limits by electrocoagulation. The concentration of oil and

grease was reduced from 35 mg/L to 10.2 mg/L in just 20 min. A further long duration of the

experiment till 30 min suggested the concentration of oil and grease to be below the permissible

limits of 10 mg/L. A further increase in the electrocoagulation time (30 mins) resulted in the

lowering of oil and grease content to as low as 2 mg/L. However, since the present work is for an

operation time of 20 min, the result obtained for 30 min has not been included.

Table 2 here

It was found that a large number of work has been done for the treatment of oily water in the past

using the EC process. X. Xu and X. Zhu performed electrocoagulation on refectory oily

wastewater having and initial oil concentration of 100-250 mg/L. M. Basibuyuk and D.G. Kalat

18
did electrocoagulation on Vegetable oil refinery wastewater with initial oil concentration of

2648-3880 mg/L. Similar work was done by N. Oswal et al., X. Chen et al., P. Canizares, U.B.

Ogutveren and S. Koparal with initial oil concentrations of 250000 mg/L, 100-2100 mg/L, 1500-

6000 mg/L and 200 mg/L (Table 3). Each of these samples were treated using electrocoagulation

process and a removal of 85-98 % was achieved in all the cases. However, the EC experiments

were performed for a longer period of time of 45-60 min in comparison to the present work of 20

min. Literature review suggested that although a large experiments were performed on oil

wastewater, however no work was done till date on oily wastewater from a drilling waste water.

The present sample had oil concentration of 35 mg/L which was brought down to 10 mg/L in 20

min of operation. Hence, EC process can be opted as a viable option for the treatment of oil

water from drilling sites.

Table 3 here

5. Conclusion

Effluent water from the drilling site in Barekuri, India having oil and grease along with metals

like Na, Cr, Cu, Pb and Ni were treated using the electrocoagulation technique and the treated

effluent was then filtered through a ceramic membrane to remove the flocs produced during the

EC process. Pretreatment involved varying working parameters such as current density (20-80

A/m2), electrode distance (0.005-0.2 m) and initial pH (3.6-8.7). Due to the application of a

potential difference across the electrode surface, huge amounts of flocs were produced which

were responsible for the removal of oil and grease along with the unwanted metals in the effluent

water. The concentration of oil and grease was reduced from 35 mg/L to 10.2 mg/L in just 20

min. TDS values decreased from 3230 to 2780 mg/L. Conductivity of the feed increased from

2.65 to 6.19 which was due to the addition of NaCl, salinity remained more or less the same.

19
Microfiltration was carried out to remove these flocs from the treated water. Four different

membranes prepared at sintering temperatures of 700, 800, 900 and 1000 oC were utilized for the

microfiltration studies. Trans-membrane pressures were maintained at three different pressures

of 98,196 and 194 kPa. At a pressure of 98 kPa the permeate flux increased from 264 to 423

L/h.m2 when sintering temperatures increased from 700 to 1000 oC. Again, with an increase in

pressure, permeate flux increased due to more driving force. It was also observed that 75-85 % of

the initial flux was lost during the microfiltration of the electro-coagulated samples.

Conflict of interest

The authors declare no competing financial interest.

Acknowledgement

We would like to thank the Central Instruments Facility (CIF), IITG for providing us with the

necessary equipment’s for carrying out the experiments.

References

[1]. A.M.H. Elnenary, E. Nassef, G.F. Malash, M.H.A. Magid, Treatment of drilling

fluids wastewater by electrocoagulation, Egypt. J. Pet., 26 (2017), pp. 203-208

[2]. Y.O. Fouad, Electrocoagulation of crude oil from oil-in-water emulsions using a

rectangular cell with a horizontal aluminium wire gauze anode, J. Dispers. Sci. and

Technol., 34 (2012), pp. 214-221.

[3]. P. Cañizares, F. Martínez, C. Jiménez, C.S. Manuel, A. Rodrigo, Coagulation and

electrocoagulation of oil-in-water emulsions, J. Hazard . Mater., 151 (2008), pp. 44-

51.

20
[4]. M. Carmona, M. Khemis, J.P. Lecler, F. Lapicque, A simple model to predict the

removal of oil suspensions from water using the electrocoagulation technique,

Chem. Eng. Sci., 61 (2006), pp. 1237-1246.

[5]. E. Sabah, M. Ḉnar, M.S. Ḉelik, Decolorization of vegetable oils: Adsorption

mechanism of β-carotene on acid-activated sepiolite, Food Chem., 100 (2007), pp.

1661-1668.

[6]. N. Azbar, T. Yonar, Comparative evaluation of a laboratory and full-scale treatment

alternatives for the vegetable oil refining industry wastewater (VORW), Process

Biochem., 39 (2004), pp. 869-875.

[7]. M. Basibuyuk, D.G. Kalat The use of waterworks sludge for the treatment of

vegetable oil refinery industry wastewater, Environmental Technol., 25 (2004), pp.

373-380.

[8]. N. Oswal, P.M. Sarma, S.S. Zinjarde, A. Pant, Palm oil mill effluent treatment by a

tropical marine yeast, Bioresource Technol., 85 (2002), pp. 35-37.

[9]. U.T. Un, A.S. Koparal, U.B. Ogutveren, Electrocoagulation of vegetable oil refinery

wastewater using aluminum electrodes, J. Environ. Manag., 90 (2009), pp. 428-433.

[10]. S. Sridhar, A. Kale, A.A. Khan, Reverse osmosis of edible vegetable oil industry

effluent, J. Membr. Sci., 205 (2002), pp. 83-90.

[11]. T. Mohammadi, A. Esmaeelifar, Wastewater treatment of a vegetable oil factory

by a hybrid ultrafiltration-activated carbon process, J. Membr. Sci., 254 (2005), pp.

129-137.

21
[12]. G. Rios, C. Pazos, J. Coca, Destabilization of cutting oil emulsions using

inorganic salts as coagulants, Colloids Surf. A: Physicochem. Eng. Asp., 138 (1998),

pp. 383-389.

[13]. A. Pinotti, N. Zaritzky, Effect of sluminium sulfate abd cationic polyelectrolytes

on the destabilization of emulsified wastes, Waste manag., 21(2001), pp. 535-542.

[14]. A.A. Al-Shamrani, A. James, H. Xiao, Separation of oil from water by dissolved

air flotation, Colloids Surf. A: Physicochem. Eng. Asp., 209 (2002), pp. 15-26.

[15]. K. Scott, R.J. Jachuck, D. Hall, Crossflow microfiltration of water-in-oil

emulsions using corrugated membranes, Sep. Purif. Technol., 22 (2001), pp. 431-

441.

[16]. S.E. Burns, S. Yiacoumi, C. Tsouris, Microbubble generation for environmental

and industrial separation, Sep. Purif. Technol., 16 (1997), pp.193-203.

[17]. X. Chen, G. Chen, P.L. Yue, Novel electrode system for electroflotation

wastewater, Environ. Sci. Technol., 36 (2002), pp.778-783.

[18]. N. Moulai-Mostefa, M. Tir, Coupling flocculation with electroflotation for waste

oil/water emulsion treatment. Optimization of the operating conditions,

Desalination, 161 (2004), pp. 115-121.

[19]. X. Xu, X. Zhu, Treatment of refectory oily wastewater by electro-coagulation

process, Chemosphere, 56 (2004), pp. 889-894.

22
[20]. C.L. Lai, S.H. Lin, Treatment of chemical mechanical polishing wastewater by

electrocoagulation: system performances and sludge settling characteristics,

Chemosphere, 54 (2004), pp. 235-242.

[21]. C.L. Yang, Electrochemical coagulation for oily water demulsification, Sep.

Purif. Technol., 54 (2007), pp. 388-395.

[22]. M. Kobya, O.T. Can, M. Bayramoglu, Treatment of textile wastewaters by

electrocoagulation using iron and aluminum electrodes, J. Hazard. Mater., 100

(2003), pp. 163-178.

[23]. M.Y.A. Mollah, P. Morkovsky, J.A.G. Gomes, M. Kesmez, J. Parga, D.L.

Cockec, Fundamentals, present and future perspectives of electrocoagulation, J.

Hazard. Mater., 114 (2004), pp. 199-210.

[24]. M. Kobya, H. Hiz, E. Senturk, C. Aydiner, E. Demirbas, Treatment of potato

chips manufacturing wastewater by electrocoagulation, Desalination, 190 (2006), pp.

201-211.

[25]. Q. Zuoa, X. Chena, W. Li, C. Guohua, Combined electrocoagulation and

electroflotation for removal of fluoride from drinking water, J. Hazard. Mater., 159

(2008), pp. 452-457.

[26]. A. Aouni, C. Fersi, B. Mourad, A. Sik, D. Mahmoud, Treatment of textile

wastewater by a hybrid electrocoagulation/nanofiltration process, J. Hazard. Mater.,

168 (2009), pp. 868-874.

23
[27]. W. Den, C. Wang, Removal of silica from brackish water by electrocoagulation

pretreatment to prevent fouling of reverse osmosis membranes, Sep. Purif. Technol.,

59 (2008), pp. 318-325.

[28]. D. Ghosh, M.K. Sinha, M.K. Purkait, A comparative analysis of low-cost ceramic

membrane preparation for effective fluoride removal using hybrid technique,

Desalination, 327 (2013), pp. 2-13.

[29]. D. Ghosh, C.R. Medhi, M.K. Purkait, Treatment of fluoride containing drinking

water by electrocoagulation using monopolar and bipolar electrode connections,

Chemosphere, 73 (2008), pp. 1393-1400.

[30]. R. Mukherjee, M. Mondal, A. Sinha, S. Sarkar, S. De, Application of

nanofiltration membrane for treatment of chloride rich steel plant effluent, J.

Environ. Chem. Eng., 4 (2016), pp. 1-9

[31]. U.B. Öğütveren, S. Koparal, Electrocoagulation for oil‐water emulsion treatment,

J. Environ. Sci. Health, Part A: Environ. Sci. Eng. Toxi., 32 (1997), pp. 2507-2520.

[32]. X. Chen, G. Chen, P L. Yue, Separation of pollutants from restaurant wastewater

by electrocoagulation, Sep. Purif. Technol., 19 (2000), pp. 65-76.

[33]. P. Tiwari, Water quality assessment for drinking and irrigation purpose, Indian J.

Sci. Res., 13 (2017), pp. 140-142.

[34]. Standards for Effluent Discharge Regulations, General Notice No.44. of 2003,

The Environmental protection act, 2002,Regulations made by the Minister under

sections 39 and 96 of the Environment Protection Act 2002

24
[35]. S.N. Singh, G. Srivastav, A. Bhatt, Physicochemical Determination of Pollutants

in Wastewater in Dheradun, Curr. World Environ., 7 (2012), pp. 133-138

[36]. E. Bazrafashan, H. Moein, F.K. Mostafapour, S. Nakhaie, Application of

Electrocoagulation Process for Dairy Wastewater Treatment, . J. Chem., (2013), 8

pages

[37]. Drinking Water Standards of BIS (IS: 10500: 1991).

[38]. Environmental protection agency (EPA) (1975), Patuxent Wastewater Treatment

Facilities: Environmental Impact Statement, Philadelphia, Pennsylvania.

[39]. Drinking Water specifications of BIS (IS: 10500:2012)

[40]. Integrated iron and steel industry effluent discharge standards, Central pollution

control board of India, 2012

[41]. F. N. Azad, M. Ghaedi, K. Dashtian, M. Montazerozohori, S. Hajati and E.

Alipanahpour, Preparation and characterization of MWCNTs functionalized by N-

(3-nitrobenzylidene)-N′-trimethoxysilylpropyl-ethane-1,2-diamine for the removal of

aluminum(III) ions via complexation with eriochrome cyanine R:

spectrophotometric detection and optimization, RSC Adv.75, 2015.

[42]. F. N. Azad, M. Ghaedi, K. Dashtian, S. Hajati, V. Pezeshkpour, Ultrasonically

assisted hydrothermal synthesis of activated carbon-HKUST-1-MOF hybrid for

efficient simultaneous ultrasound-assisted removal of ternary organic dyes and

antibacterial investigation: Taguchi optimization, Ultrason. Sonochem., 31 (2016),

pp. 383-393.

25
[43]. S. Mosleh, M. R. Rahimi, M. Ghaedi, K. Dashtian and S. Hajati BiPO4/Bi2S3-

HKUST-1-MOF as a novel blue light-driven photocatalyst for simultaneous

degradation of toluidine blue and auramine-O dyes in a new rotating packed bed

reactor: optimization and comparison to a conventional reactor, RSC Adv., 68

(2016).

[44]. B. Jafari, M. R. Rahimi, M. Ghaedi, K. Dashtian, S. Mosleh, CO2 capture by

amine-based aqueous solution containing atorvastatin functionalized mesocellular

silica foam in a counter-current rotating packed bed: Central composite design

modeling, Chem. Eng. Res. Des., 129 (2018), pp. 64-74

[45]. M.Ghaedi, F. N. Azad, K. Dashtian, S. Hajati, A. Goudarzi, M. Soylak, Central

composite design and genetic algorithm applied for the optimization of ultrasonic-

assisted removal of malachite green by ZnO Nanorod-loaded activated carbon,

Spectrochim. Acta A: Mol. Biomol. Spectrosc., 167 (2016), pp. 157-164.

26
Table 1: Flux decline ratio (FDR) and flux recovery ratio (FRR) at different transmembrane
pressure for the membranes prepared at 700, 800, 900 and 1000 oC

Membrane sintering Transmembrane pressure FDR FRR


temperature (oC)
(kPa)
700 98 0.659 0.131
196 0.742 0.047

294 0.746 0.041

98 0.600 0.082
800
196 0.719 0.047

294 0.75 0.040

900 98 0.778 0.054

196 0.842 0.053

294 0.837 0.049

1000 0.760 0.048


98
196 0.750 0.045

294 0.854 0.044

27
Table 2: Different properties for oily water at different operating parameters after
electrocoagulation
Parameter Permissible limits

Feed (without

water [33-

Discharged
D=0.01 cm

D=0.02 cm

wastewater
J=20 A/m2

J=50 A/m2

J=80 A/m2
salt)

Drinking
pH=3.6

pH=5.7

[33-40]
Property

40
Conductivity (mS/cm) 2.65 6.14 6.19 6.20 7.05 6.11 5.60 5.62 1.4 [WHO] 1.4 [WHO]
TDS (ppm) 3230 3130 3100 3070 3550 3040 2800 780 1000 [WHO] 2000 [WHO]
Salinity (ppt) 2.03 4.69 4.37 4.68 5.24 4.60 4.19 4.23 0.67 [WHO] 0.46 [WHO]
pH 8.72 8.80 8.80 8.15 9.70 9.04 7.20 7.59 8.5 [BIS] 9 [WHO]
D.O. (ppm) 93.7 26.1 22.4 16.7 12.0 8.5 18.6 11.2 15 [WHO] >5 [WHO]
Turbidity (NTU) 28.1 32.3 49.3 88.3 36.6 47.7 60.3 71.6 1 [WHO] 5 [WHO]
% Sodium (mg/L) 68.4 65 58 50 59 62 40 48 NA 60 [CPCB]
Oil & Grease (mg/L) 35 29.8 28 26.82 27.07 29 10.2 25.5 0.5 [BIS] 10 [CPCB]
Phenolics (mg/L) 12.2 10 7.5 - 5 10.8 1 2.8 0.002 [BIS] 1.2 [CPCB]
Cyanides (mg/L) 0.002 - - - - - - - 0.05 [BIS] 0.2 [CPCB]
Fluorides (mg/L) 0.40 - - - - - - - 1.5 [BIS] 1.5 [CPCB]
Sulphides (mg/L) 0.007 - - - - - - - 0.05 [BIS] 2 [CPCB]
Chromium (Hexa) 0.9 0.1 - - 0.3 0.45 0.1 0.3 0.1 [BIS] 0.1 [CPCB]
(mg/L)
Chromium (Total) 1.2 0.6 0.4 - 0.5 0.7 - - 1 [BIS] 1 [CPCB]
(mg/L)
Copper (mg/L) 0.17 - - - - - - - 0.2 [BIS] 0.2 [CPCB]
Lead (mg/L) 0.04 - - - - - - - 0.01 [BIS] 0.1 [CPCB]
Mercury (mg/L) - - - - - -- - - 0.001 [BIS] 0.01 [CPCB]
Nickel (mg/L) 0.36 - - - - - - - 0.02 [BIS] 3 [CPCB]

28
Table 3: Types of oil water treated using the electrocoagulation method
Type of oil water Initial feed concentration Reference
(mg/L)
Refectory oily wastewater 100-200 X. Xu and X. Zhu, 2004
Vegetable oil refinery industry wastewater 2648-3880 M. Basibuyuk and D.G.
Kalat, 2008
Palm oil mill effluent 250,000 N. Oswal et al., 2002
Vegetable oil refinery wastewater NA U.T. Un et al., 2009
Crude oil 500 Y.O. Fouad, 2012
Synthetic oil (lubricant oil (REPSOL ELITE TDI 1500-6000 P. Canizares, 2008
15W40) + soluble oil (SOL 1000))
Synthetic oil (paraffin oil+Oleic acid+ 200 U.B. Ogutveren and S.
Triethanolamine) Koparal, 2008
Restaurant oil water in Hong Kong 100-2100 X. Chen et al., 2000
Drilling wastewater from Barekuri, India 35 Present work

29
Fig.1: Setup used for Electrocoagulation-Microfiltration process

Fig.2: Variation of concentration with time at different operating parameters of (a) Current
density (b) Electrode distance and (c) Initial pH

30
Fig.3: Effect of current density, pH and electrode distance on the (a) Chloride concentration (b)
TSS concentration and (c) Sulphate concentration

100 0.022
100 0.025 (b)
(a)
0.020
90
0.020
80 0.018
Current density (A/m2)

Electrode distance (m)


80
Current density (A/m2)

Electrode distance (m)

0.016
70
0.015 60 0.014
60
0.012
50 0.010 40 0.010
40
0.008
30 0.005 20
0.006
20 0.004
0.000
20 25 30 35 40 45 50 55 60 65 70 75 2.5 3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.5
Corrosion (mg) Film thickness (m)

31
0.30 3
Total cost (US ) (c)
3
Electrode cost (US$/m )
3
0.25 Energy cost (US$/m )
Operating Parameters:
Volume: 600 ml

Operating cost(US$/m )
0.20 Time: 20 min

3
0.15

0.10

0.05

0.00
30 45 60 75 90
2
Current density(A/m )

Fig. 4: Effect of current density and electrode distance on (a) Corrosion (b) Film thickness and
(c) Operating cost variation with current density

32
60 28
(a)
24
Before filtration
After filtration
20

Lp (m/Pa.s) 109
16

50 12

4
Porosity (%)

0
700 750 800 850 900 950 1000 1050
40 Sintering temperature (oC)
30

25 Water permeation experiment


ImageJ software
20
Pore size (m)

30 15

10

20 0

700 750 800 850 900 950 1000 1050


Sintering temperature (oC)

700 750 800 850 900 950 1000 1050


Sintering temperature (oC)

33
600
(b) Membrane sintering temperature (oC)
700 800 900 1000
550

500
Flux (L/h.m2)

450

400

350

300

250
100 150 200 250 300 350
Pressure (kPa)

Fig. 5: (a) Porosity of membranes at different sintering temperatures, inset (top): Variation of
hydraulic permeability (Lp) at various sintering temperatures, inset (bottom): Variation of
average pore size obtained from FESEM images and water permeation test for different
membranes (b) Variation of permeate flux with pressure for MF membranes sintered at different
temperatures

34
13000
Membrane sintering temperature (oC) (a)
11700 700 800 900 1000

10400

9100
Flux (L/h.m2)

7800

6500 6000
Membrane sintering temperature (oC)
700 800 900 1000
5000
5200
4000

Flux (L/h.m2)
3900
3000

2600 2000

1000
1300 100 150 200 250 300 350
Pressure (kPa)

100 150 200 250 300 350


Pressure (kPa)

35
900
(b) Ceramic membrane prepared at 700oC
800 Transmembrane pressure: 98 kPa

700
Flux (L/h.m2)

600

500

400

300

200
0 5 10 15 20 25
Time (min)

Fig. 6: Pure water flux of the membranes at different sintering temperatures (a) Before
microfiltration of EC sample, inset: After microfiltration of EC sample (b) Flux declination
pattern during MF.

36
7
(a) 2
Current density (A/m )
6 20 50 80 (b)
pH
5 3.6 5.7
Intensity (%)

-1
0.1 1 10 100 1000
Flock size (m)

(c)
(111) (d)
(2 0 0)
(2 0 0)
1000C
Intensity (a.u.)

950C

900C
(2 2 2)

850C

10 20 30 40 50 60 70 80
Diffraction angle (o)

(e) (f)

Fig.7: (a) XRD figure of membrane at different sintering temperature (b) EDX analysis of flocks
filtered by membrane (c) Particle size distribution of flocks produced during EC process (d)
FESEM image of membrane after microfiltration (e) Flocks retained by ceramic membrane (f)
FESEM image of dried flocs

37
Highlights:
1. Treatment of oily wastewater was performed utilizing electrocoagulation-microfiltration in
batch mode.
2. Pretreatment was performed using electrocoagulation with working parameters of current
density (20 - 80 A/m2), electrode distance (0.005 - 0.2 m) and initial pH (3.6 - 8.7).
3. 75 - 85 % of the initial flux was lost during the microfiltration of the electro-coagulated
samples.
4. Oil and grease contents were lowered to 10.2 mg/L.

38

Anda mungkin juga menyukai