You are on page 1of 10

Desalination 465 (2019) 94–103

Contents lists available at ScienceDirect

Desalination
journal homepage: www.elsevier.com/locate/desal

Anion exchange membrane organic fouling and mitigation in salt T


valorization process from high salinity textile wastewater by bipolar
membrane electrodialysis
Yifru Waktole Berkessaa,b, Qiaolin Langa,b, Binghua Yana,b, Shaoping Kuanga, Debin Maob,
Li Shuc, Yang Zhanga,b,

a
College of Environment and Safety Engineering, Qingdao University of Science and Technology, Qingdao 266042, China
b
Waste Valorization and Water Reuse Group, Qingdao Institute of Bioenergy and Bioprocess Technology, Chinese Academy of Sciences, 189 Songling Road, Laoshan
District, Qingdao 266101, China
c
School of Engineering, RMIT University, Melbourne, VIC 3000, Australia

GRAPHICAL ABSTRACT

ARTICLE INFO ABSTRACT

Keywords: To achieve a cleaner production, textile wastewater with high organic and salt content can be treated by using
Critical salt concentration Bipolar Membrane Electrodialysis (BMED) to minimize acid and base consumption in a dyeing process. While the
Zeta potential dye molecules may foul the ion exchange membranes and strongly affect the desalination process. This work
Membrane fouling aimed to investigate the performance and fouling mechanisms of BMED during desalination of sodium sulfate
Charged foulants
from Remazol Brilliant Blue R (RBBR). Results showed that maintaining the zeta potential of RBBR above
Bipolar membrane electrodialysis
Textile wastewater
−25 mV may mitigate fouling of AEM during the BMED process. This confirms that, zeta potential of charged
foulants (RBBR) plays a key role in terms of controlling membrane fouling. Accordingly, a new parameter
“critical salt concentration” was introduced to control membrane fouling. Furthermore, energy dispersive X-ray
spectroscopy (EDS), FT-IR and electrochemical analysis confirmed that fouling of anion exchange membrane by
RBBR was due to electrostatic interaction. Finally, it was calculated that 72.02% of sodium and 66.9% of sulfate
in the feed were converted to NaOH and H2SO4, respectively. This study proves that BMED process may be an
alternative way treating textile wastewater with high salinity and the presence of dye molecules.


Corresponding author at: College of Environment and Safety Engineering, Qingdao University of Science and Technology, Qingdao 266042, China.
E-mail address: zhangyang@qibebt.ac.cn (Y. Zhang).

https://doi.org/10.1016/j.desal.2019.04.027
Received 13 November 2018; Received in revised form 7 March 2019; Accepted 23 April 2019
Available online 08 May 2019
0011-9164/ © 2019 Published by Elsevier B.V.
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

1. Introduction

The water consumption of textile industry is estimated to be be-


tween 200 and 500 m3 per ton of dyed textile product [1]. Most im-
portantly, 90% of the water consumption in textile industry goes to
processing, [2] which results in a large volume of wastewater that
mainly consists of waste dye as high as 1.5 g/L [3] and high salinity of
5–6% NaCl and 5% Na2SO4 [1,3]. A group of dyes called anthraquinone Fig. 1. Molecular structure of RBBR with molecular formula
dyes are known to have low fixation rate of ~50%, which leads to high C22H16N2Na2O11S3 (625.5 Da).
dye concentration in the effluent from textile industry [4]. The low
fixation rate of anthraquinone dyes in general and Remazol Brilliant concentration, feed surface velocity and RBBR concentration were as-
Blue R (RBBR) in particular demands addition of even higher salt sessed. Furthermore, fouling characteristics of RBBR and its interaction
concentration to enhance dye fixation rate, this results in a higher with AEM in terms of zeta potential, electrochemical impedance spec-
salinity in the effluent [2]. Direct discharge of effluent from textile troscopy (EIS), membrane resistance, and on line stack voltage mon-
industry is detrimental to the aquatic environment due to its impact on itoring were undertaken to study the fouling mechanisms in detail. As a
photosynthetic activity [5], bioaccumulation [6] and adsorption to result, a concept of “Critical Salt Concentration” was proposed to il-
microbes resulting in reaction with biologically important molecules lustrate the correlation between salt concentration and membrane
[7]. Thus, the challenge related to effluent from textile industry needs fouling tendency caused by charged organic compounds in BMED
to be addressed in view of sustainable production of textile without process. This concept is expected to provide a guideline to desalinate
polluting the environment. water containing charged organic molecules by electromembrane pro-
Conventional textile wastewater treatment methods focus on the cesses in various industrial applications.
removal and decolorization of dye molecules by using biological
treatment, advanced oxidation based processes, adsorption or coagu- 2. Method and material
lation [8–12]. Besides the aforementioned conventional treatment
methods, recent works revealed the possibility of valorizing textile 2.1. Chemicals and instruments
wastewater by standalone as well as integrated methods, e.g. tight ul-
trafiltration [13] and electrodialysis (ED) [14], nanofiltration in- Remazol Brilliant Blue R (~50% pure) was purchased from
tegrated with bipolar membrane electrodialysis [1] and reverse os- Shanghai Yuanye Bio-Technology Co., Ltd (Fig. 1). The ion exchange
mosis-advanced oxidation-bipolar membrane electrodialysis [15]. ED membranes were purchased from Hefei Chemjoy Co., Ltd., China.
was reported to be effective in desalinating sodium chloride from dye Analytical grade anhydrous sodium sulfate (99% pure), NaOH and
wastewater, yet a high energy requirement is needed as a pre-con- H2SO4 were bought from Sinopharm Chemical Reagent Co., Ltd.
centration method (to 150–200 g/L NaCl) prior to evaporation and
crystallization process to recover or dispose solid salt. In addition, solid 2.2. BMED set-up
salt disposal is not a preferable way in view of its disposal expenditures
and potential environmental impacts. BMED offers special advantage The electrodialysis stack used in this work was manufactured by
over ED since it simultaneously converts salt to acid and base, this Hefei Chemjoy Co., Ltd., China. Heterogeneous bipolar membrane
suggesting an economic value of BMED for desalination of high salinity (BPM) was manufactured by MEGA A.S., Czech Republic. The BPM had
textile wastewater. Moreover, the converted acid and base may be re- a thickness of 500 μm. Two triplets of membranes were used during all
cycled in the dyeing process and the waste salt disposal can be avoided. experiment runs with three chambers; namely, acid, feed (diluate) and
However, ion exchange membrane fouling caused by dye molecules base chambers as indicated in Fig. 2. Three, two and two pieces of
is still a main obstacle and the properties and functionalities of the heterogeneous bipolar membranes (BPM), cation exchange membranes
membranes can be shifted due to electrostatic interaction of these (CEM) and anion exchange membranes (AEM) were installed in the
charged dye molecules with membranes [14]. Furthermore, literature stack, respectively. The effective area of each membrane was 94.5cm2.
report revealed aromatic compounds may cause severe organic fouling The ED stack consisted of electrode materials made of titanium coated
of ion exchange membranes due to hydrophobic interaction [15–17]. It with iridium. The thickness of the polypropylene spacer used in the
is reported that acid blue-19 fouled anion exchange membrane (AEM) stack was 0.79 mm. Characteristics of CEMs and AEMs provided by the
showed selectivity to monovalent ions over divalent ions during desa- manufacturer were presented in the Table 1.
lination of mixture of NaCl, Na2SO4 and acid blue-19 using electro-
dialysis [18]. The selectivity developed due to organic fouling of AEM 2.3. Evaluation of BMED performance
dictates the difficulty of desalinating divalent ions from textile waste-
water mainly when the effluent contains mixture of monovalent and 2.3.1. Desalination rate
divalent salts. Likewise, Lee et al. [16] reported negatively charged The desalination ratio of ions during BMED experiment was calcu-
foulants with high adsorption capacity can significantly affect the lated:
performance of electrodialysis desalination process by forming irre-
versible fouling. Studies on fouling behavior of AEM rely on using Ci Cf
Desalination ratio (%) = × 100
standard foulants such as Bovine Serum Albumin (BSA), humate and Ci (1)
Sodium Dodecyl Benzene Sulfonate (SDBS) [16,19,20], and aromatic or where Ci and Cf refer to initial and final salt concentration, respectively.
aliphatic organic foulants with various alkyl chain lengths [21].
By far, fouling mechanism of AEM in view of desalinating high 2.3.2. Current efficiency (η)
salinity textile wastewater using bipolar membrane electrodialysis Current efficiency of the BMED process was calculated according to
(BMED) has not been given due attention. In this study, production of the following equation [23].
sulfuric acid and sodium hydroxide from direct desalination and con-
ZFV C
version of salt in the saline textile wastewater were performed. = × 100
NIt (2)
Concomitantly, fouling mechanism of the anion exchange membrane
−1
caused by negatively charged dye molecules (RBBR) was investigated. where η: current efficiency, F: Faraday constant (A-s∙mole ), z: charge
Factors that affect the membrane fouling such as feed pH, salt of the ion, V: initial volume (L), ∆C: change in diluate concentration

95
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

Fig. 2. Schematic diagram of the BMED stack configuration used in this study.

Table 1 2.6. Zeta potential of RBBR molecules


Properties of the cation exchange and anion exchange membranes.
Parameters CEM AEM
The zeta potential of RBBR molecule was determined using dynamic
light scattering technique by Zetasizer series Nano ZS from Malvern
Thickness (μm) 180–210 140–160 Instruments, UK. The samples were prepared by dissolving 0.25 g/L
Ion exchange capacity (meq/g) 0.8–1.0 0.5–0.6 RBBR and varying concentrations of Na2 SO4 (0%, 0.5%, 1%, 5% and
Water content (%) 40–50 15–20
10%) were dissolved in ultrapure water. The pH of RBBR synthetic
Blasting strength (MPa) > 0.35 > 0.35
Membrane resistance (Ωcm2) 2.5–3.5 3.5–4.5 wastewater samples were adjusted to the required pH by 0.1 M HCl and
Ion transport number > 0.93 > 0.93 0.1 M NaOH. Furthermore, water was used as dispersant during mea-
surement of RBBR zeta potential and the viscosity of RBBR solution was
0.887. The refractive index and dielectric constants of the dispersant
(mol∙L−1), N: number of cell triplets, I: applied current (A) and t: time were 1.33 and 78.5, respectively. Disposable folded capillary cells
(s). (DTS1070) were injected with 1 mL samples and installed on Nano ZS.
The zeta potential measurements were undertaken at room temperature
(25 °C) where triplicate measurements were made and average values
2.4. Nyquist plot of clean and fouled membrane were reported.

The alternating current (AC) impedance measurements were carried 2.7. Analytical methods
out in a two chamber electrodialytic cell and separated by the measured
membrane with effective area of 7.1 cm2. Flow rate of 60 mL/min was The conductivity and pH of samples at all stages of the experiment
applied to circulate solutions in the chambers. Two Ag/AgCl planar were measured by conductivity meter (FG3-ELK, Mettler-Toledo
electrodes with projected area of 1.0 cm2 were used as reference elec- Instruments Co. Ltd., USA) and pH meter (FG2, Mettler-Toledo
trodes, fixed near to IEM surface. Sinusoidal AC were supplied by an Instruments Co. Ltd., USA), respectively. Measurement of direct current
electrochemical workstation (CHI660E, CH Instrument Co., LTD, supply was recorded using DC power source (MCH-k305D, Shenzhen,
Shanghai, China), with amplitude of 10 mV and frequency range of China). Measurement of voltage difference between electrodes during
10−1-105 Hz. 0.5 M NaCl solution was used as electrolyte. resistance measurement was recorded using digital multimeters (model:
UT39A/B/C, UNI-T technology Co., Ltd., Donggua, China). The ion
transport number of AEM samples was measured using electrochemical
2.5. Ion transport number measurement work station from CH instruments (CHI660E, Shanghai, China) and ion
exchange membrane (IEM) resistances were measured using Chemjoy
The ion transport through ion exchange membranes in this work resistance cell (CJ-AMD-01, Chemjoy Polymer Materials Co., Ltd.,
was calculated using the following formula [24]: Hefei, China).
The concentrations of sulfuric acid and sodium hydroxide produced
RT C1
Em = (2t 1) ln from BMED experiment were determined using titration method. 0.1 M
F C2 (3)
NaOH and 0.1 M H2SO4 were used as titrant solutions to determine the
where, Em = voltage measured across the membrane. R, T and F were concentrations of sulfuric acid and sodium hydroxide at stoppage time
molar gas constant, temperature and Faraday's constant. of BMED experiment, respectively.
Scanning electron microscopy (SEM) technique was used to quali-
Em = E E0 (4) tatively characterize foulant layer on fouled AEM samples. Pristine and
fouled membrane samples were prepared before SEM imaging by
where, E = voltage measured between two concentrations of solution coating with a thin layer of gold in order to reduce membrane surface
(0.01 M and 0.05 M Na2SO4) and E0 was the voltage measured in the charge. SEM images of the membrane surface were taken at voltage of
same concentration of solution (0.01 M). 5 kV using the Hitachi S-4800, Japan. Moreover, Energy-dispersive x-

96
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

Fig. 3. Effect of RBBR concentration on BMED performance.

ray spectroscopy (EDS) was used to quantitatively determine the atom also found that RBBR did not present in the acid or base streams, this
composition on the pristine and fouled AEM samples at an approximate indicates that RBBR molecules (MW 625.5 Da) could not penetrate
penetration depth of 2-5 μm. Furthermore, the expected changes in through the anion/cation exchange membranes.
functional groups on pristine and fouled AEMs were analyzed using Fig. 3c shows the correlations between the conductivity of the final
Fourier transform infrared spectroscopy (FTIR) at a penetration depth diluate and RBBR concentration (5% Na2SO4 in the initial feed). A
of 0.8 μm (Thermo Scientific™ Nicolet™ iN™10, China). linear correlation (R2 = 0.99) between the RBBR concentration and the
corresponding conductivity of the final diluate was found in Fig. 3c.
This could be attributed to increased intensity of forming fouling layer
3. Result and discussion caused by accumulation of RBBR onto the membrane surface corre-
sponding to increased RBBR in the initial feed. In other words, the final
3.1. Desalination performance of BMED salt concentration in the diluate that can be achieved by BMED is a
function of the RBBR concentration in the feed. Consequently, a con-
3.1.1. Effect of RBBR concentration on BMED performance cept of “critical salt concentration” is proposed to correlate the final salt
RBBR concentration was found to adversely affect BMED perfor- concentration in BMED and the RBBR concentration in the feed. “Cri-
mance, where desalination performance for the same initial Na2SO4 tical salt concentration” refers to the salt concentration below which
concentration (5%) under 3A applied current decreased from 97.7% to desalination cannot proceed as a result of organic fouling by charged
70% (based on conductivity decrease in feed chamber) with increasing organic molecules.
RBBR concentration from 0 g/L to 1 g/L, as presented in Fig. 3a. This
can be explained by the facts that RBBR molecules are negatively
charged and also the aromatic structure interacts with AEM through 3.1.2. Effect of anion species and salt concentration
hydrophobic interaction [18]. Apart from the decreasing desalination Low salt concentration (< 0.5 M) was found to increase membrane
performance, with increasing RBBR concentration from 0 to 0.25 g/L stack resistance and in turn cause transport limitation [25]. As seen in
the operation time increased from 3.5 h to 4.0 h as indicted in Fig. 3a. Fig. 4a, the decrease in conductivity of diluate for NaCl and Na2SO4 as a
This is due to the fact that with the increase in dye concentration, a function of time showed a similar trend and endpoint (~3.5mS∙cm−1)
higher stack resistance was built up (Fig. 3b) due to increasingly in view of the diluate conductivity. On the other hand, a similar end-
dominant effect of organic fouling. It forced the system shutdown after point (~3.5mS∙cm−1) is also found when different initial salt con-
3.5 h operation since the voltage reached 30 V. The effect of RBBR centration (1% and 5% Na2SO4) was experimented, as seen in Fig. 4c.
concentration (0 g/L, 0.25 g/L and 1 g/L) on BMED desalination per- This may indicate that AEM fouling by RBBR was mainly influenced by
formance revealed that the final diluate conductivity was adversely the diluate conductivity, while the fouling tendency is regardless of
affected by the RBBR concentrations. 0.78mS/cm, 3.69mS/cm and anion spices (chloride or sulfate). The dominant effect of AEM fouling
13.02mS/cm of the final diluate were obtained with the above men- by RBBR at low salt concentration could be attributed to the decreased
tioned RBBR concentrations in the respective order at stoppage time of shielding effect of salt on RBBR allowing increased interaction with
BMED process where the stack voltage reached 30 V (Fig. 3a & b). It was AEM.

97
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

Fig. 4. Effect of initial salt concentration on BMED performance. a and c, conductivity variation with operation time. b and d, voltage profile with operation time.
The applied current for a and b was 1A, I = 3A for c and d.

This work however contradicts with a previous report where a dif- destabilization of the RBBR fouling layer and the concentration polar-
ferent fouling tendency was found on AEM fouling by acid blue-9 with ization layer at higher flow rates. Hence, 8.35 cm/s was chosen as op-
the presence of chloride over sulfate (21 and 36 mmol∙L−1, respec- timal feed surface flow velocity to run all experiments in this work.
tively) [14]. The reason for the difference may attribute to the fact that
the feed conductivity was well below the “critical salt concentration”,
therefore the negatively charged fouling layer formed by blue-9 im- 3.1.4. Effect of feed pH
proved the monovalent anion selectivity. The effect of feed pH on BMED performance revealed that relatively
On the other hand, voltage was recorded to show the evolution of low stack voltage and stable operation were achieved at lower pH
stack resistance as a function of time during the experiment, as seen in (pH = 3.0) compared to pH 7.0 and 11.0, as shown in Fig. 6. The lower
Fig. 4b. The fact that ion transport was hindered by membrane fouling stack resistance at pH 3.0 may be attributed to protonation of RBBR
can be further confirmed by comparing Fig. 3b (5% Na2SO4 with 0 g/L molecules depicting less interaction with AEM and hence low fouling
RBBR) and Fig. 4b (5% Na2SO4 with 0.25 g/L RBBR) where the stack tendency compared to pHs 7.0 and 11.0 (Fig. 6a). Moreover, there was
voltage was stable until the membrane was fouled by RBBR when the proton leakage from acid chamber to feed chamber which might have
salt concentration was lowered. This resulted in an insufficient desali- resulted in opening of ion transport channels blocked by dye molecules
nation due to the stack voltage increased above 30 V, as seen in Fig. 4 and/or dissolving the RBBR adsorbed on the membrane surface.
(5% Na2SO4 with 0.25 g/L RBBR). On the other hand, the surface charge of RBBR molecules becomes
increasingly negative at pH 7.0 and 11.0 due to less protonation at
3.1.3. Effect of surface flow velocity higher pH as opposed to pH 3. The increasingly negative surface charge
Concentration polarization in the diluate chamber of BMED is of RBBR could explain the high stack voltages for pH 7.0 and 11.0
controlled by diffusion and migration [26]. The development of con- which arise from stronger electrostatic and hydrophobic interactions
centration polarization and affinity of RBBR to AEM may have sy- between RBBR and AEM compared to similar phenomenon at pH 3 (the
nergistically acted together to increase resistance to permeation of above mentioned results will be further quantitatively analyzed and
sulfate especially at low salt concentration. Hence, feed surface flow explained in Section 3.2). Further, high pH is well known to facilitate
velocity was assessed to investigate its impact on the performance of gel formation and/or precipitation which might in this case relate to
BMED. The result reveled that low feed surface flow velocity (4.17 cm/ suppression of proton effect due to continuous addition of hydroxide
s) registered the lowest performance compared to the other two higher ion to maintain high pH (7.0 and 11.0). In conclusion, better BMED
surface flow velocities (8.35 cm/s and 12.52 cm/s) as presented in performance was observed at pH 3 compared to pH 7 and 11 due to less
Fig. 5a. Moreover, the electrochemical stack resistance presented as interaction of RBBR with AEM at pH 3 and hence showed lower fouling
voltage profile also proved higher stack voltage at lower feed surface tendency.
flow velocities and vice versa (Fig. 5b). In conclusion, a higher flow rate
was found to reduce the stack resistance which may attribute to

98
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

Fig. 5. Effect of feed flow rate on performance of BMED during desalination of sodium sulfate from RBBR based synthetic wastewater (Na2SO4 concentration in feed
chamber was 5% and the applied current was 3A).

3.2. Zeta potential of RBBR (Fig. 7b). Thus, the effect of pH on zeta potential of RBBR was sig-
nificant when concentration of Na2SO4 decreased from 5% to 0.5% at
Membrane fouling mechanisms in electromembrane processes pre- similar pH (pH = 3) where the corresponding zeta potential values
dominantly depends on electrostatic interaction between the IEMs and were −18.9 and −35 mV, respectively. According to the membrane
foulants [30]. Hence, zeta potential measurement of RBBR molecules fouling results from Section 3.1.2 and Section 3.1.4, membrane fouling
was undertaken to understand the fouling phenomena during BMED can be mitigated under a higher salt concentration or/and a lower pH.
process. Fig. 7a and b show the trends of RBBR zeta potential as a These findings may be further linked with the above mentioned zeta
function of salt concentration at constant pH and pH effect on zeta potential measurement of RBBR and a “critical zeta potential” (around
potential at constant salt concentration, respectively. The zeta potential -25 mV) can be drawn in Fig. 7 to manipulate membrane fouling from
of RBBR was found to become increasingly negative (−16.75 mV, engineering aspect.
−23.1 mV, −33.7 mV and −58.3 mV) corresponding to decreasing of Lee et al. [16] reported the zeta potentials of standard organic
Na2SO4 concentration (10%, 5%, 1%and 0%) at 0.25 g∙L−1 RBBR and foulants as humate, BSA and SDBS (measured in 10 mM KCl at constant
constant pH = 3 as indicated in Fig. 7a. The increasingly negative zeta pH values of 6.5 ± 0.3 and room temperature) were − 8.2 ± 3.1 mV,
potential of RBBR with corresponding decrease of Na2SO4 concentra- −25.3 ± 4.5 mV, and − 26.9 ± 5.3 mV, respectively. Nevertheless,
tion might be due to the decreased shielding effect of Na2SO4 on RBBR the zeta potential of RBBR measured in 35 mM Na2SO4 at pH 7 was
molecules. Further, the increasingly negative zeta potential of RBBR at −38.75 mV, which was more negative compared to the standard fou-
low Na2SO4 concentration may explain the high stack resistance of lants mentioned above. This justifying the possibility of stronger elec-
desalinating Na2SO4 from RBBR at low salt concentration due to the trostatic interaction between RBBR and AEM. Thus, it can be concluded
dominant effect of AEM fouling by RBBR as indicated in Fig. 7a. that initial salt concentration played a vital role in controlling the
On the other hand, the effect of feed solution pH on zeta potential of fouling behavior RBBR towards AEM. Furthermore, it can be further
RBBR was presented in Fig. 7b. The result revealed that effect of pH on induced that when the zeta potential of RBBR was above the critical
zeta potential was minimal at 5% Na2SO4 where zeta potential of RBBR zeta potential (−25 mV), the likelihood of AEM fouling can be miti-
showed slight change from −15.2 mV to −19.2 mV corresponding to gated. Hence, it is recommended to maintain the zeta potential of RBBR
change in pH from 1 to 11, respectively (Fig. 7b). On the contrary, above −25 mV to control AEM fouling and improve BMED perfor-
significant change in zeta potential of RBBR was observed at 0.5% mance in industrial application of this work.
Na2SO4 where the zeta potential of RBBR decreased from −18.83 mV
to -42 mV corresponding to pH increase from 1 to 11, respectively

Fig. 6. Effect of feed pH on BMED performance, (a) conductivity profile of feed chamber; (b) stack voltage profile (Na2SO4 concentration in feed chamber was 5%
and concentration of RBBR was 0.25 g/L, I = 3A).

99
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

-10
(a) (b)
-20
Zeta potential (mV)
-30
pH effect at 5% Na2SO4
Zeta potential of RBBR pH effect at 0.5% Na2SO4
-40

-50

-60

0 20 40 60 80 100 0 2 4 6 8 10 12
-1
Concentration of Na2SO4 (g L ) pH

Fig. 7. Zeta potential of RBBR a, Effect of Na2SO4 concentration on RBBR zeta potential (pH = 3 and RBBR concentration = 0.25 g/L) b, Effect of pH on RBBR zeta
potential at different Na2SO4 concentration. Red dash line: “critical zeta potential”.

3.3. Fouling mechanism and the effect on desalination On the other hand, the influence of RBBR concentration to the de-
salination of sulfate is schematically presented in Fig. 8b. As reported in
According to the above presented results, schematic diagrams of Section 3.1.1, a higher RBBR concentration resulted in a higher sulfate
membrane fouling mechanism are presented in Fig. 8a & b. As illu- rejection due to severe fouling of AEM, and this leaded to a higher stack
strated in Fig. 8a, permeation of sulfate and retention of RBBR are in- resistance (Fig. 3b). The increasing fouling tendency as a function of
fluenced by the feed pH under 5% of Na2SO4 and 0.25% of RBBR. The increasing RBBR concentration brought about increased salt con-
increase in sulfate rejection (leads to a high stack resistance) from pH 3 centration in the final diluate. The final salt concentration that varied
to 11 is attributed to the formation of a negatively charged layer as the with RBBR concentration was named “critical salt concentration”. The
increased affinity of AEM to RBBR molecules through electrostatic in- critical salt concentration was found to have a linear (first order) cor-
teractions. This mechanism can be confirmed by Section 3.1.4 and relation with RBBR concentration (Figs. 3c & 8b). This first order cor-
Fig. 6b, the stack voltage (represents the stack resistance) was sig- relation is inferred to be induced by the accumulation of RBBR mole-
nificantly higher when the feed pH was 11 than that under a lower pH. cules with a “gel layer” formation on the anion exchange membrane

Fig. 8. Schematics of RBBR fouling mechanism with the presence of sulfate ion as a function of (a) pH of the feed and (b) RBBR concentration in the feed.

100
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

Table 2 Table 3
Ion transport number of the pristine and fouled membranes. Parameters from Nyquist plot for clean and fouled AEM.
Membrane type Transport number Parameters Pristine AEM Fouled AEM

2
Pristine AEM 0.99 ± 0.04 Rs (ohm·cm ) 26.92 27.51
Fouled AEM 0.85 ± 0.04 Rt (ohm·cm2) 24.17 29.66
Wo (S·sec5·cm−2) 0.001575 0.001672

surface. The gel layer thickness was only dependent on the RBBR
concentration and thus resulted in a proportional relationship with the experiment. Likewise, the heterogeneous transfer resistance (Rt) ob-
sulfate transport and the final salt concentration. viously increased for fouled AEM. Corresponding to Nyquist plot re-
sults, this suggests that the increase in resistance of fouled AEM was not
only due to trans-membrane resistance but also due to contribution
3.4. Ion transport number and electrochemical analysis
from heterogeneous transfer of ions. Thus, it can be concluded that the
ions transfer was affected by the fouling layer both from the resistance
Ion transport number was measured on pristine and fouled AEMs to
introduced by fouling layer and from the effect on AEM surface struc-
evaluate fouling severity by RBBR. The result showed that ion transport
ture.
number decreased from 0.99 ± 0.04 to 0.85 ± 0.04 after RBBR
fouling (Table 2). Lee et al. [16] reported organic fouling of AEM by
humate, BSA and SDBS resulted in decrease of ion transport number 3.5. Membrane surface fouling analysis
from 0.96 for pristine membrane to 0.94, 0.93 and 0.92 for fouled
membranes, respectively. The decrease in ion transport number after SEM, EDS and FTIR spectroscopy were used to characterize AEM
the AEM got fouled by RBBR was found to be more significant com- before and after BMED experiment. As shown in Fig. 10b, AEM surface
pared to the literature reports. This may be attributed to strong elec- was covered and exhibited less roughness after the experiment, when it
trostatic interaction between AEM and RBBR resulting in high mem- is compared with Fig. 10a (pristine membrane). EDS analysis of pristine
brane resistance and low transport number, owing to a higher negative and fouled AEM showed clear variation of the different elements de-
zeta potential of RBBR at the operational conditions. tected on the membrane surface. In particular, the atom percentage of
Besides, electrochemical analysis was conducted and Nyquist plots nitrogen increased from 15.05% in pristine AEM to 17.97% in fouled
of the pristine and fouled membrane samples are shown in Fig. 9. Both AEM justifying adsorption of RBBR molecules to AEM surface. Fur-
the pristine and fouled AEM showed typical Nyquist curves of IEM. The thermore, sulfur which was not detected on EDS analysis of pristine
cross point with real axis in high frequency region stands for the sum of membrane appeared for the fouled AEM. Accordingly, about 0.27%
solution and trans-membrane resistance. On the other hand, the semi- (atom percent) sulfur was detected on the fouled membrane indicating
circle in middle frequency region shows heterogeneous transfer im- adsorption of RBBR to the AEM.
pedance from membrane surface to ion concentration polarization As seen in Fig. 10e and f, the FTIR spectra of pristine and fouled
layer. Further, the slope line in low frequency region corresponds to the AEMs showed broad bands at 3406.36 and 3421.85 which is the
sum of solution and trans-membrane resistance. The later stands for stretching vibration of NeH or OeH, respectively [28,29]. The NeH
Warburg impedance as indicated in Fig. 9. Compared to pristine stretching vibration represents NeH group joining aromatic structures
membrane, the semicircle of fouled AEM was relatively bigger both in in RBBR. Interestingly, unique peaks appeared for fouled AEM at
real and imaginary direction, dictating the fouling layer increased the 1570.65 and 1538.18 related to skeletal vibration of aromatic structure
resistance and capacitive reactance of fouled AEM during hetero- that probably emanate from RBBR [29]. Furthermore, a new peak was
geneous transfer process of ions. observed at 1388.99 that associated to CH3 symmetrical deformation
Furthermore, parameters obtained from fitting Nyquist plots de- mode following fouling of AEM [28]. Apart from the above mentioned
picted the sum of solution (Rs) and trans-membrane resistance (Rt) had peaks the FTIR spectra of both pristine and fouled AEM samples were
a slight increase for fouled AEM (Table. 3). The small increase in the nearly similar. Thus, the appearance of new peaks on FTIR spectra re-
sum of solution and trans-membrane resistance might be related to lated to NeH and skeletal vibration of aromatic structure confirms
application of alternating current signal during measurement process of fouling of AEM by RBBR.
Nyquist plot instead of direct current as in the case of BMED

3.6. Current efficiency and acid-base production

Accordingly, BMED desalination performance for the same initial


Na2SO4 concentration (5%) decreased from 98.30% to 74.43% with
increasing RBBR concentration from 0 g/L to 1 g/L, as presented in
Fig. 3c. Likewise, the corresponding current efficiency decreased from
44.16% to 35% when the RBBR concentration was increased from 0 g/L
to 0.25 g/L. However, the current efficiency was 39.01% when the
RBBR concentration was further increased to 1 g/L. The higher current
efficiency at the highest RBBR concentration may be attributed to early
shutdown of BMED after 3.5 h operation due to severe fouling.
The lower current efficiency reported in this work is due to the
following reasons: (i) shunt voltage on the electrodes due to redox re-
duction; (ii) proton and hydroxyl leakage through AEMs and CEMs
thereby increasing energy consumption; (iii) fouling of AEM by RBBR
resulting in increased resistance to transport of sulfate.
On the other hand, the balanced equation representing desalination
of sodium sulfate and acid-base production from RBBR wastewater is
Fig. 9. Nyquist plots of clean and fouled IEM. presented below:

101
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

Fig. 10. SEM images (a) & (b), EDS (c) & (d), and FTIR (e) & (f) results of pristine and fouled AEM, respectively.

Na2 SO4(aq) + 2H2 O(l)


Electrical energy
2NaOH(aq) + H2 SO4(aq) the acid chamber to feed chamber through the AEM.
Neutralization reaction (5)
4. Conclusion
Accordingly, theoretical yield of NaOH and H2SO4 from 50 g/L
Na2SO4 (0.352 mol) is expected to be a maximum of 0.704 mol of NaOH This work presents an investigation on membrane fouling me-
and 0.352 mol of H2SO4. The results obtained from BMED experiment chanisms of a medium size (625.5 Da) negatively charged organic
were 0.507 M NaOH and 0.236 M H2SO4. Therefore, the current effi- compound (RBBR) during BMED process. Zeta potential of RBBR under
ciencies to produce the base and acid were 72.02% and 66.9%, re- different pH and salt concentration was correlated with AEM fouling
spectively, with the desalination rate of 74.43%. from 5% Na2SO4 tendency. Specifically, maintaining the zeta potential of RBBR above
(under 0.25 g/L RBBR and 3A applied current). The unbalanced yield of -25 mV may mitigate fouling of AEM during the BMED process. This
H2SO4 compared to NaOH was due to continuous proton leakage from proves that, zeta potential of a charged foulant (RBBR) plays a key role

102
Y.W. Berkessa, et al. Desalination 465 (2019) 94–103

in terms of controlling membrane fouling. statistical investigation of some electrochemical variables, Waste Manag. 22 (2002)
Apart from this, a linear relationship between feed RBBR con- 491–499.
[6] G.A. Borges, L.P. Silva, J.A. Penido, L.R. de Lemos, A.B. Mageste, G.D. Rodrigues, A
centration and final diluate salt concentration was found. A new method for dye extraction using an aqueous two-phase system: effect of co-occur-
parameter “critical salt concentration” was introduced according to this rence of contaminants in textile industry wastewater, J. Environ. Manag. 183
correlation. More specifically, the critical salt concentration was found (2016) 196–203.
[7] I.M. Banat, P. Nigam, D. Singh, R. Marchant, Microbial decolorization of textile-
to be 0.78mS/cm, 3.6mS/cm and 13.02mS/cm at RBBR concentrations dye-containing effluents: a review, Bioresour. Technol. 58 (1996) 217–227.
of 0 g/L, 0.25 g/L and 1 g/L, respectively. This new parameter may [8] A. Asghar, A.A.A. Raman, W.M.A.W. Daud, Advanced oxidation processes for in-situ
provide a guidance for the prediction of desalination rate when medium production of hydrogen peroxide/hydroxyl radical for textile wastewater treatment:
a review, J. Clean. Prod. 87 (2015) 826–838.
size negatively charged foulants present in the feed. [9] E. Bazrafshan, M.R. Alipour, A.H. Mahvi, Textile wastewater treatment by appli-
Furthermore, the fouled AEM membranes were characterized by cation of combined chemical coagulation, electrocoagulation, and adsorption pro-
different methods in comparison with pristine ones. Energy dispersive cesses, Desalin. Water Treat. 57 (2016) 9203–9215.
[10] S. Jorfi, G. Barzegar, M. Ahmadi, R.D.C. Soltani, A. Takdastan, R. Saeedi, M. Abtahi,
x-ray spectroscopy confirmed that the atom percent of nitrogen in-
Enhanced coagulation-photocatalytic treatment of acid red 73 dye and real textile
creased from 6.56% to 17.97% following fouling of AEM by RBBR wastewater using UVA/synthesized MgO nanoparticles, J. Environ. Manag. 177
suggesting adsorption of RBBR to AEM. Likewise, sulfur was detected (2016) 111–118.
on RBBR fouled AEM which was not detected on the pristine sample. [11] M.C. Tomei, D.M. Angelucci, A.J. Daugulis, Sequential anaerobic-aerobic deco-
lourization of a real textile wastewater in a two-phase partitioning bioreactor, Sci.
Moreover, Nyquist plots showed that the ions transfer was affected by Total Environ. 573 (2016) 585–593.
the RBBR fouling layer with respect to the membrane resistance and the [12] A. Yurtsever, B. Calimlioglu, M. Görür, Ö. Çınar, E. Sahinkaya, Effect of NaCl
AEM surface structure. concentration on the performance of sequential anaerobic and aerobic membrane
bioreactors treating textile wastewater, Chem. Eng. J. 287 (2016) 456–465.
Last but not least, BMED performance showed that a good desali- [13] J. Lin, W. Ye, M.-C. Baltaru, Y.P. Tang, N.J. Bernstein, P. Gao, S. Balta, M. Vlad,
nation rate was obtained and membrane fouling was mitigated when A. Volodin, A. Sotto, Tight ultrafiltration membranes for enhanced separation of
the diluate salt concentration was controlled above the “critical salt dyes and Na 2 SO 4 during textile wastewater treatment, J. Membr. Sci. 514 (2016)
217–228.
concentration”. In particular, desalination rate of 74.43% was obtained [14] C. Xue, Q. Chen, Y.-Y. Liu, Y.-L. Yang, D. Xu, L. Xue, W.-M. Zhang, Acid blue 9
from 5% Na2SO4 under 0.25 g/L RBBR and 3A applied current. desalting using electrodialysis, J. Membr. Sci. 493 (2015) 28–36.
Meanwhile, 72.02% of sodium and 66.9% of sulfate in the feed were [15] J. Yao, D. Wen, J. Shen, J. Wang, Zero discharge process for dyeing wastewater
treatment, J. Water Proc. Eng. 11 (2016) 98–103.
converted to NaOH and H2SO4, respectively. [16] H.-J. Lee, M.-K. Hong, S.-D. Han, S.-H. Cho, S.-H. Moon, Fouling of an anion ex-
This study proves that BMED process may be an alternative way of change membrane in the electrodialysis desalination process in the presence of
treating textile wastewater with high salinity and the presence of dye organic foulants, Desalination 238 (2009) 60–69.
[17] H.-J. Lee, S.-H. Moon, Enhancement of electrodialysis performances using pulsing
molecules. Membrane fouling by medium size charged dye molecules
electric fields during extended period operation, J. Colloid Interface Sci. 287 (2005)
can be mitigated by carefully choosing operational parameters as pH 597–603.
and desalination rate. [18] S. Mikhaylin, L. Bazinet, Fouling on ion-exchange membranes: classification,
characterization and strategies of prevention and control, Adv. Colloid Interf. Sci.
229 (2016) 34–56.
Acknowledgment [19] J.-S. Park, H.-J. Lee, S.-J. Choi, K.E. Geckeler, J. Cho, S.-H. Moon, Fouling mitiga-
tion of anion exchange membrane by zeta potential control, J. Colloid Interface Sci.
The authors would like to appreciate the financial support of 259 (2003) 293–300.
[20] J.S. Park, T.C. Chilcott, H.G.L. Coster, S.H. Moon, Characterization of BSA-fouling
National Natural Science Foundation of China (21878319, 41673112), of ion-exchange membrane systems using a subtraction technique for lumped data,
the Shandong Provincial Key Science and Technology Program J. Membr. Sci. 246 (2005) 137–144.
(2018CXGC1005, 2017GSF17114), and Belt and Road Initiative Project [21] N. Tanaka, M. Nagase, M. Higa, Organic fouling behavior of commercially available
hydrocarbon-based anion-exchange membranes by various organic-fouling sub-
of the Chinese Academy of Sciences (211134KYSB20170010) and stances, Desalination 296 (2012) 81–86.
MEGA group for the free gift of heterogeneous bipolar membranes. [23] X. Sun, H. Lu, J. Wang, Recovery of citric acid from fermented liquid by bipolar
membrane electrodialysis, J. Clean. Prod. 143 (2017) 250–256.
[24] M. Reboiras, Electrochemical properties of cellulosic ion-exchange membranes II.
References
Transport numbers of ions and electro-osmotic flow, J. Membr. Sci. 109 (1996)
55–63.
[1] J. Lin, W. Ye, J. Huang, B. Ricard, M.-C. Baltaru, B. Greydanus, S. Balta, J. Shen, [25] P. Długołęcki, B. Anet, S.J. Metz, K. Nijmeijer, M. Wessling, Transport limitations in
M. Vlad, A. Sotto, Toward resource recovery from textile wastewater: dye extrac- ion exchange membranes at low salt concentrations, J. Membr. Sci. 346 (2010)
tion, water and base/acid regeneration using a hybrid NF-BMED process, ACS 163–171.
Sustain. Chem. Eng. 3 (2015) 1993–2001. [26] Y. Tanaka, Concentration polarization in ion-exchange membrane electro-
[2] Z. Carmen, S. Daniela, Textile organic dyes–characteristics, polluting effects and dialysis—the events arising in a flowing solution in a desalting cell, J. Membr. Sci.
separation/elimination procedures from industrial effluents–a critical overview, 216 (2003) 149–164.
Organic Pollutants Ten Years after the Stockholm Convention-Environmental and [28] K.S.V. Krishna Rao, B. Vijaya Kumar Naidu, M.C.S. Subha, M. Sairam,
Analytical Update, InTech, 2012. T.M. Aminabhavi, Novel chitosan-based pH-sensitive interpenetrating network
[3] A. Ammar, I. Dofan, V. Jegatheesan, S. Muthukumaran, L. Shu, Comparison be- microgels for the controlled release of cefadroxil, Carbohydr. Polym. 66 (2006)
tween nanofiltration and forward osmosis in the treatment of dye solutions, Desalin. 333–344.
Water Treat. 54 (2015) 853–861. [29] J. Fang, P.K. Shen, Quaternized poly(phthalazinon ether sulfone ketone) membrane
[4] I. Koyuncu, Reactive dye removal in dye/salt mixtures by nanofiltration membranes for anion exchange membrane fuel cells, J. Membr. Sci. 285 (2006) 317–322.
containing vinylsulphone dyes: effects of feed concentration and cross flow velocity, [30] S. Suwal, A. Doyen, L. Bazinet, Characterization of protein, peptide and amino acid
Desalination 143 (2002) 243–253. fouling on ion-exchange and filtration membranes: review of current and recently
[5] A. Gürses, M. Yalçin, C. Doğar, Electrocoagulation of some reactive dyes: a developed methods, J. Membr. Sci. 496 (2015) 267–283.

103

You might also like