You are on page 1of 36

Accepted Manuscript

Deep purification of seawater using a novel zeolite 3A incorporated polyether-


block-amide composite membrane

Filiz Ugur Nigiz, Sevil Veli, Nilufer Durmaz Hilmioglu

PII: S1383-5866(17)31511-3
DOI: http://dx.doi.org/10.1016/j.seppur.2017.07.017
Reference: SEPPUR 13874

To appear in: Separation and Purification Technology

Received Date: 11 May 2017


Revised Date: 7 July 2017
Accepted Date: 7 July 2017

Please cite this article as: F. Ugur Nigiz, S. Veli, N. Durmaz Hilmioglu, Deep purification of seawater using a novel
zeolite 3A incorporated polyether-block-amide composite membrane, Separation and Purification Technology
(2017), doi: http://dx.doi.org/10.1016/j.seppur.2017.07.017

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.
Deep purification of seawater using a novel zeolite 3A incorporated

polyether-block-amide composite membrane

Filiz Ugur Nigiz*1, Sevil Veli2, Nilufer Durmaz Hilmioglu1


1
Kocaeli University, Chemical Engineering Department, Kocaeli, 41380, TURKEY
2
Kocaeli University, Environmental Engineering Department, Kocaeli, 41380, TURKEY

*corresponding author

Email : filiz.ugur@kocaeli.edu.tr; filiz.ugurr@gmail.com

Phone : +90 2623033538

Mobile : +90 5325667117

1
Abstract

In this study, a novel hydrophilic-hydrophobic composite membrane was prepared and

used in a pervaporative desalination of NaCl-water solution and seawater. For this purpose,

polyether block amide (PEBA) was selected as membrane matrix. Zeolite 3A was

incorporated to PEBA polymer for facilitating the water permeation through the

membrane. The surface morphologies of the pristine and composite membranes were

examined by scanning electron microscopy. Thermogravimetric analyses of the pristine

and composite membranes were performed by adjusting the zeolite 3A concentration in

the PEBA polymer. The effect of the zeolite addition on membrane’s surface

hydrophilicity was analyzed using contact angle measurement. Firstly, influences of

zeolite content, feed temperature and NaCl concentration on pervaporative desalination

performances were performed in the pervaporation of NaCl-water solution. All

membranes exhibited excellent performance, and the salt rejection of >99.5 % and flux of

>2.07 kg.m-2h-1 were achieved. Secondly, seawater desalination was performed. Effect of

zeolite addition at a given temperature was also investigated. The better salt rejection was

obtained as 99.81 % accompanied with a very good flux of 4.57 kg.m-2h-1 in

pervaporative seawater desalination at 40 °C using 20 wt. % zeolite 3A incorporated

membrane.

Keywords: water purification, hydrophilic-hydrophobic membrane, polyether block

amide, desalination.

2
1. INTRODUCTION

The water scarcity in the world is growing at a fearful rate because of the increasing

human population. In addition that, the global warming and climate change effects

threaten the existing fresh water source. Today, 97 % of the world’s water (seawater,

brackish groundwater, ocean, i.e) is classified as saline water in different concentration.

Freshwater supplying from saline water by removing ions, minerals, heavy metals,

bacteria, dissolved salts, and other impurities is called as “Desalination”. The water is

called as a freshwater if the total dissolved solid (TDS) in the water is less than 1000 mg/L.

The limited TDS value for freshwater can be changed depending on the purpose of the

water consumption [1]. Desalinated water can be used for industrial purposes, irrigation,

and most importantly potable drinking water. In the current state, desalination is

estimated as an expensive technology for freshwater supplying [2,3]. However, it is

predicted that the water crisis will force to governments to improve their own desalination

technology in a close future. Especially for the countries that are located near the coast,

desalination will become a potential water supplying process.

Desalination technology has already been using in the Middle East, Africa, Europe, China,

Singapore and the USA with different freshwater capacities. There are several desalination

methods and types which preferred depending on the concentration of the saline water to

be separated and the country’s energy politics. Basically, desalination methods are

classified into two main categories as thermal separation and membrane separation.

Thermal separation techniques such as multi-effect and multistage flash distillation are

3
based on the phase separation phenomena whereby the seawater is heated and then

condensed to produce freshwater using thermal energy. Thermal desalination plants

have mostly been built in the country located in the Middle East due to the extensive

petroleum reserves. Membrane separation process is frequently constructed in the country

that has a limited energy source. Reverse osmosis (RO), forward osmosis (FO),

electrodialysis, and nanofiltration are commercialized membrane processes [4-6]. RO is a

pressure driven process in which the membrane acts as a barrier. Because of the separation

character of RO method, the size of the pores in the membrane and the differences in

component diffusivities play a key role on water purification. Although the purity of the

permeate water changes based on the membrane’s material, most of RO membranes are

able to retain dissolved ions. Hence, it is handled to obtain greater than 99 % of ion

rejections using RO technology [7-9].

Membrane distillation (MD) is a relatively new method in which mostly hydrophobic

membranes are used [10-13]. Separation phenomenon in MD depends on vapor pressure

difference between the upstream and downstream sides of the membrane. The membrane

acts as a separator in MD and the water-membrane interaction is very low due to its

hydrophobic nature.

Pervaporation (PV) offers new opportunities for alternative water desalination technology.

In this process, almost all impurities in water -from nano-scale to micro scale- are retained

on the feed side and the freshwater preferentially permeates through a dense membrane

[14-16]. By enabling the rejection of almost all impurities in seawater, greater than 99 %

4
of salt rejection results have been achieved using PV. Compared to other desalination

processes, pervaporation has some advantages. In PV, the separation factor is relatively

higher than that of RO under the same operating conditions [17]. Many of minerals, metal

elements, and ions have smaller diameters than that of the membrane with micro-pore

sized; consequently, the trace amount of undesired hazardous compounds can be non-

selectively permeated in the other membrane-based desalination technology. In PV very

sensitive separation occurs thanks to the dense and non-porous membrane usage.

While very high pressure is needed for desalination in RO, pervaporation is carried out at

the atmospheric pressure on the feed side and the vacuum on the downstream side.

Differently from the thermal desalination techniques where a great heat energy is needed

for entire water evaporation, the vacuum is used in pervaporation to evaporate the

permeate stream at low temperatures [17].

In fact, pervaporation is a long-known process for dehydration of alcohols and organic

solvents [18-20], removal of volatile organic compounds from industrial discharge [21,

22], and separation of organic-organic mixture selectively [23, 24]

The driving force in PV is the chemical potential gradient generated from a pressure

difference across the membrane. The performance of PV is directly related to the

difference in the partial pressures of the components. Separation behavior in PV is

basically explained using solution-diffusion law. When the feed solution contact with the

membrane, the target compound (water for desalination) preferentially dissolves on the

membrane’s surface, permeates through the membrane, and evaporates to the downstream

5
side. It is needed to investigate the feasibility of PV to produce large quantities of water.

In the literature, there are several successful pervaporative desalination studies by using

different types of membranes. Pristine polymeric membranes, inorganic particle loaded

polymeric membranes, hybrid membranes including more than one polymer, and inorganic

membranes have been used. The most parts of these studies depend on a hydrophilic

pervaporation membrane usage in order to increase the water flux. In a study performed

by Xie et al. [25], polyvinyl alcohol/maleic anhydride/silica (PVA/MA/Si) hybrid

membrane was prepared and used for the pervaporation of a NaCl-water solution.

Influence of the temperature, salinity of the water, and permeate pressure on desalination

performance were analyzed. They reported a flux of 11.7 kg/m2.h and rejection factor up

to 99.9 %. In another study reported by Xie et al. (2011), the same membrane was used

and the effect of membrane thickness was investigated. The water flux and rejection factor

were found as 6.93 kg/m2.h and 99.5 % respectively [26]. Chaudhri et al. (2015)

prepared a PVA-coated polysulfone membrane to be used for pervaporation with a

different range of NaCl incorporated model solution. They pointed out that the final

conductivity of the permeate water was 20 µs/cm and the flux was 7.4 L/m2.h at 344 K

[27]. Inorganic membrane desalination via pervaporation was carried out by Cho et al.

(2011). They prepared very thin NaA zeolite membrane and used for real seawater

desalination. The highest rejection was obtained as 99.9 % with a reasonable flux of 1.9

kg/m2.h [28]. Drobek and co-workers also prepared two different inorganic membranes of

silicalite-1 and ZSM-5 [29]. They investigated the performance of pervaporation at

different temperature with varying concentration of the NaCl-water solution. The highest

6
salt rejections were reported as 99 % and 96 % for ZSM-5 and silicalite-1 membranes

respectively [29]. Khajavi and co-workers synthesized a hydroxy sodalite membrane and

they achieved 99.99 % of salt rejection [30]. Additionally, almost the all impurities

retained in the feed side, and the final concentration of ions in the permeate side reduced

below a level of 0.02 ppm [30]. Feng et al. (2017) reported a grephene oxide (GO)

incorporated polyimide membrane and found a high rejection of 99.9 % with 36.1 kgm-2h-1

[31]. Xu et al. (2016) prepared a GO loaded polydopamine membrane to desalinate NaCıl-

water solution. The better flux of 48 kgm-2h- was found with a rejection greater than 99.7 %

[32]. In another study performed by Feng et al. (2016), graphene oxide framework

modified with 1, 4-phenylene diisocyanate (PDI) was synthesized and 99.9% rejection was
−2 −1
achieved with a high water flux (11.4 kg.m .h ) [33]. Hamouda et al. [34] reported a

pervaporative desalination study where the less-hydrophilic polyether-block-amide

membranes were used. They only investigated the effect of temperature and NaCl

concentration on flux and found very low flux of 3.67 g/m2.h at 28.7 °C due to the

hydrophobic nature of PEBA.

In this study, it is aimed to produce a novel zeolite 3A loaded PEBA membrane for

pervaporative seawater desalination. The main novelty of this study is converting a

hydrophobic matrix (PEBA) to a hydrophilic membrane by incorporating a water selective

zeolite (zeolite 3A). By this means, the water selectivity and passage through the

membrane is aimed to increase. Even though many researchers were fabricated

hydrophilic mixed matrix membrane, the most of them used hydrophilic filler into

hydrophilic polymer matrix. Because of the high water uptake capacity of hydrophilic

7
polymers, they can be exhibit unstable desalination performance with relatively short

lifetime. Therefore, using a hydrophobic matrix can be a key solution to reinforce the

membrane durability. Incorporating a hydrophilic zeolite into hydrophobic matrix

contributes to enhancing water flux and salt rejection. For these purposes, zeolite 3A was

incorporated into PEBA matrix. This is the first study of using hydrophilic zeolite 3A

loaded PEBA membrane based on the author’s knowledge.

Zeolite 3A is a crystalline potassium aluminosilicate, which is derivated from the sodium

form zeolite 4A. AlO2 group in the zeolite exhibits a negative charge and it is compensated

with the potassium (K+) ion in its frameworks [35]. Zeolite 3A is an appropriate zeolite

for water treatment process because of the molecular sieving effect and hydrogen bonding

capability of the water within the zeolite cages. The molecules which have kinetic

diameters greater than 0.3 nm are retained and the water is allowed to pass through the

cages of zeolite 3A. To take these advantages of the zeolite, different amounts of zeolite

3A was incorporated to PEBA matrix. Regarding the 3A incorporation, it was expected to

increase water flux coupled with higher salt rejection.

2 MATERIALS and METHODS

2.1 Materials

Polyether-block-amide (Pebax 2533) which is a block copolymer contain 80 wt%

poly(ethylene oxide) and 20 wt% polyamide was kindly obtained from the distributor of

the Arkema, Turkey. Zeolite 3A powder, acetic acid, and sodium chloride were purchased

8
from Aldrich, Turkey. The seawater was taken from the North coast of Marmara, Turkey

by the permission of the Forestry and Water Management Ministry of Turkish Republic.

2.2 Membrane preparation

The Pristine and zeolite 3A loaded mixed matrix membranes were prepared using the

solution-casting method. A PEBA-acetic acid solution with a polymer concentration of

10 wt. % was prepared and stirred for five hours at 60 ºC. This solution was defined as

pre-membrane base solution. Following the preparation of a homogeneous polymer

solution, zeolite 3A within a range of 0-50 wt. % (with respect to the pure PEBA

polymer’s weight) were added to the polymer solution and stirred for three hours at a

stirring rate of 500 rpm. In order to ensure preparation a defect-free membrane, priming

method was applied which was reported in the previous study [36,37]. According to the

priming procedure, zeolite particles were firstly dispersed in 5 ml of acetic acid and mixed

in ultrasonic mixer about 15 min. Then, 5 g of the pre-membrane base solution was added

to the zeolit-acid mixture. This procedure was applied to cover zeolite 3A particles with a

thin polymer film for obtaining a good zeolite-polymer interface. After one hour, remain

pre-membrane solution was added to zeolite-acid mixture. Afterward, the homogenization

step, the solution was poured on a Teflon plate and covered with a funnel to allow

evaporation of the solvent under the atmospheric conditions. After the membranes were

dried as a flat-sheet module, they were placed in an oven at 120 °C for three hours to

complete the thermal cross-linking reaction. The average thicknesses of the membranes

were approximately 100 µm.

9
2.3 Membrane characterization

The morphological properties and the homogeneous phase continuity of the pristine and

composite membranes were analyzed by scanning electron microscope (SEM) (JEOL

JSM-6335 F).

Fourier-Transform Infrared (FTIR) spectra (Perkin Elmer Pyris 1) analyses of the pristine

and 20 wt. % zeolite 3A incorporated membranes were performed to observe the zeolite

within membrane in the region of 650-4000 cm− 1

The effect of zeolite 3A loading on the surface hydrophilicity of the membrane was

investigated in a contact angle device (Attension KVS) using the Sessile Drop Method. To

ensure obtaining an accurate angle, measurements were repeated for four times and the

results were recorded with an accuracy of ±1-2°.

The effects of temperature on weight losses of the pristine and mixed matrix membranes

were analyzed by thermogravimetric analysis (TGA) (Mettler Toledo) in the temperature

from 25 °C to 600 °C at a heating rate of 10 °C/min under nitrogen atmosphere.

2.4 Seawater sample characterization

In the present study, the seawater was taken from the three different points of the north-

eastern Marmara seawater in Turkey. The seawater was obtained from 20 meter depth

below the ground surface and 200 meter far from the coastal. The Marmara seawater

characterization which was analyzed using Inductively Coupled Plasma Mass

Spectrometer (Perkin Elmer Elan DRC-e ICP-MS) is given in Table 1.

Table 1. Seawater chemical compositions

10
Analysis Marmara seawater
Conductivity(µs/cm) 43700
Sodium (mg/L) 13015
Magnesium (mg/L) 1411
Cloride (mg/L) 25580
Potassium (mg/L) <400
Lithium (µg/L) 115
Aluminum (µg/L) <1
Manganese (µg/L) <100
Iron (µg/L) 107
Nickel (µg/L) -
Bromide (µg/L) 58773
Arsenic (µ g/L) <5
Cadmium (µg/L) <0.2
Mercury (µ g/L) <0.2

2.5 Permeate analysis

The hydrated ion concentration of the permeated seawater was determined by Inductively

Coupled Plasma Mass Spectrometer (Perkin Elmer Elan DRC-e ICP-MS). The chloride

(Cl-) amount in the permeate water was determined using Hach Dr 5000 UV-Vis

Spectrophotometer with appropriate chloride standard. The conductivities of the feed and

permeate streams were measured using a Mettler Toledo Seven Compact S230 model

conductivity meter with an accuracy of ± 2 µs/cm.

2.6 Pervaporative desalination of the NaCl-water solution and seawater

The desalination of the NaCl-water solution and seawater feed were carried out in a

laboratory-scale pervaporation unit as shown in Fig. 1.

11
Figure 1. A representative pervaporative desalination set-up

The system consists of a home-made stainless steel membrane cell. The cell was equipped

with a mechanical stirrer to avoid a possible concentration polarization or mass-transfer

limitations. The cell was put in an oven to supply the desired temperature. The effective

separation area of the cell was 19.625 cm2. The pressure was 1 atm on the feed (upper)

side of the membrane and 1mbar pressure was applied by the vacuum pump on the

downstream side.

In the present study, pervaporative desalination performances of the pristine and zeolite

3A loaded (0, 10, 20, 30, 40, 50 wt. % 3A –respect to the polymer weight) membranes

were investigated using different concentration of NaCl-water solution and seawater under

different conditions.

Firstly, membrane performance for desalination of NaCl-water solution was performed. At

12
a given temperature of 40 °C, the effect of zeolite 3A incorporation on the performance of

desalination was determined with a constant NaCl concentration (3 wt. %). Following the

optimum zeolite determination, influences of the temperature (40, 50, 60 °C) and NaCl

concentration (1, 2, 3, 5, 10 wt. %) on the performance were investigated.

Secondly, membrane performance for Marmara Seawater desalination was investigated.

Effect of zeolite 3A concentration (in the range of 0 wt. % - 50 wt. %) on desalination

performance of seawater at 40 °C was performed.

The performance of the desalination was evaluated in terms of the water flux (J) and salt

rejection factor (R). With an hourly interval, permeated fresh water was weighted and

analyzed using conductivity method.

J= W p /A.t
(1)

R= ((C f - C p ) / C f ).100
(2)

where Wp represents the total permeate weight of the permeated sample, A is the effective

area of the flat sheet membrane and t is the operation time. Cf and Cp are conductances of

the feed and permeate stream respectively.

3 RESULTS and DISCUSSION

3.1 Membrane characterization

The morphological analysis of the pristine and mixed matrix membrane were observed

using SEM analysis. The SEM results of the membranes appeared in Figure 2.

13
Figure 2. Surface micrographs of pristine (a), 10 wt. % zeolite 3A loaded (b) and 50 wt. %

zeolite 3A loaded membrane (c)-(d)-(e)-(f)

Figure 2a represents the smooth and non-defected dense structure of the pristine PEBA

14
membrane. It is clearly indicated in Figure 2b that the zeolite was homogeneously

dispersed within the polymer. However, when the zeolite amount was increased, local

zeolite aggregations were observed as seen in Figure 2c. Although the agglomeration

tendency of excess amount zeolite, any contact-free region (between the polymer and

zeolite) was not seen (Figure 2d, 2e, 2f). Therefore, it is predicted that any non-selective

ion passage may not be occurred resulting the excess amount of zeolite loading. However,

the total permeation is expected to change by increasing amount.

The zeolite presence in PEBA matrix was analyzed using FTIR analysis. The FTIR spectra

of the pristine and 20 wt. % zeolite 3A incorporated membranes are given in Figure 3.

Figure 3. FTIR spectra of the pristine and zeolite incorporated PEBA membrane

The peak at 3312 cm- is corresponding to the characteristic N-H groups in pristine PEBA.

The peaks in the region of 2923-2852 cm- are attributed to the asymmetric and symmetric

stretching of the C-H bond. The intense of the C-H bonds reduced after zeolite addition.

15
Characteristic peaks of zeolite reveal between the band of 1019 cm- and 957 cm-. The peak

at 1019 cm- is corresponding to the Si-O-Si stretching of the zeolite within PEBA. The

peak at 3407 cm-1 indicates the Si-OH and Si-OH-Al bonding in the zeolite incorporated

membrane and confirms the presence of zeolite.

Figure 4. TGA curves and contact angle of the pristine and zeolite 3A/PEBA membrane

Figure 4 shows the thermal analysis and contact angle measurements of the pristine and

zeolite 3A loaded membranes. Although there is not a significant difference between the

first decomposition temperatures of the pristine and 20 wt. % membrane, the residual

weight of the composite membrane is higher. Meanwhile, the thermal durability of the

16
zeolite loaded membrane is expected to be relatively higher compared to pristine one

during the pervaporation experiments. However, an early decomposition is observed at

low temperature with the excess zeolite loaded membrane (50 wt. %). Figure 4 also shows

the increasing contact angle values of the membrane surfaces by zeolite addition.

Meanwhile, the hydrophilicity of the membrane enhanced with zeolite addition. This

observation was very important to fulfill the main purpose of the study. The hydrophilicity

of the hydrophobic membrane was improved by zeolite and it could be lead to enhance

both water flux and salt retention.

3.2 Membrane performance of NaCl-water solution

Figure 5 indicates the influence of zeolite loading on pervaporative desalination

performance at a constant temperature (40 °C) with a constant NaCl concentration (3

wt. %). It is clear from the figure that all membranes showed excellent NaCl rejection

results (> 99.5) with reasonable flux results (> 2 kgm-2 h-1). With increasing zeolite amount,

salt rejection firstly increased up to 99.63 %, and then gradually decreased. The better

rejection was observed using 20 wt.% zeolite 3A loaded composite membrane. Same as

the rejection, the flux value enhanced from 2.07 kgm-2 h-1 to 3.1 kgm-2h-1 with an

improvement of 49.7 % when the zeolite addition increased from 0 wt. % to 20 wt. %.

Similar increment-decrement relationship depending on the amount of filler in the

membrane was also reported in the literature [25, 26].

17
Figure 5. Effect of zeolite concentration on NaCl rejection and flux

This was due to the hydrophilic feature of the zeolite 3A. The increasing hydrophilicity of

the membrane caused a higher water-membrane interaction, consequently, water

permeation rate increased in a given operation period. Flux increment also attributed to

appropriate cage structure of zeolite 3A, which only has enough size to allow water

permeation. Although flux values decreased after 20 wt. % of zeolite loading, the decline

was not remarkable to change the overall performance of desalination. This slight

decrement could be explained by agglomeration tendency of zeolite which was also

observed in SEM micrographs. The excess amount of zeolite 3A would restrict the free

volume of the polymer and the vacant spaces of the zeolite, therefore caused a reverse

effect on flux.

Figure 6 indicates the NaCl concentration effect on flux and rejection at constant

18
temperature (40 °C) using 20 wt. % zeolite 3A loaded membrane.

Figure 6. Effect of NaCl concentration on rejection and flux

As seen in Figure 6, NaCl rejection did not affect from the concentration variations. The

NaCl rejection increased from 99.6 to 99.64 %. The decrease in flux was also very small

at the low NaCl concentration (between 1 wt. % and 3wt. %). Flux decreased from 3.33

kgm-2 h-1 to 3.1 kgm-2 h-1 when the NaCl concentration increased from 1 wt. % to 3 wt. %.

However, when the NaCl concentration increased from 3 wt. % to 10 wt. %, flux values

remarkably decreased from 3.1 kgm-2 h-1 to 1.98 kgm-2h-1. This difference was attributed to

the change in driving force in which maintained by the partial vapor pressure of the water.

NaCl was a non-volatile molecule, therefore the total vapor pressure of the feed mixture

decreased by NaCl addition. At the low temperature, this relation could be negligible, but

could not be underestimated at the high temperature. The similar rejection-flux behavior

19
was also obtained by Xie et al. [25], Drobek et al. [29] and Swenson et al. [38]. It was also

reported in the literature that the thermodynamical activity of the water decreases with the

salt addition [39]. In addition that, the excessive amount of NaCl on the surface of the

membrane would cause a concentration polarization and flux affected negatively. Another

reason of the flux decrement could be related to the free volume of the polymer. The free

volumes might be filled with non-volatile salt residuals; therefore the flux permeation

restricted during the experiments.

Figure 7 shows the influence of the feed temperature on the performance of pervaporation

experiments with a constant NaCl concentration (3 wt. %). Despite to the effect of NaCl

on performance, temperature changed the performance superiorly.

Figure 7. Effect of temperature on rejection and flux

As it is indicated in Figure 7, when the temperature increased from 40 °C to 60 °C, flux

20
enhanced from 3.1 kgm-2 h-1 to 4.33 kgm-2h-1. Increasing flux with temperature is a

prevalent observation for pervaporative separation studies [39-42]. There are many of

known reasons to increase the permeation rate of the component through the membrane.

Firstly, temperature increases the partial vapor pressure of the solvent in feed solution

exponentially, and the pressure gradient across the membrane increases. Diffusivity of the

component through the membrane and solubility of the component within the membrane

are also increased with temperature, thus the flux enhances as expected. Based on the free

volume theory of polymers, feed temperature directly affects the segmental chain

movement and permeation rates of solvents within the membrane [43]. Arrhenius equation

is helpful for better understanding the relationship between the flux and temperature [44].

It is also observed in Figure 7 that the increasing temperature adversely affected the NaCl

rejection. Salt rejection results decreased from 99.65 % to 99.16 % when the temperature

increased from 40 °C to 60 °C. This result directly related to the factors which positively

affected the water permeation. In particular, increasing free volume of the polymer would

allow for the non-selective ion passage and the rejection decreased. The similar

observation was also obtained by Chaudrhi et al. [27].

The trade-off phenomenon between the flux and rejection is a frequent observation in the

literature studies related to membrane separation process, especially in pervaporation

where the membrane-solvent interaction is remarkable.

3.3 Membrane performance of seawater

21
Salt rejection through a dense membrane depends on the size of the hydrated ions. In the

present study, an appropriate zeolite was used which allowed the only water permeation

[28]. Therefore, the ions having larger molecule diameter -than the cage size of the zeolite

or restricted free volume of the polymer- were retained in the feed side, and excellent salt

rejections were achieved (99.67 >) with all membranes (Figure 8). As shown in Figure 8,

when the zeolite ratio in polymer matrix was increased from 0 to 20 wt. %, salt rejection

enhanced from 99.67 % to 99.81 % when the temperature was 40 °C. After that point, a

slight decrement was seen due to the restricted structural mobility of the membrane.

Figure 8. Seawater desalination results of the pristine and composite membranes

On a basis of increasing hydrophilic character of the composite membrane, flux results

were improved and the highest flux of 4.57 kg.m-2h-1 was obtained. Compared to NaCl-

22
water solution, higher flux and rejection were achieved with Marmara seawater. Although

the seawater included hundreds of impurities, zeolite 3A loaded PEBA membranes

showed promising performance for saline water desalination under the mild operating

conditions. According to the charge exclusion theory, the surface of the membrane

exhibit repulsive effect on the same charge of ions [28]. In the present study, potassium

form of zeolite was used; therefore, the salt rejection increased with the increasing amount

of zeolite 3A. Additionally, compared to NaCl-water solution, the concentrations of the

ions are higher in the seawater. Hence, it was the evidence of why seawater exhibit higher

rejection performance than the NaCl-water solution.

Table 2 shows the World Health Organization's (WHO) drinking guidelines [45],

Europe/Turkish drinking water guidelines [46], Turkish irrigation water guidelines [47],

ICP-MS results of the pristine and 20 wt. % zeolite 3A loaded membrane at 40 °C. The

hydrated ions were rejected by membrane due to the repulsive charge effect of the zeolite

within the membrane. Similar results were also obtained by Cho and co-worker [28].

Because of the ionic radius of the hydrated ions are greater than cage size of the zeolite 3A,

ions would not be allowed to pass through the membrane.

Table 2. Comparison of the experimental results of permeate water with related standards

Analysis Marmara WHO Europe/Turkish Turkish Pristine Wt. 20 %


seawater drinking drinking water irrigation PEBA Zeolite
water standard water loaded
standard standard PEBA
Conductivity 43700 - 2500 - 116 73
(µ s/cm)
Sodium 13015 200 200 - 75.8 3.5
(mg/L)
Magnesium 1411 - - - 2.1 0.5

23
(mg/L)
Cloride 25580 250 250 - 2 0
(mg/L)
Potassium <400 - - - 3.3 0.96
(mg/L)
Lithium 115 - - 2500 <1 <1
(µ g/L)
Aluminum <1 200 200 1000 78 34
(µ g/L)
Manganese <100 100 50 200 2.6 2.3
(µ g/L)
Iron 107 300 200 5000 27.7 <5
(µ g/L)
Nickel - - 20 200 182 87
(µ g/L)
Bromine 58773 - - - 24 18
(µ g/L)
Arsenic <5 - 10 500 0.3 <0.1
(µ g/L)
Cadmium <0.2 - 5 10 0.1 <0.1
(µ g/L)
Mercury <0.2 - 1 - 0.1 <0.1
(µ g/L)

Another observation seen from the Table 2 that the chloride, cadmium, bromide

concentrations in desalinated water were very low where cannot be achieved in another

processes such as single pass RO or FO. This level is far below to be allowed for irrigation

or potable water standard.

In order to determine the membrane stability for seawater desalination, repeatability

experiments were carried out using 20 wt. % 3A incorporated membrane at 40 °C during

40 hours. For this purpose, 300 ml seawater (with an average salt concentration of 2.7 %)

was continuously desalinated with the same membrane. The flux and rejection stability of

the membrane is illustrated in Figure 9.

24
Figure 9. Flux and rejection stability of the 20 wt. % zeolite 3A incorporated membrane

According to the rejection results, membrane preserved 99.9 % of its stability. In the first

period (until 20 hours) no significant variations were observed in flux. However, totally

21.7 % flux decrement was seen at the end of 40 hours due to the increasing salt

concentration (from 2.7 % to 5.6 %) on the feed side with the time.

Conclusions

In this study, it was aimed to produce a novel hydrophilic-hydrophobic membrane by

dispersing the zeolite 3A particles in continues phase of PEBA polymer. Thus, while the

water permeation was increased by zeolite 3A addition, excess swelling degree –which

caused a reduction in rejection- was restricted by using PEBA polymer. To the author’s

knowledge, this is the first study of using hydrophilic zeolite 3A loaded PEBA membrane

25
for pervaporative seawater desalination. Despite to the fact that the only hydrophilic dense

membrane is appropriate type for pervaporative water separation, the zeolite 3A loaded

PEBA membrane shows superior desalination performance. The main findings are listed

below;

• Effects of zeolite loading on the performances of the NaCl-water solution and

seawater desalination were investigated, and the optimum zeolite addition was

observed as 20 wt. % zeolite 3A in NaCl-water and seawater experiments.

• Flux value enhanced from 2.07 kgm-2 h-1 to 3.1 kgm-2h-1 and 49.7 % flux

improvement was achieved when the zeolite content in membrane increased from

0 wt. % to 20 wt. %.

• All membranes exhibited excellent performance with a salt rejection of >99.5 %

and flux of >2 kg.m-2h-1 .

• Both the flux and NaCl rejection did not significantly change by increasing

concentration of NaCl in water, hence this finding confirmed that this membrane

could be used for separation of the different source of water (groundwater,

seawater, ocean, brackish).

• When the temperature of the 3 wt. % of NaCl-water solution increased from 40 °C

to 60 °C, flux enhanced from 3.1 kgm-2 h-1 to 4.33 kgm-2 h-1 and the rejection

slightly decreased from 99.65 % to 99.16 %.

• The better salt rejection was obtained as 99.81 % accompanied with a very good

flux of 4.57 kg.m-2h-1 in pervaporative seawater desalination experiments using 20

26
wt. % zeolite 3A incorporated membrane at 40 °C. The flux of pristine membrane

for seawater desalination was found as 3.33 at the same conditions, and 37 % flux

improvement was achieved.

• The 20 wt. % 3A zeolite incorporated membranes showed a stable rejection results

and preserved 99.9 % of its rejection performance during 40 hours experiments

period for seawater desalination. Regarding the flux increment and salt rejection

enhancement, 20 wt. % zeolite incorporated PEBA membrane could be consider

more feasible than pristine PEBA for pervaporative desalination.

Additionally, ICP-MS analysis of purified water confirmed that the zeolite 3A loaded

PEBA membrane was a proper candidate for being a high-performance dense

membrane to obtain potable pure water from the saline water.

Acknowledgements

The authors would like to thank colleagues at Polymer Laboratory from Kocaeli

University for permitting to use of their characterization equipment. The authors also

gratefully acknowledge Hayim Pinhas Group (Distributor of Arkema in Turkey) for kindly

supplying of the Pebax 2533. The study was financially funded by the Scientific Research

Center of Kocaeli University (2017/009)

References

27
[1] L. F. Greenlee, D. F. Lawler, B. D. Freeman, B. Marrot, P. Moulin, P. Ce, “Reverse

osmosis desalination : Water sources , technology , and today’s challenges,” Water

Res., vol. 43, no. 9, pp. 2317–2348, 2009.

[2] K. P. Lee, T. C. Arnot, D. Mattia, “A review of reverse osmosis membrane materials

for desalination — Development to date and future potential,” J. Memb. Sci., vol.

370, no. 1–2, pp. 1–22, 2011.

[3] C. Fritzmann, J. Löwenberg, T. Wintgens, T. Melin, “State-of-the-art of reverse

osmosis desalination,” Desalination, vol. 216, pp. 1–76, 2007.

[4] Y. Cai, W. Shen, S. Leng Loo, W. B. Krantz, R. Wang, A. G. Fane, X. Hu, “Towards

temperature driven forward osmosis desalination using Semi-IPN hydrogels as

reversible draw agents,” Water Research, vol. 47, pp. 3773 -3781, 2013.

[5] R.V. Linares, Z. Li, S. Sarp, Sz.S. Bucs, G. Amy, J.S. Vrouwenvelder, “Forward

osmosis niches in seawater desalination and wastewater reuse,” Water Research, vol.

66, pp. 122-139, 2014.

[6] G. Kang, Y. Cao, “Development of antifouling reverse osmosis membranes for water

treatment : A review,” Water Res., vol. 46, no. 3, pp. 584–600, 2011.

[7] D. Li and H. Wang, “Recent developments in reverse osmosis desalination membranes,”

Journal of Materials Chemistry, vol. 20, pp. 4551–4566, 2010.

[8] B. Penate and L. García-rodríguez, “Current trends and future prospects in the design

of seawater reverse osmosis desalination technology,” Desalination, vol. 284, pp. 1–

8, 2012.

28
[9] T. Suzuki, R.Tanaka, M. Tahara, Y. Isamu, M. Niinae, L. Lin, J.Wang, J. Luh, O.

Coronell, “Relationship between performance deterioration of a polyamide reverse

osmosis membrane used in a seawater desalination plant and changes in its

physicochemical properties,” Water Research, vol. 100, pp. 326-336, 2016.

[10] Y.D. Kim, K. Thu, K. Choon, G.L. Amy, N. Ghaffour, “A novel integrated thermal-

/membrane-based solar energy-driven hybrid desalination system: Concept

description and simulation results,” Water Research, vol. 100, pp. 7-19, 2016.

[11] M. Bhadra, S.Roy, S. Mitra, “Enhanced desalination using carboxylated carbon

nanotube immobilized membranes,” Separation and Purification Technology,

vol.120, pp. 373–377, 2013.

[12] S. Roy, M. Bhadra, S. Mitra, “Enhanced desalination via functionalized carbon

nanotube immobilized membrane in direct contact membrane distillation,”

Separation and Purification Technology, vol. 136, pp. 58–65, 2014.

[13] S. Bano, A. Mahmood, S. J.Kim, K. Lee, “Chlorine resistant binary complexed

NaAlg/PVA composite membrane for nanofiltration,” Separation and Purification

Technology, vol. 137, pp. 21–27, (2014).

[14] M.C. Duke, S. Mee, J.C. Diniz da Costa, “Performance of porous inorganic

membranes in non-osmotic desalination,” Water Research, vol. 41, pp. 3998-4004,

2007.

[15]A. Subramani, J. G. Jacangelo, “Emerging desalination technologies for water

treatment: A critical review,” Water Research, vol. 75, pp. 164-187, 2015.

29
[16] Z. Xie, M. Hoang, D. Ng, C. Doherty, A. Hill, S.Gray, “Effect of heat treatment on

pervaporation separation of aqueous salt solution using hybrid PVA/MA/TEOS

membrane,” Separation and Purification Technology, vol. 127, pp. 10–17, 2014.

[17] A. Gao, “Desalination of High-salinity Water by Membranes” Master Thesis,

University of Waterloo, Canada, 2016.

[18] A. Basile, A. Figoli M. Khayet, “Pervaporation, Vapour Permeation and Membrane

Distillation, Principles and Applications” First ed. Woodhead Publishing, 2015.

[19] H. E. A. Brüschke, “State-of-Art of Pervaporation Processes in the Chemical Industry,

in Membrane Technology: in the Chemical Industry” (eds S. P. Nunes and K.-V.

Peinemann), Wiley-VCH Verlag GmbH, Weinheim, 2001.

[20] Y. K. Ong, G. M. Shi, N. L. Lea, Y. P. Tang, J. Zuo, S. P. Nunes, T. Chung, “Recent

membrane development for pervaporation processes,” Progress in Polymer Science,

vol. 57, pp. 1–31, 2016.

[21] L. Aouinti, D. Roizard, M. Belbachir, “PVC–activated carbon based matrices: A

promising combination for pervaporation membranes useful for aromatic–alkane

separations,” Separation and Purification Technology, vol. 147, pp. 51–61, 2015.

[22] H. Zhou, Y.Su,Y. Wan, “Phase separation of an acetone–butanol–ethanol (ABE)–

water mixture in the permeate during pervaporation of a dilute ABE solution,”

Separation and Purification Technology, vol. 132, pp.354–361, 2014.

[23] S. B. Kuila, S. K. Ray, “Separation of benzene – cyclohexane mixtures by filled blend

membranes of carboxymethyl cellulose and sodium alginate,” Separation and

30
Purification Technology, vol. 123, pp. 45–52, 2014.

[24] G. Genduso, H. Farrokhzad, Y. Latré, S. Darvishmanesh, P. Luis, B. Van der Bruggen,

“Polyvinylidene fluoride dense membrane for the pervaporation of methyl acetate–

methanol mixtures,” Journal of Membrane Science, vol. 482, pp. 128–136, 2015.

[25] Z. Xie, D. Ng, M. Hoang, T. Duong, S. Gray, “Separation of aqueous salt solution by

pervaporation through hybrid organic – inorganic membrane : Effect of operating

conditions,” Desalination, vol. 273, no. 1, pp. 220–225, 2011.

[26] Z. Xie, M. Hoang, T. Duong, D. Ng, B. Dao, S. Gray, “Sol – gel derived poly ( vinyl

alcohol )/ maleic acid / silica hybrid membrane for desalination by pervaporation,” J.

Memb. Sci., vol. 383, no. 1–2, pp. 96–103, 2011.

[27] S. G. Chaudhri, B. H. Rajai, P. S. Singh, “Preparation of ultra-thin poly (vinyl alcohol)

membranes supported on polysulfone hollow fiber and their application for

production of pure water from seawater,” Desalination, vol. 367, pp. 272–284, 2015.

[28] C. H. Cho, K. Y. Oh, S. K. Kim, J. G. Yeo, P. Sharma, “Pervaporative seawater

desalination using NaA zeolite membrane : Mechanisms of high water flux and high

salt rejection,” J. Memb. Sci., vol. 371, no. 1–2, pp. 226–238, 2011.

[29] M. Drobek, C. Yacou, J. Motuzas, A. Julbe, L. Ding, J.C.D. da Costa “Long term

pervaporation desalination of tubular MFI zeolite membranes,” Journal of

Membrane Science, vol. 416, pp. 816–823, 2012.

[30]S. Khajavi, J. C. Jansen, F. Kapteijn, “Production of ultra pure water by desalination

of seawater using a hydroxy sodalite membrane,” J. Memb. Sci., vol. 356, no. 1–2,

31
pp. 52–57, 2010.

[31] B. Feng, K. Xu, A. Huang, “Synthesis of graphene oxide/polyimide mixed matrix

membranes for desalination,” RSC Adv., vol. 7, pp. 2211–2217, 2016.

[32] K. Xu, B. Feng, C. Zhou, A. Huang, “Synthesis of highly stable graphene oxide

membranes on polydopamine functionalized supports for seawater desalination,”

Chem. Eng. Sci., vol. 146, pp. 159–165, 2016.

[33] B. Feng, K. Xu, A. Huang, “Covalent synthesis of three-dimensional graphene oxide

framework (GOF) membrane for seawater desalination,” Desalination, vol. 394, pp.

123–130, 2016.

[34] S. B. Hamouda, A. Boubakri, Q. T. Nguyen, aM. Ben Amor, “PEBAX membranes for

water desalination by pervaporation process,” High Performance Polymers, vol.

23(2) pp. 2–5, 2011.

[35] M. L. Restrepo, M. A. Mosquera, “Accurate correlation, thermochemistry, and


structural interpretation of equilibrium adsorption isotherms of water vapor in
zeolite 3A by means of a generalized statistical thermodynamic adsorption model,”
Fluid Phase Equilibria, vol.283, pp. 73-88, 2009.
[36] F. Ugur, H. Dogan, N. Durmaz, “Pervaporation of ethanol/water mixtures using

clinoptilolite and 4A fi lled sodium alginate membranes,” Desalination, vol. 300, pp.

24–31, 2012.

[37] F. U. Nigiz, N. D. Hilmioglu, “Pervaporation of ethanol/water mixtures by zeolite

filled sodium alginate membrane,” Desalination and Water Treatment,vol. 51, pp.

637-643, 2013.

32
[38] P. Swenson, B. Tanchuk, A. Gupta, W. An, S. M. Kuznicki, “Pervaporative

desalination of water using natural zeolite membranes,” Desalination, vol. 285, pp.

68–72, 2012.

[39] Q. Wang, N. Li, B. Bolto, M. Hoang, Z. Xie, “Desalination by pervaporation : A

review,” Desalination, vol. 387, pp. 46–60, 2016.

[40] D. Sun, B. Li, Z. Xu, “Pervaporation of ethanol / water mixture by organophilic nano-

silica filled PDMS composite membranes”, Desalination, vol. 322, pp. 159–166,

2013.

[41] Y. Fu, C. Lai, J. Chen, C. Liu, S. Huang, “Hydrophobic composite membranes for

separating of water – alcohol mixture by pervaporation at high temperature,” Chem.

Eng. Sci., vol. 111, pp. 203–210, 2014.

[42] Q. Wang, Y. Lu, N. Li, “Preparation , characterization and performance of sulfonated

poly (styrene-ethylene/butylene-styrene) block copolymer membranes for water

desalination by pervaporation,” Desalination, vol. 390, pp. 33–46, 2016.

[43]A. Mafi, A. Raisi, M. Hatam, A. Aroujalian, “A Comparative Study on the Free

Volume Theories for Diffusivity Through Polymeric Membrane in Pervaporation

Process,” Journal of Applied Polymer Science, vol. 40581, pp. 1–12, 2014.

[44] H. G. Premakshi, K. Ramesh, M. Y. Kariduraganavar, “Modification of crosslinked

chitosan membrane using NaY zeolite for pervaporation separation of water –

isopropanol mixtures,” Chem. Eng. Res. Des., vol. 94, pp. 32–43, 2014.

[45] URL-1. http://www.who.int/en/ (visiting date: 24.03.2017)

33
[46] URL-2: http://www.resmigazete.gov.tr/eskiler/2013/03/20130307-7.htm (visiting date:

24.03.2017)

[47] URL-3: http://www.cmo.org.tr/mevzuat/mevzuat_detay.php?kod=230 (visiting date:

24.03.2017)

34
HIGHLIGHT

• A novel hydrophilic-hydrophobic composite membrane was prepared.


• 3A zeolite loaded-polyether-block-amide membrane was used for desalination.
• All membranes exhibited excellent desalination performance
• The better salt rejection was obtained as 99.81 % accompanied with a very good
flux of 4.57 kg.m-2h-1

35

You might also like