Vous êtes sur la page 1sur 146

Fouling-resistant surface

modification of membranes used


in water treatment

by

Sara Azari
B. Eng (Textile Chemistry Engineering)
M. Eng (Textile Chemistry Engineering)

A thesis submitted for the degree of

Doctor of Philosophy

SA Centre for Water Management and Reuse


School of Natural and Built Environment
Division of Information Technology, Engineering and the Environment

2014
Abstract
In recent decades, membrane technology has emerged as a highly viable water purification
technique. However, membrane fouling is a challenge in widespread application. One
approach to control fouling is through surface modification of membranes. My research
explores novel materials for low fouling surface modifications of membranes used in water
treatment.

It firstly focuses on incorporating the amino acid 3-(3,4-Dihydroxyphenyl)-L-alanine


(L-DOPA) onto commercial polyamide reverse osmosis (RO) membranes (SW 30 XLE) to
create a zwitterionic surface that resists membrane fouling. The top layer of the membrane
surface was modified by the deposition of L-DOPA from its alkaline solution. Contact
angle measurements indicated that the hydrophilicity of the coated membranes was
significantly improved. While the salt rejection remained unchanged, a systematic increase
in water flux was observed for the samples coated up to 12 hours which was explained by
the increase in membrane hydrophilicity. In cross-flow filtration tests, using a mixed
bovine serum albumin (BSA) and alginic acid sodium salt solution as the feed, the
modified membranes exhibited lower fouling rate than the virgin untreated membrane. The
coated membranes also showed an improvement in their fouling resistance when tested
with a solution containing dodecyltrimethyl ammonium bromide (DTAB), a positively
charged surfactant model foulant. To further investigate the low fouling effect of L-DOPA
on different membrane types and materials, modification of a cellulose acetate
microfiltration membrane as an example of a porous membrane type was also studied.
Porosity measurement showed a slight decrease in membrane porosity due to the
modification. Static adsorption experiments revealed an improved resistance of the
modified membranes towards the adhesion of bovine serum albumin (BSA) as the model
foulant. Membrane filtration tests using a dead end cell, confirmed the enhanced fouling
resistance of the modified membrane.

To expand the understanding about fouling resistance effect of zwitterionic aminoacids, the
amino acid L-cysteine was grafted onto the RO membrane. The modification was carried
out through bonding the thiol group of the L-cysteine to the allyl functionalized membrane.
Contact angle measurement indicated the higher hydrophilicity of the modified membrane.
Although membrane modification enhanced salt rejection, membrane permeability
declined which can be due to the tighter structure of the modified membrane. Cross-flow
filtration of solutions of BSA and DTAB, as model foulants, displayed a lower fouling
propensity of the modified membrane.
ii
Both L-DOPA coating and L-cysteine grafting could successfully mitigate membrane
fouling. However, L-DOPA coating was identified as the more promising approach. Since
membrane modification with L-DOPA can be easily performed even in a membrane
module assembly, this approach is also of particular interest practically. To obtain a deeper
insight into the low fouling effect of L-DOPA on the membrane, free energies of adhesion
between the modified RO membrane and model foulants were determined. These free
energies were evaluated based on contact angle results using the Extended
Derjaguin-Landau-Verwey-Overbeek (XDLVO) theory. L-DOPA modification
significantly lowers the affinity toward adhesion of both the foulants studied. The
accelerated fouling tests confirmed well the results obtained from theory.

The outcomes of this study suggest that L-DOPA can be considered as a promising
material for industrial scale fouling-resistant membrane surface modification.

iii
Declaration
I hereby declare the following to be my own work, and to the best of my knowledge and
belief, this thesis contains no materials published or written by another person, except
where due acknowledgement and reference are made, as defined by the University's policy
on plagiarism. Any contribution made to the research by others, with whom I have worked
at UniSA or elsewhere, is explicitly acknowledged in the thesis.

Sara Azari February 2014

iv
Acknowledgments
First of all, I would like to express my sincere appreciation to my principal supervisor,
Professor Linda Zou for her support and guidance throughout this entire PhD project. Her
efforts to provide her research group with a well-equipped laboratory for a high quality
research are really appreciated.

Thanks also extended to Dr. Emile Cornelissen for his insightful advice and comments.

Great appreciation to my associate supervisor Professor Dennis Mulcahy, and also to Dr.
Monica Behrend and Dr. Judy Ford for their kind assistance on revising my thesis draft.

Thanks also go to all my friends in the lab; Yasodinee, Chaijuan, Juan, Anh, Hibo, Wei,
Fransoic and Pejman for sharing the thought, laugh and hardship. I also would like to thank
my entire friends in the CWMR centre with whom I shared my happiness and sadness
moments. Their friendships will always be remembered.

I would also like to take this opportunity to express my gratitude to all the staff members in
the SA Centre for Water Management and Reuse; particular thanks go to Professor Chris
Saint and Ms. Allison Price for their generous helps in many respects.

I gratefully thank to Mawson Lakes Fellowship program for providing me an opportunity


to conduct a visiting research in Nagoya University, Japan during my PhD studies. Special
thanks to Professor Yasushito Mukai for accepting me to work in his group.

Particular thanks go to people in the workshop, Tim and Wes for their assistance on
technical issues.

I am forever grateful to my parents and my sister for their love, support and
encouragement, through entire my life.

Last but not least, I would like to express my deepest gratitude to my husband Nasser, for
all his emotional and technical supports during my PhD study. I am forever thankful to him
for all his efforts to make my PhD journey peaceful and enjoyable.

v
Table of Contents
Abstract ii

Declaration iv

Acknowledgments v

Table of Contents vi

List of Figures x

List of Tables xiii

List of Publications xiv

Glossary xvi

1 Introduction 1

1.1 Motivation of the work 1


1.2 Membranes for water treatment 1
1.3 Membranes for desalination of seawater and brackish water 3
1.4 Problem statement 4
1.5 Strategies to control fouling 5
1.5.1 Pre-treatment of feed water 5
1.5.2 Membrane fouling control by mechanical means 5
1.5.3 Membrane surface modification 7
1.6 Objectives and the scope of the thesis 8
1.7 Thesis organization 9
1.8 Contribution to knowledge 10
1.9 Concluding remarks 11

2 Literature review 14

2.1 Introduction 14
2.2 Classification of fouling agents and their properties 14
2.2.1 Mineral scales 14
2.2.2 ‎Colloidal foulants 15
2.2.3 Biofoulants 17
2.3 Effect of membrane fouling on membrane performance 18
2.4 General factors affecting membrane organic fouling 20
vi
2.4.1 Effect of feed water characteristics on organic fouling 20
2.4.2 Effect of operational conditions on organic fouling 22
2.4.3 Effect of membrane properties 23
2.4.4 Effect of membrane–foulant and foulant–foulant interactions on fouling 26
2.5 Membrane surface modification approach to control foulant adhesion 27
2.5.1 Hydrophilic modification 28
2.5.2 Zwitterionic modification 32
2.6 Conclusion: challenges and future opportunities 34

3 Using zwitterionic amino acid L-DOPA to modify the surface of thin film
composite polyamide reverse osmosis membranes to increase their fouling
resistance 46

3.1 Introduction 47
3.2 Experimental 49
3.2.1 Materials 49
3.2.2 Surface modification of the membrane 49
3.2.3 Characterisation of the membrane surface 50
3.2.4 The permeation and salt rejection experiments 51
3.2.5 Cross-flow filtration experiments and fouling characterisation 52
3.3 Results and discussion 53
3.3.1 Surface characterization 54
3.3.2 Water flux and salt rejection of membranes 59
3.3.3 Filtration experiment 61
3.4 Conclusion 64
References 65

4 Facile fouling resistant surface modification of microfiltration cellulose


acetate membranes by using amino acid L-DOPA 69

4.1 Introduction 70
4.2 Experimental 71
4.2.1 Materials 71
4.2.2 Methods 71
4.3 Results and discussion 75
4.3.1 Characterization of the membranes 75
4.3.2 BSA adsorption and fouling behaviour of the membranes 78
4.4 Conclusions 81

vii
5 Fouling resistant zwitterionic surface modification of reverse osmosis
membranes using amino acid L-cysteine 85

5.1 Introduction 86
5.2 Materials and methods 88
5.2.1 Materials 88
5.2.2 Membrane surface modification 88
5.2.3 Characterization of modified membranes 89
5.2.4 Measurement of water permeability and salt rejection 91
5.2.5 Evaluation of antifouling efficiency of modified membranes 91
5.3 Results and discussion 92
5.3.1 Surface characterization 92
5.3.2 Membrane performance 99
5.4 Conclusion 102

6 Assessing the effect of surface modification of polyamide RO membrane


by L-DOPA on the short range physiochemical interactions between the
membrane and biopolymers: elucidation of the role of L-DOPA surface
modification on membranes fouling behaviour 107

Abstract 107
6.1 Introduction 108
6.2 Theory 109
6.2.1 Interaction energies 109
6.2.2 Surface tension components 109
6.2.3 Determination of free energy of interaction 110
6.3 Materials and method 111
6.3.1 Membranes and foulant 111
6.3.2 Membrane modification 112
6.3.3 Probe liquids 113
6.3.4 Membrane characterization 113
6.3.5 Foulants characterization 114
6.3.6 Membrane fouling experiments 114
6.4 Results and discussion 114
6.4.1 Surface zeta potential 114
6.4.2 Surface energy determination 116
6.4.3 Free energy of adhesion at contact 117
6.4.4 Fouling test 118

viii
6.5 Conclusion 121

7 Conclusions and Future Work Recommendations 125

7.1 Conclusions 125


7.2 Recommendations for future work 129

ix
List of Figures
Figure ‎1.1 Schematic illustration of cross-flow filtration versus dead end filtration [22] ................ 6
Figure ‎1.2 Increasing research interest on membrane surface modification reflected in annual
academic publications (based on the web of science) ........................................................................ 8
Figure ‎2.1 SEM of calcium sulfate (gypsum) precipitate from a supersaturated solution on RO
membrane at 50 °C (left) and after addition of antiscalant (right) (from [5]) .................................. 15
Figure ‎2.2 Biofouled membrane module [15] .................................................................................. 18
Figure ‎2.3 Schematic illustration of the effect of feed water chemistry on Humic acid
conformation and the consequence effect on membrane fouling (adapted from [28]) .................... 22
Figure ‎2.4 Schematic illustration of the effect of roughness on the deposition rate of fouants in a
cross flow filtration system (less foulant deposition on smooth surface) (adapted from [62]) ........ 25
Figure ‎2.5 Factors affecting organic fouling of RO and NF membranes (adapted from [7]) .......... 25
Figure ‎2.6 Two main structures of zwitterionic materials a) poly(sulfobetaine) b)
poly(carboxybetaine) (adapted from [103]) ..................................................................................... 33
Figure ‎3.1 Pictures of method and Apparatus used for top surface modification: (a) in the first
minutes of coating (b) after 12 hour of coating................................................................................ 50
Figure ‎3.2 Schematic of the cross-flow test unit .............................................................................. 52
Figure ‎3.3 Oxidative polymerisation of L-DOPA and schematic of surface adsorption resistance for
organic matter imparted by the hydrated zwitterionic poly L-DOPA coated surface ...................... 54
Figure ‎3.4 Measured zeta potential results ....................................................................................... 55
Figure ‎3.5 UV–visible spectra for the virgin, 4 and 24-hour modified membranes ........................ 56
Figure ‎3.6 Water contact angle of the membranes as a function of modification time (test
temperature: 22°C) ........................................................................................................................... 57
Figure ‎3.7 Atomic force microscopy images (including morphological statistics) ......................... 58
Figure ‎3.8 The measured amount of BSA adsorbed on the surface of virgin and modified
membranes from BSA solutions of 100 mg/L ................................................................................. 59
Figure ‎3.9 Water permeability and salt rejections of the virgin XLE membrane and the membranes
modified with different modification times ..................................................................................... 60
Figure ‎3.10 Resistance of the virgin and modified membranes as a function of time during
BSA/alginate (100 mg/L of each) fouling, initial flux 21± 0.4 L/m2.h ............................................ 61
Figure ‎3.11 Resistance of the virgin and modified membranes as a function of permeate volume
during BSA/alginate (100 mg/L of each) fouling, initial flux 21 ± 0.4 L/m2.h ............................... 62
Figure ‎3.12 Resistance of the virgin and modified membranes as a function of time during DTAB
(50 mg/L) fouling. initial flux 21.5± 0.5 L/m2.h .............................................................................. 63
Figure ‎3.13 Resistance of the virgin and modified membranes as a function of permeate volume
during DTAB (50 mg/L) fouling, initial flux 21.5± 0.5 L/m2.h ....................................................... 63
Figure ‎4.1 Schematic diagram of the filtration setup ....................................................................... 74
Figure ‎4.2 ATR-FTIR spectra of the original and three- step coated membranes ........................... 76
x
Figure ‎4.3 The percentage of flux loss evaluated by measuring the flux of each membrane, at 20
kPa, before and after static BSA adsorption .................................................................................... 78
Figure ‎4.4 The plot of membrane resistance versus time during the filtration of 0.2 g/L of BSA
solution ............................................................................................................................................. 79
Figure ‎4.5 The plot of membrane resistance versus permeate volume during the filtration of 0.2 g/L
of BSA solution................................................................................................................................ 79
Figure ‎4.6 The plot of membrane resistance versus time during the filtration of 0.2 g/L of 0.04 g/L
BSA solution .................................................................................................................................... 80
Figure ‎4.7The plot of membrane resistance versus permeate volume during the filtration of 0.2 g/L
of BSA solution................................................................................................................................ 80
Figure ‎5.1 Schematic of surface adsorption resistance for organic matter imparted by the hydrated
zwitterionic L-cysteine coated membrane surface ........................................................................... 87
Figure ‎5.2 Schematic of chemical reactions used for membrane modification: firstly activation with
allyl glycidyl ether then reaction with thiol group of zwittterionic L-cysteine ................................ 89
Figure ‎5.3 ATR-FTIR spectra of the virgin RO, AGE activated and L-cysteine modified
membranes a) wave numbers between 1000 −1800 cm–1 −3700
cm–1 .................................................................................................................................................. 93
Figure ‎5.4 High-resolution XPS scans for S2p: (a) virgin RO membrane, (b) AGE activated RO
membrane and (c) L-cysteine grafted membrane ............................................................................. 95
Figure ‎5.5 Atomic force microscopy images (including morphological statistics) of (a,b) polyamide
membranes ....................................................................................................................................... 96
Figure ‎5.6 Measured water contact angle results ............................................................................. 97
Figure ‎5.7 Measured zeta potential results ....................................................................................... 98
Figure ‎5.8 Water permeability and salt rejections of the membrane samples (control membrane is a
sample of virgin membrane which has been subjected to the same modification conditions as the
modified membrane but without any reactant) ................................................................................ 99
Figure ‎5.9 Resistance of the original and modified membranes as a function of time using
(100 mg/L of BSA, 10 mM NaCL, pH 6.4) and (20 mg/L DTAB, 10 mM NaCl, pH 6.2) as the
model foulants. Initial flux was 21.5±0.5 L/h.m2. Temperature was 22±2°C. ............................... 100
Figure ‎5.10 Resistance of the original and modified membranes as a function of permeate volume
using (100 mg/L of BSA, 10 mM NaCL, pH 6.4) and (20 mg/L DTAB, 10 mM NaCl, pH 6.2) as
the model foulants. Initial flux was 21.5±0.5 L/h.m2. Temperature was 22±2°C. ........................ 101
Figure ‎6.1 Probable polymerization pathway of L-DOPA during the modification process ......... 112
Figure ‎6.2 Measured zeta potential of the membranes [9] ............................................................. 115
Figure ‎6.3 Resistance of the original and modified membranes as a function of time using (100
mg/L of BSA, 10 mM NaCL, pH 6.4 ) and (100 mg/L alginate, 10mM NaCl, ) as model
biopolymer, Temperature: 20 ± 2°C, initial flux 21.2± 0.4 L/m2.h ............................................... 119

xi
Figure ‎6.4 Resistance of the original and modified membranes as a function of time using
(100 mg/L of BSA, 10 mM NaCL, pH 6.4 ) and (100 mg/L alginate, 10mM NaCl, ) as model
biopolymer, Temperature: 20 ± 2°C, initial flux 21.2± 0.4 L/m2.h................................................ 120

xii
List of Tables
Table ‎1.1 Characteristics of different membrane types used for water treatment [5] ................... 2
Table ‎1.2 chemical pre-treatment options for RO and NF membranes [20] ................................ 5
Table ‎2.1 Properties of selected inorganic foulants (from [7]) ................................................... 16
Table ‎2.2 Properties of selected organic foulants (from [7]) ...................................................... 17
Table ‎2.3 Selected coating methods ............................................................................................ 29
Table ‎2.4 Membrane modification using grafting approach ....................................................... 30
Table ‎2.5 Chemical structure of selected hydrophilic materials applied for membrane modification
..................................................................................................................................................... 31
Table ‎4.1 Water permeability, contact angle and surface charge of the original and coated
membranes .................................................................................................................................. 77
Table ‎5.1 Elemental composition of membranes ........................................................................ 94
Table ‎5.2 Relative abundance for different types of carbon in the membrane samples (C1: 287, C2:
289 and C3: 290.7 eV B.E. ) ....................................................................................................... 94
Table ‎5.3 similarities and differences of two methods used for surface modification of polyamide
thin film composite membranes ................................................................................................ 103
Table ‎6.1 Surface tension components (mJ/m2) of model liquids [11, 21] ............................... 113
Table ‎6.2 Measured zeta potential and particle size of the BSA and alginate .......................... 115
Table ‎6.3 Average values of the measured contact angle (in degrees) of water, glycerol and
diiodomethane as well as calculated surface tension parameters (in units of mJ/m2) of membranes
and model foulants .................................................................................................................... 116
Table ‎6.4 Calculated interaction energies (mJ/m2) ................................................................... 117

xiii
List of Publications

This thesis is partly based on the following publications.

 S. Azari and L. Zou 2012 "Using zwitterionic amino acid L-DOPA to modify the
surface of thin film composite polyamide reverse osmosis membranes to increase
their fouling resistance " Journal of Membrane Science, Vol 401-402, pp. 68-75.

 S. Azari and L. Zou 2013 "Fouling resistant zwitterionic surface modification of


reverse osmosis membranes using small amino acid L-cysteine" Desalination, Vol
324, pp. 79-86.

 S. Azari, L. Zou, E. Cornelissen and Y. Mukai 2013"Facile fouling resistant surface


modification of microfiltration cellulose acetate membranes by using amino acid
L-DOPA" Water Science and Technology, Vol 68, pp. 901-908

Other works produced during this PhD include:

 Nguyen, S. Azari, L. Zou 2013 "Coating zwitterionic amino acid L-DOPA to


increase fouling resistance of forward osmosis membrane" Desalination, Special
Issue: Advances in FO. Vol 312, pp.82-87.

 S. Azari, L. Zou, E. Cornelissen, "Assessing the effect of surface modification of


polyamide RO membrane by L-DOPA on the short range physicochemical
interactions between the membrane and biopolymers: elucidation of the role of L-
DOPA surface modification on membranes fouling behaviour" (manuscript
prepared to be submitted to Colloids and Surfaces B: Biointerfaces).

 S. Azari, L. Zou, E. Cornelissen, “Evaluating the effect of surface modification of


RO membrane on the interfacial interactions affecting initial stage of fouling"
accepted for oral presentation at the IWA Membrane Technology Conference
(IWA-MTC 2013), Toronto, Canada on August 25 - 29, 2013.

 S. Azari, L. Zou, E. Cornelissen, "Assessing the effect of L-DOPA surface


modification on the physicochemical interactions between RO membrane surface
and organic foulants" Accepted for oral presentation in International Membrane
Science and Technology Conference on 25 - 29 November 2013, at The University
of Melbourne, Australia.

xiv
 S. Azari, L. Zou, Y. Mukai, K. Takiguchi, "Protein Fouling of surface modified
Cellulose Acetate Microfiltration Membranes" Accepted for poster presentation
Euromembrane 2012 conference, September 2012, London, UK

xv
Glossary

Abbreviation Meaning

AB acid-base

ATR attenuated total reflection

BSA bovine serum albumin

CA cellulose acetate

EL electrostatic double layer

EPS extracellular polymeric substance

f foulant

FTIR fourier transform infrared

h hour

IEP isoelectric point

kDa kilo Dalton

kPa kilo Pascal

L litre

L-DOPA L-3,4-dihydroxy-phenylalanine

LW Lifshitz-van der Waals

mg/L milligram per litre

NF nanofiltration

NF nanofiltration

nm nanometer

NOM natural organic matter

PA polyamide

pzc point of zero charge

RO reverse osmosis

s solid

TDS total dissolved salt

xvi
UF ultrafiltration

UV ultraviolet light

w water

XPS X-ray photoelectron spectroscopy

XDLVO Extended Derjaguin-Landau-Verwey-Overbeek

Greek symbol

Lifshitz-Van der Waals free energy of interaction

electrostatic double layer free energy of interaction

Lewis acid-Base free energy of interaction

total free energy of interaction

cohesive free energy

surface charge

vacum permittivity

relativepermitivity of water

surface free energy

Lifshitz-Van der Waals component of the surface free enrgy

Lewis acid-Base component of the surface free energy

electron-acceptor component of the surface free energy

electron-donor component of the surface free energy

xvii
1 Introduction
1.1 Motivation of the work
Water scarcity is among the major problems that will impact many societies and the
world in the current century. Every continent is already faced with water shortage
issues. Approximately 1.2 billion people live in water stressed areas, and around 500
million are approaching this condition [1]. "Economic water shortage" where the
countries lack appropriate infrastructure to take water from the available sources, is also
i fl ci g 1.6 illio p opl ’ li .

The United Nations predicts that by 2030 the world population will reach 8.2 billion,
with 60% then living in cities and towns [2, 3]. Increase in world population and
urbanisation as well as waste, pollution and unsustainable management of water
resources will place significant pressure on the traditional water supply resources.
Th fo i i c i ic l o look fo l i ol io o pply fo h pl ’
ever-increasing population.

Brackish water and seawater as well as produced water from oil and natural gas are
recognised as alternative water supply sources; however, development of efficient low
energy processes to recover clean water from these sources has been a major challenge.
To resolve this issue, membrane technology has attracted a great deal of attention.

1.2 Membranes for water treatment


Membrane technology has emerged as a viable water and wastewater treatment
solution. For the past few decades, there has been a significant increase in interest in
application of membrane technology in the water industry. The membrane technology
market in water and wastewater treatment in Southeast Asia earned 249.1 million US
dollars in 2011 and this number is forecast to reach $398.0 million by 2017 [4].

Membrane processes require less addition of aggressive chemicals and produce no


hazardous or problematic by-products. Membrane filtration enables the removal of
contaminants which cannot be removed by alternative technologies. Moreover,
membrane technology can replace multiple unit treatment processes with a single

1
CHAPTER ONE

process thereby being more economic compared to the competing techniques. New
regulations imposed on potable water quality and industrial water treatment have also
turned membrane technology into a promising water treatment technique.

The most common membrane types used in the water industry are microfiltration (MF),
ultrafiltration (UF), nanofiltration (NF) and reverse osmosis (RO) [5]. MF membranes
are capable of removing only suspended particles, colloids and bacteria; UF membranes
can remove viruses, macromolecules and natural organic matter. Dissociated acids,
divalent ions and pharmaceuticals are removed by NF membranes and finally RO
membranes are able to retain almost all the solutes. Table 1.1 displays the
characteristics of different membrane types used for water treatment [5].

Table 1.1 Characteristics of different membrane types used for water treatment [5]

Membrane Pressure
Pore size Separation capability Application examples
type (bar)

RO <1nm Monovalent ions 10-60 Drinking water from


seawater

NF 1-10 nm Molecules with MWs of 5-25 Surface water or


200-20,000 kDa groundwater treatment
(multivalent ions, natural organic ma
tter)

UF 5-100 nm Macromolecules with MWs of 0.5-5 Wastewater treatment


10-500 kDa (Viruses, Bacteria)

MF 50 nm-5μ Bacteria and colloids 0.5-3 Pre-filtration in water


treatment

Forward osmosis (FO) has recently emerged as a new membrane process for water
desalination and wastewater treatment. Contrary to other membrane processes, forward
osmosis does not require hydraulic pressure to operate. In a typical forward osmosis
membrane process, water is transported due to the osmotic pressure difference between

2
CHAPTER ONE

the feed solution on one side and the draw solution on the other side of the membrane.
Selection of an appropriate draw solution and fabrication of high performance
membranes are still among the challenges for wide scale application of FO membrane
technology for water treatment. For more information about FO membrane process
readers may refer to the review by Zhao et al. [6].

1.3 Membranes for desalination of seawater and brackish


water
Seawater is obviously the most abundant source of water that can be desalinated for
potable water production and the production of high quality water for industrial
applications. Desalination has had a long history and has been used for thousands of
years. 4th century BC Greek sailors used to boil seawater to desalinate it and the
Romans used clay filters to separate salt [7]. Recently desalination technology has being
used for l g c l po c io . To y’ ophi ic li io ho f ll
into two main categories viz. thermal (distillation) processes and membrane processes.

A significant study on the application of a semi permeable cellulose acetate membrane


for water desalination was conducted in the 1950s at the University of Florida [8];
however, it did not achieve enough success. At that time water desalination was a
national high priority technology target. There was a slogan "Go to the moon and make
the desert bloom" [8]. Research supported by state and federal funding advanced the
science and technology of fresh water production from saline water [9]. In 1959
researchers in University of California-Los Angeles became the first to use the reverse
osmosis membrane process for saline water conversion [8, 9]. The Loeb-Sourirajan [10]
cellulose acetate membranes were the industry standard untill mid-1970s, when
Cadotte, at North Star Research, fabricated the first polyamide thin film
composite(TFC) membrane by the interfacial polymerization method[11] . TFC
membranes had higher permeability and salt rejection than cellulose acetate membranes.
Owing to the advancement of thin film composite membrane elements, today the
reverse osmosis industry can desalinate sea water with 10-15% less energy demand than
the early RO membranes [12].

Nowadays there are many membrane desalination plants around the world, with the
largest ones generally in the United Arab Emirates, Saudi Arabia and Israel. When the

3
CHAPTER ONE

expansion of the Ras Azzour plant in Saudi Arabia is accomplished, it will be the
largest thermal and membrane plant in the world, producing more than one billion litres
of p y. Co pl io of I l’ So q pl ill k i h l g
reverse osmosis membrane plant with a delivering capacity of 510 million litres water
per day [13].

In Australia, desalination plants have been constructed in major coastal cities in order to
deliver sufficient amounts of potable water for urban populations. In 2006 the first
major Australian seawater desalination plant was commissioned in Perth. The plant
could produce 45 billion litres of water per year. Later a second plant was built to add a
f h 1 illio li p y o P h’ ppli . O h A li jo
desalination plants have been established in Adelaide, Sydney and Gold Coast [7].

To sum up, membrane processes can play a considerable role in reducing water scarcity.
Their capabilities to treat water form resources such as ocean that were previously
inaccessible due to technical or economic considerations have been significant in
driving their use in water treatment. They can be utilised to treat wastewaters before
discharge to surface water, to recover materials from industrial water before they enter
waste streams and, of course to produce potable water.

1.4 Problem statement


The biggest technical issue with the use of membranes for water treatment which
challenges their economically sustainable development is the high potential for fouling.
Membrane fouling can be caused by colloids and suspended solids [14] , soluble
organic macromolecules [15] and microorganisms [16] that are typically not well
removed with conventional pre-treatment methods. Membrane fouling increases the
necessary feed pressure, causes decline in water quality and results in the need for
frequent membrane cleaning. This leads to reduced efficiency and shorter membrane
life. Frequent membrane replacement due to fouling costs the industry millions of
dollars. Understanding the causes of membrane fouling and finding strategies to prevent
fouling has therefore been an interesting and high relevant research topic over the past
decades [17-19].

4
CHAPTER ONE

1.5 Strategies to control fouling


1.5.1 Pre-treatment of feed water

The proper pre-treatment of raw water is necessary to approach a continuous and


reliable membrane operation. The type and extent of pre-treatment depends on the
source of the water (i.e. well water, seawater or wastewater). Pre-treatment can be
mechanical, chemical or a combination of both. Table 1.2 displays the potential
pre-treatment options for RO and NF membrane systems [20]

Table 1.2 chemical pre-treatment options for RO and NF membranes [20]

Chemical
Purpose
technique

Coagulants Improve precipitation and removal of solids

Antiscalants Reduce the crystallization of scale compounds on the membrane by being adsorbed
to the surface of microcrystals and increasing the solubility of the scale compounds

Dispersant Keep iron and other compounds from being suspended

Acids Changing the pH to the acidic range to reduce the scaling and adsorption potential
of some compounds

Bisulfites Typical chlorine reducing agent

UV To reduce biological activities

Ozone To remove organic compounds and also to reduce biological activities

1.5.2 Membrane fouling control by mechanical means

Pre-treatment steps not only fail to fully remove all the foulants from the feed water, but
can even contribute to membrane fouling by releasing some compounds into the feed
water. Some of the compounds used for pre-tratment actually cause membrane fouling.
Therefore additional strategies are also required to control membrane fouling.
Mechanical methods such as cross flow filtration, vibration enhanced membrane
filtration and rotation enhanced filtration [21] are among the mechanical techniques
which are used to mitigate membrane fouling.

5
CHAPTER ONE

Figure 1.1 Schematic illustration of cross-flow filtration versus dead end filtration [22]

Cross-flow filtration is one of the most well-established methods applied [23]. In this
method feed solution is pumped at high speed in relation to the membrane. High shear
and dynamic mass transport prevent the deposition of particles on the membrane
surface. Figure 1.1 shows the role of cross-flow filtration in the prevention of particle
deposition [22]. One of the biggest drawbacks of cross-flow filtration is its high energy
consumption. Due to the energy cost issues there is a limit on cross-flow velocity.
Therefore, cross-flow filtration cannot provide enough shear force to remove particles
retained between the turbulent boundary layer and the membrane surface. Consequently
fouling still occurs.

Vibration enhanced membrane filtration was introduced as an alternative method of


creating increased shear rate [21]. This method develops a shear at the membrane
surface which is not dependent upon the flow rate. Another approach is rotation
enhanced filtration which applies a rotating shaft in between the membrane plates to
create a high cross-flow velocity [21]. The rotation enhanced membrane separation and
membrane vibration techniques are normally used when a high concentration of feed is
applied. For water treatment and desalination purposes cross-flow filtration has
substantially been considered in the design of membrane modules.

6
CHAPTER ONE

1.5.3 Membrane surface modification

Based on what has been explored earlier, optimization of operation conditions


(pre-treatment and cross-flow filtration) cannot completely inhibit membrane fouling.
An alternative approach to mitigate fouling is modification of the membrane in a way to
reduce adhesion of particles, organic macromolecules and microorganisms to its
surface. Low fouling surface modification has been successfully used for biomedical
and nautical applications. This approach also has effectively been applied to MF, UF,
NF and RO membranes to reduce the membranes fouling tendency.

The potential benefits offered by membrane modification are:

 Less frequent membrane cleaning leading to cost savings in cleaning chemicals

 Lower RO pressure drops resulting in lower energy cost

 Longer RO membrane life

 Reduced environmental impact, due to the reduced need for cleaning chemicals

 Lower membrane handling and replacement costs.

Numerous studies have been conducted in developing new methods for membrane
modification [24-26]. Figure 1.2 shows the research papers published on membrane
surface modification topics which demonstrates the rising interest in this subject. Even
though many research studies have been devoted to develop new surface modification
approaches, there are limited materials and methods that can resolve the challenges of
practical applications.

7
CHAPTER ONE

Figure 1.2 Increasing research interest on membrane surface modification reflected in annual
academic publications (based on the web of science)

1.6 Objectives and the scope of the thesis


The main goal of this PhD project was to find novel and effective materials and
methods for fouling resistant surface modification of seawater desalination polyamide
RO membranes. In order to achieve this goal, two main objectives have been structured
and stated as follows:

1. To develop novel methods for low organic fouling surface modification of


polyamide RO membranes. Ideally, modification should be adaptable and
durable with the least possible effect on membrane separation performance.

2. To evaluate the performance of the surface modified membranes in terms of


water flux, salt rejection and fouling tendency.

To achieve these objectives, three phases of experimentation were designed and


conducted for each individual modification approach. The Initial phase involved design

8
CHAPTER ONE

and implementation of the modification process. It was followed in the second phase by
characterization of the modified membranes in terms of surface chemistry, surface
topography, surface charge and surface hydrophilicity. The third phase finally involved
evaluation of the modified membrane performance in terms of water flux, salt rejection
and fouling by synthetic water.

1.7 Thesis organization


This work aimed to explore surface modification strategies to control polyamide reverse
osmosis membrane fouling by organic foulants. The thesis is organized around the
following chapters:

Chapter 1 provides a basic outline of this PhD project by elaborating its significance,
research problems, objectives and potential benefits.

Chapter 2 provides a brief review of different types of fouling occurring during


membrane processes for water treatment and the effect of membrane fouling on
membrane performance with respect to permeability and separation. After a general
description of factors affecting membrane fouling, a brief review of membrane surface
modification methods is also presented. This chapter is designed to familiarize readers
with the background and basic knowledge of fouling and of antifouling surface
modification.

Chapter 3 describes a facile method of polyamide RO membrane surface modification


using a bioinspired adhesive, zwitterionic amino acid L-DOPA. The effect of this
modification on membrane surface characteristics is explored. Also membrane
performance in terms of flux, salt rejection and fouling behaviour is investigated.

Chapter 4 explores the application of the adhesive zwitterionic amino acid L-DOPA
for low fouling modification of a porous membrane viz. cellulose acetate. The effect of
modification on membrane surface characteristics, porosity and performance was
studied.

Chapter 5 presents a novel method toward fouling resistant surface modification of


polyamide RO membranes using the amino acid L-cysteine. L-cystein is grafted onto
the membrane by taking advantage of the reaction between its thiol group and the allyl

9
CHAPTER ONE

functionalized polyamide RO membrane. Sound characterizations of the membrane


surface were performed and the effect of modification on the fouling mitigation was
evaluated.

Chapter 6 describes a pioneering work in evaluating of the effect of membrane


modification on the fouling behaviour of the membranes. In this chapter the Extended
Derjaguin-Landau-Verwey-Overbeek (XDLVO) model was used to assess the effect of
modification on the free energy of adhesion between the membranes and model organic
foulants. The predictions obtained from the model are then compared with the
experimental outcomes.

Chapter 7 summarizes the outcomes of this study and puts forth a number of
recommendations for future work around this subject.

1.8 Contribution to knowledge


In order to devise the innovative fouling mitigation approach, taking into account the
complex underlying surface modification process, several technical problems need to be
tackled; experimental design, data acquisition, extensive data analysis and eventually
validation through a well-established thermodynamic model were involved. The
original contributions of this research, therefore, include the following:

(i) New implementation of bioadhesive L-DOPA modification for fouling


reduction. The exclusive top layer modification setup proposed in this
project offered an efficient single side membrane modification, but
prevented the deposition of unattached particles to the membrane surface
by providing a continues flow of the solution over the surface

(ii) The application of the zwitterionic amino acid L-cysteine was proposed
for the first time toward a high performance fouling resistance surface
modification of a membrane. A special chemical bonding method was
designed to protect the zwitterionic functional groups of L-cysteine.

(iii) A new implementation of Extended Derjaguin-Landau-


Verwey-Overbeek (XDLVO) model for theoretical evaluation of the
modification performance. The employed thermodynamic model enabled

10
CHAPTER ONE

the assessment of the short range interactions between membrane and


model foulants providing a profound understanding to the underlying
physicochemical adhesion mechanism.

1.9 Concluding remarks


This chapter provided readers with a basic outline of this PhD project by elaborating its
significance, research problems, objectives and potential benefits. An overview on what
should be expected in this thesis was given. The thesis structure was also included for
readers to understand the organization of this work.

11
CHAPTER ONE

References
[1] water scarcity. http://www.un.org/waterforlifedecade/scarcity.shtml ( Accessed Jun 2012).

[2] population trend. http://www.un.org/en/development/desa/population/events/expert-


group/6/index.shtml (Accessed Jun 2012).

[3] Population Division United Nations.


http://www.un.org/en/development/desa/population/theme/urbanization/index.shtml (Accessed
Jun 2012).

[4] Membrane Technologies Market in Water and Wastewater Treatment in Southeast Asia.
http://www.environmental.frost.com (Accessed Jun 2012).

[5] Water treatment membrane processes, McGraw-Hill, New York :, 1996.

[6] S. Zhao, L. Zou, C.Y. Tang, D. Mulcahy, Recent developments in forward osmosis:
Opportunities and challenges, Journal of Membrane Science, 396 (2012) 1-21.

[7] Desalination history,


http://www.water.vic.gov.au/initiatives/desalination/desalination/desalination-history (Accessed
Jun 2012).

[8] Water Desalination Processes, in, American Membrane Technology Association, 2007.

[9] First Demonstration Of Reverse Osmosis.


http://www.engineer.ucla.edu/explore/history/major-research-highlights/first-demonstration-of-
reverse-osmosis (Accessed Jun 2012).

[10] S. Loeb, S. Sourirajan, Sea Water Demineralization by Means of an Osmotic Membrane,


in: Saline Water Conversion, AMERICAN CHEMICAL SOCIETY, 1963, pp. 117-132.

[11] R. Baker, Membrane Technology and Applications (3rd Edition), 3rd ed. ed., Wiley-
Blackwell, John Wiley & Sons Inc, GB, 2012.

[12] C.Y. Tang, Y. Zhao, R. Wang, C. Hélix-Nielsen, A.G. Fane, Desalination by biomimetic
aquaporin membranes: Review of status and prospects, Desalination, 308 (2013) 34-40.

[13] Desalination Market 2013 – 2023 Commercial, Technological & Strategic Developments
in global desalination. www.gmrdata.com/app/download/8903296/DESAL+SAMPLE.pdf
(Accessed Jun 2012).

[14] C.A.C. van de Lisdonk, B.M. Rietman, S.G.J. Heijman, G.R. Sterk, J.C. Schippers,
Prediction of supersaturation and monitoring of scaling in reverse osmosis and nanofiltration
membrane systems, Desalination, 138 (2001) 259-270.

12
CHAPTER ONE

[15] A.S. Al-Amoudi, A.M. Farooque, Performance restoration and autopsy of NF membranes
used in seawater pretreatment, Desalination, 178 (2005) 261-271.

[16] A. Matin, Z. Khan, S.M.J. Zaidi, M.C. Boyce, Biofouling in reverse osmosis membranes
for seawater desalination: Phenomena and prevention, Desalination, 281 (2011) 1-16.

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


desalination, Desalination, 216 (2007) 1-76.

[18] K.V. Peinemann, S.P. Nunes, Membranes for water treatment, Wiley-VCH, Weinheim :,
2010.

[19] G. Mannina, A. Cosenza, The fouling phenomenon in membrane bioreactors: Assessment


of different strategies for energy saving, Journal of Membrane Science, 444 (2013) 332-344.

[20] Chemical Pretreatment For RO and NF. www.membranes.com/docs/tab/TAB111.pdf


(Accessed Jun 2012).

[21] M.Y. Jaffrin, Dynamic shear-enhanced membrane filtration: A review of rotating disks,
rotating membranes and vibrating systems, Journal of Membrane Science, 324 (2008) 7-25.

[22] tangential flow vs. dead end filtration. http://www.spectrumlabs.com/filtration/Edge.html


(Accessed Jun 2012), .

[23] M. Mulder, Basic Principles of Membrane Morphology, KLUWER ACADEMIC


PUBLISHER, 1996.

[24] J. Mansouri, S. Harrisson, V. Chen, Strategies for controlling biofouling in membrane


filtration systems: challenges and opportunities, Journal of Materials Chemistry, 20 (2010)
4567-4586.

[25] G. Kang, M. Liu, B. Lin, Y. Cao, Q. Yuan, A novel method of surface modification on
thin-film composite reverse osmosis membrane by grafting poly(ethylene glycol), Polymer, 48
(2007) 1165-1170.

[26] D. Rana, T. Matsuura, Surface Modifications for Antifouling Membranes, Chemical


Reviews, 110 (2010) 2448-2471.

13
2 Literature review

2.1 Introduction
This chapter introduces the different types of fouling occurring during the membrane
processes for water treatment. The effect of fouling on membrane performance with
respect to permeability and separation is briefly explained. After a general description
of factors affecting membrane fouling, a concise review of membrane surface
modification methods is also presented.

2.2 Classification of fouling agents and their properties


Several types of fouling have been recognized in membrane processes. Foulants
available in water from different sources can be classified into the following general
groups: mineral scales, colloidal foulants and biofoulants.

2.2.1 Mineral scales

Mineral scale formation usually occurs in reverse osmosis membranes particularly when
treating surface water or groundwater. Due to the concentration polarization effect, the
concentration of mineral salts near the membrane–water interface surpasses their
solubility limit. Therefore mineral salts become insoluble in water and get deposited on
the membrane surface [1, 2]. Deposition and crystallization of these mineral salts on the
membrane surface causes a decline in membrane performance. Various mineral salts
such as CaCO3, CaSO3. 2.H2O, BaSO4, Fe(OH)3 and SrSO4 have been reported to cause
membrane surface scale formation; however, scaling by CaSO3.2.H2O (calcium sulfate
dihydrate or gypsum) is the most common when treating surface and ground water [3].
Scale formation can be reduced by combining optimization of operating conditions to
minimise concentration polarization of scale formation and addition of antiscalant to the
feed solution to delay the crystallization of mineral salts [4]. Error! Reference source
not found. Figure 2.1 shows the SEM image of the effect of antiscalant on gypsum
crystallization and deposition on membrane surface [5].

14
CHAPTER TWO

Figure 2.1 SEM of calcium sulfate (gypsum) precipitate from a supersaturated solution on RO
membrane at 50 °C (left) and after addition of antiscalant (right) (from [5])

2.2.2 Colloidal foulants

Membrane fouling with colloidal and suspended particles in the size range of a few
nanometres to a few micrometers is referred to as colloidal fouling. Colloids can be
classified into the following general groups: inorganic colloids and organic
macromolecules.

Inorganic colloids are usually compact rigid particles with a variety of sizes and shapes.
The most common inorganic particles available in natural water resources are silica,
aluminium silicate and iron (oxy) hydroxide. Some inorganic colloids are naturally
occurring (silica and aluminium silicates), whereas others, are produced during the
water pre-treatment steps (e.g. aluminium or ferric hydroxide due to the use of
ferric/aluminium based coagulant) [6]. Table 2.1 shows the properties of selected
inorganic foulants [7].

15
CHAPTER TWO

Table 2.1 Properties of selected inorganic foulants (from [7])

Type of inorganic Size and shape charge


foulant

Silica Round Negatively charged (pH1pzc~3)

Aluminium silicate Angular Negatively charged at Neutral pH

Varies depending on crystalline or Positively charged (pHpzc~9 for


Ferric oxide/hydroxide
amorphous FeO(OH))

Organic macromolecules or organic foulants are described as those derived from living
materials, such as humic and fluvic acids and dissolved natural organic matters (NOM)
as well as those organic compounds from manufacturing operations, such as detergents,
fats, oils and greases. Organic materials cause strong and irreversible fouling. They may
have a rigid structure (e.g. large molecular weight polysaccharides) or a flexible one
(e.g. proteins). Dissolved natural organic matters have been implicated as the main
organic foulant when the membranes are used in the treatment of seawater, brackish
water and surface water [8]. Table 2.2 shows a brief summary of properties of selected
organic foulants available in natural waters [7]. Colloidal properties including their size,
shape, surface charge and their specific interaction with water or membrane surface
determine their behaviour in the solution and consequently affecting membrane fouling
propensity.

1
pHpzc is the pH value at which a colloidal particle submerged in an electrolyte exhibits zero net
electrical charge

16
CHAPTER TWO

Table 2.2 Properties of selected organic foulants (from [7])

Type of organic
Size and shape Charge
macromolecule

Humic acid
Globular (linear under high pH) Negatively charged
substances

Schizophyllan 400-500 KDa-rigid rod like Neutral

Polysaccharides Xanthan and gellan 100-2500 KDa-linear Negatively charged

Alginate 200-2000KDa-extended random coil Negatively charged

Bovine serum 64 kDa-globular pHIEP=4.7


albumin
proteins
Bovine hemogolobin 68 kDa-globular pHIEP=7.1

Lysozome 14.4 kDa-globular pHIEP=11

2.2.3 Biofoulants

Damage to the membrane and decrease in the membrane performance due to biofouling
is a major concern in water purification membrane systems [9, 10]. Biofouling
formation involves adhesion and growth of microorganisms onto the membrane surface.
These microorganisms include bacteria, viruses, protozoa, fungi and algae. It is hard to
control biofouling because bacteria are ubiquitous in water. If nutrients are available in
water, they can multiply even after their number has been significantly reduced by
disinfection or microfiltration (MF)/ultrafiltration (UF) membrane pre-treatment.
Extracellular polymeric substances (EPS) which form as a result of bacteria metabolic
activity enhance the survival and robustness of the already built up biofilm. EPS matrix
also adsorbs dissolved feedwater organics which enable bacteria to survive and grow.
Th , h iofil lif yl f cili h ic oo g i ’ survival and multiplication
even in extremely low nutrient environments. Some studies have been conducted to
evaluate the fundamental mechanisms of biofouling formation [11-13] resulting in a
common view that membrane physicochemical properties and feed solution
characteristics affect the initial stage of bacterial adhesion and biofilm development

17
CHAPTER TWO

[14]. However, biofilm formation still needs to be further investigated. Figure 2.2 shows
a biofouled membrane module.

Figure 2.2 Biofouled membrane module [15]

2.3 Effect of membrane fouling on membrane performance


Fouling can affect membrane performance through formation of a dynamic film
(surface layer or filter cake) on the front face of the membrane or by blocking the pores.
Both mechanisms introduce an additional hydraulic resistance and thus decrease
membrane flux. In a membrane process, either of these mechanisms can be dominant,
depending on the membrane structure or foulant types. In a reverse osmosis (RO) and
nanofiltration (NF) membrane system, cake layer formation dominantly controls
membrane fouling. The resistance of the formed cake layer depends on the applied
pressure, foulant type and solution conditions [16, 17]. For example, in terms of foulant
type, the resistance of cake layer formed by large mono dispersed inorganic particles
was found to be negligible [18] compared to the significant resistance of cake layer
formed from macromolecules [19].

The resistance of the cake layer alone cannot explain the large decline in the membrane
permeability. There is another phenomenon which also contributes towards the RO
membrane flux loss called cake enhanced osmotic pressure (CEOP). CEOP arises from
the greatly increased osmotic pressure due to the solutes accumulation on membrane

18
CHAPTER TWO

fluid interface caused by the cake layer. Actually the cake layer hinders back transport
of solutes from the membrane surface to the bulk solution [20]. It also prevents the
solutes to be exposed to shear forces resulting in a higher concentration of solutes on
membrane surface rather than in the bulk solution. Where the hydraulic resistance of
cake layer is relatively small (e.g. cake layer of large particles), CEOP effect can be the
dominant mechanism of flux loss [18].

In addition to membrane flux, membrane separation performance is also affected by


fouling. Fouling alters membrane properties that potentially affect solute removal,
including surface charge, hydrophobicity, pore size, functionality and surface
morphology. Fouling can either improve or deteriorate membrane performance [21-23].
For example, Lipp et al.[24] found that fouling of RO membranes by iron hydroxide
resulted in a decrease in salt rejection, whereas fouling by humic substances caused an
enhancement of salt rejection. Another study by van Ores et al. [21] showed that bovine
serum albumin (BSA) fouling improved the polyethylene glycol (PEG) rejection of the
UF membrane, but silica fouling decreased the PEG rejection of that membrane. They
proposed relative solute selectivity of the membrane and that of the fouling layer
determines the solute rejection. First state is when the membrane itself rejects solutes
better than the deposited layer and the back diffusion of rejected solutes is hindered by
the fouling layer. In this condition, a high concentration gradient across the membrane
forms which results in higher solute passage. Second state happens when the fouling
layer rejects solutes better than membrane. In this circumstance, the fouling layer
controls solute rejection thus rejection improves. When RO and NF membranes get
fouled with colloidal particles, first state is most likely to occur. For example, reported
increase in salt passage through the RO membrane fouled by iron oxide is in agreement
with the above statements [22, 24].

Fouling does not affect membrane performance in a consistent way. For example,
Agenson and Urase [25] observed an increase of high molecular weight solutes passage
through RO membrane fouled with activated sludge and landfill leachate. On the other
hand, they reported the decreased passage of low MW solutes by the fouled membrane.
The increased passage of the high molecular weight solutes was attributed to the effect
of cake-enhanced concentration polarization. Where, the decreased passage of low MW
ol l o h o po iz of h fo l ki g h “ iz

19
CHAPTER TWO

exclusion" responsible for the solute passage. Opposite to Agenson and Urase statement
about the effect of fouling on the high molecular weight solutes passage, Elimelech et
al. [26] found fouling minimally affects the large MW inert organic compounds
transport.

2.4 General factors affecting membrane organic fouling


It has been well recognized that membrane fouling can be extensively affected by
membrane properties and feed water composition. Appropriate understanding of these
factors can contribute to more affective fouling control. This section briefly explains the
effect of feed water composition, membrane properties and operational parameters on
membrane fouling by organic macromolecules. For a comprehensive review of factors
affecting membrane fouling readers refer to references [27, 28].

2.4.1 Effect of feed water characteristics on organic fouling

Membrane fouling is strongly impacted by foulant type, concentration and


physicochemical properties. Increasing foulant concentration generally results in a
decrease in the permeate flux [29] . A linear correlation has been obtained for fouling
indexes with the increase of concentration [30]. Some of the foulants physicochemical
properties (size, charge, conformation and functionality) are dependent on the solution
chemistry [31]. Macromolecules surface charge and conformation varies with solution
pH and ionic strength. The most common instance for the effect of solution pH and
ionic strength on the foulants physicochemical properties can be found in the case of
proteins, where changes in their conformation and charge strongly affect their stability
in water besides influencing protein–membrane interaction parameters [32, 33].

The ionic strength and pH value of the feed water solution not only impact the foulants,
but also can alter the membrane surface characteristics [34]. Hong and Elimelech [35]
conducted a study on humic acid fouling of a RO membrane associated with various
ionic strengths. Their work indicated a higher fouling rate, at higher ionic strength. This
was attributed to the lower electrostatic repulsion due to the double layer compression.

In addition to the aforementioned factors, existence of some divalent cations can also
drastically change the membrane fouling rate. The effects of divalent cations on NOM
fouling were examined by several researchers [31, 36, 37]. Membrane fouling by

20
CHAPTER TWO

alginate showed greater flux decline rate in the presence of Ca2+. This was caused by the
rise of alginate mass accumulation on membrane surface due to the formation of a
bridge between two alginate molecules [38]. Membrane fouling by humic acid
substances is also influenced by divalent cations. Interaction of humic carboxyl
functional groups with divalent cations can reduce the charge. This in turn would
decrease the electrostatic repulsion between humic macromolecules leading to higher
humic matters deposition rate [36].

Most of existing water resources contain mixed components such as solute organics,
particulate matters, colloidal particles and multiple electrolytes. The interaction of these
components with each other can also affect membrane fouling rate and propensity. For
instance, adsorption of some proteins such as BSA onto a negatively charged silica
colloid can reduce the surface charge of silica. This would consequently reduce the
electrostatic repulsion between silica and membrane as well as the repulsion between
silica particles [39] . Some studies revealed a synergic effect of the mixed foulant
systems on membrane fouling [40-42]. In this case fouling rate is remarkably greater
than the sum of the fouling rates corresponding to the individual foulants. For instance,
Ang and Elimelech [41] reported a significant increase in the BSA fouling rate of RO
membranes in the presence of alginate as the co-foulant. Synergistic fouling effect could
be explained by interactions between foulants. For example, adsorption of humic acid
onto kaolinite resulted in a stabilized cake layer of considerably reduced porosity
compared to individual particle filtration cake layer [40]. Wang and Tang [42]
investigated the protein fouling of NF, RO and UF membranes by BSA and lysozyme
(LYS) and their mixture. In the pH range between the isoelectric points (IEPs) of these
two proteins (i.e. pH 4.7–10.4), the mixed protein system experienced a more severe
flux decline compared to the respective single protein system. This can be attributed to
the electrostatic attraction between the negatively charged BSA and the positively
charged LYS molecules. Figure 2.3 shows a schematic of the effect of feed water
characteristics on membrane fouling by humic acid as a model NOM foulant [28].

21
CHAPTER TWO

Low ionic strength High ionic strength

High pH Low pH

Absence of divalent cations Presence of divalent cations

Stretched linear Coiled and compact


configuration in solution configuration in solution

Loose and thin fouling layer Compact, dense and thick fouling layer

Figure 2.3 Schematic illustration of the effect of feed water chemistry on Humic acid conformation
and the consequence effect on membrane fouling (adapted from [28])

2.4.2 Effect of operational conditions on organic fouling

Operational conditions such as hydrodynamic parameters can significantly affect the


membrane fouling rate. Cross-flow velocity and flux are two typical hydrodynamic
conditions that strongly influence fouling.

The effect of cross-flow velocity on membrane fouling has been widely studied [43-45].
In RO and NF membrane systems, particle transport toward the membrane surface
induces a force perpendicular to the surface (permeation drag force). Cross-flow also
induces a force tangential to surface. Therefore increasing the cross-flow velocity would
cause reduction of the boundary layer thickness and results in the decrease of
concentration polarization by enhancing the mass transfer and back transport
mechanism. Consequently, lower fouling is generally observed at higher cross-flow

22
CHAPTER TWO

velocity. For example, Chong et al. [18] noted an increase in membrane fouling by
silica when reducing the cross-flow velocity.

Several studies have been conducted to observe the effect of flux on the membrane
fouling. At high fluxes, large permeate volume causes a severe concentration
polarization and large hydrodynamic drag force toward the surface. Therefore high flux
leads to a more severe fouling on the membrane. Elimelech and co-workers [46, 47]
demonstrated the dependence of inorganic colloidal fouling of reverse osmosis
membranes on the permeate volume. Their other studies on membrane fouling by
macromolecules also consistently showed promotion of membrane fouling at higher
flux rates [31, 41]. Field and co- o k ’ [48] study on a MF membrane system
revealed that no fouling happens if the flux is under a certain value. Their work along
with some other studies led to the development of the "critical flux" concept. Critical
flux is described as the flux below which a decline of flux with time does not occur.
This concept was also applicable to RO and NF membrane systems [42][18]. Particles
transport phenomenon determines the value of critical flux. Therefore, parameters such
as hydrodynamic conditions, membrane morphology and feed water ionic strength,
which control particles transport, could affect the critical flux [49-51].

2.4.3 Effect of membrane properties

RO and NF Membrane fouling is greatly governed by membrane surface and


morphological properties. Hydrophobic and electrostatic interaction between organic
materials in the feedwater and membrane surface can potentially affect fouling.
Membrane surface morphology can also impact fouling via affecting the hydrodynamic
and surface interactions at liquid-membrane interface. The role of each of these factors
on membrane fouling will be further explained in the next two sections.

2.4.3.1 Surface hydrophilicity

There is a common point of view that hydrophilic membranes are less prone to fouling.
This can be attributed to high energy required to replace surface attached water with
foulant. Several research studies were conducted to study the effect of membrane
hydrophilicity on membrane fouling potential [52-54]. For example, Park et al. [55]
observed higher fouling resistance of hydrophilic piperaizine-based polyamide
membrane compared to the hydrophobic m-phenylene diamine-based membranes.

23
CHAPTER TWO

Investigation on commercially available fouling resistant PVA coated polyamide RO


membranes also revealed a smooth and hydrophilic surface of PVA coated RO
membranes [56].

2.4.3.2 Surface charge

Solute deposition on the membrane can be remarkably reduced if there is a repulsive


force between the charged membrane surface and particles in the feed solution. Most of
the colloidal particles such as NOM are negatively charged around neutral pH values
[31]. Therefore, a negatively charged surface is desirable to mitigate NOM to approach
the membrane. Since most proteins are negatively charged around neutral pH, a
negative surface charge is also desirable when separating proteins [57]. However, when
the feed water contains both negatively and positively charged foulants, a neutral
surface is preferred. Furthermore, where the feed water contains calcium ions, the
i gi g h c oxylic f c io l g o p of h fo l ’
functional groups with opposite polarity may increase the foulant adhesion rate [58, 59].
This suggests that a hydrophilic neutral membrane surface is ideal for a fouling resistant
membrane.

2.4.3.3 Surface roughness

Membrane surface roughness has been claimed to be one of the most important
parameters affecting membrane performance, particularly in respect of pemeability and
fouling behaviour. The influence of the membrane surface morphology on fouling with
organic and colloidal foulant has been widely investigated [46, 60]. Elimelech et al.
[61] found a strong correlation between the reduction in the energy barrier and the
magnitude of surface roughness. They elucidated that valleys in rough surfaces
represent wells of low interaction energy in which particles preferentially accumulate. A
study comparing cellulose acetate RO membrane with thin film composite (TFC)
polyamide RO membrane reported that CA acetate membrane with a smoother surface
is less prone to fouling. Li and colleagues [62] investigated the effect of membrane
surface properties on membrane fouling by biopolymers; they indicated that membrane
surface roughness dominantly controls fouling of the RO membrane. Tang et al. [63]
used RO and NF membranes with different surface roughness to treat a wastewater
sample containing toxic and persistent compound of perfluorooctane sulfonate; flux
reduction was found to correlate to membrane roughness, with the rougher membranes

24
CHAPTER TWO

showed more flux reduction than the smoother ones. Figure 2.4 schematically
compares the rough membrane surface with smooth surface in terms of foulant
accumulation rate. The foulant molecules accumulated in the valleys of rough surface
are less subject to hydraulic shear force, resulting in faster deposition of foulants on the
rough surface [62].

Figure 2.4 Schematic illustration of the effect of roughness on the deposition rate of fouants in a cross
flow filtration system (less foulant deposition on smooth surface) (adapted from [62])

According to literature It was discussed that feed water characteristics, operation


condition and membrane properties can affect the extent and rate of fouling. Figure 2.5
provides a comprehensive outline of the aforementioned factors affecting membrane
organic fouling.

Figure 2.5 Factors affecting organic fouling of RO and NF membranes (adapted from [7])

25
CHAPTER TWO

2.4.4 Effect of membrane–foulant and foulant–foulant interactions on fouling

Membrane fouling is affected by both foulant–membrane and foulant–foulant


interaction forces. Generally speaking, membrane fouling involves a sequence of two
stages: (i) foulant deposition onto the membrane surface and (ii) further foulant
deposition on the first deposited layer. Adhesion forces between clean membrane and
foulants determine the initial stage of fouling, whereas long term fouling behaviour is
controlled by interactions between the fouled membrane and foulants [64]. In this
connection, Tang et al. [19] studied humic acid fouling of a variety of RO and NF
membranes using a streaming potential analyser. They observed that while considerable
differences existed between the surface charges of the virgin membranes, humic acid
fouled membranes displayed identical surface charge properties. It was concluded that
humic acid completely cover the surface and thus the surface properties of the fouled
membranes was controlled by the humic acid layer. They also found that under mild
fouling conditions, RO membranes with smoother and more hydrophilic surface than
that of the NF membranes had less significant fouling by humic acid. However, at
severe fouling conditions and long fouling duration, this effect was hardly observed
[65]. Wang and Tang [42] studied protein fouling of NF, RO and UF membranes by
BSA, lysozyme and their mixture under cross-flow conditions. Under severe fouling
conditions membrane fouling was conceivably controlled by the interactions between
foulant and fouled membrane, whereas it was likely dominated by the interactions
between foulant and clean membrane under mild fouling conditions. With regard to
membrane biofouling, Lee et al. [66] studied the effect of membrane surface properties
on the initial stage of biofouling. They found that although the initial bacterial cell
adhesion was dependent on the membrane surface properties, further biofouling growth
was independent to the membrane surface properties. This provided a clue to the minor
role of membrane surface properties on the biofilm growth.

The interactions leading to membrane fouling by colloidal foulants are best described
by the Extended Derjaguin-Landau-Verwey-Overbeek (XDLVO) theory developed by
van Oss [67]. This theory considers three types of interactions, namely Lifshitz-van der
Waals (LW), acid-base (AB) and electrostatic (EL) double layer interactions. In aquatic
systems, the polar AB interaction can be more significant than the non-polar LW
interaction [67]. All these three interactions decrease as the colloid–membrane or

26
CHAPTER TWO

colloid–colloid separation distance (ds) increases. This theory has been frequently used
to describe colloidal attachment to the membrane surface [52, 53]. Lee et al. [68] used
this theory to study the NOM fouling of UF membranes. They also used it to elucidate
the mechanism of organic matter fouling on a commercial RO membrane [69].

The interaction energies are strongly dependent on various physicochemical parameters


such as particle size and zeta potential, membrane surface properties and feed water
chemistry (pH and ionic strength) [58]. In high ionic strength environments (e.g.
seawater), electrostatic forces are lowered due to the electrostatic double layer
compression. In agreement with this statement, higher fouling rates were observed for
negatively charged RO membranes during fouling by negatively charged biopolymers
in feed solutions with higher ionic strengths [41, 70].

2.5 Membrane surface modification approach to control


foulant adhesion
The major strategies used to control membrane fouling can be divided into two
categories: (i) optimizing the operational conditions such as cross-flow velocity and
pre-treatment of the feed solution and (ii) development of new membranes or
modification of existing membranes. In the following sections, the membrane surface
modification approach is explained with the examples of the methods and materials
commonly utilized for this purpose.

In the previous sections, it was indicated that membrane surface properties such as
roughness, charge and hydrophilicity are important in determining the susceptibility of
membranes to fouling. There have been few studies about the modification of surface
charge or surface roughness in order to minimise membrane fouling. For example, by
interposing an appropriate polymer layer between the polysulfone substrate and the
polyamide layer, a smooth membrane surface was obtained. This was the idea behind
forming the commercial low fouling GE osmonics Duraslick membrane. Ulbricht et al.
[71] modified the surface charge of polyacrylonitrile UF membrane by photo-initiated
graft polymerization with carboxylic acid and sulfonic acid. They observed a lower
protein fouling rate for a negatively charged modified membrane. In another study a
polyamide thin film composite RO membrane was modified by electrostatic

27
CHAPTER TWO

self-assembly of polyethyleneimine on the membrane surface. The modified membrane


showed significantly improved fouling resistance to cationic foulant [72] .

There are numerous reports available on the hydrophilic surface modification of the
membranes. These are explained briefly in the following section. For a comprehensive
review of membrane surface modification methods readers refer to references [73, 74].

2.5.1 Hydrophilic modification

The most common method employed for surface modification of membranes, especially
those used for water treatment, is hydrophilic modification. This process can be carried
out either through attachment of a hydrophilic polymer layer to the membrane surface
or through graft polymerization on the membrane surface. Hydrophilic modification
approaches including polymer coating, polymer grafting and polymer blending will be
described in this section.

2.5.1.1 Polymer coating

Polymer coating can be achieved by either covalent bonding of the modifying polymer
to the membrane substrate or by noncovalent attachment. Table 2.3 presents examples
of the use of coating approach for low fouling modification. Unlike grafting, membrane
coating is usually easy to apply and can be simply adapted to deal with different
membrane types.

28
CHAPTER TWO

Table 2.3 Selected coating methods

Modification Base membrane Function of the membrane

Water insoluble polyether–polyamide Polyamide RO Lower fouling rate of the modified


block copolymer adsorbed on the membrane membrane when exposed to
membrane (non covalent attachment) oil/surfactant/water feed emulsion
[75]

Hydrophilic poly (vinyl alcohol) coated on PES ultrafiltration Reduction of total and irreversible
the membrane through an adsorption- membrane fouling ratio of modified membranes
crosslinking process (non covalent [76]
attachment)

Coating with chitosan by dip and surface PVDF Modified membrane displayed
flow method (non covalent attachment) microfiltration antifouling properties when tested by
membrane BSA adsorption and filtration [77]

Epoxide functional PEG (poly Polyamide RO Resistance to fouling by surfactants


ethyleneglycol diglycidyl ether) reacted membrane (anionic and cationic) improved [78]
with the amine functional group of the
membrane (covalent attachment).

2.5.1.2 Graft polymerization

The Grafting approach to produce antifouling surfaces has been widely studied. An
advantage of this technique compared to the polymer deposition lies in its capability to
control the density of chains growing from the surface. Table 2.4 presents examples of
membrane modification utilizing the grafting approach.

29
CHAPTER TWO

Table 2.4 Membrane modification using grafting approach

Modification Base membrane Function of the membrane

Hydrophilic polymers (sulphopropyl Polyamide RO Modified membrane adsorbed


methacrylates and PEGMA) formed by (SWRO) membrane lower organic foulant when used
redox initiated grafting with vinyl for sea water/surface water
monomers filtration [79]

N-2-vinyl pyrolidinone (NVP) and acrylic PES ultrafiltration High protein retention, high protein
acid (AA) were grafted using UV membrane solution flux, and low irreversible
modification fouling were obtained with the
modified membrane [80]

PEG pre-coated on the membrane (dipping PVDF microfiltration protein adsorption experiments
method) followed by argon membrane revealed good antifouling
plasma-induced grafting of PEG on the properties for the modified
surface membrane [81]

PEG grafted onto the membrane using Ultrafiltration Modified membrane showed
sodium persulphate as oxidising agent for cellulose acetate higher fluxes, lower flux declines,
free radical development membrane and a more reversible fouling layer
[82]

2.5.1.3 Polymer blending

Blending the membrane polymer material with hydrophilic compounds in order to


improve membrane fouling resistance is also considered an attractive method for
modification of microporous polymer membranes (UF and NF membranes). The
blending approach can alter the microporous membrane surface characteristics with
only a slight influence on its bulk morphology and properties. As an example,
hydrophilic modification of polyethersulfone (PES) membranes was achieved by
blending a variety of Pluronic surfactants with different content and PEO chain lengths.
Increasing both Pluronic content or PEO chain length significantly improved the BSA
fouling resistance of the blended membranes [83].

30
CHAPTER TWO

Table 2.5 displays the chemical structures of the most common materials used for
hydrophilic modification.

Table 2.5 Chemical structure of selected hydrophilic materials applied for membrane modification

Polymer material Structure

Poly(vinyl alcohol)

Polyethylene glycol (PEG)

PEG diepoxide

Polyether-polyamide block
copolymer

Pluronic surfactant

Chitosan

Even though there have been numerous studies of materials and methods for low
fouling surface modification, there are still very few materials which can meet the
challenges of practical application. Poly (ethylene glycol) (PEG) has been used as the
most powerful antifouling material [84]. Nevertheless, undesirable oxidation of PEG in

31
CHAPTER TWO

the presence of oxygen or transition metal ions [85-87] besides loss of protein resistance
of PEG modified surfaces at above 35°C [88, 89] makes it necessary to look for an
alternative material.

2.5.2 Zwitterionic modification

Learning from nature, investigating the antifouling properties of biomaterials offers a


new approach for design and development of surface modification methods. The outer
surfaces of lipid bilayer membrane of biological cells have a high amount of
zwitterionic ligands [90]. Th c ll ’ i c o po i o p io oi h
researchers to use zwitterionic materials for reducing protein adsorption to synthetic
surfaces [91]. Surfaces modified with zwitterionic material can minimise nonspecific
protein adsorption to ultralow levels [92]. They also show high resistance to bacterial
adhesion and biofilm formation [93].

The non-fouling characteristic of zwitterionic materials is strongly associated with a


hydration layer near the surface. Strongly bound water forms a barrier layer which
effectively prevents close contact of protein molecules with the surface [92, 94]. In fact
water is attached to hydrophilic materials through hydrogen bonding, whereas it binds
to zwitterionic materials through ionic solvation. Therefore, compared to hydrophilic
materials, higher energy is required to remove bound water from zwitterionic materials.
Since adsorption of a protein to the surfaces involves expulsion of water from both
protein and the surface, zwitterionic materials are more stable against protein adsorption
than hydrophilic materials [95].

More recently, researchers have reported the potential application of zwitterionic


materials to modify the surface of filtration membranes. Shi et al. [96] first examined
the idea of using zwitterionic materials to reduce membrane fouling. They showed that
ultrafiltration membranes from the copolymer of sulfobetaine and polyethersulfone
were highly resistant to BSA fouling. In another study Sun et al. [97] reported the BSA
fouling resistance of ultrafiltration membranes fabricated from a copolymer blend of
sulfobetaine and acrylonitrile. Subsequent studies showed a great fouling resistance of
the microporous polypropylene membrane modified by grafting zwitterionic polymers
[98-100]. Due to the delicate structure of polyamide reverse osmosis and nanofiltration
membranes, only a few studies were found on the zwitterionic modification of these

32
CHAPTER TWO

membrane types. Chiang et al. [101] developed a method for interfacial


zwitterionization of NF polyamide membranes by N-alkylation of a secondary amine
group using iodopropionic acid and further quarternization with iodomethane. They
performed the adsorption test in phosphate buffer saline (neutral pH). An excellent
resistance to the adsorption of negatively charged BSA (IEP 4.7) as well as positively
charged egg white lysozyme (IEP 11.3) was observed. Yang et al. [102] synthesized a
copolymer containing sulfobetaine zwitterionic groups onto a RO membrane using an
initiated chemical vapour deposition technique. Compared with the bare RO membrane,
the coated one showed superior resistance against Escherichia coli adhesion. Zhao et al.
[99] tethered poly (sulfobetaine methacrylate) onto a polypropylene membrane surface
through a UV-induced surface graft polymerization with a subsequent surface-initiated
atom transfer radical polymerization. Experimental results demonstrated the good
antifouling performance of the modified polypropylene membrane. Figure 2.6 illustrates
two main structures of zwitterionic polymers [103].

Figure 2.6 Two main structures of zwitterionic materials a) poly(sulfobetaine) b) poly(carboxybetaine)


(adapted from [103])

In addition to the aforementioned zwitterionic materials, there are a few studies on the
application of natural and synthesised peptides as antifouling materials [104-106].
Peptides are short polymers composed of amino acids linked in sequences that are
specific to the particular peptide. They possess a mixed charge characteristic,
originating from the positively and negatively charged moieties of the different
amino acid residues. When these moieties are uniformly distributed at the molecular
level, peptides are equivalent to zwitterionic polymers. Successful application of natural
and synthesised peptides as antifouling materials [104-106] led to the idea of using
amino acids as zwitterionic surface modifying substances. Even though each amino acid

33
CHAPTER TWO

alone bears zwitterionic properties, only a few studies on the use of individual amino
acids for low fouling surface modifications have been published. Rosen et al. [107]
investigated the functionalization of silica nanoparticles with L-cysteine to create a low
fouling zwitterionic surface. The modified particles displayed excellent stability in
protein solutions. They used L-cysteine due to its low cost and ready availability and the
ability to conjugate additional molecules to it.

There are limited studies on the application of amino acids for low fouling surface
modification of polymeric membranes. Shi et al. [108] covalently bonded three types of
natural amino acids, including lysine, glycine and serine onto poly (acrylonitrile)
ultrafiltration membranes via carbodimide chemistry. Only the lysine retained its
zwitterionic properties after bonding to the membrane. They observed that although the
hydrophilicity of all the coated samples increased, it was only the lysine coated
membrane which showed superior protein fouling resistance. Rohani and Zydney [109]
studied zwitterionic modification of cellulose acetate UF membranes by covalent
attachment of small zwitterionic ligands viz. the aminoacids taurine, homotaurine and
glycine. UF experiments with basic, neutral, and acidic proteins (cytochrome
c, myoglobin α-lactalbumin) proved the resistance of zwitterionic modified
membranes against the adsorption of proteins even under conditions where the protein
and membrane had opposite polarity. The outcomes of the abovementioned studies
demonstrate the potential application of amino acids as zwitterionic ligands for surface
modification of polymeric membranes.

2.6 Conclusion: challenges and future opportunities


This review of the literature has outlined different types of membrane fouling and the
parameters affecting their fouling. It also represents a general overview on the
membrane surface modification methods with a particular focus on hydrophilic and
zwitterionic modifications.

It can be concluded that, membrane surface properties including surface roughness,


surface charge and surface hydrophilicity affect membrane fouling propensity. In
general, membranes with a higher hydrophilicity are less prone to fouling. Extensive
studies are available on the hydrophilic modification of membranes. However, there are
only few hydrophilic materials which can be employed for surface modification in

34
CHAPTER TWO

practical applications. To meet the demand for higher efficiency and stability in the
antifouling performance, zwitterionic materials have emerged as a new generation of
antifouling modifiers. There are however, only few studies available on zwitterionic
modification of membranes. In particular, due to the delicate structure of polyamide RO
membranes, which are the target membranes in this project, limited studies have been
found on their zwitterionic modification. These all motivated further research to explore
novel methods and materials for zwitterionic modification of polyamide RO
membranes.

The following chapters present results and discussion of an investigation undertaken to


apply zwitterionic amino acids to polyamide RO membranes and to determine if the
application of these zwitterionic ligands can effectively inhibit membrane fouling.

35
CHAPTER TWO

References

[1] C.A.C. van de Lisdonk, B.M. Rietman, S.G.J. Heijman, G.R. Sterk, J.C. Schippers,
Prediction of supersaturation and monitoring of scaling in reverse osmosis and nanofiltration
membrane systems, Desalination, 138 (2001) 259-270.

[2] C.A.C. van de Lisdonk, J.A.M. van Paassen, J.C. Schippers, Monitoring scaling in
nanofiltration and reverse osmosis membrane systems, Desalination, 132 (2000) 101-108.

[3] W.-Y. Shih, A. Rahardianto, R.-W. Lee, Y. Cohen, Morphometric characterization of


calcium sulfate dihydrate (gypsum) scale on reverse osmosis membranes, Journal of Membrane
Science, 252 (2005) 253-263.

[4] W.-Y. Shih, K. Albrecht, J. Glater, Y. Cohen, A dual-probe approach for evaluation of
gypsum crystallization in response to antiscalant treatment, Desalination, 169 (2004) 213-221.

[5] F. Rahman, Calcium sulfate precipitation studies with scale inhibitors for reverse osmosis
desalination, Desalination, 319 (2013) 79-84.

[6] R.Y. Ning, T.L. Troyer, Colloidal fouling of RO membranes following MF/UF in the
reclamation of municipal wastewater, Desalination, 208 (2007) 232-237.

[7] C.Y. Tang, T.H. Chong, A.G. Fane, Colloidal interactions and fouling of NF and RO
membranes: A review, Advances in Colloid and Interface Science, 164 (2011) 126-143.

[8] A.S. Al-Amoudi, A.M. Farooque, Performance restoration and autopsy of NF membranes
used in seawater pretreatment, Desalination, 178 (2005) 261-271.

[9] C.M. Pang, P. Hong, H. Guo, W.-T. Liu, Biofilm Formation Characteristics of Bacterial
Isolates Retrieved from a Reverse Osmosis Membrane, Environmental Science & Technology,
39 (2005) 7541-7550.

[10] A. Matin, Z. Khan, S.M.J. Zaidi, M.C. Boyce, Biofouling in reverse osmosis membranes
for seawater desalination: Phenomena and prevention, Desalination, 281 (2011) 1-16.

[11] H. Ivnitsky, D. Minz, L. Kautsky, A. Preis, A. Ostfeld, R. Semiat, C.G. Dosoretz,


Biofouling formation and modeling in nanofiltration membranes applied to wastewater
treatment, Journal of Membrane Science, 360 (2010) 165-173.

[12] M.M.T. Khan, P.S. Stewart, D.J. Moll, W.E. Mickols, M.D. Burr, S.E. Nelson, A.K.
Camper, Assessing biofouling on polyamide reverse osmosis (RO) membrane surfaces in a
laboratory system, Journal of Membrane Science, 349 (2010) 429-437.

36
CHAPTER TWO

[13] H. Ivnitsky, I. Katz, D. Minz, E. Shimoni, Y. Chen, J. Tarchitzky, R. Semiat, C.G.


Dosoretz, Characterization of membrane biofouling in nanofiltration processes of wastewater
treatment, Desalination, 185 (2005) 255-268.

[14] W. Lee, C.H. Ahn, S. Hong, S. Kim, S. Lee, Y. Baek, J. Yoon, Evaluation of surface
properties of reverse osmosis membranes on the initial biofouling stages under no filtration
condition, Journal of Membrane Science, 351 (2010) 112-122.

[15] Isolation and characterization of microorganism involved in membrane biofouling.


http://www.wageningenur.nl/en/show/Isolation-and-characterization-of-microorganism-
involved-in-membrane-biofouling-1.htm . (Accessed July 2012).

[16] C.Y. Tang, Y.-N. Kwon, J.O. Leckie, Fouling of reverse osmosis and nanofiltration
membranes by humic acid—Effects of solution composition and hydrodynamic conditions,
Journal of Membrane Science, 290 (2007) 86-94.

[17] S.G. Yiantsios, A.J. Karabelas, The effect of colloid stability on membrane fouling,
Desalination, 118 (1998) 143-152.

[18] T.H. Chong, F.S. Wong, A.G. Fane, Implications of critical flux and cake enhanced
osmotic pressure (CEOP) on colloidal fouling in reverse osmosis: Experimental observations,
Journal of Membrane Science, 314 (2008) 101-111.

[19] C.Y. Tang, Y.-N. Kwon, J.O. Leckie, Characterization of Humic Acid Fouled Reverse
Osmosis and Nanofiltration Membranes by Transmission Electron Microscopy and Streaming
Potential Measurements, Environmental Science & Technology, 41 (2006) 942-949.

[20] E.M.V. Hoek, M. Elimelech, Cake-E h c Co c io Pol iz io :  A N Fo li g


Mechanism for Salt-Rejecting Membranes, Environmental Science & Technology, 37 (2003)
5581-5588.

[21] C.W. van Oers, M.A.G. Vorstman, P.J.A.M. Kerkhof, Solute rejection in the presence of a
deposited layer during ultrafiltration, Journal of Membrane Science, 107 (1995) 173-192.

[22] J.M. Jackson, D. Landolt, About the mechanism of formation of iron hydroxide fouling
layers on reverse osmosis membranes, Desalination, 12 (1973) 361-378.

[23] E. Staude, W. Assenmacher, A contribution to fouling in hyperfiltration induced by


chemical reaction, Desalination, 49 (1984) 215-228.

[24] P. Lipp, R. Gimbel, F.H. Frimmel, Parameters influencing the rejection properties of FT30
membranes, Journal of Membrane Science, 95 (1994) 185-197.

37
CHAPTER TWO

[25] K.O. Agenson, T. Urase, Change in membrane performance due to organic fouling in
nanofiltration (NF)/reverse osmosis (RO) applications, Separation and Purification Technology,
55 (2007) 147-156.

[26] H.Y. Ng, M. Elimelech, Influence of colloidal fouling on rejection of trace organic
contaminants by reverse osmosis, Journal of Membrane Science, 244 (2004) 215-226.

[ ] C.H. Koo, A.W. Moh , F. S j ’, M.Z. M o Talib, Review of the effect of selected
physicochemical factors on membrane fouling propensity based on fouling indices,
Desalination, 287 (2012) 167-177.

[28] A.S. Al-Amoudi, Factors affecting natural organic matter (NOM) and scaling fouling in NF
membranes: A review, Desalination, 259 (2010) 1-10.

[29] C. Park, H. Kim, S. Hong, S.-I. Choi, Variation and prediction of membrane fouling index
under various feed water characteristics, Journal of Membrane Science, 284 (2006) 248-254.

[30] L.N. Sim, Y. Ye, V. Chen, A.G. Fane, Crossflow Sampler Modified Fouling Index
Ultrafiltration (CFS-MFIUF)—An alternative Fouling Index, Journal of Membrane Science,
360 (2010) 174-184.

[31] S. Hong, M. Elimelech, Chemical and physical aspects of natural organic matter (NOM)
fouling of nanofiltration membranes, Journal of Membrane Science, 132 (1997) 159-181.

[32] A.G. Fane, C.J.D. Fell, A. Suki, The effect of ph and ionic environment on the
ultrafiltration of protein solutions with retentive membranes, Journal of Membrane Science, 16
(1983) 195-210.

[33] H. Mo, K.G. Tay, H.Y. Ng, Fouling of reverse osmosis membrane by protein (BSA):
Effects of pH, calcium, magnesium, ionic strength and temperature, Journal of Membrane
Science, 315 (2008) 28-35.

[34] Y. Shim, H.-J. Lee, S. Lee, S.-H. Moon, J. Cho, Effects of Natural Organic Matter and
Ionic Species on Membrane Surface Charge, Environmental Science & Technology, 36 (2002)
3864-3871.

[35] W.S. Ang, S. Lee, M. Elimelech, Chemical and physical aspects of cleaning of organic-
fouled reverse osmosis membranes, Journal of Membrane Science, 272 (2006) 198-210.

[36] A.I. Schäfer, A.G. Fane, T.D. Waite, Nanofiltration of natural organic matter: Removal,
fouling and the influence of multivalent ions, Desalination, 118 (1998) 109-122.

38
CHAPTER TWO

[37] C. Jarusutthirak, S. Mattaraj, R. Jiraratananon, Influence of inorganic scalants and natural


organic matter on nanofiltration membrane fouling, Journal of Membrane Science, 287 (2007)
138-145.

[38] S. Lee, W.S. Ang, M. Elimelech, Fouling of reverse osmosis membranes by hydrophilic
organic matter: implications for water reuse, Desalination, 187 (2006) 313-321.

[39] A.E. Contreras, A. Kim, Q. Li, Combined fouling of nanofiltration membranes:


Mechanisms and effect of organic matter, Journal of Membrane Science, 327 (2009) 87-95.

[40] D. Jermann, W. Pronk, R. Kägi, M. Halbeisen, M. Boller, Influence of interactions between


NOM and particles on UF fouling mechanisms, Water Research, 42 (2008) 3870-3878.

[41] W.S. Ang, M. Elimelech, Protein (BSA) fouling of reverse osmosis membranes:
Implications for wastewater reclamation, Journal of Membrane Science, 296 (2007) 83-92.

[42] Y.-N. Wang, C.Y. Tang, Fouling of Nanofiltration, Reverse Osmosis, and Ultrafiltration
Membranes by Protein Mixtures: The Role of Inter-Foulant-Species Interaction, Environmental
Science & Technology, 45 (2011) 6373-6379.

[43] Ü. D iş, B. K ki l , Ch o o l fo i g ic ll h c
crossflow filtration: Effect of transmembrane pressure and crossflow velocity, Desalination, 249
(2009) 1356-1364.

[44] L.N. Sim, Y. Ye, V. Chen, A.G. Fane, Comparison of MFI-UF constant pressure, MFI-UF
constant flux and Crossflow Sampler-Modified Fouling Index Ultrafiltration (CFS-MFIUF),
Water Research, 45 (2011) 1639-1650.

[45] W.-M. Lu, S.-C. Ju, Selective Particle Deposition in Crossflow Filtration, Separation
Science and Technology, 24 (1989) 517-540.

[46] E.M. Vrijenhoek, S. Hong, M. Elimelech, Influence of membrane surface properties on


initial rate of colloidal fouling of reverse osmosis and nanofiltration membranes, Journal of
Membrane Science, 188 (2001) 115-128.

[4 ] X. Zh , M. Eli l ch, Colloi l Fo li g of R O oi M :  M


and Fouling Mechanisms, Environmental Science & Technology, 31 (1997) 3654-3662.

[48] R.W. Field, D. Wu, J.A. Howell, B.B. Gupta, Critical flux concept for microfiltration
fouling, Journal of Membrane Science, 100 (1995) 259-272.

[49] M. Mänttäri, M. Nyström, Critical flux in NF of high molar mass polysaccharides and
effluents from the paper industry, Journal of Membrane Science, 170 (2000) 257-273.

39
CHAPTER TWO

[50] J.A. Howell, Sub-critical flux operation of microfiltration, Journal of Membrane Science,
107 (1995) 165-171.

[51] I.H. Huisman, E. Vellenga, G. Trägårdh, C. Trägårdh, The influence of the membrane zeta
potential on the critical flux for crossflow microfiltration of particle suspensions, Journal of
Membrane Science, 156 (1999) 153-158.

[52] J.A. Brant, A.E. Childress, Assessing short-range membrane–colloid interactions using
surface energetics, Journal of Membrane Science, 203 (2002) 257-273.

[53] J.A. Brant, A.E. Childress, Colloidal adhesion to hydrophilic membrane surfaces, Journal
of Membrane Science, 241 (2004) 235-248.

[54] K. Boussu, A. Belpaire, A. Volodin, C. Van Haesendonck, P. Van der Meeren, C.


Vandecasteele, B. Van der Bruggen, Influence of membrane and colloid characteristics on
fouling of nanofiltration membranes, Journal of Membrane Science, 289 (2007) 220-230.

[55] N. Park, B. Kwon, I.S. Kim, J. Cho, Biofouling potential of various NF membranes with
respect to bacteria and their soluble microbial products (SMP): Characterizations, flux decline,
and transport parameters, Journal of Membrane Science, 258 (2005) 43-54.

[56] C.Y. Tang, Y.-N. Kwon, J.O. Leckie, Effect of membrane chemistry and coating layer on
physiochemical properties of thin film composite polyamide RO and NF membranes: I. FTIR
and XPS characterization of polyamide and coating layer chemistry, Desalination, 242 (2009)
149-167.

[57] H. Susanto, M. Ulbricht, Performance of surface modified polyethersulfone membranes for


ultrafiltration of aquatic humic substances, Desalination, 199 (2006) 384-386.

[58] X. Jin, X. Huang, E.M.V. Hoek, Role of Specific Ion Interactions in Seawater RO
Membrane Fouling by Alginic Acid, Environmental Science & Technology, 43 (2009) 3580-
3587.

[59] Q. Li, Z. Xu, I. Pinnau, Fouling of reverse osmosis membranes by biopolymers in


wastewater secondary effluent: Role of membrane surface properties and initial permeate flux,
Journal of Membrane Science, 290 (2007) 173-181.

[60] W. Richard Bowen, T.A. Doneva, Atomic Force Microscopy Studies of Membranes: Effect
of Surface Roughness on Double-Layer Interactions and Particle Adhesion, Journal of Colloid
and Interface Science, 229 (2000) 544-549.

40
CHAPTER TWO

[61] M. Elimelech, Z. Xiaohua, A.E. Childress, H. Seungkwan, Role of membrane surface


morphology in colloidal fouling of cellulose acetate and composite aromatic polyamide reverse
osmosis membranes, Journal of Membrane Science, 127 (1997) 101-109.

[62] Q. Li, Z. Xu, I. Pinnau, Fouling of reverse osmosis membranes by biopolymers in


wastewater secondary effluent: Role of membrane surface properties and initial permeate flux,
Journal of Membrane Science, 290 (2007) 173-181.

[63] C.Y. Tang, Q.S. Fu, C.S. Criddle, J.O. Leckie, Effect of Flux (Transmembrane Pressure)
and Membrane Properties on Fouling and Rejection of Reverse Osmosis and Nanofiltration
Membranes Treating Perfluorooctane Sulfonate Containing Wastewater, Environmental Science
& Technology, 41 (2007) 2008-2014.

[64] I.H. Huisman, P. Prádanos, A. Hernández, The effect of protein–protein and protein–
membrane interactions on membrane fouling in ultrafiltration, Journal of Membrane Science,
179 (2000) 79-90.

[65] C.Y. Tang, J.O. Leckie, Membrane Independent Limiting Flux for RO and NF Membranes
Fouled by Humic Acid, Environmental Science & Technology, 41 (2007) 4767-4773.

[66] W. Lee, S. Lee, C.H. Ahn, S. Hong, S. Kim, Y. Baek, J. Yoon, Evaluation of surface
properties of reverse osmosis membranes on the initial biofouling stages under no filtration
condition, Journal of Membrane Science, 351 (2010) 112-122.

[67] C.J. Van Oss, Interfacial forces in aqueous media, 2nd. ed., Taylor & Francis, Boca Raton,
Fla. :, 2006.

[68] S. Lee, S. Kim, J. Cho, E.M.V. Hoek, Natural organic matter fouling due to foulant–
membrane physicochemical interactions, Desalination, 202 (2007) 377-384.

[69] S. Kim, E.M.V. Hoek, Interactions controlling biopolymer fouling of reverse osmosis
membranes, Desalination, 202 (2007) 333-342.

[70] S. Lee, M. Elimelech, Relating Organic Fouling of Reverse Osmosis Membranes to


Intermolecular Adhesion Forces, Environmental Science & Technology, 40 (2006) 980-987.

[71] M. Ulbricht, K. Richau, H. Kamusewitz, Chemically and morphologically defined


ultrafiltration membrane surfaces prepared by heterogeneous photo-initiated graft
polymerization, Colloids and Surfaces A: Physicochemical and Engineering Aspects, 138
(1998) 353-366.

41
CHAPTER TWO

[72] Y. Zhou, S. Yu, C. Gao, X. Feng, Surface modification of thin film composite polyamide
membranes by electrostatic self deposition of polycations for improved fouling resistance,
Separation and Purification Technology, 66 (2009) 287-294.

[73] J. Mansouri, S. Harrisson, V. Chen, Strategies for controlling biofouling in membrane


filtration systems: challenges and opportunities, Journal of Materials Chemistry, 20 (2010)
4567-4586.

[74] D. Rana, T. Matsuura, Surface Modifications for Antifouling Membranes, Chemical


Reviews, 110 (2010) 2448-2471.

[75] J.S. Louie, I. Pinnau, I. Ciobanu, K.P. Ishida, A. Ng, M. Reinhard, Effects of polyether–
polyamide block copolymer coating on performance and fouling of reverse osmosis membranes,
Journal of Membrane Science, 280 (2006) 762-770.

[76] X. Ma, Y. Su, Q. Sun, Y. Wang, Z. Jiang, Enhancing the antifouling property of
polyethersulfone ultrafiltration membranes through surface adsorption-crosslinking of
poly(vinyl alcohol), Journal of Membrane Science, 300 (2007) 71-78.

[77] S. Boributh, A. Chanachai, R. Jiraratananon, Modification of PVDF membrane by chitosan


solution for reducing protein fouling, Journal of Membrane Science, 342 (2009) 97-104.

[78] E.M. Van Wagner, A.C. Sagle, M.M. Sharma, Y.-H. La, B.D. Freeman, Surface
modification of commercial polyamide desalination membranes using poly(ethylene glycol)
diglycidyl ether to enhance membrane fouling resistance, Journal of Membrane Science, 367
(2011) 273-287.

[79] S. Belfer, J. Gilron, Y. Purinson, R. Fainshtain, N. Daltrophe, M. Priel, B. Tenzer, A.


Toma, Effect of surface modification in preventing fouling of commercial SWRO membranes at
the Eilat seawater desalination pilot plant, Desalination, 139 (2001) 169-176.

[80] M. Taniguchi, G. Belfort, Low protein fouling synthetic membranes by UV-assisted


surface grafting modification: varying monomer type, Journal of Membrane Science, 231
(2004) 147-157.

[81] P. Wang, K.L. Tan, E.T. Kang, K.G. Neoh, Plasma-induced immobilization of
poly(ethylene glycol) onto poly(vinylidene fluoride) microporous membrane, Journal of
Membrane Science, 195 (2002) 103-114.

[82] T. Gullinkala, Evaluation of poly (ethylene glycol) grafting as a tool for improving
membrane performance, ProQuest, UMI Dissertations Publishing, 2010.

42
CHAPTER TWO

[83] Y.-Q. Wang, Y.-L. Su, X.-L. Ma, Q. Sun, Z.-Y. Jiang, Pluronic polymers and
polyethersulfone blend membranes with improved fouling-resistant ability and ultrafiltration
performance, Journal of Membrane Science, 283 (2006) 440-447.

[84] C.A. Finch, Poly(ethylene glycol) chemistry: Biotechnical and biomedical applications.
Edited by J. Milton Harris. Plenum Publishing, New York, 1992. pp. xxi + 385, price $89.00.
ISBN 0-306-44078-4, Polymer International, 33 (1994) 115-115.

[85] E. Ostuni, R.G. Chapman, R.E. Holmlin, S. Takayama, G.M. Whitesides, A Survey of
S c −P op y R l io hip of S f c h R i h A o p io of P o i , L g i,
17 (2001) 5605-5620.

[86] S. Krishnan, C.J. Weinman, C.K. Ober, Advances in polymers for anti-biofouling surfaces,
Journal of Materials Chemistry, 18 (2008) 3405-3413.

[87] M. Shen, L. Martinson, M.S. Wagner, D.G. Castner, B.D. Ratner, T.A. Horbett, PEO-like
plasma polymerized tetraglyme surface interactions with leukocytes and proteins: in vitro and in
vivo studies, Journal of Biomaterials Science, Polymer Edition, 13 (2002) 367-390.

[88] D. Leckband, S. Sheth, A. Halperin, Grafted poly(ethylene oxide) brushes as nonfouling


surface coatings, Journal of Biomaterials Science, Polymer Edition, 10 (1999) 1125-1147.

[89] A. Roosjen, H.C. van der Mei, H.J. Busscher, W. Norde, Microbial Adhesion to
Poly( hyl oxi B h :  I fl c of Poly Ch i L g h T p ,L g i,
20 (2004) 10949-10955.

[90] A.L. Lewis, Phosphorylcholine-based polymers and their use in the prevention of
biofouling, Colloids and Surfaces B: Biointerfaces, 18 (2000) 261-275.

[91] J.A. Hayward, D. Chapman, Biomembrane surfaces as models for polymer design: the
potential for haemocompatibility, Biomaterials, 5 (1984) 135-142.

[92] S. Chen, J. Zheng, L. Li, S. Jiang, Strong Resistance of Phosphorylcholine Self-Assembled


Mo ol y o Po i A o p io :  I igh i o No fo li g P op i of Z i io ic
Materials, Journal of the American Chemical Society, 127 (2005) 14473-14478.

[93] R.E. Holmlin, X. Chen, R.G. Chapman, S. Takayama, G.M. Whitesides, Zwitterionic
SAMs that Resist Nonspecific Adsorption of Protein from Aqueous Buffer, Langmuir, 17
(2001) 2841-2850.

[94] H. Kitano, T. Mori, Y. Takeuchi, S. Tada, M. Gemmei-Ide, Y. Yokoyama, M. Tanaka,


Structure of Water Incorporated in Sulfobetaine Polymer Films as Studied by ATR-FTIR,
Macromolecular Bioscience, 5 (2005) 314-321.

43
CHAPTER TWO

[95] S. Chen, L. Li, C. Zhao, J. Zheng, Surface hydration: Principles and applications toward
low-fouling/nonfouling biomaterials, Polymer, 51 (2010) 5283-5293.

[96] Q. Shi, Y. Su, W. Zhao, C. Li, Y. Hu, Z. Jiang, S. Zhu, Zwitterionic polyethersulfone
ultrafiltration membrane with superior antifouling property, Journal of Membrane Science, 319
(2008) 271-278.

[97] Q. Sun, Y. Su, X. Ma, Y. Wang, Z. Jiang, Improved antifouling property of zwitterionic
ultrafiltration membrane composed of acrylonitrile and sulfobetaine copolymer, Journal of
Membrane Science, 285 (2006) 299-305.

[98] Y.-F. Yang, Y. Li, Q.-L. Li, L.-S. Wan, Z.-K. Xu, Surface hydrophilization of microporous
polypropylene membrane by grafting zwitterionic polymer for anti-biofouling, Journal of
Membrane Science, 362 (2010) 255-264.

[99] Y.-H. Zhao, K.-H. Wee, R. Bai, Highly hydrophilic and low-protein-fouling polypropylene
membrane prepared by surface modification with sulfobetaine-based zwitterionic polymer
through a combined surface polymerization method, Journal of Membrane Science, 362 (2010)
326-333.

[100] J. Zhao, Q. Shi, S. Luan, L. Song, H. Yang, H. Shi, J. Jin, X. Li, J. Yin, P. Stagnaro,
Improved biocompatibility and antifouling property of polypropylene non-woven fabric
membrane by surface grafting zwitterionic polymer, Journal of Membrane Science, 369 (2011)
5-12.

[101] Y.-C. Chiang, Y. Chang, C.-J. Chuang, R.-C. Ruaan, A facile zwitterionization in the
interfacial modification of low bio-fouling nanofiltration membranes, Journal of Membrane
Science, 389 (2012) 76-82.

[102] R. Yang, J. Xu, G. Ozaydin-Ince, S.Y. Wong, K.K. Gleason, Surface-Tethered


Zwitterionic Ultrathin Antifouling Coatings on Reverse Osmosis Membranes by Initiated
Chemical Vapor Deposition, Chemistry of Materials, 23 (2011) 1263-1272.

[103] S. Jiang, Z. Cao, Ultralow-Fouling, Functionalizable, and Hydrolyzable Zwitterionic


Materials and Their Derivatives for Biological Applications, Advanced Materials, 22 (2010)
920-932.

[1 4] R. Ch l o ki, S.D. K , A. K , A. P k l , C. l , T. Wi kl , N.
M zl -Nol , A. T fo , C. W ll, P p i -Based SAMs that Resist the Adsorption of Proteins,
Journal of the American Chemical Society, 130 (2008) 14952-14953.

[105] S. Chen, Z. Cao, S. Jiang, Ultra-low fouling peptide surfaces derived from natural amino
acids, Biomaterials, 30 (2009) 5892-5896.

44
CHAPTER TWO

[106] S. Tosatti, S.M.D. Paul, A. Askendal, S. VandeVondele, J.A. Hubbell, P. Tengvall, M.


Textor, Peptide functionalized poly(l-lysine)-g-poly(ethylene glycol) on titanium: resistance to
protein adsorption in full heparinized human blood plasma, Biomaterials, 24 (2003) 4949-4958.

[107] J.E. Rosen, F.X. Gu, Surface Functionalization of Silica Nanoparticles with Cysteine: A
Low-Fouling Zwitterionic Surface, Langmuir, 27 (2011) 10507-10513.

[108] Q. Shi, Y. Su, W. Chen, J. Peng, L. Nie, L. Zhang, Z. Jiang, Grafting short-chain amino
acids onto membrane surfaces to resist protein fouling, Journal of Membrane Science, 366
(2011) 398-404.

[109] M.M. Rohani, A.L. Zydney, Protein transport through zwitterionic ultrafiltration
membranes, Journal of Membrane Science, 397–398 (2012) 1-8.

45
3 Using zwitterionic amino acid L-DOPA to modify the
surface of thin film composite polyamide reverse
osmosis membranes to increase their fouling
resistance

Abstract

This study focuses on the incorporation of redox functional amino acid


3-(3,4-Dihydroxyphenyl)-L-alanine (L-DOPA) onto commercial reverse osmosis (RO)
membranes (SW 30 XLE) to create a zwitterionic surface that resists membrane fouling.
The top layer of the membrane surface was modified by the deposition of L-DOPA from
its alkaline solution. Streaming potential and contact angle measurements were
conducted to characterise the membrane surface. Zeta potential data showed little or no
change in the surface charge of 24-hour-coated membrane relative to the virgin
membrane. The contact angle measurements indicated that the hydrophilicity of the
coated membranes was significantly improved. UV–visible spectroscopy confirmed the
presence of poly DOPA film on the coated membranes. Whilst the salt rejection
remained unchanged, a systematic increase in water flux was observed for the samples
coated up to 12 hours. Static BSA adhesion experiments revealed that L-DOPA coating
has reduced the amount of BSA adsorbed to the surface. To investigate the dynamic
fouling resistance of the membranes, a series of cross-flow filtration tests were carried
out using bovine serum albumin (BSA) and alginic acid sodium salt (alginate) solution
as the feed. The modified membranes exhibited lower fouling rate than the virgin
untreated membrane. The coated membranes also showed an improvement in their
fouling resistance when testing with the solution containing dodecyltrimethyl
ammonium bromide (DTAB), a positively charged surfactant model foulant.

This chapter is based on a published article (journal of membrane science 2012)

46
CHAPTER THREE

3.1 Introduction
I h f c of o l ’ co i i g fo i c pply i ili y,
there are growing interests in the reverse osmosis technique being used in water
desalination and waste water treatment. It is projected that water produced by
desalination will exceed 38 billion m3 per year in 2016 [1]. However, fouling is still a
major hindrance in the extensive application of reverse osmosis technology for water
treatment processes [2]. Dissolved organic matter and colloidal particles participate in
the formation of a fouling layer on the membrane surface, leading to a decline in
permeate flux and permeate quality, and reducing the membrane life time. To reduce the
operational costs, it is necessary to control membrane fouling. Membrane surface
modification is one proposed technique to reduce membrane fouling. It is common
knowledge that membranes with neutral, smooth and hydrophilic surfaces are less prone
to fouling [3, 4]. Therefore, most surface modification methods are designed to fulfil
these requirements [5]. Poly (ethylene glycol) (PEG)-based materials have proved to be
effective in improving fouling resistance of the membranes; employing both surface
grafting [6, 7] and coating [8, 9]. Surface hydration and steric repulsion are generally
considered the keys of PEG resistance to non-specific protein adsorption. Although
PEG showed excellent protein resistance ability, PEG-based materials are susceptible to
damage caused by oxidation and they have been reported to be only resistant to
short-term bacterial adhesion; however, the long-term biofilm formation still occurs
[10]. Moreover, the hydrated layer forms via a hydrogen bond network and this bond
can break easily. Therefore, the PEG-coated surface is subject to the transition from
non-fouling to fouling by any change in surface hydration [11].

Zwitterionic materials have been introduced as a new generation of non-fouling


materials and have been used to prepare fouling resistance surfaces since 1990 [12, 13].
They possess both negative and positive charged units and can create stronger and more
stable electrostatic bonds with water than hydrophilic materials [14, 15]. For example,
polysulfobetaine methacrylates, which is a zwitterionic compound comprising anionic
sulphate and cationic quaternary ammonium groups, was grafted on the poly(vinylidene
fluoride), polysulfone, polyethersulphone and microporous polypropylene membranes
and surface resistance to proteins has been achieved [16]. In recent years, some
innovative methods have been developed to anchor zwitterionic ligands on the surfaces

47
CHAPTER THREE

[17, 18] . However, these methods are rather chemistry-intensive and are not easy to
apply on the delicate structure of polyamide RO membranes [19]. A comprehensive
review of the application of zwitterionic meterials for surface modification of polymeric
membranes is presented in chapter two.

3-(3,4-Dihydroxyphenyl)-L-alanine, a zwitterionic (redox functional amino acid) that


was inspired from the adhesive proteins found in mussel, has been successfully
incorporated into various synthetic polymers in mild aqueous environments [20].
L-DOPA contains functional groups such as hydroxyl, amino, and carboxylate or acid
groups. Their state changes depend on the solution pH. It varies from cationic to
zwitterionic and then anionic as the solution acidity decreases. However, at the pH
values around its isoelectric point, there are always fractions available in the
zwitterionic form. The catechol side chain of L-DOPA is capable of many different
types of chemical interactions. In an elevated pH, such as in the marine environment, it
is particularly susceptible to take part in the crosslinking reaction. The oxidation of
L-DOPA to the reactive o-quinone allows the formation of di-DOPA crosslinks [21]. It
was found that upon oxidative crosslinking, robust and water resistant bonds form,
which bind firmly to diverse substrate surfaces. Thus, compared to the other methods,
the coating is simple and easily carried out in large-scale industrial applications.

3,4-Dihydroxyphenethylamine (dopamine), a catecholic derivatives of L-DOPA, is also


able to self-polymerize in aqueous solutions and form a strong attachment to a wide
g of . F ’ go p h p h of op i o h
membrane for antifouling purposes [22]. In their study, L-DOPA is neither mentioned
nor investigated [23]. Comprising the negatively charged carboxyl group and positively
charged amine group, qualifies L-DOPA as a zwitterionic molecule, whereas dopamine
does not. Until now, most studies on L-DOPA have been focused on using it as a spacer
for substrate functionalisation [24, 25]. There are quite few studies on its application as
the main membrane modifier agent [26]. However, there is no evidence of investigating
its fouling resistance characteristics as a zwitterionic compound and also to our
knowledge, the L-DOPA modification of thin film composite RO membranes has not
been reported yet. In this research, our main aim is to utilise amino acid L-DOPA to
create zwitterionic surfaces on reverse osmosis (RO) membranes to improve their
organic fouling resistance.

48
CHAPTER THREE

3.2 Experimental
3.2.1 Materials

SW30XLE polyamide thin film composite membranes manufactured by DOW Water


and Process Solutions were used in this study. SW30XlE is a low-energy seawater
desalination membrane. 3-(3,4-Dihydroxyphenyl)-L-alanine (L-DOPA),
Tris(hydroxymethyl)aminomethane (Tris), albumin from bovine serum (BSA), dodecyl
trimethyl-ammonium bromide (DTAB) and alginic acid sodium salt (from brown algae)
were obtained from Sigma–Aldrich and used as received.

3.2.2 Surface modification of the membrane

L-DOPA solution was prepared by dissolving in the Tris–HCL (10 mM, PH 8.3) buffer
solution. The concentration of the L-DOPA solution was 2.0 g/L. In order to prevent
adsorption of L-DOPA into the reinforcing fabric and polysulfone supporting layer of
the membrane, a controlled top-layer surface treatment method was employed. The
membrane coupons were soaked in DI water overnight to wash out the protective
agents. Then, the clean membrane was placed into a flat sheet cell that was connected to
a peristaltic pump (Aqua, Italy). The pump sucks the L-DOPA solution from a beaker
and then gently delivers it to the cell, flowing over the membrane surface. From the
outlet of the cell, the solution returns to the beaker and this circulation continues for the
required coating hours. During the coating time, the cell was kept in a vertical position
to prevent particle deposition on the membrane surface. Figure 3.1 shows the top side
coating method applied. The coated membranes were named according to the coating
time, for example, 4 h XLE refers to the samples that have been coated for 4 hours. It is
o h io i g h ho gh McClo k y’ g o p [23] obtained good results with
only a few minutes deposition of dopamine on the RO membranes, our preliminary
studies proved that L-DOPA modification cannot be effective in short coating times.
Consequently in this study the minimum coating time reported is 4 hours.

49
CHAPTER THREE

(a) (b)

Figure 3.1 Pictures of method and Apparatus used for top surface modification: (a) in the first
minutes of coating (b) after 12 hour of coating

3.2.3 Characterisation of the membrane surface

3.2.3.1 Streaming potential

The zeta potential of coated and uncoated membranes was determined using an Anton
Paar SurPASS Electrokinetic Analyser (Anton Paar USA, Ashland, VA) based on the
streaming potential measurements. Helmholtz-Smoluchowski approach was used to
evaluate the zeta potential from the streaming potential data. The results were obtained
using the clamping cell apparatus and the 10mM KCl as the background solution. The
initial pH was adjusted around 9 using 0.1 M NaOH. Afterward, the tests were carried
out by automatically dosing the solution with 0.1 M HCl until the pH was reduced to 3.
The instrument was set to measure the data at 0.5 pH intervals. At each pH value, four
measurements were recorded and the average was reported.

3.2.3.2 UV–visible spectroscopy

Varian Cary 100 UV–Vis spectrophotometer equipped with diffuse reflectance


accessories was used to investigate and prove the deposition of poly L-DOPA on
membrane surface. Scanning mode in the range of 200–800 was applied.

3.2.3.3 Contact angle measurement

Contact angle analysis was used to evaluate the effect of the coating on the membrane
hydrophilicity. Static water contact angle determined by the sessile drop method using a
OCA15 Contact angle analyser (Data-Physics GmbH, Filderstadt, Germany). A water
op of μL ih h of 1 μL/ pl c o h f c

50
CHAPTER THREE

side-view picture was captured after 10 s. Each measurement was made at least ten
times and the average was reported.

3.2.3.4 Atomic force microscopy

Topographical images of the membrane surface were obtained by atomic force


microscopy (AFM). The experiments were performed using NT-MDT NTEGRA SPM
atomic force microscope operated in non-contact mode. Silicon nitride non-contact tip
coated with gold on the reflective side (NT-MDT, NSG03) were used and had
resonance frequencies between 65 and 100 kHz. Images of 10µm x 10µm were obtained
with 10nm amplitude of oscillation and the scan rate of 1 Hz. The scanner had a range
of 100µm and was calibrated using 1.5µm standard grids. Images were processed using
WSxM software.

3.2.3.5 Static BSA surface adhesion measurement

The amount of protein adsorbed on virgin RO membranes and L-DOPA coated


membrane was measured by the solution depletion method. The method involves
placing the membrane into a dead end cell with the effective surface area of 28.3 cm2.
Then it was incubated with 12 mL of BSA solution with a concentration of 100 mg/L.
During the test, the cell was sealed to prevent solution evaporation. After the incubation
period of 3 hours, the amount of protein in the solution was determined by UV
measurement at 280 nm (Cary 100, Varian Inc.). Before each measurement, the cell was
gently swirled to return loosely coagulated proteins to the solution. The amount of
protein adsorbed on polymer surfaces was calculated from the total amount of protein in
the solution before and after adsorption. Each membrane type was tested twice and for
each test, adsorption experiments were repeated three times.

3.2.4 The permeation and salt rejection experiments

Salt rejections and water permeability of the bare and coated membranes were
measured by a stirred batch cell (Hp4750, Sterlitech Corp., WA, USA). The details of
the apparatus and the procedure for the data acquisition are described in chapter 5. All
experiments were performed under 50 bar transmembrane pressure difference and the
data were recorded following 1 hour filtration in order to avoid membrane compaction
affects the results. To calculate the membrane salt rejection, a conductivity metre (Hach
HQ 40 d) was used to measure the ion concentration in the feed and permeate solution.

51
CHAPTER THREE

For each type of membrane, three pieces were tested and the average values of the
membranes flux and salt rejection were reported.

3.2.5 Cross-flow filtration experiments and fouling characterisation

The fouling experiments were carried out in a laboratory scale cross-flow system. The
cross-flow test unit includes a Sepa CF II high-pressure membrane module (GE
Osmonics, Minnetonka, MN) with an active membrane area of 140 cm2 (14.6×9.5 cm),
high-pressure pump (Hydracell), feed tank, heat exchanger to control temperature, and a
digital balance and PC to collect data. The schematic of the unit is shown in Figure 3.2.

Figure 3.2 Schematic of the cross-flow test unit

Before each test, the system was cleaned with circulation of 200ppm sodium
hypochlorite solution for 30 minutes. Afterwards, the system was rinsed several times
with pure water.

For testing the virgin membrane, the applied pressure and the flow rate were adjusted to
18 bar and 0.75 L/min, respectively. Reynolds number for this module geometry was
estimated from the formula applied for a wide duct and was calculated to be 2600. This
Reynolds number falls within the transition region. The feed temperature was
maintained at 20oC by circulating chilled water. Permeate was collected in a beaker
placed on a digital balance (Mettle Toledo), which was connected to an associate

52
CHAPTER THREE

software (balance link) and was periodically (after any 4 hours) recycled back to the
tank to avoid concentrating of the solution. Following 12 hours of filtration of the pure
water, the solution with model organic foulant was introduced to the tank and the
membrane flux was recorded for the next 16 hours.

3.3 Results and discussion


L-DOPA, a catecholic amino acid, is one of the components that play a critical role in
enabling the mussels to attach to almost all types of inorganic and organic materials;
even to adhesion-resistant materials such as poly (tetrafluoroethylene) (PTFE). The
exact polymerisation mechanisms of L-DOPA in the basic environment have not yet
been determined. But according to the literature the reaction progresses via spontaneous
oxidation of L-DOPA to DOPA quinone form [21, 27].

Due to the exceptional coordination affinity of the catechol groups and the hydrophobic
backbone composed of benzene and indole rings, the poly L-DOPA have strong
adhesion in principle to various substrates despite the surface hydrophilicity of the
substrates. Although inner-sphere attachment through the phenolic oxygens has been
indicated in some studies [28, 29], the detailed bonding mechanism for its adhesion still
remains unclear. It is believed that poly L-DOPA comprise an equal number of
positively and negatively charged moieties within a molecule, thus maintaining overall
electrical neutrality, which can be strongly hydrated through ionic solvation.
Predictably, the formed hydrated surface effectively resists the membrane fouling.
Figure 3.3 illustrates the oxidative polymerization of L-DOPA and fouling resistance
behaviour of the poly L-DOPA-coated polyamide RO membrane. The surface
characteristics and fouling behaviour of the modified samples will be explained in
detail.

53
CHAPTER THREE

Figure 3.3 Oxidative polymerisation of L-DOPA and schematic of surface adsorption resistance for
organic matter imparted by the hydrated zwitterionic poly L-DOPA coated surface

3.3.1 Surface characterization

3.3.1.1 Surface charge

Membrane surface charge occurs from ionisable groups on the surface and has been
proven to promote membrane fouling when the foulants of opposite charge exist in the
feed solution. Thus, surface electrostatic properties and zeta potential can help to
anticipate the fouling behaviour. The zeta potential data of 24-hour-modified and
unmodified XLE membranes were measured in the presence of 1mM KCl solution and
the results are shown in Figure 3.4. As a result of presence of the residual unreacted
carboxylic acid and amine groups, the polyamide membranes possess a charged
property that is apparent from the results. The data for virgin XLE membrane are in the
range of values reported for polyamide membranes in the literature [30, 31]. Having
numbers of ionisable groups, poly L-DOPA is considered a charged material. However,
the measured zeta potential for the L-DOPA-coated membrane is equal to that of virgin
XLE membranes. For membrane coating with neutral materials it has been reported that
if the coating thickness is in nanometre scale, there would be no or little change in the
charge of the coated membrane [8]. A correlation between the graft density or the
coating layer thickness and zeta potential has been reported [32, 33]. Thus, from the
acquired data and literature reports, we may speculate that: (1) the coating layer is too

54
CHAPTER THREE

thin to cover the underlying layer charge, or (2) the coated material has similar zeta
potential to the virgin membrane.

20
15
10
virgin XLE
Zeta potential (mV)

5
0 24 h XLE

-5
-10
-15
-20
-25
-30
2 3 4 5 6 7 8 9 10
pH

Figure 3.4 Measured zeta potential results

3.3.1.2 UV–visible analysis of membrane surface

On visual inspection, modification of membranes with L-DOPA solution leads to the


formation of a dark brown coating on the membranes surface. However the film growth
during the coating cannot be detected visually. UV–visible results for the virgin
membrane samples and coated membranes are shown in Figure 3.5. It can be seen that
in the visible range virgin XLE membrane has almost zero absorption because of its
white and shiny appearance. Whilst in the UV range the absorption increase rapidly
from 370 nm with a broad peak at around 280 nm. In comparison with the virgin
membrane, 4 h XLE shows higher and monotonically increasing absorption at visible
range. In the UV range it also keeps rising though the absorbance of the virgin
membrane is still dominated and the peak at 280 nm is observed. The spectrum of
24 h XLE is similar to 4 h XLE but show higher absorption. However the peak seems a
bit flattened which means the DOPA film has more coverage on the membrane surface.
It has been reported that DOPA and dopamine films deposited from Tris buffer
solutions do not show any distinct peak at UV–visible range and only a monotone short

55
CHAPTER THREE

wavelength increase take place [27]. Therefore, the results for the L-DOPA coated
membranes are consistent with the literature.

1.6

1.4
virgin XLE
1.2 4 h XLE
24 h XLE
1
Absorbance

0.8

0.6

0.4

0.2

0
200 300 400 500 600 700 800
wavelength (nm)

Figure 3.5 UV–visible spectra for the virgin, 4 and 24-hour modified membranes

3.3.1.3 Surface hydrophilicity

A correlation between membrane surface hydrophilicity and membrane fouling


resistance has been addressed in literature [34, 35]. Higher fouling resistance is
expected for more hydrophillic surfaces. The existence of two full charges within the
poly L-DOPA molecule greatly increases its hydrophilicity. These positively and
negatively charged moieties strongly interact with water via ionic–dipole interactions.
To evaluate the hydrophilisation effect of coated poly L-DOPA, the water contact angle
of the membranes was measured. The results are reported in Figure 3.6. As shown, in
the first hours of coating, a rapid and significant decrease in water contact angle is
observed. The measured contact angle for the virgin XLE membrane is around 55
degrees. Whilst for the 4 h XLE, the water contact angle has been dropped to nearly half
of the XLE membrane. After about 5 hours of coating the contact angle still decrease
but with lower rate than the initial hours. It seems the contact angle is reaching a plateau
after about 12 hours coating.

56
CHAPTER THREE

60

50
Water contact angle (°)

40

30

20

10

0
0 5 10 15 20 25
Treatment time (h)

Figure 3.6 Water contact angle of the membranes as a function of modification time (test
temperature: 22°C)

As contact angle is affected by the surface top layer, the coating of the first monolayer
will significantly affect the contact angle, but further multiple layer coating will be built
on the first monolayer of poly DOPA, instead of membrane surface, so has much less
effect on contact angle. According to prior investigation [36] about polydopamin
deposition, it also may be speculated that factors such as depletion of the reactive
L-DOPA in the solution and intermolecular repulsion force between the first coated
layer of poly L-DOPA can hinders the film growth and consequently affect the
measured contact angle.

3.3.1.4 Surface roughness

The AFM images and the attached table in Figure 3.7 illustrate the surface
morphologies of the membranes tested. A smoother membrane surface is obtained for
24 h XLE membrane. It is clearly observed that the peak to valley distance on the
modified membrane is lower than that of virgin membrane. Attached table also indicates
a 33% and a 24% reduction in the surface root mean square (RMS) roughness and peak
to peak distance of the modified membrane, respectively.

Lower surface roughness of the modified membrane can be explained by the fact that
L-DOPA more readily deposits in the surface valleys. In this regard, Elimelech et al.
[37] found a strong correlation between the reduction in the energy barrier and the

57
CHAPTER THREE

magnitude of surface roughness. They elucidated that valleys in rough surface present
wells of low interaction energy in which particles preferentially accumulate.

Virgin
Virginmembrane
membrane 24
24hour
hourmodified
modified

Sample RMS roughness (nm) Peak-peak (nm)

Virgin membrane 113.29 832.1

24 hour modified membrane 75.72 630.1

Figure 3.7 Atomic force microscopy images (including morphological statistics)

3.3.1.5 Surface BSA adhesion

The amounts of proteins adsorbed on the virgin membrane and L-DOPA coated
membranes are presented in Figure 3.8. All membranes showed small amounts of BSA
adsorption. This is due to the fact that at the pH of deionised water, BSA carries a net
negative charge which repulsively interacts with negatively charged membrane surfaces.
It was found that both modified membranes adsorb less amounts of protein in
comparison to the virgin membrane. This observation agrees with the previous reports
in that under identical adsorption conditions, smaller amounts of protein adhere to the
more hydrophilic surfaces [38, 39]. However, surface hydrophilicity alone may not be
the best criteria as to whether proteins will adhere or not. The structure and orientation

58
CHAPTER THREE

of surface functional groups may also be involved in protein adsorption via direct
surface–protein interactions [40, 41].

5
4.5
Protein adsorption amount (µg/cm²)

4
3.5
3
2.5
2
1.5
1
0.5
0
virgin XLE 12 h XLE 24 h XLE

Figure 3.8 The measured amount of BSA adsorbed on the surface of virgin and modified membranes
from BSA solutions of 100 mg/L

3.3.2 Water flux and salt rejection of membranes

Th ’ p ili y l j c io i h i batch cell


unit are illustrated in Figure 3.9. It is observed that following the membrane coating
with L-DOPA, the salt rejection remained almost intact. Conversely, for the modified
membranes, the water flux increases with increasing the coating duration and reaches
the maximum in 12 hours coating time, which is measured to be 1.1 L/m2.h.bar, almost
1.27 times higher than the permeance of virgin membrane. Poly L-DOPA deposition
occurs via a mild oxidation mechanism and also according to the information catalogue
provided by the membrane manufacturer, for continuous operational procedures,
polyamide thin film composite membranes are stable in pH values between 2 and 11.
Accordingly, the increase in water flux may not be the reason of any damage to the
c .M hil , h fig fo ’ l j c ion also confirm
this statement. Yang et al. [19] also observed an increase in the water permeability of
RO membranes after they have been coated with ultrathin zwitterionic materials. They
reported that the water permeation through an RO membrane consists of two main

59
CHAPTER THREE

steps. First, water molecules are adsorbed onto the membrane surface, and then these
molecules will diffuse through the coating layer and polyamide active layer. So a very
thin coating does not have a considerable effect on diffusion step but can accelerate the
adsorption step. For L-DOPA coating a thickness of 10s nanometre have been reported
[36]. Therefore it is concluded that the increase in water permeability is caused by
increase in membrane hydrophilicity. However, with further increase of the coating
time, a slight reduction in membranes water permeability is observed. As a result of RO
’ o po o nature of the active layer, deposition occurs on the membrane
active surface and not into the pores. Thus, following the modification, no pore
blockage occurs; as the coating proceeds, a diffusion resistant layer gradually forms on
the membrane active surface and adds to the water diffusion resistance of the barrier
layer.

1.09
membrane permeance (L/m².h.bar)
membrane salt rejection (%)
1.02
0.99

0.979 0.97 0.97


0.97 0.97 0.97

0.86

virgin 4 h XLE 12 h XLE 18 h XLE 24 h XLE

Figure 3.9 Water permeability and salt rejections of the virgin XLE membrane and the membranes
modified with different modification times

60
CHAPTER THREE

3.3.3 Filtration experiment

To investigate the fouling behaviour of the membranes on dynamic operation, a series


of cross-flow filtration experiments were conducted. Since the initial water flux can
affect the extent of fouling observed, it is preferred to compare the performance of the
membranes with similar initial water flux. In order to achieve this, the pressure was
adjusted so that all the membranes have the same initial flux as the unmodified
membrane. In the first set of experiments, BSA and alginate were used as the model
protein and polysaccharide for organic foulants. Using BSA and alginate together has
been proven to have an accelerating effect on the membrane fouling [42]. The
BSA/alginate fouling data of virgin XLE membrane as well as 12 h XLE and 24 h XLE
membranes are summarised in Figure 3.10 and Figure 3.11.

6
membrane resistance×10-12 (1/m)

5.5 virgin XLE 12 h XLE 24 h XLE

4.5

3.5

3
0 200 400 600 800 1000
Time (min)

Figure 3.10 Resistance of the virgin and modified membranes as a function of time during
2
BSA/alginate (100 mg/L of each) fouling, initial flux 21± 0.4 L/m .h

61
CHAPTER THREE

6
membrane resistance × 10-12 (1/m)
5.5 virgin membrane

5 12 h XLE

4.5
24 h XLE

3.5

3
0 200 400 600 800 1000 1200 1400
permeate volume

Figure 3.11 Resistance of the virgin and modified membranes as a function of permeate volume
during BSA/alginate (100 mg/L of each) fouling, initial flux 21 ± 0.4 L/m2.h

It can be seen that in the first hours of the test, the resistance of virgin membrane
increases at a higher rate than that of either the 12 h XLE or 24 h XLE membranes. For
virgin and 12 hour modified membrane, two stages of fouling can be distinguished.
During the early stage of fouling, the resistance of the membranes increased rapidly,
whereas in the second stage, the resistance increase rate gradually slowed down. It can
be explained by the fact that in the initial hours of filtration, the adsorption of BSA and
alginate is governed by the membrane-foulant interaction where at the second stage; it is
controlled by the foulant–foulant interactions. 24-hour modified membrane o ’
display similar trend of that of virgin and 12 h XLE membranes; a small decrease in the
resistance of the 24-hour modified membrane can be attributed to a small increase in
water temperature during the initial stage of the test. Even though the 24-hour modified
membrane has a higher initial resistance than that of the virgin membranes, its final
flow resistance is lower than that of both the 12 h XLE and virgin membrane which
proves its higher fouling resistance.

To compare the antifouling behaviour of the surface modified membranes relative to


unmodified membranes in the presence of either negatively charged or positively
charged foulants, DTAB which is a common cationic surfactant present in the waste
water was also applied as a model foulant. A series of fouling experiments were
performed using the aqueous solutions containing 50 mg/L DTAB as the feed solution.

62
CHAPTER THREE

Membrane resistance vs. time and permeate volume were plotted and the results are
presented in Figure 3.12 and Figure 3.13, respectively.

4.8

4.6
membrane resistance ×10-12 (1/m)

4.4

4.2

4 24 h XLE

3.8
12 h XLE

3.6
virgin membrane
3.4
0 50 100 150 200 250 300 350
Time (min)

Figure 3.12 Resistance of the virgin and modified membranes as a function of time during DTAB
2
(50 mg/L) fouling. initial flux 21.5± 0.5 L/m .h

4.8
membrane resistance × 10-12 (1/m)

4.6

4.4

4.2

3.8 24 h XLE

3.6 12 h XLE

3.4 virgin membrane

3.2
0 100 200 300 400 500 600
permeate volume

Figure 3.13 Resistance of the virgin and modified membranes as a function of permeate volume
2
during DTAB (50 mg/L) fouling, initial flux 21.5± 0.5 L/m .h

63
CHAPTER THREE

As shown, in case of DTAB fouling, the resistance increase is faster than BSA/Alginate
fouling. It can be clearly seen that even though the initial resistance of the virgin
membrane is lower than that of the 24 hour modified membrane, after around 5 hours of
filtration its resistance becomes higher than those of both modified membranes.
Aforementioned results prove that the surface modification of the membrane with
L-DOPA can also improve the membrane resistance to the adsorption of amphiphilic
cationic molecule of DTAB.

3.4 Conclusion
Multifunctional amino acid L-DOPA has been applied as a zwitterionic coating to
SW30XLE RO membranes to alter their surface properties for fouling-resistant
purposes. A controlled top-layer surface treatment method was employed to deposit
L-DOPA on the membrane surface. Compared to virgin membrane, modified membrane
possessed a smoother surface. After L-DOPA modific io , h ’
hydrophilicity has been remarkably improved. Due to the increase in hydrophilicity, the
water permeability of the membranes increased, while the salt rejection was maintained.
The enhancement in water flux was observed for up to 12 hours of modification.
However, further deposition of L-DOPA caused a slight decrease in water permeability,
which is attributed to the formation of a diffusion resistant layer. Static BSA adsorption
measurements did show an improvement in the BSA adhesion resistance of the coated
membrane. Dynamic filtration experiments proved that modification has enhanced the
organic and surfactant fouling resistance of the modified membranes. Since membrane
modification can be performed in situ, this approach is of particular interest from a
practical point of view, as the membranes can be treated in their virgin module
assembly post-manufacture process.

64
CHAPTER THREE

References
[1] M. Elimelech, W.A. Phillip, The Future of Seawater Desalination: Energy, Technology, and
the Environment, Science, 333 (2011) 712-717.

[2] J.S. Baker, L.Y. Dudley, Biofouling in membrane systems -- A review, Desalination, 118
(1998) 81-89.

[3] M. Pasmore, P. Todd, S. Smith, D. Baker, J. Silverstein, D. Coons, C.N. Bowman, Effects of
ultrafiltration membrane surface properties on Pseudomonas aeruginosa biofilm initiation for
the purpose of reducing biofouling, Journal of Membrane Science, 194 (2001) 15-32.

[4] E.M. Vrijenhoek, S. Hong, M. Elimelech, Influence of membrane surface properties on


initial rate of colloidal fouling of reverse osmosis and nanofiltration membranes, Journal of
Membrane Science, 188 (2001) 115-128.

[5] J. Mansouri, S. Harrisson, V. Chen, Strategies for controlling biofouling in membrane


filtration systems: challenges and opportunities, Journal of Materials Chemistry, 20 (2010)
4567-4586.

[6] S. Belfer, Y. Purinson, R. Fainshtein, Y. Radchenko, O. Kedem, Surface modification of


commercial composite polyamide reverse osmosis membranes, Journal of Membrane Science,
139 (1998) 175-181.

[7] E.M. Van Wagner, A.C. Sagle, M.M. Sharma, Y.-H. La, B.D. Freeman, Surface
modification of commercial polyamide desalination membranes using poly(ethylene glycol)
diglycidyl ether to enhance membrane fouling resistance, Journal of Membrane Science, 367
(2011) 273-287.

[8] A.C. Sagle, E.M. Van Wagner, H. Ju, B.D. McCloskey, B.D. Freeman, M.M. Sharma, PEG-
coated reverse osmosis membranes: Desalination properties and fouling resistance, Journal of
Membrane Science, 340 (2009) 92-108.

[9] J.S. Louie, I. Pinnau, I. Ciobanu, K.P. Ishida, A. Ng, M. Reinhard, Effects of polyether–
polyamide block copolymer coating on performance and fouling of reverse osmosis membranes,
Journal of Membrane Science, 280 (2006) 762-770.

[10] A. Roosjen, H.C. van der Mei, H.J. Busscher, W. Norde, Microbial Adhesion to
Poly( hyl oxi B h :  I fl c of Poly Ch i L g h T p ,L g i,
20 (2004) 10949-10955.

65
CHAPTER THREE

[11] R.E. Holmlin, X. Chen, R.G. Chapman, S. Takayama, G.M. Whitesides, Zwitterionic
SAMs that Resist Nonspecific Adsorption of Protein from Aqueous Buffer, Langmuir, 17
(2001) 2841-2850.

[12] S. Chen, L. Li, C. Zhao, J. Zheng, Surface hydration: Principles and applications toward
low-fouling/nonfouling biomaterials, Polymer, 51 (2010) 5283-5293.

[13] L. Andrew L, Phosphorylcholine-based polymers and their use in the prevention of


biofouling, Colloids and Surfaces B: Biointerfaces, 18 (2000) 261-275.

[14] H. Kitano, T. Mori, Y. Takeuchi, S. Tada, M. Gemmei-Ide, Y. Yokoyama, M. Tanaka,


Structure of Water Incorporated in Sulfobetaine Polymer Films as Studied by ATR-FTIR,
Macromolecular Bioscience, 5 (2005) 314-321.

[15] S. Chen, J. Zheng, L. Li, S. Jiang, Strong Resistance of Phosphorylcholine Self-Assembled


Monolayers to Protein Adsorptio :  I igh i o No fo li g P op i of Z i io ic
Materials, Journal of the American Chemical Society, 127 (2005) 14473-14478.

[16] Y.-F. Yang, Y. Li, Q.-L. Li, L.-S. Wan, Z.-K. Xu, Surface hydrophilization of microporous
polypropylene membrane by grafting zwitterionic polymer for anti-biofouling, Journal of
Membrane Science, 362 (2010) 255-264.

[17] J. Zhao, Q. Shi, S. Luan, L. Song, H. Yang, H. Shi, J. Jin, X. Li, J. Yin, P. Stagnaro,
Improved biocompatibility and antifouling property of polypropylene non-woven fabric
membrane by surface grafting zwitterionic polymer, Journal of Membrane Science, 369 (2011)
5-12.

[18] Y.-H. Zhao, K.-H. Wee, R. Bai, Highly hydrophilic and low-protein-fouling polypropylene
membrane prepared by surface modification with sulfobetaine-based zwitterionic polymer
through a combined surface polymerization method, Journal of Membrane Science, 362 (2010)
326-333.

[19] R. Yang, J. Xu, G. Ozaydin-Ince, S.Y. Wong, K.K. Gleason, Surface-Tethered Zwitterionic
Ultrathin Antifouling Coatings on Reverse Osmosis Membranes by Initiated Chemical Vapor
Deposition, Chemistry of Materials, 23 (2011) 1263-1272.

[20] H. Lee, S.M. Dellatore, W.M. Miller, P.B. Messersmith, Mussel-Inspired Surface
Chemistry for Multifunctional Coatings, Science, 318 (2007) 426-430.

[21] M. Nicolai, G. Gonçalves, F. Natalio, M. Humanes, Biocatalytic formation of synthetic


melanin: The role of vanadium haloperoxidases, L-DOPA and iodide, Journal of Inorganic
Biochemistry, 105 (2011) 887-893.

66
CHAPTER THREE

[22] B.D.A. Freeman, TX, US) , H.B.A. Park, TX, US) , B.D.A. Mccloskey, TX, US) Water
purification membranes with improved fouling resistance in, Board of Regents The University
of Texas System (Austin, TX, US) United States, 2011.

[23] B.D. McCloskey, H.B. Park, H. Ju, B.W. Rowe, D.J. Miller, B.J. Chun, K. Kin, B.D.
Freeman, Influence of polydopamine deposition conditions on pure water flux and foulant
adhesion resistance of reverse osmosis, ultrafiltration, and microfiltration membranes, Polymer,
51 (2010) 3472-3485.

[24] J.L. Dalsin, B.-H. Hu, B.P. Lee, P.B. Messersmith, Mussel Adhesive Protein Mimetic
Polymers for the Preparation of Nonfouling Surfaces, Journal of the American Chemical
Society, 125 (2003) 4253-4258.

[25] C. Gao, G. Li, H. Xue, W. Yang, F. Zhang, S. Jiang, Functionalizable and ultra-low fouling
zwitterionic surfaces via adhesive mussel mimetic linkages, Biomaterials, 31 (2010) 1486-1492.

[26] Z.-Y. Xi, Y.-Y. Xu, L.-P. Zhu, Y. Wang, B.-K. Zhu, A facile method of surface
modification for hydrophobic polymer membranes based on the adhesive behavior of
poly(DOPA) and poly(dopamine), Journal of Membrane Science, 327 (2009) 244-253.

[27] F. Bernsmann, V. Ball, F.d.r. Addiego, A. Ponche, M. Michel, J.J.d.A. Gracio, V.r.
To i zzo, D. R ch, Dop i −M l i Fil D po i io D p o h U Oxi
Buffer Solution, Langmuir, 27 (2011) 2819-2825.

[28] S. Bahri, C.M. Jonsson, C.L. Jonsson, D. Azzolini, D.A. Sverjensky, R.M. Hazen,
Adsorption and Surface Complexation Study of L-DOPA o R il (α-TiO2) in NaCl Solutions,
Environmental Science & Technology, 45 (2011) 3959-3966.

[29] D.A. Stern, G.N. Salaita, F. Lu, J.W. McCargar, N. Batina, D.G. Frank, L. Laguren-
Davidson, C.H. Lin, N. Walton, Studies of L-DOPA and related compounds adsorbed from
aqueous solutions at platinum(100) and platinum(111): electron energy-loss spectroscopy,
Auger spectroscopy, and electrochemistry, Langmuir, 4 (1988) 711-722.

[30] M. Elimelech, W.H. Chen, J.J. Waypa, Measuring the zeta (electrokinetic) potential of
reverse osmosis membranes by a streaming potential analyzer, Desalination, 95 (1994) 269-286.

[31] A.E. Childress, M. Elimelech, Effect of solution chemistry on the surface charge of
polymeric reverse osmosis and nanofiltration membranes, Journal of Membrane Science, 119
(1996) 253-268.

[32] Y.-H.M. Chan, R. Schweiss, C. Werner, M. Grunze, Electrokinetic Characterization of


Oligo- and Poly(ethylene glycol)-Terminated Self-Assembled Monolayers on Gold and Glass
Surfaces, Langmuir, 19 (2003) 7380-7385.

67
CHAPTER THREE

[33] H. Susanto, M. Ulbricht, Photografted Thin Polymer Hydrogel Layers on PES


Ul fil io M :  Ch c iz io , S ili y, I fl c o S p io
Performance, Langmuir, 23 (2007) 7818-7830.

[34] D.S. Wavhal, E.R. Fisher, Hydrophilic modification of polyethersulfone membranes by


low temperature plasma-induced graft polymerization, Journal of Membrane Science, 209
(2002) 255-269.

[35] H. Ju, B.D. McCloskey, A.C. Sagle, V.A. Kusuma, B.D. Freeman, Preparation and
characterization of crosslinked poly(ethylene glycol) diacrylate hydrogels as fouling-resistant
membrane coating materials, Journal of Membrane Science, 330 (2009) 180-188.

[36] B. Li, W. Liu, Z. Jiang, X. Dong, B. Wang, Y. Zhong, Ultrathin and Stable Active Layer of
Dense Composite Membrane Enabled by Poly(dopamine), Langmuir, 25 (2009) 7368-7374.

[37] M. Elimelech, Z. Xiaohua, A.E. Childress, H. Seungkwan, Role of membrane surface


morphology in colloidal fouling of cellulose acetate and composite aromatic polyamide reverse
osmosis membranes, Journal of Membrane Science, 127 (1997) 101-109.

[38] J. Kim, W. Qian, Z.Y. Al-Saigh, Measurements of water sorption enthalpy on polymer
surfaces and its effect on protein adsorption, Surface Science, 605 (2011) 419-423.

[39] A. Sethuraman, M. Han, R.S. Kane, G. Belfort, Effect of Surface Wettability on the
Adhesion of Proteins, Langmuir, 20 (2004) 7779-7788.

[40] Y. Arima, H. Iwata, Effect of wettability and surface functional groups on protein
adsorption and cell adhesion using well-defined mixed self-assembled monolayers,
Biomaterials, 28 (2007) 3074-3082.

[41] M.S. Wang, L.B. Palmer, J.D. Schwartz, A. Razatos, Evaluating Protein Attraction and
Adhesion to Biomaterials with the Atomic Force Microscope, Langmuir, 20 (2004) 7753-7759.

[42] W.S. Ang, M. Elimelech, Protein (BSA) fouling of reverse osmosis membranes:
Implications for wastewater reclamation, Journal of Membrane Science, 296 (2007) 83-92.

68
4 Facile fouling resistant surface modification of
microfiltration cellulose acetate membranes by
using amino acid L-DOPA

Abstract

A major obstacle in the widespread application of microfiltration (MF) membranes in


the wet separation processes such as wastewater treatment is the decline of permeates
flux as a result of fouling. This study reports on the surface modification of cellulose
acetate microfiltration membrane with amino acid 3-(3,4-Dihydroxyphenyl)-L-alanine
(L-DOPA) to improve fouling resistance of the membrane. The membrane surface was
characterised using Fourier transform infrared spectroscopy (FTIR), water contact angle
and zeta potential measurement. Porosity measurement showed a slight decrease in
membranes porosity due to coating. Static adsorption experiments revealed an improved
resistance of the modified membranes towards the adhesion of bovine serum albumin
(BSA) as the model foulant. Membrane filtration tests using a stirred batch cell
exhibited that the fouling resistance of the modified membranes was improved.
However, the effect of the modification depended on the foulant solution concentration.
It is concluded that L-DOPA modification is a convenient and non-destructive approach
to enable low-BSA adhesion surface modification of CA microfiltration membranes.

This chapter is based on published article (Water Science and Technology2013)


69
CHAPTER FOUR

4.1 Introduction
The microfiltration (MF) process is an attractive method for the pre-treatment of
seawater and wastewater. It can remove bacteria and particulate matter and can partially
remove viruses from water[1, 2]. Separation by MF membranes is also employed in the
pharmaceutical and food industries [3]. However, MF processes suffer from a decline in
performance due to membrane fouling by proteins and other biomolecules [4]. Even
though an optimal design of operating conditions and careful selection of the membrane
type to the feed solution can help to reduce the degree of fouling, the performance of the
MF membrane system eventually declines [2, 3].

In recent years, there has been an increased interest in developing techniques to alter the
physical and chemical properties of MF membrane surfaces in order to minimise
membrane fouling by reducing foulants adhesion to the membrane surface [5-7].
Proteins are one group of foulants which cause a major challenge in MF membranes
fouling. Therefore there have been a number of studies for low protein adsorption
surface modification of these membranes. Akhtar et al.[8] reported on membrane
modification by grafting 2-methacryloyloxyethyl phosphorycholine onto
polyvinylidene difluoride (PVDF) and cellulose acetate (CA) MF membranes surfaces.
After two hours of filtration using the bovine serum albumin (BSA) solution as the
model protein foulant, they observed 70 and 30 percent less flux decline for the coated
PVDF and CA membranes, respectively. Liu et al.[9] immobilised
poly(n-vinyl-2-pyrrolidone) (PVP) on polypropylene microfiltration membranes. They
observed that as more PVP immobilised on the surface, less static and dynamic BSA
fouling occurred. However, some other researchers have found membrane fouling to be
independent of the membrane surface chemistry, as it occurs primarily due to the
physical deposition of foulant components. Wang et al. [10] grafted hydroxyethyl
methacrylates onto the polypropylene (PP) membranes to make the membranes more
hydrophilic. Despite the fact that the coated membranes could reach nearly zero BSA
adsorption, a similar flux decline trend was observed for both the coated and uncoated
membranes. Mueller and Davis [11] investigated the hydrophilic modification of
polyethylene and PP membranes by polyvinyl alcohol coating. Their results also
showed similar BSA fouling behaviour for both the original and modified membranes.

70
CHAPTER FOUR

This study presents the chemical deposition of 3-(3,4-Dihydroxyphenyl)-L-alanine


(L-DOPA) onto CA microfiltration membranes. This is done to investigate the effect of
L-DOPA coating on the membrane fouling caused by BSA solutions. The present study
was motivated by the past work that reported on the development of a low fouling
zwitterionic interface on the thin film composite reverse osmosis membrane surface
following L-DOPA deposition [12].

L-DOPA is a critical functional element in forming the adhesive properties of marine


mussle proteins [13]. A very significant characteristic of L-DOPA is its ability to self-
polymerize in aqueous solutions then forming a strong and water resistance attachment
to the substrate via their catechol group [14]. Previous studies applied L-DOPA onto
hydrophobic porous membranes in order to increase the membrane hydrophilicity [15,
16]. It also has been used as spacer for further functionalization of the substrates [17,
18]. Even though L-DOPA has been previously applied on the membranes, to the
ch ’k o l g , h i o po o L-DOPA coating for reducing the protein
fouling of microfiltration membranes. Despite the fact that most membrane
modification methods are complex and difficult to adopt in industrial applications, the
deposition of poly L-DOPA from aqueous solutions of catecholamine L-DOPA offers a
convenient way to functionalize surfaces, regardless of the surface chemistry of the
substrate [14].

4.2 Experimental
4.2.1 Materials

Experiments were performed using cellulose acetate (CA) microfiltration (MF)


membranes, obtained from Toyo Roshi Corp, with a nominal pore size of 0.2 µm.
Bovine serum albumin (BSA) (fraction V) was supplied by Sigma Aldrich and used in
all experiments as the model protein foulant. L-DOPA and
tris(hydroxymethyl)aminomethane (Tris) were used for membrane surface modification.
These were also obtained from Sigma Aldrich and used as received.

4.2.2 Methods

4.2.2.1 Membrane surface modification

The CA membranes were pre-treated by soaking in ultrapure water for two hours to
remove the glycerin present in virgin membranes. L-DOPA solution was prepared by

71
CHAPTER FOUR

dissolving L-DOPA in the Tris–HCl (10 mM, pH 7.9) buffer solution. The modification
was performed by soaking the membranes in a 2 g /L L-DOPA solution for eight, 16 and
24 hours. During the coating, the samples were subjected to a gentle shaking in order to
prevent non-specific deposition of microparticles.

4.2.2.2 Membrane surface characterization

Attenuated total reflectance spectroscopy (ATR-FTIR) was used to analyse the chemical
structure of the membranes. An FTIR analyser (Perkin Elmer), equipped with an ATR
accessory (diamond crystal), was applied to collect the infrared spectra. Each spectrum
was taken using four scans at a resolution of 4 cm-1 in the region between 650 and
4,000 cm-1. Spectrum software provided with the instrument was used to analyse the
data.

The hydrophilicity of the membranes was assessed using static water contact angle
measurement. Contact angle results were determined by the captive bubble method
using a OCA15 contact angle analyser (Data-Physics GmbH, Filderstadt, Germany).
Membrane sample was immersed in a transparent water container upside down and
placed on the goniometer. An air droplet was placed on the inverted membrane surface
and the contact angle was measured. The values for the contact angle were averaged
over 18 data points by using three identical membrane samples with three air bubbles on
each sample. For the porous and hydrophilic CA membrane, captive bubble method is
preferred because it better represents the state of the membrane in actual wet operating
conditions and is less sensitive to the moisture content of the membrane.

The electrostatic interactions between the protein molecule and the membrane surface
have a considerable role in the BSA adsorption behaviour [19]. Therefore the
’ f c ch g i g S PASS El c oki ic A ly
(Anton Paar GmbH). The Helmholtz-Smoluchowski approach was employed to
evaluate the zeta potential from the streaming potential data [20]. The tests were carried
out using an adjustable-gap cell apparatus. Measurements were performed from alkaline
to acidic pH, in a 10 mM KCl solution as the background solution at ambient
temperature. The initial pH was adjusted at 8.5 using 0.1 M NaOH. After that, the tests
were carried out by automatically dosing the solution with 0.1 M HCl until the pH was
reduced to 5.5. In each pH point four measurements was recorded.

72
CHAPTER FOUR

Porosity of the membrane samples was obtained by water evaporation method using the
following formula.

Where ww and wd denote the weight of wet and dry sample respectively. and
represent the water and cellulose acetate polymer density. To measure the weights of the
wet and dry sample, firstly, a pre-soaked membrane (quarter-circle with the surface area
of 35.3 mm2) was pat dried using a lint free wipe in order to remove the excess water
from the surface. The membrane was then weighted with a high precision electronic
scale (Metler Toledo AL 204). In the next stage, the membrane was heated in the oven
at 40°C for 2 hours to evaporate the water. The dried sample was later cooled down in a
desiccator and weight measurement was performed afterwards. The test was repeated
for two samples of each membrane type and the average was reported.

4.2.2.3 Filtration apparatus and experiments

Measurements of the pure water permeability of the membrane samples and


microfiltration experiments were carried out at room temperature using a dead end cell
set-up, as is demonstrated in Figure 4.1. The filtration cell fits a membrane with the
surface area of 7.65 cm2. The water permeabilities of the membranes were measured
using de-ionised water as the feed. A constant transmembrane pressure of 20 kPa was
applied and the membranes were pre-compacted for one hour. The flux was measured
every five minutes until it reached a constant value in at least three readings. The
g of fi i g po h ’ fl x.

73
CHAPTER FOUR

Figure 4.1 Schematic diagram of the filtration setup

4.2.2.4 Static BSA adsorption test

As indicated in previous studies, the adsorption and deposition of foulant components


are two distinct mechanisms that simultaneously affect the flux decline of
microfiltration membranes [21, 22]. The static BSA adsorption experiment examines
the effect of surface modification on the BSA adsorptive fouling without the
interference of filtration-induced deposition. In general this experiment is performed by
simply submerging the membrane material into a protein solution. In this study, the
hydraulic permeability of the membrane sample was first measured in the filtration cell.
Then the membrane was removed from the filtration cell and immersed in the BSA
solution for four hours. Directly after the adsorption period, the membrane was returned
to the filtration cell and the hydraulic permeability of the membrane was measured
again. Two solutions of 0.04 g/L BSA and 0.2 g/L BSA were applied for this test. The
pH of both solutions was adjusted to seven by adding 0.1 M NaOH solution. Each test
was repeated twice and the average results were reported. The differences in the

74
CHAPTER FOUR

permeability of the clean and pre-adsorbed membranes were investigated and the extent
of the flux decline associated with protein adsorption was evaluated.

To compare the fouling tendencies of the modified and unmodified membranes in a


dynamic filtration condition, microfiltration of BSA solution was performed in the dead
end membrane filtration set-up. The transmembrane pressure was set at 20 kPa for the
virgin membranes. To test the modified membranes, the pressure was adjusted to the
same initial flux as that of the unmodified membranes. This was done in order to have a
similar fouling load on the membrane in both experiments. In order to make sure that
the results are producible each test was repeated twice.

4.3 Results and discussion


4.3.1 Characterization of the membranes

ATR-FTIR was used to examine the presence of the functional groups present on the
membrane surface. Figure 4.2 shows the FTIR spectra for the coated and original
membrane samples in the range of 1,000 cm−1 to 1,900 cm-1. The original CA
membrane exhibited several peaks in this region. The most intense peaks in this region
were located in 1,234 cm−1 and 1,047 cm−1, which were assigned to C-O single bond
stretching mode. The peak at 1,736 cm−1 was specific to CA and was attributed to the
vibration of the carbonyl group [23-24]. The spectra for the eight-, 16- and 24- hour
coated samples were similar to the spectrum of the original membrane, and no change in
the spectrum was detectable. Although L-DOPA deposition causes an eye-visible grey
colour change, the deposited film thickness seemed to be much thinner than the infrared
beam penetration depth; thus, the spectrum of the underlying material was still
dominant.

To make the coating detectable by FTIR, the membranes were coated in a three-step
process. Each step took 16 hours and was carried out in a freshly prepared DOPA
solution. The multi-step deposition thickened the film to make it detectable by IR
method. For the multi-step coated membrane, a moderately intense band at 1,550 cm−1
appeared. This was attributed to the asymmetric stretching vibration of the anionic
carboxyl group -CO2−. The band for un-ionised carboxyl should be in the region
between 1,800 cm−1 to 1,750 cm−1. Therefore, the position of this band in the spectrum
of the multi-step coated sample confirmed that the carboxyl group of the deposited poly

75
CHAPTER FOUR

L-DOPA was ionised. This band could also be attributed to the bending vibration of
ammonium cations [25, 26]. Furthermore the peak at 1,640 cm−1corresponds to the
stretching vibration of C=C of phenylene rings.

three- step coated CA


membrane

Original CA membrane

Absorbance
1800 1700 1600 1500 1400 1300 1200 1100 1000 900
Wave number (cm-1)

Figure 4.2 ATR-FTIR spectra of the original and three- step coated membranes

The contact angle results obtained by captive bubble method are shown in Table 4.1 in.
The results showed a decrease in water contact angle in the eight-hour coated membrane
compared to the virgin membrane. Contrary to the theoretical reasoning, the contact
angle for the 16- and 24- hour coated membranes were slightly higher than for the
eight-hour coated samples. Since the microfiltration membranes surfaces are rough and
porous these observations in the apparent contact angle can be explained by the change
in membrane surface roughness.

76
CHAPTER FOUR

Table 4.1 Water permeability, contact angle and surface charge of the original and coated membranes

Membrane type Water flux Contact angle zeta potential Membrane porosity (%)

in 20 kPa
pH 8 pH 6
(L/m2h)

Original CA 920 36±1.5 -45±1 -35±1 67.8 ±0.2

8 hours coated CA 980 29±1 -49±4 -38±3 67.5± 0.3

16 hours coated CA 960 32±1 -50±3 -42±2 67.2±0.1

24 hours coated CA 870 34±2 -53±2 -44±1 66.3±0.3

The results of zeta potential of the membranes surface estimated by the streaming
potential measurement are indicated in Table 4.1. It has been generally observed that the
zeta potential of the membranes correlates well with its protein fouling [27, 28]. The
results for the membrane surface charge in the tested pH points showed that the
modified membranes had higher negative values than the original membrane (see
Table 4.1). This could result in a higher electrostatic repulsion between the negatively
charged BSA and the membrane surface. Dissociation of the carboxyl functional group
of poly L-DOPA caused development of a higher negative charge on the coated
membranes. The virgin CA membrane contained acetyl (CH3CO-) and hydroxyl (OH-),
which cannot dissociate in the tested conditions. Their negative zeta potential can be
attributed to the adsorption of anions from the solution [20].

Table 4.1 also shows the water flux of the original CA membrane and coated
membranes. The results demonstrated an increase in the average of water flux measured
for the eight and 16 hours coated samples. An improvement in the membrane
wettability (without blocking the pores) could result in an increase in the water
permeability. However, for the 24 hours coated sample, a slight decrease in water flux

77
CHAPTER FOUR

was observed. This can be attributed to the increased membrane flow resistance due to a
slight decrease in membrane porosity.

4.3.2 BSA adsorption and fouling behaviour of the membranes

The effect of static adsorption of BSA on the permeability of the membranes is


displayed in the form of the percentage of flux loss in Figure 4.3. The results
demonstrate a lower flux loss for the 24 hour coated membranes than for the
unmodified membranes. This indicates less protein adsorption onto the 24 hour coated
membranes. These results also confirm that BSA adsorption rate depends on BSA
concentration.

35%

30%
L-DOPA coated
25% membrane

uncoated membrane
flux loss ( %)

20%

15%

10%

5%

0%
0.04 g/L BSA 0.2 g/L BSA

Figure 4.3 The percentage of flux loss evaluated by measuring the flux of each membrane, at 20 kPa,
before and after static BSA adsorption

Figure 4.4 and Figure 4.6 display the plots of resistance versus time and permeate
volumes during the filtration of 0.2 g/L BSA. The membrane resistance can be
calculated from the membrane flux data. Since the permeate flux through the membrane
is proportional to the transmembrane pressure (kPa) and is inversley proportional to
membrane resistance (1/m) and the viscosity of the fluid (N.s/m2) thus having the
membrane flux, the transmembrane pressure and the viscosity of the fluid, membrane
resistance can be easily estimated. It can be observed that fouling occured quickly for
the uncoated membrane and its resistance sharply rose to 14 m-1 after 100 minutes of

78
CHAPTER FOUR

filtration.Where, the resistance of the coated membrane only showed a slight increase
with time and reached about 1.6 m-1 after 100 minutes of filtration. Both resistance
versus time and resistance versus permeate volume plots proved the superior
performance of the coated membrane.

16

14
original membrane
12
Resistance (1/m)

10 24-hour coated
membrane
8

0
0 20 40 60 80 100
Time (min)

Figure 4.4 The plot of membrane resistance versus time during the filtration of 0.2 g/L of BSA solution

5 24-hour coated
membrane
4 original membrane
Resistance (1/m)

0
0 2000 4000 6000 8000 10000
permeate volume(L/h.m2)

Figure 4.5 The plot of membrane resistance versus permeate volume during the filtration of 0.2 g/L of
BSA solution

79
CHAPTER FOUR

Compared to test results from filtration of 0.2 g/L BSA, Lower resistance increase rate
is observed for samples tested with 0.04 g/L BSA feed solution (Figure 4.6 and
Figure 4.6). It can be seen that in the first minutes of the filtration, almost the same
amount of resistance is added to both coated and uncoated membranes. However, in the
later stage of fouling, the resistance of the uncoated membrane increases with a higher
rate compared to that of the uncoated membrane.

2.2

2 24-hour coated
membrane
1.8
Resistance (1/m)

original membrane
1.6

1.4

1.2

0.8
0 50 100 150 200 250

Time (min)

Figure 4.6 The plot of membrane resistance versus time during the filtration of 0.2 g/L of 0.04 g/L BSA
solution

2
24-hour coated
1.8 membrane
original membrane
1.6
Resistance (1/m)

1.4

1.2

0.8
0 5000 10000 15000 20000 25000 30000 35000

permeate volume (L/h.m2)

Figure 4.7The plot of membrane resistance versus permeate volume during the filtration of 0.2 g/L of
BSA solution

80
CHAPTER FOUR

The filtration results of the less concentrated BSA solution (0.04 g/L) seem to indicate
that in early stage of fouling, membrane fouling depends more strongly on the
membrane structure than on the membrane surface properties.

4.4 Conclusions
Cellulose acetate microfiltration membranes were modified by the chemical deposition
of amino acid L-DOPA onto the membranes. The ATR-FTIR results indicated that the
deposited polymer was in the zwitterionic form. Contact angle results associated with
increased water permeability of the 8- and 16- hour modified membranes, confirmed a
small increase in the wettability of the modified membranes. The zeta potential results
revealed a higher negative surface charge of the coated membranes. This can cause an
additional electrostatic repulsion of BSA molecules by the modified membrane.
Through the static BSA adsorption experiments it was found that the modification has
reduced the BSA adsorption. Further investigation employing dynamic BSA filtration
demonstrated that the fouling resistance of the modified membranes was improved.
However, it was observed that early stage of fouling was more governed by the
membrane structure rather than the membrane surface properties.

81
CHAPTER FOUR

References
[1] Water treatment membrane processes, McGraw-Hill, New York, 1996.

[2] P.E. Hillis, Membrane Technology in Water and Wastewater Treatment, Royal Society of
Chemistry, TheIngram Publisher Services, Cambridge , 2000.

[3] Z.F.A. Cui, H.S.A. Muralidhara, Membrane Technology: A Practical Guide to Membrane
Technology and Applications in Food and Bioprocessing, Butterworth-Heinemann Elsevier
Science & Technology Books, San Diego , 2010.

[4] G. Belfort, R.H. Davis, A.L. Zydney, The behavior of suspensions and macromolecular
solutions in crossflow microfiltration, Journal of Membrane Science, 96 (1994) 1-58.

[5] G.N.B. Baroña, B.J. Cha, B. Jung, Negatively charged poly(vinylidene fluoride)
microfiltration membranes by sulfonation, Journal of Membrane Science, 290 (2007) 46-54.

[6] H. Guo, M. Ulbricht, Surface modification of polypropylene microfiltration membrane via


entrapment of an amphiphilic alkyl oligoethyleneglycolether, Journal of Membrane Science,
349 (2010) 312-320.

[7] M.-G. Yan, L.-Q. Liu, Z.-Q. Tang, L. Huang, W. Li, J. Zhou, J.-S. Gu, X.-W. Wei, H.-Y.
Yu, Plasma surface modification of polypropylene microfiltration membranes and fouling by
BSA dispersion, Chemical Engineering Journal, 145 (2008) 218-224.

[8] S. Akhtar, C. Hawes, L. Dudley, I. Reed, P. Stratford, Coatings reduce the fouling of
microfiltration membranes, Journal of Membrane Science, 107 (1995) 209-218.

[9] Z.-M. Liu, Z.-K. Xu, L.-S. Wan, J. Wu, M. Ulbricht, Surface modification of polypropylene
microfiltration membranes by the immobilization of poly(N-vinyl-2-pyrrolidone): a facile
plasma approach, Journal of Membrane Science, 249 (2005) 21-31.

[10] Y. Wang, J.-H. Kim, K.-H. Choo, Y.-S. Lee, C.-H. Lee, Hydrophilic modification of
polypropylene microfiltration membranes by ozone-induced graft polymerization, Journal of
Membrane Science, 169 (2000) 269-276.

[11] J. Mueller, R.H. Davis, Protein fouling of surface-modified polymeric microfiltration


membranes, Journal of Membrane Science, 116 (1996) 47-60.

[12] S. Azari, L. Zou, Using zwitterionic amino acid l-DOPA to modify the surface of thin film
composite polyamide reverse osmosis membranes to increase their fouling resistance, Journal of
Membrane Science, 401–402 (2012) 68-75.

[13] M. Yu, J. Hwang, T.J. Deming, Role of l-3,4-Dihydroxyphenylalanine in Mussel Adhesive


Proteins, Journal of the American Chemical Society, 121 (1999) 5825-5826.

82
CHAPTER FOUR

[14] H. Lee, S.M. Dellatore, W.M. Miller, P.B. Messersmith, Mussel-inspired surface chemistry
for multifunctional coatings, Science, 318 (2007) 426-430.

[15] L.-P. Zhu, J.-Z. Yu, Y.-Y. Xu, Z.-Y. Xi, B.-K. Zhu, Surface modification of PVDF porous
membranes via poly(DOPA) coating and heparin immobilization, Colloids and Surfaces B:
Biointerfaces, 69 (2009) 152-155.

[16] Z.-Y. Xi, Y.-Y. Xu, L.-P. Zhu, Y. Wang, B.-K. Zhu, A facile method of surface
modification for hydrophobic polymer membranes based on the adhesive behavior of
poly(DOPA) and poly(dopamine), Journal of Membrane Science, 327 (2009) 244-253.

[17] Y. Liao, Y. Wang, X. Feng, W. Wang, F. Xu, L. Zhang, Antibacterial surfaces through
dopamine functionalization and silver nanoparticle immobilization, Materials Chemistry and
Physics, 121 (2010) 534-540.

[18] J.M. Hong, B.J. Kim, J.-H. Shim, K.S. Kang, K.-J. Kim, J.W. Rhie, H.J. Cha, D.-W. Cho,
Enhancement of bone regeneration through facile surface functionalization of solid freeform
fabrication-based three-dimensional scaffolds using mussel adhesive proteins, Acta
Biomaterialia, 8 (2012) 2578-2586.

[1 ] A. M , F. M z, .I. C l o, P. P o , L. P l cio, A. H z, P o i
adsorption onto an inorganic microfiltration membrane: Solute–solid interactions and surface
coverage, Journal of Membrane Science, 207 (2002) 199-207.

[20] M. Elimelech, W.H. Chen, J.J. Waypa, Measuring the zeta (electrokinetic) potential of
reverse osmosis membranes by a streaming potential analyzer, Desalination, 95 (1994) 269-286.

[21] H. Ma, C.N. Bowman, R.H. Davis, Membrane fouling reduction by backpulsing and
surface modification, Journal of Membrane Science, 173 (2000) 191-200.

[22] A.L.Z. L. J. Zeman, Microfiltration and Ultrafiltration - Principles and Applications, 69


(1997) 1479-1479.

[23] S. Rauf, A. Ihsan, K. Akhtar, M.A. Ghauri, M. Rahman, M.A. Anwar, A.M. Khalid,
Glucose oxidase immobilization on a novel cellulose acetate–polymethylmethacrylate
membrane, J Biotechnol, 121 (2006) 351-360.

[24] J.L. Koenig, Infrared and Raman Spectroscopy of Polymers, Smithers Rapra Publishing,
Shrewsbury, 2001.

[25] A. Barth, The infrared absorption of amino acid side chains, Progress in Biophysics and
Molecular Biology, 74 (2000) 141-173.

83
CHAPTER FOUR

[26] M. Weinhold, S. Soubatch, R. Temirov, M. Rohlfing, B. Jastorff, F.S. Tautz, C. Doose,


Structure and Bonding of the Multifunctional Amino Acid l-DOPA on Au(110), The Journal of
Physical Chemistry B, 110 (2006) 23756-23769.

[27] S. Yeu, J.D. Lunn, H.M. Rangel, D.F. Shantz, The effect of surface modifications on
po i ic ofil io p op i of A opo ™ , o l of M Sci c , 3
(2009) 108-117.

[28] K. Nakamura, K. Matsumoto, Properties of protein adsorption onto pore surface during
microfiltration: Effects of solution environment and membrane hydrophobicity, Journal of
Membrane Science, 280 (2006) 363-374.

84
5 Fouling resistant zwitterionic surface modification of
reverse osmosis membranes using amino acid
L-cysteine

Abstract

Chapter three explained the successful application of zwitterionic aminoacid L-DOPA


in the suppression of membrane fouling. This chapter describes the application of
zwittterionic amino acid L-cysteine on to the commercial thin-film composite (TFC)
polyamide reverse osmosis membranes in order to improve membrane fouling
resistance. The modification was carried out through bonding the thiol group of the
L-cysteine to the allyl functionalized membrane. Attenuated total reflection
spectroscopy (ATR-FT-IR) and X-ray photoelectron spectroscopy (XPS) tests
confirmed the success of grafting reaction. Contact angle measurement and atomic force
microscopy (AFM) images indicated higher hydrophilicity and smoother surface of the
modified membrane. Water permeability and salt rejection of the membranes were
measured using a stirred batch cell. Although membrane modification tended to enhance
membrane salt rejection, membrane permeability declined. Cross-flow filtration of
bovine serum albumin (BSA) solution and dodecyltrimethylammonium bromide
(DTAB) as model foulants displayed a lower fouling propensity of the modified
membrane. Improved fouling resistance of the modified membrane was attributed to its
enhanced hydrophilicity and lower surface roughness.

This chapter is based on a published article (Desalination 2013)


85
CHAPTER FIVE

5.1 Introduction
Membrane fouling represents a major challenge in efficient and large-scale application
of polymeric membranes in water purification systems. For instance, in a membrane
desalination plant, fouling of the reverse osmosis membranes increases the energy
consumption and reduces the salt rejection. It also negatively affects the plant
productivity by requiring frequent membrane cleanings and premature membrane
replacement. Therefore, inhibiting the membrane fouling becomes a key issue in a plant
of this type [1].

Initial stage of membrane fouling is affected by the interaction between the foulants and
the membrane surface [2, 3]. Prevention of non-specific attachment of biomolecular and
microorganisms has been an effective approach to inhibit membrane fouling [4]. There
has been extensive research interest in membrane surface modification methods and
materials that effectively prevent the undesirable adhesion of foulants [5, 6].
Nevertheless, as described previously in chapters two and three there are still just very
few materials which can meet the challenges of practical applications. Poly (ethylene
glycol) (PEG) and oligo (ethylene glycol) have been used as the most powerful
antifouling materials [7]. However, undesirable oxidation of PEG in the presence of
oxygen or transition metal ions [8-10] besides loss of protein resistance of PEG at above
35°C [11, 12] makes it necessary to look for an alternative material.

Results presented in chapter 3 showed that deposition of zwitterionic amino acid


3- (3,4-Dihydroxyphenyl)-L-alanine (L-DOPA) on the polyamide RO [13] and cellulose
acetate forward osmosis membranes [14] can effectively mitigate biopolymer fouling of
those membranes. This inspired us to expand our understanding of the effect of
membrane modification with zwitterionic amino acids on the improvement of the
’ fo li g i .

Recently, Rosen and Gu [15] demonstrated an excellent BSA stability for the silica
nanoparticles functionalized by conjugating amino acid L-cysteine to the particle
surface via thiol group. Duan and Lewis [16] covalently immobilized L-cysteine onto
two biomedical polymers, polyurethane and polyethylene terephthalate. They observed
a reduced platelet adhesion onto the modified polymers. These works motivated us to

86
CHAPTER FIVE

employ amino acid L-cysteine as a candidate for fouling resistant surface modification
of polyamide RO membranes.

In this chapter, we report on the synthesis of zwitterionic L-cysteine modified reverse


osmosis membrane and evaluation of its fouling resistance. Modification occurs in a
two-step process, involving the activation of membrane surface with an allyl compound
and followed by selective reaction with L-cysteine thiol group via thiol-ene reaction.
Th ’ fo li g i c l c o -flow filtration conditions
using BSA solution, as the model biopolymer foulant, and the DTAB solution, as the
model cationic surfactant foulant. Surfactants often exist in treated industrial wastewater
because mixtures of anionic and cationic surfactants are being used to remove organic
contaminates from waste water [17]. Figure 5.1 schematically displays the hypothesis of
fouling resistance of modified membranes by hydrated zwitterionic L-cysteine. To the
best of our knowledge, this study represents the first application of L-cysteine for low
fouling surface modification of a polymeric membrane.

Figure 5.1 Schematic of surface adsorption resistance for organic matter imparted by the hydrated
zwitterionic L-cysteine coated membrane surface

87
CHAPTER FIVE

5.2 Materials and methods


5.2.1 Materials

SW30XLE polyamide thin film composite membrane (a low-energy seawater


desalination membrane) manufactured by DOW Water and Process Solutions was used
in this study. Allyl glycidyl ether (AGE), L-cysteine, potassium persulfate (KPS),
Tris(hydroxymethyl)aminomethane (Tris), bovine serum albumin (BSA) and
dodecyltrimethylammonium bromide (DTAB) were all obtained from Sigma-Aldrich
and used as received.

5.2.2 Membrane surface modification

In order to prepare a low-fouling zwitterionic surface from L-cysteine, amino and


carboxylate groups of L-cysteine should be kept intact. Therefore it is necessary to
control the functional groups through which the L-cysteine is attached to the membrane
surface. This selectivity can be achieved by using the thiol as the bonding functional
group. In this study, thiol-ene reaction was employed as an approach to covalently bond
L-cysteine to the membrane surface. Thiol-ene reaction displays high rates with near
quantitative, regioselective yields and high tolerance towards various functional groups
[18]. In addition, the formed thiol ether bonds provide high chemical stability [18].
Allyl glycidyl ether was applied as the activating agent on membrane. Its epoxy end
group is capable of reacting with the free carboxylic acid and primary amine groups on
the polyamide membranes, while the alkene end group can react with thiol group of
L-cysteine.

Amine group is strongly hydrophilic; therefore, it can desirably react with the epoxide.
However, to facilitate the reaction of carboxylic acid group with the epoxide, the
reaction was carried out in the alkaline environment (a negatively charged carboxylate
has greater nucleophilicity than its uncharged acid)[19]. Therefore, the AGE solution
was prepared by dissolving the required amounts of AGE into the water with its pH
value adjusted to 8.5 by titrating 1 M NaOH solution. The membrane activation step
was carried out with soaking the clean membrane in the AGE solution for 15 min at
ambient temperature. The membranes were then removed from the solution and rinsed
with deionized water several times. Using high concentrations of AGE did adversely

88
CHAPTER FIVE

affect membrane salt rejection. Therefore, the optimum concentration of AGE solution
was investigated by measuring salt rejection of the activated membranes.

The general procedure for bonding of L-cysteine to the activated membrane is as


follows; 733 mg L-cysteine and 60 mg of initiator (KPS) were dissolved in 150 ml of
15 mM Tris buffer (pH 8) solution. The reaction mixture with the membrane was
placed in the shaking bath at 60 °C for 40 min. Then the membranes were rinsed and
flushed several times with pure water to remove any residual unreacted L-cysteine.
Finally, the membranes were stored in pure water until use. The schematic of the
surface grafting of L-cysteine onto RO membranes is displayed in Figure 5.2.

Figure 5.2 Schematic of chemical reactions used for membrane modification: firstly activation with
allyl glycidyl ether then reaction with thiol group of zwittterionic L-cysteine

5.2.3 Characterization of modified membranes

5.2.3.1 FTIR

Attenuated total reflectance spectroscopy (ATR-FTIR) was used to analyse the chemical
structure of the modified and original membranes. Perkin Elmer FTIR analyser

89
CHAPTER FIVE

equipped with an ATR accessory (diamond crystal) was applied to collect the spectra.
Each spectrum was taken using 128 scans at a resolution of 4 cm–1 in the region
between 650 and 4000 cm–1.

5.2.3.2 XPS

The surface elemental composition was determined by X-ray photoelectron microscopy


to evaluate the success of surface modification. The XPS analysis was performed with
an Omicron ESCA Probe available at Mawson Institute, University of South Australia.
D o i i g o och o ic AlK i io (hν =1486.6 eV) as the X-ray
source. The XPS data were analysed using the Casa XPS software (version
2.3.12Dev9). A linear background subtraction was applied to estimate the elemental
composition.

5.2.3.3 AFM

Atomic force microscopy was applied to characterize the surface roughness of the
membranes. The experiments were performed using NT-MDT NTEGRA SPM atomic
force microscope operated in non-contact mode. Silicon nitrile coated with gold on the
reflective side was used as the non-contact tip and had the resonance frequencies
between 65 and 100 kHz with 10 nm amplitude of oscillation. 10 μ ×1 μ i g
were obtained with a scan rate of 0.5 Hz.

5.2.3.4 Contact angle measurement

The measurement of the contact angle of the membranes was performed in ambient
conditions using a OCA15 contact angle analyser (Data- Physics Instruments GmbH,
Filderstadt ,Germany). Contact angle was measured using sessile drop method in which
a water drop of 2 μL ih h of 1 μL/S l o h f c y
approaching the stage (substrate) parallel to the needle direction. A side-view picture
was captured after 10 seconds. The contact angle on both sides of the water droplet was
measured. Each measurement was made at least ten times and the average is reported.

5.2.3.5 Zeta potential

The surface charge of the membranes was measured using an Anton Paar SurPASS
Electrokinetic Analyser (Anton Paar GmbH, Austria). The adjustable gap cell apparatus
was applied to hold the membrane. The zeta potential data were computed from

90
CHAPTER FIVE

streaming potential slope versus pressure plots based on the Helmholtz–Smoluchowski


approach. 10 mM KCl was used as the background solution.

5.2.4 Measurement of water permeability and salt rejection

Water permeability and salt rejection of the membranes were evaluated in a stirred
Batch cell (Hp4750, Sterlitech Corp., WA, USA). The cell has an active filtration area
of 14.6 cm2 and working volume of 200 ml. All experiments were performed under
420 rpm stirring speed and 50 bar transmembrane pressure difference which is
controlled by a high pressure gas regulator connected to a high pressure nitrogen vessel.
D co i g o c l fl x o . Th ’ l
rejection was measured using a feed solution of 2.5 g NaCl. A conductivity metre
(Hach) was used to measure the total dissolved salt (TDS) in the feed and permeate
solutions. When 50 g solution was collected in the permeate vessel, the pressure was
released and the TDS values of the permeate, feed solution left in the cell and original
feed solution were measured. The average l of o igi l f ol io ’ TDS and
that of the feed left in the cell was considered as the feed TDS. The salt rejection can be
calculated as:

( )

For each type of membrane, three samples were tested and the averages and standard
deviations of flux and salt rejection were reported.

5.2.5 Evaluation of antifouling efficiency of modified membranes

The fouling experiments were performed in a laboratory scale cross-flow test unit. The
unit includes a Sepa CF II high-pressure membrane module (GE Osmonics,
Minnetonka, MN) with an effective membrane surface area of 140 cm2, a high-pressure
pump (Hydracell), a feed tank, a heat exchanger to control temperature, and a digital
balance and PC to collect data. The details of the unit and the procedure for the data
acquisition have been described earlier in chapter 3. Before starting the fouling test,
membranes were compacted for 12 hours by pure water only. Following 12 hours
filtration of the pure water, the foulant solution was poured into to the tank and the
membrane flux was recorded for the next 12 hours.

91
CHAPTER FIVE

5.3 Results and discussion

5.3.1 Surface characterization

5.3.1.1 ATR-FTIR

ATR-FTIR was used to examine the presence of functional groups present in the top
layer of the membranes. Figure 5.3 (a and b) contains FTIR spectra for virgin, AGE
activated and L-cysteine grafted membranes. In Figure 5.3 (a), both polyamide and
polysulfone layers of the tested membranes can be identified due to the penetration
depth higher than 300 nm in this wave number range (1000–1800 cm−1) [20]. For
example, the peaks at 1241 cm−1 and 1150 cm−1are attributed to the C-O-C stretch and
asymmetric O=S=O stretch of the PS support, respectively, whereas, the peaks at
1609 cm−1and 1537 cm−1 are assigned to N-H stretching and C-N stretching of amide
group, respectively. The peaks at 1585, 1488, 1105 cm−1, are correspondent to C-C
stretch of the aromatic rings of both PS and polyamide [5, 21]. Due to the fact that in
wave number range lower than 2500 cm–1, the IR beam has a penetration depth around
500 nm, the spectrum of underling materials has the major effect on the whole graph.
Therefore, no noticeable change is observed in the spectrum of the modified membrane.
At the wave numbers greater than 2500 cm–1 (Figure 5.3 b) ATR-FTIR has shallower
depth of penetration so it is much more surface sensitive with a penetration depth
around 200 nm. Thus AGE activated membrane shows a peak at around 2950 cm–1,
which can be assigned to stretching of aliphatic C–H bonds, consistent with an AGE
material abundance in aliphatic carbons. No more change in the spectrum is observed
after L-cysteine was grafted to the surface. It may be due to the fact that L-cysteine has
the similar functional groups as the membrane itself has. Moreover, no indication of
damage is observed in the FTIR spectrum of the modified membrane which may
suggest that modification process conditions do not have any adverse effect on the
membrane chemical structure. This can be further supported by membrane performance
test results.

92
CHAPTER FIVE

Figure 5.3 ATR-FTIR spectra of the virgin RO, AGE activated and L-cysteine modified membranes a)
–1 –1
wave numbers between 1000 −1800 cm −3700 cm

5.3.1.2 XPS analysis

Relative atomic concentrations of the elements present in the surface of the membranes
are indicated in Table 5.1. The results are in accordance with the data reported in the
literature for the low energy RO membranes which have a higher percentage of oxygen

93
CHAPTER FIVE

compared to the other types of polyamide RO membranes [20, 22]. Upon AGE
activation, an increase in oxygen content and a decrease in nitrogen content are
observed. The increase in the sulphur concentration for the L-cysteine grafted sample
confirmed the L-cysteine attachment to the membrane surface.

Table 5.1 Elemental composition of membranes

Membrane C 1s O 1s N 1s S 2p

unmodified 77.4 15.7 6.9 ------

Allyl glycidyl activated 76.5 16.6 5.6 ------

L-cysteine modified membrane 76.4 14.9 7.8 0.9

XPS spectra of C1s were deconvoluted to investigate the corresponding binding


energies of carbon atoms in the polyamide structure. Deconvolution of the C1s forms a
major peak centred at 287 eV B.E. and two minor peaks centred at 289 and 290.7 eV
B.E. [23]. The major peak is assigned to the C-C and C-H of aliphatic and aromatic
environments and the minor peaks correspond to carbons in a relatively higher electron
withdrawing environment (C-O and C-N) [24, 25]. Table 5.2 indicates the relative
abundance of different types of carbon in membrane samples. A growth in the minor
peak at 289 eV B.E. for Allyl glycidyl ether activated sample can be attributed to
the -C-O-C- and -C-O-H- of the allyl glycidyl ether. For L-cysteine modified sample, a
growth in the minor peak at 290.7 eV B.E. is observed which corresponds to carboxylic
acid group of L-cysteine attached. A growth in the peak at 287 eV B.E. can also be
associated with the C-S and amine group of the L-cysteine.

Table 5.2 Relative abundance for different types of carbon in the membrane samples (C1: 287, C2: 289
and C3: 290.7 eV B.E. )

Membrane C1 C2 C3

Virgin membrane 68 17 15

AGE activated membrane 63 23 14

L-Cysteine modified
69 12 19
membrane

94
CHAPTER FIVE

The high resolution spectra of sulphur (Figure 5.4) show two separate peaks for the
L-cysteine modified sample and one peak for the virgin membrane. The sulfone peak at
170.5 eV B.E. is related to polysulfone support layer which has a minor contribution to
the XPS spectra of the tested membranes. The peak at 166.5 eV B.E. represents
the -C-S-C- formed from the reaction of thiol group of the L-cysteine with alkene group.

S2p S2p
1 1
x 10 x 10
60
46
(a) (b)
55 44

42
50
40

45 38

36
40
34
35
32

162 164 166 168 170 172 174 176 162 164 166 168 170 172 174 176
Binding Energy (eV) Binding Energy (eV)
S2p
1
65 x 10

(c)
60

55

50

45

40

162 164 166 168 170 172 174 176


Binding Energy (eV)

Figure 5.4 High-resolution XPS scans for S2p: (a) virgin RO membrane, (b) AGE activated RO
membrane and (c) L-cysteine grafted membrane

5.3.1.3 Surface roughness

Previous works have reported that the physical roughness of the membrane significantly
influences the extent and rate of membrane fouling [26]. Therefore AFM was used to
investigate the effect of surface modification on membrane surface roughness.
Three-dimensional 10 μ c i g of h o ifi o ifi
were taken, and the root mean square roughness (RMS) and peak-to-peak distance of
the membrane surface were calculated and listed at the bottom table in

95
CHAPTER FIVE

Figure 5.5.

a b

Sample RMS roughness (nm) Peak to peak distance(nm)

a. Virgin membrane 113 832

b. L-cysteine modified membrane 90 750

Figure 5.5 Atomic force microscopy images (including morphological statistics) of (a,b) polyamide
membranes

After L-cysteine grafting, the RMS and peak to peak distance of the modified membrane
decreased. AFM results demonstrated that the modification process had made the
membrane surface smoother. The decrease in the membrane surface roughness can be
explained in the light of deposition of L-cysteine in the surface valleys as well as
increase in the packing density of the polymer chains. During the modification process,
high temperature can raise the mobility of the polymer chains in the amorphous regions.
Therefore, reorientation and increase in the packing density of the polymer chains occur
causing a decrease in membrane surface roughness. It can also be hypothesized that
L-cysteine deposition readily occurs in the surface valleys which can reduce the
membrane surface roughness. These two factors resulted in an overall reduced
membrane surface roughness.

96
CHAPTER FIVE

5.3.1.4 Water contact angle

It is proved that fouling resistant characteristics of zwitterionic materials are well


correlated with a tightly bound water layer near the surface [27]. In this context, water
contact angle measurements were conducted to identify the hydrophilicity of the surface
modified membranes. Figure 5.6 displays the measured water contact angle results for
the tested membranes.

Figure 5.6 Measured water contact angle results

The static water contact angle for the water droplet on the virgin membrane was found
to be around 50° which falls in the range reported for other seawater desalination
membranes [28].

Activation of membranes with AGE increased the water contact angle which is owing to
the addition of hydrophobic unsaturated (alkene) group to the membrane surface. It is
observed that after grafting of L-cysteine onto the membrane, the water contact angle
has dropped to ~ 34° which is far smaller than the AGE coated sample and
comparatively smaller than the virgin RO membrane. To examine the effect of
o ific io co i io (i. . 4 i 6 °C o h ’ co c gle, a
sample of virgin membrane was subjected to the same modification condition without
any reactant (control membrane). Compared to the L-cysteine grafted membrane, a very

97
CHAPTER FIVE

small reduction in water contact angle was observed which confirms the negligible
effect of modification condition.

5.3.1.5 Zeta potential

It has been proved that electrostatic interactions between the membrane and foulants
have a significant impact on the adhesion rate of these materials to the membrane [29].
Therefore, when characterizing antifouling property of a membrane, it is essential to
consider the membrane surface charge. Figure 5.7 displays the measured zeta potential
for both virgin RO and the modified membranes. The measured profile for the zeta
potential of the virgin RO membrane shows an amphoteric behaviour which is due to
the residual unreacted carboxylic acid and amine g o p o h poly i ’
surface. The zeta potential values of the virgin RO membranes are 1 mV at pH 3 and
−1 V pH 1 . Th l i g ih p io l po i the
literature [22, 30]. Compared to virgin RO membrane, modified membrane shows a
higher net negative charge at the alkaline pH values and a higher positive charge at the
acidic pH values (i.e. 6 mV at pH 3 − 3.5 mV at pH 9.3). This can be explained by
addition of equal amounts of amine and acid groups of the L-cysteine to the membrane
surface. Compared to the isoelectric point (IEP) of the virgin membrane a small right
shift toward higher pH was observed for the IEP of the modified membrane.

10

0
Zeta potential (mV)

-5

-10

-15

-20 Cystein modified membrane


virgin RO membrane
-25

-30
1 3 5 7 9 11
pH

Figure 5.7 Measured zeta potential results

98
CHAPTER FIVE

5.3.2 Membrane performance

5.3.2.1 Water permeability and salt rejection

Figure 5.8 presents water permeabilities and salt rejections of the tested membranes.
Modification caused a noticeable increase in the salt rejection of membranes. However,
an unfavourable decline in their permeability was observed. These alterations in salt
rejection and water permeability may be attributed to either L-cysteine attachment or
reaction conditions. In this relation, higher salt rejection and lower permeability of the
control membrane prove these changes to be due to the reaction conditions. Some
studies also have reported an increase in the salt rejection of the heat treated polyamide
RO membranes. For instance, Mänttäri et al. [31] proposed that high temperature
permanently reorients the polymer chains to a tighter structure causing an increase in
the salt rejection.

98.7% 98.4%
95.3% 94.7%

0.85 0.83
0.78 0.79

virgin membrane AGE activated control cystein modified

water permeability (L/m².h.bar) salt rejection (%)

Figure 5.8 Water permeability and salt rejections of the membrane samples (control membrane is a
sample of virgin membrane which has been subjected to the same modification conditions as the
modified membrane but without any reactant)

5.3.2.2 Anti-fouling performance

A series of cross-flow filtration experiments were conducted to evaluate the effect of


modification on the fouling resistance of the membranes. Since a rise in fouling rate

99
CHAPTER FIVE

with increasing water flux has been observed previously for RO membranes [32], it is
preferred to compare the performance of the membranes with a similar initial water
flux. To achieve this, the pressure applied for the test of virgin membrane was fixed at
18 bar. This pressure was set at 19.6 bar in the case of modified membrane to obtain
almost the same initial flux as that of the virgin membrane. In such applied pressures, an
initial flux of 21.5+.5 L/h.m2 was obtained for both virgin and modified membranes.

4.5 virgin membrane (DTAB as foulant)


modified membrane (DTAB as foulant)
4.3 modified membrane (BSA as foulant)
Membrane resistance ×10-12 (1/m)

virgin membrane (BSA as foulant)


4.1

3.9

3.7

3.5

3.3
0 100 200 300 400 500 600 700
Time (min)

Figure 5.9 Resistance of the original and modified membranes as a function of time using (100 mg/L of
BSA, 10 mM NaCL, pH 6.4) and (20 mg/L DTAB, 10 mM NaCl, pH 6.2) as the model foulants. Initial flux
2
was 21.5±0.5 L/h.m . Temperature was 22±2°C.

100
CHAPTER FIVE

4.6
virgin membrane (BSAas foulant)
membrane resistance × 10-12 (1/m) 4.4 virgin membrane (DTAB as foulant)
modifed membrane (BSA as foulant)
4.2 modified membrane (DTAB as foulant)

3.8

3.6

3.4

3.2

3
0 50 100 150 200 250 300 350 400 450
permeate volume (L/h.m2.bar)

Figure 5.10 Resistance of the original and modified membranes as a function of permeate volume
using (100 mg/L of BSA, 10 mM NaCL, pH 6.4) and (20 mg/L DTAB, 10 mM NaCl, pH 6.2) as the model
foulants. Initial flux was 21.5±0.5 L/h.m2. Temperature was 22±2°C.

The resistance behaviour of the membranes using solutions of 100 mg/L BSA and
20 mg/L DTAB, as the model foulants, is illustrated in Figure 5.9 and Figure 5.10. It
can be seen that BSA causes a slight resistance increase within the time frame of the
experiment which agrees with the other studies about BSA fouling of the polyamide RO
membranes [32]. Even though DTAB solution is much more diluted than BSA solution,
a quicker resistance increase is observed for it. This higher fouling rate for DTAB
solution can be attributed to the attractive electrostatic interaction force between
negatively charged membrane (at the pH value of DTAB solution i.e. 6.2) and
positively charged DTAB molecule. At the pH value of BSA solution (i.e. 6.4), both
BSA and membranes are negatively charged. As result, there is an electrostatic
repulsion force between membranes and BSA. Moreover, the amount of energy required
to remove bound water from hydrophilic BSA molecule is higher compared to the
energy needed to expulse water from the hydrophobic tail of DTAB molecule. Such
higher energy level, in addition to the aforementioned electrostatic forces, is responsible
for lower adhesion rate of BSA to the membrane surface.

The modified membrane displayed lower fouling propensity in both BSA and DTAB
fouling tests. At the pH values of the fouling solutions, both modified and original

101
CHAPTER FIVE

membrane showed almost equal zeta potentials. It can therefore be deduced that
electrostatic interactions have no role in the improvement of modified membranes
fouling resistance. Higher fouling resistance of modified membrane are indebted to the
enhanced membrane hydrophilicity as well as to the smoother membrane surface.

5.4 Conclusion
In this study, the fouling-resistant surface modification of the RO membranes using the
zwitterionic amino acid L-cysteine was investigated. Surface modification was first
performed through the activation of the membrane with an epoxy compound (allyl
glycidyl ether). L-cysteine was then grafted onto the AGE activated membrane via the
thiol-ene reaction. The surface properties and performance of the membranes were
evaluated. The membrane characterization using FTIR and XPS, supported the
successful incorporation of L-cysteine onto the polyamide reverse osmosis membrane.
Contact angle measurements indicated that the modification resulted in a more
hydrophilic membrane surface. AFM topographical images revealed that the
modification has resulted in a smoother membrane surface. Although modification was
found to enhance salt rejection of the membranes, a slight deterioration in the
permeability of the modified membrane was observed. The increase in the salt retention
by the modified membrane was attributed to the tighter structure of polyamide layer
caused by the modification condition. The fouling behaviour of the membranes was
studied using 100 mg/L BSA and 20 mg/L DTAB solutions. BSA fouling rate was
found to be very slow. Contrary to BSA fouling, a significant fouling rate was observed
in the case of DTAB. The discrepancy observed between the fouling rates of BSA and
DTAB was explained in the ground of hydrophilic–hydrophobic interactions as well as
electrostatic interactions between membranes and foulants. In both BSA and DTAB
tests, modified membranes displayed lower fouling propensity compared to the
unmodified ones. Since at the pH values of foulant solutions, both virgin and modified
membranes displayed almost equal surface charges, it was thus concluded that
electrostatic interactions have no effect on the improved performance of the modified
membrane. The superior fouling resistance of the modified membrane was explained in
the light of its improved hydrophilicity and surface morphology.

102
CHAPTER FIVE

Both L-DOPA coating and L-cysteine grafting could mitigate membrane fouling.
However, due to the higher initial resistance (lower permeability) of L-cystein modified
membrane, it can hardly compete with L-DOPA modified membrane. Therefore, due to
the facile modification process and no adverse effect on membranes permeability,
L-DOPA is believed to be a more promising candidate for antifouling surface
modification of polyamide RO membranes. Table 5.3 summarises similarities and
differences of these two methods.

To verify L-DOPA’ ff c o fo li g i ig io , a thermodynamic model


will be used in the next chapter to calculate the total energy needed for bringing the
model foulant molecules from an infinite distance to the surface of a water immersed
membrane.

Table 5.3 Similarities and differences of two methods used for surface modification of polyamide thin
film composite membranes

SIMILARITIES DIFFRENCES

a) Both L-cystein and L-dopa are natural a) L-DOPA coating is carried out easily
amino acids. in mild alkaline pH and room
temperature whereas L-cystein
b) Both L-cystein grafting and L-DOPA
grafting is a two-step method
coating methods create a zwitterionic
including a heating stage.
layer on the membrane surface.
b) L-DOPA coating does not have any
c) Both methods improve the membrane
adverse effect on the membrane
hydrophilicity.
permeability whereas L-cystein
d) Both methods reduce the membrane grafting reduces the membrane
surface roughness. permeability which can be due to the
tighter structure of the membrane
e) Both methods improve membrane
formed by the heat treatment step.
fouling resistance when tested with
BSA and DTAB as model foulants.

103
CHAPTER FIVE

References

[1] M. Elimelech, W.A. Phillip, The Future of Seawater Desalination: Energy, Technology, and
the Environment, Science, 333 (2011) 712-717.

[2] W. Lee, C.H. Ahn, S. Hong, S. Kim, S. Lee, Y. Baek, J. Yoon, Evaluation of surface
properties of reverse osmosis membranes on the initial biofouling stages under no filtration
condition, Journal of Membrane Science, 351 (2010) 112-122.

[3] M. Pasmore, P. Todd, S. Smith, D. Baker, J. Silverstein, D. Coons, C.N. Bowman, Effects of
ultrafiltration membrane surface properties on Pseudomonas aeruginosa biofilm initiation for
the purpose of reducing biofouling, Journal of Membrane Science, 194 (2001) 15-32.

[4] J. Mansouri, S. Harrisson, V. Chen, Strategies for controlling biofouling in membrane


filtration systems: challenges and opportunities, Journal of Materials Chemistry, 20 (2010)
4567-4586.

[5] S. Belfer, Y. Purinson, R. Fainshtein, Y. Radchenko, O. Kedem, Surface modification of


commercial composite polyamide reverse osmosis membranes, Journal of Membrane Science,
139 (1998) 175-181.

[6] A.C. Sagle, E.M. Van Wagner, H. Ju, B.D. McCloskey, B.D. Freeman, M.M. Sharma, PEG-
coated reverse osmosis membranes: Desalination properties and fouling resistance, Journal of
Membrane Science, 340 (2009) 92-108.

[7] C.A. Finch, Poly(ethylene glycol) chemistry: Biotechnical and biomedical applications.
Edited by J. Milton Harris. Plenum Publishing, New York, 1992.

[8] E. Ostuni, R.G. Chapman, R.E. Holmlin, S. Takayama, G.M. Whitesides, A Survey of
S c −P op y R l io hip of S f c h R i h A o p io of P o i , L g i,
17 (2001) 5605-5620.

[9] S. Krishnan, C.J. Weinman, C.K. Ober, Advances in polymers for anti-biofouling surfaces,
Journal of Materials Chemistry, 18 (2008) 3405-3413.

[10] M. Shen, L. Martinson, M.S. Wagner, D.G. Castner, B.D. Ratner, T.A. Horbett, PEO-like
plasma polymerized tetraglyme surface interactions with leukocytes and proteins: in vitro and in
vivo studies, Journal of Biomaterials Science, Polymer Edition, 13 (2002) 367-390.

[11] D. Leckband, S. Sheth, A. Halperin, Grafted poly(ethylene oxide) brushes as nonfouling


surface coatings, Journal of Biomaterials Science, Polymer Edition, 10 (1999) 1125-1147.

104
CHAPTER FIVE

[12] A. Roosjen, H.C. van der Mei, H.J. Busscher, W. Norde, Microbial Adhesion to
Poly( hyl oxi B h :  I fl c of Poly Ch i L g h T p ure, Langmuir,
20 (2004) 10949-10955.

[13] S. Azari, L. Zou, Using zwitterionic amino acid l-DOPA to modify the surface of thin film
composite polyamide reverse osmosis membranes to increase their fouling resistance, Journal of
Membrane Science, 401–402 (2012) 68-75.

[14] A. Nguyen, S. Azari, L. Zou, Coating zwitterionic amino acid l-DOPA to increase fouling
resistance of forward osmosis membrane, Desalination, 312 (2013) 82-87.

[15] J.E. Rosen, F.X. Gu, Surface Functionalization of Silica Nanoparticles with Cysteine: A
Low-Fouling Zwitterionic Surface, Langmuir, 27 (2011) 10507-10513.

[16] X. Duan, R.S. Lewis, Improved haemocompatibility of cysteine-modified polymers via


endogenous nitric oxide, Biomaterials, 23 (2002) 1197-1203.

[17] P. Weschayanwiwat, O. Kunanupap, J.F. Scamehorn, Benzene removal from waste water
using aqueous surfactant two-phase extraction with cationic and anionic surfactant mixtures,
Chemosphere, 72 (2008) 1043-1048.

[18] A.B. Lowe, Thiol-ene "click" reactions and recent applications in polymer and materials
synthesis, Polymer Chemistry, 1 (2010) 17-36.

[19] G.T. Hermanson, Bioconjugate Technique, 2nd ed. Academic Press, San Diego, USA,
2008.

[20] O. Akin, F. Temelli, Probing the hydrophobicity of commercial reverse osmosis


membranes produced by interfacial polymerization using contact angle, XPS, FTIR, FE-SEM
and AFM, Desalination, 278 (2011) 387-396.

[21] J.L. Koenig, Infrared and Raman Spectroscopy of Polymers, Smithers Rapra Publishing
Shrewsbury, 2001.

[22] C.Y. Tang, Y.-N. Kwon, J.O. Leckie, Probing the nano- and micro-scales of reverse
osmosis membranes--A comprehensive characterization of physiochemical properties of
uncoated and coated membranes by XPS, TEM, ATR-FTIR, and streaming potential
measurements, Journal of Membrane Science, 287 (2007) 146-156.

[23] J. Benavente, M.I. Vázquez, Effect of age and chemical treatments on characteristic
parameters for active and porous sublayers of polymeric composite membranes, Journal of
Colloid and Interface Science, 273 (2004) 547-555.

[24] G. Zschornack, Handbook of X-ray data, Springer, New York, 2006.

105
CHAPTER FIVE

[25] D. Briggs, Surface analysis of polymers by XPS and static SIMS, Cambridge University
Press, Cambridge, U.K. ;New York, 1998.

[26] E.M. Vrijenhoek, S. Hong, M. Elimelech, Influence of membrane surface properties on


initial rate of colloidal fouling of reverse osmosis and nanofiltration membranes, Journal of
Membrane Science, 188 (2001) 115-128.

[27] S. Chen, L. Li, C. Zhao, J. Zheng, Surface hydration: Principles and applications toward
low-fouling/nonfouling biomaterials, Polymer, 51 (2010) 5283-5293.

[28] X. Wei, Z. Wang, Z. Zhang, J. Wang, S. Wang, Surface modification of commercial


aromatic polyamide reverse osmosis membranes by graft polymerization of 3-allyl-5,5-
dimethylhydantoin, Journal of Membrane Science, 351 (2010) 222-233.

[29] Y. Zhou, S. Yu, C. Gao, X. Feng, Surface modification of thin film composite polyamide
membranes by electrostatic self deposition of polycations for improved fouling resistance,
Separation and Purification Technology, 66 (2009) 287-294.

[30] A.E. Childress, M. Elimelech, Effect of solution chemistry on the surface charge of
polymeric reverse osmosis and nanofiltration membranes, Journal of Membrane Science, 119
(1996) 253-268.

[31] M. Mänttäri, A. Pihlajamäki, E. Kaipainen, M. Nyström, Effect of temperature and


membrane pre-treatment by pressure on the filtration properties of nanofiltration membranes,
Desalination, 145 (2002) 81-86.

[32] Q. Li, Z. Xu, I. Pinnau, Fouling of reverse osmosis membranes by biopolymers in


wastewater secondary effluent: Role of membrane surface properties and initial permeate flux,
Journal of Membrane Science, 290 (2007) 173-181.

106
6 Assessing the effect of surface modification of
polyamide RO membrane by L-DOPA on the short
range physiochemical interactions between the
membrane and biopolymers: elucidation of the role of
L-DOPA surface modification on membranes fouling
behaviour

Abstract
Theoretical predictions of interaction energies for membrane–biopolymer foulant pairs
were used to compare the fouling tendencies of a virgin commercial polyamide reverse
osmosis (RO) membranes with a amino acid 3-(3,4-Dihydroxyphenyl)-L-alanine
(L-DOPA) coated RO membrane. Lifshitz-van der Waals (LW) and Lewis acid-base (AB)
surface tension components of the membranes were determined based on contact angle
results using the van Oss approach. From these values the LW and AB components of the
free energy of adhesion between membrane and foulants were calculated. Electrostatic
(EL) double layer interaction energies between the membrane and foulants were also
estimated using the measured surface charge data of the membranes and fouling agents.
Bovine serum albumin (BSA) and alginic acid sodium salt (alginate) were used as model
biopolymers causing membrane fouling. Based on the calculated adhesion free energies,
acid-base interactions were found to have the strongest impact on the adhesion of both
BSA and alginate to either membranes surface. It was found that L-DOPA modification has
significantly lowered acid-base interaction affinity toward the adhesion of both foulants
studied. On the basis of calculated free energies of adhesion results, lower fouling
tendency of the L-DOPA modified membrane was expected. The fouling tests indicated a
lower flux decline rate for the modified membrane and confirmed the results obtained from
theory.

This Chapter is based on a Manuscript prepared to be submitted to the Colloids and Surfaces B:
Biointerfaces
107
CHAPTER SIX

6.1 Introduction
Fouling of reverse osmosis membranes is one of the barriers against their efficient
application to produce fresh water from seawater or other water resources. Among
different types of materials which cause membrane fouling, organic matter and
biopolymers are of the most problematic, because they can pass through the pre-treatment
steps and reach the reverse osmosis (RO) membranes [1, 2]. Adsorbed organic matter and
biopolymers also provide a suitable environment for bacterial growth. Several researchers
described that fouling of the membranes by biopolymers and organic matter is influenced
by operating conditions [3], feed solution chemistry [4] as well as physicochemical
properties of organic matter and membranes [5].

Prevention of adhesion of foulants has been an effective strategy to control membrane


fouling. Therefore, modification of the membrane surface to make it unfavorable for
adhesion has been proposed as a strategy to reduce the membrane fouling rate [6-8]. In
chapter three, surface modification of the RO membrane with L-DOPA was explined [9].
Modified membranes showed a lower flux decline rate compared to the virgin membrane
in an accelerated fouling test using a mixture of BSA and alginate as model biopolymers.
Since modification with L-DOPA can be performed in situ, this approach is of particular
interest practically. The goal of this investigation is to obtain a more in depth insight into
the effect of L-DOPA modification on the interfacial interactions leading to the inception
of fouling.

In this study, the thermodynamic work of adhesion which quantifies the energy available
for adhesion is calculated using the concept developed by Fowkes [10] and later by van
Oss [11]. The calculated energy of adhesion is then related to the propensity of the adhered
biopolymers. This theory has been frequently used to describe colloidal attachment to the
membrane surface [12, 13]. Lee et al. [14] used this theory to study fouling of
ultrafiltration membranes by natural organic matter (NOM). They also used this method to
elucidate the mechanism of organic matter fouling on a commercial RO membrane [15].
Application of this approach for NOM is based on the hypothesis that dissolved natural
organic matter behave like colloidal particles. Botton et al. [16] used this theory to predict
the pharmaceuticals rejection by virgin and biofouled nanofiltration membranes. They
found that pharmaceuticals rejection directly correlates with the free energy of interaction
between membranes studied and pharmaceuticals dissolved in the water phase.

108
CHAPTER SIX

The objectives of the present research are therefore two-fold: (i) to characterize the effect
of L-DOPA modification on the surface properties of the membrane and consequently on
the interactions between membrane and biopolymers, and (ii) to investigate how the
alterations in interaction parameters affect membrane fouling behaviour.

6.2 Theory
6.2.1 Interaction energies

The Extended Derjaguin-Landau-Verwey-Overbeek (XDLVO) theory suggested by van


Oss was applied to calculate the total energy needed to bring organic matter from an
infinite distance to a membrane surface immersed in water. According to the XDLVO
theory, the interfacial energies in an aqueous system can be described as the sum of:

i. Apolar or Lifshitz-van der Waals (LW) energies

ii. Polar or Lewis acid-base (AB) energies

iii. Electrostatic double layer (EL) energies

iv. Brownian movement (BR) energies.

The contribution of Brownian motion energy, resulting from thermal energy of molecules,
is very small compared to the other interaction energies. Therefore the total free energy of
interaction is usually written as:

(1)

where represents the Lifshitz-van der Waals energy, the acid-base interaction
energy and refers to the electrostatic energy. From a thermodynamic point of view,
when is negative, adhesion or attraction occurs whereas repulsion occurs
when is positive.

Application of this approach to estimate the free energy of adhesion requires the
determination of surface tension components and measurement of surface charges of the
membranes and relevant fouling agents.

6.2.2 Surface tension components

Van Oss expressed the total surface tension as the sum of Lifshitz-van der Waals ( )
and Lewis acid-base ( ) components of the surface tension [11].
109
CHAPTER SIX

(2)

The AB component of a material surface energy is given by:

√ (3)

is the electron acceptor parameter and is the electron donor parameter.

The Young equation links the contact angle of a drop of liquid (l) deposited on a flat solid
surface (s) with the surface tension of the liquid and the solid surface. It is expressed as:

(4)

Dupre expressed the work which is required to separate these two surfaces from contact as:

(5)

The Young-Dupre equation is then obtained from equations (4) and (5):

( ) (6)

The complete Young-Dupre equation in terms of the interfacial tension components is


obtained as discussed by van Oss [11]:

( ) (√ √ √ ) (7)

Consequently, the unknown surface tension components of a solid surface can be


determined by performing contact angle measurements using three probe liquids with
known surface tension component parameters.

6.2.3 Determination of free energy of interaction

Using the Young-Dupre equation, the surface tension components of the membranes (m)
and model foulants (f) can be determined. These surface tension parameters are then used
to calculate the free energy of adhesion between membrane and foulants immersed into
water. The LW and AB components of free energy per unit area can be expressed as [11]:

(√ √ ) (√ √ ) (8)

110
CHAPTER SIX

√ (√ √ √ ) √ (√ √ √ ) √ √

(9)

The electrostatic (EL) free energy per unit area between a membrane and a foulant is
furthermore given by [17].

( ) [ ( ) ( )] (10)

where is the dielectric permittivity of the liquid, is the inverse Debye screening
length and is proportional to the square root of the solution ionic strength. and are
surface potentials of the membranes and foulants, respectively. Parameter represents the
separation distance. Usually a value of 0.158 nm is considered for the minimum
equilibrium cut off distance. It may be regarded as the distance between outer electron
shells of interacting molecules [18]. In this study, the measured surface zeta potential was
used as the surface potential.

As discussed in Section 6.2.1, the summation of these free energies ( ) can be used
as a criterion of the tendency of fouling agents to adhere to the membrane surface.

Free energy of adhesion alone cannot provide sufficient information about the membrane
fouling behaviour. Also the calculation of free energy of cohesion which provides
quantitative insight into the stability of foulant molecules in water is required. Free energy
of cohesion predicts the interaction between two surfaces of a single substance when
immersed in a solvent and brought into contact. Free energy of cohesion in water as a
media is expressed as [11]:

(√ √ ) [(√ √ ) (√ √ )] (11)

where s and w stand for solid surface and water, respectively.

6.3 Materials and method


6.3.1 Membranes and foulant

The RO membrane selected for this investigation was a SW30XLE polyamide thin film
composite membrane manufactured by DOW Water and Process Solutions.

111
CHAPTER SIX

3-(3,4-Dihydroxyphenyl)-L-alanine (L-DOPA) from Sigma Aldrich was utilized to modify


the membrane surface. Tris(hydroxymethyl) aminomethane (Tris) was used to prepare the
alkaline buffer solution. Bovine serum albumin (BSA) and sodium alginate were
employed as the representative biopolymer foulants. BSA and sodium alginate were both
supplied by the sigma Aldrich and were used as received.

6.3.2 Membrane modification

Membrane modification was carried out under the same protocol explained in chapter
three. It is hypothesized that modification is carried out via in situ adhesion and
polymerization of L-DOPA onto the membrane. This adhesion to the surfaces has been
explained by several strong interactions involving the unique catechol group of the
L-DOPA side-chain such as hydrogen bonding, Michael-type addition from o-quinones,
π-π o i g [19, 20]. It has been anticipated that oxidation of L-DOPA under elevated
pH may result in covalent bonding (Michael-type addition reactions) with amine-
containing polymers [21]. Nevertheless, the detailed bonding mechanism for its adhesion
remained elusive. Also oxidation of the catechol group of L-DOPA to Dopaquinone, a
highly reactive group, can lead to the formation of poly L-DOPA oligomers. However,
clear evidence for such reactions occurring has been lacking. The probable polymerization
pathway of L-DOPA in alkaline pH during the modification process is illustrated in Figure
6.1.

Figure 6.1 Probable polymerization pathway of L-DOPA during the modification process

112
CHAPTER SIX

6.3.3 Probe liquids

Ultra pure water (polar), glycerol (polar) and diiodomethane (apolar) were selected to
determine the two polar components and apolar component of the membranes surface
tension. Surface tension components of the probe liquids are shown in Table 6.1.

2
Table 6.1 Surface tension components (mJ/m ) of model liquids [11, 21]

Probe liquid

Water 21.8 25.5 25.5 51.0 72.8

Glycerol 34.0 3.9 57.4 30.0 64.0

Diiodomethane 50.8 0.0 0.0 0.0 50.8

Ultra pure water was obtained from a Millipore water purification system. Glycerol and
diiodomethane were supplied by Sigma Aldrich and Merck, respectively and used as
received.

6.3.4 Membrane characterization

6.3.4.1 Surface energy determination

The surface energy parameters of the membranes were evaluated employing the measured
contact angles of the aforementioned probe liquids on the membrane surface.
Measurements were conducted with a OCA15 contact angle analyser (Data-Physics
Instruments GmbH, Filderstadt, Germany). Contact angles were measured using the sessile
drop method in which a probe liquid drop of 2 μL l o h y
surface by approaching the stage (substrate) parallel to the needle direction. A side-view
picture was captured after 10 s and the left and right side contact angles of the droplet were
measured. For each probe liquid an average of at least 10 repeats on two membrane strips
was reported.

6.3.4.2 Surface charge analysis

The surface charge of the membranes was measured using an Anton Paar SurPASS
Electrokinetic Analyzer (Anton Paar USA, Ashland, VA). The measurement protocol has
been explained in chapter three.

113
CHAPTER SIX

6.3.5 Foulants characterization

6.3.5.1 Surface energy determination

BSA and alginate were separately deposited onto the RO membrane surface using a stirred
batch cell (Hp4750, Sterlitech Corp., WA, USA). The membrane was then removed from
the cell to allow the deposited foulants (BSA or alginate) to dry in ambient conditions for 8
hour. The sessile drop method, as previously explained, was employed to measure the
contact angles of the probe liquids on the deposited foulants.

6.3.5.2 Measurement of the particle size and the surface charge of the foulants

Zeta potentials and the mean hydrodynamic diameter of BSA and alginate were measured
using Laser Doppler Velocimetry (LDV) and Dynamic Light Scattering techniques,
respectively with a Malvern ZetaSizer Nano ZS (Malvern Instruments Limited, UK).
Solutions of BSA and alginate were freshly prepared with the same ionic strength and
concentrations as the solutions used for fouling experiments. Each solution was tested
twice and each time the measurement was repeated three times.

6.3.6 Membrane fouling experiments

A laboratory scale cross-flow unit was used to carry out the fouling experiments. The
details of the filtration test unit and the method used for testing the membranes have been
explained in chapter three.

6.4 Results and discussion


6.4.1 Surface zeta potential

The measured zeta potential and particle size of BSA are displayed in Table 6.2. BSA is a
globular macromolecule; its hydrodynamic size varies with the solution pH and ionic
strength. Similar to other proteins BSA has an amphoteric charge property with an
isoelectric point value at pH 4.7. It is positively charged at a pH below its isoelectric points
(pHIEP), and it becomes negatively charged at pH higher than its pH IEP [22]. At the pH
studied in this work, a moderate negative zeta potential and a 21 nm hydrodynamic radius
was measured for BSA. The measured zeta potential and mean hydrodynamic radius of
alginate are also presented in Table 2. A large hydrodynamic radius and a high negative
zeta potential was measured for alginate which are consistent with its extended coil
conformation and polyanionic structure, respectively [23].

114
CHAPTER SIX

Table 6.2 Measured zeta potential and particle size of the BSA and alginate

Zeta potential (mv) Mean hydrodynamic radius(nm)

BSA (100 mg/L, pH 6.4) -31 ± 3 21 ± 6 nm

Alginate (100 mg/L, pH 6.1) -58 ± 4 410 ± 25 nm

The zeta potential data of membranes are displayed in Figure 6.2. As a result of presence
of the residual unreacted carboxylic acid and amine groups, the unmodified polyamide
membranes showed amphoteric property. For water purification applications, polyamide
RO membranes are normally operated at pH values between 6 to 8. In this pH range, a
negative zeta potential is observed for XLE membranes. The data for virgin XLE
membrane are in the range of values reported for polyamide membranes in the literature
[24, 25]. The L-DOPA coated membrane exhibited almost identical charge characteristics
to the virgin membrane [9]. Thus it may be speculated that the coating layer is too thin to
cover the underlying layer charge, or the coated material has similar zeta potential to the
original membrane.

20
15
10
virgin membrane
5
Zeta potential (mV)

0 24 hour modified membrane


-5
-10
-15
-20
-25
-30
2 3 4 5 6 7 8 9 10
pH

Figure 6.2 Measured zeta potential of the membranes [9]

115
CHAPTER SIX

6.4.2 Surface energy determination

The measured contact angles of the membranes surfaces as well as their calculated surface
tension parameters are listed in Table 6.3. It can be observed that contact angles of
L-DOPA coated membrane are lower than those of virgin (unmodified) membrane in the
case of all three probe liquids, particularly for water and diiodomethane.

The calculated surface tension parameters are also presented in Table 6.3. Virgin RO
membrane exhibits monopolar functionality with relatively high electron donor (-) and
zero electron acceptor (+) parameters which can be attributed to the significant presence of
electron donor sites on the membrane. These results agree with previous studies [12], that
reported low electron acceptor and high electron donor components of polyamide RO and
NF membranes. After modification, both Lifshitz-van der Waals and electron donor
components of the surface increased; in particular, the electron donor component was
changed dramatically by L-DOPA modification.

Table 6.3 Average values of the measured contact angle (in degrees) of water, glycerol and
2
diiodomethane as well as calculated surface tension parameters (in units of mJ/m ) of membranes and
model foulants

∆Gsws

Membrane

Virgin membrane 49.5 ± 4.5 68 ± 3 37±4 41.1 0.0 20.0 11.41

L-DOPA modified 19 ± 1 50 ± 6 19 ± 3 48.0 0.0 46.0 45.19


membrane

Model foulants

BSA 35 ± 1.5 61 ± 2 49 ± 4 34.9 0.0 42.5 26.56


a a a
26.71 0.00 39.51

Alginate 73 ± 2 60 ± 3 55 ± 2 31.4 1.2 13.0 -24.58


a a a
27.44 0.19 19.22
a
values from the literature [15]

Table 6.3 also indicates the measured contact angles and the calculated surface tension
parameters of deposited foulants. A relatively low electron acceptor and a medium electron
donor surface tension component were obtained for alginate. BSA was characterized by
116
CHAPTER SIX

relatively high electron donor and zero electron acceptor components. The data obtained
for alginate and BSA in this study, agreed with the literature [15].

Free energy of cohesion per unit area for each of the membranes and foulants studied are
also shown in Table 6.3. As was discussed, cohesive energy provides information
regarding the hydrophilicity/hydrophobicity of membranes and foulant molecules. A
surface is considered to be hydrophilic when the cohesive energy is positive, while
hydrophobic surfaces are known by negative values of cohesive energy. In this regard,
based on the values given in Table 6.3, the virgin RO membrane with a positive value of
∆Gsws is by implication hydrophilic; the modified membrane with the ∆Gsws value higher
than the cohesive free energy of the virgin XLE membrane is considered strongly
hydrophilic. BSA with the ∆Gsws value of 26 mJ/m2 is also hydrophilic, whereas alginate
with a negative cohesive energy value is known as a hydrophobic material. This result
means that in water BSA is more stable than alginate.

6.4.3 Free energy of adhesion at contact

Table 6.4 shows the calculated free energy of adhesion per unit area for the various
membrane–foulant pairs tested. For both membrane–BSA pairs, the LW free energy of
adhesion is slightly negative indicating attractive LW force, whereas the AB free energy of
adhesion is positive indicating a repulsive AB force between membranes and BSA.
Particularly, the L-DOPA modified membrane exhibited a strong repulsive acid-base free
energy which originates from the strong affinity of membranes to water and high of the
modified membrane surface. Minor and attractive electrostatic double layer free energies
were obtained for both membrane–BSA pairs.

2
Table 6.4 Calculated interaction energies (mJ/m )

Virgin XLE-BSA -4.32 9.09 -0.07 4.7

Modified XLE-BSA -5.67 32.33 -0.07 26.59

Virgin XLE-alginate -3.27 -19.11 -3.08 -25.46

Modified XLE-alginate -4.27 -0.98 -3.08 -8.33

117
CHAPTER SIX

In the case of hydrophobic alginate-membrane pairs, the calculated LW interactions were


relatively similar to those obtained for BSA-membrane pairs. AB components of free
energy of adhesion between membranes and alginate were found to be attractive for virgin
membrane and almost zero for the modified one. EL interaction was found to be attractive
and in the range of LW interactions for membrane-alginate pairs.

Since BSA, alginate and membranes are all negatively charged at the studied pH values
(pH values of feed solutions i.e. 6.4 for BSA and 6.2 for alginate solution), repulsive
electrostatic double layer interactions between membranes and model foulants are
expected. However, attractive EL interactions were obtained. In this connection, Usui and
Hachisu [17] xpl i h clo pp o ch h ψ1 ψ2 are of the same sign and
different magnitudes, attraction occurs because the sign of the surface charge of the plane
with lower surface potential is reversed when the two plates come close together.

It should be noted that in this study the results for EL double layer interactions were
obtained considering ionic strength of 10 mM NaCl. In very high ionic strength conditions
such as seawater desalination (ionic strength 5 M), EL interaction will be significantly
diminished. Therefore, in desalination of seawater or brackish water, fouling will be
predominantly controlled by AB interactions.

6.4.4 Fouling test

Accelerated fouling tests were conducted to validate the results obtained by theory.
Figure 6.3 and Figure 6.4 demonstrate membrane flow resistance profile during the
filtration of BSA and alginate solutions.

118
CHAPTER SIX

4.4

virgin membrane (alginate)


4.2 modified membrane (alginate)
membrane resistance × 10-12 (1/m)

virgin membrane (BSA)


4
modified membrane (BSA)

3.8

3.6

3.4

3.2
0 100 200 300 400 500 600 700
Time (min)

Figure 6.3 Resistance of the original and modified membranes as a function of time using (100 mg/L of
BSA, 10 mM NaCL, pH 6.4 ) and (100 mg/L alginate, 10mM NaCl, ) as model biopolymer, Temperature: 20
± 2°C, initial flux 21.2± 0.4 L/m2.h

119
CHAPTER SIX

4.4
virgin membrane (alginate)

4.2
membrane resistance × 10-12 (1/m)

modified membrane (alginate)

virgin membrane (BSA)


4

modified membrane (BSA)


3.8

3.6

3.4

3.2
0 50 100 150 200 250 300 350 400 450
permeate volume

Figure 6.4 Resistance of the original and modified membranes as a function of time using (100 mg/L of
BSA, 10 mM NaCL, pH 6.4 ) and (100 mg/L alginate, 10mM NaCl, ) as model biopolymer, Temperature:
2
20 ± 2°C, initial flux 21.2± 0.4 L/m .h

The BSA fouling diagrams represent an initial stage of slow rate of resistance increase
followed by a moderate resistance increase rate. The modified membrane shows a slower
resistance increase rate than the virgin membrane, however, at a later stage, the resistance
increase rate is almost similar for both membranes tested. This observed trend can be
explained on the grounds of foulant–membrane and foulant–foulant interactions. Where
foulant–membrane interaction forces (adhesion forces) affect the initial stage of fouling,
long term fouling behaviour is controlled by foulant–foulant interactions (cohesive forces)
[26, 27]. Thereby, slow resistance increase rates observed in the initial fouling stage,
particularly for the modified membrane, are consistent with the calculated repulsive free
energies of adhesion i.e. 4.7 mJ/m2 and 26.59 mJ/m2 for virgin membrane–BSA and
modified membranes–BSA pairs, respectively. With a positive cohesive energy of
26.56 mJ/m2, BSA molecules appear to be thermodynamically stable in water. This high
positive value of cohesive energy causes minimal deposition of additional BSA on the
membrane surface after the initially deposited BSA covers the surface. Based on the above

120
CHAPTER SIX

discussion, later stage of fouling which is governed by BSA cohesive forces is expected to
be slower than the initial stage (for modified membrane) or to be at an almost similar rate
to the initial stage (for virgin membrane); on the contrary, it is shown to be faster (faster
increase in resistance). Faster deposition rate of BSA in the second stage can be explained
by extra protein–protein interactions due to intermolecular forces. In regard to this, Kelly
and Zydney [28] found that thiol oxidation and intermolecular thiol-disulfide bonds play a
significant role in protein-protein interactions.

Alginate fouling diagrams are also depicted in Figure 6.3 and Figure 6.4. In comparison,
the resistance of membranes increase more rapidly which means alginate fouls the
membrane faster than BSA. Similar to BSA fouling, the modified membrane shows a
lower resistance increase rate than virgin membrane in fouling by alginate. Obtained
results are consistent with calculated attractive adhesion energies between membranes and
alginate, with values of –25.46 mJ/m2 and –8.33 mJ/m2 for virgin and modified
membranes, respectively. An attractive cohesive energy with the value of –24.58 mJ/m2
was calculated for alginate. Based on calculated attractive adhesive and cohesive energies,
alginate is expected to cause a continual increase in membrane resistance (due to alginate
deposition) that never levels off. Nevertheless, fouling by alginate slows down after about
6 hours of fouling. This level-off trend of the membrane resistance can be explained in the
light of membrane surface topography (AFM image of membranes are shown in Chapter
4). Once the surface valleys are clogged up, any additional deposited foulant will be much
more exposed to hydrodynamic shear forces, causing deposition rate to slow down.

6.5 Conclusion
In this work, the Extended Derjaguin-Landau-Verwey-Overbeek XDLVO theory
suggested by van Oss was applied to obtain a more comprehensive picture of the effect of
membrane surface modification by 3-(3,4-Dihydroxyphenyl)-L-alanine (L-DOPA) on the
interaction energies between membrane and biopolymers as model organic foulants. This
approach to predict the interactions between membranes and biopolymers is based on the
hypothesis that dissolved biopolymers behave similar to colloidal particles. Lifshitz-van
der Waals, Lewis acid-base and electrostatic double layer components of free energy of
adhesion were calculated using measured contact angles of three probe liquids and
measured zeta potential of membranes and model organic foulants (bovine serum albumin
and alginate). It was found that L-DOPA mainly affected the acid-base component of free

121
CHAPTER SIX

energy of adhesion between modified membrane and model foulants. Thus, due to the
enhanced acid-base interaction energy, lower affinities toward the adhesion of both bovine
serum albumin and alginate were calculated for the L-DOPA modified membrane. Actual
fouling tests with feed solutions of studied biopolymers were in good agreement with the
theoretical predictions. Analysis of the adhesion energies also indicated that in solutions
with moderate or high ionic strengths where the electrostatic double layer interactions
decline, the acid-base interaction energy has the highest effect on biopolymers adhesion.
Therefore, it can be expected that L-DOPA modified membranes can maintain their fouling
resistance performance even in solution with high ionic strengths.

The outcome of this study suggests that XDLVO theory can be successfully used to assess
the effect of membrane surface modification on its fouling propensity. Furthermore, owing
to the simple modification procedure and desirable surface properties, L-DOPA is
considered as a promising material for industrial scale anti-fouling membrane
modifications.

122
CHAPTER SIX

References
[1] A.I. Schäfer, A.G. Fane, T.D. Waite, Fouling effects on rejection in the membrane
filtration of natural waters, Desalination, 131 (2000) 215-224.
[2] A.R. Costa, M.N. de Pinho, M. Elimelech, Mechanisms of colloidal natural organic
matter fouling in ultrafiltration, Journal of Membrane Science, 281 (2006) 716-725.
[3] T.H. Chong, F.S. Wong, A.G. Fane, Implications of critical flux and cake enhanced
osmotic pressure (CEOP) on colloidal fouling in reverse osmosis: Experimental
observations, Journal of Membrane Science, 314 (2008) 101-111.
[4] S. Hong, M. Elimelech, Chemical and physical aspects of natural organic matter
(NOM) fouling of nanofiltration membranes, Journal of Membrane Science, 132 (1997)
159-181.
[5] W. Lee, S. Lee, C.H. Ahn, S. Hong, S. Kim, Y. Baek, J. Yoon, Evaluation of surface
properties of reverse osmosis membranes on the initial biofouling stages under no filtration
condition, Journal of Membrane Science, 351 (2010) 112-122.
[6] J. Mansouri, S. Harrisson, V. Chen, Strategies for controlling biofouling in membrane
filtration systems: challenges and opportunities, Journal of Materials Chemistry, 20 (2010)
4567-4586.
[7] S. Belfer, Y. Purinson, R. Fainshtein, Y. Radchenko, O. Kedem, Surface modification
of commercial composite polyamide reverse osmosis membranes, Journal of Membrane
Science, 139 (1998) 175-181.
[8] A.C. Sagle, H. Ju, B.D. Freeman, M.M. Sharma, PEG-based hydrogel membrane
coatings, Polymer, 50 (2009) 756-766.
[9] S. Azari, L. Zou, Using zwitterionic amino acid L-DOPA to modify the surface of thin
film composite polyamide reverse osmosis membranes to increase their fouling resistance,
Journal of Membrane Science, 401–402 (2012) 68-75.
[10] F.M. Fowkes, Role of acid-base interfacial bonding in adhesion, Journal of Adhesion
Science and Technology, 1 (1987) 7-27.
[11] C.J. Van Oss, Interfacial forces in aqueous media, 2nd ed., Taylor & Francis, Boca
Raton, Fla., 2006.
[12] J.A. Brant, A.E. Childress, Colloidal adhesion to hydrophilic membrane surfaces,
Journal of Membrane Science, 241 (2004) 235-248.
[13] J.A. Brant, A.E. Childress, Assessing short-range membrane–colloid interactions
using surface energetics, Journal of Membrane Science, 203 (2002) 257-273.
[14] S. Lee, S. Kim, J. Cho, E.M.V. Hoek, Natural organic matter fouling due to foulant–
membrane physicochemical interactions, Desalination, 202 (2007) 377-384.
[15] S. Kim, E.M.V. Hoek, Interactions controlling biopolymer fouling of reverse osmosis
membranes, Desalination, 202 (2007) 333-342.
[16] S. Botton, A.R.D. Verliefde, N.T. Quach, E.R. Cornelissen, Influence of biofouling on
pharmaceuticals rejection in NF membrane filtration, Water Research, 46 (2012) 5848-
5860.
[17] A. Kitahara, A. Watanabe, Electrical phenomena at interfaces, M. Dekker, New York
:, 1984.

123
CHAPTER SIX

[18] S.H. Lee, E. Ruckenstein, Adsorption of proteins onto polymeric surfaces of different
hydrophilicities—a case study with bovine serum albumin, Journal of Colloid and Interface
Science, 125 (1988) 365-379.
[19] J.H. Waite, Nature's underwater adhesive specialist, International Journal of Adhesion
and Adhesives, 7 (1987) 9-14.
[20] M. Yu, J. Hwang, T.J. Deming, Role of l-3,4-Dihydroxyphenylalanine in Mussel
Adhesive Proteins, Journal of the American Chemical Society, 121 (1999) 5825-5826.
[21] H. Lee, N.F. Scherer, P.B. Messersmith, Single-molecule mechanics of mussel
adhesion, Proceedings of the National Academy of Sciences, 103 (2006) 12999-13003.
[22] H.J. Jacobasch, J. Schurz, Characterization of polymer surfaces by means of
electrokinetic measurements, in: K. Hummel, J. Schurz (Eds.) Dispersed Systems,
Steinkopff, 1988, pp. 40-48.
[23] F. Fairbrother, H. Mastin, CCCXII.-Studies in electro-endosmosis. Part I, Journal of
the Chemical Society, Transactions, 125 (1924) 2319-2330.
[24] Y. Zhao, F. Li, M.T. Carvajal, M.T. Harris, Interactions between bovine serum
albumin and alginate: An evaluation of alginate as protein carrier, Journal of Colloid and
Interface Science, 332 (2009) 345-353.
[25] C.Y. Tang, T.H. Chong, A.G. Fane, Colloidal interactions and fouling of NF and RO
membranes: A review, Advances in Colloid and Interface Science, 164 (2011) 126-143.
[26] M. Elimelech, W.H. Chen, J.J. Waypa, Measuring the zeta (electrokinetic) potential of
reverse osmosis membranes by a streaming potential analyzer, Desalination, 95 (1994)
269-286.
[27] A.E. Childress, M. Elimelech, Effect of solution chemistry on the surface charge of
polymeric reverse osmosis and nanofiltration membranes, Journal of Membrane Science,
119 (1996) 253-268.
[28] I.H. Huisman, P. Prádanos, A. Hernández, The effect of protein–protein and protein–
membrane interactions on membrane fouling in ultrafiltration, Journal of Membrane
Science, 179 (2000) 79-90.
[29] S.T. Kelly, A.L. Zydney, Mechanisms for BSA fouling during microfiltration, Journal
of Membrane Science, 107 (1995) 115-127.
[30] S.T. Kelly, A.L. Zydney, Effects of intermolecular thiol–disulfide interchange
reactions on BSA fouling during microfiltration, Biotechnology and Bioengineering, 44
(1994) 972-982.

124
7 Conclusions and Future Work Recommendations

Due to fresh water scarcity and ever increasing o l pop l io , i g h ol ’


future water demands will be a serious challenge. However, immediate relief could be
provided if water treatment technologies could economically purify water from
alternative water sources. With the recent advances in membrane technology, it has
emerged as a viable method for this purpose. However, all membrane processes suffer
from a decline in their performance due to fouling. To address this issue, surface
modification methods have been introduced as an approach to reduce the adhesion of
foulants to the membrane surface. Several technical concerns need to be tackled for a
successful modification. It should (i) be cost-effective (ii) not cause damage to the
membrane structure and performance (iii) possess a long life span (iv) maintain its
performance under different operational conditions. Given these critical requirements,
there are a few materials capable of meeting the challenges of practical applications. In
this study, we aimed to explore new approaches for design and development of surface
modification methods. This project mainly focused on the surface modification of
commercial polyamide reverse osmosis (RO) membranes commonly used for drinking
water purification from seawater.

7.1 Conclusions
This study adopted the application of the bioinspired adhesive zwitterionic aminoacid
3-(3,4-Dihydroxyphenyl)-L-alanine (L-DOPA) as a coating on SW30XLE RO
membranes to alter their surface properties for fouling-resistance purposes. A controlled
top-layer surface treatment method was employed to deposit L-DOPA on the membrane
surface. Visual inspection of modified membrane, revealed a dark brown coating on the
membrane surface. Atomic force microscopy (AFM) topographical images showed that
the modification resulted in a smoother membrane surface. Compared to the virgin
membrane, 33% reduction in root mean square (RMS) roughness of the 24-hour-coated
membrane was observed which can be explained by the fact that L-DOPA more readily

125
CHAPTER SEVEN

deposits in the surface valleys. Measurement of the zeta potential of the membrane
surface showed that L-DOPA modification had little or no effect on the surface charge
of the membrane. A 56 % reduction in the water contact angle of the 24-hour-coated
membrane indicated a remarkable increase in the hydrohphilicity. While the salt
rejection was maintained, a 15% increase in the water permeability of the
12-hour-coated sample was observed which can be attributed to the increase in
membrane wettability.

After 12 hours of cross-flow filtrations of an 100 g/L bovine serum albumin (BSA)
solution and an 100 g/L alginic acid sodium salt (alginate) solution, 0.32×1012 m-1 and
0.65×1012 m-1 were added to the resistance of 12-hour-coated membrane, respectively.
In comparison, virgin membrane showed 0.5×1012 m-1 and 0.87×1012 m-1 increase in its
resistance, respectively. Cross-flow filtration of dodecyltrimethylammonium (DTAB),
which represents a model positively, charged polymer foulant, also showed the
improved fouling resistance of the coated membrane. After 300 minutes filtration of a
50 mg/L DTAB solution, the 24-hour coated membrane displayed 0.74×1012 m-1
resistance increase while 1.4×1012 m-1 was added to the resistance of virgin membrane.
The coated membrane also exhibited a significant improvement in the fouling tests
using the mixed foulant solution. After 17 hours filtration of the mixed foulant solution
of 100 g/L BSA and 100 g/L alginate, the 12-hour coated membrane displayed 1.2×1012
m-1 resistance increase while 1.96×1012 m-1 increase in the resistance of virgin
membrane was observed. This demonstrated that L-DOPA can be successfully applied
for low fouling surface modification of commercial polyamide RO membranes. Since
membrane modification is facile and can be performed in situ, this approach is of
particular interest from a practical stand point, as the membranes can be treated in their
original module assembly.

In order to further investigate the fouling reduction effect of L-DOPA coating on


another type of membrane, a cellulose acetate (CA) microfiltration membrane was also
chosen to be coated with L-DOPA. Contact angle results associated with the increase
water permeability of the 8- and 16-hour coated membrane, confirmed a small increase
in the wettability of the modified membranes. Zeta potential measurement revealed a
considerable higher negative surface charge of the 24-hour-coated CA membrane.
24-hour-coated membrane also showed 5% less flux than the virgin CA membrane. This

126
CHAPTER SEVEN

can be attributed to a 2% reduction in the measured porosity. Static BSA fouling tests
indicated improved resistance of the 24-hour-coated membrane to BSA adhesion; after
immersing the membranes in a 0.2 g/L BSA solution for 3 hours, the modified
membrane lost 19 % of its flux which was lower than 26% loss in the flux of the virgin
membrane. Further investigation employing dynamic BSA filtration demonstrated that
the fouling resistance of the modified membranes was improved. The extent of the
improvement depended on the BSA solution concentration.

It was concluded that L-DOPA modification improved the fouling resistance of both
commercial polyamide RO and cellulose acetate microfiltration membranes; however,
two distinct interactions seems to be responsible for these improvements for these two
membrane types. The low fouling effect of the coated polyamide RO membrane can be
attributed to its superior hydrophilicity. On the other hand, the low fouling behaviour of
the coated CA microfiltration membrane can be rather explained on the grounds of its
higher net negative charge which raises the electrostatic repulsion between the BSA
molecule and the membrane surface.

Successful application of the zwitterionic amino acid L-DOPA to achieve fouling


reduction inspired to expand the understanding of the effect of membrane modification
with other zwitterionic amino acids on membrane fouling resistance. Therefore the
fouling-resistant surface modification of the RO membranes using the zwitterionic
amino acid L-cysteine was also investigated. This modification was first performed
through the activation of the membrane with an epoxy compound (allyl glycidyl ether).
L-cysteine was then grafted onto the activated membrane via the thiol-ene reaction.
Membrane characterization using FTIR and XPS supported the successful incorporation
of L-cysteine onto the membrane. The water contact angle of the modified membrane
was 30 % smaller than that of the virgin membrane. This indicates a more hydrophilic
surface of the modified membrane. AFM topographical images revealed the
modification had resulted in a smoother membrane surface with 20% reduction in RMS
roughness. Although the modification process was found to favourably enhance salt
rejection of the membrane by 3.2%, 7% deterioration in the permeability was observed.
These variations in salt rejection and membrane permeability were linked to a tighter
structure of the polyamide layer. In both BSA and DTAB fouling tests, the modified
membrane displayed lower fouling propensity compared to the virgin one. After 12

127
CHAPTER SEVEN

hours cross flow filtration of 100 g/L BSA solution, 0.5×1012 m-1 was added to the
resistance of the modified membrane whereas the resistance of virgin membrane was
increased by 0.58×1012 m-1. Also using 20 mg/L DTAB solution as model foulant, the
modified membrane displayed 0.77×1012 m-1 resistance increase whereas 0.95×1012 m-1
resistance increase was observed for virgin membrane. Since at the pH values of the
foulant solutions, both virgin and modified membranes displayed almost equal surface
charge, it was concluded that electrostatic interactions had no effect on the improved
performance of the modified membrane. The superior fouling resistance of the
L-cysteine grafted membrane must thus be explained in the light of its improved
hydrophilicity and surface morphology.

Both L-DOPA coating and L-cysteine grafting mitigated membrane fouling; however,
the enhancement effect of L-DOPA was more pronounced. Moreover, L-DOPA
modification did not adversely affect the membrane flux. Therefore, due to the versatile
and easy modification process of L-DOPA, it deems to be a superior candidate for
fouling resistant surface modification of delicate RO membranes.

Finally to obtain a more comprehensive picture of the effect of L-DOPA surface


modification on the interactions between polyamide RO membrane and model foulants,
the Extended Derjaguin-Landau-Verwey-Overbeek (XDLVO) theory suggested by van
Oss was applied to evaluate the free energy of adhesion between membranes and model
foulants. This theory suggests that free energy of adhesion in aqueous systems is
comprised of three components viz. apolar or Lifshitz-van der Waals (LW) energy,
polar or Lewis acid-base (AB) energy and electrostatic double layer (EL) energy. The
role of L-DOPA in improving the membrane fouling resistance was attributed to the
increasing the acid-base component of the free energy of adhesion. Contrary to the
electrostatic double layer energy, this energy does not depend on the ionic strength of
the solution. Therefore, it can be concluded that L-DOPA modified polyamide RO
membranes are able to maintain their low fouling performance even in solutions of very
high ionic strength. The predictions of XDLVO theory were in good agreement with the
findings of the filtration experiments. The outcome of this study also suggests the
potential application of XDLVO theory for assessing the effect of membrane surface
modification on membrane fouling propensity.

128
CHAPTER SEVEN

7.2 Recommendations for future work


This research study mainly focused on the investigation of the effect of surface
modification with L-DOPA on mitigation of membrane fouling by organic foulants.
However, fundamental cross disciplinary research is still needed to discover the
deposition and polymerization mechanism of L-DOPA. Understanding the exact
structure of poly L-DOPA and the way it attaches to the membrane surface can help to
more widespread application of this material for surface modification or surface
functionalization. Therefore, detailed material chemical structure characterization may
be recommended. Also details on coating thickness and uniformity may be investigated
in future studies.

In this work, the methods utilized to test the antifouling effect of the modified
membranes were based on the equipment available in our laboratory; it would be
interesting if supplementary studies could be carried out to precisely measure the
amount of foulants adhered to the surface. Further expansion to the measurement of
adhesion of bacteria to the membrane surface would be highly desirable.

L-DOPA modification was successful in reducing fouling in laboratory scale cross-flow


filtration of synthetic water. It would then be a valuable contribution to evaluate its
performance in real seawater applications.

More thoughts need to be given on the long term stability of the coating. Investigation
of the L-DOPA stability to vigorous cleaning chemicals such as acid, base and
surfactants as well as its resistance to chlorine would also be advantageous.
Furthermore, studies on the resistance of coating to biological degradation in long term
operation are recommended.

129

Vous aimerez peut-être aussi