103
BIOFOULING CONTROL OF REVERSE OSMOSIS MEMBRANES USING FREE NITROUS ACID Jingshi Wang Bachelor of Chemical Engineering (Honours) A thesis submitted for the degree of Master of Philosophy at The University of Queensland in 2015 School of Chemical Engineering Advanced Water Management Centre

BIOFOULING CONTROL OF REVERSE OSMOSIS …379468/s43348651_mphil... · Reverse osmosis (RO) membranes have been widely applied in membrane filtration processes for water purification,

Embed Size (px)

Citation preview

BIOFOULING CONTROL OF REVERSE OSMOSIS

MEMBRANES USING FREE NITROUS ACID

Jingshi Wang

Bachelor of Chemical Engineering (Honours)

A thesis submitted for the degree of Master of Philosophy at

The University of Queensland in 2015

School of Chemical Engineering

Advanced Water Management Centre

i

ABSTRACT

Reverse osmosis (RO) membranes have been widely applied in membrane filtration

processes for water purification, since the high selective RO membranes are designed to

reject all materials with particle diameter larger than 10 angstrom (Å) [1]. However, this

optimal selectivity leads to fouling that can greatly affect the performance and productivity of

RO membranes. Biofouling remains as one of the major operational problems in RO

processes and is caused by unwanted deposit and growth of microorganisms on the

membrane. Numerous biofouling control strategies have been developed to restore the

performance of RO membranes, but none of them are able to prevent or remove biofouling

completely. A novel cleaning technique using a weak and monobasic acid (pKa=3.34, 25℃)

named free nitrous acid (FNA) in combined with hydrogen dioxide (H2O2) was proposed.

The effects of FNA with or without H2O2 on biofouling of RO membranes were investigated

in Chapter 4, five RO membranes with different degree of biofouling were cleaned using

FNA solutions (10, 35 and 47 mg HNO2-N/L) at pH 2.0, 3.0 and 4.0 under cross-flow

conditions for 24 hours. The cleaning efficiency of FNA solutions was compared with

conventional cleaning solution sodium hydroxide (NaOH, pH 11). The cleaning tests

demonstrated that FNA cleaning solutions were more efficient than NaOH at biomass

removal and inactivation. At the optimum cleaning conditions (35 mg HNO2-N/L at pH 3.0),

FNA has achieved higher biomass removal than NaOH for both heavily fouled (86-96%

versus 41-83%) and moderately fouled (92-95% against 89-92%) membranes, respectively.

In accordance to the biomass removal, 6-32% of viable cells remained on the moderately

fouled RO membranes under the impact of FNA cleaning (pH 3), whereas 38-58% of viable

cells stayed on the heavily fouled RO membranes. These results revealed that FNA cleaning

is more effective for moderately fouled membranes, implying that early cleaning for

biofouling is preferable. Although applying FNA alone, or combining it with H2O2 have

shown better efficiency at biofouling removal than NaOH, the cleaning efficiency has not

been significantly improved (<1% of enhancement) by adding H2O2 to FNA cleaning

solutions. The effects of FNA on scaling of RO membranes were also studied using the same

cleaning protocol developed for biofouling control. The results showed that FNA solutions at

pH 2.0 and 3.0 were as efficient as conventional cleaning acids (hydrochloric acid and citric

acid). The scaling layers which contain 32.4±1.7 g/cm2 of calcium were completely removed

by all acidic cleaning solutions. Based on the results, FNA is shown to be a promising

ii

cleaning agent for RO membrane biofouling and scaling removal.

Further investigation focused on the effectiveness of FNA for biofouling prevention in RO

processes (Chapter 5). The results showed that weekly FNA cleanings were unable to prevent

fouling in the RO filtration systems, as the hydraulic performances (permeability and salt

rejection) of RO membranes have gradually declined over two to three weeks filtration period.

Although FNA cleaning was able to restore the permeability of RO membranes for one to

two days, continuing declined permeability implied that the fouling rate was greater than the

inhibition rate of FNA. The results of prevention tests also showed that FNA was more

efficient at biomass inactivation and removal. The biomass contents and viable cells of the

fouling layers formed in the experiment filtration unit (with FNA weekly cleaning) were less

than half of that in the control filtration unit (without FNA weekly cleaning). Moreover, the

results of live/dead cell staining revealed the abundance of viable cells in the control unit

(57±5%) was four times higher than that in the experiment unit (13±2%). However, there was

no significant difference in the concentration of macromolecules such as proteins and

polysaccharides between control and experiment filtration units.

iii

Declaration by author

This thesis is composed of my original work, and contains no material previously published

or written by another person except where due reference has been made in the text. I have

clearly stated the contribution by others to jointly-authored works that I have included in my

thesis.

I have clearly stated the contribution of others to my thesis as a whole, including statistical

assistance, survey design, data analysis, significant technical procedures, professional

editorial advice, and any other original research work used or reported in my thesis. The

content of my thesis is the result of work I have carried out since the commencement of my

research higher degree candidature and does not include a substantial part of work that has

been submitted to qualify for the award of any other degree or diploma in any university or

other tertiary institution. I have clearly stated which parts of my thesis, if any, have been

submitted to qualify for another award.

I acknowledge that an electronic copy of my thesis must be lodged with the University

Library and, subject to the policy and procedures of The University of Queensland, the thesis

be made available for research and study in accordance with the Copyright Act 1968 unless a

period of embargo has been approved by the Dean of the Graduate School.

I acknowledge that copyright of all material contained in my thesis resides with the copyright

holder(s) of that material. Where appropriate I have obtained copyright permission from the

copyright holder to reproduce material in this thesis.

iv

Publications during candidature

No publications

Publications included in this thesis

No publications included

v

Contributions by others to the thesis

I would like to thank Professor Jin Zou and Mr Zhi Zhang from the Centre for Microscopy &

Microanalysis (CMM), Australian Institute for Bioengineering and Nanotechnology (AIBN),

University of Queensland (UQ) assisted in SEM-EDS analysis.

Statement of parts of the thesis submitted to qualify for the award of another degree

None.

vi

ACKNOWLEDGEMENTS

I wish to express my first and sincere gratitude to my supervisors Professor Zhiguo Yuan and

Dr Emmanuelle Filloux for their intellectual and invaluable advice in the course of this

research project. I would also like to acknowledge our industrial partners Australian Water

Recycling Centre of Excellence, National Centre of Excellence in Desalination, Seqwater,

and Veolia water, who generously sponsor this research project.

I further extend my gratitude to Dr Wolfgang Gernjak, Dr Phillip Bond, Dr Bogdan Donose

and Dr Liu Ye for reviewing the progress of my work and providing valuable feedbacks. I am

grateful to Dr Kenn Lu for his assistance in the microbiological analysis. I would like to

acknowledge Professor Jin Zou and Mr Zhi Zhang for SEM imaging. I would like to thank Dr.

Beatrice Keller-Lehman, Jianguang Li and Nathan Clayton for their analytical support, and

also to the administration staff for all their indispensable assistance through the entire process.

I am very grateful to all the members of the drinking water group, my friends and colleagues

at AWMC, Thank you to Dang, Apra, Shao, Elisabet, Guillermo and Zanina for providing

countless assistance on both academic and personal levels. I am also thankful to all my

friends outside the university, particularly Alfreda, Angela, Joy, Clare, Jeomo and Penny, for

their continued friendship, support, chats and laughs along the journey.

My final thanks are reserved for my parents and my family who have been a continual source

of support, strength and motivation and for that I am forever grateful.

vii

Keywords

Free nitrous acid, Biofouling control, Scaling control, Membrane cleaning, Reverse osmosis

membrane filtration.

Australian and New Zealand Standard Research Classifications (ANZSRC)

ANZSRC code: 090404 Membrane and Separation Technologies, 50%

ANZSRC code: 090301 Analytical Chemistry, 25%

ANZSRC code: 090605 Microbiology, 25%

Fields of Research (FoR) Classification

FoR code: 0904, Chemical Engineering, 100%

viii

TABLE OF CONTENTS

ABSTRACT ................................................................................................................................................................... I

ACKNOWLEDGEMENTS .................................................................................................................................... VI

TABLE OF CONTENTS ...................................................................................................................................... VIII

LIST OF FIGURES .................................................................................................................................................... X

LIST OF TABLES .................................................................................................................................................... XII

LIST OF ABBREVIATIONS AND NOMENCLATURE ............................................................................... XIII

1. INTRODUCTION ............................................................................................................................................. 1

§ 1.1. RESEARCH MOTIVATION ................................................................................................................ 1

§ 1.2. GENERAL RESEARCH OBJECTIVES ............................................................................................. 2

2. LITERATURE REVIEW ................................................................................................................................ 3

§ 2.1. INTRODUCTION .................................................................................................................................. 3

§ 2.2. RO MEMBRANES FILTRATION ...................................................................................................... 3

§ 2.3. FOULING ON RO MEMBRANES ..................................................................................................... 4

§ 2.3.1. Biofouling ....................................................................................................................................... 5

§ 2.3.2. Scaling ............................................................................................................................................. 7

§ 2.4. RO MEMBRANES BIOFOULING PREVENTION AND CLEANING METHODS ............... 9

§ 2.4.1. Prevention Methods ..................................................................................................................... 9

§ 2.4.2. Cleaning Methods ...................................................................................................................... 10

§ 2.5. FREE NITROUS ACID (FNA) ........................................................................................................... 13

§ 2.5.1. FNA and Its Biocidal Effect ....................................................................................................... 13

§ 2.5.2. FNA versus Other Biocides ....................................................................................................... 15

§ 2.6. RESEARCH GAP .................................................................................................................................. 16

3. RO MEMBRANE FOULING CHARACTERISATION .........................................................................19

§ 3.1. INTRODUCTION .................................................................................................................................19

§ 3.2. MATERIALS AND METHODS......................................................................................................... 20

§ 3.2.1. Membranes................................................................................................................................... 20

§ 3.2.2. Sample Preparation .................................................................................................................... 20

§ 3.3. MEMBRANE AUTOPSY .................................................................................................................... 20

§ 3.3.1. Biological Analysis ...................................................................................................................... 20

§ 3.3.2. Organic and Molecular Analysis ............................................................................................. 22

§ 3.3.3. Elementary Analysis ................................................................................................................... 24

§ 3.3.4. RO Membrane Hydraulic Performances ............................................................................... 24

§ 3.4. RESULTS AND DISCUSSION .......................................................................................................... 26

ix

§ 3.4.1. Visual Inspection ........................................................................................................................ 26

§ 3.4.2. Biofouling Characterisation ..................................................................................................... 27

§ 3.4.3. Scaling Characterisation ........................................................................................................... 32

§ 3.4.4. Hydraulic Performances of RO Membranes ......................................................................... 34

§ 3.5. CONCLUSION ..................................................................................................................................... 34

4. RO MEMBRANE CLEANING USING FNA........................................................................................... 36

§ 4.1. INTRODUCTION ................................................................................................................................ 36

§ 4.2. MATERIALS AND METHODS......................................................................................................... 37

§ 4.2.1. Cleaning Set-Up and Operation .............................................................................................. 37

§ 4.2.2. Design of Cleaning Tests ........................................................................................................... 37

§ 4.2.3. Post-Cleaning Analyses ............................................................................................................. 40

§ 4.3. RESULTS AND DISCUSSION .......................................................................................................... 42

§ 4.3.1. Effects of FNA Cleaning on RO Biofouling under Cross-Flow Conditions .................. 42

§ 4.3.2. Descaling Efficiency of FNA ...................................................................................................... 51

§ 4.3.3. Hydraulic Performances of RO Membranes after Cleaning ............................................. 53

§ 4.4. CONCLUSION ..................................................................................................................................... 55

5. RO BIOFILM PREVENTION USING FNA ............................................................................................. 57

§ 5.1. INTRODUCTION ................................................................................................................................ 57

§ 5.2. MATERIALS AND METHODS......................................................................................................... 58

§ 5.2.1. Membranes................................................................................................................................... 58

§ 5.2.2. Bench-Scale RO Filtration Units ............................................................................................ 58

§ 5.2.3. Operation Conditions and Analyses ...................................................................................... 59

§ 5.3. RESULTS AND DISCUSSION ........................................................................................................... 61

§ 5.3.1. Effects of FNA Dosing on the Hydraulic Performances of RO Membranes ................... 61

§ 5.3.2. Effect of FNA on the membrane fouling layer ...................................................................... 65

§ 5.4. CONCLUSIONS ................................................................................................................................... 67

6. CONCLUSIONS AND RECOMMENDATIONS ................................................................................... 68

§ 6.1. CONCLUSIONS ................................................................................................................................... 68

§ 6.2. RECOMMENDATIONS..................................................................................................................... 69

REFERENCES .......................................................................................................................................................... 72

APPENDIXES ............................................................................................................................................................ 80

APPENDIX A. CALIBRATION CURVE FOR ATP MEASUREMENT ..................................................................... 80

APPENDIX B. ADDITIONAL RESULTS OF FISH ANALYSIS .................................................................................. 81

APPENDIX C. LIVE/DEAD CELLS REPRESENTED BY CLSM IMAGES ................................................................ 84

APPENDIX D. ICP-OES AND SEM-EDS RESULTS OF DESCALING TESTS ...................................................... 87

x

LIST OF FIGURES

Figure 2-1. Schematic diagrams of cross-flow pattern (left) and configuration of the spiral-

wound RO module (right) [20]. ................................................................................................. 3

Figure 2-2. Five stages of biofilm development [52]................................................................ 6

Figure 2-3. Schematic demonstration of scale formation [37] .................................................. 8

Figure 3-1. Cross flow filtration unit. ..................................................................................... 25

Figure 3-2. Visual inspection on fouled membranes (RO1-6). ............................................... 26

Figure 3-3. Biomass (ATP) contents in the fouling layers of each RO membranes. The error

bars show the standard errors of the replicate samples (n=5). ................................................. 28

Figure 3-4. The abundances of proteobacteria (Alpha, Beta, and Gamma) and archaea in the

biofouling layers of membrane RO1 (a), RO3 (b), RO4 (c) and RO5 (d). .............................. 29

Figure 3-5. Total solid, volatile solid and organic fraction of fouling layers deposited on

membranes. The error bars show the standard errors of the replicate samples (n=5). ............ 30

Figure 3-6. SEM image (left) and EDS element weight percentage (right) of RO6 scaling

layer.......................................................................................................................................... 33

Figure 4-1. The cleaning cell (left), and the entire cleaning system with five cleaning cells in

parallel, fed with a single peristaltic pump with five pump heads........................................... 37

Figure 4-2. Biomass residual after 24 hours cleaning tests performed in cross-flow conditions

with the membranes (a) heavily fouled RO1, RO2 and RO3, and (b) moderately fouled RO4

and RO5. The cross-flow velocity applied was 0.1 m/s. The error bars show the standard

errors of three replicate experiments. The results without error bars are based on three

measurements from each experiment....................................................................................... 44

Figure 4-3. Proportion of viable cells in membrane biofilm before and after 24 hours

cleaning tests for membranes RO4 and RO5. A cross-flow velocity of 0.1 m/s was applied.

The error bars show the standard errors of 15 to 60 CLSM images. ....................................... 46

Figure 4-4. The abundance of proteobacteria (Alpha, Beta, and Gamma) and archaea in the

biofouling layers before and after 24 hours cleaning tests for the membranes RO1 and RO5,

respectively. Standard test conditions: FNA (50 mgN-NO2/L), pH 3, cross-flow velocity 0.1

m/s. The abundances of each microbe were calculated based on the FISH images (n=20±5).

.................................................................................................................................................. 47

Figure 4-5. Biomass removal (%, based on ATP values), protein and polysaccharide removal

(%) after 24 hours cleaning tests for the membranes (a) RO4 and (b) RO5. A cross-flow

xi

velocity 0.1 m/s was applied. The error bars in the plot show the standard errors of 2-3

replicate experiments. .............................................................................................................. 49

Figure 4-6. Biomass removal after 24 hours cleaning tests for the membranes heavily fouled

RO2 and RO3 and moderately fouled RO4. A cross-flow velocity 0.1 m/s was applied. The

error bars show the standard errors of three replicate experiments. The results without error

bars are based on three measurements from each experiment. ................................................ 50

Figure 4-7. Dissolved calcium content removed from the membrane surface after 24 hours

cleaning tests with membrane RO6. A cross-flow velocity 0.1 m/s was applied in all tests.

The error bars show the standard errors of four measurements from two replicate experiments.

.................................................................................................................................................. 52

Figure 5-1. A schematic diagram of the crossflow membrane filtration unit ......................... 59

Figure 5-2. Normalised permeability (Kw/Kw0) of filtration test 1 (a) and test 2 (b). Standard

test conditions: SE was circulating at 0.1 m/s in both filtration systems. For the experiment

filtration, the data points highlighted in red represent that the membrane was cleaned by FNA

(10 mg-N/L) at pH 3 for 6 hours on these specific. The scatter points are based on two

measurements from each experiment on each day. ................................................................. 62

Figure 5-3. Salt rejection of the filtration test 1 (a) and test 2 (b). Standard test conditions: SE

was circulating at 0.1 m/s in both filtration systems. For the experiment filtration, the data

points highlighted in red represent that the membrane was cleaned by FNA (10 mg-N/L) at

pH 3 for 6 hours on these specific dates. The scatter points are based on two measurements

from each experiment on each day. ......................................................................................... 64

Figure 5-4. CLSM images of fouling samples from control (a) and experimental (b) filtration

units. Live cells were stained in green with the DNA specific dye SYTO®

9 and the dead cells

are stained in red with PI. Standard test conditions: a cross-flow velocity 0.1 m/s was applied.

FNA (10 mg FNA-N/L) at pH 3 was cleaned for 6 hours in the experimental filtration unit on

a weekly base. .......................................................................................................................... 66

xii

LIST OF TABLES

Table 2-1: Major categories of membrane cleaning chemicals [19] .................................. 11

Table 2-2: Studies of biocidal effects of FNA on microorganisms..................................... 13

Table 3-1: RO membranes used in this study. .................................................................... 20

Table 3-2: Oligonucleotide probes used in FISH analysis .................................................. 22

Table 3-3: Protein and polysaccharides deposited on RO membranes. ........................... 31

Table 3-4: Metal concentration of fouling layers for each membrane (ICP-OES results)

.................................................................................................................................................. 32

Table 3-5: Hydraulic performances of RO membrane coupons (n=4). ............................ 34

Table 4-1: Design of the biofouling control tests. ................................................................ 38

Table 4-2: The repetition of the cleaning tests .................................................................... 39

Table 4-3: Design of the descaling tests with RO6. ............................................................. 39

Table 4-4: Comparison of FNA and NaOH cleaning effects on proteins and

polysaccharides. ..................................................................................................................... 48

Table 4-5: Hydraulic performances of membranes after cleaning tests. .......................... 54

Table 5-1: The composition of synthetic nutrient [146]. .................................................... 60

Table 5-2: The operation conditions of filtration tests and foulant analyses performed

after the filtration tests. ......................................................................................................... 60

Table 5-3: Summary results of foulant analyses ................................................................. 65

xiii

LIST OF ABBREVIATIONS AND NOMENCLATURE

Abbreviation

ATP: adenosine triphosphate

BSA: bovine serum albumin

CA: cellulose acetate

CFV: cross flow velocity

CLSM: confocal laser scanning microscope

CP: concentration polarisation

Cy: cyanine

DABCO: 1,4-Diazabicyclo[2.2.2]octane, Sigma-Aldrich

DBNPA: 2,2-dibromo-3-nitrilopropionamide

DI:

DNA: deoxyribonucleic acid

EDS: energy dispersive spectroscopy

EDTA: ethylenediaminetetraacetic acid

EPS: extracellular polymeric substances

FISH: fluorescence in-situ hybridization

FITC: fluorescein isothiocyanate

FNA: free nitrous acid

ICP: inductively coupled plasma

LOD: limit of detection

LOI: loss on ignition

N.D.: not determined

MF microfiltration

NF: nanofiltration

NP: normalised permeability

OES: optical emission spectrometry

PBS: phosphate-buffered saline

PFA: paraformaldehyde

PI: propidium iodide

PV: polyamide

RLU: relative light units

xiv

RNA: ribonucleic acid

RNS: reactive nitrogen species

RO: reverse osmosis

SBS: sodium bisulfite

SDS: sodium dodecyl sulfate

SE: secondary effluent

SEM: scanning electron microscopy

TFC: thin film composite

UF: ultrafiltration

UV: ultraviolet

Nomenclature

Å Angstrom

Cf: feed conductivity

Cp: permeate conductivity

J: flux (l∙h-1

∙m-2

)

K: permeability (l∙h-1

∙ bar-1

∙m-2

)

Ka: ionization constant

Kw: temperature corrected permeability (l∙h-1

∙ bar-1

∙m-2

, 25 ℃)

OF: organic fraction (%)

∆𝑃: transmembrane pressure (bar)

Qp: permeate flow rate (l∙h-1

)

S: membrane surface area (m2)

T: temperature (℃)

TS: total solids concentration(g.m-2

)

VS: volatile solids concentration (g.m-2

)

1

1. INTRODUCTION

§ 1.1. RESEARCH MOTIVATION

Due to freshwater scarcity across Australia and the world, membrane technologies have

gained enormous attention for water purification applications such as seawater desalination

and wastewater recycling. Compare to micro, ultra, and nanofiltration, reverse osmosis (RO)

filtrations can achieve high rate of contaminant removal using low energy consumption [2].

Due to this reason, RO membrane filtration has been widely applied for water purification in

recent years [3]. However, membrane fouling and specially biofouling are the major obstacles

hindering the full potential of RO purification processes [4]. Biofouling is defined as the

undesired development of microbial layers on RO membranes [5], and well known by its

adverse effects to the membranes. Studies have reported biofouling is likely to cause the

increase of energy and chemical costs, loss of water production and quality, and eventually

membrane deterioration [6, 7].

To restore the performances of RO membranes, chemical cleaning is regularly required.

Chemical cleanings involve alkali cleaning (i.e. sodium hydroxide) for organics and biofilm

removal, and acid cleaning (i.e. hydrochloric acid, citric acid) for scaling removal. However,

many studies have reported that biofouling cannot be removed effectively using the standard

chemical cleaning method [8-10]. The application of chemical cleaning agents in large

quantities has also caused significant operational costs and environmental issues for their

disposal [11-13].

Free nitrous acid (FNA) has been reported to have a strong biocidal effect on sewer biofilms

and waste activated sludge [14-18]. Studies have reported FNA potentially induce cell death

and biofilm detachment at parts per million levels (0.2 mg HNO2-N/L), and the biocidal

effect of FNA was increased by 43–51% when FNA is combined with hydrogen peroxide

(H2O2) [14, 15]. Based on these studies, it is anticipated that FNA not only can damage the

structure of biofouling layers but also can inactivate the microbes in the biofilm formed on

RO membranes. FNA as acid can hydrolyse organic constituents of biofouling layers such as

proteins and polysaccharides [19], resulting a loose biofilm that may be susceptible to

biocidal attracts. As a biocidal agent, inactivation of bacteria induced by FNA can inhibit the

development or the regrowth of biofilm. Moreover, the synergistic biocidal effect of FNA

and H2O2 has been well demonstrated on sewer biofilms and waste activated sludge [14-18],

2

as aforementioned, H2O2 is a strong oxidant agent that can also cause the death of bacteria.

Therefore, an alternative anti-fouling strategy using FNA with or without H2O2 for RO

membranes was studied. Additionally, it is expected FNA as an acid could potentially remove

scaling from RO membranes. The descaling efficiency of FNA was also investigated.

The potential inactivation and cleaning effects of FNA on RO membrane biofilms and the

application of FNA to replace the conventional two-stage cleaning strategy under cross-flow

conditions have formed the motivation for the work in this thesis.

§ 1.2. GENERAL RESEARCH OBJECTIVES

The main objective of this study is to use FNA with or without H2O2 for RO membrane

biofouling and scaling removal. The specific goals are:

1. To characterise biofouling of RO membrane from different full-scale plants.

2. To design a cleaning protocol under cross-flow conditions at lab-scale.

3. To determine the cleaning effects of FNA with or without H2O2 on different biofouling

matrix.

4. To reveal the optimal cleaning conditions for RO biofouling removal by studying the

impacts of FNA concentration, pH, and H2O2 concentration.

5. To determine the descaling efficiency of FNA in comparison with standard acid cleaning

solutions (i.e. hydrochloric acid and citric acid).

6. To investigate the ability of FNA to prevent the accumulation of fouling on RO

membranes at the bench-scale.

3

2. LITERATURE REVIEW

§ 2.1. INTRODUCTION

In this chapter, literatures about reverse osmosis (RO) membrane filtration and biofouling

have mainly been reviewed. The review includes the principles and functions of RO

membrane filtration, the principles and adverse effects of fouling and biofouling on RO

membrane filtration, prevention and cleaning methods for biofouling in RO processes.

§ 2.2. RO MEMBRANES FILTRATION

Flat sheet RO membranes are commonly packed in the spiral-wound configuration for large-

scale applications. The spiral-wound configuration is considered better than plate and cushion

configurations due to its high component density (high surface area to volume ratio) [20]. In

the spiral-wound configuration, RO membranes usually operate under cross-flow conditions

also known as dynamic filtrations. Figure 2-1 demonstrates the flow pattern of dynamic

filtration and the general configuration of the spiral-wound RO module. During the filtration

process, the feed flows across RO membranes in parallel, and is then separated into two

streams. The purified stream that passes through the membrane is called permeate and the

other stream that contains rejected solutes by RO membranes is called retentate or

concentrate [21, 22]. RO membranes have been widely used to produce purified water in the

seawater desalination, wastewater treatment and recycling plants, since RO membranes

produce high capacity and quality products with less requirement of energy [23].

Figure 2-1. Schematic diagrams of cross-flow pattern (left) and configuration of the spiral-

wound RO module (right) [20].

4

The asymmetric membrane made of cellulose acetate (CA) polymer is the first commercially

available RO membrane, which was developed by Loeb and Sourirajean in late 1950’s [24].

There are some disadvantages limit applications of CA membranes. CA membranes can be

easily degraded at high temperature and alkaline conditions. Therefore, the operating

temperature has to below 30℃ and pH is restricted to 4-6 for CA membranes applications

[22]. CA membranes have a low tolerance to chlorination, as the structure and selectivity of

CA membranes are likely to be degraded when continuously exposed to free chlorine (<1

mg/L) [25]. It has been reported that the cellulose backbone of CA membranes can be

consumed by the organisms in the RO feed water, which also leads to the membrane

degradation [26]. Thin film composite (TFC) membranes are the second generation of RO

membranes, which are composed of an active layer and a porous supportive layer. The active

layer is responsible for the filtration and commonly made of cross-linked aromatic polyamide

and the porous supportive layer is commonly made of polysulfone [27, 28]. TFC membranes

are more popular than CA membranes due to its higher performances in permeate flux

production and salt rejection [23, 28]. TFC membranes can be operated over a wider

temperature and pH range (pH 2-12) than CA membranes [29], hence TFC membranes have

been widely used in full-scale processes [30].

§ 2.3. FOULING ON RO MEMBRANES

Due to the high efficiency and selectivity of RO membranes, rejected solutes and particles

that are not flushed away by the cross-flow are likely to accumulate at membrane surfaces.

This phenomenon is generally referred as membrane fouling which can hinder membrane

performances [31, 32]. RO membrane fouling can be classified into four groups: biofouling,

inorganic fouling or scaling, colloidal fouling and organic fouling [33-39]. The formation and

development of fouling are influenced by feed water properties, hydrodynamic conditions,

temperature and pH [40, 41]. In reality, the performance loss of membranes is likely due to

the combination of different fouling phenomena [42-44]. RO fouling are generally initiated

by concentration polarization (CP) which is caused by the build-up of solutes near the

membrane surface [11, 31]. A cake layer induced by CP is formed by attaching more

microorganisms, colloidal particles, dissolved natural organic matter from feed water [45].

Cake-layers further enhance CP, hydraulic resistances and osmotic pressure near membrane

surface, resulting the decline of membrane performances in term of permeability and salt

rejection [42]. To recovery the performance of RO membranes, extra energy input might be

5

required to raise the operating pressure and to meet the production rate. Cleaning is an

alternative method to restore the capacity of membranes, especially in the case of heavy

fouling. Additional units or processes need to be installed to neutralise the cleaning chemicals

and to avoid the product contamination [12]. However, frequent or inefficient cleaning

normally leads to the replacement of membrane.

Among different types of fouling, biofouling have been considered as one of the most serious

operational problems in RO process and found in 70% of the seawater RO membrane plants

in the Middle East [46]. Literature suggests that effective anti-biofouling methods have not

yet been discovered or implemented yet, so there is an urgent need to define a simple and

effective method for biofouling control in RO applications. Mineral scales (i.e. calcium

carbonate) presented in almost any feedwater are more likely to interact with biofilm and

form a complex fouling network. Hence, the following literature review focus on biofouling

and scaling in RO processes and the respective control methods for biofouling and scaling.

§ 2.3.1. Biofouling

§ 2.3.1.1. Formation

Biofilm is a layer of microorganisms deposited at the membrane surface, and biofouling is

defined as the undesirable growth and proliferation of biofilm [12, 47]. The formation of

biofilm on RO membranes is inevitable, since the membrane filtration process provides water

and a surface for biofilm growth [48]. The feedwater is the source of microorganisms and

nutrients such as organic matters and salt ions which are essential for the growth of

microorganisms [49]. The RO membranes provide the platform for microbes to deposit.

Moreover, biofilm formation is greatly influenced by hydrodynamic operating conditions,

such as flow rate of feed (cross-flow rate) and product (flux) [41]. The effects of cross-flow

velocity (CFV) on biofilm remains unclear, as reports revealed that increasing CFV can either

facilitate or reduce the accumulation of biofilm on RO membranes [50]. It has been reported

that high CFV may pose a strong shear rate to reduce the CP and the rate of biofilm formation

[50, 51], but the other studies reported that high CFV or shear rate may attribute to the

formation of a denser and thinner biofilm [50]. High flux is achieved normally by applying

high pressure, which is likely to accelerate the development of biofilm. Suwarno et al.

revealed severe biofouling is caused by a higher flux rate [51], suggested that the growth rate

of biofilm is proportionally related to the flux rate.

6

A schematic diagram of biofilm development is given in Figure 2-2. A conditioning film

composed of inorganic solutes and organic molecules is formed on the membrane surface

initially. At the second stage, the conditioning film starts to attract free-floating cells from

feedwater [52]. Biofilm is formed with the embedment of microcolonies [22]. As the product

of microcolonies, extracellular polymeric substances (EPS) attach cells irreversibly to the

biofilm [53]. The number of embedded microorganisms has been reported to be 500 to

50,000 times higher than the free-floating cells in feedwater [48]. The biofilm is matured with

additional EPS production, and then is capable for cellular motility and reproduction [54, 55].

The reproduction of microorganisms in biofilm is problematic, as it has been reported that

although 99.9% of biofilm are removed, the remaining microorganisms are still able to

initiate the regrowth of biofilm by consuming the nutrients from feedwater [56].

Figure 2-2. Five stages of biofilm development [52].

§ 2.3.1.2. Structure and composition

The random structure and composition of biofilm is the main obstacle to study

biofilm/biofouling in RO processes, since the structure and composition of biofilms varies

widely depended on the process water and operating parameters [40, 41]. Biofilm is mainly

made up of 70-95% water. Within 5-30% of dry weight, 70-95% of EPS hold microbial cells

[57]. Beside the embedded microorganisms, EPS consist 40-95% of polysaccharides, up to

60 % of proteins and small amount of nucleic acids, lipids and other biopolymers [58].

Extracellular polymeric substances (EPS)

EPS are normally composed of polysaccharides, proteins, nucleic acids and lipids [59]. The

primary functionality of EPS is to facilitate the growth of biofilm by attaching microbes and

7

nutrients from feed water [53]. EPS participate in microbial activities. For instance, EPS store

nutrients and distribute them under starvation conditions [60]. EPS also contribute to the

formation of a complex and robust biofilm. The stability of EPS depends on the

hydrophobicity interaction with RO membranes, the cross-linking network with mineral ions

and the entanglements of the biopolymers within EPS matrix [61]. The complex network of

EPS protects embedded microbes to against shear force and cleaning attacks, and

substantially limits the diffusion of biocides and other disinfectants into the biofilm [62].

Nucleic acids and lipids occupy a small portion of EPS, which are contributed to the

stabilization of the biofilm structure and hydrophobic properties of EPS [55, 63].

Proteins and Polysaccharides

The main biological functions of proteins can be divided into five major classes: metabolism,

biosynthesis, secretion, adaptation and protection [55, 64]. Proteins have been reported to

play an important role in the initial adherence of biofilm onto the surface of membrane [55].

Proteins also serve as enzymatic catalysts for chemical reactions within the cell [65].

Structural proteins form parts of cell walls and ribosomes [65]. If the protein is successfully

removed or denatured, the functionality and structure of biofilm are expected to be damaged.

The main functionality of polysaccharides is glucose storage for energy generation [27].

Structural polysaccharides form parts of cell walls, like the peptidoglycan layer in bacterial

cell walls [27]. As major part of EPS, polysaccharides support the structural stability and

architecture of EPS. It was reported that exopolysaccharide composition induces the

filamentous networks and gel-like structures within biofilm [59, 65]. Polysaccharides also

facilitate the extension of EPS structure. For instance, polysaccharides contribute to the

formation of EPS backbone where other components of EPS can bind [66]. It is anticipated

successful removal of polysaccharides would not only cut-off the energy supply for the

microbial cells within biofilm, but also damage the structure of EPS and entire biofilm.

§ 2.3.2. Scaling

Inorganic fouling or scaling is caused by the precipitation of supersaturated inorganic ions

from feed water [37]. RO membranes are especially at high risk to experience scaling

problem, since a large number of rejected inorganic ions are likely contribute to the

concentration polarization on the membrane surface. When the concentrate stream becomes

supersaturated and exceeds the solubility limit of inorganic ions, dissolved mineral ions start

8

to precipitate and form scales in RO systems [37, 67]. Mineral ions such as calcium are

present in almost any feedwater, which are like to form calcium carbonate, calcium sulphate,

and calcium phosphate scales in high pressure water treatments. Silicate scale is another

common scale in RO systems, since silica is rich in nature [37]. Scale can be formed through

two schemes: surface crystallisation and bulk crystallisation [68, 69]. Surface crystallisation

is induced by the growth of the scale on the membrane surface. Bulk crystallisation is the

deposition of homogeneously formed crystals from the bulk solutions. Bulk and surface

crystallisation can also occur simultaneously [37]. Like the other types of fouling, scaling can

hinder RO membrane performances and cause the decline of flux and salt rejection, loss of

production, membrane deterioration and elevated operating costs.

Figure 2-3. Schematic demonstration of scale formation [37]

Scaling can be effectively controlled by acidification, antiscalant addiction and ion-exchange

softening. Acidification is commonly applied to prevent calcium carbonate (CaCO3) and

calcium phosphate by adjusting the pH of the feed water to 5-7 [70]. Acid addition increases

the solubility of calcium ions and keeps calcium salts soluble in the concentrate stream.

Acidification is also effective for removing scales from membrane surfaces. DOW

recommends to use 0.2 wt% of hydrochloric acid, or alternatively 2.0 wt% of citric acid for

severely CaCO3 fouled RO membranes [71].

Antiscalants are surface active materials that disrupt crystallisation process in three basic

ways: threshold inhibition, crystal modification and dispersion [37]. Threshold inhibition is

one of the functionalities of antiscalants, which enables sparingly salts ions soluble in their

supersaturated solutions without forming any crystals. Antiscalants worked as crystal

modifiers can distort crystal shapes and produce less adherent scales. Antiscalants such as

polyphosphates can interrupt crystal growth by attacking positive charged calcium and

9

magnesium scales at the sub-microscopic level. Antiscalants with high anionic charged

functional groups such as polycarboxylates can cause the dispersion of crystals by imparting

an anionic charge on the crystals, and the separation between the anionic charged crystals and

negatively changed RO membranes. Antiscalants should be applied with caution as overdose

of antiscalants may lead to the precipitation of themselves [72]. Moreover, monitoring the

quantities of antiscalant in the RO is more complicated than monitoring acidification [73]. In

addition, it has been reported the application of polyacrylic acid and phosphonates based

antiscalants for scaling removal could stimulate biofouling in RO systems [74].

Ion-exchange softening commonly use sodium attached resin to exchange with magnesium

and calcium ions based on the reactivity of these metal ions [73]. Consequently, the

concentration of magnesium and calcium ions is declined and so is the formation potential for

their associated scales. Brine solutions are required to regenerate the resin, which is a

disadvantage of this method since it is problematic to disposal brine regenerate [73]. In order

to decide the adequate decaling method among aforementioned methods, characterisation of

feed water, compatibility of the RO elements with chemical additives and costs associated

with the cleaning all have to take into consideration [37].

§ 2.4. RO MEMBRANES BIOFOULING PREVENTION AND CLEANING

METHODS

Biofouling control for RO membranes has generally been performed in two stages.

Stage one: Prevention is carried out to alleviate the formation of biofouling when the

capacity of RO membranes is yet to be influenced by biofilm. During the downtime of

RO processes, prevention is occasionally required to preserve the properties of RO

membranes [56].

Stages two: Cleaning is required to restore the performance loss of membranes caused

by biofouling [75, 76].

§ 2.4.1. Prevention Methods

In the past, prevention methods focused on removing nutrients and inactivating

microorganisms in the upstream of RO processes. Membrane pre-treatment is one of common

methods used to prevent the biofilm [12, 56]. By comparing to the other pre-treatment

technologies such as coagulation and flocculation, micro- and ultra-filtrations (MF and UF

10

respectively) are more efficient in bacteria, colloidal particles and nutrients removal with less

space requirements and chemical usages [12, 56]. However, membranes used in the pre-

treatment also have to face fouling and performance loss issues. The leakage of unwanted

fouling materials from pre-treatment units can cause fouling in RO units. It has been reported

fouling is likely to occur in RO units if MF and UF fail to reject sub-micro (> 1 𝜇m) materials

[56, 77].

Biocidal disinfection is commonly used to sterilise microorganisms in the upstream of RO

processes [12, 56]. Free chlorine, a strong oxidant, has been reported to be the most effective

biocidal agent [12, 56, 78], as it was reported that process water containing 0.04-0.05 mg/L

free chlorine can effectively prevent biofilm [79]. However, the application of free chlorine

has been limited due to its potential negative effect on membrane integrity. Many studies

reported that free chlorine can damage the active filtration layers of polyamide composite RO

membranes, and eventually lead to membrane degradation and replacement [80-83]. In order

to protect the integrity of polyamide composite RO membranes, process water containing free

chlorine should always be dechlorinated before entering the RO instalments. However,

chlorination and dechlorination using sodium bisulfite (SBS) have been reported to

occasionally enhance biofouling [56]. Alternative biocides such as chloramines,

monochloramine and chlorine dioxide have been applied to replace free chlorine [84-86].

Although these biocides have less effects on membrane, all of them have been found to be

less efficient at disinfecting than free chlorine [80].

Recent research has focused on membrane surface modification in order to produce chlorine-

resistance or low fouling membranes. Studies reported that bacteria are less likely to deposit

on the hydrophilic, negatively charged and smooth membrane surfaces [12, 56]. However,

membrane surfaces with these modification might attract hydrophilic and positively charged

fouling materials [12]. Although studies have suggested that low fouling membrane might

reduce the deposition of fouling materials [46], there is no guarantee that low fouling

membrane can completely prevent biofouling. Chemical cleaning is commonly required once

biofouling formed on RO membranes.

§ 2.4.2. Cleaning Methods

Membrane cleaning methods can be generally divided into two groups: physical and chemical

11

cleaning. Physical cleaning methods such as forward and reverse flushing, sponge ball

cleaning, permeate back pressure, vibration and etc. have been widely applied to clean hollow

fibre or tubular RO modules [87]. However, physical cleaning methods are not suitable for

spiral-wound RO modules. Chemicals are commonly employed to clean spiral-wound RO

modules. During cleaning processes, chemical agents are used to disperse biofilm by

disrupting the bonds between foulants, and between foulants and membrane surfaces [56]. A

shear force created by the cross-flow is normally applied along with chemical agents to

remove the detached fouling materials. There are five groups of chemicals commonly used to

clean RO membranes, as given in Table 2-1 [19].

Table 2-1: Major categories of membrane cleaning chemicals [19]

Category Functions

Caustic Hydrolysis, solubilisation

Biocides Oxidation, disinfection

Acid Solubilisation

Chelating Agents Chelation

Surfactants Dispersion, surface conditioning

Caustic solutions (i.e. sodium hydroxide, NaOH) are commonly used to clean organic and

biological fouled RO membranes through hydrolysis and solubilisation [19]. Caustic

solutions can potentially disrupt biofouling by hydrolysing its major organic constituents

such as polysaccharides and proteins. Moreover, caustic solutions not only can transform fats

and oils into water-soluble micelles, but also can increase the solubility of organic molecules

such as phenolic functional group [19]. Addition of NaOH creates the electrostatic repulsion

between negatively charged organic matters and membranes [88, 89], resulting a loose

fouling layer or even the dispersion of the fouling layers. Many studies have combined NaOH

with less harmful biocides such as 2,2-dibromo-3-nitrilopropionamide (DBNPA) [90],

chelating agents (e. g. ethylenediaminetetraacetic acid, EDTA) or detergents (e. g. sodium

dodecyl sulfate, SDS) to improve its cleaning efficiency [36, 91]. However, no evidence of

completely biofouling removal was found.

The cleaning effects of weak biocidal agents such as chloramines, monochloramine and

chlorine dioxide on biofouling have been widely studied. Weak biocides as oxidising agents

12

can reduce the adhesive forces between fouling materials (i.e. organic polymers) and

membranes [19]. As mentioned in the prevention methods, biocidal agents must applied with

caution due to their oxidising effect on the active filtration layers of RO membranes [12, 56,

90]. Hydrogen peroxide (H2O2) is another biocide commonly used for membrane disinfection

and storage [40]. H2O2 formed from hydroxyl free radicals causes the death of

microorganisms by breaking the cell wall of the microorganisms [12]. A non-oxidising

biocide, DBNPA (2, 2-dibromo-3-nitrilopropionamide) is efficient at inhibiting a large range

of aerobic and anaerobic bacteria [90]. Since DBNPA is not a non-oxidising agent, it does not

affect the membrane surface as chlorine. However, DBNPA has showed same shortcomings

as the other weak biocides, as DBNPA has demonstrated inefficient inhibition against algae

and fungi, and it is unstable in the solution at pH higher than 8 [12, 92].

Acids such as hydrochloric acid, citric acid and sulphuric acid are commonly applied to clean

scaling rather than biofouling, but the organic constituents (such as proteins and

polysaccharides) of the fouling layers can be hydrolysed by acid [19]. Acids and the chelating

agent EDTA are able to dissolve or chelate the divalent ions in the fouling layer, respectively

[19], which might result a less dense and adhesive fouling layer. Surfactant (i.e. SDS) can

interfere with fouling layers in three ways [19]: first, surfactants can raise the solubility of

fouling materials such as fat, oil and proteins by changing the hydrophobicity of these

materials, resulting the dispersion of these fouling materials. Second, surfactants are likely to

impede the adhesive force between bacteria and membranes. Lastly, the cell walls of bacteria

can be damaged by surfactants. There is a chance however those surfactants might bind to

fouling material that has an affinity to them and in turn enhance fouling. Hence, surfactants

such as SDS have been applied in combination with NaOH to ensure and improve the

cleaning efficiency.

Recently, biochemical enzymes have been examined aiming at cell lysis and biofilm

dispersion [93, 94]. For the purpose to destroy the structure of biofilm, polysaccharides lyases

and proteases have been invented to cleave the structure of polysaccharides and proteins,

respectively [12, 93, 94]. However, this method cannot provide a comprehensive control for

biofouling, since the enzymes are designed specifically to target on polysaccharides and

proteins [13, 95]. In addition, the production of biochemical enzymes is a costly process, and

degrading enzymes cannot be reuse after cleaning [12].

13

Overall, literature review reveals that the current prevention and cleaning methods for

biofouling are not efficient enough. None of the novel techniques have proved dramatically

improvement for biofouling control, and it is difficult to implement these techniques based on

the shortcomings listed above. Therefore, there is still a need to develop efficient methods for

biofouling control. In this study, the ability of free nitrous acid (FNA) to clean biofouling was

investigated.

§ 2.5. FREE NITROUS ACID (FNA)

§ 2.5.1. FNA and Its Biocidal Effect

FNA is the protonated form of nitrite, it is a weak and monobasic acid (pKa=3.34, 25℃)

which only presents in solution [96]. The concentration of FNA in solution is calculated

based on the following equation which is extracted from [97]: FNA = NO2--N / (Ka x 10

pH),

where Ka is the ionization constant of the nitrous acid (Ka=e-2300/(T+273)) and T is the

temperature (℃). Over the past 40 years, FNA has shown its strong biocidal effect on various

types of microorganisms, as summarised in Table 2-2.

Table 2-2: Studies of biocidal effects of FNA on microorganisms.

Year Remarks Reference

1976 FNA rather than nitrite inhibits nitrification [98]

2006 FNA inhibition on the metabolism of nitrifying

organisms in the nitrification

[99, 100]

2010 FNA inhibition on the aerobic metabolism of poly-

phosphate accumulating organisms (PAOs)

[16]

2010 FNA inhibition on the aerobic metabolism of glycogen

accumulating organisms (GAOs); on the anaerobic

metabolism of PAOs and GAOs.

[17, 101]

2012 FNA inhibition on aerobic and anoxic metabolism of

PAOs.

[102]

2012c FNA treatment improves the biodegradability of

secondary sludge

[103]

2013 FNA (with hydrogen peroxide) has a strong biocidal

effect on microbes in anaerobic sewer biofilms

[14, 15]

2014 FNA treatment can inactivate nitrite oxidizing bacteria

(NOB) and ammonium oxidizing bacteria (AOB).

[104]

2015 FNA effectively deactivates sulfide and sulfur oxidizing

bacteria (SOB) in the sewer corrosion layer.

[105]

14

Many studies have reported that the strong biocidal effect of FNA is likely derived from FNA

and its reactive nitrogen species (RNS) such as nitric oxide (NO), nitric dioxide (NO2) and

peroxynitrite [14, 106, 107].

2[HNO2 (aq) ↔ H+ + NO2

-] Equation 1

2HNO2 (aq) ↔ N2O3 (aq) + H2O (l) ↔ NO (aq) + NO2 (aq) + H2O (l) Equation 2

•NO + •O2- ↔ ONOO

- Equation 3

HNO2 (aq)+ H2O2 (aq) ↔ ONOO- + H3O

+ Equation 4

Zhou et al. [107] revealed that FNA hinders ATP synthesis by acting as an uncoupler agent,

however they acknowledged that this inhibition effect does not occur in every organism. NO

is one of intermediate products of the HNO2 reaction (Equation 2). NO has been reported to

react with heme and metal centres of proteins [107]. The product of this reaction, metal-

nitrosyl complexes (e.g. Fe-NO-R), can destruct the catalytic site of the enzymes, and then

inhibits electron transport and ATP generation [107]. NO also interferes with the oxygen

respiration and hence inhibits oxygen uptake rate for the cells [108]. The other two RNS of

FNA, NO2 and peroxynitrite have also been reported to play important roles during the

biocidal processes. NO2 can induce lipid peroxidation, resulting in cell membrane damage

[109]. Peroxynitrite (ONOO-) can oxidise protein, DNA and lipids, and lead to the death of

microbial cells [110]. Moreover, recent studies have revealed that FNA is able to break the

structure of biofilm extracted from waste activated sludge. Du et al. reported that FNA can

break the bond between organic materials and metals (cooper and zinc) and the structure of

EPS [111]. Zhang et al. also demonstrated the breakdown of macromolecules such as

proteins in EPS under the impact of FNA [18]. Based on the inhibition mechanisms of FNA

and its reactive nitrogen species (RNS) on microorganisms and biofilm structure, it is

anticipated that FNA can be effective for biofouling control. However, disinfection using

FNA involves risks, as it reported that sodium nitrate can react with organic substances to

produce carcinogenic nitrosamines under acidic conditions [112]. FNA is generally prepared

by mixing sodium nitrite and hydrochloric acid. Hence, using FNA at low pH level may lead

to the formation of disinfection by-products (DBPs) such as carcinogenic nitrosamines in the

RO systems. The formation of nitrosamines is specifically not desired in water treatment, as

human exposure to these may have serious health effects such as cancer [113, 114]. Hence,

15

the formation potential of carcinogenic nitrosamines should be investigated before applying

FNA at real plants.

As previously mentioned in the cleaning method, H2O2 is a strong oxidant that can cause the

death of microorganisms. The synergistic biocidal effect of FNA and H2O2 has been

demonstrated in the anaerobic wastewater system [15]. Jiang et al reported that applying

H2O2 can enhance the biocidal efficiency of FNA on anaerobic sewer biofilm from 90% to 99%

[15]. Many studies have reported that H2O2 can lethally effect on bacteria by oxidising

proteins, DNA and bacterial cell membranes [15, 115]. Hence, it is postulated that combining

H2O2 with FNA would enhance the effect of FNA on biofouling.

In addition, FNA as an acid should be capable of removing inorganic scaling from RO

membranes as commercial acids, such as hydrochloric acid (HCl) and citric acid. This could

allow one-stage cleaning for biofouling and scaling, reduce the requirement and hence cost of

two stages cleaning: alkaline cleaning for organics and biofouling removal and acid cleaning

for scaling removal [116].

§ 2.5.2. FNA versus Other Biocides

In addition to the strong biocidal effects of FNA on microorganisms, it was anticipated

applying FNA has more advantages than applying other biocides. The first advantage was

that free nitrous acid will not damage the active filtration layers of RO membrane as other

biocides do. The ageing effect of FNA on TFC polyamide membranes were studied in our

research team. Results of ageing tests showed that polyamide membranes are compatible with

FNA. In comparison, application of free chlorine is generally not recommended to protect the

integrity of polyamide composite RO membranes as aforementioned (§ 2.4.1 Prevention

Methods). Though alternative biocides such as monochloramine has less effect on polyamide

composite membrane in compare to free chlorine, the efficiency of monochloramine in

disinfection was found to be reduced [80]. By applying FNA, there is no need to evaluate the

tradeoff between the effect of FNA on membrane and its efficiency in disinfection.

Moreover, the residual control for the new technique using FNA was anticipated to be more

convenient. FNA is prepared from the commonly available sodium nitrite and hydrochloric

acid, and can be simply discharged after dilutions, as the residues of FNA solutions were

16

reported to be biodegradable. Hence, its disposal after dilutions will not cause environmental

problems. In comparison, monochloramine is formed by mixing aqueous ammonia and

sodium hypochlorite. As a consequence, a large amount of ammonia is generally required to

balance monochloramine residual in the reverse osmosis concentration stream, which is

unwanted for the environment. For example, this is under strictly control at Luggage Point

Luggage point wastewater treatment plant (Brisbane, Australia) [117].

Another benefit of using FNA is it works as an acid which can remove scaling as well. If

FNA could remove biofouling and scaling simultaneously, it not only benefits the cleaning

process but also reduces the costs associated with conventional two-step cleaning. In

comparison, common biocides such as chloramines, monochloramine and chlorine dioxide

can only be used for biofouling mitigation. Therefore, new cleaning method using FNA was

proposed for RO membranes in this study.

§ 2.6. RESEARCH GAP

The literature review has shown the principles of RO membrane filtration, the adverse effects

of fouling and biofouling on RO membrane filtration, prevention and cleaning methods for

biofouling in RO processes. Due to the highly varied features of biofouling in RO processes,

no adequate method can be applied among conventional and novel cleaning methods.

Therefore, the major objective of this study is to examine the efficiency of an alternative

chemical agent named FNA for RO membrane biofouling control. Additionally, RO

membrane fouled at different full-scale plants were characterised and the preliminary study

on the effect of FNA for fouling prevention in RO processes were carried out. The specific

approaches are given as following.

Objective 1: To characterise RO membranes biofouling fouled at different full-scale plants.

(Chapter 3)

Five RO membranes fouled at different full-scale plants were used in this study, the degree

and composition of different biofouling matrices were anticipated to be highly varied. As

reviewed in section 2.3.1, biofouling is mainly composed of microorganisms, proteins,

polysaccharides plus the other organic and inorganic materials. The purpose of this objective

was to reveal the characteristics of five different fouling matrices before undergoing FNA

cleaning. Different fouling matrices were characterised by quantifying the major constituents

17

such as bacterial cells, inorganic ions, proteins and polysaccharides in the biofouling layers.

In addition, the live and dead bacterial cells of biofouling layer were measured in order to

investigate the biocidal effects of FNA on RO biofouling.

Objective 2: To use FNA with or without H2O2 in a single stage for RO membrane

biofouling and scaling removal. (Chapter 4)

A cleaning protocol using cross-flow conditions was designed for the cleaning tests. As

previously reviewed, the shear forces generated by the cross-flow might be potentially

facilitate biofouling cleaning by removing fouling materials detached from biofouling layers

and promoting the diffusion of cleaning solutions. Hence, RO membranes were cleaned by

FNA cleaning solutions and NaOH (pH 11) at a cross-flow rate of 0.1 m/s for 24 hours. As

the main objective of this study, the efficiency of FNA with or without H2O2 for membrane

biofouling control was reflected by the reduction of fouling materials (bacterial cells, viable

cells, proteins and polysaccharides) in the biofouling layers after cleaning tests.

In addition, the effect of FNA on the scale deposited on RO membrane was tested using the

cleaning protocol developed for biofouling control. The descaling efficiency of FNA was

assessed along with standard acid cleaning solutions (i.e. hydrochloric acid and citric acid),

all FNA and standard acid cleaning solutions were adjusted pH at 2 and 3 for the cleaning

tests, as these pH levels were recommended by the membrane manufacturer [116]. The

descaling efficiency was determined by evaluating the concentration of dissolved salt ions

that have been removed by each cleaning solution.

Objective 3: To investigate the ability of FNA to prevent the accumulation of fouling on

RO membranes at bench-scale. (Chapter 5)

Biofouling control in RO membrane processes is generally divided into two groups,

prevention and cleaning, as aforementioned, but the ideal condition for biofouling control is

to prevent the initial adsorption of fouling materials on RO membranes. Thus, the efficiency

of FNA for RO membrane biofouling prevention was preliminarily investigated in this study.

Two bench-scale crossflow RO filtration units were applied to simulate fouling accumulation

processes, except only one of filtration units was exposed to FNA cleaning for 6 hours on a

weekly base. As previously reviewed, fouling is the main reason that causes the performances

loss of RO membranes, and the development of fouling can be reflected by the loss of

permeability and salt rejection in RO processes. Hence, the hydraulic performances of RO

18

membranes filtrations with or without FNA treatment have been continuously monitored for

up to three weeks, in order to justify the ability of FNA to prevent fouling on RO membranes.

After the filtration tests, the effect of FNA on fouling layers was revealed by comparing the

characteristics of fouling material deposited on the RO membranes with or without FNA

treatments.

19

3. RO MEMBRANE FOULING CHARACTERISATION

§ 3.1. INTRODUCTION

The fouling is notorious to the membrane process, as it hinders the process to perform their

full capacity [4]. When the performance decline cannot be restored due to irreversible fouling,

membrane autopsy can be conducted to discover the causes. In this study, membrane

autopsies were performed to characterise membrane foulants prior cleaning trials. Six fouled

RO membranes from different full-scale plants were used.

Membrane autopsy involves visual inspection, fouling characterisations and hydraulic

performance of RO membranes. Fouling was characterised in three fractions: biological,

organics and inorganics. The biological and organics fractions were determined via biological

and molecular analyses. In these analyses, the quantities of biomass were measured to reveal

the fraction of biofouling. The organic fraction was determined by measuring protein,

polysaccharides and volatile solid. The proportion of inorganic material was revealed via

elemental analysis. Based on the results of membrane autopsy, fouling characterisations not

only revealed the composition of fouling layers, but also the degree of fouling (the quantities

of foulant deposits) on RO membranes. The hydraulic performance was determined in terms

of permeability and salt rejection. Sample preparation and the procedures of each analysis are

given in this chapter.

Results reported in this chapter formed part of the following submitted paper

E. Filloux, J.Wang, M. Pidou, W. Gernjak, Z. Yuan. 2015. Biofouling and scaling control of

reverse osmosis membrane using one-step cleaning - potential of acidified nitrite solution as

an agent.

20

§ 3.2. MATERIALS AND METHODS

§ 3.2.1. Membranes

Membrane autopsies have been done on six fouled RO membranes. All RO membranes are

commercial thin-film composite polyamide membranes, which were collected from full-scale

plants (Table 3-1) and stored in the cold room at 4℃ until membrane autopsy took place. In

membrane autopsy, biological fouled RO modules (RO1-5) were undergone all the analyses

listed in section 3.3 Membrane Autopsy. The RO6 was expected to be mainly fouled with

calcium carbonate scales and characterised by the elementary analysis and hydraulic tests

only.

Table 3-1: RO membranes used in this study.

Membrane # Source Fouling Type Membrane

Autopsy

Date

RO1

RO2 Industrial wastewater recycling plant Biofouling

2013/10/16

2014/01/13

RO3 Water reclamation plant Biofouling 2014/05/07

RO4 Water reclamation plant Biofouling 2014/08/13

RO5 Seawater desalination plant Biofouling 2014/10/30

RO6 Coal seam gas water recycling plant Scaling 2014/07/09

§ 3.2.2. Sample Preparation

Foulant samples were collected in two ways: (1) the in situ method, which membrane

coupons with foulant attachments were cut directly from RO modules, (2) the destructive

extraction methods, which foulant was physically scraped or brushed off the membranes [91].

Size of membrane coupons and extra preparation procedures are varying depending on the

limit of detection of each analysis. For microscopy based analyses, in situ biofilm samples on

RO membrane coupons were prepared. In order to conduct comprehensive membrane

autopsy, five biomass samples were collected from different location of RO modules for each

analysis.

§ 3.3. MEMBRANE AUTOPSY

§ 3.3.1. Biological Analysis

21

§ 3.3.1.1. Adenosine tri-phosphate (ATP)

ATP is an energy-rich compound present in all living microorganisms [118]. The method for

analysing ATP was adapted from Hammes et al. [119]. ATP was determined using the

BacTiter-GloTM reagent (Promega Corporation, USA). For ATP measurements, biomass

samples were prepared from a membrane coupon (2x2.5 cm2) using the destructive extraction

method. A set volume of biomass sample (300 µL) was mixed with 50 µL of the reagent in

the 96 well plate (Greiner Bio-One, Germany). The luminescence response was then

measured at 38℃ after 20s orbital shaking by a DTX 880 multiplate reader (Beckman coulter,

USA). Samples were prepared in triplicate with at least three controls. MilliQ water (control)

and standard controls prepared from pure rATP (Promega Corporation, USA) were measured

for each batch of sample. Data were acquired as relative light units (RLU) and then converted

to ATP concentrations (nM) using a calibration curve made from pure rATP standard, as

given in the Appendix Figure A-1.

§ 3.3.1.2. Fluorescence in situ hybridization (FISH)

FISH can hybridise targeted microbes with fluorescently labelled oligonucleotide probes and

was applied to provide visual examination and quantification for the targeted microbes [120,

121]. FISH were performed as described in Amann et al. [122]. 0.5 mL biomass was fixed in

2 volumes of 4% paraformaldehyde (PFA) for 2 h at 4℃. Fixed biomasses were washed with

phosphate-buffered saline (PBS). Fixed cells were then well suspended in the mixture of 100%

ethanol-PBS (1:1) and stored in the freezer (-30℃) until the hybridisation.

Approximately 3 μL of fixed cell suspension was applied to each well of the glass slide

(coated with 0.1% gelatin solution). After air-drying, the slide was dehydrated in an ethanol

series of 50, 80 and 98% ethanol (3 min each). Hybridisation buffer (2 ml) containing 30%

formamide. 9uL of hybridisation buffer was added to each well, followed by the addition of 1

μL fluorescently labelled probes (0.5ng/μL). Since proteobacteria and archaea has been found

in many reverse osmosis application plants for water treatment [123-129], FISH probes

targeting major groups of proteobacteria were applied as well as general bacteria and archaea

in this study (Table 3-2).

22

Table 3-2: Oligonucleotide probes used in FISH analysis

Specificity FISH probe-

fluorochrome

Probe Sequence (5’-3’)

References

Bacteria EUB338c-Cy5

d GCTGCCTCCCGTAGGAGT [120]

EUB338+c-Cy5

d GCWGCCACCCGTAGGTGT [120]

Alphaproteobacteria ALF1b-Cy3d CGTTCG(C/T)TCTGAGCCAG [130]

Betaproteobacteria BET42a-FITCd GCCTTCCCACTTCGTTT [130]

Gammaproteobacteria GAM42a-FITCd GCCTTCCCACATCGTTT [130]

Most Archaea ARC915-FITCd GTGCTCCCCCGCCAATTCCT [130]

a Base on Escherichia. coli rRNA numbering [131].

b Percentage formamide in the hybridization buffer.

c EUB338 and EUB338+ are used in a mixture called EUBMIX.

d Fluorescein isothiocyanate (FITC), cyanine (Cy) 3 and 5 are green, red and blue labelled

FISH probes, respectively.

The slide was incubated at 46℃ for 2 h and then washed for 15 min at 48℃. Extra salts (from

buffers) were removed by dipping the slide in the cold water (4℃) for 2–3 s. After air-dried,

the slide was mounted with DABCO (1,4-Diazabicyclo[2.2.2]octane, Sigma-Aldrich) and

viewed under the Zeiss LSM510 confocal laser scanning microscope (CLSM) (School of

Chemistry and Molecular Biosciences at UQ). 20 ± 5 images were randomly taken for each

sample using the CLSM equipped with a Krypton–Argon laser (488 nm) and two He–Ne

lasers (543 and 633 nm). The images were imported to DAIME (Centre for Organismal

Systems Biology, Austria) for biovolume fraction analysis.

§ 3.3.2. Organic and Molecular Analysis

§ 3.3.2.1. Loss on ignition (LOI)

LOI is a gravimetric method for measuring organic and inorganic fractions in biofouling

layers [132].The fouling samples were collected from a known surface area of membrane

(40×40 cm2) using destructive extraction methods. The samples were transferred to crucibles,

23

and subsequently dried in an oven at 105°C overnight and a furnace at 500°C for another 3~4

hours. The total and volatile solids concentrations were then calculated as follows:

Total solids concentration(g/m2) : 𝑇𝑆 =

𝑤2 − 𝑤1

𝑆 Equation 5

Volatile solids concentration (g/m2): 𝑉𝑆 =

𝑤2 − 𝑤3

𝑆 Equation 6

Organic fraction (%): 𝑂𝐹 = 𝑉𝑆 𝑇𝑆⁄ × 100 Equation 7

Where w1 is the weight of the crucible, w2 is the weight of the crucible and deposits after

oven, w3 is the weight of the crucible and deposits after furnace and S is the surface area of

membrane coupons.

§ 3.3.2.2. Polysaccharide quantification with phenol-sulfuric acid method

Phenol-Sulfuric Acid method, also known as DuBois’s method was used to determine

polysaccharides in aqueous solutions [133]. Biomass samples were collected through the

destructive extraction method, 1mL aliquot of biomass samples (n=2) was mixed with 1 mL

of phenol 5 % (w/v) (Sigma-Aldrich, reagent plus >99 %). Samples were mixed carefully and

incubated for 10 min. 5 mL of sulfuric acid 98 % (Merck, 98 %) were added to the mixture

and incubated for 30 min. Mixture samples and blank solution were then analysed with a UV

Spectrophotometer (Varian Cary 50 UV-Visible) at wavelength 490 nm. The signals were

calibrated with glucose, and the amount of polysaccharides was represented as g glucose per

unit area.

§ 3.3.2.3. Protein quantification with the BCA method

The BCA-method was used to quantify protein in solutions, which has been introduced by

Smith et al. in 1985 [134]. A BCA Assay Kit (Sigma, USA) contains three regents:

QA (alkaline buffer), a 250 mL of mixture of sodium carbonate, sodium tartate, and

sodium bicarbonate in 0.2 M NaOH, pH 11.25.

QB (BCA solution) contains 250mL of a 4% (w/v) bicinchoninic acid solution, pH 8.5.

QC contains 12mL of a 4% (w/v) copper (II) sulfate, pentahydrate solution.

Triplicate biomass samples were prepared through the destructive extraction method, 2 mL

aliquot of sample was mixed with 2 mL protein reagents in a ratio of 100 QA: 100 QB : 4 QC.

The mixture samples and then transferred into a preheated oven (37℃ , Thermo Fisher

Scientific Australia) for 2 hours. The samples and blank solution were then immediately

24

measured by the UV Spectrophotometer at wavelength 562 nm. The signals were calibrated

with bovine serum albumin (BSA), and the amount of proteins was reported as g BSA per

unit area.

§ 3.3.3. Elementary Analysis

§ 3.3.3.1. Inductively coupled plasma-optical emission spectrometry (ICP-OES)

Vista-PRO simultaneous ICP-OES (Varian Inc.) was used to measure the metal contents in

fouling layers [106]. A set area of fouled membrane coupon was cut and placed in a 50 mL

yellow capped tube with 20 mL of 10% nitric acid solution. Samples were then recovered

after 1-2 min vortex and kept at room temperature until the digestion process. Aggregate

compounds were further broken and dissolved during the digestion process. Final filtered

samples were used for the ICP-OES measurements and reported.

§ 3.3.3.2. Scanning electron microscopy-energy dispersive spectroscopy (SEM-EDS)

A Philips XL30 scanning electron microscope (secondary electrons) was used to directly

visualise the foulant from the membrane surface. Samples of approximately 3x10 mm were

cut from the RO modules and dried in a desiccator for at least 24h. Before the analysis,

samples were placed on metal stubs and then Pt coated. Typical imaging has been done at

500x magnification rates, 15 kV accelerating voltage. Energy Dispersive X-ray Spectroscopy

was performed on the previously SEM imaged areas with a typical working distance of 10.3

mm. Elemental percentage has been obtained after background subtraction and ZAF

optimization.

§ 3.3.4. RO Membrane Hydraulic Performances

Hydraulic performances of RO membranes were evaluated by measuring both water

permeability and salt rejection. Hydraulic performance tests were carried out in a laboratory

scale cross-flow filtration unit (Figure 3-1). The unit is composed of a 15 L feed tank, a

diaphragm pump (Metering pump Z series, Tacmica), a dampener (PD36, Neptune), two

filtration cells (CF042, Sterlitech, Figure 3-1) and two needle valves.

The pressure and flow rate of this system required constant monitoring during the

measurements. Pressure of the system was indicated by a sensor (Cerabar M, Endress &

25

Hauser) located in the feed line just before the filtration cells. A flow meter is fitted in bypass

line indicating the total flow rate of two inlet streams in the filtration units. The system was

operated in batch mode and consequently both permeate and concentrate were recirculated

back into the feed tank. Once the system reaches steady state, an operating pressure of 5 bars

and total flow rate of 80 L/h could be applied based on the overall setting. The following

parameters can be calculated by recording the permeate flow rate, temperature, pressure and

the conductivity of salt feed and permeate streams.

𝐽 =𝑄𝑃𝑆

Equation 8

where 𝐽: Flux [𝐿

𝑚2×ℎ], Qp: Permeate flow rate [

𝐿

ℎ], S: Membrane surface Area [𝑚2]

𝐾 =𝐽

∆𝑃

Equation 9

where K: Permeability [𝐿

𝑚2×ℎ×𝑏𝑎𝑟], 𝐽: Flux [

𝐿

𝑚2×ℎ], ∆𝑃: Transmembrane pressure [bar]

𝐾𝑤 =𝐾

𝑒𝛽×(1

273.15+𝑇−

1298.15

)

Equation 10

where Kw: Temperature corrected permeability [𝐿

𝑚2×ℎ×𝑏𝑎𝑟, 25℃], K: Permeability [

𝐿

𝑚2×ℎ×𝑏𝑎𝑟], T:

Temperature[℃] 𝛽: Membrane coefficient

Salt Rejection = (1 −𝐶𝑝

𝐶𝑓) × 100

Equation 11

where 𝐶𝑝: permeate conductivity, 𝐶𝑓: feed conductivity

The pure water permeability was monitored for 1-2 hours and was recorded every 30 minutes.

For salt rejection tests, the DI water was subsequently replaced by a 1500 mg/L sodium

chloride (NaCl) solution and operational data points were recorded at same interval for 1-2

hours.

Figure 3-1. Cross flow filtration unit.

26

§ 3.4. RESULTS AND DISCUSSION

§ 3.4.1. Visual Inspection

In Figure 3-2, brown viscous deposits were observed on RO1 and RO2. The color of fouling

layers was darker on RO2 than RO1, and so did the density of the fouling layers. The

observation implied that concentration of foulant on RO2 was higher than that on RO1.

Brown silky deposits were spotted on RO3, which were not as viscous as fouling deposited

on RO1 and RO2. Deposits were spread across these membranes’ surfaces. However, the

abundance of fouling has decreased from the feed side over to the concentration side of the

membranes. Smell could be noticed when unfold these membranes. Dark brown and white

gelatin-like deposits were observed on RO4 and RO5, respectively. Based on the visual

inspection, it appeared the abundance of deposits on RO4 and RO5 were less than that on RO

1 to RO3. RO6 was covered with hard, dry, dark brown fouling deposits.

RO1

RO2

RO3

Figure 3-2. Visual inspection on fouled membranes (RO1-6).

27

RO4

RO5

RO6

Figure 3-2 (continued). Visual inspection on fouled membranes (RO1-6).

§ 3.4.2. Biofouling Characterisation

§ 3.4.2.1. Biomass quantification

The biomass deposited on RO membranes was measured in term of ATP (Figure 3-3).

Biomass was detected on all membranes, and the biomass load from high to low is following

this order: RO2 (4234±1543 pgATP/cm²)>RO1 (950±288 pgATP/cm2)>RO3 (919±477

pgATP/cm2)>RO5 (339±159 pgATP/cm²)>RO4 (204±153 pgATP/cm²). In membrane

systems, fouling layer is considered to have high biomass contents when the ATP

measurement exceeds 1000 pg/cm² [135, 136]. Hence, the results suggested that RO

membranes (RO1 to RO3) were severely fouled with biomass. In agreement to the visual

observation, the results suggested that RO2 was most fouled membrane, and RO4 and RO5

were least fouled membranes.

28

Figure 3-3. Biomass (ATP) contents in the fouling layers of each RO membranes. The error

bars show the standard errors of the replicate samples (n=5).

§ 3.4.2.2. Microbial community of biofouling layers

Based on FISH images (n=20±5), the relative quantities of each microbe and the total

community (all bacteria and archaea) were estimated using DAIME (Appendix Table B-1).

The representative FISH images are given in the Appendix Figure B-1. Figure 3-4 presents

the distribution of each microbe on different RO membranes. The Alphaproteobacteria,

Betaproteobacteria and archaea presented in all biofouling matrixes. More than 60% of

communities were identified by the FISH probes for RO1 and RO5. Alphaproteobacteria

(43%) dominated in RO1 with significant amount of Betaproteobacteria (21%), while the

opposite observed in RO3. Gammaproteobacteria emerged in RO4 and RO5. Within the 25%

detection range, 17% of Gammaproteobacteria was the dominant group in RO4. Community

in RO5 were even distributed within 4 groups with slightly higher abundance of Archaea

(28%).

The results showed that the quantities of detected microbial communities are not related to

the degree of biofouling on membranes, since more than 60% of microbial communities were

identified for both heavily fouled membrane RO1 and moderately fouled membrane RO5.

29

Although proteobacteria and archaea has been mainly found in many reverse osmosis

application plants, proportion of each microbe depends on RO feed water quality, operation

conditions, and sampling location [123-129]. It explains the substantial difference in

distribution and abundances of each microbe in different biofouling matrixes. Less than half

of microbial communities have not been defined by chosen FISH for RO3 and RO4. The

results suggested that more comprehensive FISH analysis is needed by introducing new FISH

probes other than Alphaproteobacteria, Betaproteobacteria, Gammaproteobacteria and

archaea.

(a)

(b)

(c)

(d)

Figure 3-4. The abundances of proteobacteria (Alpha, Beta, and Gamma) and archaea in the

biofouling layers of membrane RO1 (a), RO3 (b), RO4 (c) and RO5 (d).

§ 3.4.2.3. Loss on ignition (LOI)

The organic fractions of the fouling layers were measured by LOI. LOI estimated the

percentage of volatile solids in the total solids measured on membranes. Figure 3-5 reveals

there was more than 88% of organic material deposited in all fouling matrixes. It has been

30

reported that membranes are likely biofouled if the fouling layers contain a large quantity of

microorganisms and more than 70% of organics substances [57, 137]. Hence, the LOI results,

along with the ATP measurement, confirmed the presence of large amount of organic matters

on the surface of membranes, suggesting these membranes (RO1-5) were covered by a

mixture of organic and biological fouling.

In accordance to biomass quantification, the total solid contents of RO1 to RO3 (1.9±0.1 to

4.0±0.1 g/m2) was more than two times of RO4 and RO5 membranes (0.8±0.1 and 0.5±0.2

g/m2 respectively). The quantities of biomass and total solids have suggested that RO1-RO3

were heavily biofouled, RO4 and RO5 were moderately biofouled.

Figure 3-5. Total solid, volatile solid and organic fraction of fouling layers deposited on

membranes. The error bars show the standard errors of the replicate samples (n=5).

§ 3.4.2.4. Proteins and polysaccharides

Proteins and polysaccharides were determined as the organic constituents of fouling layers. In

addition to the biomass deposits, protein (0.04-0.07 gBSA/m²) and polysaccharide (0.07-0.16

gGlucose/m²) were found in the fouling layers (Table 3-3). The detection of biomass and

biopolymers implied that RO membranes were biofouled. However, the amounts of protein

0

20

40

60

80

100

120

0

1

2

3

4

5

RO1 RO2 RO3 RO4 RO5

Org

an

ic F

ract

ion

(%

)

Soli

d D

eposi

t (g

/m2)

Reverse Osmosis Membranes

Total solid (g/m2)

Volatile solid (g/m2)

Organic fraction

31

are not in accordance to biomass contents in the fouling layers. It has been reported that

protein contents in biofouling are highly influenced by the properties of feed water [138].

Hence, the results of protein could not reflect the degree of fouling. The trend of

polysaccharide deposits was in the opposite order as the biomass content for each RO

modules. The polysaccharide deposits on RO4 (0.16±0.03 gGlucose/m2) and RO5 (0.15±0.08

gGlucose/m2) were two time higher than that on RO1 (0.07±0.01 gGlucose/m

2). According to

the life cycle of biofilm development [52], a conditional layer composed of inorganic and

organic is formed on the membrane surfaces at the beginning, which attracts more bacterial

cells for biofilm growth. These results suggested that RO4 and RO5 are covered with less

matured biofilm, since less biomass and high concentration of biopolymers were found in

their fouling layers.

Table 3-3: Protein and polysaccharides deposited on RO membranes.

Membrane # RO1 RO2 RO3 RO4 RO5

Proteins

(gBSA/m2, n=2)

0.06±0.01 n.d. n.d. 0.04±0.01 0.07±0.03

Polysaccharides

(gGlucose/m2, n=3)

0.07±0.01 n.d. n.d. 0.16±0.03 0.15±0.08

n.d.: not determined. The deviation ranges show the standard errors of the replicate samples.

§ 3.4.2.5. Elemental analysis

Metals concentrations, representing less than 12% of total solid deposits, were determined by

ICP-OES. Table 3-4 shows high concentrations of calcium (Ca) ions were found in all

fouling matrixes. Iron (Fe) was second abundant ion, which was dominant in the fouling

layers of RO1 and RO2 (0.04±0.01 g/m2). Divalent Ca and Fe ions primarily influence fouling

in two aspects. First, Ca and Fe ions promote the adsorption of organic matters such as

polysaccharides either to the membrane surfaces or biofouling layers [7, 34, 36]. Second, Ca

and Fe ions interact with proteins and polysaccharides, resulting complex and bridge

structures in the fouling layers [34, 36, 139].

The concentration of sodium (Na) up to 0.1 g/m2 was found in the fouling layers of RO5,

which 1 to 2-log higher than the amounts on RO1-4. Since the membrane was fouled at

seawater desalination plant, the high amount of sodium deposits must from the feed water.

32

Studies have reported phosphorus (P) is essential for all living organisms as it participates in

the synthesis of deoxyribonucleic acid (DNA), ribonucleic acid (RNA), and ATP [140, 141].

P was found on all membranes except RO4, implying biofilm is likely to be formed on these

membranes.

Table 3-4: Metal concentration of fouling layers for each membrane (ICP-OES

results)

Membrane

# RO1 RO2 RO3 RO4 RO5

Metals

(n=5)

Calcium

(g/m2)

0.01±0.00 0.03±0.00 0.03±0.00 0.01±0.01 0.01±0.00

Iron

(g/m2)

0.04±0.01 0.04±0.01 0.004±0.001 <LOD <LOD

Sodium

(g/m²) 0.01±0.00 0.01±0.00 0.02±0.00 0.003±0.000 0.1±0.0

Phosphate

(g/m2)

0.005±0.001 0.02±0.01 0.01±0.00 <LOD 0.004±0.002

Sulfur

(g/m2)

0.002±0.000 0.01±0.00 0.02±0.00 <LOD 0.01±0.00

Magnesium

(g/m²) 0.006±0.001 0.01±0.00 0.005±0.001 <LOD 0.02±0.00

Potassium

(g/m2)

0.003±0.000 0.006±0.002 <LOD <LOD 0.006±0.001

Manganese

(g/m²) 0.001±0.000 0.002±0.000 <LOD <LOD 0.001±0.001

LOD: limit of detection. The deviation ranges show the standard errors of the replicate samples.

§ 3.4.3. Scaling Characterisation

Elemental analyses (ICP-OES and SEM-EDS) were performed on the foulant characterisation

for membrane RO6. ICP-OES results revealed that the fouling layer was dominated by

calcium (21.7±3.8 gCa/m2) and smaller concentrations of barium, potassium and magnesium.

33

As the RO6 was collected from a full-scale coal seam gas water treatment plant, large amount

of calcium was found based on ICP results, suggesting the presence of calcium carbonate.

SEM-EDS analysis has supported the ICP-OES results. SEM images (n=14) showed crystals

deposited on the membrane surfaces (Figure 3-6). EDS elemental analysis (n=18) have

suggested the presence of calcium carbonate (CaCO3) on membrane RO6, as the signals of C

(15±2 % as element weight percentage, wt%) and Ca (24±5 %) are approximatively 2 to 4

times lower than signal of O (60±3 %).

Figure 3-6. SEM image (left) and EDS element weight percentage (right) of RO6 scaling

layer.

34

§ 3.4.4. Hydraulic Performances of RO Membranes

Hydraulic performances of RO membranes were measured in terms of pure water

permeability and salt rejections. No information has been provided on the operation

conditions and hydraulic performances of these RO before fouling, so no correlation between

the degree of fouling and hydraulic performances of RO modules can be made. Baseline

values for the hydraulic performances of RO membrane before cleaning are given in Table

3-5. The effect of different chemical agents on RO membranes was studied by comparing the

hydraulic performances of membrane coupons after chemical cleaning to these baseline

values (section 4.3.3). Hydraulic tests were unable to perform on RO6, as RO6 was very dry

when it was received.

Table 3-5: Hydraulic performances of RO membrane coupons (n=4).

Permeability

(L/m2.h.bar, 25℃)

Rejection

(%)

RO1 4.1±0.2 95.1±1.2

RO2 4.4±0.2 95.7±1.7

RO3 2.3±0.1 98.1±0.7

RO4 3.6±0.2 98.4±0.2

RO5 2.5±0.1 99.0±0.2

RO6 - -

§ 3.5. CONCLUSION

Membrane autopsy was carried out on six fouled RO membranes. In the membrane autopsy,

the biological, organic and inorganic fractions of fouling were determined for each RO

module. It has been concluded that membrane RO1 to RO5 were mainly biofouled, while

RO6 was scale fouled. By comparing the quantities of biomass and total solid in different

fouling matrixes, RO1 to RO3 were considered as heavily fouled and RO4 and RO5 were

moderately fouled.

The FISH analysis revealed that more than 60% of RO1 and RO5 microbial communities

have been represented by the major proteobacteria (alpha, beta and gamma) and archaea.

Additional FISH probes are required to fully demonstrate the microbial communities formed

on the RO membranes.

35

Although the hydraulic performances were no comparable between the RO modules, the

performances can be used as baseline for the cleaning tests. The results obtained in

membrane autopsy can also be the baseline of fouling characterisation. The effects of

cleaning solutions on foulants and membranes would then be revealed by comparing to these

baseline conditions.

36

4. RO MEMBRANE CLEANING USING FNA

§ 4.1. INTRODUCTION

The cleaning effect of free nitrous acid (FNA) on biological fouled RO membranes has been

studied at lab-scale in static conditions. The static cleaning test results showed a stronger

biocidal effect of FNA than the conventional cleaning sodium hydroxide (NaOH) on

microorganisms on fouled RO membrane [106]. However, the inactivation of FNA on RO

biofilm was not as efficient as its effect on sewer biofilm. It has been reported that 0.2 mg

HNO2-N/L FNA can achieve approximately 80-90% inactivation in sewer biofilm, and the

addition of 30 mg/L H2O2 to the FNA solution could further enhance its inactivation to 99%

[15]. To improve FNA cleaning efficiency on RO biofilm, dynamic cleaning tests with cross-

flow were carried out in this study. Hence, the main objective of the cleaning experiments is

to investigate and demonstrate the effectiveness of FNA for RO membrane biofouling control.

In the dynamic cleaning tests, the efficiency of FNA alone or in combination with hydrogen

peroxide (H2O2) was assessed along with the commercial cleaning agent NaOH. The analyses

focussed on the biomass residual on the membrane surfaces and potential recovery of

membrane performance in terms of permeability and salt rejection. The remaining biofilm

(biofouling residual) on membranes were analysed with previously used autopsy-based

methods. The quantities of biofouling residual were used to compare the cleaning efficiency

of FNA and the other cleaning solutions.

In addition, as an acid, it is anticipated that FNA will also be able to remove inorganics

fouling or scaling on the membrane surface. The effectiveness of FNA for RO membrane

scaling control was also tested. The cleaning procedures developed for biofouling control

were applied for the descaling cleaning tests. The descaling efficiency of FNA cleaning

solutions adjusted pH at 2 and 3 were compared with the conventional acidic cleaning

solutions such as hydrochloric acid and citric acid. Elemental analyses, ICP-OES and SEM-

EDS, were used to evaluate the descaling efficiency of FNA and the other acidic cleaning

solutions.

Results reported in this chapter formed part of the following submitted paper

E. Filloux, J.Wang, M. Pidou, W. Gernjak, Z. Yuan. 2015. Biofouling and scaling control of

reverse osmosis membrane using one-step cleaning - potential of acidified nitrite solution as

an agent.

37

§ 4.2. MATERIALS AND METHODS

§ 4.2.1. Cleaning Set-Up and Operation

Cleaning trials were carried out at lab-scale under dynamic (with cross-flow recirculation)

conditions. Five cleaning cells made of Perspex were operated in parallel with cross-flow,

without permeate production, as they were designed to simulate the configuration of RO

filtration system only (Figure 4-1). In a test, membrane coupons (150 cm2 of membrane active

surface) cut out from a fouled membrane and the respective feed spacers were placed in the

cleaning cells. All cleaning cells were connected to the same pump (Cole Parmer, Masterflex

L/S economy drive pump), this pump was assembled with five pump heads allowing cleaning

cells to run in parallel with similar flows. In order to simulate industry cleaning practice, a

cross flow velocity of 0.1 m/s was applied for the cleaning trials [142]. To conduct a cleaning

test, the following operational procedure was applied:

Tap water was pumped through the cell for 2 hours to remove the loose, external

biofilm layer that is removable using shear force. The external layer of the marine

biofilm on RO5 could further be affected by tap water due to osmotic shock.

A cleaning solution (to be described in section § 4.2.2) was pumped through the cell

for 22 hours;

Tap water was pumped through for 15 minutes to remove chemicals.

Figure 4-1. The cleaning cell (left), and the entire cleaning system with five cleaning cells in

parallel, fed with a single peristaltic pump with five pump heads.

§ 4.2.2. Design of Cleaning Tests

Based on the preliminary results of static cleaning tests (data not shown), the nitrite

concentration used in the dynamic tests was 50 mgNO2--N/L. Sodium nitrite (≤99%, Sigma

38

Aldrich) and hydrochloric acid, HCl (32%, Univar) were used to prepare the FNA solution at

different pH level. The concentration of FNA is related to the total nitrite concentration, the

pH and the temperature, and is calculated based on the following equation which is extracted

from [97]: FNA = NO2--N / (Ka x 10

pH), where Ka is the ionization constant of the nitrous

acid (Ka=e-2300/(T+273)) and T is the temperature (°C). In this study, the FNA

concentration was varied by adjusting the pH level of cleaning solutions. FNA cleaning

solutions at concentrations of 47, 35 and 10 mg HNO2-N/L were prepared by acidifying the

nitrite solution (50 mgNO2--N/L) to pH 2, 3 and 4, respectively (T=20

oC). The cleaning

efficiency of FNA was compared with that of sodium hydroxide solution (NaOH, Univar) at

pH 11 and deionised (DI) water, as controls. Although DOW and LENNTECH (RO

membrane manufacturers) indicate that caustic solutions at pH level higher than pH 11 are

more effective at biofouling cleaning, it is generally not recommended to use such harsh

cleaning solutions since NaOH at pH 11.5 can shorten the membrane life due to hydrolysis

[116, 143]. Hence, NaOH at pH 11 was selected to be the control solution in this study.

Table 4-1: Design of the biofouling control tests.

Cell #1 Cell #2 Cell #3 Cell #4 Cell #5

RO1

Test 1

Test 2

Test 3

47 mgFNA-N/L,

pH 2.0

35 mgFNA-N/L,

pH 3.0

10 mgFNA-N/L,

pH 4.0

NaOH,

pH 11.0

DI

water

RO2

Test 4 35 mgFNA-N/L,

pH 3.0

NaOH,

pH 11.0

DI

water

Test 5 35 mgFNA-N/L,

pH 3.0

35 mgFNA-N/L,

pH 3.0;

50 mg/L H2O2

35 mgFNA-N/L,

pH 3.0;

150 mg/L H2O2

NaOH,

pH 11.0

DI

water

RO3 Test 6

Test 7

35 mgFNA-N/L,

pH 3.0 50 mg/L H2O2

35 mgFNA-N/L,

pH 3.0;

50 mg/L H2O2

NaOH,

pH 11.0

DI

water

RO4

Test 8 47 mgFNA-N/L,

pH 2.0

35 mgFNA-N/L,

pH 3.0

10 mgFNA-N/L,

pH 4.0

NaOH,

pH 11.0

DI

water

Test 9 35 mgFNA-N/L,

pH 3.0

NaOH,

pH 11.0

Test 10 35 mgFNA-N/L,

pH 3.0 50 mg/L H2O2

35 mgFNA-N/L,

pH 3.0;

50 mg/L H2O2

NaOH,

pH 11.0

DI

water

RO5 Test 11

Test 12

47 mgFNA-N/L,

pH 2.0

35 mgFNA-N/L,

pH 3.0

10 mgFNA-N/L,

pH 4.0

NaOH,

pH 11.0

DI

water

39

The replication of the cleaning tests (cleaning solutions) was summarised in the following

table.

Table 4-2: The repetition of the cleaning tests

Cleaning

Solutions

47

mgFNA

-N/L,

pH 2.0

35

mgFNA

-N/L,

pH 3.0

10

mgFNA

-N/L,

pH 4.0

35

mgFNA

-N/L,

pH 3.0;

50 mg/L

H2O2

35

mgFNA-

N/L, pH

3.0;

150

mg/L

H2O2

50 mg/L

H2O2

NaOH,

pH 11.0

DI

water

RO1 n=3 n=3 n=3 n=3 n=3

RO2 n=2 n=1 n=1 n=2 n=2

RO3 n=2 n=2 n=2 n=2 n=2

RO4 n=1 n=3 n=1 n=1 n=1 n=3 n=3

RO5 n=2 n=2 n=2 n=2 n=2

Based on membrane autopsy results, the fouling layer of RO6 was mainly composed of

calcium carbonate. Therefore, the efficiency of acidified nitrous solutions (50 mgNO2--N/L,

pH 2.0, 3.0) for scaling removal was compared with these two standard cleaning solutions,

i.e., HCl (pH 2.0, 3.0), citric acid (pH 2.0, .3.0, 99.5%, Chem supply). Along with these

cleaning agents, DI water and 10 v/v % of nitric acid (HNO3, pH 0.5, Univar) were applied as

additional controls. It is reasonable to assume that CaCO3 scaling would be completely

removed from the membrane by HNO3 at this extreme pH level. Design of the descaling tests

is given in Table 4-3.

Table 4-3: Design of the descaling tests with RO6.

Cell #1 Cell #2 Cell #3 Cell #4 Cell #5

Test 1 47mgFNA-N/L;

Citric acid pH 2.0

47 mgFNA-N/L;

HCl pH 2.0

Citric acid

pH 2.0

HCl pH 2.0 DI water

Test 2 35 mgFNA-N/L;

Citric acid pH 3.0

35 mgFNA-N/L;

HCl pH 3.0

Citric acid

pH 3.0

HCl pH 3.0 10 v/v%

HNO3

Test 3 47 mgFNA-N/L;

Citric acid pH 2.0

47 mgFNA-N/L;

HCl pH 2.0

Citric acid

pH 2.0

HCl pH 2.0 DI water

Test 4 35 mgFNA-N/L;

Citric acid pH 3.0

35 mgFNA-N/L;

HCl pH 3.0

Citric acid

pH 3.0

HCl pH 3.0 10 v/v%

HNO3

40

§ 4.2.3. Post-Cleaning Analyses

§ 4.2.3.1. Post-cleaning analysis for biofouling cleaning tests

The characteristics of the biofouling layer after different cleaning tests were determined using

following methods. Samples preparation and methodology for each analysis were given in

Chapter 3. The method of live/dead cells is described in the following section.

Biomass (ATP) measurement (section 3.3.1.1)

Live/dead cells staining method (section 4.2.3.1)

FISH analysis (section 3.3.1.2)

Protein and polysaccharides measurement (section 3.3.2.2 and 3.3.2.3)

Hydraulic performances of RO membranes (section 3.3.4)

Based on the biofouling conditions before cleaning (referring to Chapter 3), the percentage of

biomass residuals and removal percentage of protein and polysaccharides under the impact of

different cleaning agents were obtained based on following equations:

Biomass residuals (%) = ATP (pg/cm²) of biofilm after cleaning / ATP (pg/cm²) of biofilm

before cleaning×100

Protein removal (%) = (1-BSA/m2 of biofilm after cleaning / BSA/m

2 of biofilm before

cleaning) ×100

Polysaccharides removal (%) = (1- Glucose/m2 of biofilm after cleaning / Glucose/m

2 of

biofilm before cleaning) ×100

From the FISH analysis, the distribution and relative abundance of major groups of

proteobacteria (Alpha-, and Beta-proteobacteria) and archaea were determined and compared

with the biofilm communities before cleaning. The detail procedure and oligonucleotide

probes which were applied for FISH analysis are given in Chapter 3, section 3.3.1.2. The

percentage of each microbe was estimated using relative abundance of the microbe (i.e.

Alphaproteobacteria) over that of total targeted microbes (bacteria + archaea).

Live/dead cell staining

The microbial viability of biofilm was determined using the LIVE/DEAD® BacLight

TM

Bacterial Viability Kits (Molecular Probes, L-7012). The viability kits contains with two

41

nucleic acid stains, the green-fluorescent SYTO-9 labels viable cells, whereas red-fluorescent

Propidium iodide (PI) stains dead cells [14]. 2×2 cm2 membrane coupons (n=5) with in situ

biofilm attachment were cut, and then submerged into 1 ml MilliQ water in the centrifuge

tubes (1 membrane coupon per tube). After mixing 1 uL of each SYTO-9 and PI stains in the

tubes, samples were incubated for 25 mins under dark condition. The stained membranes

were air dried first, then mounted onto a glass slide and observed under the Zeiss 510

Confocal Laser Scanning Microscopy (CLSM) (School of Chemistry and Molecular

Biosciences at UQ). Two excitation/emission wavelengths were used for the two fluorescent

stains: 488 nm/500 nm for SYTO-9 and 510 nm/635 nm for PI. CLSM images (n=15-60)

were randomly taken from each sample. CLSM images were used to demonstrate the

distribution and abundance of live and dead cells in the biofilm before and after the cleaning

tests. DAIME (Center for Organismal Systems Biology, Austria) was applied to estimate the

relative abundance of live and dead cells by counting green and red pixels respectively. The

proportion of viable cell was obtained using relative abundance of viable cells (green pixels)

over that of total cells (red + green pixels).

§ 4.2.3.2. Post-cleaning analysis for descaling tests

From membrane autopsy, the concentration (21.7±3.8 gCa/m2) of calcium carbonate scale

was determined. Given the size of membrane used in the cleaning tests, the concentration of

calcium carbonate scale that was removed can be estimated by measuring the calcium

contents in the cleaning solutions via ICP-OES. The descaling efficiency of different cleaning

solutions can be determined by comparing the calcium contents in the cleaning solution with

membrane autopsy result. SEM-EDS analysis was also performed on the membrane coupons

and its results can be used to justify the ICP-OES results. The characteristics of the biofouling

layer after different cleaning tests were determined using following methods. Samples

preparation and methodology for following elementary analyses are given in Chapter 3.

ICP-OES (section 3.3.3.1)

SEM-EDS (section 3.3.3.2)

42

§ 4.3. RESULTS AND DISCUSSION

§ 4.3.1. Effects of FNA Cleaning on RO Biofouling under Cross-Flow

Conditions

§ 4.3.1.1. Effects of FNA cleaning on active biomass (ATP)

ATP was used to quantify the biomass contents within biofouling layers after cleaning tests.

Figure 4-2 shows the biomass residual in percentage of the pre-cleaning level for five

different fouled RO membranes after the cleaning tests. The percentage of biomass residual

was estimated using ATP measurements after cleaning tests over initial ATP contents of

biofilm, which quantitatively demonstrates biomass residual under impacts of different

cleaning solutions.

For all RO membranes tested in the cleaning tests, the FNA cleaning solutions were more

efficient than conventional cleaning solution, namely NaOH (pH11), for removing biomass.

Acidified nitrite cleaning solutions removed 7-45% more biomass than NaOH for heavily

fouled membranes, RO1-3 (referring to Chapter 3). For moderately fouled membranes, RO4

and RO5 (referring to Chapter 3), 3-5% and 2-3% more biomass were removed by nitrite

cleaning solution at low pH level for RO4 and RO5 respectively. Although the RO2 biofilm

contained the highest biomass load (4234±15423 pgATP/cm2) (referring to Figure 3-3), FNA

solution at pH 3 has achieved the best cleaning efficiency for RO2 among heavily fouled

membranes. Hijnen et al. reported the resistance of biofilm against cleaning varies based on

the structure of biofilm [91]. It is likely that the natural structure of biofilm formed on RO2

has created less resistance to FNA cleaning solution. Hijnen et al.’s theory also applies to

cleaning tests for moderately fouled membranes. The biomass residual results have shown

FNA cleaning was more efficient for moderately fouled membranes, implying that cleaning

efficiency of cleaning agents is influenced by the degree of fouling.

The results showed all cleaning solutions were more efficient for cleaning moderately fouled

membranes than heavily fouled membranes, suggesting cross-flow conditions have created

great effects on the loose biofilm layers. It is likely that cross-flow conditions helped to

facilitate the diffusion of cleaning solutions into biofouling layers, detach bacterial cells and

flushed them away [88]. However, it is evident that biomass removal was not mainly caused

by hydrodynamic shear, since removal caused by FNA cleaning solutions is significantly

43

higher compared to removal with DI water only. The results have suggested that applying

cross-flow enhances the cleaning efficiency of FNA.

Figure 4-2 (a) also demonstrated the effect of pH level on the cleaning efficiency of nitrite

cleaning solutions. For heavily fouled membrane RO1, the best cleaning efficiency of nitrite

cleaning solutions is observed at pH 3. Based on the three cleaning tests, more than 86% of

biomass was removed by nitrite cleaning solutions at pH 3. The concentration of FNA is

inversely correlated to the pH level of nitrite solutions. The results have proved higher FNA

concentration at pH 3 has resulted in better cleaning efficiency. However, there was no

obvious further improvement to biomass removal for nitrite solutions at pH 2, likely due to

the loss of nitrite concentration induced by the conversion of nitrite to nitrate [106]. Hence,

50 mg NO2-N/L at pH 3.0 as optimum conditions were used to clean RO2 and RO3.

For moderately fouled membranes, there was no obvious difference between the cleaning

efficiency of nitrite cleaning solutions at different pH levels. The results have suggested that

lower concentration of nitrite can be applied to clean moderately fouled membranes.

44

a.

b.

Figure 4-2. Biomass residual after 24 hours cleaning tests performed in cross-flow conditions

with the membranes (a) heavily fouled RO1, RO2 and RO3, and (b) moderately fouled RO4 and

RO5. The cross-flow velocity applied was 0.1 m/s. The error bars show the standard errors of

three replicate experiments. The results without error bars are based on three measurements from

each experiment.

45

§ 4.3.1.2. Influence of FNA on viability of biomass

The biocidal effect of FNA has already been demonstrated on anaerobic sewer biofilm and waste

activated sludge applications [14, 144]. Thus, it is anticipated FNA would affect biofilm formed

on RO membranes in a similar way. The potential inhibition of FNA was assessed by measuring

live and dead bacterial cells after cleaning tests. The CLSM image analysis was carried out

directly on biofilm on the RO membranes. Due to the density and thickness of the biofilm on the

heavily fouled membranes (RO1 to RO3), it was difficult to reveal the accurate quantification of

live and dead cells. CLSM imaging could not give quantitative results due to the density of

biofilm changing from location to location on the membrane [138]. Therefore, live and dead

analysis did not appear to be applicable to heavily fouled biofilm.

In situ CLSM observation and image analysis could be performed on the moderately fouled

membranes (RO4&5) due to lower biofilm density. The CLSM images of RO4&5 before and

after cleaning are given in Appendix Figures C-1 and C-2. Comparing to the biomass before

cleaning, 13-39% and 22-54% of viable cells were inactivated and removed by chemical

cleaning solutions for RO4 and RO5, respectively (Figure 4-3). The trends of viable cell revealed

the best inactivation effect of nitrite cleaning solutions is at pH 3 for RO4, and there was no

distinct difference between the inactivation effects of nitrite cleaning solutions at different pH

levels on RO5 biofilm.

In accordance with the biomass residual measured as ATP, the inactivation efficiency of FNA is

better than NaOH (pH 11). 32±5% of total cells remained activated in the biofilm on RO4 under

the impact of FNA at pH 3, which is 26% less than that under impact of NaOH. For RO5, the

biofilm contained only 6-7% of viable cells after cleaning with nitrite solutions, whereas 38±4%

of viable cells remained in the biofilm which was cleaned by NaOH. CLSM images (n=15-45)

also explain the more superior inactivation effect of FNA solutions on RO5 biofilms in

comparison to that on RO4. CLSM images revealed that the biofilm formed on RO4 was denser

than biofilm on RO5. It appears the denser biofilm on RO4 creates resistance to the

hydrodynamic flow, which then reduces the interaction between cleaning chemicals and biofilm,

hence hinders the cleaning efficiency. The results of live/dead cells staining confirm that the

bactericidal efficiency of FNA on bacteria formed on RO membranes. The results additionally

46

demonstrated FNA is more efficient than NaOH in killing bacteria.

Figure 4-3. Proportion of viable cells in membrane biofilm before and after 24 hours cleaning

tests for membranes RO4 and RO5. A cross-flow velocity of 0.1 m/s was applied. The error bars

show the standard errors of 15 to 60 CLSM images.

§ 4.3.1.3. Biofilm community structure changes after FNA cleaning

The selected FISH probes have revealed that more than 60% of biofilm communities on RO1

and RO5 were Alphaproteobacteria, Betaproteobacteria, Gammaproteobacteria and archaea, so

the impacts of FNA on these microbes were further investigated using the same probes. The

relative quantity of each detected bacterium was determined by DAIME based on FISH images,

which is given in the appendix (Appendix Table B-2).

FISH results showed that the total amount of detected bacteria has been dropped after the

cleaning tests. Figure 4-4 (a) and (b) show targeted bacteria in the biofilm of RO1 decreased 70%

after FNA cleaning at its optimal condition (pH 3). Figure 4-4 (b) and (d) indicate that 38% of

total detected bacteria of RO5 biofilm were removed under the same treatment. This result also

revealed that the microbe communities of RO1 were more susceptible to the impact of FNA

47

cleaning than that of RO5. The rise of Alphaproteobacteria in the population of RO5 biofilm is

likely due to the drastic removal of Betaproteobacteria and archaea, the relative quantity of all

bacteria detected and Alphaproteobacteria has been reduced under the impact of FNA cleaning

(Appendix table B-2). FISH analysis results show that Betaproteobacteria and archaea were

greatly affected by FNA cleaning for both membranes. However, the FNA effects were less

consistent on Alphaproteobacteria. Similar phenomenon has been reported by Bereschenko et al.

[123], the weekly chemical cleaning (30% sodium bisulfite solutions and mixed acid detergent

descaler) has more impact on Beta- and Gamma-proteobacteria than Alpha-proteobacteria on

the membranes after 3-6 month operation. Due to limited availability of biofilm residues after

cleaning, limited oligonucleotide probes were chosen for FISH analysis. Large proportion of

bacteria was not detected by the selected probes, especially for RO5, suggested that the results of

FISH analysis are not comprehensive for the purpose of this study.

(a) RO1 Before Cleaning

(b) RO1 After Cleaning

(c) RO5 Before Cleaning

(d) RO5 After Cleaning

Figure 4-4. The abundance of proteobacteria (Alpha, Beta, and Gamma) and archaea in the

biofouling layers before and after 24 hours cleaning tests for the membranes RO1 and RO5,

respectively. Standard test conditions: FNA (50 mgN-NO2/L), pH 3, cross-flow velocity 0.1 m/s.

The abundances of each microbe were calculated based on the FISH images (n=20±5).

48

§ 4.3.1.4. Effect of FNA on protein and polysaccharide removal

In addition to biomass analyses, the impact of FNA and other cleaning solutions on the removal

of the organic constituents of biofilm was studied for moderately fouled membranes (RO4&5).

This impact was reflected by evaluating the removal of proteins and polysaccharides by the

cleaning. By comparing the average removal efficiency of proteins and polysaccharides achieved

by the FNA cleaning solutions with that by NaOH (Table 4-4), FNA cleaning solutions have

shown similar or better efficiency for proteins and polysaccharides removal. The removal of

proteins and polysaccharides was likely due to the interaction between FNA or its reactive

nitrogen species (RNS) and biofilm, as FNA is able to damage chemical bond and structure of

EPS in waste activated sludge [18]. Based on results obtained from all analyses, FNA can be a

substitute cleaning chemical for NaOH.

Table 4-4: Comparison of FNA and NaOH cleaning effects on proteins and polysaccharides.

Cleaning Agents FNA NaOH

Membrane # RO4 RO5 RO4 RO5

Protein removal rate (%) 93.1±3.4 67.1±2.5 60.3±14.9 59.16±0.0

Polysaccharides removal rate (%) 72.9±2.0 86.5±1.8 61.7±7.5 79.36±0.0

The deviation ranges show the standard errors of the FNA cleaning solutions (at pH 2.0, 3.0 and 4.0)

The comparison of the effects of all cleaning solutions on biomass, protein and polysaccharides

is given in Figure 4-5. Figure 4-5 (b) shows the removal rate for biomass and polysaccharides

and proteins presented a similar trend. However, FNA cleaning solution was still more efficient

on biomass removal (94-95%) than on protein and polysaccharides removal, respectively (Table

4-4). Many studies have reported that bacteria are more susceptible than EPS to the cleaning

chemicals, in which polysaccharides and proteins have been used as proxy of EPS [91, 123, 138].

Hijnen et al. and Bereschenko et al. [91, 123] revealed that the biofilm matrix consisting of

polysaccharides and proteins creates greater resistance to the cleaning attack, though the bacteria

cells are removed during the cleaning processes.

49

a.

b.

Figure 4-5. Biomass removal (%, based on ATP values), protein and polysaccharide removal

(%) after 24 hours cleaning tests for the membranes (a) RO4 and (b) RO5. A cross-flow

velocity 0.1 m/s was applied. The error bars in the plot show the standard errors of 2-3

replicate experiments.

50

§ 4.3.1.5. Synergistic cleaning effects of FNA and H2O2 on RO membrane biofouling

Jiang et al. reported adding H2O2 can enhance the biocidal efficiency of FNA by 43-51%

compared with FNA alone [15]. Hence, the synergistic cleaning effects of FNA and H2O2 on RO

biofouling were studied and compared with effects of FNA and H2O2 individually, as shown in

Figure 4-6. For both heavily fouled RO2 and moderately fouled RO4, minimal enhancement

(less than 1%) was obtained after adding H2O2 to FNA. Increasing the concentration of H2O2 to

150 mg/L showed no improvement in cleaning for RO2. However, FNA and FNA/H2O2 are still

more efficient than NaOH on biomass (ATP) removal for all three RO membranes. Given that

the cleaning efficiency of H2O2 alone is lower than FNA or the combination of FNA and H2O2,

this implies that FNA is still the main cleaning agent for the biofouling removal.

Figure 4-6. Biomass removal after 24 hours cleaning tests for the membranes heavily fouled

RO2 and RO3 and moderately fouled RO4. A cross-flow velocity 0.1 m/s was applied. The error

bars show the standard errors of three replicate experiments. The results without error bars are

based on three measurements from each experiment.

51

§ 4.3.2. Descaling Efficiency of FNA

Since the FNA cleaning solutions are formed by combining nitrite with acid, it is expected that

FNA cleaning solutions would act like acid to remove scaling from membrane surfaces. The

same cleaning procedures used for biofouling were carried out to remove scaling from RO

membrane (RO6). After 24 hour cleaning tests, the descaling efficacy of FNA and the other

cleaning solutions (as controls) were assessed by evaluating the dissolved inorganic elements in

the cleaning solutions. Figure 4-7 presents the dissolved calcium content based on ICP-OES

results. 28.5±4.6 to 34.3±1.4 g/cm2 calcium was removed by all acidic solutions, which suggests

that all cleaning solutions are efficient at calcium scaling removal. The results suggested that

nitrite has no influence on the descaling efficiency of acidic solutions, since the descaling

efficient of FNA at pH 2 and 3 (either adjusted by HCl or citric acid) was similar to conventional

descaling agents (HCl and citric acid). In accordance to the cleaning tests for biofouling removal,

it was proven again that hydrodynamic shear is beneficial for the cleaning tests as it improves the

solubility of the fouling layer. HNO3 (10 v/v % at pH 0.5) the control cleaning solution

successfully removed 32.4±1.7 g/cm2 versus the autopsy results 21.7±3.8 g/cm

2. The calcium

removal rate is higher in cross-flow conditions. However, scaling removal was not mainly

caused by hydrodynamic shear, since only 4.4±0.1 g/cm2 of calcium was removed from the

membrane surface by water washing.

52

Figure 4-7. Dissolved calcium content removed from the membrane surface after 24 hours

cleaning tests with membrane RO6. A cross-flow velocity 0.1 m/s was applied in all tests. The

error bars show the standard errors of four measurements from two replicate experiments.

In addition, SEM-EDS analyses were conducted on the membrane after cleaning for nitric acid

and all cleaning solutions at pH 3, which is also the optimal pH level selected for biofouling

removal. SEM images and element distribution (wt%) are available in the appendix (Appendix

Figures D-1 and D-2). In agreement with the ICP-OES results (Figure 4-7), the SEM-EDS results

revealed that the calcium contents of scaling layers reduced from 24% to 1% under impacts of all

cleaning solutions adjusting pH to 3, implying that calcium carbonate scaling has been

effectively removed. This result is supported by SEM images of calcium carbonate scaled RO

membrane before and after cleaning. SEM images revealed that the crystal structure of calcium

carbonate scaling was removed by all cleaning solutions at pH 3. Overall, the results of the

cleaning tests for biofouling and scaling removal suggested that FNA can be used as a single

cleaning agent for both biofouling and scaling removal.

53

§ 4.3.3. Hydraulic Performances of RO Membranes after Cleaning

As result of effective cleaning for RO membrane, it is anticipated that permeability would be

improved and salt rejection would be increased. However, based on the filtration results, there

was no significant difference in permeability and salt rejection for all tested RO membranes after

cleaning. According to RO membrane manufacturers, the permeability of RO module is normally

±15–20% of its nominal rate due to membrane manufacturing and experimental error[145]. Since

all permeability changes remained in the range of this deviation (Table 4-5), no conclusive

statement could be made based on the hydraulic results for this study. In addition, fouling is

likely non-uniformly spread out on membranes surface. The small coupons (active area 42 cm2)

used in the filtration trials do not completely represent the full scale RO module. Applying small

coupons would have also caused large variations in performance of the membrane coupons. The

results have suggested that the lab-scale filtration system was not suitable for permeability and

salt rejection tests. It is necessary to study the cleaning impact of FNA on the hydraulic

parameters at larger-scale filtration units.

54

Table 4-5: Hydraulic performances of membranes after cleaning tests.

Membrane #

Membrane autopsy Cleaning Tests

Permeability

(L/m2.h.bar, 25

oC)

Water Rinse

(2hr) Water

NaOH

pH 11

FNA

pH 2

FNA

pH 3

FNA

pH 4

RO1 4.1 1.3

3.7

4.7 2.9

4.8 4.6

RO2 4.3 4.0 3.8 4.7

4.3

4.0 4.2 4.2

3.9

RO3 2.3 2.2 1.9 2.2

2.2

RO4 3. 6

3.6 3.6 3.4 3.7 3.2

4.0

2.7

3.7 2.9

3.7

RO5 2.5 2.8 2.7 2.7 2.8 2.8

2.6 2.6

Rejection (%)

Water Rinse

(2hr) Water

NaOH

pH 11

FNA

pH 2

FNA

pH 3

FNA

pH 4

RO1 95.1 95.9

95.1

95.4 96.5

96.1 95.6

RO2 95.7 95.4 95.5 96.4

97.5

92.0 96.3 95.9

97.6

RO3 98.1 98.6 98.5 98.7

98.2

RO4 98.4

97.3 98.8 99.4 96.3 99.1

99.0

98.0

98.1 98.5

98.3

RO5 99.0 99.2 98.9 99.3 99.4 98.8

98.1 98.7

55

§ 4.4. CONCLUSION

Five biofouled membranes (RO1-5) were used in 12 dynamic cleaning tests in which 0.1 m/L

cross-flow was applied. The cleaning tests have shown that FNA solutions are more efficient at

biofouling cleaning than NaOH (pH 11) and 35 mgFNA-N/L at pH 3.0 is the optimum cleaning

conditions in this study, based on the results of biomass, protein and polysaccharides analysis,

and live/dead cells staining.

The ATP results showed that all nitrite cleaning solutions (pH 2.0, 3.0 and 4.0) removed 7-45%

more biomass than NaOH for heavily fouled membranes, RO1-3 (referring to Chapter 3).

Although the superior cleaning efficiency of nitrite cleaning solutions was not obvious for

moderately fouled membranes (referring to Chapter 3), there were still 3-5% and 2-3% more

biomass removed by nitrite cleaning solutions than NaOH for RO4 and RO5, respectively. This

result suggested that all chemical cleaning solutions are more efficient for cleaning moderately

fouled membranes than heavily fouled membranes, under cross-flow cleaning conditions.

Live/dead cells, protein and polysaccharides measurements were able to be performed on

moderately fouled membranes (RO4&5), due to their less dense biofilm structure. In accordance

to ATP results, live/dead cells staining has revealed that less viable cells remain in biofouling

layers after FNA cleaning at pH 3.0 (32±5% for RO4 and 7±2% for RO5) than NaOH (57±5%

for RO4 and 38±4% for RO5). The removal rate (%) of protein and polysaccharides showed a

similar trend as the results of ATP and CLSM analysis. However, FNA cleaning appeared more

efficient for biomass removal.

FISH analysis has demonstrated that the overall abundances of targeted bacteria on RO1 and

RO5 have been reduced under the impact of FNA cleaning at pH 3.0. The results of FISH

analysis also revealed that Betaproteobacteria and archaea were greatly affected by FNA

cleaning than Alphaproteobacteria.

Although applying FNA alone, or combine FNA and H2O2 have shown better efficient at

biofouling removal than NaOH, the percentage of biomass residual showed combining FNA with

H2O2 was not able to improve the cleaning/removal efficiency of FNA significantly (less than 1%

56

of enhancement).

The scaling cleaning tests revealed that FNA solutions at pH 2.0 and 3.0 were as efficient as

conventional descaling acids based on elemental analyses (ICP-OES and SEM-EDS). The results

of both analyses showed that the scaling layers were effectively cleaned by all acidic cleaning

solutions. Based on the outcomes of biofouling and scaling cleaning tests, FNA has showed its

better at both biofouling and scaling cleaning. Hence, FNA can be a promising cleaning agent to

achieve biofouling and scaling removal at a single stage.

For all cleaning tests, performances of tested membrane coupons were examined in terms of

permeability and salt rejection. No significant difference was observed for all membrane, which

was likely due to the small size of filtration cells used in this study.

57

5. RO BIOFILM PREVENTION USING FNA

§ 5.1. INTRODUCTION

The results obtained in previous chapter suggested that FNA cleaning was more effective for

moderately fouled membranes, suggesting that early cleaning is preferable. Based on this finding,

the application of FNA cleaning for biofouling prevention, rather than for biofouling removal,

was investigated. In this chapter, RO membranes were used for filtrating secondary effluent (SE).

Additional nutrients were introduced to the filtration systems in order to generate fouling. Two

bench-scale crossflow RO filtration units were applied in parallel: one unit was fed with SE only

(the control unit), while the second one was fed with SE and exposed to FNA cleaning (10 mg

FNA-N/L, pH 3) for 6 hours (the experimental unit) on a weekly base.

The permeability and salt rejection of both filtration systems have been continuously monitored

for three weeks. The operational data of filtrations with or without FNA cleanings were used to

reveal the effects of fouling and FNA on the hydraulic performances of RO membranes. After

last FNA cleaning at the end of the filtration experiments, membrane coupons were retrieved

from the filtration units, and foulant analyses such as ATP, protein, polysaccharides

measurements, elemental analysis (ICP-OES) and live/dead staining were carried out on the

fouling material deposited on the membranes. The results of foulant analyses were used to

disclose the effect of FNA on the fouling layers.

58

§ 5.2. MATERIALS AND METHODS

§ 5.2.1. Membranes

Membrane coupons were cut from a polyamide thin-film composite RO membrane, ESPA 2

(Hydranautics, USA) and used for the filtration tests. New RO membranes were compacted with

10 L DI water overnight. The general performances of membranes were characterised in terms of

pure water permeability and salt rejection prior to the filtration tests. 1500 mg/L NaCl salt

solution was used in the salt rejection measurements.

§ 5.2.2. Bench-Scale RO Filtration Units

Two bench-scale crossflow filtration units were operated in parallel to foul the RO membranes.

Both filtration units were made of stainless steel, a schematic diagram of the filtration unit is

presented in Figure 5-1. Within the filtration cell body, the shim, the feed spacer and permeate

spacer were installed to give a 0.86 mm channel height. Since filtration cells have dimensions of

14.6 and 9.5 cm for flow channel length and width respectively, so the active area of membrane

is 139 cm2 and the cross-section flow area is 0.82 cm

2. Base on the cross-section flow area, the

feed flow was kept at 30 L/h to maintain the cross-flow velocity at 0.1 m/s. The active membrane

area was used to estimate permeability based on the Equation 10 given in Chapter 3. The feed

pressure was maintained at 8 bar and was not particularly adjusted to compensate the fouling

during the experiments. The conductivities of the feed and permeate were obtained using a

SevenEasy conductivity meter (Mettler Toledo, USA) for salt rejection measurements (referring

to Equation 11 in Chapter 3). The temperature of feed SE was recorded with an IC9237

thermometer (Synotronics, Australia), and all filtration experiments were carried out at ambient

temperature.

59

Figure 5-1. A schematic diagram of the crossflow membrane filtration unit

§ 5.2.3. Operation Conditions and Analyses

Two short-term filtration tests were performed. For all the tests, one of filtration units was used

as the control system, in which only secondary effluent (SE) was circulating for the entire

filtration trials. The other unit was operated as the experimental system, in which the membrane

coupons were cleaned by FNA for 6 hours on a weekly base. A cleaning duration of 6 hours was

chosen in order to complete the experiments within one day, as otherwise full monitoring and

potential intervention of the experiments would not be possible. This is mainly due to the

limitations associated with the operation of laboratory systems. The duration and frequency of

the cleaning applied in this study is not directly comparable to those applied to full-scale plants.

Hence, the cleaning conditions should be verified at full-scale and under more realistic

conditions in the future.

The FNA (10 mg FNA-N/L) solutions were prepared by dissolving the sodium nitrite (71 mg/L)

in SE and the pH of FNA solutions were adjusted to pH 3 using HCl. 10 L of SE was weekly

collected from the Luggage point wastewater treatment plant (Brisbane, Australia) and pre-

60

filtered through 1.2 µm filters (nylon filter membrane, PM separations, Australia) before to be

used. Synthetic nutrient was added to the feed every second day, in order to speed up the

biofouling process and make the thesis possible within the timeframe available. The composition

of the synthetic nutrient is given in Table 5-1.

Table 5-1: The composition of synthetic nutrient [146].

Substance [Supplier] Concentration

[mg/10 L]

Sodium Acetate (NaCH3CO2) [Chem Supply, Australia] 136

Urea (CH4N2O) 3

Di-Potassium Hydrogen Phosphate (K2HPO4) [Univar, Australia] 2.8

The operation conditions and analyses are given in

Table 5-2. Permeate flow rate of both filtration units, and the conductivity of feed and permeate

were measured on daily base. The permeability and salt rejection of membranes were calculated

based on the Equation 10 and Equation 11 given in Chapter 3.

Table 5-2: The operation conditions of filtration tests and foulant analyses performed after

the filtration tests.

Operating conditions

Filtration Test 1 Filtration Test 2

Control Experiment

(FNA Cleaning) Control

Experiment

(FNA Cleaning)

Period of run (days) 14 21 22 22

Foulant sampling After last FNA cleaning, at end of filtration tests

Foulant analysis ATP analysis, protein and polysaccharides measurements, elemental

analysis (ICP-OES), and live/dead cells staining method

Foulant analyses with replicated samples (n=2-4) were carried out based on the procedures given

in Chapter 3 and 4. CLSM and live/dead staining method were used to reveal the viability of

bacterial cells on the membrane surface at the end of the filtration trials. 35 CLSM images were

61

taken on the in situ foulants deposited on the membranes. The relative counts of live and dead

cells were obtained via DAIME, and the viability of bacterial cells in the fouling layers was

estimated using the percentage of live/total.

Biomass (ATP) measurement (section 3.3.1.1)

Live/dead cells staining method (section 4.2.3.1)

Protein and polysaccharides measurement (section 3.3.2.2 and 3.3.2.3)

ICP-OES (section 3.3.3.1)

§ 5.3. RESULTS AND DISCUSSION

§ 5.3.1. Effects of FNA Dosing on the Hydraulic Performances of RO Membranes

§ 5.3.1.1. Normalised permeability (NP)

Figure 5-2 shows the NP of control and experiment filtration systems has decreased along the

filtration tests, indicating fouling has occurred. After 14 days operation, the decline rate of

permeability was similar in the control (54±1%) and the experiment (56±8%) filtration units,

based on the average NP data of two tests. In test 2, there is only 7% of NP difference between

two filtration systems after 21 days operation. The continuing decline of NP was likely caused by

the accumulation of fouling material at the surface of RO membranes. The performances of RO

could be initially hindered by salt concentration polarisation (CP). The salt CP is occurred when

rejected ionic species accumulate at the membrane surface [39]. The cake layer induced by CP

was likely formed to create great hydraulic resistances, which is normally responsible for the

osmotic pressure rise near the membrane surface [35, 39, 42]. The NP of both filtration systems

has gradually declined over the experiments, which was likely caused by the enhanced hydraulic

resistances and increased osmotic pressures.

Figure 5-2 also shows that NP has been controlled or recovered during the weekly FNA cleaning.

After the second FNA cleaning, the NP was maintained at the similar level for three days in test

1, while 13% of NP was recovered in test 2. After the third time cleaning, 12% and 15% of NP

were recovered in test 1 and 2, respectively. The short-term recoveries of NP reflected that the

development of fouling was alleviated under the impact of FNA cleaning. However, it was

unable to deter the fouling completely. Therefore, the results indicated the inhibition rate of FNA

was slower than the growing rate of fouling and the continuing decline of NP suggested that

FNA cleaning were inefficient at preventing fouling.

62

a.

b.

Figure 5-2. Normalised permeability (Kw/Kw0) of filtration test 1 (a) and test 2 (b). Standard

test conditions: SE was circulating at 0.1 m/s in both filtration systems. For the experiment

filtration, the data points highlighted in red represent that the membrane was cleaned by FNA

(10 mg-N/L) at pH 3 for 6 hours on these specific. The scatter points are based on two

measurements from each experiment on each day.

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

20

15

/3/1

1

20

15

/3/1

2

20

15

/3/1

3

20

15

/3/1

4

20

15

/3/1

5

20

15

/3/1

6

20

15

/3/1

7

20

15

/3/1

8

20

15

/3/1

9

20

15

/3/2

0

20

15

/3/2

1

20

15

/3/2

2

20

15

/3/2

3

20

15

/3/2

4

20

15

/3/2

5

20

15

/3/2

6

20

15

/3/2

7

20

15

/3/2

8

20

15

/3/2

9

20

15

/3/3

0

20

15

/3/3

1

No

rmal

ise

d P

erm

eab

ility

Kw

/Kw

0

Test 1 Control Filtration

Test 1 Experiment Filtration with Weekly FNA Cleaning

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

1.1

20

15

/4/1

20

15

/4/2

20

15

/4/3

20

15

/4/4

20

15

/4/5

20

15

/4/6

20

15

/4/7

20

15

/4/8

20

15

/4/9

20

15

/4/1

0

20

15

/4/1

1

20

15

/4/1

2

20

15

/4/1

3

20

15

/4/1

4

20

15

/4/1

5

20

15

/4/1

6

20

15

/4/1

7

20

15

/4/1

8

20

15

/4/1

9

20

15

/4/2

0

20

15

/4/2

1

20

15

/4/2

2

No

rmal

ise

d P

erm

eab

ility

Kw

/Kw

0

Test 2 Control Filtration

Test 2 Experiment Filtration with Weekly FNA Cleaning

63

§ 5.3.1.2. Salt rejection

Figure 5-3 presents the salt rejection data of both filtration systems. In accordance with the NP

plots, salt rejection of all RO membranes used in filtration tests has also gradually decreased

along the filtration tests, suggesting membranes were fouled. In test 2, the salt rejection of RO

membranes from control systems dropped from 98% to 95%, whereas the RO membranes

exposed to FNA cleanings had less than 2% decline in salt rejection. The decline of salt rejection

was likely caused by the same fouling mechanisms which lead to the decline of NP. Studies have

reported that concentration polarisation (CP) raises the salt passage across RO membranes and

cake layers also enhance the passage of salt solutes [31, 39, 42]. It is noteworthy that the salt

rejection trend obtained in the experiment is not in accordance to that observed in real practice.

In practice, the feed pressure is increased to compensate the fouling in order to ensure the

productivity of RO membranes. Water molecules are pushed through the membrane at a faster

rate than the salts under such circumstance. As a consequence, the salt rejection increases as

fouling progresses. In contrast, the feed pressure was maintained at the same level throughout the

experiments in this study. That explains why the salt rejection trend observed in this study is

different to what observed at real plants. As mentioned before, the cleaning condition applied in

this study is not directly comparable to those applied to full-scale plants. Hence, the cleaning

conditions should be verified under more realistic conditions at full-scale.

In Figure 5-3, it appeared that FNA cleanings have applied a short-term reversible effect on RO

membranes. During FNA cleaning, the salt rejection of RO membranes dropped 7 to 11%

compared to the rejection data before FNA cleaning (unfilled data points in Figure 5-3), which

quickly came back after switching the FNA solution back to SE in both tests. Since the FNA

cleaning solutions were prepared by new SE, the sharp drops of salt rejection were likely due to

the sudden salinity changes of the feed water [147, 148].

64

a.

b. Figure 5-3. Salt rejection of the filtration test 1 (a) and test 2 (b). Standard test conditions: SE was

circulating at 0.1 m/s in both filtration systems. For the experiment filtration, the data points highlighted in

red represent that the membrane was cleaned by FNA (10 mg-N/L) at pH 3 for 6 hours on these specific

dates. The scatter points are based on two measurements from each experiment on each day.

80%

82%

84%

86%

88%

90%

92%

94%

96%

98%

100%

20

15

/3/1

1

20

15

/3/1

2

20

15

/3/1

3

20

15

/3/1

4

20

15

/3/1

5

20

15

/3/1

6

20

15

/3/1

7

20

15

/3/1

8

20

15

/3/1

9

20

15

/3/2

0

20

15

/3/2

1

20

15

/3/2

2

20

15

/3/2

3

20

15

/3/2

4

20

15

/3/2

5

20

15

/3/2

6

20

15

/3/2

7

20

15

/3/2

8

20

15

/3/2

9

20

15

/3/3

0

20

15

/3/3

1

Salt

Rej

ecti

on

(%

)

Test 1 Control Filtration

Test 1 Experiment Filtration with Weekly FNA Cleaning

Test 1 Salt Rejection Data before Weekly FNA Cleaning

80%

82%

84%

86%

88%

90%

92%

94%

96%

98%

100%

20

15

/4/1

20

15

/4/2

20

15

/4/3

20

15

/4/4

20

15

/4/5

20

15

/4/6

20

15

/4/7

20

15

/4/8

20

15

/4/9

20

15

/4/1

0

20

15

/4/1

1

20

15

/4/1

2

20

15

/4/1

3

20

15

/4/1

4

20

15

/4/1

5

20

15

/4/1

6

20

15

/4/1

7

20

15

/4/1

8

20

15

/4/1

9

20

15

/4/2

0

20

15

/4/2

1

20

15

/4/2

2

Salt

Rej

ecti

on

(%

)

Test 2 Control Filtration

Test 2 Experiment Filtration with Weekly FNA Cleaning

Test 2 Salt Rejection Data before Weekly FNA Cleaning

65

§ 5.3.2. EFFECT OF FNA ON THE MEMBRANE FOULING LAYER

After the last FNA cleaning in test 2, fouling deposits were collected from membrane coupons

and undergone foulant analyses. Table 5-3 presents the biomass contents, the percentage of

viable cells, and the concentration of proteins, polysaccharides and metal ions in the fouling

layers formed on the RO membranes. The results showed that all fouling layers were composed

of microorganisms, organic (such as proteins and polysaccharides) and inorganic materials.

However, total metal deposits were less than 1 g/m2 in all fouling layers, suggesting that there

were mainly biological and organic foulants deposited on the RO membrane. All fouling layers

contained more than 1000 pgATP/cm2, implying biomass contents were high in both filtration

units [149]. Table 5-3 also indicates the percentage of viable cells from experiment filtration was

four times higher than that from control filtration unit. Moreover, in accordance to the biomass

measurement (ATP), the relative count of total cells after exposing to FNA cleaning was less

than half of that from the control unit, as given in Appendix Table C-1. Figure 5-4 presents

distribution of live and dead cells on the membranes collected from control and experimental

filtration units, respectively.

Table 5-3: Summary results of foulant analyses

Control Experiment

(FNA Cleaning)

Biological Analysis

Adenosine triphosphate (pg/cm2) (n=6-9) 23112±386 9967±608

Viable/Total cells (%) (n=35) 57±5 13±2

Organic Analysis (n=4-6)

Proteins (gBSA/m2) 1.23±0.00 0.95±0.02

Polysaccharides (gGlucose/m2) 0.98±0.03 0.78±0.02

Elemental Analysis (n=4)

Calcium (g/m2) 0.19±0.01 0.06±0.00

Iron (g/m2) 0.06±0.00 0.10±0.00

Sodium (g/m2) 0.07±0.00 0.09±0.00

Phosphate (g/m2) 0.16±0.01 0.06±0.00

ICP results (0.01-0.1 g/m2) K, Mg, Mn, S, Zn Cu, Mg, S, Zn

ICP results (0.001-0.01 g/m2) Cu K

Sum of total metal ions (g/m2) 0.63±0.01 0.45±0.02

Note: The deviation ranges show the standard errors based on the number of the data points of

each analysis.

66

(a)

(b)

Figure 5-4. CLSM images of fouling samples from control (a) and experimental (b) filtration

units. Live cells were stained in green with the DNA specific dye SYTO®9 and the dead cells

are stained in red with PI. Standard test conditions: a cross-flow velocity 0.1 m/s was applied.

FNA (10 mg FNA-N/L) at pH 3 was cleaned for 6 hours in the experimental filtration unit on a

weekly base.

The results of biological analysis suggested that FNA cleanings were efficient at inhibiting

bacteria. In contrast, the concentration of proteins and polysaccharides were similar for both

filtration systems. This outcome is in accordance to the discussion regarding the effect of FNA

on proteins and polysaccharides (section 4.3.1.4), implying FNA cleanings were not effective

for the macromolecules such as proteins and polysaccharides cleaning. The inefficient removal

of these macromolecules might be responsible for declined performance of membrane in both

filtration systems, as these macromolecules were likely to impede the flow of feed water and

enhance the fouling. Overall, the results of foulant analyses suggested that FNA was more

efficient at killing bacteria than disrupting proteins and polysaccharides in the fouling layers,

which might explain the continuing loss of membranes’ hydraulic performance.

67

§ 5.4. CONCLUSIONS

Two bench-scale crossflow filtration units were used to test the new cleaning method using FNA

to prevent fouling on RO membranes. SE and additional nutrient were added to the filtration

units to generate fouling. Although FNA cleanings have showed the ability to restore the

permeability of RO membranes for a short period, the result revealed that FNA cleanings were

unable to deter the performance (NP and salt rejection) decline of RO membranes and the

formation of fouling on RO membranes. The results of foulant analysis indicated that the fouling

materials collected from both filtration systems were mainly composed of microbes and organic

matters. Although the composition of proteins and polysaccharides in the fouling layers were

found to be similar, the biomass contents and viable cells of the fouling layers formed in the

experiment filtration (with FNA cleanings) unit were less than half of that in the control filtration

unit. The abundance of viable cells in the control unit (57±5%) was found to be four times higher

than that in the experiment unit (13±2%). Although the FNA cleaning was efficient at

inactivating bacteria, it was inefficient to clean proteins and polysaccharides. The accumulation

of protein and polysaccharides was likely attributed to the continuing loss of the hydraulic

performance of RO membranes.

68

6. CONCLUSIONS AND RECOMMENDATIONS

§ 6.1. CONCLUSIONS

This chapter summarises the main conclusions from this study and recommends direction for

future study.

Overall, the biofouling cleaning tests have suggested that FNA is more efficient at biomass

removal and inactivation than NaOH (pH 11). 50 mg NO2-N/L at pH 3 (corresponding to 35 mg

HNO2-N/L) was the optimum cleaning conditions for biofouling. Achieving higher biomass

removal than NaOH for both heavily fouled (86-96% versus 41-83%) and moderately fouled

(94-95% against 86-92%) membranes, respectively. In accordance to ATP results, live/dead

staining results have revealed that less viable cells remain in biofouling layers after FNA

cleaning. For moderately fouled RO membranes, 6-32% of total cells remained activated under

the impact of FNA at pH 3, whereas 38-58% of viable cells remained in the biofilm cleaned by

NaOH. However, the difference between the efficiency of FNA and NaOH at protein (67-93%

versus 59-60%) and polysaccharides (73-87% versus 62-79%) removal was not as significant as

biomass removal for moderate fouled RO membranes. In addition, FISH analysis has

demonstrated that FNA had effectively reduced the quantities of targeted bacteria deposited on

heavily and moderately fouled RO membranes. The results of FISH analysis also revealed that

Betaproteobacteria and archaea were greatly affected by FNA cleaning. The benefit of using

FNA is more obvious for cleaning moderately fouled membranes, implying that early cleaning

for biofouling is preferable. Although applying FNA alone, or combining FNA and H2O2 have

shown better efficiency at biofouling removal than NaOH, the cleaning efficiency of FNA has

not been significantly improved (<1% of enhancement).

Descaling tests have shown that FNA was as efficient as HCl and citric acid for scaling control.

Elemental analyses, such as ICP-OES and SEM-EDS, showed that the scaling layer with a

calcium concentration of 32.4±1.7 g/cm2 was completely removed by the FNA solutions at pH 2

and 3 adjusted with HCl or citric acid and these standard acids alone. Based on the outcomes of

biofouling and scaling cleaning tests, FNA has been shown to be a promising cleaning agent to

achieve biofouling and scaling removal at a single stage.

69

Although the characteristics of fouling layers formed in filtration tests are different from the

fouling matrix used in cleaning tests, similar observation was obtained as FNA is more efficient

at biomass inactivation and removal. The biomass contents and viable cells of the fouling layers

formed in the experiment filtration (with FNA cleaning) unit were less than half of that in the

control filtration unit. Moreover, the results of live/dead staining revealed the abundance of

viable cells in the control unit (57±5%) was four times higher than that in the experiment unit

(13±2%). However, the concentration of proteins and polysaccharides were similar in both

control and experiment filtration units (referring to Table 5-3). In term of hydraulic performances,

FNA cleaning showed the ability to restore the permeability of RO membranes for a short period,

but the overall result showed that FNA cleanings were unable to prevent the decline of NP and

salt rejection, and consequently the accumulation of fouling on RO membranes.

§ 6.2. RECOMMENDATIONS

This research work also opens up several interesting directions that can be taken for further

research. These include:

Protocol optimisation for RO biofilm cleaning

The cleaning protocol used in this study was very basic. Different concentrations of FNA

cleaning solutions were the only variables, operating parameters such as cross flow velocity

(CFV, 0.1 m/s), cleaning time (24 hr) and temperature (20℃) were all applied as constants.

Optimising operational conditions could potentially improve the cleaning efficiency of FNA and

simplify the cleaning procedures. The optimal CFV should provide sufficient shear force to

detach biofilm, and also allow the cleaning solution to fully contact with biofouling layers [50].

The effects of CFV on the accumulation of foulants have been critically reviewed by Dreszer et

al. [50], it remains unclear whether increasing CFV can facilitate (increasing nutrient load) or

hinder (enhancing shear force) the biofilm accumulation. For the purpose of optimising cleaning

protocol, actual experiments need to be conducted to justify the effects of CFV.

Regarding the cleaning time for biofouling removal, Jiang et al. have reported that 6-24 h

exposure time would be sufficient for sewer biofilms inactivation at lab-scale [150]. Protocol

optimisation should test if cleaning efficiency of FNA presented in this study could be achieved

70

less than 24 h.

The effect of cleaning temperature on cleaning efficiency for membranes was generally studied

at a temperature range of 20–35 °C. These studies showed that better cleaning efficiency was

obtained at higher temperature (35 °C), as the higher temperature led to higher reaction rate

between cleaning chemicals and fouling deposits. Moreover, higher cleaning temperature

facilitated the transport of foulants from the fouling layer to the bulk solution [151, 152]. The

cleaning efficiency of FNA at a higher temperature (e.g. 35 °C) should be studied in future

research and compared to caustic cleaning under similar temperatures.

Protocols validation for both cleaning and prevention at large scale

Due to randomly distributed amorphous biofouling or general fouling on the surface of RO

membranes [3]. The small filtration cells (referring to Figure 3-1) were unable to reveal the

hydraulic performances (permeability and salt rejection) of RO under impacts of different

cleaning solutions. After the cleaning protocol is fully optimised, it is recommended to conduct

cleaning tests and hydraulic measurements at large-scale filtration units.

In the short-term biofilm prevention tests (referring to Chapter 5), the impacts of FNA weekly

cleaning on fouling accumulation have been demonstrated in terms of permeability and salt

rejection. However, the pressure changes of filtration units were not representative for full scale

RO module due to the size of filtration cells. Membrane manufacturers has revealed fouling can

be detected by monitoring permeability, salt rejection and pressure changes [71, 143]. Hence, it

is still important to monitor pressure changes (either over the feed channel or through the

membranes) during the filtration tests under the impacts of FNA. In order to validate prevention

protocol using FNA, pressure changes should monitor along with permeability and salt rejection

at a larger scale.

Study the cleaning mechanisms between FNA cleaning solutions and RO biofilm

The impacts of FNA on RO biofouling were inferred from the mechanisms behind the

applications of FNA on sewer anaerobic biofilm and waste activated sludge [14-18]. The

cleaning mechanisms between FNA and RO biofilm are yet to be fully revealed, highlighting the

71

need for quantitative molecular tools to monitor the abundance and the dynamics of RO biofilm

under the impacts of FNA cleaning solutions. For instance, EPS can be extracted from RO

biofilms and its interaction with FNA can be studied individually. New FISH probes which are

not limiting to the proteobacteria should also be applied for further microbial communities study.

Hence, the phylogenetic diversity in the biofilm and interaction between FNA and each bacterial

group can be fully investigated.

72

REFERENCES

[1] K. Scott, Handbook of industrial membranes, Elsevier, 1995.

[2] R. Franks, C.R. Bartels, K. Andes–Hydranautics, O.C.W. District, T.X.Y.-K. Seghers,

Implementing energy saving RO technology in large scale wastewater treatment plants, in:

Proceedings of the International Desalination and Water Reuse Conference, Las Palmas, Spain,

2007.

[3] S. Jamaly, N.N. Darwish, I. Ahmed, S.W. Hasan, A short review on reverse osmosis

pretreatment technologies, Desalination, 354 (2014) 30-38.

[4] A. Płatkowska-Siwiec, M. Bodzek, The influence of membrane and water properties on

fouling during ultrafiltration, Desalination and Water Treatment, 35 (2011) 235-241.

[5] H.-C. Flemming, P.S. Murthy, R. Venkatesan, K. Cooksey, Marine and industrial biofouling,

Springer, 2009.

[6] H.C. Flemming, Biofouling in water systems – cases, causes and countermeasures, Applied

Microbiology and Biotechnology, 59 (2002) 629-640.

[7] M. Herzberg, S. Kang, M. Elimelech, Role of Extracellular Polymeric Substances (EPS) in

Biofouling of Reverse Osmosis Membranes, Environmental Science & Technology, 43 (2009)

4393-4398.

[8] S.A. Creber, J.S. Vrouwenvelder, M.C.M. van Loosdrecht, M.L. Johns, Chemical cleaning of

biofouling in reverse osmosis membranes evaluated using magnetic resonance imaging, Journal

of Membrane Science, 362 (2010) 202-210.

[9] H.C. Flemming, G. Schaule, T. Griebe, J. Schmitt, A. Tamachkiarowa, Biofouling—the

Achilles heel of membrane processes, Desalination, 113 (1997) 215-225.

[10] J.S. Vrouwenvelder, J.C. Kruithof, M.C.M. Van Loosdrecht, Integrated approach for

biofouling control, Water Science and Technology, 62 (2010) 2477-2490.

[11] D.E. Potts, R.C. Ahlert, S.S. Wang, A critical review of fouling of reverse osmosis

membranes, Desalination, 36 (1981) 235-264.

[12] R.A. Al-Juboori, T. Yusaf, Biofouling in RO system: Mechanisms, monitoring and

controlling, Desalination, 302 (2012) 1-23.

[13] H.-C. Flemming, Microbial Biofouling: Unsolved Problems, Insufficient Approaches, and

Possible Solutions, in: H.-C. Flemming, J. Wingender, U. Szewzyk (Eds.) Biofilm Highlights,

Springer Berlin Heidelberg, 2011, pp. 81-109.

[14] G. Jiang, O. Gutierrez, Z. Yuan, The strong biocidal effect of free nitrous acid on anaerobic

sewer biofilms, Water Research, 45 (2011) 3735-3743.

[15] G. Jiang, Z. Yuan, Synergistic inactivation of anaerobic wastewater biofilm by free nitrous

acid and hydrogen peroxide, Journal of hazardous materials, 250 (2013) 91-98.

[16] M. Pijuan, L. Ye, Z. Yuan, Free nitrous acid inhibition on the aerobic metabolism of poly-

phosphate accumulating organisms, Water Research, 44 (2010) 6063-6072.

[17] L. Ye, M. Pijuan, Z. Yuan, The effect of free nitrous acid on the anabolic and catabolic

processes of glycogen accumulating organisms, Water Research, 44 (2010) 2901-2909.

[18] T. Zhang, Q. Wang, J. Khan, Z. Yuan, Free nitrous acid breaks down extracellular

polymeric substances in waste activated sludge, RSC Advances, 5 (2015) 43312-43318.

[19] C. Liu, S. Caothien, J. Hayes, T. Caothuy, T. Otoyo, T. Ogawa, Membrane chemical

cleaning: from art to science, Pall Corporation, Port Washington, NY, 11050 (2001).

[20] J. Pinnekamp, H. Friedrich, Membrane technology for waste water treatment, FiW Verlag,

2006.

73

[21] D. Lapierre, Water purification, in, Google Patents, 1987.

[22] M. Mulder, Basic principles of membrane technology, Springer Science & Business Media,

1996.

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

desalination—Development to date and future potential, Journal of Membrane Science, 370

(2011) 1-22.

[24] H. Lonsdale, The growth of membrane technology, Journal of Membrane Science, 10 (1982)

81-181.

[25] M. Cheryan, Ultrafiltration and microfiltration handbook, CRC press, 1998.

[26] M.M. Pendergast, E.M. Hoek, A review of water treatment membrane nanotechnologies,

Energy & Environmental Science, 4 (2011) 1946-1971.

[27] J. Kucera, Wiley-Scrivener : Reverse Osmosis : Design, Processes, and Applications for

Engineers, Wiley-Scrivener, Hoboken, NJ, USA, 2010.

[28] R.J. Petersen, Composite reverse osmosis and nanofiltration membranes, Journal of

membrane science, 83 (1993) 81-150.

[29] I.L. Alsvik, M.-B. Hägg, Pressure retarded osmosis and forward osmosis membranes:

materials and methods, Polymers, 5 (2013) 303-327.

[30] A. Simon, L.D. Nghiem, P. Le-Clech, S.J. Khan, J.E. Drewes, Effects of membrane

degradation on the removal of pharmaceutically active compounds (PhACs) by NF/RO filtration

processes, Journal of Membrane Science, 340 (2009) 16-25.

[31] S.S. Sablani, M.F.A. Goosen, R. Al-Belushi, M. Wilf, Concentration polarization in

ultrafiltration and reverse osmosis: a critical review, Desalination, 141 (2001) 269-289.

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

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

[33] J. Baker, L. Dudley, Biofouling in membrane systems—a review, Desalination, 118 (1998)

81-89.

[34] Q. Li, M. Elimelech, Organic Fouling and Chemical Cleaning of Nanofiltration

Membranes:  Measurements and Mechanisms, Environmental Science & Technology, 38 (2004)

4683-4693.

[35] M.M.T. Khan, P.S. Stewart, D.J. Moll, W.E. Mickols, S.E. Nelson, A.K. Camper,

Characterization and effect of biofouling on polyamide reverse osmosis and nanofiltration

membrane surfaces, Biofouling, 27 (2011) 173-183.

[36] 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.

[37] A. Antony, J.H. Low, S. Gray, A.E. Childress, P. Le-Clech, G. Leslie, Scale formation and

control in high pressure membrane water treatment systems: A review, Journal of Membrane

Science, 383 (2011) 1-16.

[38] A. Rahardianto, W.-Y. Shih, R.-W. Lee, Y. Cohen, Diagnostic characterization of gypsum

scale formation and control in RO membrane desalination of brackish water, Journal of

Membrane Science, 279 (2006) 655-668.

[39] E.M.V. Hoek, A.S. Kim, M. Elimelech, Influence of Crossflow Membrane Filter Geometry

and Shear Rate on Colloidal Fouling in Reverse Osmosis and Nanofiltration Separations,

Environmental Engineering Science, 19 (2002) 357-372.

[40] T. Nguyen, F.A. Roddick, L. Fan, Biofouling of water treatment membranes: a review of the

underlying causes, monitoring techniques and control measures, Membranes, 2 (2012) 804-840.

74

[41] X. Chen, S.R. Suwarno, T.H. Chong, D. McDougald, S. Kjelleberg, Y. Cohen, A.G. Fane,

S.A. Rice, Dynamics of biofilm formation under different nutrient levels and the effect on

biofouling of a reverse osmosis membrane system, Biofouling, 29 (2013) 319-330.

[42] E.M. Hoek, M. Elimelech, Cake-enhanced concentration polarization: a new fouling

mechanism for salt-rejecting membranes, Environmental science & technology, 37 (2003) 5581-

5588.

[43] H.-C. Flemming, Reverse osmosis membrane biofouling, Experimental thermal and fluid

science, 14 (1997) 382-391.

[44] E. Curcio, X. Ji, G. Di Profio, A.O. Sulaiman, E. Fontananova, E. Drioli, Membrane

distillation operated at high seawater concentration factors: Role of the membrane on CaCO3

scaling in presence of humic acid, Journal of Membrane Science, 346 (2010) 263-269.

[45] C. Park, Y.H. Lee, S. Lee, S. Hong, Effect of cake layer structure on colloidal fouling in

reverse osmosis membranes, Desalination, 220 (2008) 335-344.

[46] J.S. Vrouwenvelder, M.C.M. Van Loosdrecht, J.C. Kruithof, A novel scenario for

biofouling control of spiral wound membrane systems, Water Research, 45 (2011) 3890-3898.

[47] H.-C. Flemming, G. Schaule, H.F. Ridgway, Living with biofilms-elements of an integrated

anti-fouling strategy, Annu Rev Microbiol, 41 (2001) 435-464.

[48] T. Mattila‐Sandholm, G. Wirtanen, Biofilm formation in the industry: A review, Food

Reviews International, 8 (1992) 573-603.

[49] S. Pandey, V. Jegatheesan, K. Baskaran, L. Shu, Fouling in reverse osmosis (RO)

membrane in water recovery from secondary effluent: a review, Reviews in Environmental

Science and Bio/Technology, 11 (2012) 125-145.

[50] C. Dreszer, H.C. Flemming, A. Zwijnenburg, J.C. Kruithof, J.S. Vrouwenvelder, Impact of

biofilm accumulation on transmembrane and feed channel pressure drop: Effects of crossflow

velocity, feed spacer and biodegradable nutrient, Water Research, 50 (2014) 200-211.

[51] S.R. Suwarno, X. Chen, T.H. Chong, D. McDougald, Y. Cohen, S.A. Rice, A.G. Fane,

Biofouling in reverse osmosis processes: The roles of flux, crossflow velocity and concentration

polarization in biofilm development, Journal of Membrane Science, 467 (2014) 116-125.

[52] D. Monroe, Looking for Chinks in the Armor of Bacterial Biofilms, PLoS Biol, 5 (2007)

e307.

[53] S.U. Gerbersdorf, B. Westrich, D.M. Paterson, Microbial extracellular polymeric substances

(EPS) in fresh water sediments, Microbial ecology, 58 (2009) 334-349.

[54] P. Stoodley, K. Sauer, D.G. Davies, J.W. Costerton, BIOFILMS AS COMPLEX

DIFFERENTIATED COMMUNITIES, Annual Review of Microbiology, 56 (2002) 187-209.

[55] S. Andersson, Characterization of bacterial biofilms for wastewater treatment, (2009).

[56] 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.

[57] H.-C. Flemming, Biofilms and environmental protection, Water Science & Technology, 27

(1993) 1-10.

[58] H. Flemming, J. Wingender, Relevance of microbial extracellular polymeric substances

(EPSs)-Part I: Structural and ecological aspects, Water Science & Technology, 43 (2001) 1-8.

[59] H.C. Flemming, J. Wingender, Extracellular Polymeric Substances (EPS): Structural,

Ecological and Technical Aspects, in: Encyclopedia of Environmental Microbiology, John

Wiley & Sons, Inc., 2003.

75

[60] G.M. Wolfaardt, J.R. Lawrence, R.D. Robarts, D.E. Caldwell, In situ Characterization of

Biofilm Exopolymers Involved in the Accumulation of Chlorinated Organics, Microbial ecology,

35 (1998) 213-223.

[61] H.-C. Flemming, T.R. Neu, D.J. Wozniak, The EPS Matrix: The “House of Biofilm Cells”,

Journal of Bacteriology, 189 (2007) 7945-7947.

[62] D.G. Allison, T. Maira-Litran, and P. Gilbert, Antimicrobial resistance of biofilms, in: L.V.

Evans (Ed.) In Biofilms: recent advances in their study and control, Harwood Academic

Publishers, Amsterdam, 2000, pp. 149-166.

[63] A. Conrad, M. Kontro, M.M. Keinänen, A. Cadoret, P. Faure, L. Mansuy-Huault, J.-C.

Block, Fatty acids of lipid fractions in extracellular polymeric substances of activated sludge

flocs, Lipids, 38 (2003) 1093-1105.

[64] P. Stoodley, K. Sauer, D. Davies, J.W. Costerton, Biofilms as complex differentiated

communities, Annual Reviews in Microbiology, 56 (2002) 187-209.

[65] I.W. Sutherland, Biofilm exopolysaccharides: a strong and sticky framework, Microbiology,

147 (2001) 3-9.

[66] S.S. Adav, D.-J. Lee, J.-H. Tay, Extracellular polymeric substances and structural stability

of aerobic granule, Water Research, 42 (2008) 1644-1650.

[67] A. Rahardianto, J. Gao, C.J. Gabelich, M.D. Williams, Y. Cohen, High recovery membrane

desalting of low-salinity brackish water: Integration of accelerated precipitation softening with

membrane RO, Journal of Membrane Science, 289 (2007) 123-137.

[68] J. Gilron, D. Hasson, Calcium sulphate fouling of reverse osmosis membranes: flux decline

mechanism, Chemical Engineering Science, 42 (1987) 2351-2360.

[69] S. Lee, C. Lee, Scale formation in NF/RO: mechanism and control, Water Science &

Technology, 51 (2005) 267-275.

[70] N. Prihasto, Q.-F. Liu, S.-H. Kim, Pre-treatment strategies for seawater desalination by

reverse osmosis system, Desalination, 249 (2009) 308-316.

[71] Dow, FILMTEC™ Reverse Osmosis Membranes Technical Manual, in, 2008.

[72] A. Plottu-Pecheux, B. Houssais, C. Democrate, D. Gatel, C. Parron, J. Cavard, Comparison

of three antiscalants, as applied to the treatment of water from the River Oise, Desalination, 145

(2002) 273-280.

[73] E.H. Kelle Zeiher, B. Ho, K.D. Williams, Novel antiscalant dosing control, Desalination,

157 (2003) 209-216.

[74] J.S. Vrouwenvelder, S.A. Manolarakis, H.R. Veenendaal, D. van der Kooij, Biofouling

potential of chemicals used for scale control in RO and NF membranes, Desalination, 132 (2000)

1-10.

[75] Dow, Liquid Separation, FILMTEC™ Reverse Osmosis Membranes Technical Manual, in,

2008.

[76] HYDRANAUTICS, RO Membrane Foulants and Their Removal from Composite

Polyamide (ESPA, ESNA, CPA, LFC, and SWC) RO Membrane Elements, in, May 2000.

[77] A. Al-Amoudi, R.W. Lovitt, Fouling strategies and the cleaning system of NF membranes

and factors affecting cleaning efficiency, Journal of Membrane Science, 303 (2007) 4-28.

[78] D. Kim, S. Jung, J. Sohn, H. Kim, S. Lee, Biocide application for controlling biofouling of

SWRO membranes—an overview, Desalination, 238 (2009) 43-52.

[79] V. Lund, K. Ormerod, The influence of disinfection processes on biofilm formation in water

distribution systems, Water research, 29 (1995) 1013-1021.

76

[80] J. Yu, Y. Baek, H. Yoon, J. Yoon, New disinfectant to control biofouling of polyamide

reverse osmosis membrane, Journal of Membrane Science, 427 (2013) 30-36.

[81] N.P. Soice, A.C. Maladono, D.Y. Takigawa, A.D. Norman, W.B. Krantz, A.R. Greenberg,

Oxidative degradation of polyamide reverse osmosis membranes: studies of molecular model

compounds and selected membranes, Journal of applied polymer science, 90 (2003) 1173-1184.

[82] G.-D. Kang, C.-J. Gao, W.-D. Chen, X.-M. Jie, Y.-M. Cao, Q. Yuan, Study on hypochlorite

degradation of aromatic polyamide reverse osmosis membrane, Journal of membrane science,

300 (2007) 165-171.

[83] Y.-N. Kwon, J.O. Leckie, Hypochlorite degradation of crosslinked polyamide membranes:

II. Changes in hydrogen bonding behavior and performance, Journal of membrane science, 282

(2006) 456-464.

[84] C. Bartels, M. Wilf, K. Andes, J. Iong, Design considerations for wastewater treatment by

reverse osmosis, Water Science & Technology, 51 (2005) 473-482.

[85] M.K. da Silva, I.C. Tessaro, K. Wada, Investigation of oxidative degradation of polyamide

reverse osmosis membranes by monochloramine solutions, Journal of Membrane Science, 282

(2006) 375-382.

[86] G. Petrucci, M. Rosellini, Chlorine dioxide in seawater for fouling control and post-

disinfection in potable waterworks, Desalination, 182 (2005) 283-291.

[87] S. Ebrahim, Cleaning and regeneration of membranes in desalination and wastewater

applications: State-of-the-art, Desalination, 96 (1994) 225-238.

[88] W.S. Ang, N.Y. Yip, A. Tiraferri, M. Elimelech, Chemical cleaning of RO membranes

fouled by wastewater effluent: Achieving higher efficiency with dual-step cleaning, Journal of

Membrane Science, 382 (2011) 100-106.

[89] N. Porcelli, S. Judd, Chemical cleaning of potable water membranes: a review, Separation

and Purification Technology, 71 (2010) 137-143.

[90] HYDRANAUTICS, Biocides for Disinfection and Storage of Hydranautics Membrane

Elements, in, March 2015.

[91] W.A.M. Hijnen, C. Castillo, A.H. Brouwer-Hanzens, D.J.H. Harmsen, E.R. Cornelissen, D.

van der Kooij, Quantitative assessment of the efficacy of spiral-wound membrane cleaning

procedures to remove biofilms, Water Research, 46 (2012) 6369-6381.

[92] C. Frayne, The selection and application of nonoxidizing biocides for cooling water systems,

The Analyst (Spring), (2001) 1-10.

[93] J.B. Kristensen, R.L. Meyer, B.S. Laursen, S. Shipovskov, F. Besenbacher, C.H. Poulsen,

Antifouling enzymes and the biochemistry of marine settlement, Biotechnology advances, 26

(2008) 471-481.

[94] B. Thallinger, E.N. Prasetyo, G.S. Nyanhongo, G.M. Guebitz, Antimicrobial enzymes: An

emerging strategy to fight microbes and microbial biofilms, Biotechnology Journal, 8 (2013) 97-

109.

[95] J.S. Webb, L.S. Thompson, S. James, T. Charlton, T. Tolker-Nielsen, B. Koch, M. Givskov,

S. Kjelleberg, Cell death in Pseudomonas aeruginosa biofilm development, Journal of

bacteriology, 185 (2003) 4585-4592.

[96] N.N.E. Greenwood, Alan, Chemistry of the Elements, Butterworth-Heinemann, 1997.

[97] A.C. Anthonisen, R.C. Loehr, T.B.S. Prakasam, E.G. Shinath, Inhibition of nitrification by

ammonia and nitrous acid, J. Water Pollut. Control Fed, 48 (1976) 835-852.

[98] A. Anthonisen, R. Loehr, T. Prakasam, E. Srinath, Inhibition of nitrification by ammonia

and nitrous acid, Journal (Water Pollution Control Federation), (1976) 835-852.

77

[99] V.M. Vadivelu, J. Keller, Z. Yuan, Effect of free ammonia and free nitrous acid

concentration on the anabolic and catabolic processes of an enriched Nitrosomonas culture,

Biotechnology and Bioengineering, 95 (2006) 830-839.

[100] V.M. Vadivelu, Z. Yuan, C. Fux, J. Keller, The Inhibitory Effects of Free Nitrous Acid on

the Energy Generation and Growth Processes of an Enriched Nitrobacter Culture, Environmental

Science & Technology, 40 (2006) 4442-4448.

[101] L. Ye, M. Pijuan, Z. Yuan, The effect of free nitrous acid on key anaerobic processes in

enhanced biological phosphorus removal systems, Bioresource Technology, 130 (2013) 382-389.

[102] Y. Zhou, L. Ganda, M. Lim, Z. Yuan, W.J. Ng, Response of poly-phosphate accumulating

organisms to free nitrous acid inhibition under anoxic and aerobic conditions, Bioresource

Technology, 116 (2012) 340-347.

[103] M. Pijuan, Q. Wang, L. Ye, Z. Yuan, Improving secondary sludge biodegradability using

free nitrous acid treatment, Bioresource Technology, 116 (2012) 92-98.

[104] Q. Wang, L. Ye, G. Jiang, S. Hu, Z. Yuan, Side-stream sludge treatment using free nitrous

acid selectively eliminates nitrite oxidizing bacteria and achieves the nitrite pathway, Water

Research, 55 (2014) 245-255.

[105] X. Sun, G. Jiang, P.L. Bond, J. Keller, Z. Yuan, A novel and simple treatment for control

of sulfide induced sewer concrete corrosion using free nitrous acid, Water Research, 70 (2015)

279-287.

[106] E.F. B. Donose, W. Gernjak, M. Pidou and Z. Yuan, Preliminary results of the membrane

biofouling removal with Free Nittous Acid (FNA), in, 2013.

[107] Y. Zhou, A. Oehmen, M. Lim, V. Vadivelu, W.J. Ng, The role of nitrite and free nitrous

acid (FNA) in wastewater treatment plants, Water Research, 45 (2011) 4672-4682.

[108] T. Casey, M. Wentzel, G. Ekama, Filamentous organism bulking in nutrient removal

activated sludge systems: Paper 9: Review of biochemistry of heterotrophic respiratory

metabolism, Water SA, 25 (1999) 409-424.

[109] M. Repetto, J. Semprine, A. Boveris, Lipid Peroxidation: Chemical mechanism, biological

implications and analytical determinations, Lipid peroxidation, (2012) 1-28.

[110] W. Heaselgrave, P.W. Andrew, S. Kilvington, Acidified nitrite enhances hydrogen

peroxide disinfection of Acanthamoeba, bacteria and fungi, Journal of Antimicrobial

Chemotherapy, (2010).

[111] F. Du, S. Freguia, Z. Yuan, J. Keller, I. Pikaar, Enhancing Toxic Metal Removal from

Acidified Sludge with Nitrite Addition, Environmental Science & Technology, 49 (2015) 6257-

6263.

[112] M. Greenblatt, S. Mirvish, B.T. So, Nitrosamine studies: Induction of lung adenomas by

concurrent administration of sodium nitrite and secondary amines in Swiss mice, Journal of the

National Cancer Institute, 46 (1971) 1029-1034.

[113] U. EPA, Guidelines for carcinogen risk assessment. Retrieved from, in, 1986.

[114] E. Ireland, Water Treatment Manual: Disinfection, in, 2011.

[115] E. Linley, S.P. Denyer, G. McDonnell, C. Simons, J.-Y. Maillard, Use of hydrogen

peroxide as a biocide: new consideration of its mechanisms of biocidal action, Journal of

Antimicrobial Chemotherapy, 67 (2012) 1589-1596.

[116] DOW, Cleaning Procedures for DOW FILMTEC FT30 Elements, in.

[117] C. Bele, Y. Kumar, T. Walker, Y. Poussade, V. Zavlanos, Operating boundaries of full-

scale advanced water reuse treatment plants: many lessons learned from pilot plant experience,

Water Science and Technology, 62 (2010) 1560-1566.

78

[118] B. Alberts, D. Bray, K. Hopkin, A. Johnson, J. Lewis, M. Raff, K. Roberts, P. Walter,

Essential cell biology, Garland Science, 2013.

[119] F. Hammes, F. Goldschmidt, M. Vital, Y. Wang, T. Egli, Measurement and interpretation

of microbial adenosine tri-phosphate (ATP) in aquatic environments, Water Research, 44 (2010)

3915-3923.

[120] R.I. Amann, B.J. Binder, R.J. Olson, S.W. Chisholm, R. Devereux, D.A. Stahl,

Combination of 16S rRNA-targeted oligonucleotide probes with flow cytometry for analyzing

mixed microbial populations, Applied and environmental microbiology, 56 (1990) 1919-1925.

[121] G. Talbot, E. Topp, M.F. Palin, D.I. Massé, Evaluation of molecular methods used for

establishing the interactions and functions of microorganisms in anaerobic bioreactors, Water

Research, 42 (2008) 513-537.

[122] R. Amann, Fluorescently labelled, rRNA‐targeted oligonucleotide probes in the study of

microbial ecology, Molecular Ecology, 4 (1995) 543-554.

[123] L.A. Bereschenko, H. Prummel, G.J.W. Euverink, A.J.M. Stams, M.C.M. van Loosdrecht,

Effect of conventional chemical treatment on the microbial population in a biofouling layer of

reverse osmosis systems, Water Research, 45 (2011) 405-416.

[124] C.M. Pang, W.-T. Liu, Community structure analysis of reverse osmosis membrane

biofilms and the significance of Rhizobiales bacteria in biofouling, Environmental Science &

Technology, 41 (2007) 4728-4734.

[125] C.-L. Chen, W.-T. Liu, M.-L. Chong, M.-T. Wong, S. Ong, H. Seah, W. Ng, Community

structure of microbial biofilms associated with membrane-based water purification processes as

revealed using a polyphasic approach, Applied Microbiology and Biotechnology, 63 (2004) 466-

473.

[126] J. Lee, I.S. Kim, Microbial community in seawater reverse osmosis and rapid diagnosis of

membrane biofouling, Desalination, 273 (2011) 118-126.

[127] A. Al Ashhab, O. Gillor, M. Herzberg, Biofouling of reverse-osmosis membranes under

different shear rates during tertiary wastewater desalination: Microbial community composition,

Water Research, 67 (2014) 86-95.

[128] C. Ayache, C. Manes, M. Pidou, J.P. Croué, W. Gernjak, Microbial community analysis of

fouled reverse osmosis membranes used in water recycling, Water Research, 47 (2013) 3291-

3299.

[129] A. Al Ashhab, M. Herzberg, O. Gillor, Biofouling of reverse-osmosis membranes during

tertiary wastewater desalination: Microbial community composition, Water Research, 50 (2014)

341-349.

[130] W. Manz, R. Amann, W. Ludwig, M. Wagner, K.-H. Schleifer, Phylogenetic

oligodeoxynucleotide probes for the major subclasses of proteobacteria: problems and solutions,

Systematic and applied microbiology, 15 (1992) 593-600.

[131] J. Brosius, T.J. Dull, D.D. Sleeter, H.F. Noller, Gene organization and primary structure of

a ribosomal RNA operon from< i> Escherichia coli</i>, Journal of molecular biology, 148 (1981)

107-127.

[132] L. Nieradzik, Biofouling removal from RO membranes, in, University of Duisburg-Essen,

2012.

[133] M. Dubois, K. Gilles, J. Hamilton, P. Rebers, F. Smith, A colorimetric method for the

determination of sugars, (1951).

79

[134] P. Smith, R.I. Krohn, G. Hermanson, A. Mallia, F. Gartner, M. Provenzano, E. Fujimoto,

N. Goeke, B. Olson, D. Klenk, Measurement of protein using bicinchoninic acid, Analytical

biochemistry, 150 (1985) 76-85.

[135] J. Vrouwenvelder, S. Manolarakis, H. Veenendaal, D. Van der Kooij, Biofouling potential

of chemicals used for scale control in RO and NF membranes, Desalination, 132 (2000) 1-10.

[136] J.S. Vrouwenvelder, D. van der Kooij, Diagnosis, prediction and prevention of biofouling

of NF and RO membranes, Desalination, 139 (2001) 65-71.

[137] H.-C. Flemming, Mechanistic aspects of reverse osmosis membrane biofouling and

prevention, (1993).

[138] E.L. Farias, K.J. Howe, B.M. Thomson, Spatial and temporal evolution of organic foulant

layers on reverse osmosis membranes in wastewater reuse applications, Water research, 58 (2014)

102-110.

[139] M.T. Khan, M. Busch, V.G. Molina, A.-H. Emwas, C. Aubry, J.-P. Croue, How different

is the composition of the fouling layer of wastewater reuse and seawater desalination RO

membranes?, Water Research, 59 (2014) 271-282.

[140] W. Fang, J. Hu, S. Ong, Influence of phosphorus on biofilm formation in model drinking

water distribution systems, Journal of applied microbiology, 106 (2009) 1328-1335.

[141] RepetaGroup, Speciation of dissolved organic phosphorus in seawater, in, February 11,

2014.

[142] Dow, Liquid Separations, FILMTECTM

Reverse Osmosis Membranes. Technical manual.

Form no. 609-00771-0705:1-182., in, 2008.

[143] LENNTECH, Foulants and Cleaning Procedures for composite polyamide RO Membrane

Elements (ESPA, ESNA, CPA, LFC, and SWC), in, Octorber 2011.

[144] M. Pijuan, Q. Wang, L. Ye, Z. Yuan, Improving secondary sludge biodegradability using

free nitrous acid treatment Bioresource Technology, 116 (2013) 92-98.

[145] T. Schipolowski, A. Jeżowska, G. Wozny, Reliability of membrane test cell measurements,

Desalination, 189 (2006) 71-80.

[146] L. Nieradzik, Biofouling removal from RO membranes, in, University of Duisburg-Essen,

Essen, 2012, pp. 96.

[147] C. Bartels, R. Franks, S. Rybar, M. Schierach, M. Wilf, The effect of feed ionic strength on

salt passage through reverse osmosis membranes, Desalination, 184 (2005) 185-195.

[148] Lewabrane, Principles of Reverse Osmosis Membrane separation, in, 2012.

[149] J. Vrouwenvelder, D. Van der Kooij, Diagnosis of fouling problems of NF and RO

membrane installations by a quick scan, Desalination, 153 (2003) 121-124.

[150] G. Jiang, A. Keating, S. Corrie, K. O'Halloran, L. Nguyen, Z. Yuan, Dosing free nitrous

acid for sulfide control in sewers: Results of field trials in Australia, Water Research, 47 (2013)

4331-4339.

[151] L.H. Kim, A. Jang, H.-W. Yu, S.-J. Kim, I.S. Kim, Effect of chemical cleaning on

membrane biofouling in seawater reverse osmosis processes, Desalination and Water Treatment,

33 (2011) 289-294.

[152] M.R. Sohrabi, S.S. Madaeni, M. Khosravi, A.M. Ghaedi, Chemical cleaning of reverse

osmosis and nanofiltration membranes fouled by licorice aqueous solutions, Desalination, 267

(2011) 93-100.

80

APPENDIXES

Appendix A. Calibration curve for ATP measurement

Figure A-1. Calibration curve for ATP quantification. Error bars show the standard deviation

on n measurements (n= [2-14]). The detection limit was defined as the lowest ATP

concentration of which a three-fold error bar extension downward did not overlap with a

three-fold error bar extension of the negative control, and was set at 0.01 nM ATP, upwards

of which a linear correlation of R2=0.99 was obtained

y = 13397x + 10181 R² = 0.9943

1.E+02

1.E+03

1.E+04

1.E+05

1.E+06

1.E-03 1.E-02 1.E-01 1.E+00 1.E+01 1.E+02

Res

po

nse

(R

LU)

rATP (nM)

81

Appendix B. Additional results of FISH analysis

Table B-1: The relative quantification of each specific bacteria detected with the respective FISH

probes (n= 20±5) in each biofouling matrix.

Membrane # All Alpha-

proteobacteria

Beta-

proteobacteria

Gamma-

proteobacteria Archaea

RO1 6086±1004 2603±769 1292±387 n.d. 108±30

RO3 44844±5598 4467±693 12215±1564 n.d 3150±613

RO4 7291±632 119±40 154±51 1252±267 238±104

RO5 5752±642 720±152 804±467 477±125 1583±290

n.d.: not determined

Table B-2: The relative abundances of each specific bacteria detected with the respective FISH

probes (n=20±5) in the biofouling matrix of RO1 and RO5 before and after cleaning.

Membrane # All Alpha-

proteobacteria

Beta-

proteobacteria

Gamma-

proteobacteria Archaea

RO1-Before

cleaning 6086±1004 2603±769 1292±387 n.d. 108±30

RO1-After

cleaning 1816±360 313±122 67±25 n.d. 13±9

RO5-Before

cleaning 5752±642 720±152 804±467 477±125 1583±290

RO5-After

cleaning 3575±772 551±193 141±39 n.d. 3±2

n.d.: not determined

82

(a) (b)

(c) (d)

(e) (f)

Figure B-1. FISH images of biofilm formed on the membranes RO1 (a); RO3 (b); RO4(c and

d) and RO5 (e and f). In images (a), (b), (c), and (e), bacteria (EUBmix) are in blue,

alphaproteobacteria are in pink, betaproteobacteria are in cyan and archaea in green. For (d)

and (f), Gammaproteobacteria are in pink. 20±5 FISH images were taken for this analysis.

83

(a)

(b)

(c)

(d)

Figure B-2. FISH images of biofilm formed on membranes RO1 before (a) and after cleaning

tests (b); RO5 before (c) and after cleaning tests (d), respectively. Bacteria (EUBmix) are in

blue, Gammaproteobacteria are in pink, Betaproteobacteria are in cyan and archaea in green.

A cross-flow velocity 0.1 m/s was applied. 20±5 FISH images were taken for this analysis.

84

Appendix C. Live/dead cells represented by CLSM images

(a) (d)

(b) (e)

(c) (f)

Figure C-1. CLSM images of RO4, before cleaning (a) and after cleaning with Control

(Water) (b), NaOH, pH 11(c), FNA, pH 2 (d), FNA, pH 3 (e) and FNA, pH 4 (f). Standard

test conditions: a cross-flow velocity 0.1 m/s was applied. Live cells were stained in green

with the DNA specific dye SYTO®9 and the dead cells are stained in red with PI.

85

(a) (d)

(b) (e)

(c) (f)

Figure C-2. CLSM images of RO5, before cleaning (a) and after cleaning with Control

(Water) (b), NaOH, pH 11(c), FNA, pH 2 (d), FNA, pH 3 (e) and FNA, pH 4 (f). Standard

test conditions: a cross-flow velocity 0.1 m/s was applied. Live cells were stained in green

with the DNA specific dye SYTO®9 and the dead cells are stained in red with PI.

86

Table C-1: The relative quantities of bacteria (live and dead) (n=35) after the filtration tests.

Control FNA

Live Cells 113148±11417 11789±1580

Dead Cells 137227±21319 103175±12171

Total Cells 250375±25482 114964±12206

Live/Total Cells (%) 57±5 13±2

87

Appendix D. ICP-OES and SEM-EDS results of descaling tests

(a) (b)

(c) (d)

(e) (f)

Figure D-1. SEM images of calcium carbonate scaled RO membrane (a) before and (b-d) after cleaning.

Test conditions: (b) Nitric Acid; (c) HCl, pH 3; (c) FNA, HCl pH 3; (d) citric acid, pH 3; (e) FNA, citric

acid pH 3; Nitrite concentration is 50 mgNO2--N/L, cross-flow velocity 0.1 m/s.

88

a) b)

c) d)

e) f)

Figure D-2. Element wt% (from SEM-EDS analysis) of calcium carbonate scaled RO membrane (a)

before and (b-d) after cleaning. Test conditions: (b) Nitric Acid; (c) HCl, pH 3; (c) FNA, HCl pH 3; (d)

citric acid, pH 3; (e) FNA, citric acid pH 3; Nitrite concentration is 50 mgNO2--N/L, cross-flow velocity

0.1 m/s, No. of analysis = 5.