forward osmosis for resource recovery from anaerobically

96
University of Wollongong Research Online University of Wollongong esis Collection 2017+ University of Wollongong esis Collections 2019 Forward osmosis for resource recovery from anaerobically digested sludge centrate Truong Minh Vu University of Wollongong Unless otherwise indicated, the views expressed in this thesis are those of the author and do not necessarily represent the views of the University of Wollongong. Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library: [email protected] Recommended Citation Vu, Truong Minh, Forward osmosis for resource recovery from anaerobically digested sludge centrate, Master of Philosophy thesis, School of Civil, Mining and Environmental Engineering, University of Wollongong, 2019. hps://ro.uow.edu.au/theses1/494

Upload: others

Post on 28-Apr-2022

2 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Forward osmosis for resource recovery from anaerobically

University of WollongongResearch Online

University of Wollongong Thesis Collection 2017+ University of Wollongong Thesis Collections

2019

Forward osmosis for resource recovery fromanaerobically digested sludge centrateTruong Minh VuUniversity of Wollongong

Unless otherwise indicated, the views expressed in this thesis are those of the author and donot necessarily represent the views of the University of Wollongong.

Research Online is the open access institutional repository for the University of Wollongong. For further information contact the UOW Library:[email protected]

Recommended CitationVu, Truong Minh, Forward osmosis for resource recovery from anaerobically digested sludge centrate, Master of Philosophy thesis,School of Civil, Mining and Environmental Engineering, University of Wollongong, 2019. https://ro.uow.edu.au/theses1/494

Page 2: Forward osmosis for resource recovery from anaerobically

School of Civil, Mining and Environmental Engineering

Faculty of Engineering and Information Sciences

University of Wollongong Australia

FORWARD OSMOSIS FOR RESOURCE RECOVERY FROM

ANAEROBICALLY DIGESTED SLUDGE CENTRATE

A thesis submitted in fulfilment of the requirements for the award of the degree of

Master of Philosophy

from

UNIVERSITY OF WOLLONGONG

By

Truong Minh Vu

February 2019

Page 3: Forward osmosis for resource recovery from anaerobically

i

CERTIFICATION

I, Truong Minh Vu, hereby declare that this thesis, submitted in fulfilment of the requirements

for the award of Master of Philosophy to the School of Civil, Mining and Environmental

Engineering, Faculty of Engineering and Information Sciences, University of Wollongong is

wholly my own work unless otherwise referenced or acknowledged. The document has not

been submitted for qualification at any other academic institution

Truong Minh Vu

February 2019

Thesis supervision:

Assoc. Prof. Faisal I. Hai, School of Civil, Mining and Environmental Engineering

UOW, Wollongong, NSW 2522. Tel: +612 42213054. Email: [email protected]

Prof. Long Nghiem, School of Civil and Environmental Engineering

UTS, Sydney, NSW 2522. Tel: +612 95142625. Email: [email protected]

Dr. Ashley J. Ansari, School of Civil, Mining and Environmental Engineering

UOW, Wollongong, NSW 2522 Tel: +612 42214590. Email. [email protected]

Page 4: Forward osmosis for resource recovery from anaerobically

ii

THESIS RELATED PUBLICATIONS

Peer-reviewed Journal Paper

1. M.T. Vu, A.J. Ansari, F.I. Hai, L.D. Nghiem, Performance of a seawater-driven forward

osmosis process for pre-concentrating sludge centrate: organic enrichment and membrane

fouling, Environmental Science: Water Research & Technology, 4 (2018) 1047-1056.

Conference Presentation

1. M.T. Vu, A.J. Ansari, F.I. Hai, L.D. Nghiem, Fouling behaviour and performance of

seawater-driven forward osmosis for pre-concentrating sludge centrate. Oral presentation at

the 11st conference of the Aseanian Membrane Society, AMS 11, Brisbane, Queensland,

Australia, 3 – 6 July 2018.

Page 5: Forward osmosis for resource recovery from anaerobically

iii

ABSTRACT

Digested sludge centrate also known as sludge centrate is a by-product of the anaerobic

digestion process obtained from sludge dewater. Sludge centrate contains most of the

nutrients and some organics from the initial sewage sludge. Nutrients, namely ammonia and

phosphorus can be extracted from sludge centrate in the forms of calcium phosphate or

struvite precipitate. However, the concentration of organic matter and nutrients in raw sludge

centrate is too low for for direct anaerobic digestion and precipitation. Thus, a novel approach

of employing seawater-driven forward osmosis to pre-concentrate organic matter and

nutrients in sludge centrate is vital to render sludge centrate anaerobically digestible and less

chemical-intensively precipitable.

This thesis aimed to demonstrate for the first time the efficiency of organic matter enrichment

in sludge centrate using a seawater-driven forward osmosis process. In particular, this

research elucidated the impacts of membrane materials, prewetting procedures and draw

solution on the performance of the seawater-driven forward osmosis process. The results

indicated that the cellulose triacetate membrane (CTA) offered better performance than the

polyamide membrane (PA) in terms of organic materials enrichment, fouling resistance and

membrane cleaning efficiency. The prewetting protocol using 70% alcohol solution led to the

significantly increased pure water flux of the PA membrane, while this step appeared to be

ineffective to improve that of the CTA membrane. Membrane fouling decreased the

enrichment efficiency of organic matter since the deposition of suspended particulate matter

on the membrane surface caused fouling and loss of organic matter from the concentrated

sludge centrate. The results showed that increasing the draw solution concentration increased

flux but did not aggravate membrane fouling, however, it could reduce the efficiency of

physical flushing to recover the flux. Seawater showed comparable forward osmosis

performance to that of analytical grade NaCl as draw solutes in terms of flux and organic

enrichment. The results also showed that seawater as the draw solution resulted in more

membrane fouling and lower flux recovery compared to NaCl.

This study also proposed and evaluated techniques to mitigate membrane fouling during

nutrient enrichment in sludge centrate by forward osmosis. Phosphorus precipitation in the

bulk feed solution and on the membrane surface was systematically quantified and compared

with respect to solution pH and operation time. The results indicated that low efficiency of

nutrients enrichment when using seawater as the draw solution was attributed to formation of

phosphorus precipitates. Moreover, increase in pH during the filtration also caused low

Page 6: Forward osmosis for resource recovery from anaerobically

iv

efficiency of ammonium enrichment. Formation of inorganic precipitates (i.e. calcium

phosphate and struvite) on the membrane surface led to severe fouling. The results also

showed a relationship among pH, duration of filtration, fouling and phosphorus precipitation.

This study demonstrated that membrane fouling could be minimized by encouraging

phosphorus precipitation in the bulk solution (avoid feed solution mixing) and increasing the

membrane surface area to shorten operation time. Furthermore, proper ratio of membrane area

over permeate volume to enrich nutrients was 175 m-1. The reduction in membrane fouling

also resulted in a proportional increase of phosphate and ammonia in the final concentrated

feed solution.

Page 7: Forward osmosis for resource recovery from anaerobically

v

ACKNOWLEDGEMENTS

I would like to express the deepest gratitude to Prof. Long Nghiem in the School of Civil and

Environmental Engineering at the University of Technology Sydney for his supervision and

support. Prof. Long Nghiem has been a very dedicated and thoughtful mentor. He has

provided me with massive helpful advice on my research. He has always taught me right from

wrong. I am grateful for all opportunities he has given me. Without him, my thesis would not

be successfully completed.

I would like to acknowledge my principal supervisor, Assoc. Prof. Faisal I. Hai in the School

of Civil, Mining and Environmental Engineering at the University of Wollongong for his

supervision and encouragement during my study here.

My special thanks are sent to my co-supervisor, Dr. Ashley J. Ansari for her support and

encouragement during the difficult times. As for me, she is like a friend rather than a

supervisor. Her stories in terms of research really inspire me and keep me motivated to pursue

my research career.

I gratefully acknowledged the University of Wollongong Australia for financial scholarship

support to assist me to carry out my research.

I would like to thank all my friends in the environmental engineering lab. They are so

friendly, kind-hearted and trustworthy. They have helped me settle down my life as well as

get on well with the new life in Australia. I will never forget what they have done for me.

Words are powerless to express my gratitude to my parents and relatives for their love and

encouragement throughout my life. From the bottom of my heart, I wish them health and

happiness.

Finally, I would like to spend the most sincere and special thanks to my wife, Ngoc (Emily),

especially for her patience, encouragement and love. Without her, my study could be never

possible.

Page 8: Forward osmosis for resource recovery from anaerobically

vi

CONTENTS

ABSTRACT ................................................................................................................... iii

ACKNOWLEDGEMENTS ............................................................................................ v

CONTENTS ................................................................................................................... vi

LIST OF FIGURES ....................................................................................................... ix

LIST OF TABLES .......................................................................................................... x

ACRONYMS ................................................................................................................ xii

CHAPTER 1. INTRODUCTION ................................................................................... 1

1.1. Background .............................................................................................................. 2

1.2. Objectives and scope of thesis ................................................................................. 3

1.3. Thesis outline ........................................................................................................... 4

CHAPTER 2. LITERATURE REVIEW ........................................................................ 6

2.1. Forward osmosis....................................................................................................... 7

2.1.1. Osmotic process ................................................................................................. 7

2.1.2. Transport phenomena ........................................................................................ 7

2.1.2.1. Mechanisms of mass transfer ...................................................................... 7

2.1.2.2. External concentration polarization ............................................................ 8

2.1.2.3. Internal concentration polarization ........................................................... 10

2.1.2.4. Reverse solute flux .................................................................................... 12

2.1.3. Materials .......................................................................................................... 14

2.1.4. Forward osmosis operation .............................................................................. 15

2.1.4.1. Temperature .............................................................................................. 15

2.1.4.2. Feed solution pH ....................................................................................... 16

2.1.4.3. Cross-flow velocities and spacer designs .................................................. 17

2.1.4.4. Draw solution ............................................................................................ 18

2.1.4.5. Membrane orientation ............................................................................... 21

2.1.4.6. Module configurations .............................................................................. 21

2.2. Fouling phenomena in the FO membrane operation .............................................. 22

2.2.1. FO membrane fouling ...................................................................................... 22

2.2.2. Colloidal and organic fouling .......................................................................... 23

Page 9: Forward osmosis for resource recovery from anaerobically

vii

2.2.2.1. Characteristics of colloids and macromolecules ....................................... 23

2.2.2.2. Colloidal interactions ................................................................................ 26

2.2.3. Inorganic fouling.............................................................................................. 26

2.2.4. Biofouling ........................................................................................................ 27

2.2.5. Factors affecting FO membrane fouling .......................................................... 28

2.2.5.1. Feedwater characteristics .......................................................................... 28

2.2.5.2. Hydrodynamic conditions ......................................................................... 29

2.2.5.3. Membrane properties ................................................................................ 30

2.2.5.4. The composition of draw solution ............................................................. 31

2.2.5.5. Membrane orientation ............................................................................... 32

2.2.6. Effects of membrane fouling on the performance of FO process.................... 32

2.2.7. Mitigation of fouling in the FO process .......................................................... 33

2.2.7.1. Fouling characterization ............................................................................ 33

2.2.7.2. Reduced-fouling methods ......................................................................... 34

2.2.7.3. Membrane fouling control ......................................................................... 38

CHAPTER 3 ................................................................................................................. 42

PERFORMANCE OF A SEAWATER-DRIVEN FORWARD OSMOSIS PROCESS

FOR PRE-CONCENTRATING SLUDGE CENTRATE: ORGANIC ENRICHMENT

AND MEMBRANE FOULING ................................................................................... 42

3.1. Introduction ............................................................................................................ 43

3.2. Materials and methods ........................................................................................... 44

3.2.1. Forward osmosis system and membranes ....................................................... 44

3.2.2. Feed solution and draw solution ...................................................................... 45

3.2.3. Membrane prewetting ...................................................................................... 46

3.2.4. Experimental protocols .................................................................................... 46

3.2.5. Calculations ..................................................................................................... 46

3.2.6. Analytical methods .......................................................................................... 47

3.3. Results and discussions .......................................................................................... 47

3.3.1. Pure water flux under different conditions ...................................................... 47

3.3.2. Pre-concentration of sludge centrate by forward osmosis .............................. 49

3.3.3. Factors affecting fouling behaviour ................................................................ 51

Page 10: Forward osmosis for resource recovery from anaerobically

viii

3.3.3.1. Membrane properties and orientations ...................................................... 51

3.3.3.2. Draw solution ............................................................................................ 52

3.3.3.3. Fouling layer characteristics ..................................................................... 53

3.3.4. Cleaning and flux reversibility ........................................................................ 54

3.4. Conclusions ............................................................................................................ 57

CHAPTER 4 ................................................................................................................. 58

SEAWATER-DRIVEN FORWARD OSMOSIS PRE-CONCENTRATION OF

NUTRIENTS IN SLUDGE CENTRATE: PERFORMANCE AND IMPLICATIONS ..

................................................................................................................................ 58

4.1. Introduction ............................................................................................................ 59

4.2. Materials and methods ........................................................................................... 60

4.2.1. Forward osmosis system.................................................................................. 60

4.2.2. Feed solution and draw solution ...................................................................... 61

4.2.3. Experimental protocols .................................................................................... 61

4.2.4. Analytical methods .......................................................................................... 62

4.3. Results and discussions .......................................................................................... 62

4.3.1. Performance of FO for enriching nutrients in sludge centrate ........................ 62

4.3.1.1. Water flux .................................................................................................. 62

4.3.1.2. Nutrient enrichment .................................................................................. 63

4.3.2. Performance of FO for enriching nutrients in sludge centrate with buffered

seawater ..................................................................................................................... 66

4.3.2.1. Effects of using buffered seawater as the DS on water flux ..................... 66

4.3.2.2. Improvement in nutrient enrichment using buffered seawater as the DS . 67

4.4. Conclusions ............................................................................................................ 68

CHAPTER 5 ................................................................................................................. 70

CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE WORK ................ 70

5.1. Conclusions ............................................................................................................ 71

5.2. Recommendations for Future Work ...................................................................... 71

REFERENCES ............................................................................................................. 73

Page 11: Forward osmosis for resource recovery from anaerobically

ix

LIST OF FIGURES

Figure 1. The outline of thesis ........................................................................................ 5

Figure 2. Structure of an FO membrane. ........................................................................ 8

Figure 3. (a) ECP phenomena and dilutive ICP in AL-FS orientation and (b) ECP

phenomena and concentrative ICP in AL-DS orientation in the FO process. ................ 9

Figure 4. Geometric structure of ladder (a) and diamond (b) configurations of spacer 17

Figure 5. Flow direction in forward and reverse flushing in FO mode ........................ 38

Figure 6. The permeate flow direction during backwashing in PRO mode ................. 39

Figure 7. Schematic diagram of the lab-scale FO system. ........................................... 45

Figure 8. Pure water fluxes of CTA and PA FO membranes using seawater as DS in

AL-FS and AL-DS orientations with and without prewetting. ..................................... 47

Figure 9. The performance of FO for enrichment of COD in sludge centrate using

seawater as the DS. The theoretical COD concentration factor was calculated assuming

100% COD retention by the FO membrane.................................................................. 49

Figure 10. The performance of FO for enrichment of COD in sludge centrate using the

NaCl solution as the DS and the CTA membrane in the AL-FS orientation. Note: The

theoretical COD concentration factor was calculated assuming 100% COD retention

by the FO membrane. .................................................................................................... 50

Figure 11. Fouling behaviour of the FO membranes in (A) AL-FS orientation and (B)

AL-DS orientation using seawater as DS. .................................................................... 51

Figure 12. Comparison of fouling behaviour towards CTA membranes in the AL-FS

orientation using seawater and the NaCl solutions at different concentration as DSs. 52

Figure 13. SEM images of the AL of (A) the fouled CTA membrane and (B) the

fouled TFC PA membrane in the AL-FS orientation using seawater DS..................... 53

Figure 14. Flux recovery of the fouled CTA, PA and prewetted PA membranes using

seawater DS after physical flushing. ............................................................................. 54

Figure 15. Comparison of pure water fluxes and flux recovery of CTA membranes in

the AL-FS orientation using seawater and NaCl DS. ................................................... 55

Figure 16. SEM images of (A) the AL of the cleaned CTA membrane in the AL-FS

orientation using seawater as DS, (B) the AL of the cleaned CTA membrane in the

AL-FS orientation using the 0.5 M NaCl solution as DS, (C) the SL of the cleaned

CTA membrane in the AL-DS orientation using seawater as DS and (D) the AL of the

cleaned PA membrane in the AL-FS orientation using seawater DS. .......................... 56

Page 12: Forward osmosis for resource recovery from anaerobically

x

Figure 17. Water flux profile as a function of water recovery at different Am/VP ratios

with and without FS stirring during the pre-concentration of nutrients in sludge

centrate by FO. Seawater was used as the DS. ............................................................. 62

Figure 18. Effects of stirring up the FS on sludge centrate pH and nutrients enrichment

during the filtration. The theoretical values of phosphate and ammonium concentration

factor were calculated assuming 100% nutrient retention by the FO membrane. The

Am/VP ratio in this experiment was 175 m-1. Seawater was used as the DS. ................ 63

Figure 19. Phosphate and ammonium rejection towards different membrane areas and

stirring conditions at water recovery of 70%. Seawater was used as the DS ............... 64

Figure 20. Correlation between duration of filtration and pH of the FS towards

different experimental conditions at water recovery of 70%. Raw seawater was used as

the DS. ........................................................................................................................... 65

Figure 21. Changes in water flux at different ratios of Am/VP without FS stirring when

seawater buffered at pH 4.4 was used as the DS. ......................................................... 66

Figure 22. The performance of FO for enrichment of nutrients in sludge centrate

without FS stirring at different ratios of Am/VP. The theoretical values of phosphate

and ammonia concentration factor were calculated assuming 100% nutrients retention

by the FO membrane. Buffered seawater was used as the DS. .................................... 67

Figure 23. Differences in mass balance of phosphate between with and without FS

stirring, as well as between using raw seawater and buffered seawater as the DSs. The

used ratio of Am/VP was 175 m-1. .................................................................................. 68

LIST OF TABLES

Table 1. Comparison between CTA and TFC-PA FO membranes regarding

performance parameters and stability ........................................................................... 15

Table 2. Typical non-responsive draw solutes for FO applications ............................. 19

Table 3. Typical responsive draw solutes for FO applications ..................................... 20

Table 4. Advantages and disadvantages of different membrane module configurations

used for water applications ........................................................................................... 22

Table 5. Properties of some typical colloids in membrane fouling .............................. 25

Table 6. Characteristics of typical scalants ................................................................... 27

Table 7. Properties of several typical bio-foulants in FO operation ............................. 28

Table 8. FO pretreatment technologies in fouling control ........................................... 35

Page 13: Forward osmosis for resource recovery from anaerobically

xi

Table 9. Several modification techniques of the FO membrane surface to mitigate the

membrane fouling ......................................................................................................... 37

Table 10. Several typical cleaning chemicals according to the type of foulant. .......... 40

Table 11. Key properties of the AL of the FO membranes [100, 101]......................... 44

Table 12. Composition of seawater and sludge centrate (Values indicated average ±

standard deviation of three samples). ........................................................................... 45

Table 13. Composition of seawater and sludge centrate (values indicated average ±

standard deviation of at least five samples). ................................................................. 61

Page 14: Forward osmosis for resource recovery from anaerobically

xii

ACRONYMS

AL Active layer

AHL Acyl homoserine lactone

ATP Adenosine triphosphate

CECP Concentrative external concentration polarization

CICP Concentrative internal concentration polarization

COD Chemical oxygen demand

CP Concentration polarization

CTA Cellulose triacetate

DECP Dilutive external concentration polarization

DI Deionized

DICP Dilutive internal concentration polarization

DS Draw solution

ECP External concentration polarization

EDTA Ethylenediaminetetraacetic acid

EPSs Extracellular polymeric substances

FO Forward osmosis

FS Feed solution

HTI Hydration technology innovation

ICP Internal concentration polarization

IEP Isoelectric point

NF Nanofiltration

PA Polyamide

PRO Pressure retarded osmosis

PZC Point of zero charge

RO Reverse osmosis

Page 15: Forward osmosis for resource recovery from anaerobically

xiii

SEM Scanning electron microscopy

SL Supporting layer

TFC Thin film composite

WWTP Wastewater treatment plant

Page 16: Forward osmosis for resource recovery from anaerobically

1

CHAPTER 1. INTRODUCTION

Page 17: Forward osmosis for resource recovery from anaerobically

2

1.1. Background

In wastewater treatment facilities, anaerobic treatment is extensively employed to assimilate

biosolids released during wastewater treatment. These processes aim to decrease the volume

of the solid waste, thereby reducing transportation and disposal costs. Moreover, anaerobic

digestion of sewage sludge can illustrate the enormous potential of energy recovery via biogas

production. Biogas can be burnt to generate heat and electricity so as to compensate the

amount of consumed energy in wastewater treatment plants (WWTPs). Ultimate products of

anaerobic digestion processes are composed of biogas and anaerobically digested sludge

which is afterwards dewatered to separately create biosolids and a liquid fraction called

sludge centrate. Biosolids can be considered as a source of fertilizer for agriculture production

or be discarded, while sludge centrate is returned to the head of work and mixed with the

influent raw sewage to be treated [1, 2]. Sludge centrate is nutrient-rich wastewater containing

>150 mg/L orthophosphate, >1100 mg/L ammonia and around 400 mg/L biological oxygen

demand [3]. The recirculation of raw sludge centrate results in negative impacts on biological

treatment processes and water quality of the final effluent due to the significant addition of

organic and nutrient load. Among methods to overcome the challenges posed by sludge

centrate, pre-concentration of sludge centrate can be a promising approach. The reduced

volume of sludge centrate with high content of organic matter and nutrients such as ammonia

and phosphorus after pre-concentrated can lay the foundation of subsequent nutrient and

energy recovery.

Phosphorus is an essential fertilizer ingredient for agriculture production [4, 5]. These days,

the depletion of phosphorus reservation due to the over-exploitation of minable phosphate

rocks for agricultural production has threaten food security [6]. Phosphorus preservation is

particularly important in Australia, where the soils are mostly phosphorous deficient [7].

Besides, the excess of phosphorus in aquatic environment can cause ecological disasters such

as eutrophication and bloom algae [8]. Thus, recycling and recovery of phosphorus from

nutrient-rich sources such as municipal wastewater, urine and especially from sludge centrate

to produce fertilizers can be promising approaches to ensure the sustainable development of

agriculture and the environmental protection. Furthermore, nutrient recovery from sludge

centrate can help WWTPs reduce the treatment costs and comply with the regulation of

stringent phosphorus and ammonia concentrations in the effluent.

Nutrients such as nitrogen and phosphorus can be reportedly extracted from nutrient-rich

sources under the forms of precipitation, namely calcium phosphate (Ca3(PO4)2) and struvite

Page 18: Forward osmosis for resource recovery from anaerobically

3

(MgNH4PO4) [1, 9]. Recently, membrane technologies have exhibited as perspective

candidates to pre-concentrate nutrients prior to phosphate mineral precipitation [1, 9].

Nanofiltration (NF) and reverse osmosis (RO) have been effectively applied for nutrient

retention from wastewater such as urine and sludge centrate [10, 11]. However, these

processes have a high membrane fouling tendency and consume much energy [12]. Compared

to pressure-driven membrane technologies such as RO, NF used to recover nutrients from

nutrient-rich sources, forward osmosis (FO) filtration is supposed to overcome the challenges

in terms of operational cost, energy consumption and membrane fouling [13]. The high

potential of using FO for nutrient recovery from treated municipal wastewater and sludge

centrate has been reported in some previous studies [1, 9, 13-16]. The high rate of nutrient

recovery from the effluent of a membrane bioreactor using a flat sheet FO membrane was

reported [13, 16]. A combined FO/RO membrane system using NaCl solution as a draw

solution exhibited the effective production of clean water from anaerobic digester centrate

[14]. A hybrid FO/MD membrane system using MgCl2 solution as a draw solution was

developed to effectively extract phosphorus and clean water from sludge centrate [9].

However, these systems require an extra system to re-generate draw solutes because both

NaCl and MgCl2 are costly. The requirement of draw solute regeneration system is one of

obstacles to full scale deployment of FO. In this case, seawater, which is cheap and readily

available in coastal areas is considered as a potential draw solution (DS). The diluted seawater

after FO filtration can be discharged into the ocean. Thus, regeneration of DS is unnecessary.

A seawater-driven FO process was applied to directly extract phosphorus via calcium

phosphate precipitation from sludge centrate. However, the content of phosphorus inside

obtained precipitate was low [1].

1.2. Objectives and scope of thesis

This research thesis is conducted to comprehensively evaluate a seawater-driven FO system

for resource recovery from sludge centrate. The feasibility of this system is considered based

on the efficiency of organic matter and nutrients enrichment and the magnitude of decline in

water flux of FO membranes as well as flux reversibility after physical cleaning. Expected

results of this thesis thus entail:

• Demonstrating the potential of using a seawater-driven FO process for organic matter and

nutrients enrichment in sludge centrate and identifying possible solutions to address key

challenges associated with the seawater-driven FO process for organic matter and nutrients

enrichment.

Page 19: Forward osmosis for resource recovery from anaerobically

4

• Comprehensively evaluating factors affecting the stable performance of a seawater-driven

FO process for pre-concentrating organic matter and nutrients in sludge centrate for

subsequent resource recovery. These factors include membrane material properties,

prewetting protocols, membrane orientation, DS and operating parameters such as stirring

conditions and effective membrane area.

• Characterizing membrane fouling during operation of a seawater-driven FO system for

pre-concentration of sludge centrate to understand more about the nature and mechanism of

fouling phenomenon.

• Evaluating efficiency of fouling preventative methods to find out a feasible cleaning

method as well as assessing impacts of cleaning techniques on the performance of the system.

1.3. Thesis outline

The thesis outline is schematically described in Figure 1. This thesis consists of five chapters.

Chapter 1 presents the introduction. Chapter 2 – Literature review summarizes fundamentals

of FO and membrane fouling. Chapter 3 investigates effects of membrane materials,

membrane orientation, prewetting protocols and DS on the performance of seawater-driven

FO for organic matter enrichment in sludge centrate. The performance of the system is

evaluated through the efficiency of chemical oxygen demand enrichment, membrane fouling

and flux reversibility after physical cleaning. In addition, the potential and challenges of using

seawater for pre-concentrating organic materials in sludge centrate are highlighted via a

comparison between seawater and analytical grade NaCl as DSs. In Chapter 4, key challenges

associated with seawater-driven FO for enriching nutrients in sludge centrate are identified

and potential solutions are proposed to address these challenges. Furthermore, effects of

operating conditions such as stirring and effective membrane area on the efficiency of nutrient

enrichment and membrane fouling are examined and optimized for the first time. Chapter 5

summarizes key findings of this research thesis before suggesting further work in the future.

Page 20: Forward osmosis for resource recovery from anaerobically

5

Figure 1. The outline of thesis

Page 21: Forward osmosis for resource recovery from anaerobically

6

CHAPTER 2. LITERATURE REVIEW

Page 22: Forward osmosis for resource recovery from anaerobically

7

2.1. Forward osmosis

2.1.1. Osmotic process

Forward osmosis has attracted significant attention both academically and commercially

because of its potentials in water purification and nutrient recovery [17]. Osmosis is the

transport of a solvent (normally water) across a semi-permeable membrane from a solution of

high water chemical potential referred to as a feed solution (lower osmotic pressure) to

another solution of low water chemical potential referred to as a draw solution (high osmotic

pressure) [18]. This semi-permeable membrane selectively allows for water permeation, while

restricting the passage of dissolved solutes [18]. Several recent studies have demonstrated the

potentials of FO for the treatment of difficult wastewaters such as fracking fluids [19], landfill

leachate [18], sludge centrate [1, 18] and for resource recovery from these waste solutions [1,

20, 21].

2.1.2. Transport phenomena

2.1.2.1. Mechanisms of mass transfer

In FO, the transport of water, contaminants, and draw solutes is governed by diffusion [17,

18, 22]. Due to the chemical potential difference of water between the feed and DS, water is

diffused from the feed solution (FS) (low concentration solution) to the DS (high

concentration solution) [18]. In practice, FO membranes are not completely impermeable to

all dissolved molecules, thus, some contaminants can permeate through the FO membrane to

the DS [20]. Similarly, the diffusion of draw solutes from the DS to the FS can also occur

[23].

The structure of FO membranes resembles that of RO membranes but without the backing

layer. Most RO and FO membranes are from an advanced membrane generation called thin-

film composite (TFC) membranes [17, 18]. These membranes are composed of a dense

ultrathin active layer and a porous supporting layer as illustrated in Figure 2 [24]. The dense

ultrathin active layer is responsible for separation and is made of polyamide. The porous

support layer is made of polysulfone and provides mechanical support to the FO membrane

[17]. Unlike RO membranes, FO membranes do not have a backing layer because they are

operated at a low hydraulic pressure and this backing layer can exacerbate the concentration

polarization during FO operation [17].

Page 23: Forward osmosis for resource recovery from anaerobically

8

Figure 2. Structure of an FO membrane.

In addition to the diffusion of water through the membrane, there are several other mass

transfer phenomena, namely concentration polarization (CP) and reverse solute diffusion [17,

18]. CP refers to the high concentration of solutes at the membrane surface compared to that

in the bulk solution [25]. The reason for CP is that during the membrane separation processes,

draw solutes are retained on the membrane surface. Moreover, the forward solute diffusion

rate from the bulk solution to the membrane surface is significantly higher than that in the

reverse direction [25]. Thus, draw solutes become increasingly accumulated on the membrane

surface, which results in increase in the concentration of draw solutes near the FO membrane

surface [25]. In the FO process, there are two types of CP, namely external CP (ECP) and

internal CP (ICP) [17, 18]. Reduction of permeate flux, membrane fouling aggravation, and

decrease in membrane rejection can be significant consequences of CP phenomena on FO

membrane performance [17, 18, 26, 27]. In addition to CP, reverse draw solute diffusion

should also be taken into consideration during the FO process because this phenomenon

results in harmful effects on FO performance such as the salt accumulation in the FS, and loss

of draw solutes in the DS [17, 20, 22, 27].

2.1.2.2. External concentration polarization

The ECP phenomenon in FO is similar to that encountered in the RO process [18]. In RO, CP

implies the build-up of solute at the surface of the membrane dense active layer under an

applied pressure, leading to higher concentration of solute in the bulk feed compared to solute

concentration on the membrane surface [18, 25]. Similarly, during osmotic-driven membrane

processes, ECP phenomena can occur on both sides of the membrane without dependence on

membrane orientation, as described in Figure 3 [18]. In Figure 3, Cb is the concentration of

the bulk feed; Cal and Csl are the concentrations of draw solute on the active layer and

supporting layer, respectively; Ci is the solute concentration at the active layer – supporting

Page 24: Forward osmosis for resource recovery from anaerobically

9

layer interface; Cds is the concentration of DS [18]. When the FS flows on, either the active

layer (AL) (Figure 3a, AL-FS orientation), or the supporting layer (SL) of the FO membrane

(Figure 3b, AL-DS orientation), draw solutes can accumulate and be concentrated near the

surface of membrane layers [18]. It is called concentrative ECP (CECP) [18]. On the other

hand, the solute concentration near the membrane surface on the draw side is diluted because

of the permeate flux [18]. This is called dilutive ECP (DECP) [18].

Figure 3. (a) ECP phenomena and dilutive ICP in AL-FS orientation and (b) ECP phenomena

and concentrative ICP in AL-DS orientation in the FO process.

Both CECP and DECP can cause negative impacts on the FO membrane performance [18].

Firstly, the accumulation of solutes in CECP increases the osmotic pressure of FS near the

AL, resulting in reducing the osmotic pressure difference across the membrane. This outcome

also results from the dilution of DS near the permeate side in DECP. Thus, both CECP and

DECP contribute to the reduction of the effective osmotic driving force which causes the

water flux decline [18]. Secondly, similar to scaling fouling in RO, inorganic scaling can

occur in the FO membrane process due to either the deposition and growth of salt

crystallization, or precipitation of sparingly salt ions on the membrane surface [26]. The

extent of scaling-fouling is greatly dependent on the concentration of solutes [26]. According

to the thermodynamic theory, solutes will commence crystalizing when their concentration is

higher than the saturation limit, or metal ions in the FS can be able to precipitate if the product

of ions is greater than the soluble product [26]. Thus, ECP with the high increasing solute

concentrations near the membrane surface can contribute to the enhancement of scaling-

fouling. Finally, in FO mode with the occurrence of ECP, the solution on the permeate side

Page 25: Forward osmosis for resource recovery from anaerobically

10

near the active-support layer interface is less diluted due to flux decline [18]. This can be a

favorable condition for the diffusion of solutes from the DS to the SL, thus aggravating ICP

[17]. Consequently, solutes can permeate through the membrane to the FS, leading to worse

reverse solute flux.

The above adverse effects of ECP on osmotic-driven membrane process can be tackled by

several solutions. General principle to reduce ECP is to abolish the accumulation of solutes on

the membrane surface. The first solution is that increase in cross-flow velocities and

turbulence can minimize ECP because greater cross-flow velocity causes greater shear force

on the membrane surface, resulting in lower accumulation of solutes [18, 26]. Secondly, a

study indicated that using spacers, especially in wastewater applications, could alleviate ECP

as well as fouling. The inserted spacer in the feed channel can promote mixing, the effective

cross-flow velocity, turbulence and mass transport near the membrane surface [17, 26].

However, in most flux models for FO, ECP is assumed to be neglected because of low fluxes

and a high mass transfer [17].

2.1.2.3. Internal concentration polarization

The occurrence of ICP is exclusive to FO because of typical characteristics of the FO process

[17]. In general, ICP is a type of CP phenomena taking place within the porous SL of the TFC

FO membrane [17]. In the FO process, there are always two solutions of different

concentration flowing on the both sides of the membrane. Moreover, the hydraulic pressure

and water flux are low in FO processes [17]. Therefore, ICP always occurs that is independent

of membrane orientation. Different membrane orientations will establish different ICP

phenomena. If the porous SL of the FO membrane faces the DS in AL-FS orientation (Figure

3a), the polarized layer of draw solutes will be diluted by water flux from the FS across the

membrane [18]. This case can be called as dilutive ICP (DICP) [18]. On the other hand, when

the FS streams on the SL in AL-DS orientation (Figure 3b), and water permeates the AL,

draw solutes within the porous layer will be concentrated [18]. This is referred to as

concentrative ICP (CICP) [18]. Both DICP and CICP phenomena occur within the membrane

porous SL, so they cannot be mitigated by conventional hydrodynamic conditions [18].

The detrimental impacts of ICP on the FO membrane process are reportedly more severe than

that of ECP in terms of water flux decline, increase in membrane clogging phenomena and

reverse draw solute diffusion [17, 18, 26]. First of all, either the concentrated draw solute flow

in AL-DS orientation, or the diluted DS in the SL in AL-FS orientation due to ICP leads to

the reduction of the discrepancy of osmotic pressure across the FO membrane [17, 18, 26].

Page 26: Forward osmosis for resource recovery from anaerobically

11

This causes significant loss of the driving force of the mass transfer, which results in water

flux decline [17]. Secondly, the membrane fouling deriving from ICP is similar to that from

ECP, except that it happens within the porous layer behaving as an unstirred layer, and

therefore, is almost irreversible [18]. This phenomenon causes increase in overall membrane

resistance for water transport [26]. Thus, the ultimate consequence of ICP is still a decrease of

water flux. In the end, ICP can exacerbate the reverse solute flux [17]. In AL-FS orientation,

the accumulation of draw solutes within the porous layer leads to significant increase in the

difference of solute concentration across the membrane. This enhances the driving force of

reverse solute diffusion. In addition, CICP occurs in AL-DS orientation that can aggravate

DECP [17]. Thus, the possibility of reverse solute transport can be higher because of the

higher solute concentration near the active layer on the draw side.

The modifications in fabrication of FO membrane, along with optimizing operating conditions

and changes in membrane orientation and materials can be applied to alleviate ICP. The

extent of ICP in FO processes is associated with the degree of the resistance to solute

diffusion (K) within the membrane porous SL (Equation 1) [17, 18]. A larger value of K

implies more severe ICP [18]. Thus, by modifying the structural parameter to reduce the K

value, ICP can be minimized. As can be seen from Equation 1 that a thinner, more porous and

less tortuous SL is indicative of decrease of K, resulting in the less ICP. In addition, the pore

clogging can result from the deposition of the foulants into the membrane porous SL [26].

This contributes to the reduced porosity of the SL, which enhances ICP. This phenomenon

can be deteriorated in AL-DS orientation when the FS containing many different

contaminants flows on the porous SL. Therefore, the selection of appropriate membrane

orientation and hydrodynamic adjustments such as increase in cross-flow velocities and more

frequent flushing can help diminish the foulant accumulation, so alleviate ICP [17, 18, 26].

Moreover, temperature is referred as an important operating condition in FO processes

because of its effects on the diffusion coefficient of the draw solute (Ds) [26]. The increase in

temperature can lead to the intensification of the solute diffusivity because of the reduced

viscosity of the solution [26]. As a result, ICP can be alleviated. Lastly, the membrane

materials can influence the degree of ICP [17]. Hydrophilic SLs can improve water flux and

decrease ICP because of increasing the wetting of small pores within the SL [17].

Page 27: Forward osmosis for resource recovery from anaerobically

12

Equation 1

s

KD

=

In which K is the resistance of solute diffusion; ℓ, τ and ε are the membrane thickness,

tortuosity and porosity, respectively; Ds is the diffusion coefficient of the solute [17, 18].

2.1.2.4. Reverse solute flux

Reverse solute leakage is a unique phenomenon which can be mathematically quantified in

the FO process. In fact, no FO membranes can remove all solutes completely [27]. Thus, both

forward and reverse solute diffusion can simultaneously occur in the FO process. The reverse

permeation of solutes from the DS into the FS is called as reverse solute diffusion [17, 27]. In

contrast, the transport of dissolved solute through the membrane from the feed side to the

draw side is referred as the forward solute diffusion [17]. The low hydraulic pressure and

water flux in FO process can facilitate these phenomena. To determine the extent of reverse

permeation of solute, Js/Jw ratio is introduced to depict the amount of draw solute reversely

diffusing into the FS per unit volume of water flux into the DS (Equation 2) [17, 26, 28]. It

can be seen clearly from Equation 2 that the ratio is constant for a certain membrane, and only

dependent on temperature and the membrane transport properties of the AL, namely A and B

values. A membrane with the superior selectivity which means a higher value of A and a

lower value of B can mitigate the reverse solute flux. Theoretically, Js/Jw is independent of the

properties of the SL, membrane orientation, operating conditions, and hydrodynamic

conditions [17, 26, 28]. In practice, however, the membrane separation properties can vary

with the high concentration of the DS in contact with the dense AL [26, 29], and various types

of solution [26].

Equation 2

w

s

g

J B

J A R T=

Where Js is the reverse solute flux; Jw is the permeate flux; A is the pure water permeability

constant; B is the solute permeability constant; β is the Van’t Hoff coefficient; Rg is the ideal

gas constant; T is the absolute temperature.

The solute in the draw side diffusing into the feed side can lead to disadvantageous effects on

the FO process. The first adverse impact is that the reverse solute leakage causes the loss of

draw solute in the DS [17, 28]. This leads to the reduced osmotic pressure of the DS, and the

Page 28: Forward osmosis for resource recovery from anaerobically

13

enhanced osmotic pressure of the FS [17]. Thus, the osmotic driving force of mass transport

will decrease significantly, which results in the acceleration of water flux decline. To solve

this problem, the replenishment of draw solute to maintain the effective osmotic pressure of

the DS can be necessary [27]. This increases the cost of treatment process [17, 27]. Secondly,

the accumulation of draw solute in the FS can harm microorganisms in osmotic membrane

bioreactors because these species are susceptible to the high salt concentration [17]. When

using some types of DS such as NH3-CO2 solution, the discharge of the FS after accumulating

the draw solute can lead to damage of the environment, namely the eutrophication

phenomenon [28]. Thirdly, the reverse draw solute permeation can be responsible for the

enhancement of CP in the FO process [26]. The deteriorated CECP and CICP will occurs with

increase in the concentration of draw solute near the AL and the porous layer due to the

reverse solute leakage in AL-FS orientation and AL-DS orientation, respectively. Finally,

reverse solute diffusion not only can aggravate CP, but also may contribute to changes in

chemical composition of the FS, which may detrimentally affect the membrane fouling

behavior [26]. The draw solutes such as divalent cations can play a role as fouling promoters

[26]. The appearance of these solutes in the FS can exacerbate the organic and colloidal

fouling which results from the deposition of colloidal particles on the membrane [26].

Moreover, some draw solutes, namely Ca2+, Mg2+, SO42-, CO3

2- and PO43- are considered as

scaling precursors [26]. Thus, the accelerated inorganic fouling in the feed side can happen

due to reverse diffusion of these scaling precursor ions [26]. In many cases, the draw solutes

can behave as nutrients, and the reverse diffusion of these solutes can lead to the biofilm

development which causes biofouling [26].

Reverse solute diffusion can be minimized by changing the membrane intrinsic properties,

appropriate draw solute selection, and hydrodynamic conditions. At first, as mentioned above,

the extent of reverse draw solute diffusion is dependent on the membrane intrinsic separation

properties [17, 26]. A FO membrane with a dense AL of higher water permeability and lower

solute permeability can result in less solute diffusing from the DS into the FS [26]. Thus,

some advanced modifications of the AL such as sealing defects which leads to the reduced B

value will help alleviate the reverse solute leakage. Another factor affecting the rate of reverse

solute leakage is the DS composition [17, 26]. In FO process, the high concentration of DS is

vital for the osmotic driving force of water transport across the membrane [17, 18]. However,

higher DS concentration can lead to greater rate of reverse draw solute diffusion [26]. Thus,

an appropriate value of DS concentration to ensure the harmony between water flux and

Page 29: Forward osmosis for resource recovery from anaerobically

14

reverse solute flux should be taken into consideration. In addition to the DS concentration, the

draw solute type can also influence the draw solute reverse diffusion noticeably [17, 26]. A

study indicated that using monovalent ions as draw solutes demonstrates the higher solute

leakage despite releasing higher permeate fluxes compared to using bivalent ions [17]. The

reason is that the size exclusion and electrostatic repulsion of monovalent ions is low [17].

Meanwhile, bivalent ions illustrate lower reverse solute fluxes because of their higher steric

hindrance and electrostatic repulsion [17]. At last, it is reported that the reverse solute

diffusion can be deteriorated by CP [17]. Thus, some adjustments of hydrodynamic conditions

such as increase in cross-flow velocities to decrease CP can lead to the reduced reverse solute

diffusion.

2.1.3. Materials

FO membranes are similar to those used in RO applications regarding materials and

fabrication methods [24]. Currently, most FO and RO membranes are commercially made of

two typical polymeric materials including cellulose triacetate (CTA) and polyamide (PA)

[24]. Particularly, PA FO and RO membranes are made on a support of polysulfone. The

methods of FO membrane fabrication are dependent on the specific materials. CTA FO

membranes are fabricated by phase inversion, followed by heat annealing treatment [24]. In

this method, the characteristics of these CTA FO membranes can be significantly affected by

polymer concentration, type of solvent, evaporation time, annealing temperature, casting

substrate and coagulant bath [24]. For example, a more dense CTA FO membrane can result

from using higher concentration of polymeric solution [25]. When it comes to PA TFC FO

membranes, these membranes are composed of a porous SL and a selective AL [24]. The

porous SL made of polysulfone is fabricated via phase inversion with/without a thin non-

woven fabric [24]. Meanwhile, the dense AL made of crosslinked aromatic PA are generally

synthesized via interfacial polymerization reaction between meta phenylenediamine monomer

in water and trimesoyl chloride monomer in organic solvent [24]. In this process, factors such

as concentrations of monomers and solvents, types of monomers and solvents, and

polymerization conditions need to be optimized to create high performance PA TFC

membranes [24].

Membrane materials can directly influence the FO performance in terms of separation

properties, permeability, fouling tendency and chemical and biological stability. Firstly, using

FO membranes with superior intrinsic separation properties which are indicative of higher

water permeability and selectivity can lead to greater water flux and minimize the negative

Page 30: Forward osmosis for resource recovery from anaerobically

15

impacts of draw solute leakage [26, 28]. When utilizing these membranes, it is more

compatible with using small and fast-diffusing draw solutes such as sodium chloride,

magnesium chloride and ammonia-carbon dioxide [28]. The reason is that these draw solutes

can establish a high osmotic driving force and alleviate ICP, which results in higher water

flux, but they are also susceptible to draw solute diffusion due to their small size [28].

Secondly, the dense AL of smaller permeability and lower rejection can cause the enhanced

forward and reverse draw solute diffusion. As a consequence, this may deteriorate the risk of

fouling as mentioned above [26]. A study indicated that hydrophobic membranes are more

vulnerable to adsorptive fouling than hydrophilic membranes such as CTA because of higher

repulsive acid-base interaction that deters the accumulation of colloidal particles on the

membrane surface [17, 26]. Also, the chemical and biological endurance of FO membranes

can be noticeably impacted by used materials [17, 26]. This depends on the interaction

between chemical components of the FO membrane material and contaminants in wastewater

[17]. Table 1 compares the performance and properties of TFC PA and CTA membranes [17,

18, 26, 28].

Table 1. Comparison between CTA and TFC-PA FO membranes regarding performance

parameters and stability.

Parameters CTA TFC-PA

Water permeability Low High

Rejection Low High

Membrane fouling High fouling propensity Low fouling tendency

Chemical stability

A narrow range of pH

tolerance (3-8)

Highly resistant to chlorine

Stability at broad pH values (2-12)

Sensitive to chlorine

Biological stability High biodegradable tendency Low biodegradable propensity

2.1.4. Forward osmosis operation

2.1.4.1. Temperature

The feed and draw solution temperature and transmembrane temperature difference can

significantly affect the performance of FO process [30, 31]. Regarding membrane properties,

an increase in the FS temperature can increase the solute diffusivity and decrease the viscosity

of water, thereby raising the pure water and solute permeability coefficients (A and B

coefficients, respectively) of the FO membrane [31]. In addition, the membrane polymeric

matrix can be deformed at excessively elevated temperature [17]. When it comes to water and

Page 31: Forward osmosis for resource recovery from anaerobically

16

reverse solute fluxes, a rise in the DS temperature can lead to decrease in DS viscosity and

increase in osmotic pressure as well as the draw solute diffusivity [30-32]. Thus, water and

reverse salt fluxes can be enhanced with the high temperature of the DS [30, 31]. This is

consistent with the increase in the A and B values of the FO membrane mentioned above.

When the FS temperature increases, the FS viscosity decreases, thereby enhancing the

diffusivity of water through the membrane [30]. Moreover, the improved mass transfer

coefficient of the FS can result from the increase in the diffusion coefficient due to high

temperature, resulting in alleviating ECP [30, 31]. Notably, the temperature difference

between the feed and draw solution can cause certain effects on the rejection of neutral

contaminants [31]. When the temperature of the FS is higher than that of the DS, the

decreased rejection of neutral contaminants can be attributed to an increase in their diffusivity

at an elevated temperature [31]. By contrast, the rejection of neutral pollutants can be

improved when the DS temperature is higher than the FS temperature [31]. This is because

the increased reverse salt flux which is induced by an increase in the draw solute diffusivity

can hinder the forward diffusion of contaminants [31]. Another reason is the increase in water

flux can directly account for an increase in rejection, which is similar to that observed in

nanofiltration and reverse osmosis processes [31].

2.1.4.2. Feed solution pH

The FS pH can influence water flux, reverse salt flux in the FO process through its effects on

FO membrane structure [20, 26, 33, 34]. Firstly, the increased water flux can be a result of

increasing the FS pH [33, 34]. This phenomenon can be attributed to changes in cross-linked

membrane polymer structure and membrane wettability [33]. It is assumed that an increase in

the FS pH can accelerate the electrostatic repulsion among ionizable functional groups of the

membrane polymeric structure, leading to an increased average pore size and higher permeate

flux [33]. Moreover, more dissociation of carboxyl functional group (COO-) of the AL can

occur at higher FS pH, so the FO membrane surface can become more hydrophilic [33, 35]. A

more hydrophilic membrane can facilitate water transport [33]. Secondly, the reverse salt flux

can decrease as the FS pH increases [33]. An increase in the FS pH can not only cause a rise

in water flux, but also make the FO membrane more negatively charged [33]. Both of these

consequences can suppress the salt flux [33]. Finally, the FO membrane damage can occur at

either excessively low or high pH values because they can break peptide bonds in polymer

chains of the membrane material such as polyamide [17, 18]. As a consequence, the

membrane separation properties such as permeability and selectivity can be affected.

Page 32: Forward osmosis for resource recovery from anaerobically

17

2.1.4.3. Cross-flow velocities and spacer designs

The cross-flow velocity and spacer design can significantly impact on the level of CP and

mass transfer near the membrane surface in the FO process [26]. Increased cross-flow

velocities can encourage the back diffusion of those solutes and particles as well as enhance

mas transfer coefficient, resulting in mitigating the accumulation of retained solutes and

colloids near the membrane surface [26]. As a result, CP and membrane fouling, especially

external fouling can be significantly alleviated [26]. In practice, increase in cross-flow

velocities aims to reduce the membrane fouling, not to improve water flux [18]. However, the

intensified cross-flow velocities are mostly beneficial for the reduction of external fouling and

ECP, but may not be effective to reduce internal fouling and ICP occurring within the SL

[26]. In addition to the increase of cross-flow velocities, the use of spacers can also affect

cross-flow conditions [17]. In particular, spacers are necessary in spiral-wound membrane

configurations [36]. They are alternately positioned among membrane flat sheets to form the

flow channels [36]. Therefore, spacers can be utilized in both the feed and permeate channel

[36]. Commercial spacer consists of two layers of cylindrical filaments creating a ladder-type

or diamond-type spacer geometry (Figure 4) [36]. In operation, these spaces can cause the

enhanced turbulence of liquid streams, which can encourage mixing and improve mass

transfer [17]. This can help to restrict the formation of the membrane boundary layer and CP,

especially ECP [17]. Thus, membrane fouling can be also alleviated, and water flux can be

enhanced.

Figure 4. Geometric structure of ladder (a) and diamond (b) configurations of spacer

Top layer

Bottom layer

Top layer

Bottom layer

(a) (b)

Page 33: Forward osmosis for resource recovery from anaerobically

18

2.1.4.4. Draw solution

Draw solution is a highly concentrated solution which has low chemical potential of water on

the permeate side of the FO membrane to induce the transport of water through the membrane

[18, 37]. A suitable DS needs to satisfy the following requirements [17, 18, 37, 38]: (i)

generating higher osmotic pressure compared to the FS; (ii) causing no damage to the

membrane structure of the active and support layers; (iii) minimal reverse diffusion of the

draw solute; (iv) chemical stability for re-use; (v) non-toxicity with no harmful impacts on

human and the environment; (vi) low cost; (vii) easy re-concentration of the diluted DS; (viii)

low viscosity and small molecular weight. Herein commonly-used types of draw solute will

be characterized and categorized into non-responsive and responsive solutes.

Non-responsive draw solutes including inorganic salts, non-functionalized polymers and

organic molecules, and seawater and RO brine indicate draw solutes that cannot change their

properties with external stimulants such as pH, temperature, electro-magnetic field or light

[17, 37, 38] (Table 2). Responsive draw solutes can include nanoparticles, hydrogels,

metathesis precipitable salts, soluble gases and volatile liquids, NH3-CO2 draw solutes, highly

charged compounds and polarity-switchable draw solutes. They respond to external stimuli

(Table 3) [17, 37, 38]. In practical, the most common draw solute is sodium chloride (NaCl)

because it has high solubility, creates high osmotic pressure and is somewhat simple to re-

concentrate to high concentration with RO without risk of scaling [17, 26].

The dilution of DS, the accumulation of dissolved contaminants in the DS and reverse draw

solute leakage can simultaneously occur during the FO operation [17, 18, 20, 26]. In the FO

process, water is continually transferred from the feed side to the draw side over time, thereby

leading to the dilution of DS [18]. This can diminish the osmotic driving force of mass

transfer. Thus, a re-concentration system to extract water from the diluted DS is required [20,

26]. Currently, draw solutes can be re-concentrated by hybrid membrane systems combining

the FO membrane and other types of membrane such as RO membranes and membrane

distillation membranes [20]. In addition to the water transfer through the FO membrane,

pollutants especially trace organic compounds can also pass through the membrane and enter

the DS because FO membranes are not completely impermeable to all dissolved solutes [20].

The accumulation of these contaminants can result in the degradation of product water and

even membrane fouling in the draw solute recovery system [20, 26]. Therefore, the removal

of these contaminants should be considered. Meanwhile, the reverse diffusion of draw solute

from the draw side to the feed side can cause the deficit of draw solute which reduces the

Page 34: Forward osmosis for resource recovery from anaerobically

19

osmotic pressure of the DS [17, 26]. Thus, a replenishment of draw solute needs to be

periodically conducted to ensure the necessary osmotic driving force of mass transfer [17, 20].

Table 2. Typical non-responsive draw solutes for FO applications

Draw

solute Examples Advantages Disadvantages

Non-r

espo

nsi

ve

solu

tes

Inorganic salts

(e.g. NaCl, MgCl2,

Na2SO4, (NH4)2SO4,

Ca(NO3)2)

• High solubility

• Low cost

• High osmotic

pressure

potential

• Sensitive to scaling/clogging

• Higher reverse draw solute flux

with small draw solutes

• Difficult to recover in the re-

concentration system

Polymers/ macro-

organic molecules (e.g.

poly-sodium-acrylate,

glycine)

• Reduce reverse

diffusion

• High water flux

• Aggravate CP

• Difficulties in regeneration

process and circulation

• Limited storage time due to

biodegradation

Seawater and RO brine • Abundant source • Sensitive to fouling

• May need to be pre-treated

Page 35: Forward osmosis for resource recovery from anaerobically

20

Table 3. Typical responsive draw solutes for FO applications

Draw

solute Examples Advantages Disadvantages

Res

ponsi

ve

dra

w s

olu

tes

Nanoparticles such as Fe3O4

• Very low reverse draw solute flux

• High osmotic pressure potential at low

concentration

• Easy to re-concentrate using a magnet

• Agglomeration during magnetic separation

• Ultrasonication weakens magnetic properties

• Viscosity of solution diminishes effective

driving force and the flux

Hydrogels • No draw solute reverse leakage • Low water flux

• Poor water recovery

Metathesis precipitable salts such

as CuSO4, MgSO4, Al2(SO4)3

• Only be applicable for brackish water

treatment

• Not suitable for drinking water production

• Use of large amount of chemicals and formation

of byproducts during generation

Soluble gases and volatile liquids

such as SO2, CO2, NH3 • High osmotic pressure

• Loss of draw solutes

• Need external pressure to aid the dissolution of

the draw solute

• Complication on actual implementation

NH3-CO2 draw solutes • High solubility in water

• Recovery by moderate heat

• Toxic thermolytic product

• High diffusive loss

Polarity-switchable draw solutes

such as hydrophobic amines • High osmotic pressure

• Tend to swell and damage the AL

• Difficulty in regeneration process

Highly-charged compounds such

as EDTA

• High water flux

• Low reverse diffusion

• Less energy consumption during re-

concentration

• High cost

• pH dependency

Page 36: Forward osmosis for resource recovery from anaerobically

21

2.1.4.5. Membrane orientation

Membrane orientation is indicative of the position of an FO membrane relative to the DS and

the FS [17]. As noted in section 2.1.2, the FO membrane comprises of a dense selective AL

and a porous SL [17, 18]. Thus, the FO membrane can be positioned with the AL facing either

the feed side (AL-FS or FO mode), or the draw side (AL-DS or PRO mode) [17, 18, 39].

Because of the characteristics of the AL and the SL, membrane orientation can directly

influence the performance of FO process [17].

Permeate flux and salt rejection of the FO process should be taken into comprehensive

account during the selection of membrane orientation for FO applications [17, 18, 33, 39].

Previous studies reveal that the permeate flux of the FO membrane in PRO mode is higher

than that in FO mode as the DS concentration is high. This is because CICP can be less severe

in PRO mode [17, 39]. Salt rejection in FO mode is usually higher than that in PRO mode

[33, 39]. This can be attributed to the smoother AL surface that can mitigate the possibility of

salt accumulation on the membrane surface in FO mode, thereby reducing the salt diffusion

from the feed side to the draw side [26, 33]. Moreover, the diminished turbulence of the FS

can accelerate ICP in PRO mode, resulting in the reduced salt rejection [26, 33].

2.1.4.6. Module configurations

In water purification and wastewater treatment applications, FO module configurations share

similarities with those used in RO. Module configurations imply the ways of packing or

holding membranes into a module so as to maximize the surface to volume area and reduce

particles deposition by sufficient cross-flow [17]. In practice, there are three prevalent FO

membrane module configurations including spiral-wound, hollow fiber and plate and frame

employed to meet various treatment requirements [17, 18]. Each module configuration has

different benefits and drawbacks in terms of economic considerations, the complication of

fabrication into the module design, the flexible applicability, fouling tendency and CP

phenomena (Table 4) [17, 18, 40].

Page 37: Forward osmosis for resource recovery from anaerobically

22

Table 4. Advantages and disadvantages of different membrane module configurations used

for water applications

Modules Advantages Disadvantages

Spir

al-w

ound

• High packing density

• Easy cleaning of fouling

agents

• Limited pressures in membrane envelop

and clogging of spacers

• Inability to induce flow in the permeate

channels (inner side of the envelopes) of

the membrane elements for cleaning or

backwashing

Holl

ow

fib

er

• Self-supported

characteristics

• Appropriate flow

• Simplicity of fabrication

• High packing density

• Limited mixing at membrane surface

• In open channels, the CP film grows

undisturbed in the channel

Pla

te a

nd f

ram

e

• Well-suited to wastewater

applications

• Less complicated in design

• Better backwashing

• Higher cross-flow velocities

• Lack of adequate membrane supports

and problems with internal and external

sealing

• Low packing density (leads to a larger

system footprint, higher capital costs,

and higher operating costs

2.2. Fouling phenomena in the FO membrane operation

2.2.1. FO membrane fouling

Similar to RO, fouling is inevitable during FO membrane operation. Membrane fouling in the

FO process refers to the deposition and accumulation of constituents such as solutes, particles

or microorganisms on the FO membrane surface, which can deteriorate the performance of

the membrane [17, 26, 34]. Flux decline, changes in intrinsic separation properties and even

damage to the FO membrane can be consequences of membrane fouling [17, 26, 41]. It

appears that FO fouling is more reversible than that in RO due to the lack of hydraulic

pressure [17, 26].

In the FO process, membrane fouling can be classified into four groups in terms of foulant

types [26, 29]. Colloidal fouling is indicative of the deposition of colloids such as silica and

hydroxides on the membrane surface [26, 42]. Organic fouling refers to the deposition and

Page 38: Forward osmosis for resource recovery from anaerobically

23

adsorption of organic macromolecules such as polysaccharides, humic acid and protein [26,

42]. In addition, once the precipitation or crystallization of dissolved inorganic substances

takes place on the membrane, inorganic scaling can be established [26, 29]. Biofouling

implies the adhesion and accumulation of microorganisms forming biofilm on the membrane

[26, 29]. In practice, all fouling types can occur during FO operation [17, 26].

Membrane fouling in the FO process can also be divided into surface fouling (external

fouling) and internal fouling in terms of the fouling places [26, 29]. Both types of fouling can

be likely to occur on the dense AL and within the porous SL at the same time. In FO mode,

once the dense AL faces the feed side, numerous foulants in the FS can accumulate on the

surface of this layer, thereby building up a cake layer which resembles fouling in RO.

Membrane fouling in this case can be referred to as external fouling [26, 29]. In the meantime,

the possibility of membrane fouling within the porous SL can exist under the form of

inorganic fouling in DS side due to forward diffusion of fouling precursors [26]. Membrane

fouling behavior can be more complicated in PRO mode as the porous SL keeps into contact

with the feed side. If the size of foulants is small enough, they can go into the SL through

convection [26]. Some foulants can be absorbed on the pore wall [26]. Others can be finally

retained by the dense AL [26]. As a result, the pore can be filled up with foulants over time.

This phenomenon is called as internal fouling [26]. Under the real wastewater condition,

foulants still keep depositing on the outer surface of the porous layer, which can cause the

formation of external fouling after the pore is replete [26]. Thus, both external fouling and

internal fouling can be significant in PRO mode.

2.2.2. Colloidal and organic fouling

2.2.2.1. Characteristics of colloids and macromolecules

Colloids can be divided into two groups namely inorganic colloids and organic

macromolecules [26, 42]. The former include some common inorganic substances such as

silica, iron hydroxide and clay minerals [42]. The latter include rigid biopolymers (large

molecular weight polysaccharides), fulvic compounds (fulvic and humic acid), proteins and

other macro-organic molecules [42]. In FO operation, both colloidal fouling and organic

fouling can be affected by similar factors as well as explained by the same fouling

mechanisms because inorganic colloids and organic macromolecules share many similarities

in terms of their properties [26].

Colloids can be fully described through their size and conformation and their charge in the

feed water (Table 5) [26, 42]. In terms of inorganic colloids, the size and shape of inorganic

Page 39: Forward osmosis for resource recovery from anaerobically

24

particles are various, and they are mostly rigid [42]. In water environment, most inorganic

colloids are negatively charged at around neutral pH [42].

Regarding organic colloids, polysaccharides are large macromolecules with average

molecular weight range from hundred to a few thousand kDa [42]. They are weakly

negatively charged [42]. One of typical polysaccharides causing fouling is alginate which has

the charge density of 3 meq/g (up to 6 meq/g) [42]. In particular, polysaccharides can exist

under as a gel-like structure (a three-dimensional cross-linked substance in the presence of

multivalent metal ions) [42]. Another organic foulant in wastewater is fulvic substances,

which are produced by biodegradation of dead organic matter [26, 42]. Fulvic substances are

abundant in wastewater, thus are a major cause of organic fouling during FO operation.

Compared to polysaccharides, the molecular weight of humic acid is lower (just from several

kDa to a few hundred kDa) [42]. This acid is highly polydispered [42]. Moreover, humic acid

is a negatively charged compound at a broad range of pH due to the deprotonation of

carboxylic and phenolic functional groups, and its charge density will increase as pH

increases [42]. Most proteins are amphoteric because of the presence of both carboxylic (-

COO-) and amine (-NH3+) groups in their chemical structure. When pH is lower than

isoelectric point pH (pHIEP), protein molecules are positively charged. By contrast, these

molecules are negatively charged once pH of the solution is higher than pHIEP. In aqueous

solution, protein are globular macromolecules.

Page 40: Forward osmosis for resource recovery from anaerobically

25

Table 5. Properties of some typical colloids in membrane fouling

Colloids Examples Molar mass

(kDa) Shape Charge in water pHIEP or pHzpc

Inorganic fouling

am-SiO2 Flexible Round Negative 3-3.5

am-Al2O3 Flexible Angular Negative 7

am-FeOOH Flexible Crystalline or amorphous Positive 7.9-8.1

Polysaccharide

Alginate 200 - 2000 Extended random coil Negative

3-6 meq/g 5.4

Xanthan and Gellan 100 - 2500 Single-stranded or double-helical Negative 4.5

Schizophyllan 400 - 500 Rigid rod-like Neutral -

Humic substance

Humic acid in International

Humic Substance Society

(IHSS) standards

Of a few kDa to a

few hundred kDa

globular molecule (linear under

high pH, low ionic strength and

low concentration)

Negative

5-10 meq/g 3

Protein

Bovine serum albumin 67 Flexible Negative

1.5 meq/g 4.7

Bovine immunoglobulin G 155 Flexible Negative 6.6

Bovine hemoglobin 68 Flexible Negative 7.1

Bovine pancreas ribonuclease

A 13.7 Flexible Negative 7.8

Lysozyme 14.4 Flexible Positive 11

Note: “-” indicates not available; “am” stands for amorphous; pHpzc stands for the point of zero charge.

Page 41: Forward osmosis for resource recovery from anaerobically

26

2.2.2.2. Colloidal interactions

Membrane fouling can be governed by colloidal-colloidal and colloidal-surface interactions

[26, 42]. The colloidal-surface interaction can enable colloids to approach the membrane

surface, and then deposit on it, while colloidal-colloidal interactions will allow foulants to

attach the deposited foulants and develop into a cake layer on the membrane surface [26, 42].

According to a classical theory, particle-particle and particle-surface interactions can be

dominated by synergic impacts of Van der Waals force and electrostatic interaction force

known as the electrical double layer force [42]. The sum of these forces can establish an

interaction energy called an energy barrier [26, 42]. This energy barrier must be overcome if

any particles wish to attach onto the membrane surface [26]. Once a colloidal system is stable

(i.e. it has a large energy barrier), it can be difficult for particles to affix onto membrane

surface. This interaction energy is dependent on physico-chemical parameters such as particle

size, surface charge and surface roughness as well as solution chemistry such as pH, ionic

strength and types of ions [26, 42]. For example, electrostatic interaction can decrease at high

solution ionic strength. In addition to the classical theory, a state-of-the art theory called the

Lewis acid-base theory, has been introduced to explain for the aggregation of colloids [26,

42]. In accordance with this novel approach, a colloid in aqueous solution can interact with

others through an electron-acceptor-electron-donor interaction [26]. There will be a repulsion

among hydrophilic colloids because of hydration force, or an attraction among hydrophobic

colloids [26]. In summary, both theories can be utilized to characterize the colloidal

interactions entirely [26].

2.2.3. Inorganic fouling

The occurrence and intensity of scaling is dependent on the concentration and solubility of

scalants in the solutions [26]. Scaling is the deposition of sparingly soluble inorganic salts on

the membrane surface at certain conditions [26, 41]. In general, scaling happening in the FO

process can be a consequence of CP which causes increase in the concentration of insoluble

salts as well as the possibility of precipitation [41]. When the product of concentration of

minerals is higher than the solubility product, the precipitation of these minerals will occur.

The accumulation of these crystals can lead to scale occurrence on the membrane surface.

Mineral salt crystal deposition from the bulk solution, along with the direct crystallization

onto membrane surface can form a scaling cake layer [26]. Unlike colloidal and organic

fouling that can happen at low concentration, inorganic scaling only take place as the

Page 42: Forward osmosis for resource recovery from anaerobically

27

supersaturation exceeds a critical value [26]. Several typical scalants can be listed and

characterized in Table 6 [26, 41].

Table 6. Characteristics of typical scalants

Examples Size and shape Alkalinity

/Acidity

Factors affecting the

formation

Calcium sulphate

(Gypsum) Needle-like or plate-like Neutral

Concentrations of ions

and temperature

Calcium carbonate

(CaCO3)

Calcite: rhombohedral

Aragonite clusters:

outward oriented needles

Vaterite: spherical

Alkaline

Calcium hardness,

bicarbonate alkalinity,

temperature and pH

Calcium phosphate

(Ca3(PO4)2) Amorphous Alkaline

Phosphate concentration,

temperature and pH

Silicate scale Crystalline or amorphous Weakly acid trivalent metal cations and

pH

2.2.4. Biofouling

Biofouling is associated with the development of biofilm on the FO membrane surface [26,

29, 41]. During FO operation, live microbes such as bacteria, fungi and microalgae attached

on the membrane surface can utilize nutrients existing in the FS to propagate [26, 29]. Over

time, these microorganisms can quickly proliferate and form the microbe layer colonizing the

membrane surface [26, 29]. This layer can be referred to as biofilm. Basically, two essential

components of biofilm consist of microorganisms and extracellular polymeric substances

(EPSs) comprising of polysaccharides, proteins, lipid and nucleic acid [26, 29]. EPSs are

excreted from the metabolism process of microbes to protect them from the external harsh

environment as well as enhance the adhesion to the membrane surface [26, 29, 41]. Compared

to the other fouling types, biofouling is more complex due to the interaction of live matters

[29]. Characteristics of representative foulants causing biofouling can provide further insights

into the nature of biofilm (Table 7) [26, 29, 41, 43].

The occurrence and severity of biofouling can be controlled by biofouling potential which is

represented by microorganisms and nutrient concentration in the aqueous environment [26].

Each factor plays a crucial role in each stage of the development of biofilm on the membrane

surface. The biofilm development can be divided into three stages, namely induction,

Page 43: Forward osmosis for resource recovery from anaerobically

28

logarithmical growth and plateau [26, 29]. The induction stage can be corresponding to the

attachment of microbes onto membrane surface [29]. Therefore, the concentration and type of

microbes can be pivotal in this stage [26, 29]. The second stage can imply the reproduction

and proliferation of microorganisms by consuming nutrients in the FS [29]. The third stage

can be indicative of detachment of bacteria due to lack of nutrients and increase of population

density [29]. Thus, the nutrient concentration and type can strongly influence the second and

third stages. To decrease biofouling, the reduction of nutrients supply is necessary [26].

Conventionally, nutrients existing in the FS can stem from three sources including pollutants

in the original bulk FS, biodegradable membrane materials and the reverse diffusion of draw

solutes as nutrients from the DS [26]. To conclude, the control of microorganisms and

nutrients in the FS can contribute to the biofouling mitigation.

Table 7. Properties of several typical bio-foulants in FO operation

Foulants Examples Size and shape Charge

Bacteria

Pseudomonas, E. coli,

Corynebacterium, B.

subtilis

Rod-shaped Neutral

Fungi Penicillium Brush-like and flask-shaped

- Trichoderma Divergent and flask-shaped

Microalgae Chlorella sorokiniana

Cell diameter ~ 5μm Negative Chlorella pyrenoidosa

Note: “-” indicates not available.

2.2.5. Factors affecting FO membrane fouling

2.2.5.1. Feedwater characteristics

The type and concentration of foulants and their physicochemical properties are pivotal

factors affecting the membrane fouling. As mentioned in Section 2.2.1, each type of foulant

will be responsible for a single fouling phenomenon. However, membrane fouling can be

deteriorated by inter-foulant-species interactions among mixed-foulants in the real FS due to

their physicochemical properties [26, 44]. The coexistence of both alginate and gypsum

foulants can result in the enhancement of membrane fouling because of the accelerated

gypsum crystal nucleation and growth in the presence of alginate [45]. Another example is

that the fouling behavior can increase in the appearance of alginate and lysozyme due to the

electrostatic interaction between the oppositely charged foulants [44]. In addition, the high

Page 44: Forward osmosis for resource recovery from anaerobically

29

concentration of foulants can contribute to the enhanced extent of initial fouling [26, 29, 42].

The reason is the increased foulant concentration can enhance the frequency of foulant

collision with the membrane, thereby increasing the rate of foulant aggregation on the

membrane surface [26].

The FS chemistry such as pH, ionic strength and ionic composition can cause significant

effects on fouling behavior [26, 42]. Similar to effects of pH and ionic strength on membrane

fouling in pressure-driven membrane processes, the FO membrane fouling caused by

negatively charged alginate, or humic acid can be aggravated at low pH and high ionic

strength [26, 42]. Also, protein fouling in the FO process can be more severe at pHIEP [26,

42]. The negative impacts of pH and high ionic strength on the FO membrane fouling can be

attributed to the reduced electrostatic repulsion among foulants and that between foulants and

the membrane surface [26]. Furthermore, the membrane fouling can be affected by the

existence of divalent ions such as Ca2+ and Mg2+ in the bulk FS due to foulant-ion and

membrane-ion interactions [26, 42, 46, 47]. Multivalent cations such as Ca2+ and Mg2+ can

bind to macromolecules strongly because these ions have an affinity for -COO- groups in the

molecular structure of organic compounds such as alginate, protein and organic acids [26, 42,

46]. This can partially neutralize the negative charge of these substances, thereby facilitating

their aggregation [26, 46]. Furthermore, divalent ions such as Ca2+ can connect two

macromolecules through the bridging effect [26, 42, 46]. For example, Ca2+ can play a role as

a bridging agent which can bind to carboxyl groups in extracellular polymeric substances to

link molecules [46]. Finally, these cations can influence the charge of polyamide – based

membrane surface through their interaction with -COO- groups in the membrane structure

[26].

2.2.5.2. Hydrodynamic conditions

Initial permeate flux and cross-flow velocities can influence fouling behavior through changes

in the structure of fouling layer, the influences of CP and mass transfer near the membrane

surface [26, 42, 47-49]. In general, the more severe membrane fouling can occur at higher

initial water permeate flux [26, 47, 49]. Three assumptions can be referred to explain for this

phenomenon [26, 42, 48]. Firstly, higher water permeate flux can increase the volume of

water permeating through the membrane, thereby increasing a number of foulants towards the

membrane surface due to water convection [26, 42]. Secondly, the exacerbated fouling can be

a consequence of the enhanced CP at high initial permeate flux [26, 48]. In the end, increase

in permeate flux can lead to greater hydrodynamic drag force towards the membrane surface,

Page 45: Forward osmosis for resource recovery from anaerobically

30

which can enhance the aggregation and compaction of foulants [26, 48]. In addition to initial

permeate flux, the increase of cross-flow velocity can mitigate the fouling propensity [26, 42,

48]. The reason is a high cross-flow velocity can increase the shear force and back diffusion

of foulants from the membrane surface to the bulk solution [26, 42, 48]. This can promote

mass transfer as well as reduce CP and the accumulation of colloids and salts on the

membrane surface [26, 42, 48].

Temperature can alter thermodynamic characteristics of feed and draw solutions as well as

hydrodynamic conditions, thereby causing impacts on the membrane fouling tendency [26,

50]. The increase of DS temperature which lead to the increased the osmotic pressure can

deteriorate the membrane fouling due to increase in the initial permeate flux [26, 48, 50]. In

the meantime, the decreased membrane fouling can be observed when heating the FS [26, 42,

48, 50]. This can be ascribed to the improvement of back diffusion of colloids from the

membrane surface as well as the increased foulant solubility at elevated temperature [26, 42,

48, 50]. Furthermore, temperature can be also an important parameter affecting the solubility

product of sparingly soluble salts in the scaling phenomenon [26, 48].

2.2.5.3. Membrane properties

The membrane intrinsic properties can indirectly affect membrane fouling via changes in

initial permeate flux and solution chemistry [17, 18, 26, 46]. A FO membrane with superior

separation properties (high water permeability and rejection) can induce high initial permeate

flux [17, 18]. As mentioned in section 2.2.5.1, this can exacerbate the membrane fouling due

to the increased hydrodynamic drag force. On the other hand, a membrane with inferior

intrinsic properties (low water permeability and rejection) can lead to increase in forward and

reverse draw solute diffusion between the FS and the DS [17, 18] . In many cases, the draw

solutes from the DS can be considered as nutrients for microorganisms [51]. This can

facilitate the biofouling development. Moreover, the draw solutes such as Ca2+ and Mg2+ can

raise the risk of scaling if they accumulate in the FS through the reverse solute diffusion [26,

46]. Also, effects of these ions on organic fouling are specified in section 2.2.5.1. Besides,

forward diffusion can result in the enhanced membrane scaling potential near the membrane

surface on the draw side [26].

In addition to the membrane separation properties, the surface properties of the FO membrane

such as morphology, wettability, surface functional groups and charge can directly impact

fouling behavior through the influence of foulant-surface interactions [26, 29, 42, 52, 53].

Membrane morphology implies the roughness of membrane surface. It is indicated that a

Page 46: Forward osmosis for resource recovery from anaerobically

31

smoother membrane surface may be less sensitive to fouling for two reasons [52]. The first

reason is the surface area of a smooth membrane for colloid-membrane interaction is

presumably smaller, compared to that of a rough membrane [26, 52]. Another reason is the

reduction of the shear force caused by the frictional resistance of the rough membrane can be

greater than that of the smooth membrane [26, 52]. Thus, the rate of foulants swept away from

the rough membrane surface can be low. In addition, the less severe fouling of the FO

membrane can be a result of using hydrophilic membrane materials [26, 29]. This can be

because of the large repulsive acid-base interaction that can prevent the attachment of colloids

on the membrane surface [26, 29]. Furthermore, functional groups such as –COO- groups

within the material structure of the FO membrane can bind to divalent cations such as Ca2+ to

form complexes [26, 29, 42, 46]. Consequently, a significant increase in the concentration of

calcium ions can promote the growth of gypsum on the membrane surface [42, 46]. Similarly,

the adhesion of alginate and the membrane surface can be intensified due to the bridging

effect of Ca2+ ions [42, 46]. Thus, organic fouling can become more severe in this case.

Noticeably, once most colloids existing in the FS are negatively charged, negatively charged

membranes can facilitate the removal of fouling because of the electrostatic repulsion [26, 29,

53]. However, these membranes can also be prone to potential ions such as Ca2+ or positively

charged colloids in the FS [26]. In summary, a FO membrane with a smooth, hydrophilic and

neutral surface and without carboxyl groups should be developed for the fouling prevention.

2.2.5.4. The composition of draw solution

The effects of DS composition on membrane fouling are via the impact of flux and changes in

the chemistry of the FS [26, 27, 38, 54]. The high concentration of DS can increase initial

permeate flux, which can lead to the increased membrane fouling as explained in section

2.2.5.2 [26, 38]. Moreover, the reverse draw solute diffusion can be more severe at a high DS

concentration, thereby potentially exacerbating the fouling tendency [26, 27, 38]. In addition

to the DS concentration, the type of draw solute can also elevate the membrane fouling via the

degree of reverse draw solute diffusion [26, 27, 38]. For example, a FO system operated with

CaCl2 solution as a DS can be more susceptible to biofouling, compared to that using a DS of

MgCl2 [55]. This is because of the complexation of calcium ions to EPSs of bacteria [55].

Another example is the utilization of glucose or ammonia carbonate as a DS can be a cause of

the enhanced biofouling because these draw solutes can play roles as nutrients for the microbe

development [26, 54].

Page 47: Forward osmosis for resource recovery from anaerobically

32

2.2.5.5. Membrane orientation

Membrane orientation can influence fouling behavior through the synergic effects of surface

properties, hydrodynamic conditions and CP [26, 56-58]. Many studies pointed out that the

less austere fouling was observed in AL-FS orientation compared to that in AL-DS

orientation [56-58]. The reason is the foulant deposition on the membrane surface can be

alleviated as the smooth dense AL faces the FS in the AL-FS orientation [17, 26]. In addition,

the disappearance of cross-flow velocity within the pores of the porous SL can aggravate the

internal fouling as well as ICP in AL-DS orientation [47, 56, 58]. Therefore, AL-FS

orientation can be superior compared to AL-DS orientation regarding the fouling mitigation.

However, AL-FS orientation can experience the more severe ICP, thereby reducing permeate

flux when compared with AL-DS orientation under otherwise identical conditions [59]. Thus,

the selection of appropriate membrane orientation should be taken account of the synergic

effects of membrane fouling and ICP rather than any single impact.

2.2.6. Effects of membrane fouling on the performance of FO process

The occurrence of membrane fouling can noticeably affect the FO membrane properties and

reverse solute flux [26, 60]. The membrane surface can be more negatively charged in the

presence of a humic acid fouling layer because of the plentiful carboxylic and hydroxyl

functional groups in humic acid molecules [60]. This fouled membrane can hinder the

negative ions dissociated from the draw solute molecules because of the electrostatic

repulsion [60]. Therefore, the reverse draw solute diffusion can be alleviated. Moreover, the

formation of this fouling layer can decrease the membrane solute permeability coefficient, but

insignificantly influence the membrane water permeability coefficient [60]. The reason is the

diffusion of water through the membrane is not significantly affected by the hydrated humic

acid layer on the membrane surface, while the solute diffusion can be restricted by steric

hindrance due to this layer [60]. The relieved reverse solute diffusion in the existence of the

membrane fouling layer can also be explained based on this mechanism [26]. In some cases,

membrane damage can result from the high concentration of acidic by-products at the

membrane surface [61, 62]. For example, cellulose acetate used to fabricate FO and RO

membranes can be biodegraded by fungi and bacteria [61].

In addition to the effects on the membrane properties, the membrane fouling can cause

impacts on water flux [26, 48, 60]. Two adverse consequences of membrane fouling are the

additional hydraulic resistance and the cake enhanced CP phenomenon and are considered as

major causes of flux decline [26, 41]. The former comes from a foulant cake layer established

Page 48: Forward osmosis for resource recovery from anaerobically

33

by the accumulation and deposition of foulants on the membrane surface [26, 60]. This can

reduce the overall water permeability of the fouled membrane [26]. The latter can also be

considered a result of this cake layer because this layer can hinder back diffusion of solutes

within it [26]. Thus, the solute concentration near the membrane surface can significantly

increase, which can enhance CP [26, 42]. As a result, the net driving force of water transport

is reduced due to the increased osmotic pressure of the solution near the membrane surface at

the feed side [26].

The accumulation of foulants on the FO membrane surface can either improve or decrease the

removal of pollutants and solutes due to different mechanisms [26, 49, 60, 63]. Several

explanations for the increased membrane rejection because of fouling can be elucidated as

follows. The first hypothesis is the membrane fouling can result in the pore blockage or seal

the molecular-scale defects of the FO separation layer [49, 63]. These changes can vary the

effective pore size of the FO membrane and impede the passage of foulants and solutes [49,

63]. Another hypothesis is related to the presence of an additional barrier known as a cake

layer on the membrane surface [49, 63]. This layer can enhance the steric hindrance to

solutes, resulting in the increase of membrane rejection [60]. On the other hand, the

membrane fouling can also decrease the membrane rejection due to the deteriorated fouling-

enhanced CP [26, 49]. Xie et al observed that the membrane rejection of trace organic

compounds witnessed a decline when the FO system was operated at high initial permeate

flux [49]. It is possible that the more compact and adhesive membrane fouling layer at high

initial permeate flux can exacerbate fouling-enhanced CP [49].

2.2.7. Mitigation of fouling in the FO process

2.2.7.1. Fouling characterization

Similar to fouling detection in RO, calculation of fouling potentials of the FS can help

forecast fouling during FO membrane operation. Foulant adhesion forces measurement can be

effective to evaluate the potential of organic fouling [64]. In addition, the determination of

carboxylic acidity can expose the concentration of carboxylic functional groups in foulants

such as alginate, humic acid and proteins [65]. As a result, the extent of interaction between

these foulants and calcium which can potentially aggravate the membrane fouling, can be

predicted [65]. Furthermore, the biofouling potential can be assessed by using bioassays that

include various techniques to determine the microbial growth potential of the FS [66].

However, once membrane fouling takes place on the FO membrane surface, ex-situ and off-

line methods can be essential for specifying the nature of fouling cake, facilitating the further

Page 49: Forward osmosis for resource recovery from anaerobically

34

preventative methods [17, 67]. Zeta potential and contact angle measurements can be used to

characterize the chemical properties of a fouled membrane such as surface charge and

wettability [67]. In addition, the morphology of the fouled membrane surface can be

determined by using ex-situ methods such as scanning electron microscopy and atomic force

microscopy [17, 67]. Also, some techniques namely adenosine triphosphate measurements,

extracellular polymeric substances quantification and confocal laser scanning microscopy can

be useful for biofilm characterization [17, 67].

Online and non-invasive monitoring of the performance of FO process is one of effective

methods to detect early signs of fouling in real-time [67, 68]. In-situ monitoring techniques

for membrane fouling can be divided into two groups [69]. The first group includes optical

methods such as direct observation technique, laser-based techniques, photo interrupt sensor,

particle image velocimetry, fluorescence-based technique, optical coherence tomography, and

membrane fouling simulator [67-69]. The other is associated with non-optical techniques

namely nuclear magnetic resonance imaging, X-ray micro-imaging, ultrasonic time-domain

reflectometry, electrical impedance spectroscopy and streaming potential measurement [67-

70].

2.2.7.2. Reduced-fouling methods

2.2.7.2.1. Pretreatment of the FS

Pretreatment methods applied to RO membrane system can be applicable to pre-treat the FS

during FO membrane operation [26]. These methods either pay attention to the removal of

foulants from the FS, or focus on the control of the feedwater chemistry such as pH, ionic

strength and specific ion composition to reduce the risk of membrane fouling in the FO

process [26, 71]. Many conventional and advanced technologies such as screening,

coagulation, flocculation, sedimentation, media filtration, microfiltration and ultrafiltration

can be applied to extract particles and microorganisms from the feedwater [26, 29, 71]. In

addition, the potential of membrane fouling can also be alleviated through the adjustment of

pH, disinfection with chlorine and the supplement of anti-scalants into the feedwater [26, 29,

71]. However, the use of anti-scalants can enhance the development of biofilm on the

membrane surface in some cases [71]. For example, polyphosphate-based anti-scalants can

play a role as phosphorous suppliers for the growth of microbe, thereby promoting biofouling

[72]. Thus, using anti-scalants in the FO system should be cautiously considered.

Characteristics and roles of each pretreatment method can be illustrated in Table 8 [26, 29, 71,

72].

Page 50: Forward osmosis for resource recovery from anaerobically

35

Table 8. FO pretreatment technologies in fouling control

Pretreatment

techniques Typical chemicals used Major roles

Screening No chemicals

Only mechanically separate

Remove large particles and

suspended solids

Disinfection Chlorine, sodium hypochloride

Prevent the growth of

microorganisms

Control biofouling

Coagulation-

flocculation and

sedimentation

Aluminum and ferric salts

Remove colloids, suspended solids

and dissolved organic carbon

Prevent organic and colloidal fouling

and biofouling

Media filtration Activated carbon, anthracite,

coal, glass fiber, expanded clay

Remove suspended solids and

dissolved organic carbon

Prevent organic and colloidal fouling

Membrane

filtration

Using microfiltration and

ultrafiltration

Can use coagulation and

flocculation simultaneously in

this technique

Remove suspended solids, colloids,

microorganisms, and

macromolecules

Prevent organic fouling, colloidal

fouling and biofouling

Anti-scalants

Polyacrylates,

organophosphonates, sodium

hexametaphosphate

Control scaling

2.2.7.2.2. Optimization of hydrodynamic conditions

Appropriate adjustments of initial flux, cross-flow velocity and membrane orientation can

alleviate risk of fouling. As elucidated in section 2.2.5.2, the membrane fouling propensity

can be more severe at high initial permeate flux. Meanwhile, the membrane fouling was

reported not to occur when the flux was under a certain value [56]. This initial flux value can

be referred to as a critical flux value [26]. Thus, the FO system should be operated at this

critical permeate flux. Furthermore, the intensification of cross-flow velocity is also proved to

reduce the membrane fouling and CP in section 2.2.5.2, and thus is preferred in the operation.

In addition to the increased cross flow velocity, the presence of spacers can also encourage

the effective cross flow velocity, turbulence and mass transfer near the membrane surface,

Page 51: Forward osmosis for resource recovery from anaerobically

36

thereby alleviating the membrane fouling [26]. As mentioned in section 2.2.5.5, the

membrane fouling occurred at FO mode is less severe than that at PRO mode. Nevertheless,

the initial water permeate flux attained at PRO mode is higher than that at FO mode. Thus, the

tradeoff between the reduction of membrane fouling and the achievement of high initial water

flux should be evaluated before selecting the most appropriate membrane orientation for

practical applications.

2.2.7.2.3. The use of relevant draw solution

Based on effects of the type and concentration of draw solute on fouling behavior, selection of

a suitable DS for the FO process can be beneficial for fouling control. As discussed in section

2.2.5.4, the increased concentration of DS can deteriorate the fouling tendency due to the

higher permeate flux and reverse draw solute flux. Thus, the introduction of critical DS

concentration which is corresponding to critical permeate flux can be useful for mitigating the

membrane fouling [73]. The critical DS concentration is indicative of the threshold DS

concentration below which fouling is not noticeable [73]. In addition, selected draw solutes

should not behave as promoters, fouling precursors and nutrients which can facilitate the

enhancement of fouling. Furthermore, draw solutes should possess low solute permeability to

minimize the reverse draw solute diffusion through the membrane, which can potentially

exacerbate the membrane fouling propensity.

2.2.7.2.4. Modification of membrane properties and structure

Two aspects including surface modification and renovation of membrane materials can be

considered to create a superior anti-fouling FO membrane [17, 26, 29, 48]. To date, most

studies have focused on modifying the dense AL of FO membrane, while very few studies

have paid attention to the modification of the porous SL as well as found out novel membrane

materials [17, 26, 48]. The attempts to alter the AL can be listed as increasing surface

hydrophilicity, reducing the surface roughness, decreasing charge density, reducing carboxyl

groups in the material structure, embedding nanoparticles to prevent biofouling and forming a

second nanofiltration layer on the support layer side [17, 26, 48]. Details of some membrane

alteration techniques can be found in Table 9 [74-79].

Page 52: Forward osmosis for resource recovery from anaerobically

37

Table 9. Several modification techniques of the FO membrane surface to mitigate the membrane fouling

Membrane Modification technique Characteristics of the modified membrane surface

Cel

lulo

se

Tri

acet

ate

The porous side of the membrane is coated by poly amino acid 3-(3,4-

Dihydroxyphenyl)-L-alanine, a zwitterionic polymer Increase the membrane hydrophilicity

Silver particles are deposited on the membrane surface by a photoinduced

approach before charge-driven self-assembly of titanium dioxide (TiO2)

nanoparticles are attached onto the layers of silver nanoparticles

Increase anti-biofouling

TF

C p

oly

amid

e

In-situ grafting Jeff amine (an amine-terminated poly-ethylene glycol

derivative) to dangling acyl chloride surface groups on the nascent PA AL

More hydrophilic surface

Decrease the membrane roughness

Decrease the number of carboxyl groups

Less negative surface charge

Functionalize the AL of a TFC PA membrane with poly-ethylene glycol

(PEG)

Increase the membrane hydrophilicity

The PEG layer can impede foulant adsorption.

Reduce the number of carboxyl groups

Grafting polyzwitterions onto the PA membrane via click chemistry

Specific binding sites such as carboxyl groups on the

membrane surface are shielded

Enhance hydrophilic and steric repulsion

Using a dip coating method to bind free carboxyl groups of the membrane

surface with super hydrophilic silica nanoparticles which are

functionalized

Super hydrophilicity

Neutralized carboxylic groups

Covered by positively charged nanoparticles

Page 53: Forward osmosis for resource recovery from anaerobically

38

2.2.7.3. Membrane fouling control

2.2.7.3.1. Physical cleaning

Flushing and osmotic backwashing are the most widely-studied methods to physically clean

the fouled FO membrane. The basis of flushing is to utilize the high shear force generated by

increasing cross-flow velocity of cleaning solution to dislodge and remove fouling layers

from the membrane surface [67, 80, 81]. The introduction of air bubbles into the cleaning

solution for air scouring may also help to enhance the efficiency of cleaning process [26, 80,

81]. Schematic diagram of flushing process can be depicted in Figure 5 [80]. The rate of

cross-flow is an important parameter affecting the effectiveness of flushing. Flux recovery can

reach up to 98% after 15 min of flushing with deionized water (DI water) at the increased

cross-flow velocity from 8.5 cm/s to 21 cm/s without any chemicals used [81]. Apart from

cross-flow velocity, the efficiency of the flushing process can be influenced by other factors

such as flushing solution, cleaning duration, use of air bubble and cleaning frequency of the

FO system [80, 81]. However, flushing can only remove foulants on the membrane surface,

except for those within the porous support layer [26]. In this case, backwashing can be

effective to clean the fouled membrane [26].

Figure 5. Flow direction in forward and reverse flushing in FO mode

Osmotic backwashing has justified its efficiency of membrane cleaning in lab-scale tests [26,

82]. Osmotic backwashing can be performed by replacing the original FS with a high salinity

solution and the DS with a low salinity solution [80, 82]. This act aims to create a reverse

permeate flow from the draw side to the feed side to detach foulants both on the membrane

Page 54: Forward osmosis for resource recovery from anaerobically

39

surface and within the porous SL and move them away from the membrane (Figure 6) [26, 40,

80, 82]. It is reported that the implementation of osmotic backwashing can allow recovery of

more than 85% of the initial flux [82]. The cleaning efficiency of the osmotic backwash

process is dependent on the backwashing frequency and duration as well as intensity of

backwash and cross-flow velocity [40, 80, 82]. In addition to osmotic backwashing, hydraulic

backwashing is supposed to remove the foulant layer effectively. A recent study demonstrated

that the permeate flux restoration could reach nearly 100% when hydraulic backwash was

applied to clean gypsum scaling within the porous SL [83]. However, the FO membrane is not

designed for high-pressure operation. Therefore, further studies need to be conducted to verify

the applicability of hydraulic backwashing.

Figure 6. The permeate flow direction during backwashing in PRO mode

2.2.7.3.2. Chemical cleaning

Chemical agents such as acids, bases, chelators, surfactants and disinfectants can be utilized

to remove and prevent the FO membrane fouling effectively [26, 67, 80]. Although physical

cleaning methods can be justified having the promising efficiency of fouling mitigation

during FO operation, they still face difficulties in removal of the irreversible fouling [26, 67,

80]. The foulant-surface interaction such as adhesion force is quite high in the occurrence of

organic fouling and biofouling [26, 67]. Thus, it is difficult for physical cleaning to break this

interaction and completely dislodge foulants from the membrane surface. In this case,

membrane fouling can be overcome by chemical cleaning. In this type of cleaning, the choice

of cleaning agent is critical and dependent on the membrane material and type of foulant [80].

An appropriate chemical agent must ensure both the effectiveness of removal of fouling and

Page 55: Forward osmosis for resource recovery from anaerobically

40

no damage to the membrane surface [80]. In practice, the cleaning agents for FO can resemble

that for RO. Some acids such as nitric, phosphoric, hydrochloric, sulfuric and citric can be

used to remove precipitate and scaling, while some bases namely NaOH, KOH can be useful

for removing organic fouling [26, 67, 80]. In addition, some oxidants and disinfectants such

as NaOCl, ozone, H2O2, KMnO4 and EDTA can be suitable for biofouling removal [26, 67,

80]. Several common cleaning agents according to fouling type can be illustrated in Table 10

[26, 67, 80].

Table 10. Several typical cleaning chemicals according to the type of foulant.

Type of foulant Cleaning agents

Colloids NaOH, chelating agents and surfactants

Organic compounds NaOH, chelating agents and surfactants

Metal oxides Citric acid or Na2S2O4

Silica NaOH

Scalants such as CaCO3, CaSO4 Citric acid, HCl and EDTA

Biofilm NaOH, chelating agents, EDTA, surfactants and

disinfectants

2.2.7.3.3. Novel fouling prevention approaches

Three novel solutions of fouling control including quorum quenching, enzymatic cleaning and

energy un-coupling can be adopted to mitigate the membrane fouling in FO operation [26, 84-

88]. Even though physical and chemical cleaning methods can remove reversible foulants

effectively, they can also cause changes in physicochemical properties of the membrane as

well as even reduce its mechanical stability. For example, oxidizing agents such as chlorine

can damage polyamide TFC FO membranes [89]. Moreover, the implementation of flushing,

backwashing and chemical cleaning require the use of clean water, energy consumption as

well as cause negative effects of used chemicals on the environment [84-86]. Therefore, three

above-mentioned advanced cleaning methods can be expected to overcome these detrimental

impacts. Quorum quenching is a technique to control biofouling through the interruption of

quorum sensing [84, 86]. Quorum sensing is a biological communication mechanism among

microorganisms via signal molecules such as Acylhomoserine lactone (AHL) [84, 86]. These

molecules are produced from microbe’s metabolism [84, 86]. This mechanism can facilitate

microbe’s ties and thus enhance the formation of biofilm on the membrane surface [84, 86].

Therefore, quorum quenching techniques to interrupt or eliminate these signal molecules can

be effective to prevent biofouling [26, 86]. Several approaches to disrupt the process of

Page 56: Forward osmosis for resource recovery from anaerobically

41

quorum sensing can be listed as: (i) suppressing the production of signal molecules; (ii)

degrading AHL; (iii) reducing the activity of AHL protein or AHL synthase; (iv) mimic the

signal molecules by synthetic or natural compounds [86]. For example, natural inhibitors such

as AHL-lactonase from Bacillus cereus and AHL-acylase from Tenacibaculum can degrade

AHLs successfully [86].

In addition to quorum quenching, enzymatic cleaning can be applied to control the membrane

fouling, especially organic fouling. According to this method, various species of enzyme such

as protease, lipase and α-chymotrypsin can be employed to decompose organic compounds

such as proteins, lipids and humic acid on the membrane surface [85, 88]. The type of

enzyme, enzyme concentration and cleaning time are three major factors affecting the

efficiency of enzymatic cleaning [85, 88]. Furthermore, energy uncoupling method is also

useful for biofouling control through the disruption of microbial metabolism energy [87]. It

appears that the development of biofilm is closely connected to bioenergy existing under the

form of adenosine triphosphate (ATP) [87]. This means biofilm can be eliminated by using

chemical un-couplers such as 2,4-dinitrophenol to inhibit ATP synthesis [87]. The halted ATP

synthesis can not only stop any microbial activities, but also contribute to the disruption of

signal molecules which are means of biological communication [87]. As a result, the

membrane biofouling can be controlled effectively [87].

Page 57: Forward osmosis for resource recovery from anaerobically

42

CHAPTER 3

PERFORMANCE OF A SEAWATER-DRIVEN FORWARD OSMOSIS

PROCESS FOR PRE-CONCENTRATING SLUDGE CENTRATE:

ORGANIC ENRICHMENT AND MEMBRANE FOULING

Page 58: Forward osmosis for resource recovery from anaerobically

43

3.1. Introduction

Forward osmosis is a robust separation platform capable of treating highly complex solutions

that are not suitable for conventional membrane processes [9, 17]. In FO, mass transfer

through the membrane is osmotically driven. Thus, when a DS is readily available, the FO

process can occur with very low energy input [17, 18, 90, 91]. The absence of external

hydraulic pressure can also explain the low membrane fouling tendency and excellent fouling

reversibility of FO. In recent years, the potential of FO to treat many complex streams has

been demonstrated in the literature. These complex solutions include drilling and fracking

fluids from oil and gas exploration [92-95], sludge [96, 97], sludge centrate [1, 9, 14], and

municipal wastewater [13, 91, 98].

A major obstacle to full scale deployment of FO is the lack of a suitable DS [17]. Issues

associated with cost of the draw solutes, regeneration, and loss of draw solutes due to reverse

diffusion can increase the operating cost, thereby hindering the feasibility of FO applications

[17, 28, 37]. In this context, seawater, which is abundant and cheaply available in coastal

areas, has been increasingly considered as a potential DS [17, 37]. The diluted seawater

released from the process can be returned to the sea, and thus, DS regeneration is not

necessary.

In a typical wastewater treatment plant, the sludge is anaerobically digested. The digested

sludge is then dewatered to obtain biosolids for land application. The liquid from this

dewatering process is called sludge centrate, which has a high content of suspended solids,

nutrients, and organic matter [14]. Due to the difficulties associated with the treatment of this

sludge centrate, in most cases, it is returned to the headworks of the treatment plant. The

recirculation of untreated sludge centrate to the headworks leads to additional organic and

nutrient loading, and deprive the plant from any opportunities for energy and nutrient

recovery [1, 2].

The use of FO to pre-concentrate sludge centrate has been investigated in several recent

studies [9, 14]. However, draw solutes such as MgCl2 and NaCl are expensive and must be

regenerated. On the other hand, seawater appears to be a particularly promising DS for pre-

concentrating sludge centrate. Ansari et al. [1] has recently demonstrated a seawater-driven

FO process for phosphorus recovery from sludge centrate with a specific focus on evaluating

the efficiency of nutrient recovery. Results from the previous study suggest that identifying

the most suitable membrane and orientation is necessary to ensure the best performance of

seawater-driven FO process for pre-concentrating sludge centrate. More importantly,

understanding of the fouling process and developing strategies to control fouling need to be

discussed to guarantee the long-term operation of the FO system. It is also necessary to

consider all other factors affecting water flux and membrane fouling, such as membrane pre-

wetting and DS.

This study aims to elucidate the effects of membrane materials, prewetting procedures and DS

on the performance of seawater-driven forward osmosis for pre-concentrating organic matter

Page 59: Forward osmosis for resource recovery from anaerobically

44

in the sludge centrate. The performance of the FO process is observed in terms of chemical

oxygen demand (COD) enrichment, membrane fouling and flux recovery by physical

cleaning. Comparison between seawater and analytical grade NaCl as the DS was made to

highlight the potential and challenges of using seawater for enriching COD in sludge centrate.

3.2. Materials and methods

3.2.1. Forward osmosis system and membranes

A lab-scale cross-flow FO system (Figure 7) was used. The FO system included a membrane

cell, two variable speed gear pumps (Micropump, Vancouver, Washington, USA),

conductivity and temperature controllers, and a digital balance to measure the flux. The FO

membrane cell consisted of two symmetric rectangular chambers for the feed and draw

solutions, respectively. The internal dimensions of each chamber were 10 cm in length, 5 cm

in width and 0.2 cm in height. The system was operated in the counter-current mode. The FO

membranes were positioned either in AL-FS orientation, or AL-DS orientation.

Flat sheet FO membranes were obtained from Hydration Technology Innovations (HTI,

Albany, OR) and Porifera, Inc. (Hayward, California, USA). The HTI membrane had an

asymmetric structure and was made of CTA with an embedded polyester mesh for mechanical

support. The Porifera membrane was a TFC membrane consisting of a PA layer on a

microporous polysulfone SL. Key properties of the HTI and Porifera membranes were

summarised in Table 11.

Table 11. Key properties of the AL of the FO membranes [99, 100].

Membrane HTI-CTA Porifera-PA

Water permeability (A) (L/m2.h.bar) 0.84 3.2

Salt (NaCl) permeability (B) (L/m2.h) 0.32 0.41

Structural parameter (S) (mm) 0.57 0.46

Contact angle (o) 61 49.5

Surface roughness (nm) 3.8 57.4

Zeta potential at pH = 7 (mV) -5 -16

Page 60: Forward osmosis for resource recovery from anaerobically

45

Figure 7. Schematic diagram of the lab-scale FO system.

3.2.2. Feed solution and draw solution

In this study, seawater and NaCl solutions were used as DSs. Seawater was collected from

North Wollongong Beach (Wollongong, New South Wales, Australia) and was filtered using

filter paper with a pore size of 1 μm prior to the experiments. NaCl solutions of 0.25, 0.5, 0.75

and 1 M were prepared using analytical grade NaCl and deionised (DI) water. Anaerobically

sludge centrate was collected from a high-speed centrifuge at the Shellharbour Wastewater

Treatment Plant (Shellharbour, New South Wales, Australia). The sludge centrate was pre-

filtered using a 0.2 mm plastic screen to remove any large objects. The compositions of

seawater and sludge centrate were summarised in Table 12.

Table 12. Composition of seawater and sludge centrate (Values indicated average ± standard

deviation of three samples).

Parameters Unit Seawater Sludge centrate

pH - 7.3 ± 0.2 7.1 ± 0.1

Electrical conductivity mS/cm 44.2 ± 0.3 6.8 ± 0.1

Osmotic pressure bar 28.1 ± 0.6 6.5 ± 0.1

Total solids g/L 31.7 ± 2.8 1.6 ± 0.1

COD mg/L - 420.3 ± 15.5

Ammonia mg/L - 520 ± 2.6

Phosphorus mg/L - 371.5 ± 1.4

Page 61: Forward osmosis for resource recovery from anaerobically

46

3.2.3. Membrane prewetting

Prewetting was conducted by soaking in a solution containing 70% ethanol and 30% water for

45 min. Following soaking, the membrane was rinsed and preserved in DI water overnight

prior to filtration experiments.

3.2.4. Experimental protocols

All FO experiments were conducted in four steps. In the first step, DI water was used as the

FS for 30 min to determine the pure water flux. In the second step, DI water was substituted

with sludge centrate and the FO experiment was conducted until a water recovery of 55% had

been achieved. Duration of this second step varied from experiment to experiment ranging

from 24 to 120 hours. At specified time intervals, 10 mL samples were collected from the

feed for analysis. In the third step, the draw and feed solutions were replaced by DI water to

facilitate membrane cleaning by flushing at a cross-flow velocity of 24 cm/s for 20 min. In the

last step, the pure water flux was determined again using DI water under identical

experimental conditions as in the first step. In all experiments, initial volumes of the feed and

draw solutions were 1 and 3 L, respectively. The circulation flow rate of the feed and the draw

solution was 0.8 L/min (i.e., cross flow velocity of 13 cm/s).

3.2.5. Calculations

Water flux (Jw) was calculated based on the change in weight of DS, and expressed as in

Equation 3:

( )1

1

i it t iw

i i m i m

m m mJ

t t A t A −

− = =

− (Equation 3)

In which: Δmi: the change in weight of DS over a time interval (g)

Δti: a time interval (h)

ρ: the solution density (g/cm3)

Am: the effective membrane area (m2)

Water recovery was determined based on the ratio of the cumulative permeate volume and the

initial volume of FS, and presented as in Equation 4:

Water recovery (%) = 0 0

0100

t

m w

initial

A J dt

V

(Equation 4)

In which: Jw: the observed water flux at time t

Vinitial: the initial volume of FS

Page 62: Forward osmosis for resource recovery from anaerobically

47

The draw solute flux (Js) was calculated by Equation 5:

2 1, 2 , 1f t f f t f

s

m

C V C VJ

t A

− =

(Equation 5)

In which: Cf,t2: The concentration of draw solute in FS at time t2 (g/L)

Vf2: The volume of FS at time t2 (L)

Cf,t1: The concentration of draw solute in FS at time t1 (g/L)

Vf1: The volume of FS at time t1 (L)

Δt: a certain period of filtration time (h)

3.2.6. Analytical methods

Key water quality parameters of sludge centrate and seawater were measured according to

standard methods. COD was determined using a Hach DRB200 COD Reactor and Hach

DR3900 spectrophotometer following the US-EPA Standard Method 5220. Temperature and

pH of solutions were measured by an Orion 4-Star Plus pH/conductivity meter (Thermo

Scientific, Waltham, MA).

The surface characteristics of the FO membranes were characterized using scanning electron

microscopy (SEM) (JOEL, JSM-6400LV, Japan). Prior to taking SEM images, coupon

membrane samples were coated with a thin layer of gold.

3.3. Results and discussions

3.3.1. Pure water flux under different conditions

Figure 8. Pure water fluxes of CTA and PA FO membranes using seawater as DS in AL-FS

and AL-DS orientations with and without prewetting.

Page 63: Forward osmosis for resource recovery from anaerobically

48

Under the same experimental conditions, the TFC PA membrane showed a higher pure water

flux than that of the CTA membrane (Figure 8). This observation could be explained by the A

value (water permeability under a hydraulic pressure) and the structure parameter (S value) of

these two membranes (Table 11). Indeed, the A value of the TFC PA membrane was

approximately four times higher than that of the CTA membrane. The structural parameter

value of the TFC PA membrane (460 μm) was slightly lower than that of the CTA membrane

(570 μm) [99]. Previous studies have demonstrated that a smaller structural parameter resulted

in less severe ICP and thus higher water flux [17, 99]. It is noteworthy that in the FO process,

CP could also influence the water flux. Thus, the difference in pure water flux between the

TFC PA and CTA membranes was not necessarily proportional to the difference in their A

and S values.

The AL-DS orientation exhibited a higher pure water flux compared to the AL-FS orientation.

The difference in water flux between these two orientations was considerably less than the

comparison between the TFC PA and CTA membranes discussed above. Indeed, the higher

water flux under the AL-DS orientation compared to the AL-FS orientation could be solely

attributed to a less severe ICP condition [17, 33, 101].

Prewetting significantly improved the pure water flux of the TFC PA membrane but had a

negligible impact on the CTA membrane. As a result of prewetting, water flux of the PA

membrane increased by 29.7% and 59.7% under the AL-FS and AL-DS orientation,

respectively (Figure 8). There were two possible reasons for this notable increase in flux by

the TFC PA membrane after prewetting, namely swelling of the PA skin layer and prewetting

of the polysulfone SL. The PA skin layer could swell in alcohol causing an increase in the

effective pore size, and thus, increased the water flux [100]. However, there was no

discernible increase in the reverse salt flux due to prewetting. The reverse draw solute fluxes

of the PA membrane and the prewetted PA membrane were all 3.38 g/m2.h. Thus, swelling

was not expected to be a major reason for the improvement in water flux observed here. The

polysulfone SL was hydrophobic and could not be completely wetted by water [102].

Compared to water, ethanol had a lower surface tension, thus, could easily penetrate into the

porous structure and prewet the pores of the polysulfone SL for subsequent water permeation

[102, 103]. On the other hand, the CTA membrane was asymmetric and readily hydrophilic

because of the presence of hydroxyl groups. Thus, the membrane can be fully hydrated by

soaking in the DI water overnight (Figure 8) [104].

Page 64: Forward osmosis for resource recovery from anaerobically

49

3.3.2. Pre-concentration of sludge centrate by forward osmosis

Figure 9. The performance of FO for enrichment of COD in sludge centrate using seawater as

the DS. The theoretical COD concentration factor was calculated assuming 100% COD

retention by the FO membrane.

As the sludge centrate was concentrated by FO, COD concentration increased proportionally

(Figure 9). In all cases, the COD concentration factor was lower than the theoretical value

assuming complete COD retention. The observed difference between the COD concentration

factor and theoretical value could possibly be attributed to the deposition of particulate COD

materials on the membrane surface. In fact, there was a correlation between membrane

fouling and the COD enrichment results in Figure 9 as discussed further in section 3.3.3.

The best performance in terms of COD enrichment was from the CTA membrane with the

AL-FS orientation (Figure 9). The AL of the CTA membrane had a lower surface roughness

than its own SL as well as that of the TFC PA membrane [26, 52, 100, 105]. Thus, due to the

hydrodynamic drag force from the cross flow, the deposition of organic substances on the

CTA membrane was expected to be less compared to the TFC PA membrane. In addition,

surface chemistry interaction between organic matter and the membrane surface could also be

a reason for the lower organic enrichment of using the PA membrane. The PA membrane with

the considerable number of highly polar carboxylic functional groups on its surface had

significant affinity with organic colloids in the FS [100]. As a result, the accumulation of

organic matter on the PA membrane surface could be enhanced, thus reducing COD

concentration in the bulk feed. It is noted that this combination (CTA under AL-FS

orientation) also had the lowest initial water flux. However, it appears that water flux did not

affect COD enrichment performance as demonstrated below.

Page 65: Forward osmosis for resource recovery from anaerobically

50

Figure 10. The performance of FO for enrichment of COD in sludge centrate using the NaCl

solution as the DS and the CTA membrane in the AL-FS orientation. Note: The theoretical

COD concentration factor was calculated assuming 100% COD retention by the FO

membrane.

DS concentration did not affect the enrichment of COD by FO (Figure 10). No significant

difference in COD enrichment was observed when the NaCl DS concentration increased from

0.25 to 1 M (Figure 10). The initial water flux was proportional to the DS concentration.

Thus, the results in Figure 10 also suggested that water flux did not affect COD enrichment as

discussed above.

Seawater showed comparable COD enrichment performance to that of analytical grade NaCl

as the draw solutes. The osmotic potential of seawater was similar to that of the 0.5 M NaCl

solution [106, 107]. Seawater was readily available in coastal areas and thus it was a low-cost

DS. However, in addition to NaCl, seawater contained many other salts. Some of them were

sparingly soluble and might cause membrane scaling as further discussed in section 3.3.3.2.

Page 66: Forward osmosis for resource recovery from anaerobically

51

3.3.3. Factors affecting fouling behaviour

3.3.3.1. Membrane properties and orientations

Figure 11. Fouling behaviour of the FO membranes in (A) AL-FS orientation and (B) AL-DS

orientation using seawater as DS.

In all cases, membrane fouling was significant as indicated by significant water flux decline

during COD enrichment (Figure 11). The decrease in water flux could be mostly attributed to

the formation of a cake layer on the membrane surface. This cake layer caused an additional

hydraulic resistance and increased CP, thus reducing water flux.

In good agreement to the data in Figure 9, the CTA membrane was less susceptible to fouling

than the TFC PA membrane regardless of the membrane orientation. As described above, this

result was likely due to the smooth surface of the CTA membrane. The higher roughness and

prominent ridge-and-valley structure of the PA membrane surface could exacerbate the

deposition of foulants, thus more severe fouling [100]. Additionally, a high density of

carboxylic functional groups in the structure of the PA membrane could be potentially

vulnerable to fouling [108]. In the presence of Ca2+ ions, carboxyl acid functional groups

Page 67: Forward osmosis for resource recovery from anaerobically

52

could lead to bridging of membrane surface and Ca2+-organic foulants, thus probably

exacerbating organic fouling.

The AL-FS orientation showed less fouling than the AL-DS orientation. There were several

possible explanations. As mentioned above, the lower roughness of the AL in the AL-FS

orientation could alleviate the accumulation of foulants on the membrane surface. FO

operation under the AL-DS orientation was susceptible to internal clogging since organic

molecules could readily penetrate the porous SL. In addition, the high reverse solute diffusion

in the AL-DS orientation could increase the osmotic potential of the feed, decreased effective

driving force, and thus, reduced the water flux [33].

3.3.3.2. Draw solution

Figure 12. Comparison of fouling behaviour towards CTA membranes in the AL-FS

orientation using seawater and the NaCl solutions at different concentration as DSs.

As discussed in the previous sections and further discussed in section 3.3.4, the CTA

membrane under the AL-FS orientation demonstrated the best suitability for enriching organic

matter in sludge centrate. Thus, only the CTA membrane under the AL-FS orientation was

used for further investigating the effect of draw solution on the performance of the process.

Increasing the DS concentration led to a higher initial water flux (Figure 15), but no

significant impacts on membrane fouling (Figure 12). The elevated initial water flux was a

result of the increased driving force of water transport due to an increase in the concentration

gradient along the membrane. However, the extent of fouling in all cases was nearly the same.

This was probably because of the smoothness of the CTA membrane surface that could

effectively minimize the accumulation and deposition of foulants. In addition, this could be

explained using a theory of ‘critical DS concentration’. According to this concept, fouling

could be less severe at below the critical DS concentration [58, 73]. It is possible that the used

Page 68: Forward osmosis for resource recovery from anaerobically

53

DS concentrations in this study were lower than the critical DS concentration, thus, the

impacts of DS concentration on fouling were insignificant.

Seawater exhibited similar fouling to that from the 0.5 M NaCl solution as the DSs. This

observation was consistent with the data shown in Figure 10, and thus could also be explained

by the same reason as referred to earlier. However, it is noted that the flux profile when using

seawater was less stable than that of using the 0.5 M NaCl solution. Multivalent ions in

seawater, such as Ca2+, Mg2+, SO42- and PO4

3- could act as fouling promoters, and scaling

precursors through the reverse solute diffusion, thus increasing fouling potential.

3.3.3.3. Fouling layer characteristics

Figure 13. SEM images of the AL of (A) the fouled CTA membrane and (B) the fouled TFC

PA membrane in the AL-FS orientation using seawater DS.

A notable contrast in the morphology of the fouling layer on the CTA and TFC PA

membranes could be observed (Figure 13). The fouling layer on the AL surface of the CTA

membrane was loose and soft (Figure 13A) while that of the PA membrane was dense and

firm (Figure 13B). The observed irregular shape and size of crystals and organic cake-layer on

the membrane surface suggested the presence of both organic and inorganic foulants.

Page 69: Forward osmosis for resource recovery from anaerobically

54

3.3.4. Cleaning and flux reversibility

Figure 14. Flux recovery of the fouled CTA, PA and prewetted PA membranes using seawater

DS after physical flushing.

As expected, flux recovery by flushing in the AL-FS orientation was higher than that in the

AL-DS orientation (Figure 14). Flux recovery was proportional to the efficiency of foulant

removal from the membrane surface. In the AL-FS orientation, the deposition of foulants on

the membrane was a surface phenomenon and the fouling cake could be readily removed by

shear force from flushing. On the other hand, in the AL-DS orientation, the FS was in contact

with the SL and due to pore clogging of the SL, flux recoverability was much lower than in

the AL-FS orientation (Figure 14).

The highest flux recovery (95%) was observed with the CTA membrane in the AL-FS

orientation (Figure 14). Together with the high COD enrichment performance shown in

Figure 9, this result suggested that the CTA membrane in the AL-FS orientation was the most

suitable for sludge centrate. This result was also consistent with the loose and soft fouling

layer of the CTA membrane under the AL-FS orientation previously shown in Figure 13A.

Flushing was not efficient in restoring the water flux of the PA and prewetted PA membranes.

There were two possible reasons. Firstly, foulants deposited on a rough surface of the PA

membrane could be sheltered from cross-flow shear force, thus decreasing the number of

foulants detached from the membrane surface. Secondly, the highly polar carboxylic

functional groups on the PA membrane surface were available for ionic bonding with

foulants, thus improving foulant adhesion [100]. In contrast, the CTA membrane was only

slightly negatively charged and did not have free carboxyl functional groups that could

interact with the foulants [100].

Page 70: Forward osmosis for resource recovery from anaerobically

55

Figure 15. Comparison of pure water fluxes and flux recovery of CTA membranes in the AL-

FS orientation using seawater and NaCl DS.

Increasing the DS concentration resulted in a higher pure water flux but also reduced flux

recovery by flushing (Figure 15). Indeed, as the NaCl concentration increased from 0.25 to 1

M, flux recovery decreased from nearly 100% to 70%. Since the DS concentration was

proportional to water flux, results in Figure 15 showed that the extent of irreversible fouling

(by flushing) was inversely correlated to the initial flux. A denser and more compact fouling

was formed at higher initial flux, thereby, impairing the efficiency of flushing.

Seawater had a similar osmotic potential to a 0.5 M NaCl solution. However, a slightly lower

flux recovery was observed when using seawater as the DS compared to the 0.5 M NaCl

solution. The complex composition of seawater could result in a less reversible fouling layer.

Indeed, multivalent cations (such as Ca2+ and Mg2+) in seawater could exacerbate fouling and

render fouling layer more adhesive, thus lowering the flux recoverability by flushing.

Page 71: Forward osmosis for resource recovery from anaerobically

56

Figure 16. SEM images of (A) the AL of the cleaned CTA membrane in the AL-FS

orientation using seawater as DS, (B) the AL of the cleaned CTA membrane in the AL-FS

orientation using the 0.5 M NaCl solution as DS, (C) the SL of the cleaned CTA membrane in

the AL-DS orientation using seawater as DS, (D) the AL of the cleaned PA membrane in the

AL-FS orientation using seawater as DS and (E) the SL of the cleaned PA membrane in the

AL-DS orientation using seawater as DS.

The complex composition of seawater and the membrane characteristics were observed to

govern the cleaning efficiency of flushing (Figure 16). After flushing, the CTA membrane

active surface in the AL-FS orientation (seawater as DS) still had some salt crystals and

organic particles (Figure 16A). In contrast, when using the 0.5 M NaCl solution, the

Page 72: Forward osmosis for resource recovery from anaerobically

57

membrane was mostly clean after flushing (Figure 16B). This was consistent with the data in

Figure 15 that indicated how the complex composition of seawater resulted in a more

adhesive fouling layer.

In addition, SEM images indicated that both for CTA membrane under the AL-DS orientation

(Figure 16C) and PA membrane in the AL-FS orientation (Figure 16D) flushing removed

crystals significantly, but the organic layers only partly. Also, simple flushing seemed to be

inefficient in removing both crystals and organic layers towards the PA membrane in the AL-

DS orientation. These observations resulted from synergistic effects of membrane roughness

and chemical structure of these membranes on fouling as discussed above.

3.4. Conclusions

Results from this study demonstrated the potential of the FO process to enrich COD in sludge

centrate. Compared to the TFC PA membrane, the CTA membrane in the AL-FS orientation

showed a much better COD enrichment efficiency, lower fouling, and higher flux recovery by

simple flushing. There was a correlation between membrane fouling and COD enrichment

efficiency. In other words, COD enrichment efficiency decreased when organic matter

accumulated on the membrane surface, causing fouling. The results also showed that

membrane fouling was not affected by the DS concentration (or initial water flux) possibly

because of the low initial flux in this study, however, flux recovery by membrane flushing

decreased as the initial water flux increased. Seawater was a potentially low-cost and effective

DS for COD enrichment. However, compared to NaCl, seawater as the DS led to more severe

membrane fouling and lower flux recovery by flushing.

Page 73: Forward osmosis for resource recovery from anaerobically

58

CHAPTER 4

SEAWATER-DRIVEN FORWARD OSMOSIS PRE-CONCENTRATION

OF NUTRIENTS IN SLUDGE CENTRATE: PERFORMANCE AND

IMPLICATIONS

Page 74: Forward osmosis for resource recovery from anaerobically

59

4.1. Introduction

Phosphorus is an essential element for all living organisms on earth [109]. It is a vital

component of biological molecules, such as adenosine triphosphate, deoxyribonucleic acid

and ribonucleic acid [109]. Also, phosphorus can be found in all living beings, especially in

bones and teeth [109]. Human beings intake phosphorus via daily meals with agricultural

products predominantly cultivated using phosphate fertilizers [1]. However, minable

phosphorus will be exhausted in the near future [110]. This threatens food security and a

source of renewable phosphorus supply is essential for future generations. On the other hand,

phosphorus, along with ammonia are also potential pollutants at their excessive concentration

in water bodies [109]. Thus, the phosphorus and ammonia concentrations in the effluent of

WWTPs are strictly regulated. Consequently, environmental concerns, regulatory need for

phosphorus and ammonia removal and the severe depletion of global phosphorus reserves

lead to the idea of recovery of nutrients from wastewater and its by-products in WWTPs [1,

111, 112].

In a typical WWTP, sludge centrate contains phosphate and ammonia concentrations in the

ranges of 75 – 371 mg/L and 250 – 532 mg/L, respectively [109, 111, 113]. This stream is

usually returned to the head of work in WWTPs, resulting in gradually building up

phosphorus and ammonia in the treatment process [114]. This not only decreases the

efficiency of biological treatment, but also leads to gradual struvite precipitation causing

blockages in the pipelines and equipment scaling [115]. Thus, recycling and recovery of

phosphorus from sludge centrate can simultaneously help WWTPs abide by the stringent

regulations of nutrient concentration in effluent, reduce troubleshooting costs caused by

blockages and create an input for the subsequent production of fertilizers. Accordingly, sludge

centrate is likely to be a promising candidate for nutrients recovery.

Although nutrients recovery from sludge centrate obviously brings many benefits,

conventional recovery technologies have been still facing challenges regarding economical

and environmentally-friendly feasibility. Recovery of phosphorus from wastewater is

commonly achieved by precipitation and crystallization processes [109]. These processes are

chemically-intensive and expensive because of the requirement of addition of magnesium and

calcium salts from external sources [109]. Typical products of these processes are in the forms

of calcium phosphate and magnesium ammonium phosphate (struvite) [1, 109]. The

efficiency of precipitation as well as phosphorus content inside precipitate which defines the

value of fertilizer are controlled by the initial concentration of phosphate. Therefore,

phosphorus recovery by precipitation can be improved by firstly enriching phosphate in

sludge centrate [9, 116]. In summary, preconcentrating sludge centrate can simultaneously

increase the precipitation kinetics of nutrients recovery and decrease energy consumption in

WWTPs.

Several advanced separation technologies, namely reverse osmosis and membrane distillation

have been used to pre-concentrate minerals in wastewater for subsequent recovery [117, 118].

Page 75: Forward osmosis for resource recovery from anaerobically

60

However, these membranes are not appropriate to concentrate sludge centrate due to high

susceptibility to fouling and energy intensity. Forward osmosis (FO) has been recently

identified as a robust separation platform capable of treating highly complex solutions [9, 17].

FO is an osmotically-driven membrane separation process and thus, this process can occur

with less or without external energy input [17, 18]. Moreover, the absence of external

hydraulic pressure can account for low fouling tendency and high fouling reversibility in FO

[81, 119]. Therefore, the FO pre-concentration of sludge centrate would be of interest. To

date, the use of FO to preconcentrate nutrients in sludge centrate has been reported in several

studies [1, 9, 14]. Holloway et al. used a hybrid FO-RO system with NaCl solution as DS to

concentrate sludge centrate [14]. In their study, they focused on evaluating flux stability and

water recoverability from sludge centrate rather than showing efficiency of nutrients

enrichment. Ansari et al. demonstrated possibility of using a seawater-driven FO process to

recover only phosphate via direct precipitation of calcium phosphate in sludge centrate

without any chemical addition [1]. Seawater was demonstrated as a promising DS for

preconcentrating sludge centrate. However, phosphate precipitate only made up a small

percentage of total solids recovered [1]. Furthermore, these studies have yet to refer to effects

of operating parameters on the performance of a FO system for nutrients recovery. The results

from these studies, thus, suggest the need to comprehensively and deeply evaluate the use of

seawater-driven FO process for nutrient recovery from sludge centrate.

In this study, we explore challenges of using a seawater-driven FO process for pre-

concentrating sludge centrate for subsequent nutrients recovery. A novel feasible solution to

deal with these challenges is demonstrated herein. In addition, significant effects of stirring

conditions and ratio of membrane area over permeate volume on the performance of

enrichment process are discovered for the first time. These operating conditions are then

optimized, thereby rendering this study highly practical.

4.2. Materials and methods

4.2.1. Forward osmosis system

A TFC PA membrane from Porifera was used. The membrane was soaked into DI water over

night before each experiment. Key properties of this membrane have been reported elsewhere

[113].

A lab-scale cross-flow FO system was used. The details of this cross-flow FO system are

available elsewhere [113]. Two different membrane cells were employed for comparison to

evaluate the role of membrane area for enriching centrate. They have the same channel height

of 2 mm and their effective membrane areas are 50 and 123 cm2, respectively. All

experiments were operated at cross-flow velocity of 12 cm/s and in the counter-current FO

mode. In other words, the membrane AL was facing the FS.

Page 76: Forward osmosis for resource recovery from anaerobically

61

4.2.2. Feed solution and draw solution

Seawater from Fairy Meadow Beach, New South Wales, Australia was used as the DS. The

collected seawater was filtered by 0.45 μm filter paper before use. A portion of this seawater

was buffered at pH 4.4 by dissolving CH3COONa and CH3COOH in seawater to obtain 0.1 M

of each buffer reagent. Sludge centrate from the centrifuge system in the Shellharbour

wastewater treatment plant was used as the FS. The composition of sludge centrate was

presented in Table 13. Composition of seawater and sludge centrate (values indicated average

± standard deviation of at least five samples). HNO3 solution (1 M) was used to dissolve

phosphorus precipitate for mass balance calculation.

Table 13. Composition of seawater and sludge centrate (values indicated average ± standard

deviation of at least five samples).

Parameters Unit Seawater Sludge centrate

pH - 7.35 ± 0.1 7.6 ± 0.2

Electrical conductivity mS/cm 38.6 ± 0.4 6.4 ± 0.2

Osmotic pressure bar 28.0 ± 0.4 6.3 ± 0.1

Total solids g/L 29.5 ± 3.2 1.7 ± 0.2

Ammonia (NH3-N) mg/L - 486.0 ± 10.3

Phosphate (PO43-) mg/L - 284.0 ± 5.6

4.2.3. Experimental protocols

The volume of water transferred through the membrane was continuously monitored using a

digital balance connected to a computer for water flux calculation. To investigate the effects

of stirring up the FS on the performance of enrichment of sludge centrate, a magnetic stirrer at

200 rpm was used to agitate the FS during the experiment.

The initial FS and DS volumes were 1 and 3 L, respective. This high DS to FS volume ratio is

to ensure minimum dilution effect of the DS. As noted in section 2.1, two membrane cells

with effective area of 50 and 123 cm2 were used in this study. Since each experiment was

conducted until 70% water recovery, the corresponding membrane surface area (m2) over

permeate volume (m3) abbreviated as Am/ VP from these two configurations were 70 and 175

m-1, respectively. This Am/VP ratio could be used to evaluate the impact of available

membrane surface area on fouling.

At the beginning of each experiment, DI water was used as the FS to determine its pure water

flux. Then, sludge centrate was substituted to commence the enrichment process until a water

recovery of 70% was achieved, or water flux was decreased to around 90% of initial water

flux. The conductivity and pH of the FS and DS were regularly measured. At specified time

interval, a 5 mL sample was taken from the FS for ammonia and phosphate analyses. At the

end of the filtration, amounts of phosphate precipitated on the membrane surface and in the

pipeline of the system were determined using 500 mL 1 M HNO3 acid solution. The amount

Page 77: Forward osmosis for resource recovery from anaerobically

62

of phosphate penetrating the FO membrane to the DS was also measured. As a result, mass

balance of phosphate in the whole system was quantified.

4.2.4. Analytical methods

pH, electrical conductivity and temperature were measured using an Orion 4 – Star

pH/conductivity meter (Thermo Scientific, Waltham, MA). Ammonia (NH3-N) was analysed

using salicylate US-EPA Standard Method 10205 and Hach DR3900 spectrophotometer.

Orthophosphate (PO43-) was determined using molybdovanadate US-EPA Standard Method

10214 and Hach DR3900 spectrophotometer.

4.3. Results and discussions

4.3.1. Performance of FO for enriching nutrients in sludge centrate

4.3.1.1. Water flux

0 10 20 30 40 50 60 70

0.0

0.2

0.4

0.6

0.8

1.0

1.2

No

rma

lize

d w

ate

r flu

x

Water recovery (%)

175 m-1 & no stirring

175 m-1 & stirring

70 m-1 & no stirring

70 m-1 & stirring

Figure 17. Water flux profile as a function of water recovery at different Am/VP ratios with

and without FS stirring during the pre-concentration of nutrients in sludge centrate by FO.

Seawater was used as the DS.

During FO pre-concentration of sludge centrate, fouling appeared to be driven mostly by the

deposition of inorganic precipitate on the membrane surface (Figure 17). Results in Figure 17

also showed that precipitate deposition occurred both in the bulk solution and on the

membrane surface. Indeed, by agitating the FS, thus preventing chemical precipitation in the

bulk solution and encouraging deposition on the membrane surface, membrane fouling was

significantly more severe (Figure 17). In addition, at the same initial water flux, larger

membrane area reduced the filtration time and the deposition of inorganic precipitate, thus,

resulting in significantly less fouling (Figure 17). At water recovery of 50% without FS

stirring, when the membrane area over permeate volume ratio was 175 m-1, water flux

Page 78: Forward osmosis for resource recovery from anaerobically

63

declined by just 20%. On the other hand, without FS stirring, when the membrane area over

permeate volume ratio was 70 m-1, water flux decreased by 90%. Flux decline with FS stirring

was significantly more severe. These values were 60% and 99% for membrane area over

permeate volume ratios of 175 m-1 and 70 m-1, respectively.

4.3.1.2. Nutrient enrichment

0 10 20 30 40 50 60 70-0.5

0.0

0.5

1.0

1.5

2.0

2.5

3.0

Stirring

Phosphate

Ammonium

Am

mo

nia

an

d p

ho

sp

ha

te

co

nce

ntr

atio

n f

acto

r

pH

Water recovery (%)

Theoretical values of phosphate/ammonia

No stirring

Phosphate

Ammonium

7.6

8.0

8.4

8.8

9.2

9.6

10.0

pH of FS

pH of FS

Figure 18. Effects of stirring up the FS on sludge centrate pH and nutrients enrichment during

the filtration. The theoretical values of phosphate and ammonium concentration factor were

calculated assuming 100% nutrient retention by the FO membrane. The Am/VP ratio in this

experiment was 175 m-1. Seawater was used as the DS.

The precipitation of key nutrient constituents during FO pre-concentration of sludge centrate

was also evidenced when examining their fate in the aqueous phase (Figure 18). In this study,

phosphorous and nitrogen existed in sludge centrate as soluble ortho-phosphate and

ammonium and were retained by the FO membrane by over 99% (Figure 19). If they

remained in the aqueous phase, the concentrations of phosphate and ammonium were

expected to increase as water recovery increased based on a mass balance calculation.

However, experimental results did not show any significant increase in phosphate and

ammonium concentrations in the aqueous phase (Figure 18). In fact, a notable decrease in

phosphate and ammonium concentrations was observed as water recovery increased from

50% to 70%, coinciding with a considerable increase in pH of the sludge centrate FS. The

disappearance of phosphate and ammonium from the pre-concentrated aqueous phase

indicated the formation of struvite (i.e. MgNH4PO4) and calcium phosphate (i.e. Ca3(PO4)2)

precipitates that also caused membrane fouling as discussed in section 4.3.1.1.

Page 79: Forward osmosis for resource recovery from anaerobically

64

Figure 19. Phosphate and ammonium rejection towards different membrane areas and stirring

conditions at water recovery of 70%. Seawater was used as the DS

Results in this study suggested that nutrient precipitation during FO pre-concentration of

sludge centrate was both pH and time dependent. This was evidenced when examining

phosphate concentration in the aqueous phase at different experimental conditions. Without

FS stirring, a small but nevertheless discernible increase in phosphate concentration in sludge

centrate as a function of water recovery up to 50% could be observed (Figure 18). By

contrast, with FS stirring, there was a notable decline in the concentration of phosphate in

sludge centrate. As discussed in section 4.3.1.1, FS stirring caused more membrane fouling,

thus prolonging the experimental time, allowing for more phosphorous precipitation.

Moreover, FS stirring also led to a significant pH increase, thus, promoting precipitation or

the transfer of phosphate from the aqueous phase to the solid phase that caused membrane

fouling. Indeed, there was a strong correlation between pH increase in the FS and FO

filtration time to achieve 70% water recovery (Figure 20).

Page 80: Forward osmosis for resource recovery from anaerobically

65

Figure 20. Correlation between duration of filtration and pH of the FS towards different

experimental conditions at water recovery of 70%. Raw seawater was used as the DS.

It is noteworthy that ammonium could be transferred away from the aqueous phase in the

sludge centrate feed via both co-precipitation with phosphate and volatilization (or air

stripping). The latter occurred when ammonium was converted to ammonia at approximately

pH 9 or higher [120]. The decrease in ammonium concentration without FS stirring was less

severe than that with FS stirring because increase in pH of the FS without FS stirring was less

significant, compared to with FS stirring (Figure 18).

Page 81: Forward osmosis for resource recovery from anaerobically

66

4.3.2. Performance of FO for enriching nutrients in sludge centrate with buffered

seawater

4.3.2.1. Effects of using buffered seawater as the DS on water flux

0 10 20 30 40 50 60 700.0

0.2

0.4

0.6

0.8

1.0

1.2

No

rma

lize

d w

ate

r flu

x

Water recovery (%)

175 m-1 - buffered seawater

175 m-1 - raw seawater

70 m-1 - buffered seawater

Figure 21. Changes in water flux at different ratios of Am/VP without FS stirring when

seawater buffered at pH 4.4 was used as the DS.

Compared to raw seawater as the DS, buffered seawater significantly reduced membrane

fouling (Figure 21). At Am/VP ratio of 175 m-1 and the same water recovery of 70%, water

flux using buffered seawater decreased by only 40% instead of 90% when unbuffered

seawater was used as the DS. Similarly, at Am/VP ratio of 70 m-1, the decrease in water flux

was only 55% when using buffered seawater (Figure 21), compared to 99% when unbuffered

seawater was used as the DS (Figure 17). This phenomenon could be ascribed to the

inhibition of phosphorus precipitate at low pH due to using buffered seawater as the DS as

discussed in section 4.3.2.2. Less phosphorus precipitate led to less membrane fouling,

thereby less decline in water flux during the filtration.

Page 82: Forward osmosis for resource recovery from anaerobically

67

4.3.2.2. Improvement in nutrient enrichment using buffered seawater as the DS

0 10 20 30 40 50 60 70

1.0

1.5

2.0

2.5

3.0A

m/V

P of 70 m

-1

Phosphate

Ammonium

pH

Theoretical values of phosphate/ammonia

Am/V

P of 175 m

-1

Phosphate

AmmoniumP

ho

sp

ha

te a

nd

am

mo

nia

co

nce

ntr

atio

n f

acto

r

Water recovery (%)

4

5

6

7

8

9

10

pH of FS

pH of DS

pH of FS

pH of DS

Figure 22. The performance of FO for enrichment of nutrients in sludge centrate without FS

stirring at different ratios of Am/VP. The theoretical values of phosphate and ammonia

concentration factor were calculated assuming 100% nutrients retention by the FO membrane.

Buffered seawater was used as the DS.

In contrast to the low nutrient enrichment efficiency previously described in section 4.3.1.2,

buffered seawater resulted in a substantial improvement in nutrient enrichment (Figure 22). At

the Am/VP ratio of 175 m-1, the enrichment of both phosphate and ammonium was almost

identical to the theoretical enrichment curve (Figure 22). This observation could be due to

decreased precipitation of phosphate and lower amount of ammonium transferred away from

sludge centrate via volatilization at decreased pH of the FS. Throughout the filtration process,

pH of the FS gradually declined, whereas that of buffered seawater was stable as expected

(Figure 22). This could be explained by a bidirectional transport phenomenon of ion

hydrogens in FO [24]. It is also noted that larger membrane area was more effective in

enriching nutrients as indicated by higher nutrient enrichment efficiency towards using the

ratio of Am/VP of 175 m-1, compared to using that of 70 m-1.

Page 83: Forward osmosis for resource recovery from anaerobically

68

Stirring - raw seawater

No stirring - raw seawater

No stirring - buffered seawater

0

10

20

30

40

50

60

70

80

90

100

Ma

ss b

ala

nce

of

ph

osp

ha

te (

%)

Phosphate in other parts

Phosphate in concentrated centrate

Phosphate on the membrane surface

Figure 23. Differences in mass balance of phosphate between with and without FS stirring, as

well as between using raw seawater and buffered seawater as the DSs. The used ratio of

Am/VP was 175 m-1.

Mass balance calculation of the fate of phosphorus during the preconcentration process

showed that when seawater was employed as the DS and the FS was agitated, phosphate

precipitation occurred mostly on the membrane surface. This was illustrated by the highest

proportion of phosphate (75%) found on the membrane surface at the end of the filtration

(Figure 23). The results also indicated that less soluble phosphate was precipitated on the

membrane surface (60%) without the agitation of the FS (Figure 23). This observation was in

a good agreement with the results shown in section 4.3.1 since phosphorus precipitation led to

membrane fouling and decrease in concentration of nutrients in the bulk FS.

By contrast, soluble phosphate was predominantly retained in concentrated centrate instead of

being precipitated on the membrane surface when buffered seawater was used as the DS.

(Figure 23). Up to 85% of soluble phosphate has been detected in concentrated centrate at the

end of the filtration (Figure 23). These results gave strong support to low fouling propensity

and high phosphate enrichment efficiency as presented in Figure 21 and Figure 22,

respectively when buffered seawater was utilized as the DS. Once concentrated centrate was

expected to be used for subsequent nutrients recovery via outer precipitation, these results

were really promising.

4.4. Conclusions

The results from this study highlighted obstacles in the path of enriching nutrients in sludge

centrate using seawater-driven forward osmosis and a possible way to overcome them. This

study indicated that when using seawater as the DS low efficiency of nutrients enrichment and

significant membrane fouling were closely correlated to the formation and deposition of

Page 84: Forward osmosis for resource recovery from anaerobically

69

precipitates in the bulk FS and on the membrane surface. Complex interactions among pH,

precipitation potential, membrane fouling and duration of filtration have been pointed out

herein. Furthermore, the results showed that no FS stirring led to less significant decrease in

concentration of nutrients and less severe membrane fouling, compared to FS stirring. Also,

the membrane area over permeate volume ratio of 175 m-1 was suitable for pre-concentration

of sludge centrate. It is noteworthy that an innovative approach of using buffered seawater as

the DS led to a remarkable improvement in enriching nutrients and mitigating fouling. At

Am/VP ratio of 175 m-1 using buffered seawater as the DS without FS stirring, the

concentrations of nutrients closely increased to theoretically calculated values, and water flux

only decreased by 40% at water recovery of 70%.

Page 85: Forward osmosis for resource recovery from anaerobically

70

CHAPTER 5

CONCLUSIONS AND RECOMMENDATIONS FOR FUTURE WORK

Page 86: Forward osmosis for resource recovery from anaerobically

71

5.1. Conclusions

The results from this thesis demonstrated that the CTA membrane was superior over the TFC

PA membrane in terms of organic matter enrichment, fouling resistance and reversibility after

cleaning. Prewetting was efficient in significantly increasing the pure water flux of the TFC

PA membrane but causing negligible impacts on that of the CTA membrane. The AL-FS

orientation exhibited higher COD enrichment and lower fouling propensity compared to the

AL-DS orientation. Also, the study indicated that membrane fouling led to the decreased

efficiency of COD enrichment due to the retention and deposition of organic suspended solids

on the membrane surface. In terms of flux and the efficiency of organic enrichment, seawater

and analytical grade NaCl showed the comparable performance. However, it is noted that

seawater as DS resulted in more fouling and lower flux recovery compared to NaCl.

This thesis work showed that during operation of the FO system using seawater as the DS for

enrichment, phosphate and ammonium concentrations in the aqueous phase decreased

significantly. These observations were attributed to formation of phosphorus precipitates (e.g.

struvite and calcium phosphate) and increase in pH of the FS during the filtration. Low

efficiency of nutrients enrichment and severe membrane fouling were closely correlated to

each other via the formation and deposition of precipitates in the bulk FS and on the

membrane surface. In addition, the results indicated that FS stirring resulted in more

significant decrease in concentration of nutrients and more severe fouling than no FS stirring.

Additionally, the membrane surface area over permeate volume ratio of 175 m-1 was right for

pre-concentration of sludge centrate. A major discovery in this study is the application of

buffered seawater as DS to significantly improve nutrients enrichment and reduce membrane

fouling during the filtration.

5.2. Recommendations for Future Work

Resource recovery from waste is an inevitable pathway to ensure the sustainable development

in the future as natural resources have been excessively exploited. The application of FO

membrane to recovery resources from wastewater (e.g. sludge centrate) has been recently

considered as the future’s direction. This study demonstrated the effectiveness of using

seawater-driven FO in enriching organic matter and nutrients in sludge centrate for

subsequent recovery. However, there are still several issues to be addressed prior to the full-

scale deployment of this technology.

Fouling is inevitable in the realm of FO-based wastewater treatment. As mentioned in Chapter

3 and Chapter 4, membrane fouling hindered FO membrane performance and decreased the

efficiency of organic matter and nutrients enrichment. Thus, further research is necessary to

address the issue of membrane fouling when using seawater as the DS for COD and nutrients

enrichment in sludge centrate. A deep understanding of foulants (e.g. colloids, organic

foulants and bio-foulants) in sludge centrate needs to be developed for more effective control

of membrane fouling. In Chapter 4, the chemicals employed to buffer seawater are costly.

Therefore, further investigations need conducting to optimize proper dosages of these

Page 87: Forward osmosis for resource recovery from anaerobically

72

chemicals as well as to find out other possible chemicals for the purpose of buffering

seawater. Furthermore, the results from these chapters suggest further evaluations of the

effectiveness and stability of the seawater-driven FO system for organic matter and nutrients

recovery in long-term operation.

Page 88: Forward osmosis for resource recovery from anaerobically

73

REFERENCES

[1] A. J. Ansari, F. I. Hai, W. E. Price, and L. D. Nghiem, "Phosphorus recovery from

digested sludge centrate using seawater-driven forward osmosis," Separation and

Purification Technology, vol. 163, pp. 1-7, 2016.

[2] W. Mo and Q. Zhang, "Energy–nutrients–water nexus: Integrated resource recovery in

municipal wastewater treatment plants," Journal of Environmental Management, vol.

127, no. Supplement C, pp. 255-267, 2013/09/30/ 2013.

[3] G. Galvagno, C. Eskicioglu, and M. Abel-Denee, "Biodegradation and chemical

precipitation of dissolved nutrients in anaerobically digested sludge dewatering

centrate," Water Research, vol. 96, no. Supplement C, pp. 84-93, 2016/06/01/ 2016.

[4] R. J. Jackson D., English B., Holmes B., Matthews R., Dixon R., Smith P.,

MacDonald N. , "Phosphorus management of beef cattle in Northern Australia," Meat

and Livestock Australia Limited2012.

[5] R. Norton, "4R Canola Nutrition Guide," International Plant Nutrition Institution

2013.

[6] J. Elser and E. Bennett, "Phosphorus cycle: A broken biogeochemical cycle," Nature,

10.1038/478029a vol. 478, no. 7367, pp. 29-31, 10/06/print 2011.

[7] S. T. Broad and R. Corkrey, "Estimating annual generation rates of total P and total N

for different land uses in Tasmania, Australia," Journal of Environmental

Management, vol. 92, no. 6, pp. 1609-1617, 2011/06/01/ 2011.

[8] D. J. Conley et al., "Controlling Eutrophication: Nitrogen and Phosphorus," Science,

10.1126/science.1167755 vol. 323, no. 5917, p. 1014, 2009.

[9] M. Xie, L. D. Nghiem, W. E. Price, and M. Elimelech, "Toward Resource Recovery

from Wastewater: Extraction of Phosphorus from Digested Sludge Using a Hybrid

Forward Osmosis–Membrane Distillation Process," Environmental Science &

Technology Letters, vol. 1, no. 2, pp. 191-195, 2014.

[10] C. Blöcher, C. Niewersch, and T. Melin, "Phosphorus recovery from sewage sludge

with a hybrid process of low pressure wet oxidation and nanofiltration," Water

Research, vol. 46, no. 6, pp. 2009-2019, 2012/04/15/ 2012.

[11] M. Maurer, W. Pronk, and T. A. Larsen, "Treatment processes for source-separated

urine," Water Research, vol. 40, no. 17, pp. 3151-3166, 2006/10/01/ 2006.

[12] M. Xie, H. K. Shon, S. R. Gray, and M. Elimelech, "Membrane-based processes for

wastewater nutrient recovery: Technology, challenges, and future direction," Water

Research, vol. 89, pp. 210-221, 2016/02/01/ 2016.

[13] W. Xue, T. Tobino, F. Nakajima, and K. Yamamoto, "Seawater-driven forward

osmosis for enriching nitrogen and phosphorous in treated municipal wastewater:

effect of membrane properties and feed solution chemistry," Water Res, vol. 69, pp.

120-30, Feb 01 2015.

[14] R. W. Holloway, A. E. Childress, K. E. Dennett, and T. Y. Cath, "Forward osmosis for

concentration of anaerobic digester centrate," Water Res, vol. 41, no. 17, pp. 4005-14,

Sep 2007.

[15] G. Qiu, Y. M. Law, S. Das, and Y. P. Ting, "Direct and complete phosphorus recovery

from municipal wastewater using a hybrid microfiltration-forward osmosis membrane

bioreactor process with seawater brine as draw solution," Environ Sci Technol, vol. 49,

no. 10, pp. 6156-63, May 19 2015.

Page 89: Forward osmosis for resource recovery from anaerobically

74

[16] W. Xue, K. Yamamoto, and T. Tobino, "Membrane fouling and long-term

performance of seawater-driven forward osmosis for enrichment of nutrients in treated

municipal wastewater," Journal of Membrane Science, vol. 499, pp. 555-562, 2016.

[17] K. Lutchmiah, A. R. Verliefde, K. Roest, L. C. Rietveld, and E. R. Cornelissen,

"Forward osmosis for application in wastewater treatment: a review," Water Res, vol.

58, pp. 179-97, Jul 01 2014.

[18] T. Cath, A. Childress, and M. Elimelech, "Forward osmosis: Principles, applications,

and recent developments," Journal of Membrane Science, vol. 281, no. 1-2, pp. 70-87,

2006.

[19] A. Altaee and N. Hilal, "Dual-stage forward osmosis/pressure retarded osmosis

process for hypersaline solutions and fracking wastewater treatment," Desalination,

vol. 350, pp. 79-85, 10/1/ 2014.

[20] A. J. Ansari, F. I. Hai, W. E. Price, J. E. Drewes, and L. D. Nghiem, "Forward osmosis

as a platform for resource recovery from municipal wastewater - A critical assessment

of the literature," Journal of Membrane Science, vol. 529, pp. 195-206, 5/1/ 2017.

[21] W. Luo et al., "Phosphorus and water recovery by a novel osmotic membrane

bioreactor–reverse osmosis system," Bioresource Technology, vol. 200, pp. 297-304,

1// 2016.

[22] B. Kim, S. Lee, and S. Hong, "A novel analysis of reverse draw and feed solute fluxes

in forward osmosis membrane process," Desalination, vol. 352, pp. 128-135, 11/3/

2014.

[23] N. T. Hancock, P. Xu, M. J. Roby, J. D. Gomez, and T. Y. Cath, "Towards direct

potable reuse with forward osmosis: Technical assessment of long-term process

performance at the pilot scale," Journal of Membrane Science, vol. 445, pp. 34-46,

10/15/ 2013.

[24] L. Shu et al., "Why does pH increase with CaCl2 as draw solution during forward

osmosis filtration," Process Safety and Environmental Protection, vol. 104, pp. 465-

471, 2016/11/01/ 2016.

[25] R. W. Baker, Membrane Technology and Applications, Third ed. John Wiley and

Sons, Ltd., 2012.

[26] Q. She, R. Wang, A. G. Fane, and C. Y. Tang, "Membrane fouling in osmotically

driven membrane processes: A review," Journal of Membrane Science, vol. 499, pp.

201-233, 2/1/ 2016.

[27] J. S. Y. William A.Phillip, and Menachem Elimelech, "Reverse Draw Solute

Permeation In Forward Osmosis: Modeling and Experiments," Environ. Sci. Technol.,

vol. 44, pp. 5170-5176, 2010.

[28] D. L. Shaffer, J. R. Werber, H. Jaramillo, S. Lin, and M. Elimelech, "Forward

osmosis: Where are we now?," Desalination, vol. 356, pp. 271-284, 1/15/ 2015.

[29] S. Jiang, Y. Li, and B. P. Ladewig, "A review of reverse osmosis membrane fouling

and control strategies," Sci Total Environ, vol. 595, pp. 567-583, Oct 01 2017.

[30] S. Phuntsho, S. Vigneswaran, J. Kandasamy, S. Hong, S. Lee, and H. K. Shon,

"Influence of temperature and temperature difference in the performance of forward

osmosis desalination process," Journal of Membrane Science, vol. 415–416, pp. 734-

744, 10/1/ 2012.

[31] M. Xie, W. E. Price, L. D. Nghiem, and M. Elimelech, "Effects of feed and draw

solution temperature and transmembrane temperature difference on the rejection of

Page 90: Forward osmosis for resource recovery from anaerobically

75

trace organic contaminants by forward osmosis," Journal of Membrane Science, vol.

438, pp. 57-64, 7/1/ 2013.

[32] A. H. Hawari, N. Kamal, and A. Altaee, "Combined influence of temperature and flow

rate of feeds on the performance of forward osmosis," Desalination, vol. 398, pp. 98-

105, 11/15/ 2016.

[33] M. Xie, W. E. Price, and L. D. Nghiem, "Rejection of pharmaceutically active

compounds by forward osmosis: Role of solution pH and membrane orientation,"

Separation and Purification Technology, vol. 93, pp. 107-114, 6/1/ 2012.

[34] P. Zhao, B. Gao, Q. Yue, P. Liu, and H. K. Shon, "Fatty acid fouling of forward

osmosis membrane: Effects of pH, calcium, membrane orientation, initial permeate

flux and foulant composition," Journal of Environmental Sciences, vol. 46, pp. 55-62,

8// 2016.

[35] M. Sadrzadeh, J. Hajinasiri, S. Bhattacharjee, and D. Pernitsky, "Nanofiltration of oil

sands boiler feed water: Effect of pH on water flux and organic and dissolved solid

rejection," Separation and Purification Technology, vol. 141, pp. 339-353, 2/12/ 2015.

[36] J. Schwinge, P. R. Neal, D. E. Wiley, D. F. Fletcher, and A. G. Fane, "Spiral wound

modules and spacers: Review and analysis," Journal of Membrane Science, vol. 242,

no. 1–2, pp. 129-153, 10/15/ 2004.

[37] Q. Ge, M. Ling, and T.-S. Chung, "Draw solutions for forward osmosis processes:

Developments, challenges, and prospects for the future," Journal of Membrane

Science, vol. 442, pp. 225-237, 9/1/ 2013.

[38] Y. Cai and X. M. Hu, "A critical review on draw solutes development for forward

osmosis," Desalination, vol. 391, pp. 16-29, 8/1/ 2016.

[39] Y. Wang, C. Xie, N. Zhou, and S. Liu, "Study on the membrane orientation in FO

process," Sichuan Daxue Xuebao (Gongcheng Kexue Ban)/Journal of Sichuan

University (Engineering Science Edition), Article vol. 46, no. SUPPL.2, pp. 49-53,

2014.

[40] M. M. Motsa, B. B. Mamba, J. M. Thwala, and A. R. D. Verliefde, "Osmotic

backwash of fouled FO membranes: Cleaning mechanisms and membrane surface

properties after cleaning," Desalination, vol. 402, pp. 62-71, 2017.

[41] Y. Liu, "Fouling in FO Membrane Processes: Characterisation, Mechanisms and

Mitigation," Doctor of Philosophy, Department of Civil and Environmental

Engineering, University of Maryland, University of Maryland, 2013.

[42] C. Y. Tang, T. H. Chong, and A. G. Fane, "Colloidal interactions and fouling of NF

and RO membranes: a review," Adv Colloid Interface Sci, vol. 164, no. 1-2, pp. 126-

43, May 11 2011.

[43] F. Zhao et al., "Using axial vibration membrane process to mitigate membrane fouling

and reject extracellular organic matter in microalgae harvesting," Journal of

Membrane Science, vol. 517, pp. 30-38, 2016.

[44] Y. Gu, Y.-N. Wang, J. Wei, and C. Y. Tang, "Organic fouling of thin-film composite

polyamide and cellulose triacetate forward osmosis membranes by oppositely charged

macromolecules," Water Research, vol. 47, no. 5, pp. 1867-1874, 2013.

[45] Y. Liu and B. Mi, "Combined fouling of forward osmosis membranes: Synergistic

foulant interaction and direct observation of fouling layer formation," Journal of

Membrane Science, vol. 407-408, pp. 136-144, 2012.

Page 91: Forward osmosis for resource recovery from anaerobically

76

[46] M. Zhang, H. Lin, L. Shen, B.-Q. Liao, X. Wu, and R. Li, "Effect of calcium ions on

fouling properties of alginate solution and its mechanisms," Journal of Membrane

Science, vol. 525, pp. 320-329, 2017.

[47] P. Zhao, B. Gao, Q. Yue, P. Liu, and H. K. Shon, "Fatty acid fouling of forward

osmosis membrane: Effects of pH, calcium, membrane orientation, initial permeate

flux and foulant composition," J Environ Sci (China), vol. 46, pp. 55-62, Aug 2016.

[48] P. Xiao, J. Li, Y. Ren, and X. Wang, "A comprehensive study of factors affecting

fouling behavior in forward osmosis," Colloids and Surfaces A: Physicochemical and

Engineering Aspects, vol. 499, pp. 163-172, 2016.

[49] M. Xie, L. D. Nghiem, W. E. Price, and M. Elimelech, "Impact of organic and

colloidal fouling on trace organic contaminant rejection by forward osmosis: Role of

initial permeate flux," Desalination, vol. 336, pp. 146-152, 2014.

[50] Y. Kim, S. Lee, H. K. Shon, and S. Hong, "Organic fouling mechanisms in forward

osmosis membrane process under elevated feed and draw solution temperatures,"

Desalination, vol. 355, pp. 169-177, 2015.

[51] J. R. McCutcheon, R. L. McGinnis, and M. Elimelech, "Desalination by ammonia–

carbon dioxide forward osmosis: Influence of draw and feed solution concentrations

on process performance," Journal of Membrane Science, vol. 278, no. 1–2, pp. 114-

123, 7/5/ 2006.

[52] M. Elimelech, Z. Xiaohua, A. E. Childress, and H. Seungkwan, "Role of membrane

surface morphology in colloidal fouling of cellulose acetate and composite aromatic

polyamide reverse osmosis membranes," Journal of Membrane Science, vol. 127, no.

1, pp. 101-109, 1997/04/30/ 1997.

[53] S. Qi, C. Q. Qiu, Y. Zhao, and C. Y. Tang, "Double-skinned forward osmosis

membranes based on layer-by-layer assembly—FO performance and fouling

behavior," Journal of Membrane Science, vol. 405-406, pp. 20-29, 2012.

[54] J. R. McCutcheon, R. L. McGinnis, and M. Elimelech, "A novel ammonia—carbon

dioxide forward (direct) osmosis desalination process," Desalination, vol. 174, no. 1,

pp. 1-11, 2005/04/01/ 2005.

[55] M. Xie, E. Bar-Zeev, S. M. Hashmi, L. D. Nghiem, and M. Elimelech, "Role of

Reverse Divalent Cation Diffusion in Forward Osmosis Biofouling," Environmental

Science & Technology, vol. 49, no. 22, pp. 13222-13229, 2015/11/17 2015.

[56] C. Y. Tang, Q. She, W. C. L. Lay, R. Wang, and A. G. Fane, "Coupled effects of

internal concentration polarization and fouling on flux behavior of forward osmosis

membranes during humic acid filtration," Journal of Membrane Science, vol. 354, no.

1-2, pp. 123-133, 2010.

[57] J. Zhang, W. L. C. Loong, S. Chou, C. Tang, R. Wang, and A. G. Fane, "Membrane

biofouling and scaling in forward osmosis membrane bioreactor," Journal of

Membrane Science, vol. 403-404, pp. 8-14, 2012.

[58] S. Zou et al., "Direct microscopic observation of forward osmosis membrane fouling

by microalgae: Critical flux and the role of operational conditions," Journal of

Membrane Science, vol. 436, pp. 174-185, 2013.

[59] G. T. Gray, J. R. McCutcheon, and M. Elimelech, "Internal concentration polarization

in forward osmosis: role of membrane orientation," Desalination, vol. 197, no. 1-3,

pp. 1-8, 2006.

Page 92: Forward osmosis for resource recovery from anaerobically

77

[60] M. Xie, L. D. Nghiem, W. E. Price, and M. Elimelech, "Impact of humic acid fouling

on membrane performance and transport of pharmaceutically active compounds in

forward osmosis," Water Res, vol. 47, no. 13, pp. 4567-75, Sep 01 2013.

[61] A. P. Murphy, C. D. Moody, R. L. Riley, S. W. Lin, B. Murugaverl, and P. Rusin,

"Microbiological damage of cellulose acetate RO membranes," Journal of Membrane

Science, vol. 193, no. 1, pp. 111-121, 2001/10/31/ 2001.

[62] T. Nguyen, A. F. Roddick, and L. Fan, "Biofouling of Water Treatment Membranes:

A Review of the Underlying Causes, Monitoring Techniques and Control Measures,"

Membranes, vol. 2, no. 4, 2012.

[63] X. Jin, Q. She, X. Ang, and C. Y. Tang, "Removal of boron and arsenic by forward

osmosis membrane: Influence of membrane orientation and organic fouling," Journal

of Membrane Science, vol. 389, pp. 182-187, 2012.

[64] S. Lee and M. Elimelech, "Relating Organic Fouling of Reverse Osmosis Membranes

to Intermolecular Adhesion Forces," Environmental Science & Technology, vol. 40,

no. 3, pp. 980-987, 2006/02/01 2006.

[65] B. Mi and M. Elimelech, "Chemical and physical aspects of organic fouling of

forward osmosis membranes," Journal of Membrane Science, vol. 320, no. 1, pp. 292-

302, 2008/07/15/ 2008.

[66] D. Van der Kooij, Hijnen W.A.M., Comelissen E.R., "Elucidation of Membrane

Biofouling Processes Using Bioassays for Assessing the Microbial Growth Potential

of Feedwater," presented at the Membrane Technology Conference 2007.

[67] Y. Chun, D. Mulcahy, L. Zou, and I. S. Kim, "A Short Review of Membrane Fouling

in Forward Osmosis Processes," Membranes (Basel), vol. 7, no. 2, Jun 12 2017.

[68] X. Li, Y. Mo, J. Li, W. Guo, and H. H. Ngo, "In-situ monitoring techniques for

membrane fouling and local filtration characteristics in hollow fiber membrane

processes: A critical review," Journal of Membrane Science, vol. 528, pp. 187-200,

2017.

[69] L. N. Sim and A. G. Fane, "Advanced Monitoring of Membrane Fouling and Control

Strategies," in Reference Module in Chemistry, Molecular Sciences and Chemical

Engineering: Elsevier, 2017.

[70] E. W. Tow, M. M. Rencken, and J. H. Lienhard, "In situ visualization of organic

fouling and cleaning mechanisms in reverse osmosis and forward osmosis,"

Desalination, vol. 399, pp. 138-147, 2016.

[71] S. Jamaly, N. N. Darwish, I. Ahmed, and S. W. Hasan, "A short review on reverse

osmosis pretreatment technologies," Desalination, vol. 354, pp. 30-38, 2014.

[72] A. Sweity, Y. Oren, Z. Ronen, and M. Herzberg, "The influence of antiscalants on

biofouling of RO membranes in seawater desalination," Water Research, vol. 47, no.

10, pp. 3389-3398, 2013/06/15/ 2013.

[73] Q. She, Y. K. W. Wong, S. Zhao, and C. Y. Tang, "Organic fouling in pressure

retarded osmosis: Experiments, mechanisms and implications," Journal of Membrane

Science, vol. 428, pp. 181-189, 2013/02/01/ 2013.

[74] X. Lu, S. Romero-Vargas Castrillon, D. L. Shaffer, J. Ma, and M. Elimelech, "In situ

surface chemical modification of thin-film composite forward osmosis membranes for

enhanced organic fouling resistance," Environ Sci Technol, vol. 47, no. 21, pp. 12219-

28, 2013.

Page 93: Forward osmosis for resource recovery from anaerobically

78

[75] A. Nguyen, S. Azari, and L. Zou, "Coating zwitterionic amino acid l-DOPA to

increase fouling resistance of forward osmosis membrane," Desalination, vol. 312, pp.

82-87, 2013.

[76] A. Nguyen, L. Zou, and C. Priest, "Evaluating the antifouling effects of silver

nanoparticles regenerated by TiO2 on forward osmosis membrane," Journal of

Membrane Science, vol. 454, pp. 264-271, 2014.

[77] S. Romero-Vargas Castrillón, X. Lu, D. L. Shaffer, and M. Elimelech, "Amine

enrichment and poly(ethylene glycol) (PEG) surface modification of thin-film

composite forward osmosis membranes for organic fouling control," Journal of

Membrane Science, vol. 450, pp. 331-339, 2014.

[78] A. Tiraferri, Y. Kang, E. P. Giannelis, and M. Elimelech, "Superhydrophilic thin-film

composite forward osmosis membranes for organic fouling control: fouling behavior

and antifouling mechanisms," Environ Sci Technol, vol. 46, no. 20, pp. 11135-44, Oct

16 2012.

[79] H.-Y. Yu, Y. Kang, Y. Liu, and B. Mi, "Grafting polyzwitterions onto polyamide by

click chemistry and nucleophilic substitution on nitrogen: A novel approach to

enhance membrane fouling resistance," Journal of Membrane Science, vol. 449, pp.

50-57, 2014.

[80] B. G.-F. Jose Miguel Arnal, Maria Sancho, "Membrane cleaning, Expanding issues in

desalination ", I. corporation, Ed., ed, 2011.

[81] B. Mi and M. Elimelech, "Organic fouling of forward osmosis membranes: Fouling

reversibility and cleaning without chemical reagents," Journal of Membrane Science,

vol. 348, no. 1-2, pp. 337-345, 2010.

[82] G. Blandin, H. Vervoort, P. Le-Clech, and A. R. D. Verliefde, "Fouling and cleaning

of high permeability forward osmosis membranes," Journal of Water Process

Engineering, vol. 9, pp. 161-169, 2016.

[83] E. Arkhangelsky, F. Wicaksana, S. Chou, A. A. Al-Rabiah, S. M. Al-Zahrani, and R.

Wang, "Effects of scaling and cleaning on the performance of forward osmosis hollow

fiber membranes," Journal of Membrane Science, vol. 415-416, pp. 101-108, 2012.

[84] T. Ergön-Can, B. Köse-Mutlu, İ. Koyuncu, and C.-H. Lee, "Biofouling control based

on bacterial quorum quenching with a new application: Rotary microbial carrier

frame," Journal of Membrane Science, vol. 525, pp. 116-124, 2017.

[85] H. Petrus, H. Li, V. Chen, and N. Norazman, "Enzymatic cleaning of ultrafiltration

membranes fouled by protein mixture solutions," Journal of Membrane Science, vol.

325, no. 2, pp. 783-792, 2008.

[86] M. F. Siddiqui, M. Rzechowicz, W. Harvey, A. W. Zularisam, and G. F. Anthony,

"Quorum sensing based membrane biofouling control for water treatment: A review,"

Journal of Water Process Engineering, vol. 7, pp. 112-122, 2015.

[87] H. Xu and Y. Liu, "Control and Cleaning of Membrane Biofouling by Energy

Uncoupling and Cellular Communication," Environmental Science & Technology, vol.

45, no. 2, pp. 595-601, 2011/01/15 2011.

[88] C. H. Yu, L. C. Fang, S. K. Lateef, C. H. Wu, and C. F. Lin, "Enzymatic treatment for

controlling irreversible membrane fouling in cross-flow humic acid-fed

ultrafiltration," J Hazard Mater, vol. 177, no. 1-3, pp. 1153-8, May 15 2010.

Page 94: Forward osmosis for resource recovery from anaerobically

79

[89] R. C. M. a. C. P. B. Lucinda F. Silva, "Modification of polyamide reverse osmosis

membranes seeking for better resistance to oxidizing agents," Membrane Water

Treatment, vol. 3, pp. 169-179, 2012.

[90] A. A. Alturki, J. A. McDonald, S. J. Khan, W. E. Price, L. D. Nghiem, and M.

Elimelech, "Removal of trace organic contaminants by the forward osmosis process,"

Separation and Purification Technology, vol. 103, pp. 258-266, 2013/01/15/ 2013.

[91] A. J. Ansari, F. I. Hai, W. Guo, H. H. Ngo, W. E. Price, and L. D. Nghiem, "Factors

governing the pre-concentration of wastewater using forward osmosis for subsequent

resource recovery," Sci Total Environ, vol. 566-567, pp. 559-66, Oct 01 2016.

[92] G. Chen et al., "Treatment of shale gas drilling flowback fluids (SGDFs) by forward

osmosis: Membrane fouling and mitigation," Desalination, vol. 366, pp. 113-120,

2015/06/15/ 2015.

[93] Y. Chun, S.-J. Kim, G. J. Millar, D. Mulcahy, I. S. Kim, and L. Zou, "Forward

osmosis as a pre-treatment for treating coal seam gas associated water: Flux and

fouling behaviour," Desalination, vol. 403, pp. 144-152, 2017/02/01/ 2017.

[94] B. D. Coday et al., "The sweet spot of forward osmosis: Treatment of produced water,

drilling wastewater, and other complex and difficult liquid streams," Desalination, vol.

333, no. 1, pp. 23-35, 2014/01/15/ 2014.

[95] X.-M. Li et al., "Water reclamation from shale gas drilling flow-back fluid using a

novel forward osmosis–vacuum membrane distillation hybrid system," Water Science

and Technology, 10.2166/wst.2014.003 vol. 69, no. 5, p. 1036, 2014.

[96] S. Liyanaarachchi, V. Jegatheesan, I. Obagbemi, S. Muthukumaran, and L. Shu,

"Effect of feed temperature and membrane orientation on pre-treatment sludge volume

reduction through forward osmosis," Desalination and Water Treatment, vol. 54, no.

4-5, pp. 838-844, 2015/05/01 2015.

[97] N. C. Nguyen, H. T. Nguyen, S.-S. Chen, N. T. Nguyen, and C.-W. Li, "Application

of forward osmosis (FO) under ultrasonication on sludge thickening of waste activated

sludge," Water Science and Technology, 10.2166/wst.2015.341 vol. 72, no. 8, p. 1301,

2015.

[98] Z. Wang, J. Zheng, J. Tang, X. Wang, and Z. Wu, "A pilot-scale forward osmosis

membrane system for concentrating low-strength municipal wastewater: performance

and implications," Scientific Reports, Article vol. 6, p. 21653, 02/22/online 2016.

[99] W. Luo, M. Xie, F. I. Hai, W. E. Price, and L. D. Nghiem, "Biodegradation of

cellulose triacetate and polyamide forward osmosis membranes in an activated sludge

bioreactor: Observations and implications," Journal of Membrane Science, vol. 510,

no. Supplement C, pp. 284-292, 2016/07/15/ 2016.

[100] N. M. Mazlan et al., "Organic fouling behaviour of structurally and chemically

different forward osmosis membranes – A study of cellulose triacetate and thin film

composite membranes," Journal of Membrane Science, vol. 520, pp. 247-261, 2016.

[101] S. Zhao, L. Zou, and D. Mulcahy, "Effects of membrane orientation on process

performance in forward osmosis applications," Journal of Membrane Science, vol.

382, no. 1, pp. 308-315, 2011/10/15/ 2011.

[102] J. Kochan, T. Wintgens, R. Hochstrat, and T. Melin, "Impact of wetting agents on the

filtration performance of polymeric ultrafiltration membranes," Desalination, vol. 241,

no. 1, pp. 34-42, 2009/05/31/ 2009.

Page 95: Forward osmosis for resource recovery from anaerobically

80

[103] J. R. McCutcheon and M. Elimelech, "Influence of membrane support layer

hydrophobicity on water flux in osmotically driven membrane processes," Journal of

Membrane Science, vol. 318, no. 1, pp. 458-466, 2008/06/20/ 2008.

[104] T. Y. Cath et al., "Standard Methodology for Evaluating Membrane Performance in

Osmotically Driven Membrane Processes," Desalination, vol. 312, no. Supplement C,

pp. 31-38, 2013/03/01/ 2013.

[105] E. M. V. Hoek, S. Bhattacharjee, and M. Elimelech, "Effect of Membrane Surface

Roughness on Colloid−Membrane DLVO Interactions," Langmuir, vol. 19, no. 11, pp.

4836-4847, 2003/05/01 2003.

[106] A. Poisson and A. Papaud, "Diffusion coefficients of major ions in seawater," Marine

Chemistry, vol. 13, no. 4, pp. 265-280, 1983/10/01/ 1983.

[107] L. Yuan-Hui and S. Gregory, "Diffusion of ions in sea water and in deep-sea

sediments," Geochimica et Cosmochimica Acta, vol. 38, no. 5, pp. 703-714,

1974/05/01/ 1974.

[108] M. Xie and S. R. Gray, "Gypsum scaling in forward osmosis: Role of membrane

surface chemistry," Journal of Membrane Science, vol. 513, no. Supplement C, pp.

250-259, 2016/09/01/ 2016.

[109] S. Daneshgar, A. Callegari, A. Capodaglio, and D. Vaccari, "The Potential Phosphorus

Crisis: Resource Conservation and Possible Escape Technologies: A Review,"

Resources, vol. 7, no. 2, 2018.

[110] D. C. J.J. Schroder, A.L. Smit and A. Rosemarin, "Sustainable Use of Phosphorus ",

Netherland2010.

[111] E. Desmidt et al., "Global Phosphorus Scarcity and Full-Scale P-Recovery

Techniques: A Review," Critical Reviews in Environmental Science and Technology,

vol. 45, no. 4, pp. 336-384, 2015/02/16 2015.

[112] T. Schütte, C. Niewersch, T. Wintgens, and S. Yüce, "Phosphorus recovery from

sewage sludge by nanofiltration in diafiltration mode," Journal of Membrane Science,

vol. 480, pp. 74-82, 2015/04/15/ 2015.

[113] M. T. Vu, A. J. Ansari, F. I. Hai, and L. D. Nghiem, "Performance of a seawater-

driven forward osmosis process for pre-concentrating digested sludge centrate: organic

enrichment and membrane fouling," Environmental Science: Water Research &

Technology, 10.1039/C8EW00132D vol. 4, no. 7, pp. 1047-1056, 2018.

[114] L. E. de-Bashan and Y. Bashan, "Recent advances in removing phosphorus from

wastewater and its future use as fertilizer (1997–2003)," Water Research, vol. 38, no.

19, pp. 4222-4246, 2004/11/01/ 2004.

[115] K. S. Le Corre, E. Valsami-Jones, P. Hobbs, and S. A. Parsons, "Phosphorus Recovery

from Wastewater by Struvite Crystallization: A Review," Critical Reviews in

Environmental Science and Technology, vol. 39, no. 6, pp. 433-477, 2009/06/01 2009.

[116] A. T. K. Tran et al., "P-recovery as calcium phosphate from wastewater using an

integrated selectrodialysis/crystallization process," Journal of Cleaner Production,

vol. 77, pp. 140-151, 2014/08/15/ 2014.

[117] A. Ali, A. C. Quist-Jensen, F. Macedonio, and E. Drioli, "Application of Membrane

Crystallization for Minerals’ Recovery from Produced Water," Membranes, vol. 5, no.

4, 2015.

[118] C. Quist-Jensen, A. Ali, S. Mondal, F. Macedonio, and E. Drioli, Membrane

crystallization for minerals recovery from brine. 2015.

Page 96: Forward osmosis for resource recovery from anaerobically

81

[119] B. Mi and M. Elimelech, "Silica scaling and scaling reversibility in forward osmosis,"

Desalination, vol. 312, no. Supplement C, pp. 75-81, 2013/03/01/ 2013.

[120] C. Eskicioglu, G. Galvagno, and C. Cimon, "Approaches and processes for ammonia

removal from side-streams of municipal effluent treatment plants," Bioresource

Technology, vol. 268, pp. 797-810, 2018/11/01/ 2018.