150
ENGINEERED BIOFILTRATION FOR ULTRAFILTRATION FOULING CONTROL AND DBP PRECURSOR REMOVAL by Jamal Azzeh A thesis submitted in conformity with the requirements for the degree of Master of Applied Science Graduate Department of Civil Engineering University of Toronto © Copyright by Jamal Azzeh 2014

ENGINEERED BIOFILTRATION FOR … ENGINEERED BIOFILTRATION FOR ULTRAFILTRATION FOULING CONTROL AND DBP PRECURSOR REMOVAL Jamal Azzeh Master of Applied Science, 2014 Graduate Department

  • Upload
    buihanh

  • View
    223

  • Download
    2

Embed Size (px)

Citation preview

ENGINEERED BIOFILTRATION FOR

ULTRAFILTRATION FOULING CONTROL AND DBP

PRECURSOR REMOVAL

by

Jamal Azzeh

A thesis submitted in conformity with the requirements

for the degree of Master of Applied Science

Graduate Department of Civil Engineering

University of Toronto

© Copyright by Jamal Azzeh 2014

ii

ENGINEERED BIOFILTRATION FOR ULTRAFILTRATION

FOULING CONTROL AND DBP PRECURSOR REMOVAL

Jamal Azzeh

Master of Applied Science, 2014

Graduate Department of Civil engineering

University of Toronto

ABSTRACT

Recently, treatment plants have adopted biofiltration to compliment conventional treatment

and ozonation. Previous literature has focused on passive applications of biofiltration. In this study,

several pilot-scale biofiltration trains were designed in parallel to conventional treatment to

investigate the impact of nutrient addition (nitrogen and phosphorus), use of hydrogen peroxide,

aluminum sulphate (alum), and different filtration media (anthracite vs. granular activated carbon

(GAC)) on biofiltration performance. Parameters measured included organic removal, reduction

of DBP precursor, improvements in filter runtimes and ultrafiltration (UF) fouling control. Nutrient

addition did not improve biofiltration performance. Supplementing hydrogen peroxide (<1 mg/L)

decreased headloss, DBP formation potentials while adversely affecting UF fouling. In-line alum

addition (<0.5 mg/L) improved biofilter’s ability to control fouling and DBP precursor without

adversely impacting headloss. GAC provided superior performance when compared to anthracite.

Conventional treatment provided higher DOC, and DBP precursor removal, as well as better UF

fouling control compared to biofiltration.

iii

ACKNOWLEDGMENTS

This work was funded by the Natural Sciences and Engineering Research Council of

Canada (NSERC) Chair in Drinking Water Research at the University of Toronto, and by the

Ontario Research Fund (ORF).

I would like to thank my supervisor, Professor R.C. Andrews, for his continuous and

invaluable guidance, support, and encouragement. Also, I would like to extend my thanks to Prof.

R. Hofmann and Prof. S. Andrews in particular and my other course instructors in general. I

sincerely appreciate the assistance of John Armour, Liz Taylor-Edmonds, Sara Sadreddini, Jim

Wang, Isabelle Netto, Mike McKie, Jennifer Lee, Pulin Mondal, Sabrina Diemert, Joshua Elliott

and other members of the Drinking Research Group. My thanks also goes to Giovanni Buzzeo for

the effort in constructing the experimental set-up.

My gratitude to the management and staff at the Peterborough utilities company for their

support, accommodation, and hospitality especially Rene Gagnon, Keven Light, and Graham Bill.

This research would not have been possible without their help.

Finally, to my dear parents, family and friends, thank you for your encouragement, advice

and love which inspires me at every stage of my life.

iv

TABLE OF CONTENTS

ABSTRACT .................................................................................................................................... ii

ACKNOWLEDGMENTS ............................................................................................................. iii

TABLE OF CONTENTS ............................................................................................................... iv

LIST OF TABLES ....................................................................................................................... viii

LIST OF FIGURES ........................................................................................................................ x

NOMENCLATURE .................................................................................................................... xiii

1. Introduction ................................................................................................................................ 1

1.1 Background ......................................................................................................................... 1

1.2 Research Objectives ............................................................................................................ 2

1.3 Description of Chapters ...................................................................................................... 3

2. Literature Review ....................................................................................................................... 4

2.1 Natural Organic Matter (NOM) .......................................................................................... 4

2.2 Biofiltration ......................................................................................................................... 5

2.2.1 Factors Affecting Biofilter Performance ................................................................ 5

2.2.1.1 Contact Time ............................................................................................ 6

2.2.1.2 Water Matrix............................................................................................. 7

2.2.1.3 Filter Media .............................................................................................. 7

2.2.1.4 Temperature .............................................................................................. 8

2.2.1.5 Backwashing ............................................................................................. 9

2.2.1.6 Biological Activity ................................................................................. 10

2.2.1.7 Ozone Pre-treatment and Hydrogen Peroxide Addition ......................... 11

2.2.2 NOM Monitoring Techniques for Biofiltration Studies ....................................... 12

2.2.3 Biomass Quantification and Qualification Methods ............................................. 16

2.3 Biofiltration for UF fouling control .................................................................................. 17

2.4 Coagulation as a Pre-treatment for UF ............................................................................. 18

2.5 Research Gaps and Needs ................................................................................................. 19

3. Materials and Methods ............................................................................................................. 21

3.1 Research Objective ........................................................................................................... 21

3.2 Experimental Configuration .............................................................................................. 21

3.2.1 Biofilter Size and Media Depth ............................................................................ 21

3.2.2 Pilot-scale Plant .................................................................................................... 22

v

3.2.3 Biofiltration Treatment Train ................................................................................ 24

3.2.4 Sampling Locations .............................................................................................. 27

3.2.5 Ultrafiltration Unit ................................................................................................ 28

3.3 Experimental Design ......................................................................................................... 29

3.3.1 Variables ............................................................................................................... 29

3.3.2 Monitored Parameters ........................................................................................... 30

3.3.3 Sampling Time ...................................................................................................... 30

3.3.4 Media Type ........................................................................................................... 31

3.3.5 Chemical Addition ................................................................................................ 31

3.3.5.1 Nutrient Enhancement ............................................................................ 32

3.3.5.2 Hydrogen Peroxide Addition .................................................................. 32

3.3.5.3 Alum Addition ........................................................................................ 32

3.4 Analytical Methods ........................................................................................................... 34

3.4.1 Dissolved Organic Carbon .................................................................................... 34

3.4.2 UV254 ..................................................................................................................... 35

3.4.3 THMs .................................................................................................................... 36

3.4.4 HAAs .................................................................................................................... 39

3.4.5 AOX ...................................................................................................................... 42

3.4.6 ATP Analysis ........................................................................................................ 44

3.4.7 EPS Analysis ......................................................................................................... 45

3.4.8 Nitrogen and Phosphorus ...................................................................................... 49

3.4.9 Chlorine (Cl2) Demands and Residuals ................................................................ 49

3.5 Quality Assurance/ Quality Control (QA/QC) ................................................................. 50

3.6 Statistical Analysis ............................................................................................................ 51

4. Engineered Biofiltration for NOM, DBP Precursor Removal, and UF Fouling Control ......... 53

4.1 Introduction ....................................................................................................................... 53

4.2 Materials and Methods ...................................................................................................... 54

4.2.1 Source Water Quality Parameters ......................................................................... 54

4.2.2 Pilot Configuration ................................................................................................ 55

4.2.3 Ultrafiltration Units ............................................................................................... 55

4.2.4 Experimental Design ............................................................................................. 57

4.2.5 Analytical Methods ............................................................................................... 58

vi

4.2.6 Statistical Analysis ................................................................................................ 59

4.2.7 UF Fouling Quantification .................................................................................... 60

4.3 Results and Discussion ..................................................................................................... 60

4.3.1 Biofilter Acclimation and Activity ....................................................................... 60

4.3.2 Impact of Engineered Biofiltration on Turbidity Removal ................................... 61

4.3.3 Impact of Engineered Biofiltration on NOM Removal ........................................ 62

4.3.4 Impact of Enhancement Strategies on Headloss (LOH) ....................................... 65

4.3.5 DBP Precursor Reduction ..................................................................................... 67

4.3.6 Engineered Biofiltration for Ultrafiltration Fouling Mitigation ........................... 70

4.4 Summary ........................................................................................................................... 74

5. Comparison between Conventional Treatment and Biofiltration for DBP Precursor

Removal and UF Fouling Control ............................................................................................ 76

5.1 Introduction ....................................................................................................................... 76

5.2 Materials and Methods ...................................................................................................... 77

5.2.1 Source Water Quality ............................................................................................ 77

5.2.2 Pilot- and Full-scale Plant Configuration ............................................................. 77

5.2.3 Ultrafiltration Unit and Fouling Quantification .................................................... 78

5.2.4 Analytical Methods ............................................................................................... 78

5.2.5 Statistical Analysis ................................................................................................ 80

5.3 Results and Discussion ..................................................................................................... 80

5.3.1 Biofilter Acclimation and Activity ....................................................................... 80

5.3.2 Turbidity Removal ................................................................................................ 81

5.3.3 NOM Removal ...................................................................................................... 81

5.3.4 Headloss (LOH) .................................................................................................... 86

5.3.5 DBP Precursor Removal ....................................................................................... 88

5.3.6 Ultrafiltration Fouling Mitigation ......................................................................... 90

5.4 Summary ........................................................................................................................... 91

6. Overall Summary, Conclusions and Recommendations .......................................................... 93

6.1 Summary ........................................................................................................................... 93

6.2 Conclusions ....................................................................................................................... 93

6.3 Recommendations ............................................................................................................. 94

7. References ................................................................................................................................ 95

vii

8. Appendices ............................................................................................................................. 109

8.1 Appendix A (Calibration Curves) ................................................................................... 109

8.2 Appendix B (Raw Data) .................................................................................................. 112

8.3 Appendix C (QA/QC) ..................................................................................................... 127

8.3.1 THMs .................................................................................................................. 127

8.3.2 HAAs .................................................................................................................. 129

8.4 Appendix D (Additional Information) ............................................................................ 130

viii

LIST OF TABLES

Table 2.1: The degree of control and impact of factors influencing biofiltration .......................... 6

Table 2.2: LC-OCD fractions ....................................................................................................... 15

Table 3.1: Summary of the characteristics of biofilters in previous studies ................................ 22

Table 3.2: Biofilter conditions ..................................................................................................... 27

Table 3.3: Experimental design ................................................................................................... 33

Table 3.4: DOC analyzer conditions ............................................................................................ 34

Table 3.5: DOC analysis reagents ................................................................................................ 35

Table 3.6: DOC method outline ................................................................................................... 35

Table 3.7: THM instrument conditions ........................................................................................ 36

Table 3.8: THM reagents ............................................................................................................. 37

Table 3.9: THM method outline .................................................................................................. 37

Table 3.10: THM method detection limits ................................................................................... 39

Table 3.11: HAA instrument conditions ...................................................................................... 39

Table 3.12: HAA reagents ........................................................................................................... 40

Table 3.13: HAA method outline ................................................................................................. 40

Table 3.14: HAA method detection limits ................................................................................... 42

Table 3.15: AOX instrument conditions ...................................................................................... 43

Table 3.16: AOX reagents ........................................................................................................... 43

Table 3.17: AOX method outline ................................................................................................. 43

Table 3.18: ATP method outline .................................................................................................. 44

Table 3.19: EPS reagents ............................................................................................................. 46

Table 3.20: EPS method outline .................................................................................................. 46

Table 4.1: Otonabee River raw water quality parameters ............................................................ 54

Table 4.2: Expermintal design for biofiltration pilot ................................................................... 57

Table 4.3: Biopolymer and humic substance removals (%) by biofiltration. .............................. 64

Table 4.4: Average hourly headloss for biofilters during the period of chemical addition ......... 66

Table 4.5: Impact of filter media and biofiltration conditions on performance parameters ........ 75

Table 4.6: Impact of biofiltration, filter media, and chemical addition on reversible and

irreversible UF fouling .................................................................................................................. 75

Table 5.1: ATP concentrations on biofilter and conventional filters media ................................ 81

ix

Table 5.2: DOC removal pre/post pilot conventional filters with alum and PACl, and after

biofiltration ................................................................................................................................... 84

Table 5.3: Impact of biofiltration and conventional treatment on the removal of the variables in

this study ....................................................................................................................................... 92

Table 8.1: Water quality and DOC results (Raw) ...................................................................... 112

Table 8.2: THM and AOX results (Raw) ................................................................................... 116

Table 8.3: HAA results (Raw) ................................................................................................... 120

Table 8.4: Raw ATP and EPS data ............................................................................................ 125

Table 8.5: LC-OCD results (Raw) ............................................................................................. 126

Table 8.6: Sample THMs control chart warning and control limits .......................................... 127

Table 8.7: HAA control chart warning and control limits ......................................................... 129

Table 8.8: DOC concentrations in raw water and biofilter effluent ........................................... 133

x

LIST OF FIGURES

Figure 2.1: Typical LC-OCD diagram with OCD, OND and UV detectors ............................... 14

Figure 3.1: Schematic of the pilot-plant treatment processes showing the conventional treatment

trains and the biofiltration treatment trains. .................................................................................. 23

Figure 3.2: Biofiltration schematic .............................................................................................. 25

Figure 3.3: Photo of the biofiltration set-up at the PWTP showing the large biofilters along with

the conventional filters (right) and the small biofilters (left) ........................................................ 26

Figure 3.4: Sketch showing the water sampling locations at the Peterborough pilot treatment

plant ............................................................................................................................................... 28

Figure 3.5: UF unit schematic (a) and photo (b) .......................................................................... 29

Figure 3.6: DOC concentration and UV254 absorbance in the biofilter effluent after a backwash

....................................................................................................................................................... 31

Figure 3.7: Chemical addition Gantt chart ................................................................................... 33

Figure 4.1: Biofiltration pilot plant schematic ............................................................................. 56

Figure 4.2: ATP concentrations of biofilter media during the experimental period .................... 61

Figure 4.3: DOC removal (%) by biofiltration ............................................................................ 63

Figure 4.4: Protein and polysaccharide concentrations on media ............................................... 66

Figure 4.5: Relationship between ATP, proteins and polysaccharides ........................................ 67

Figure 4.6: DBP formation potentional. Single samples analyzed in duplicate, vertical bars

represent one standard deviation. .................................................................................................. 68

Figure 4.7: DBP removal as a function of biopolymer removal .................................................. 69

Figure 4.8: Normalized UF resistance profile comparing raw water vs large control biofilter

effluent .......................................................................................................................................... 70

Figure 4.9: Normalized UF resistance profile comparing nutrient enahnced vs large control

biofilter effluent ............................................................................................................................ 71

Figure 4.10: Normalized UF resistance profiles comparing 0.1 mg/L in-line alum vs large

control biofilter effluent ................................................................................................................ 72

Figure 4.11: Normalized UF resistance profile showing the impact of 0.25 mg/L in-line alum vs

small control biofilter effluent ...................................................................................................... 72

Figure 4.12: Normalized UF resistance profile the impact of 0.5 mg/L in-line alum vs small

control biofilter effluent ................................................................................................................ 73

xi

Figure 4.13: Normalized UF resistance profile showing the impact of 1 mg/L of peroxide vs the

small control biofilter effluent ...................................................................................................... 74

Figure 4.14: Normalized UF resistance profiles depict the impact of filter media type .............. 74

Figure 5.1: Pilot-plant schematic ................................................................................................. 79

Figure 5.2: Average monthly turbidity trends. ............................................................................ 82

Figure 5.3: Average monthly DOC concentrations. .................................................................... 83

Figure 5.4: Average monthly UV254 values ................................................................................. 85

Figure 5.5: UV254 removal as a function of DOC removal. ......................................................... 84

Figure 5.6: Biopolymer (BP) and humic substance (HS) removals.. .......................................... 86

Figure 5.7: Average hourly filter headloss .................................................................................. 87

Figure 5.8: Typical headloss curves for the biofilter and pilot-plant (alum and PACl) filters

during summer 2003 and winter 2013 .......................................................................................... 87

Figure 5.9: Protein and polysaccharide concentrations for biofilter and conventional filter media

....................................................................................................................................................... 88

Figure 5.10: 24 hour DBP formation potentionals. ..................................................................... 89

Figure 5.11: Normalized UF resistance profile comparing coagulation (29.1-30.1 mg/L alum) to

biofiltration ................................................................................................................................... 91

Figure 5.12: Normalized UF resistance profile comparing coagulation (30.0 mg/L alum) with

filtration to biofiltration ................................................................................................................ 91

Figure 8.1: Sample DOC calibration curve ............................................................................... 109

Figure 8.2: Sample THM calibration curve (June 2013) ........................................................... 109

Figure 8.3: HAA calibration curves (June-September 2013) .................................................... 110

Figure 8.4: Sample PS calibration curve (June 2013) ............................................................... 111

Figure 8.5: Sample Pr. calibration curve (June 2013) ............................................................... 111

Figure 8.6: TCM control chart (June 2013) ............................................................................... 127

Figure 8.7: TCM control chart (July-Nov 2013) ....................................................................... 127

Figure 8.8: BDCM control chart (June 2013) ............................................................................ 128

Figure 8.9: BDCM control chart (July-Nov 2013) .................................................................... 128

Figure 8.10: DCAA control chart (June-Nov 2013) .................................................................. 129

Figure 8.11: TCAA control chart (June-Nov 2013) .................................................................. 129

Figure 8.12: AOX control chart (Sept-Nov 2013) ..................................................................... 130

Figure 8.13: Daily coagulant dose during the study .................................................................. 130

xii

Figure 8.14: Avarage daily raw water tempreture at the peterborough water treatment plant .. 131

Figure 8.15: Average daily DOC concentration in raw water, conventional, biofilter effluent 131

Figure 8.16: Average daily UV254 in raw water, conventional, biofilter effluent ...................... 132

xiii

NOMENCLATURE

˚C Degree(s) Celsius

< Less than

> More than

α Confidence level

1,2 DBP THM internal standard

2,3,5,6 TFBA HAA internal standard

µg/g Micrograms per gram

µg/L Micrograms per liter

Alum Aluminum sulfate

AOC Assimilable organic carbon

AOX Adsorbable organic halogens

ATP Adenosine tri-phosphate

BAC Biologically active carbon

BB Building blocks

BCAA Bromochloroacetic acid

BDCM Bromodichloromethane

BDCAA Bromodichloroacetic acid

BDOC Biodegradable organic carbon

BDOCfast Easily biodegradable organic carbon

BF Biofiltration

BOM Biodegradable organic matter

BP Biopolymers

BW Backwash

C Concentration, initial concentration

C0 Initial concentration

CA Cellulose acetate

CF Conventional filtration

cm Centimetre(s)

C:N:P Carbon: nitrogen: phosphorus ratio

xiv

Coag. Coagulation

D Diameter

d10 Effective diameter

DBAA Dibromoacetic acid

DBCAA Dibromochloroacetic acid

DBCM Dibromochloromethane

DBP Disinfection by-product

DBPFP Disinfection by-product formation potential

DCAA Dichloroacetic acid

DOC Dissolved organic carbon

EBCT Empty bed contact time

EC Expanded clay

EDC Endocrine disrupting compound

EM Emission

EPS Extracellular polymeric substances

EX Excitation

FEEM Fluorescence excitation emission matrixes

FCM Flow cytometry

FSP Peterborough water treatment plant full scale filter 11

g Gram(s)

GAC Granular activated carbon

GC Gas chromatography

GC-MS Gas chromatography – mass spectrometry

GC-ECD Gas chromatography-electron capture detection

H2O2 Hydrogen peroxide

HAA(s) Haloacetic acids

HPC Heterotrophic plate counts

HS Humic substances

kDa Kilo-Dalton

kPa Kilo-Pascal

xv

L Liter

LC-OCD Liquid chromatography-organic carbon detection

LMH Liter(s) per square meter per hour

LMW Low molecular weight

LOH Headloss

m Meter(s)

M Mean

m/h Meter per hour

m-1 1/meter(s)

m2 Squared meter(s)

MCAA Monochloroacetic acid

MBAA Monobromoacetic acid

MDL Method detection limit

MF Microfiltration

mg/L Milligram(s) per liter

mg/mL Milligram(s) per milliliter

min Minute(s)

mL/min Milliliter(s) per minute

mm Millimeter(s)

MTBE Methyl-tert -butyl-ether

NaOCl Sodium hypochlorite

NaOH Sodium hydroxide

NaSO4 Sodium sulphate

NF Nanofiltration

ng/g Nanogram(s) per gram

nm Nanometer(s)

NOM Natural organic matter

NR Not reported

NS Not Sampled, off-line

NTU Nephelometric Turbidity Unit

xvi

OCD Organic carbon detection

OND Organic nitrogen detection

PACl Polyaluminum chloride

pH -log (hydrogen ions concentration)

ppb Particles per billion

PPCP Pharmaceuticals and personal care products

Pr. Proteins

PS Polysaccharides

psi Pound(s) per square inch

PWTP Peterborough water treatment plant

QA/QC Quality assurance/quality control

R Pearson correlation coefficient

R2 Coefficient of determination

RO Reverse osmosis

RW Raw water

SD Standard deviation

SEC Size exclusion chromatography

SEC-OCD Size exclusion chromatography-organic carbon detection

SEM Scan electron microscopy

SL Sample lost

SPE Solid phase extraction

SUVA Specific ultraviolet absorbance (at 254 nm)

SW Settled water

t Student t-test value or student t probability distribution

TBM Tribromomethane

TBAA Tribromoacetic acid

TCAA Trichloroacetic acid

TCM Chloroform

THM(s) Trihalomethane(s)

TMP Transmembrane pressure

xvii

TOC Total organic carbon

UC Uniformity Coefficient

UF Ultrafiltration

UFS UF experiment samples

µg/g Microgram per grams

µL Microliters

UMFI Unified membrane fouling index

UV Ultraviolet

UV254 UV absorbance at 254 nm

1

1. Introduction

1.1 Background

The goal of providing safe and aesthetically pleasing drinking water has been

traditionally accomplished using conventional treatment (coagulation/flocculation/

sedimentation, filtration and disinfection). However, since the early 1980s, detectable

concentrations (low ng/L) of pharmaceuticals, personal care products (PPCPs) and other

emerging contaminates have been observed in drinking water sources and distribution

systems (Richardson and Bowron, 1985; Richardson and Ternes, 2011). Many of these

contaminants are potentially carcinogenic, mutagenic and maybe classified as endocrine

disrupting compounds (EDCs) (Richardson, 2009; Richardson and Ternes, 2011), and

conventional treatment has shown to be ineffective for their removal (Rana et al., 2012).

As well, stricter regulations to control disinfection by-products (DBPs) and other

compounds with possible human health impacts have dictated a need for more advanced

treatment processes (Malievialle et al., 1996). Therefore, the use of membranes have

emerged as a cost effective solution to: i) improve overall water quality, ii) reduce

treatment plant size (Singh, 2006), and, iii) comply with current and potential future

regulations (Plakas and Karabelas, 2012).

In the last decade, low pressure membrane filtration, including microfiltration (MF)

and ultrafiltration (UF), has gained momentum because of its capability of providing high

quality water at competitive cost when compared to conventional treatment (Neubrand et

al., 2010; Wang and Wang, 2006). Low pressure membranes have also been used as pre-

treatment prior to high pressure membranes; nanofiltration (NF) and reverse osmosis (RO)

(Chellam et al., 1997a). However, one of the major limitations of membrane processes is

fouling (Howe and Clark, 2002) which hinders treatment efficiency, damages the

membrane surface, and increases maintenance and operation costs (Singh, 2006). The

fouling rate and type (hydraulically revisable or irreversible) is a function of the

composition of the water matrix, and the foulants’ interaction with the membrane surface

(Peiris et al., 2010b). Metals, salts, colloids, organic matter, and suspended solids have

been identified as the main foulants (Gao et al., 2011). Most of these foulants are

2

hydrophobic and/or carry a surface charge (Henderson et al., 2011; Singh, 2006). However,

studies by others have shown that hydrophilic compounds may also contribute to fouling

(Brinkman and Hozalski, 2011; Kennedy et al., 2005). Fouling may occur by many

mechanisms including adsorption to the membrane surface and pores, and gel formation

(Henderson et al., 2011; Zheng et al., 2010). Older literature identified humic substances

(HS) as the main source of organic fouling (Wang and Wang, 2006). However, more recent

studies have shown that biopolymers, hydrophobic non-humic compounds, and colloids

might be the major source of irreversible organic fouling (Hallè, 2010; Neubrand et al.,

2010; Zheng et al., 2010; Zheng et al., 2012).

Many techniques have been developed to limit fouling including modifying

membrane materials, using pre-treatment methods such as activated carbon, enhanced or

in-line coagulation, and provision of anti-scaling treatment (Huang et al., 2009).

Coagulation has been widely applied as an effective membrane pre-treatment method

(Zheng et al., 2012). However, coagulation involves the addition of chemicals and might

result in severe irreversible fouling (Malievialle et al., 1996; Neubrand et al., 2010; Wang

and Wang, 2006).

Biofiltration presents a cost effective and chemical-free process for fouling control

(Hallé et al., 2009; Huck et al., 2011; Peldszus et al., 2011). Biofilters are simply granular

media filters operated without prior chemical disinfection, which allows microorganism

growth on the filter media and the biodegradation of a portion of the organic matter (Zhu

et al., 2010). Various configurations have been used including bank, slow sand, dual media,

and activated carbon biofiltration (Hammes et al., 2011). Europe has led the use of

biofilters because of their desire to provide high quality water; while in recent years,

biofiltration has become very popular in North America as the tradition of providing high

quality water was adapted (Huck and Sozański, 2008). Recent work has focused on the

active control of biofiltration to improve its performance (Lauderdale et al. 2012).

1.2 Research Objectives

The goal of this research was to optimize the use of biofiltration at pilot-scale to

minimize UF fouling. Specific objectives of the research included:

3

1- To assess the efficiency of biofiltration with respect to natural organic matter

(NOM), DBP precursor and UF foulant removal under a range of operating

conditions: filtration media (GAC vs. anthracite), nutrient addition (phosphorus

and nitrogen), hydrogen peroxide addition, and inline coagulation with

aluminum sulphate (alum).

2- To compare biofiltration to conventional treatment as pre-treatments to improve

UF performance and DBP precursor removal

3- To study the impact of filtration media, nutrient addition, hydrogen peroxide

addition and in-line coagulation on the development of biomass within the

biofilter and on filter runtimes.

1.3 Description of Chapters

Chapter 2: provides background information regarding natural organic matter

(NOM) and biofiltration.

Chapter 3: describes the experimental design, analytical methods, quality control/

quality assurance (QA/QC) procedures, sampling techniques, and data analysis

approach.

Chapter 4: discusses the effectiveness of adding phosphorus, nitrogen, peroxide,

and alum in controlling filter headloss, mitigating UF fouling, and removing DBP

precursors in with respect to pilot-scale study results and findings.

Chapter 5: provides results for pilot- and full-scale experiments, comparing

coagulation with/without filtration and biofiltration for fouling control, DBP

precursor, and NOM removal.

Chapter 6: provides an overall summary of the study and its findings, conclusions

and recommendations.

Chapter 7: contains a list of the references.

Chapter 8: contains appendices including raw data, supplementary figures, tables,

and QA/QC charts.

4

2. Literature Review

2.1 Natural Organic Matter (NOM)

NOM is a complex mixture of organic compounds that result from various aquatic

or vegetational biological processes and is traditionally divided into humics, and non-

humics portions (Barrett et al., 2000). NOM includes a wide variety of compounds that

differ in their physical and chemical properties which complicate their detection and

removal (Leenheer et al., 2000; Yavich et al., 2004). Humic substances (HS) are usually

subdivided into humic and fulvic acids and comprise about 50% of NOM; the other half

encompasses a wide range of compounds such as carbohydrates, amino acids, and

carboxylic groups (Leenheer and Croué, 2003). NOM composition and concentration in

water sources vary spatially and temporally depending on many factors such as climate and

land use (Brinkman and Hozalski, 2011). In general, NOM concentrations increase during

the summer because of higher surface runoff and biological activity (Jarvis et al., 2004).

NOM itself does not present a direct health concern (Hozalski et al., 1999),

however, it can affect overall water quality. NOM reacts with chlorine and other

disinfectants to form disinfection by-products (DBPs) (Chaiket et al., 2002; Richardson et

al., 2007; Wassink et al., 2011), causes microbial regrowth in the distribution system

(Persson et al., 2006; Yang et al., 2011), affects treatment processes efficiency (e.g. reduces

granular activated carbon (GAC) adsorption capacity) (Hozalski et al., 1999), and may lead

to membrane fouling (Howe and Clark, 2002; Huang et al., 2007; Lee et al., 2004; Zheng

et al., 2010).

To a certain extent, NOM removal can be achieved through conventional treatment

processes including coagulation, filtration, biological processes, and adsorption (Jacangelo

et al., 1995). Membranes offer superior and more efficient NOM removal, but are

susceptible to fouling (Brinkman and Hozalski, 2011). Biofiltration has been shown to be

an effective method for NOM removal, especially when considering the easily

biodegradable portion (Hozalski et al., 1999; Huang et al., 2011; Liu et al., 2001).

5

2.2 Biofiltration

Hammes et al. (2011) defined biofiltration as a filtration process where indigenous

bacteria and microorganisms inhabit the filter media and achieve one or more of the

treatment goals. Biofilters are operated without prior chlorination allowing

microorganisms to grow and consume organic matter (Huck and Sozański, 2008). Since

the early 1900s, biofiltration has been applied for drinking water treatment (Zhu et al.,

2010). Different configurations (e.g. rapid sand filters, slow sand filters, and bank

filtration) and media (e.g. sand, anthracite, and GAC) have been used. While slow sand

filters can provide higher NOM removals (10-50%), their application is limited by their

size requirements (2.4 m2 of filter surface area per liter of water produced per day) and

hydraulic loading rates (<0.3 m/h) (Ellis, 1985); bank filtration is also limited by location

(near river and lake banks) and lack of control over the flow rates (function of the soil

permeability and the topography of the area) (Tufenkji et al., 2002). As such, rapid single

and dual media filters have gained much popularity in North America because of their

ability to provide acceptable NOM removal (up to 50% DOC removal) while operating at

higher hydraulic loading rates (>5-7 m/h) (Huck and Sozański, 2008).

Biofiltration has been used for DBP precursor removal, limiting microbial growth

in the distribution system (Van der kooij, 1992; Wert et al., 2008; Yang et al., 2011), taste

and odour control (Elhadi et al., 2006; Srinivasan and Sorial, 2011), nitrification (Wahman

et al., 2011; Wert et al., 2008), pathogens removal (Zhu et al., 2010) and membrane fouling

control (Hallé et al., 2009; Huck et al., 2011; Peldszus et al., 2011).

2.2.1 Factors Affecting Biofilter Performance

NOM removal by biofiltration is influenced by many variables including filter

volume, NOM concentration and composition (Ahmad et al., 1998; Huck and Sozański,

2008; Liu et al., 2001). Huck and Sozański (2008) stated that pre-chlorination, temperature

and operation time are the major factors in determining NOM removal, while other factors

such as the filter media type, contact time, backwashing, NOM composition and

concentration are of less importance. Table 2.1 summarizes both the impact of these factors

and the degree of control over them (Huck and Sozański 2008)

6

Table 2.1: The degree of control and impact of factors influencing biofiltration (Adapted

from Huck and Sozański, 2008)

Parameter Degree of Control Effect

Media type H M

Chlorination M H

Filtration Rate (EBCT) M M

Backwashing method H M

NOM loading L M

Temperature N H

Time since start up N H

(H= high, M= moderate, L= low, N=none)

Other studies reported that the type of microorganisms growing on the biofilter

media (McDowall et al., 2009; Zhang et al., 2011a) and the amount of essential nutrients

in the biofilter effluent (Boon et al., 2011; Lauderdale et al., 2012; Vahala et al., 1998) can

also influence biodegradation of organics. The following presents a discussion of these

factors and their impacts on biofilter performance.

2.2.1.1 Contact Time

Contact time is an important factor in determining NOM removal during

biofiltration (Elhadi et al., 2006; Liu et al., 2001) because it affects the amount of organic

carbon that is adsorbed and biodegraded by the biofilm (Huck and Sozański, 2008). Contact

time maybe expressed as empty bed contact time (EBCT) as shown in equation 2.1.

𝐸𝐵𝐶𝑇 =𝐸𝑚𝑝𝑡𝑦 𝐵𝑒𝑑 𝑉𝑜𝑙𝑢𝑚𝑒 (𝑚3)

𝐹𝑙𝑜𝑤 𝑟𝑎𝑡𝑒 (𝑚3

𝑚𝑖𝑛)

(2.1)

The EBCT varies by the filter type and operating conditions; however, rapid dual

media filters are usually operated with an empty bed contact time of 5 to 20 min (Hammes

et al., 2011). EBCT impact varies on different NOM measurements; Yavich et al. (2004)

found that an EBCT of approx. 20 min is required to remove the easily biodegradable

dissolved organic carbon (BDOCfast) following ozonation, and that the required EBCT

varied by the water source and the ozone dose (from 0.5 to 3 mg/L). Carlson and Amy

(2001) reported the optimum EBCT for DOC removal to be 5.5 min. Wert et al. (2008)

7

observed that increasing a pilot-scale biofilter (diameter (D) =25.4cm) EBCT beyond 3.2

min by increasing the filter depth from 53 to 183 cm or increasing the filtration rate from

4.8 to 14.6 m/h did not affect assimilable organic carbon (AOC) removal. Also, Hozalski

et al. (1995) found that the difference in total organic carbon (TOC) removal in a bench-

scale sand biofilters operated at EBCT 4, 10 and 20 min was statically insignificant

(α=0.05). In contrast, Hallé et al. (2009) found that increasing the biofilter’s EBCT from 5

to 14 min prior to UF membrane increased biopolymer removal (initial concentration

(C0)=0.09-0.53 mg/L) by 21% on average, and lowered the membrane’s irreversible

transmembrane pressure (TMP) development rate by approximately 70%.

2.2.1.2 Water Matrix

Biofilter influent quality and composition impacts NOM removal; some studies

have shown that a biofilter’s ability to remove organics varies by water source (Hammes

et al., 2011; Yavich et al., 2004). Persson et al. (2006) found linear correlations between

the concentrations of the biologically degradable organic carbon (BDOC) and AOC in the

biofilter effluent to their removal (R=0.8 and 0.5, respectively).

Liu et al. (2001) reported that the presence of 3 mg/L of alum residual or 1.5 mg/L

of clay (particles) in a bench-scale anthracite/sand and granular activated carbon

(GAC)/sand biofilters (D=50 mm, 0.7 m deep, and EBCT= 5.6 min) influent had an

insignificant impact (95% confidence) on the degradation of acetate (C0=300 μg/L),

formate (C0=400 μg/L), formaldehyde (C0=100 μg/L), and glyoxal (C0=30 μg/L). Influent

pH is another important parameter in determining organics removal since it can affect

microorganism growth and activity (Rittmann, 2001); Egashira et al. (1992) found a pH of

7 to 9 to be optimal for MIB removal.

2.2.1.3 Filter Media

Many studies have explored the use of different filtration media to improve biofilter

performance. Sand, anthracite, and GAC have been intensively studied (Dussert and

Tramposch, 1997; Emelko et al., 2006; Liu et al., 2001). Most of the studies found that

GAC and anthracite biofilters provided similar TOC and biodegradable organic matter

(BOM) removal (α=0.05) at 20-25˚C (Emelko et al., 2006; Liu et al., 2001); however, GAC

8

provided higher TOC and AOC removal (up to 100% more than anthracite) at low

temperatures (1-10 ˚C), when lab and pilot-scale biofilters were backwashed with water

containing 0.5 to 1 mg/L chlorine residual, and during the filter start-up (Dussert and

Tramposch, 1997; Emelko et al., 2006). The reason behind GAC superiority as biofilter

media is the high surface area (>1000 m2/g) and the porous nature (pores volume >0.15

cm3/g) of its particles which support microorganisms’ attachment and growth (Dussert and

Tramposch, 1997; Klymenko et al., 2010). Also, microorganisms have better adhesion to

GAC because they adsorb to its surface and pores which protects them from detachment

during backwash (Hammes et al., 2011; Zhu et al., 2010). Liu et al (2001) reported that

GAC is capable of supporting twice the phospholipids mass in comparison to an

anthracite/sand filter at 20 C. Likewise, many researchers have hypothesized that the GAC

superior performance may be explained by the bio-regeneration of GAC: microorganisms

consume NOM attached to GAC, which reactivate the adsorption sites for binding with

new compounds (Velten et al., 2011a; Zhu et al., 2010). Other materials such as expanded

clays (EC), ceramics and plastics have been used as a filtration media (Okei et al., 2009);

Persson et al. (2006) found that GAC and expanded clay (EC) biofilter were able to remove

34 % and 30 % of BDOC (C0= 1.06 ± 0.25 mg/L), respectively.

2.2.1.4 Temperature

Studies have shown that TOC, AOC, BDOC, and organic compound removals were

reduced at low temperatures (<10˚C) (Andersson et al., 2001; Liu et al., 2001; Moll et al.,

1999). Liu et al. (2001) noticed a reduction in acetate removal (C0 = 300 μg/L) from approx.

80% to 50% in bench-scale anthracite/sand biofilter (backwash contained 0.5 mg/L

chlorine residual) when the temperature decreased from 20 to 5˚C, but the removal in the

GAC/sand biofilter remained constant at approx. 80%. However, Emelko et al. (2006)

found the difference between TOC and oxalate removal in low (1-3˚C) and warm (21-25˚C)

temperatures to be statically insignificant (α=0.05) for both GAC and anthracite biofilters.

Moll et al. (1999), in a bench-scale study, observed constant TOC, BDOC and AOC

removals (24%, 60%, and approximately 55% respectively) between 20 and 35˚C;

however, at 5˚C, the removals decreased to 15%, 38%, and 43%, respectively, for initial

concentrations of 4±0.3, 1.6±0.2 and 1.4±0.18 mg/L, respectively. These reductions were

9

associated with changes in the microbial species from positive gram and sulphate reducing

bacteria to negative gram bacteria. Hallé et al. (2009) noted a reduction in the biopolymers

removal (C0= 0.09-0.53 mg/L) in biofiltration (EBCT= 10 min) from approx. 70% to 35%

when the temperature decreased from 10-25˚C to 1-4˚C between warm (April to

September) and cold (October to March) months. Velten et al. (2011a) stated that the lower

temperature in their study (7 ˚C) compared to others (9-20˚C) caused lower microbial

growth rate, (0.0001-0.0043 h-1) vs. (0.038-0.16 h-1).

2.2.1.5 Backwashing

The effect of using chlorine and/or air scour during backwashing on NOM removal

varies widely between different studies. Urfer (1998) observed a decrease in acetate (C0=

300 µg/L) and formate (C0= 400 µg/L) removals from approx. 90% to 40% when 1 mg/L

free chlorine was present in the biofilter backwash. Miltner et al. (1995) reported that 1

mg/L free chlorine in the backwash reduced TOC removal (C0= 1.1-2.2 mg/L) from 21%

to 16% on average and decreased trihalomethanes (THMs) and haloacetic acids (HAAs)

formation potential (at 7 days) removal (C0 not reported) from 43% and 27% to 29% and

15%, respectively. However, in later studies, Wert et al. (2008) found that 1.5 mg/L free

chlorine in the backwash did not impact AOC removal (C0 = 475±45 µg/L) in a pilot-scale

biofilter; which is in agreement with Liu et al. (2001) bench-scale studies where they

showed that 0.5 mg/L free chlorine in the filter backwash did not reduce BOM removal.

Emelko et al. (2006) found that using air scour and collapse pulsing during

backwash had an insignificant impact (α=0.05) on TOC removal (C0= 5-7 mg/L), which is

in agreement with Wert et al. (2008). Zhu et al. (2010) stated that backwashing is necessary

to preserve the steady-state condition (constant AOC, BDOC, or specific compound

removal) and avoid filter clogging. In a model proposed by Rittmann et al. (2002), it was

shown that 27% of the biomass was removed during backwash while BOM removal

(C0=2.3 mg oxygen demand/L) remained constant at 80%. In a recent study, Huck et al.

(2011) found that the standard backwashing procedure (collapse-pulsing with 50% bed

expansion) increased the TMP development rate of a UF unit downstream of the biofilter

directly after the backwash (no values reported); by eliminating air scour, they were able

to maintain a constant TMP development rate.

10

2.2.1.6 Biological Activity

Biological activity in a biofilter is controlled by the species of microorganisms

present, time from start-up, and the concentration of the nutrients (e.g. carbon,

phosphorous) in the filter influent. Biofilter microorganisms consist mostly of prokaryotes

(Moll et al., 1998), and are subjected to indigenous conditions (Hammes et al., 2011). Many

studies have attempted to characterize and classify microbial communities within biofilters

(Boon et al., 2011; Moll et al., 1998; Zhang et al., 2010; 2011a; Zheng et al., 2011);

however, their findings varied in terms of the dominant microbial species and relative

abundance. Yang et al. (2011) identified beta-proteobacteria as of the dominant species in

a full-scale biofilter (Cheng-Ching Lake, China) and hypothesized that they were

responsible for AOC removal. Zhang et al. (2011a) reported the dominant bacterial species

in a pilot-scale biological activated carbon (BAC) biofilter to be pseudomonas sp, bacillus

subtilis, and nitrospira sp. Bacillus species were also identified by Zhang et al. (2010) in a

bench-scale BAC biofilter and were hypothesized to be responsible for degrading a wide

range of organic compounds including polysaccharides which are considered a major UF

foulant (Hallé et al., 2009). McDowall et al. (2009) reported higher geosmin removal (up

to 75% in the inoculated biofilter in comparison to 25% in control) when they seeded

bench-scale biofilters (D= 2.5 cm, height= 15 cm) with previously isolated and cultured

geosmin degrading bacteria (sphingopyxis sp., novosphingobium sp. and pseudomonas

sp.); however, their experiment was run for a short time (12-16 days) and their approach

was not verified by other studies.

A carbon: nitrogen: phosphorus (C:N:P) ratio of 21:5:1 by weight has been reported

as being required for optimal bacterial growth; however this ratio is not typically attained

in a biofilter influent (Liu et al. 2001). Some studies have shown phosphorus to be the

limiting nutrient (Sang et al., 2003); moreover, Lauderdale et al. (2012) observed a 15%

decrease in headloss and a 75% increase in DOC removal (C0= 3.6±0.1 mg/L) when a

C:N:P stoichiometric ratio in a pilot-scale biofilter influent (D= 15.24 cm, height= 1.22 m)

was adjusted to more than 100:10:2.

Time to reach steady-state in terms of stabilizing the microbial community is also

a factor in controlling NOM removal. Velten et al. (2011a) reported that a consistent DOC

11

removal of 22% (C0=1.1±0.04 mg/L) and ATP biomass concentration (0.8-1.83×10-3 g

ATP/g GAC) was reached after 90 days in a pilot-scale GAC biofilter (D=1.1 m, height=

1.55 m). In their review, Hammes et al. (2011) reported that filter start-up periods varied

between different studies from 2 to 8 months. Zhang et al. (2011a) found that 9 months

were required to develop a stable microbial community and reach steady-state conditions.

Wert et al. (2008) observed a variation in the time required to achieve constant removals

of different NOM measurements (e.g. 75 days for AOC, 155 days for formaldehyde). Liu

et al. (2001) found that reaching steady-state conditions varied with respect to compound,

temperature and filter media; furthermore, at 20℃, 20 days were required to reach steady

removal of easily biodegradable BOM (acetate, formate, formaldehyde) in comparison to

40 days for non-easily biodegradable BOM (glyoxal) in both GAC and anthracite biofilters.

Similarly, at 5℃, the results did not vary for the easily biodegradable compounds in both

media; however, glyoxal was not removed by the anthracite filter at this temperature and

required 60 days to reach steady-state removal in GAC.

2.2.1.7 Ozone Pre-treatment and Hydrogen Peroxide Addition

The combination of ozone and biofiltration is widely used in drinking water

treatment because biofiltration can limit regrowth potential resulting from increased AOC

and BDOC concentrations during ozonation (Hozalski et al., 1999; Persson et al., 2006).

In a full-scale study, Yang et al. (2011) observed that AOC concretions increased by 57%

during ozonation and decreased by 83% following BAC. Yavich et al. (2004) reported an

increase in BDOC concentration by 200-240% in river water when they applied an ozone

dose of 0.5 to 1 mg O3/mg carbon. However, when the same dose was applied to lake water,

no increase in BDOC was observed. This suggests that the ozone effect varies by water

source and ozone dosage, which is in agreement with Hozalski et al. (1999) bench-scale

studies and computer model. Carlson and Amy (2001), in bench and pilot-scale studies,

found the optimum ozone dose for different water sources to be between 0.4-0.6 O3/DOC,

which approximated a typical disinfection dose required for 3-log cryptosporidium

inactivation.

Urfer and Huck (2000) did not observe a change in BOM removal (acetate and

formate removal remained >95%) when a 0.1-0.2 mg/L ozone residual or 0.5 mg/L

12

hydrogen peroxide residual was present in the influent water of a lab-scale anthracite/sand

biofilter (D= 50 mm, height= 70 cm). However, Boon et al. (2011) attributed a 30%

decrease in the biomass at the top 10 cm of a biofilter to the presence of 0.22 mg/L ozone

residual in a pilot-scale GAC biofilter influent (D= 1.1 m, height= 1.55 m). Lauderdale et

al. (2012) reported the addition of 1 mg/L hydrogen peroxide to filter influent decreased

headloss by 66% while maintaining DOC removal at 15±3%. No peroxide residual was

detected in the biofilter effluent (< 0.1 mg/L) when 1 mg/L hydrogen peroxide was added

prior to biofiltration (Lauderdale et al., 2012; Urfer and Huck, 1997).

2.2.2 NOM Monitoring Techniques for Biofiltration Studies

Researchers have developed different methods to monitor the concentrations of

NOM in water. TOC, DOC, and UV absorbance at 254 nm (UV254) were traditionally used

for this purpose and are still commonly employed. TOC is measured by oxidizing the

sample through combustion at 600-800℃ and then measuring carbon dioxide produced

using an infrared detector (Qian and Mopper, 1996). Similarly, DOC is measured by the

same technique but after passing the sample through 0.45 μm filter (Qian and Mopper,

1996). Hallé et al. (2009) noted that DOC removals are not a good indication of the

performance of the biofilter as a pre-treatment for membranes, as the removals did not vary

with EBCT. However, in most biofiltration studies DOC has been utilized as a general

measure of NOM removal. UV254 has been used to measure the concentration of the

aromatic compounds in water. Specific UV absorbance (SUVA) is the ratio of UV254 to

DOC, and it is used as a measure of the aromaticity of the organic carbon in water. SUVA

values above 4 are associated with hydrophobic NOM whereas values less than 3 are

associated with hydrophilic NOM (Matilainen et al., 2011).

The previously mentioned techniques are not representative of the biodegradation

process in biofiltration as they describe only the chemical nature of the organic compounds

and not their tendency to be metabolized. Therefore, biological techniques such as AOC

and BDOC have been developed to measure the biodegradable portion of NOM. AOC

correlates the growth of test organisms (pseudomonas fluorescens P17 and spirillum NOX)

to the concentration of easily biodegradable (assimilable) NOM; whereas BDOC measures

the consumption of DOC by a biofilm, grown on glass beads or filter media, to estimate

13

the biodegradable DOC (Escobar and Randall, 2001). AOC and BDOC concentrations in

drinking water sources have been reported to range between 0-9% (Leenheer and Croué,

2003) and 0-40% (Escobar et al., 2000) of TOC, respectively. Unlike AOC, BDOC mimics

the biofilm processes since it stimulates the biodegradation of complex compounds (i.e.

enzymatic biodegrading). However, both techniques are laborious and time consuming

(AOC, 2-14 days and BDOC, 3-5 days) (Hammes et al., 2011). Velten et al. (2008) stated

that AOC is an inadequate measurement of NOM removal since it only presented 35% of

DOC removed during biofiltration and suggested the use of BDOC instead.

Methods to fractionate NOM by hydrophobicity, size and photonic characteristics

provide researchers with the ability to identify components of organic carbon and to tackle

concerns related to specific NOM components (Matilainen et al., 2011). For example,

biopolymers as identified by liquid chromatography-organic carbon detection (LC-OCD),

as well as colloids and protein-like substances as identified by fluorescence excitation-

emission matrices (FEEM) were found to be responsible for UF fouling, and have been

shown to be readily removed by biofiltration (Hallé et al. 2009; Peldszus et al. 2011).

Polymeric adsorbents maybe used to fractionate NOM based on hydrophobicity

(Matilainen et al., 2011). Chow et al. (2004) was able to fractionate NOM as; i) very

hydrophobic, ii) slightly hydrophobic, iii) charged hydrophilic, and iv) neutral hydrophilic

acids by using three consecutive resins (DAX-8, XAD-4, and IRA-958). The

hydrophobicity of NOM can impact its removal by biofiltration and affect the degree of

fouling. For example, BDOC, which can be removed by biofiltration was correlated with

charged hydrophilic (R2= 0.78-0.87) and neutral (R2= 0.88) fractions of DOC (Tihomirova

et al., 2010). Hydrophobic NOM was found to responsible for membrane fouling (Drioli,

2009). Chen et al. (2011) found that BAC (height= 1.5 m, and flow rate= 4 m/h) after

ozonation was capable of reducing the hydrophobic (C0 approximately 1.5 mg/L) and

hydrophilic (C0 approximately 1.5 mg/L) carbon by 29% and 22%, respectively. Removals

of hydrophilic and hydrophobic fractions were attributed to biodegradation and adsorption,

respectively. Also, Tian et al. (2009) reported biologically activated carbon (D= 6 cm,

height= 1 m) to be more effective for the removal of hydrophobic neutrals, hydrophobic

acids, and hydrophilic NOM (4 to 14% higher) when compared to a membrane bioreactor.

14

Another method that is used to characterize NOM is liquid chromatography-organic

carbon detection (LC-OCD), which is a variant of size exclusion chromatography (SEC)

where the SEC column leads to a non-destructive UV analyzer (at 254 nm) and a

destructive organic carbon detector (OCD). NOM is oxidized to carbon dioxide in a

Grantzel thin-film reactor and measured by an infrared detector (Baghoth et al., 2009;

Huber and Frimmel, 1994). LC-OCD fractionates NOM into the components shown in

Table 2.1 (Huber and Frimmel, 1994; Penru et al., 2011). A typical LC-OCD

chromatogram is shown in Figure 2.1 (Huber et al., 2011a). Organic nitrogen detection

(OND) can be coupled with OCD to identify proteinic matter within each fraction

(Henderson et al., 2011; Huber et al., 2011a).

Figure 2.1: Typical LC-OCD diagram with OCD, OND and UV detectors (A=

biopolymers, B= HS, C= building blocks, D= LMW acids, E= LMW neutrals) (after Huber

et al. (2011a))

15

Table 2.2: LC-OCD fractions

Fraction

Molecular weight

(kDa) (Penru et al.,

2011)

Compounds (Huber et al.,

2011b)

Biopolymers >10-20 Polysaccharides, proteins or

amino sugars

HS 0.8-1 Humic and fulvic acid

Building blocks 0.35-6 HS breakdown products

Low molecular weight

(LMW) acids <0.35 Not applicable

LMW neutrals <0.35 LMW alcohols, aldehydes,

ketones, sugars, and amino acids

Hydrophobic organic carbon Not applicable Not applicable

Fluorescence excitation-emission matrices (FEEM) analysis is a relatively new

method of representing organic matter that has the potential of being integrated on-line to

provide real-time measurement of NOM fractions (Peiris et al., 2010a; Stedmon et al.,

2011). FEEM can be used to differentiate protein-like, tyrosine-like, humic-like, fulvic-

like, and colloidal matter (Matilainen et al., 2010). Peiris et al. (2010a) utilized FEEM and

principle component analysis to identify three fouling components: humic-like substances,

protein-like substances, and collides/ particulate matter. Using the same approach,

Peldszus et al. (2011) correlated the concentration of protein-like substances in a UF

membrane feed to irreversible fouling (R2= 0.87-0.96) when TMP was less than 6 psi. They

reported that protein-like substances can be reduced by biofiltration (no values reported),

and that their removal was found to be increasing with increasing EBCT (Peldszus et al.

2011).

The previously described techniques can also be used in tandem for further

characterization of NOM. Penru et al. (2011) utilized LC-OCD following resin

fractionation to determine the hydrophobicity of different NOM fractions. A similar

approach was applied by Kennedy et al. (2005) where they attributed irreversible fouling

to hydrophilic colloidal matter (mainly polysaccharides).

Other techniques such as mass spectrometry and membrane filtration are available

for NOM characterization but their use is limited because they require method

16

development, may damage the organic compounds, and lead to false results (Matilainen et

al., 2011).

LC-OCD and FEEM appear to be the most promising techniques in recent

biofiltration studies (Peldszus et al., 2011). Attempts to correlate their fractions (e.g.

biopolymers) with fouling and removal by biofiltration were successful (Hallé et al. 2009;

Peldszus et al. 2011). However, these correlations were not tested on different water

matrices, and recent research has shown no correlation between biopolymer concentrations

and irreversible fouling across different water matrices in Ontario (Croft, 2012).

2.2.3 Biomass Quantification and Qualification Methods

Many researchers have developed techniques to quantify and qualify the

microorganisms in biofilters (Zhu et al., 2010). Biomass measurements provide an insight

regarding the development of microbial communities within the biofilter and assist in

identifying and diagnosing potential problems (Hammes et al., 2011). Many biomass

monitoring methods have been used, such as heterotrophic plate counts (HPC),

phospholipids and scan electron microscopy (SEM). These measurements can be applied

either to the biofilter effluent or directly to the filter media.

Measuring biomass in biofilter effluent is preferable to measuring it on the filter

media since it does not involve disturbing the biofilter bed and can sometimes be integrated

on-line (e.g. Flow cytometry (FCM)) (Hammes et al., 2011). Plate counts, particle counts,

and ATP analysis are the three main techniques used to measure the biomass in the filter

effluent. HPC measures the number of cells after cultivating bacteria isolated from a

sample (for 2 to 5 days) (Allen et al., 2004). It is a widely accepted measure for microbial

regrowth potential in the distribution system, and is used for regulatory purposes in some

countries (Allen et al., 2004). A viable alternative for HPC is the use of cell counters, which

can yield faster measurements as they do not require cultivation (Wu et al., 2008). Also,

microscopy can be used to characterize the bacteria in terms of shape, size and even species

(Hammes et al., 2008; Wang et al., 2010). Adenosine triphosphate (ATP) analysis measures

the concentration in a water sample using a luminometer following the addition of an

enzyme reagent and buffer (Magic-Knezev and van der Kooij, 2004; Velten et al., 2007).

17

ATP analysis is faster than HPC, less expensive than cell counts, and can be correlated to

both HPC and cell counts (Siebel et al., 2008). However, one of the major drawbacks of

this method is that it does not yield any specific information regarding the cell size or

properties, and that it has not been widely used in biofiltration studies (Hammes et al.,

2008).

Biomass on filter media can be measured using different methods including

phospholipids and microscopy (Hammes et al., 2011). The most widely used method is

phospholipids which measures bacterial membrane constituent absorbance at 610 nm after

adding a reagent and a buffer to separate the biomass from the media (Wang et al., 1995).

ATP analysis can also be used on sample media after isolating the biofilm by chemical or

physical means. Phospholipids are considered to be a more direct measurement of the total

biomass; however, unlike ATP they are not indicative of the active biomass which is more

of an interest in biofiltration studies (Hammes et al., 2011). NOM removal was not found

to be correlated to the concentration of the biomass in terms of adenosine triphosphate

(ATP) (R2=0.03 for AOC) (Velten et al., 2011a) or phospholipids (R=0.1 for AOC and

BDOC) (Persson et al., 2006). In addition, microscopic methods has been used to

characterize the microbial communities and their distribution on the biofilter media but

their use is limited due to their high cost and complexity of their results (Madrid and Felice,

2005). Other methods such as the biomass activity and uptake of labelled substances are

also available but are not widely used due to their time requirement (5 hours to days) and

laborious nature (Urfer and Huck, 2001).

2.3 Biofiltration for UF fouling control

The concept of applying biofiltration as a UF membrane pre-treatment to reduce

organic fouling has been reported by others (Basu and Huck, 2004; Hallé et al., 2009; Huck

et al., 2011). Moreover, Tsujimoto et al. (1998) utilized a pilot-scale biologically active

carbon, with a media depth of 1 m and EBCT of 6 min, to treat water prior to a cellulose

acetate UF. They found that it could operate for 80 days before reaching a limiting TMP

of 80 kPa. Basu and Huck (2004) observed a reduction in the cleaning frequency of a

hollow fibre UF (defined by TMP exceeding 28 psi) from 53 to 28 days in one run and

from 69 to 56 days in the other when a bench-scale anthracite/sand biofilter (depth= 0.8 m,

18

EBCT of 22 min) was used as a pre-treatment. Mosqueda-Jimenez et al. (2008) reported

that using bench-scale anthracite/sand biofilter with an EBCT of 35 min prior to UF

decreased the decline in membrane permeability from 56% to 11% at 5500 L/m2 permeate

volume. Hallè et al. (2009) observed a TMP decrease from approximately 9 psi at permeate

volume of 2000 L/m2 to approximately 4 psi and approximately 1.2 at 7500L/m2 when

applying biofiltration with EBCT of 5 and 15 min, respectively to pre-treat a hollow fiber

UF. Zhang et al. (2011b) reported that combining biofiltration in conjunction with

coagulation maintained the TMP after 144 hours of operation at approximately 17 kPa in

comparison to approximately 24 kPa when coagulation was applied alone (EBCT not

reported). Peldszus et al. (2012) found that pilot-scale biofilters operated at 5, 10 and 15

min EBCTs were effective for fouling control even at low temperatures (1-25℃). However,

Geismar et al. (2012) used a lab-scale BAC (EBCT=11 min) post ozonation to evaluate its

impact on the unified membrane fouling index (UMFI), a recent method to quantify

reversible and irreversible fouling, and found that the biofilter effect on UMFI following

ozonation to be statically insignificant (α=0.05).

2.4 Coagulation as a Pre-treatment for UF

Coagulation enables the destabilization of particles by different mechanisms (e.g.

charge neutralization, and sweep flocculation) to facilitate the formation of bigger particles

that can be removed by settling or filtration (Crittenden et al., 2012). Coagulation has been

shown to be effective for removing NOM and DBP precursors (Matilainen et al., 2010).

Enhanced coagulation aims to maximize particles, TOC, and DBP precursor removals

while minimizing coagulant residual and sludge production (Edzwald and Tobiason, 1999).

Polyaluminum Chloride (PACl) can provide similar performance to alum at lower

coagulant dosages while being temperature independent (Van Benschoten and Edzwald,

1990). Different coagulation configurations have been applied as UF pre-treatment

including coagulation/flocculation with/without sedimentation (Crittenden et al., 2012),

and in-line coagulation (Neubrand et al., 2010; Wang and Wang, 2006; Zheng et al., 2012).

Some studies have observed an improvement in UF fouling mitigation when alum is used

as a pre-treatment (Choi and Dempsey, 2004; Li et al., 2013) while others did not (Howe

and Clark, 2006; Neubrand et al., 2010). Howe and Clark (2006) observed that enhanced

19

coagulation was effective for fouling mitigation but an adverse effect was observed at

lower doses. Wray et al. (Wray et al., 2014) did not observe an improvement when applying

coagulation (<15 mg/L alum, below enhanced dosages) as a pre-treatment for UF

membranes and found the point of diminishing return for biopolymer removal to be 0.1

mg/L.

2.5 Research Gaps and Needs

Biofiltration as pre-treatment to UF membrane has been previously explored as

discussed earlier. The results of these studies show that biofiltration is a promising

technique for fouling control; however, more work needs to be done to verify this potential,

to optimize biofiltration conditions, and to explore strategies to control and enhance

specific foulant removal. Also, a detailed evaluation of the process and its mechanism using

a range of water matrixes and membrane modules have not been conducted (Hallè, 2010).

In addition, a parallel comparison to other pre-treatments (e.g. coagulation, GAC) has not

been performed (Peldszus et al., 2012).

Active application of biofiltration including the use of nutrient and hydrogen

peroxide addition, pre-ozonation, and pre-aeration on the biofilter ability to limit UF

fouling has not been examined. Previous studies have shown that GAC can be more

effective than anthracite as a biofiltration media especially at low temperatures (< 3 ℃)

(Emelko et al., 2006); a study to compare GAC and anthracite as biofiltration media to

reduce ultrafiltration fouling is needed. Lauderdale et al. (2012) found that adjusting the

C:N:P ratio to beyond 100:10:1 increased DOC removal by 75 % but they did not report

the impact of this procedure on foulants (e.g. biopolymers) or DBP precursors. Also, the

impact of applying hydrogen peroxide for extended durations on fouling precursor removal

has not been evaluated.

The combination of biofiltration and in-line coagulation may offer superior fouling

control at a lower cost when compared to coagulation or biofiltration separately (i.e. lower

coagulant dose, lower EBCT, and/or longer membrane operation time). Moreover,

biofiltration has been shown to be effective for the removal of biopolymers (61 ± 22 % at

EBCT of 15 min), but not HS (e.g. 7% HS removal when biopolymer removal was 86%)

20

(Hallé et al., 2009; Nguyen and Roddick, 2010). However, coagulation has been shown to

be effective for HS removal (e.g. in-line coagulation improved the removal of natural

humic acid (<1 kDa) by UF from 6.1% to 59.3%) (Wang and Wang, 2006). Therefore, their

combination has the potential to offer higher DOC removal, and better fouling control.

Also, it may be more practical and convenient to adjust the coagulant dose as opposed to

the EBCT during severe fouling events (e.g. increased NOM or turbidity) or when

biological activity is suppressed at cold temperatures (< 10℃). Therefore, the application

of coagulation prior to or following biofiltration may impart more control and flexibility in

operation especially at low coagulant dosages (< 0.5 mg/L) which could be explored prior

and post biofiltration to improve fouling mitigation.

In terms of ultrafiltration, the correlation between biopolymer concentrations and

irreversible fouling has not been well established. Croft (2012) was able to correlate

biopolymers concentration in different source waters in Ontario, Canada to reversible

fouling but was not able to reach the same conclusion for irreversible fouling. These results

indicate that biopolymer composition and properties might alter its effect on fouling. This

problem needs to be addressed by further characterization of biopolymer properties which

can be achieved through using a different detection method following size exclusion

chromatography (such as FEEM) (Matilainen et al., 2011; Wu et al., 2003), or by using

resin fractionation prior to LC-OCD (Penru et al., 2011). FEEM may also provide insight

into the nature of the LC-OCD fractions (e.g. protein-like biopolymers); also, resin

fractionation can be used to further characterize LC-OCD fractions based on their

hydrophobicity. Additionally, the interaction of biopolymers with other water constituents

(e.g. turbidity, and particulate matter) need to examined.

12

3. Materials and Methods

3.1 Research Objective

The goal of this research was to explore the effectiveness of biofiltration and enhancement

strategies (i.e. nutrient enhancement (phosphorus and nitrogen), in-line coagulation, and hydrogen

peroxide (H2O2) addition) in terms of natural organic matter (NOM) removal (DOC, LC-OCD

fractions), disinfection by product (DBP) precursor reduction (formation potential of THMs,

HAAs, and AOX), ultrafiltration (UF) fouling control, and filter runtimes improvement (headloss

(LOH) reduction). This will be achieved by; i) supplementing phosphorus and nitrogen as a

function of DOC removed (i.e. adjusting the carbon to nitrogen to phosphorus (C:N:P) ratio), ii)

adding hydrogen peroxide (H2O2), and iii) varying the biofilter media type (granular activated

carbon (GAC) vs. anthracite) for Otonabee river water, Peterborough, ON. Finally, the impact of

in-line coagulation (0.1-0.5 mg/L alum with rapid mixing and no flocculation) prior to biofiltration

was evaluated for each of the measured parameters. A side-by-side pilot- and full-scale comparison

between biofiltration and conventional treatment was also conducted.

3.2 Experimental Configuration

3.2.1 Biofilter Size and Media Depth

To select the diameter (D) of a biofilter column, two considerations should be taken into

account; firstly, the amount of water available for filtration, and secondly, minimizing the wall

effects. Table 3.1 summarizes the dimensions and flow rates of biofilters used in previous pilot-

scale biofiltration studies. Some lab- and pilot-scale studies have utilized columns with 5-6 cm

internal diameter as biofilters (Elhadi et al., 2004; Tian et al., 2009); while others such as the ones

shown in Table 3.1 used internal diameters of up to 25.4 cm. The selection of a diameter within

the range of these studies should be adequate. For the current research, two sizes of biofilters were

selected; a large diameter (D= 15.24 cm (6 in)) which will be used for the purpose of comparing

biofiltration with or without nutrient (nitrogen and phosphorus) enhancement to conventional

filtration, and a small diameter (D= 7.62 cm (3 in)) which will be used for the evaluation of

different enhancement strategies (more discussion is provided later). The media depth and

22

configuration was matched to the conditions in the Peterborough pilot-scale conventional drinking

water plant.

Table 3.1: Summary of the characteristics of biofilters in previous studies

Study Diameter

(cm)

Media

Type

Depth

(cm)

Sand

Depth

(cm)

Gravel

Depth

(cm)

EBCT

(min)

Halle et al.

(2009)

NR Anth. 20 20 15 4.8

NR Anth. 20 97 15 14

Huck et al.

(2011)

15 Anth. 20 20 15 5

20 Anth. 20 63 15 10

15 Anth. - 40 15 15

Lauderdale

et al. (2012) 15 GAC 100 20 Cap 6.65

Wert et al.

(2008)

25.4 Anth. 53 25 - 3.2

25.4 Anth. 53 25 - 9.7

25.4 Anth. 183 20 - 8.3-25

Persson et

al. (2006)

20 GAC 100 - 15 31

20 EC 100 - - 31

Tian et al.

(2009) 6 GAC 100 - - NR

(Anth. = anthracite, GAC= granular activated carbon, EC= expanded clay, NR= not reported,

Cap= supporting cap for drainage)

3.2.2 Pilot-scale Plant

The pilot plant is located at the Peterborough Water Treatment Plant (PWTP) on the

Otonabee River (Peterborough, ON). It is fed by pre-chlorinated raw water (dose= 0.5 mg/L Cl2,

when the temperature is >12 ºC) with a DOC between 4 to 6 mg/L.

The pilot-plant consisted of two parallel conventional treatment trains (using alum and

PACl as coagulants) and two biofiltration treatment trains. The coagulation train consisted of a

rapid mix, tapered flocculation, parallel plate settlers, and dual media filtration (EBCT= 12-36

min, Diameter (D) =15.24 cm). The biofiltration train consisted of two large biofilters (D=15.24

cm) and four smaller biofilters (D = 7.62 cm), operated at an empty bed contact time (EBCT) of

11 and 10 min, respectively. Figure 3.1 shows a general schematic of the pilot plant configuration.

23

Figure 3.1: Schematic of the pilot-plant treatment processes showing the conventional treatment

trains and the biofiltration treatment trains. (Not to scale) (BF= biofilter, CF= conventional filter)

24

3.2.3 Biofiltration Treatment Train

The biofiltration and the conventional treatment trains were fed directly by a constant head

tank. Two large (D= 15.24 cm) dual-media anthracite/sand biofilters matching the size and media

depth (1 m) of the pilot-scale conventional filters were used for the study. They were constructed

of glass, and all fittings and piping were either stainless steel or Teflon®. The filters contained 50

cm of anthracite (effective size (d10)= 0.85, uniformity coefficient (UC)= 1.8) on top of 50 cm of

sand ( d10= 0.5, UC= 1.8); a wedge wire support plate was used instead of gravel. These large

filters, named biofilter 1 (BF1) and biofilter 2 (BF2) were operated in parallel with an empty bed

contact time (EBCT) of 11 minutes. The water level above each filter was maintained at 90 cm

from the top of the filter media (constant head mode). The filters were backwashed using their

unchlorinated effluent (with or without air scour) when the headloss (LOH) reached 150 cm, when

air binding occurred, or a maximum of three times a week (on Monday, Wednesday and Friday).

The backwash procedure included a slow backwash for 4 minutes (30% bed fluidization) followed

by 8 minutes of fast backwash at collapse pulsing conditions (50 % bed fluidization), and 2 minutes

of slow backwash.

The small-scale (D= 7.62 cm) biofiltration train consisted of four parallel filtration

columns. The filters were constructed of acrylic with PVC fittings. They contained 50 cm of

anthracite on top of 50 cm of sand; the filtration media was supported by a stainless steel screen

followed by a stainless steel mesh. The filters were operated at EBCT of 10 minutes. An overflow

was located at 90 cm from the top of the media to allow the filter to be operated in constant head

mode. Filter effluent storage tanks, with a volume of 57 L, were located after each filter and were

used to provide unchlorinated backwash water. Backwash was performed on each filter at the same

frequency as the large biofilters or whenever the flow could not be maintained. The backwash

sequence included a slow backwash with or without air scour for 4 minutes (30% bed fluidization)

followed by an 8 minutes of fast backwash at collapse pulsing conditions (50% bed fluidization)

and 2 minutes of slow backwash. Media sampling ports were located along the filter media depth

at 5, 45, 55, and 95 cm as measured from the bottom of the filter; also, a water sampling port was

located at 5 cm above the top of the media to obtain water samples if needed. A sketch and a photo

of the biofiltration set-up are shown in Figures 3.2 and 3.3, respectively.

25

Figure 3.2: Biofiltration schematic (BF= Biofilter, BW=Backwash and S= Sampling port) (Not

to scale)

26

Figure 3.3: Photo of the biofiltration set-up at the PWTP showing the large biofilters along with

the conventional filters (right) and the small biofilters (left)

The large-scale biofilters were used to compare biofiltration with or without nutrient

(phosphorus and nitrogen) enhancement to conventional treatment. The small biofilters were used

to explore the effect of media type (anthracite vs. GAC), peroxide enhancement, and direct

filtration (in-line alum coagulation prior to filtration) on the performance of biofiltration in terms

of LOH reduction, DBP precursor removal, UF fouling control, and other parameters measured in

this study (discussed later). The condition of each biofilter is presented in Table 3.2.

BF2 BF1 BF5 BF6

CF

Al

CF

PACl

BF4 BF3

27

Table 3.2: Biofilter conditions

Filter No. BF1

(large)

BF2

(large)

BF3

(small)

BF4

(small)

BF5

(small)

BF6

(small)

Condition Control,

comparing

biofiltration

to

conventional

treatment

Nutrient

enhancement

(phosphorus

and nitrogen)

Control,

comparing

large to

small

biofilters

H2O2

addition

Alum

addition (In-

line

coagulation)

GAC vs.

anthracite

as a filter

media

3.2.4 Sampling Locations

Samples were collected from the Peterborough pilot plant biofiltration and conventional

treatment trains as well as Peterborough full-scale Water Treatment Plant (PWTP) filter #11 (FSP).

The samples obtained were the following;

1- Raw water (RW),

2- Alum and PACl coagulated, flocculated, and settled water (Pilot pre-filter (Alum) (SW

Alum) and pilot pre-filter (PACl) (SW PACl), respectively),

3- Conventional filers effluent post Alum and PACl coagulation/flocculation/settling

treatment (post-filter (Alum) (CF Alum) and post-filter (PACl) (CF PACl), respectively),

4- The large biofilter effluents (BF1 (Large control) and BF2 (Nutrient enhancement)),

5- The small biofilter effluents (BF3 (small control), BF4 (peroxide addition), BF5 (in-line

alum), and BF6 (GAC)),

6- Full scale plant filter #11 (FSP).

Also, media samples from the pilot large and small biofilters (6 samples), conventional filters (2

samples) and FSP were collected and analyzed. Sampling locations in the pilot Peterborough

Water Treatment Plant are shown in Figure 3.4

28

Figure 3.4: Sketch showing the water sampling locations at the Peterborough pilot treatment

plant, ( = sampling point, not to scale)

3.2.5 Ultrafiltration Unit

Two automated ultrafiltration pilot systems were adapted for this research work; they used

a polyvinylidene fluoride (PVDF) hollow fiber UF, Zeeweed 1 series 500 (ZW-1) (GE, Oakville,

ON). The membrane nominal surface area was 0.047 m2, and it was operated at a flux of 30 LMH.

The membrane was immersed in a 2 L vessel and operated using outside-in dead-end mode. The

UF operation involved permeation for 30 minutes followed by backpulsing with air scour for 15

minutes; the membrane tank was then completely drained, re-filled, and placed back in operation.

The membrane module was cleaned after each run by soaking it in 750 mg/L sodium hypochlorite

(NaOCl) solution for 24-48 hours followed by permeating 750 mg/L NaOCl for 30 minutes. Clean

water recovery was then checked by permeating distilled water for 30 minutes. The membranes

29

were stored in 50 mg/L chlorine solution after cleaning. Experiments were conducted on a side by

side or a back to back basis. A sketch and a photo of the UF unit is shown in Figure 3.5.

Figure 3.5: UF unit schematic (a) and photo (b)

3.3 Experimental Design

3.3.1 Variables

Experimental variables which affect biofiltration performance are discussed in section

2.2.1. Most of these variables, including temperature and NOM concentration, are hard to control

especially in pilot- and full-scale experiments; therefore, many studies have focused on others such

as EBCT, backwash strategy and filter media type (Huck and Sozański, 2008). EBCT is usually

varied through controlling the filtration rate or the biofilter media depth. EBCT has ranged from a

few minutes to around an hour in some studies, but most researchers have used EBCT values

between 4 to 20 min (Hammes et al., 2011). On the other hand, varying the backwash procedure

is more complicated and might require a separate study to examine the long term effects on the

removal of foulants, and hydraulic operations (e.g. clogging, mud ball formation). In this study,

the EBCT was not varied since preliminary testing of lower EBCTs (5 min) caused a rapid increase

in terminal headloss (1.5 m in approx. 1 day) and increased the biofilter backwash frequency to at

a) b)

30

least once a day. Therefore, an EBCT of 11 and 10 minutes was used for the large and small filters,

respectively. For our study, we selected to vary the filter media type, C:N:P ratio, hydrogen

peroxide addition, and in-line coagulation dose.

3.3.2 Monitored Parameters

In this study, LC-OCD along with DOC and UV254 were selected as analytical techniques

to measure the organic matter in water and UF foulants. Membrane resistance (TMP/flow rate)

was also utilized for UF reversible and irreversible fouling quantification since it accounts for

variability in the membrane flux. For the purpose of DBP precursor monitoring, THM, HAA and

AOX formation potentials (after 24 hours at 20ºC and 1.5±0.5 mg/L chlorine residual) were

monitored. Additionally, turbidity and other general water parameters such as temperature and pH

were recorded.

The two large biofiltration and the two conventional treatment trains were equipped with

on-line water quality instruments and connected to the full-scale plant sophisticated supervisory

control and data acquisition (SCADA) system. These included pH (Hach DPD1P1, Mississauga,

ON), temperature (Siemens Milltronics Airanger DPL, Peterborough, ON), turbidity (Hach,

1720E, Mississauga, ON), UV absorbance at 254 (UV254) (Real Tech, UV254 M3000, Whitby,

ON), dissolved organic carbon (DOC) (General Electric, 5310 C On-Line TOC analyzer,

Colorado, USA), and differential headloss meters (Siemens, Sitrans P XP/DIP, Peterborough,

ON).

Media samples were analyzed for ATP concentration to determine the microbial activity

in each filter. Also, extracellular polymeric substances (EPS) (proteins and polysaccharides) were

measured to determine the status (stress level) of the microbial community and to determine their

correlation to headloss.

3.3.3 Sampling Time

At the beginning of the study (24th June- 12th August 2013) samples were obtained 4 hours

after the filter run. However, after obtaining on-line DOC and UV254 data from the large control

filter effluent (Figure 3.6), it was found that the biofilter requires a minimum of 24 hours to reach

steady DOC removal. Therefore, samples obtained prior to 12th August 2013 were obtained 4 hours

31

after the filter backwash, and samples obtained after 12th August 2013 were obtained on the 26

hours after the filter backwash.

Figure 3.6: DOC concentration and UV254 absorbance in the biofilter effluent after a backwash

(at 10:46 on July 26th 2013)

3.3.4 Media Type

Two filter media (anthracite and GAC) were used to evaluate the media impact on NOM,

DBP precursor, and UF foulant removal. Exhausted GAC (F300, Calgon Carbon) was obtained

from the Georgina water treatment plant, ON, and installed in BF6 (GAC biofilter) on June 4th

2013. It has been previously in service for 8 years and it was assumed to have no remaining

adsorptive capacity.

3.3.5 Chemical Addition

Chemical addition (phosphorus, nitrogen, peroxide, and alum) was divided into three

stages. The first stage started after the filters were deemed to have reached steady-state DOC

removal, and lasted for 69 days followed by 24 days and 28 days for the second and third stages,

respectively. Chemical dosing was changed after DOC removal had reached steady-state and

sampling was completed.

0.135

0.14

0.145

0.15

0.155

0.16

0.165

0.17

0.175

5.5

5.55

5.6

5.65

5.7

5.75

5.8

5.85

BF

1 E

fflu

ent

UV

25

4A

bs

(cm

-1)

BF

1 e

fflu

ent

DO

C c

once

ntr

atio

n

(mg/L

)

Series1 Series2Steady-state

DOC removal

DOC UV254

32

3.3.5.1 Nutrient Enhancement

A molar C:N:P ratio of 100:10:1 (21:5:1 by mass) has been shown to be optimal for

bacterial growth (Lauderdale et al. 2012); however, this ratio is usually not met in biofilter influent

(Liu et al. 2001). In raw Otonabee River water, nitrogen (as ammonium, NH4) and phosphorus

(orthophosphate (PO4)) concentrations were approximately 0.3 mg/L NH4-N and 0.00 mg/L PO4-

P, respectively. Since the DOC removal was approximately 0.5 mg/L before enhancement as

measured at the University of Toronto on June 24th, the C:N:P stoichiometric ratio was calculated

to be 100:30:0. This demonstrated that the water was only phosphorus limited, and that the

phosphorus addition was required to adjust the C:N:P ratio for optimal biomass activity.

Phosphorus (in the form of PO4) was as phosphoric acid (H2PO4) to the nutrient enhanced biofilter

influent; the required initial dose was calculated to be 0.025 mg/l PO4-P (C:N:P 100:30:2). DOC

removal was observed after 61 days. Then, ammonium was dosed in the form of ammonium

chloride at 0.1 mg/L NH4-N (C:N:P 100:40:2) for 24 days. The orthophosphate dose was then

increased to 0.25 mg/L PO4-P (C:N:P 100:40:20) for 28 days. Table 3.3 shows the nutrient

enhancement plan and Figure 3.7 shows a Gantt chart of the plan.

3.3.5.2 Hydrogen Peroxide Addition

Lauderdale et al. (2012) found that the addition of 1 mg/L hydrogen peroxide (H2O2)

decreased the terminal headloss by 60% when compared to a control, without affecting DOC

removal. Peroxide dosages were chosen to cover a similar range as reported in previous studies

(Lauderdale et al., 2012; Urfer and Huck, 1997). In this study, 0.1-1 mg/L H2O2 was added to BF4

(peroxide addition) influent as shown in Table 3.3 and Figure 3.7.

3.3.5.3 Alum Addition

Addition of 0.1-2 mg/L alum to the in-line coagulation biofilter influent was used

determine the optimum dose to reduce DOC, DBP precursor and UF fouling. An alum dose of

<0.5 mg/L was selected based on recent work done by Wray et al. (2014) where it was

demonstrated that 0.1 mg/L alum is the point of diminishing return for biopolymer removal. The

headloss was also monitored to determine the effect of the coagulant addition on the filter runtime.

The coagulant dose was varied as shown in Table 3.3 and Figure 3.7.

33

Table 3.3: Experimental design

Passive

Monitoring

Stage 1 (24-July-

13 to 30-Sept-13)

Stage 2 (30-Sept-

13 to 23-Oct-13)

Stage 3 (23-Oct-

13 to 21-Nov-13)

BF1 (control

large) Used as a control for the large filters (installed: October 2012)

BF2 (nutrient

Enhancement)

C:N:P

100:30:0

C:N:P

100:30:2

C:N:P

100:40:2

C:N:P

100:40:20

BF3 (control

small) Used as a control for the small filters (installed: February 2013)

BF4 (peroxide

addition) 0 mg/L 0.1 mg/L 0.5 mg/L 1 mg/L

BF5 (alum

addition) 0 mg/L 0.1 mg/L 0.5 mg/L 0.25 mg/L

BF6 (GAC) GAC column (installed: June 2013)

Figure 3.7: Chemical addition Gantt chart

43

3.4 Analytical Methods

3.4.1 Dissolved Organic Carbon

Dissolved organic carbon (DOC) was measured using the wet oxidation method based on

Standard Method 5310 D (APHA, 2012). The analysis was carried out using O-I Corporation

Model 1010 Analytical TOC Analyzer with a Model 1051 Vial Multi-Sampler. The instrument

conditions are shown in Table 3.4. Water samples were filtered using a 0.45 µm fiber glass filter,

transferred to 40 mL amber vials, and capped with Teflon®-lined septum screw caps. Samples

were stored at 4ºC and tested within 7 days of collection. DOC concentrations in water samples

were quantified using anhydrous potassium hydrogen phthalate (KHP) in Milli-Q® water

calibration solution, which was run with each sample set (Concentration= 0, 1.25, 2.5, 5, and 10

mg/L). A new calibration curve was prepared before each set of samples. Check standards (C= 3

mg/L) were run after every 10 samples, and at the end of a sample set. At least 3 blank samples

(Milli-Q® water) or a blank every tenth sample were run regularly. The reagent list and the method

outline are listed in Tables 3.5 and 3.6, respectively.

Table 3.4: DOC analyzer conditions

Parameter Description

Acid volume 500 μL of 5% phosphoric acid

Oxidant volume 1000 μL of 10% sodium persulphate

Sample volume 2 mL

Rinses per sample 1

Volume per rinse 15 mL

Replicates per sample 3

Reaction time (min:sec) TIC 01:30; TOC 02:00

Detection time (min:sec) TIC 00:00; TOC 03:00

Purge gas Nitrogen

Loop size 10 mL

35

Table 3.5: DOC analysis reagents

Reagent Source

Milli-Q® water Prepared in the laboratory

Sulphuric acid, H2SO4 Anachemia, 98+%

Sodium persulphate, Na2(SO4) Sigma Aldrich, 98+%, anhydrous

Potassium hydrogen phthalate (KHP),

C8H5KO4

Sigma Aldrich, 98+%

Phosphoric acid, H3PO4 Caledon, >85%

Nitrogen gas, N2 Praxair, Ultra high purity (UHP)

Table 3.6: DOC method outline

Blanks:

Use 40 mL of Milli-Q® water.

Stock solution:

Add 2.13 g potassium hydrogen phthalate in 1 L Milli-Q® water and acidify at pH <2

with H2SO4. Store in the fridge at 4°C

Check Standard (3.0 mg/L):

Add 300 μL of stock solution to 100 mL Milli-Q® water.

Analyze using TOC analyzer.

3.4.2 UV254

The ultraviolet absorbance at 254 nm (UV254) was determined via a CE 3055 Single Beam

Cecil UV/Visible Spectrophotometer (Cambridge, England) using 1 cm quartz cells (Hewlett

Packard, Mississauga). The spectrophotometer was zeroed with Milli-Q® water. Quartz cells were

rinsed with Milli-Q® and the water sample between measurements.

36

3.4.3 THMs

Trihalomethane (THM) (chloroform (trichloromethane, TCM), bromodichloromethane

(BDCM), dibromochloromethane (DBCM), and bromoform (tribromomethane, TBM)) analysis

was conducted using a liquid-liquid extraction gas chromatographic method based on Standard

Method 6232 B (APHA, 2012). The analysis took place at the University of Toronto laboratory

(Toronto, ON) using a Hewlett Packard 5890 Series II Plus Gas Chromatograph (Mississauga,

ON) equipped with an electron capture detector (GC-ECD) and a DB 5.625 capillary column

(Agilent Technologies Canada Inc., Mississauga, ON). The instrument conditions, required

reagents and method outline are described in Tables 3.7, 3.8, and 3.9, respectively. The

concentration of THM stock solution was 2000 μg/L. The concentrations of THM species in the

intermediate solution were 20 mg/L. Calibration standards were prepared at concentrations of 0,

5, 10, 20, 40, 60, 100, and 150 μg/L (in some cases half of these concentrations were used (i.e. 0,

2.5, 5, 10, 20, 30, 50, and 75 μg/L). The MDL for each species is shown in Table 3.10. MDLs were

determined by multiplying the standard deviation of 8 replicates, prepared in the same order of

magnitude as the expected MDL, by the Student-t value (3.0). A chloroform contaminated Methyl-

tert-butyl-ether (MTBE) was used to prepare the calibration curve and samples on September 24th,

2013. The contamination was accounted for by subtracting the blank from all samples and

standards.

Table 3.7: THM instrument conditions

Parameter Description

Injector Temperature 200°C

Detector Temperature 300°C

Temperature program 40°C for 4.0 min

4°C/min temperature ramp to 95°C

60°C/min temperature ramp to 200°C

Carrier Gas Helium

Make-up Gas P5 Argon (5% Methane)

Flow Rate 1.2 mL/min at 35°C

37

Table 3.8: THM reagents

Reagent Source

Milli-Q® water Prepared in the laboratory

Trihalomethane concentrated stock for

calibration

Supleco, 2000 μg/mL in methanol (48140-

U)

1,2- dibromopropane concentrated stock Supleco, 100 mg/L in Methyl-tert-butyl-

ether

Sodium sulphate [Na2SO4] Sigma Aldrich, ACS Grade

Methyl-tert -butyl-ether (MTBE) J.T. Baker ultra-resin-analyzed

Table 3.9: THM method outline

Collect samples in 500 mL amber bottles quenched with 0.025 g of ascorbic acid (100

mg/mL).

Store samples in the dark at 4°C for up to 7 days.

To begin preparing samples, remove from refrigerator and bring to room temperature.

Blanks:

Transfer 20 mL of Milli-Q® water into 40 mL vials and process alongside samples.

Standard Solution:

20 mg/L THM solution prepared from 2000 μg/mL stock.

Prepare calibration standards by adding the appropriate amount to get standard

concentrations of (0, 5, 10, 20, 40, 60, 100 and 150 μg/L)

** Wipe the syringe tip with a Kimwipe before measuring out the THM stock and before

adding stock to solution.

38

Table 3.9: THM method outline (cont.)

Check Standards: (20 μg/L):

Add 20 μL of calibration solution to 20 mL of Milli-Q® water in a 40 mL vial and

process alongside samples.

Include blanks and running standards every 10 samples.

Extraction:

Transfer 20 ml of each sample vial into a clean 40 mL vial.

Add 50 μL of the 1,2-dibromopropane solution

Add 2 scoops of sodium sulphate (Na2SO4) in order to increase extraction efficiency.

Add 4 mL of MTBE extraction solvent and cap with Teflon®-lined silicon septum and

screw cap.

Shake sample vial vigorously for approx. 30 seconds and place on counter on its side.

Repeat and complete for all samples, blanks and standards before proceeding.

Shake the samples by hand for 2 minutes.

Let samples stand for 15-30 minutes for phase separation.

Extract 2 mL from the organic layer using a Pasteur pipette and place in a 1.8 mL GC

vial filled with 2 small scoops of Na2SO4 (there should not be any water in the vial). Use

a clean pipette for each sample. Fill the vial to the top and cap immediately, ensuring that

there is no headspace. To ensure only the MTBE layer was taken, freeze the samples and

examine the vials after more than two hours to see if there is only one phase visible.

If not analyzing immediately, store the samples in the freezer (-11°C) for up to 21 days.

Analyze using a GC-ECD.

39

Table 3.10: THM method detection limits

Analyte Standard Deviation (μg/L) MDL (μg/L)

TCM 0.23 0.69

BDCM 0.05 0.16

DBCM 0.09 0.26

TBM 0.18 0.54

3.4.4 HAAs

Haloacetic acids (HAAs) (monochloroacetic acid (MCAA), monobromoacetic acid

(MBAA), dichloroacetic acid (DCAA), trichloroacetic acid (TCAA), bromochloroacetic acid

(BCAA), dibromoacetic acid (DBAA), bromodichloroacetic acid (BDCAA), dibromochloroacetic

acid (DBCAA), and tribromoacetic acid (TBAA)) analysis was conducted using a liquid-liquid

extraction gas chromatographic method based on Standard Method 6251 B (APHA, 2012). All

analyses were conducted at the University of Toronto laboratory facility (Toronto, ON) using a

Hewlett Packard 5890 Series II Plus Gas Chromatograph (Mississauga, ON) equipped with an

electron capture detector (GC-ECD) and a DB 5.625 capillary column (Agilent Technologies

Canada Inc., Mississauga, ON). The instrument conditions, required reagents method outline and

method MDLs are described in Tables 3.11, 3.12, 3.13, and 3.14, respectively. Few samples (BF4,

BF5, and BF5 duplicate) from September 24th were analyzed without adding an internal standard.

Table 3.11: HAA instrument conditions

Parameter Description

Injector Temperature 200°C

Detector Temperature 300°C

Temperature program 35°C for 10.0 min

2.5°C/min temperature ramp to 65°C

10°C/min temperature ramp to 85°C

20°C/min temperature ramp to 205°C, hold for 7 min

Carrier Gas Helium

Make-up Gas P5 Argon (5% Methane)

Flow Rate 1.2 mL/min at 35°C

40

Table 3.12: HAA reagents

Reagent Source

N-methyl-N-nitroso-p-toluene sulfonamide

(Diazald) [CH3C6H4SO2N(CH3)NO]

Sigma Aldrich, 99+%

Sodium hydroxide [NaOH] BDH, 85.0+%, ACD Grade

Sulphuric acid [H2SO4] Anachemia, 98+%

Haloacetic acids concentrated stock EPA 552.2 Acids Calibration Mix in

MTBE

2,3,5,6 tetrafluorobenzoic acid (TFBA)

concentrated stock

Supleco, 2000 mg/L in MTBE

Sodium sulphate [Na2SO4] Sigma Aldrich, ACS Grade

Methyl-tert -butyl-ether (MTBE) Sigma Aldrich, >99.8% Or J.T. Baker

ultra-resin-analyzed

Table 3.13: HAA method outline

Collect samples in 500 mL amber bottles quenched with 0.025 g of ascorbic acid (100

mg/mL). Ensure that samples are headspace-free.

Store samples in the dark at 4°C for up to 9 days.

To begin preparing samples, remove from refrigerator and bring to room temperature.

Blanks:

Transfer 20 mL of Milli-Q® water into 40 mL vials and process alongside samples.

Standard Solution (10 μg/mL):

Collect 20 mL of MilliQ® in 40 ml Vial

Using a 50 μL syringe and the sandwich technique, add 0, 5, 10, 20, 40, and 60 μL of

HAA stock (2000 μg/mL each)

41

Table 3.13: HAA method outline (cont.)

** Wipe the syringe tip with a Kimwipe before measuring out the HAA stock and before

adding stock.

Check Standards (40 ug/L):

Add 20 μL of working solution (10 mg/mL) to 20 mL of Milli-Q® water, process alongside

samples.

Include blanks and running standards every 10 samples.

Diazomethane Generation

Set up the generation apparatus as shown in Figure 6521:2 in Standard Methods (APHA,

2012).

Extraction

Transfer 20 ml the contents of each sample vial into a clean 40 mL vial.

Add 20 μL of the 2,3,5,6-TFBA solution

Add 3 mL of sulphuric acid (H2SO4) to reduce the pH of the sample.

Add 2 two scoop of sodium sulphate (Na2SO4) in order to increase extraction efficiency.

Add 5 mL of MTBE extraction solvent and cap with Teflon®-lined silicon septa and screw

cap.

Shake sample vial vigorously for approx. 30 seconds and place on counter on its side.

Complete this procedure for all samples, blanks and standards before proceeding.

Shake the samples by hand for 5 minutes. Let samples stand for 15-30 minutes for phase

separation.

42

Table 3.13: HAA method outline (cont.)

Transfer MTBE to GC vials to the just above 1.5 ml mark to allow for addition of 150 μL

diazomethane cap immediately, ensuring that there is no headspace. Use a clean pipette for

each sample.

To ensure only the MTBE layer was taken, freeze the samples and examine the vials after

more than two hours to see if there is only one phase visible.

Add 150 μL of diazomethane to the GC vial (submerge tip before injection) and cap

immediately.

If not analyzing immediately, store the samples in the freezer (-11°C) for up to 21 days.

Analyze using a GC-ECD.

Table 3.14: HAA method detection limits

Analyte Standard Deviation (μg/L) MDL (μg/L)

MCAA 0.34 1.03

MBAA 0.47 1.41

DCAA 0.40 1.20

TCAA 0.35 1.06

BCAA 0.39 1.16

DBAA 0.27 0.80

BDCAA 0.19 0.57

DBCAA 0.15 0.46

TBAA 0.15 0.46

3.4.5 AOX

Adsorbable Organic Halogens (AOX) analysis was conducted using a titration method

based on Standard Method 5320 (APHA, 2012). All analyses were conducted at the University of

Toronto laboratory (Toronto, ON) using a Trace Element Instruments Xplorer organic halogens

analyzer (Delft, Netherlands). Instrument conditions, reagents and method are shown in tables

43

3.15, 3.16, and 3.17. Samples were run in duplicates, and a check standard (100 μg/L) was injected

into the test cell before each analysis.

Table 3.15: AOX instrument conditions

Parameter Description

Combustion cell temperature 1000°C

Combustions gas Oxygen

Carrier gas to combustion cell Argon

Carrier gas to titration cell Oxygen

Cell scrubber Concentrated sulfuric acid 98%

Table 3.16: AOX reagents

Reagent Source

Nitrate stock solution, 1% 14 mL of nitrite stock 67% in 1 L of MQ

mixed with 14 g of Sodium Nitrate

Nitrate wash solution 50 mL of Nitrate stock solution in 1 L of MQ

Acetic Acid Sigma Aldrich, 75%

Sulphuric acid [H2SO4] Anachemia, 98+%

Standard Solution (4-Chlorophenol) Trace Elements, 200 mg/L Cl

Sodium Chloride Trace Elements, 2 mmol/L

Table 3.17: AOX method outline

Collect samples in 500 mL amber bottles. Ensure that samples are headspace-free.

Store samples in the dark at 4°C for up to 14 days.

To begin preparing samples, remove from refrigerator and bring to room temperature.

44

Table 3.17: AOX method outline (cont.)

Sample extraction:

Prepare the extraction column set by connecting two activated carbon columns with the

connector. Make sure to puncture the column cap to allow sample pass

through.

Assemble the column set onto the Xprep sample filtration unit.

Shake and homogenize samples, pour 100 mL into the extraction syringes.

Check the wash solution of sodium nitrate/nitric acid. Start the extraction. It takes about

50 minutes for the whole extraction depending on the flow rate (1-2 mL/min).

Transferring of activated carbon to the test cups:

At the end of sample extraction, take off the column sets from the Xprep.

Using the ejecting tool to transfer the activated carbon to clean test cups and place on the

sample auto sequencer and run them immediately

3.4.6 ATP Analysis

Luminaltra adenosine triphosphate (ATP) analysis kit (DSA-100C, Fredericton, NB) was

used to carry out the ATP analysis following manufacturer instructions as shown in Table 3.18.

Media samples were obtained from the top of the biofilter or conventional filter at the end of each

enhancement stage.

Table 3.18: ATP method outline

Collect in 10 mL Falcon tubes and store at 4°C and analyze within 24 hours.

To calibrate the Luminometer, add two drops (100 μL) of enzyme reagent (Luminase)

and two drops (100 μL) of tATP standard (Ultracheck 1) in a 12×55 mm test tube

45

Table 3.18: ATP method outline (cont.)

Measure the relative light units (RLU) using a Luminometer (RLUstandard).

Weight 1 g of sample media and transfer to 17×100 mm test tube

Using a micropipette, add 5 mL of (total ATP) tATP extraction reagent (UltralyseTM 7)

and shake rigorously

Incubate the sample for 5 minutes

Transfer 1 mL of the incubated sample to a new 17×100 mm test tube and add 9 mL of

tATP dilution reagent (UltraLute). Cap the tube and shake gently

Using a micropipette, transfer 100 μL from the dilution tube to a new 12×55 mm test tube

and add 2 drops (100 μL) of Luminase

Mix gently and place in the Luminometer to measure the RLU (RLUsample).

The ATP concentration is determined as follows:

• tATP (pg ATP/g) on the filter media=RLUSample

RLUStandard×

50000 (pg ATP)

masssample (g)

• tATP (pg ATP/mL) on the filter media=RLUSample

RLUStandard×

50000 (pg ATP)

volumesample (mL)

3.4.7 EPS Analysis

Extracellular polymeric substances (EPS) analysis was conducted on biofilter and

conventional filter media obtained from the top 5 cm. Protein and polysaccharide analyses were

carried out at the University of Toronto laboratory (Toronto, ON) using a method adapted from

Papineau et al. (2012) and DuBois et al. (1956), respectively. The samples were extracted using a

Tris EDTA extraction method as described by Liu and Fang (2002). A fresh calibration curve was

prepared for each analysis; a serial dilution curves (0, 3.125, 6.25, 12.5, 25, 50, and 100 mg/L) of

glucose and Bovine Serum Albumin (BSA) solution were used for polysaccharides and proteins;

46

respectively. The reagent used and the method outline are presented in Tables 3.19 and 3.20,

respectively. CE 3055 Single Beam Cecil UV/Visible Spectrophotometer (Cambridge, UK) and

Hach spectrophotometer (Hach Odyssey DR/2500 Scanning Spectrophotometer, Mississauga,

ON) were used for protein and polysaccharide analysis, respectively.

Table 3.19: EPS reagents

Reagent Source

Ethylenediaminetetraacetic acid (EDTA)

disodium salt dehydrate

Sigma Aldrich, ACS grade

Tris (Hydrohymethyl) Aminomethane,

BioUltrapure

BioShop, > 99.9 %

Bovine Serum Albumin (BSA) solution BioLabs, 10 mg/mL

D-Glucose Anhydrous, reagent grade

Pierce™ BCA Reagent A + B Thermo Scientific

Phenol Alfa Aesar, ACS grade

Sulphuric Acid Anachemia, 98+ %

Table 3.20: EPS method outline

Collect samples in 10 mL Falcon Tubes.

Store samples in the fridge at 4°C for < 48 hours.

To begin preparing samples, remove from the fridge and bring to room temperature.

Sample extraction:

Prepare Tris-EDTA buffer from concentrate in a 50 mL tube using a 1:9 ratio of 10 times

concentrated Tris-EDTA to Milli-Q® water

Add a magnetic stirrer and place in stirrer for 30 min or until reagents are fully dissolved

Sterilize in the autoclave at 121°C for 20 min and keep at room temperature

47

Table 3.20: EPS method outline (cont.)

Extraction procedure:

Incubate samples at 4°C in an environmental chamber for 4 hours on a shaker

Centrifuge samples at 12,000 g at 4°C for 15 minutes

Pipet the supernatant while avoiding the media residue

Filter samples into sterile Falcon tubes using a 0.45 μm filter syringe

If not analyzed immediately, store in freezer at -11°C.

Protein Analysis:

In a fume hood, place hot water bath in the fume hood and heat to 60°C

Prepare Pierce™ BCA working reagent by adding 50 parts reagent A to 1 part reagent B,

Calculate the volume of working reagent required by multiplying the number of samples by

1.5 (since 1.5 mL is needed per sample), including blanks, standard curve, and duplicates

Working Solutions (1 mg/mL):

Add 100 μL of the BCA standard to 900 μL of MilliQ®

Analysis Procedure:

Add 400 μL of working solution to 3600 μL of MilliQ® to prepare 100 μg/mL standard

Transfer 150 μL of the 100 μg/mL standard to a microcentrifuge tube

Add 75 μL of each concentration of the BCA standards into the next lower concentration in

microcentrifuge tubes (filled with 75 μL of MilliQ®) to prepare the serial dilution standards

Add 75 μL of sample into microcentrifuge tubes (duplicates for each sample)

Add 75 μL of Milli-Q® water into a microcentrifuge tube for the blank

48

Table 3.20: EPS method outline (cont.)

In the fume hood, add 1.5 μL of Peirce™ BCA working reagent to the blank, the BSA serial

dilution, and the samples

Incubate in water bath at 60°C for 30 min

Let cool to room temperature and read absorbance at 562 nm (OD562) with the

spectrophotometer (CE 3055 Single Beam Cecil UV/Visible Spectrophotometer).

Polysaccharides Analysis:

In a fume hood, place hot water bath in the fume hood and heat to 100°C

Standard and Working Solutions:

In a 50 mL graduated cylinder, add 2.5 g of glucose anhydrides and fill it to the 50 mL mark

with MilliQ® to prepare 50 mg/mL standard solution

In a 50 mL graduated cylinder, add 1 mL of glucose standard and fill it to the 50 mL mark

with MilliQ® to prepare 1 mg/mL working solution

Analysis Procedure:

Add 200 μL the glucose working solution (1 mg/mL) and 1800 μL of MilliQ® to a glass

vial to prepare 2 mL of 100 μg/mL standard

Add 1 mL of of each concentration of the standard into the next lower concentration in a

glass vial (already filled with 1 mL of MilliQ®) to prepare the serial dilution standards, and

cap using acid-resistant lids

Add 1 mL of sample (in triplicates) including a blank (Milli-Q®) into the glass vials

Add 1 mL of 5 % phenol solution into each vial

49

Table 3.20: EPS method outline (cont.)

In a fume hood, add 5 mL of 98 % sulfuric acid into each vial using the pump. Close the caps

but not tightly

Vortex samples (use a low vortex speed setting, do not let liquid reach tops of vials)

Place vials in water bath (100°C) for 5 minutes

Remove rack from water bath and vortex each sample

Cover racks in aluminum foil and incubate in the dark, at room temperature, for 10 minutes.

Vortex the samples gently

Re-cover the racks in foil and incubate in the dark at room temperature for additional 30

minutes and afterwards vortex the samples gently

Read absorbance at 492 nm (OD492) using the Hach spectrophotometer (Hach Odyssey

DR/2500 Scanning Spectrophotometer), vial labels aligned with arrow.

3.4.8 Nitrogen and Phosphorus

Nitrogen in the form of ammonia (NH3-N) and phosphorus in the form of orthophosphate

(PO4-P) were measured using Hach® Ammonia Reagent Set Salicylate Method (10 mL), and

PhosVer® 3 Phosphate Reagent Powder Pillows (10ml), respectively following the manufacturer

instructions. The analysis was conducted using Hach® Odyssey DR/2500 Scanning

Spectrophotometer (Mississauga, ON).

3.4.9 Chlorine (Cl2) Demands and Residuals

DBPs were formed by chlorinating raw water and biofilter effluents. The 24 hour chlorine

(Cl2) demands were estimated by dosing the conventional treatment, and biofilter samples with 5

and 7 mg/L Cl2, respectively, in 125 ml chlorine-demand-free amber bottles and measuring the

free chlorine residual after 24 hours as described in Standard Method 4500-CI G (APHA, 2012).

The free chlorine dosage was calculated to provide a residual of 1.5±0.5 mg/L after 24 hours.

50

Samples were chlorinated in 500 ml amber bottles with no head space, incubated at 20±2˚C for 24

hours, after which the chlorine residuals were measured and were quenched using ascorbic acid

(100 mg/L).

3.5 Quality Assurance/ Quality Control (QA/QC)

Water samples were stored in amber glass bottles. Samples were preserved at 4˚C and were

processed within 48 hours from the sampling period. Standards for pH, turbidity, DOC, UV254,

THMs, and HAAs were run regularly to ensure the instruments were well calibrated. LC-OCD

was calibrated at least once every six months at the University of Waterloo. DOC, UV254, turbidity,

pH, THMs, HAAs, and AOX were done in duplicates.

Media samples were collected and stored in clear falcon tubes. ATP analysis was

performed either on-site after sample collection or within 8 hours of sample collection. EPS

samples were stored at 4˚C and were analyzed within 2 days of collection. ATP samples were not

replicated. Polysaccharides and protein samples were run in duplicates and triplicates,

respectively.

Check standards were run regularly during analysis of THMs, HAAs, and AOX to establish

quality control. They were prepared to be in the expected sample concentration range and run

every tenth sample. Also, method and instrument blanks were analysed after every tenth sample.

A new calibration curve was prepared with each batch of DOC, and EPS samples. UV254 and

turbidity were zeroed with MilliQ® prior to analysis. Quality control charts were created using the

check standards, which were used to ensure the method’s accuracy and performance as per

Standard Method 1020 (APHA, 2012). The method was recalibrated if any of the following trends

were observed:

- 2 consecutive measurements outside the control limits of Mean ± 3×standard deviation(s)

(upper control limit (UCL) and lower control limit (LCL));

- 3 out of 4 consecutive measurements were outside of mean ± 2×standard deviation(s)

(upper warning limit (UWL) and lower warning limit (LWL));

- 5 out of 6 consecutive measurements were outside of mean ± standard deviation(s);

- 5 out of 6 consecutive measurements were following a trend of increasing or decreasing;

51

- 7 consecutive measurements were > the mean, or 7 consecutive measurements were < the

mean.

The mean and standard deviation were calculated by analyzing 7 or 8 individually prepared

standard standards in sequence along with a new calibration curve. Calibration curves, raw data,

and QA/QC charts are shown in Appendix A, B, and C, respectively.

3.6 Statistical Analysis

A student t-test (Equation 3.1) was used to evaluate the impact of nutrient, peroxide, and

alum addition on biofiltration ability to remove turbidity and UV254 when compared to the control.

Also, a paired t-test (Equation 3.2) was used to determine the impact of these enhancement

strategies on DOC removal and headloss compared to the control. All statistical analyses were

performed at the 95% confidence level. Also, a paired student t-test was used to compare

biofiltration (large control (BF1)) to conventional treatment (alum and PACl coagulation/

flocculation/sedimentation with/without filtration)

If 𝑡 =𝑀𝑠𝑎𝑚𝑝𝑙𝑒−𝑀𝑐𝑜𝑛𝑡𝑜𝑙

𝑆𝐷𝑠𝑎𝑚𝑝𝑙𝑒,𝑐𝑜𝑛𝑡𝑟𝑜𝑙/√2

𝑛

> |𝑡𝛼

2,𝑛−1| (3.1)

Then the difference in the historical mean of the analyte (e.g. DOC) is statistically significant

between two biofilters (or treatment options).

Where t is the t-test value, 𝑀 is the mean of each sample set, 𝑆𝐷𝑠𝑎𝑚𝑝𝑙𝑒,𝑐𝑜𝑛𝑡𝑟𝑜𝑙 is the grand standard

deviation of the two sample sets (i.e. 𝑆𝐷𝑠𝑎𝑚𝑝𝑙𝑒,𝑐𝑜𝑛𝑡𝑟𝑜𝑙 = √1

2(𝑆𝐷𝑠𝑎𝑚𝑝𝑙𝑒

2 + 𝑆𝐷𝑐𝑜𝑛𝑡𝑟𝑜𝑙2) ), and n is

the number of observations (samples).

If 𝑡 =∑ 𝐷

√𝑛(∑ 𝐷2)−(∑ 𝐷)2

𝑛−1

> |𝑡𝛼

2,𝑛−1| (3.2)

Then the difference in the historical mean of the analyte (e.g. DOC) is statistically significant

between two biofilters (or treatment options).

52

Where t is the paired t-test value, ∑ 𝐷 is the sum of difference between pairs, (∑ 𝐷2) is the sum of

the square difference between pairs, (∑ 𝐷)2 is the square of the sum of difference between pairs,

and n is the number of observations (pairs).

53

4. Engineered Biofiltration for NOM, DBP Precursor

Removal, and UF Fouling Control

4.1 Introduction

Membrane filtration has emerged as a cost effective solution to remove emerging

contaminants and to meet potential future regulations (Plakas and Karabelas, 2012; Rana et al.,

2012; Singh, 2006). However, fouling remains as one of the major limitations when considering

the implementation of membranes (Howe and Clark, 2002). Membrane fouling can hinder

treatment efficiency, damage the membrane surface, and significantly increase maintenance and

operation costs (Singh, 2006). Natural organic matter (NOM) has been identified as the major

foulant (Wang and Wang, 2006), and recent studies have shown that biopolymer and colloid

organic fractions are the major contributors to irreversible organic fouling (Hallè, 2010; Neubrand

et al., 2010; Zheng et al., 2010; Zheng et al., 2012). While NOM itself does not present a direct

health concern (Hozalski et al., 1999), it can influence overall water quality by reacting with

chlorine and other disinfectants to form disinfection by-products (DBPs) (Chaiket et al., 2002;

Richardson et al., 2007).

Biofiltration has been presented as a cost effective and chemical-free process for the

removal of NOM (Huck and Sozański, 2008). It has been reported to be effective for DBP

precursor removal, for limiting microbial growth in distribution systems (Van der kooij, 1992;

Wert et al., 2008; Yang et al., 2011), for controlling taste and odour (Elhadi et al., 2006; Srinivasan

and Sorial, 2011), as well as membrane fouling control (Hallé et al., 2009; Huck et al., 2011;

Peldszus et al., 2011). Biologically active granular media filters operated without chemical

disinfection (e.g. chlorinated backwash) can promote microorganism growth on filter media (Zhu

et al., 2010). The biological community can consume NOM. Recent research has focused on

actively improving biofiltration by way of nutrient enhancement, where adjusting the carbon to

phosphorus ratio (C:P) to 100:2 was found to increase dissolved organic carbon (DOC) removal

by 75% and decrease headloss by 15% (Lauderdale et al., 2012). Hydrogen peroxide (H2O2)

addition was reported to decrease headloss by up to 75% without impacting DOC removal

(Lauderdale et al., 2012; Urfer and Huck, 2000).

54

Granular activated carbon (GAC) and anthracite has been intensively studied as a biofilter

media (Dussert and Tramposch, 1997; Emelko et al., 2006; Liu et al., 2001). Most of the literature

suggests that exhausted GAC and anthracite biofilters provide similar TOC and organic compound

(e.g. acetate, formate) when temperatures are between 20 to 25˚C (Emelko et al., 2006; Liu et al.,

2001). It was reported that that GAC provides higher total organic carbon (TOC) and assimilable

organic carbon (AOC) removals (up to 100% more than anthracite) at low temperatures (1-10˚C)

(Dussert and Tramposch, 1997; Emelko et al., 2006). Many researchers have hypothesized that the

superior performance of GAC is due to its high surface area (>1000 m2/g) and porous nature (pores

volume >0.15 cm3/g) where particles attachment and growth of microorganisms is more favourable

(Dussert and Tramposch, 1997; Klymenko et al., 2010). Others have attributed the superior

performance to bio-regeneration, since microorganisms consume NOM attached to GAC which in

turns reactivate adsorption sites (Velten et al., 2011a; Zhu et al., 2010).

The objective of this pilot-scale study was to investigate the impact of biofiltration

enhancement strategies and filter media types on NOM removal, DBP precursor reduction, and

mitigation of ultrafiltration (UF) fouling. Six pilot-scale biofiltration trains were designed to

examine the impact of stepwise addition of nutrients (phosphorous and nitrogen), peroxide, and

in-line alum, as well as the biofilter media type (GAC vs. anthracite).

4.2 Materials and Methods

4.2.1 Source Water Quality Parameters

The source water for this study was Otonabee River (Peterborough, Canada). When

influent water temperatures were > 12˚C, the water was pre-chlorinated for zebra mussel control.

Residual chlorine was quenched prior to biofiltration by applying sodium thiosulfate (3.5:1 sodium

thiosulfate to chlorine ratio by mass). Typical water quality parameters are shown in Table 4.1.

Table 4.1: Otonabee River raw water quality parameters

Parameter (units) Yearly range

Temperature (˚C) 3-29

pH 7-9

Turbidity (NTU) 0.3-1.5

DOC (mg/L) 5.8-7.3

55

4.2.2 Pilot Configuration

The pilot consisted of parallel biofiltration treatment trains incorporating two large glass

biofilters (Diameter (D)= 15.24 cm) and four small acrylic biofilters (D= 7.62 cm). The large and

small biofilters were operated at empty bed contact times (EBCTs) of 11 and 10 minutes,

respectively. They contained 50 cm of anthracite (effective size (d10)= 0.85, uniformity coefficient

(UC)= 1.8) over 50 cm of sand (d10= 0.5, UC= 1.8). A wedge wire plate was installed below the

sand layer to support the media for the large filters, and a stainless steel mesh was used in the small

filters. Exhausted GAC (F300, Calgon Carbon) was obtained from the Georgina Water Treatment

Plant, Ontario, Canada. It has been previously in service for 8 years and adsorptive capacity was

assumed to be exhausted. The water level above each filter was maintained at 90 cm from the filter

media bed (constant head mode). Each filter was backwashed using its own effluent three times

per week (Monday, Wednesday, and Friday), or as needed to maintain the EBCT. The backwash

procedure included a slow backwash with air scour for 2 minutes (30% bed fluidization), followed

by 8 minutes of fast backwash at collapse pulsing conditions (50% bed fluidization), and a slow

backwash for 4 minutes. A general schematic of the pilot plant configuration is depicted in Figure

4.1.

4.2.3 Ultrafiltration Units

Two automated ultrafiltration pilot systems were adapted for this research work. They

consisted of a polyvinylidene fluoride (PVDF) hollow fiber UF membrane, Zeeweed 1 series 500

(ZW-1) (GE, Oakville, ON) where the membrane nominal surface area was 0.047 m2. The units

were operated at a flux of 30 LMH. The membrane was immersed in a 2 L vessel and operated

using outside-in dead-end mode. UF operation consisted of 30 minutes permeation followed by

backpulsing with air scour for 15 minutes; after which the membrane tanks were completely

drained, re-filled, and placed back in operation. The membrane module was cleaned after each UF

run by soaking it in 750 mg/L sodium hypochlorite (NaOCl) solution for 24-48 hours, followed

by permeation with 750 mg/L NaOCl for 30 minutes and distilled water for 30 minutes. The

membranes were stored in 50 mg/L chlorine solution between runs. Experiments were conducted

in parallel or back to back basis as required.

56

Figure 4.1: Biofiltration pilot plant schematic (BF= Biofilter, BW=Backwash and S= Sampling

port) (Not to scale)

57

4.2.4 Experimental Design

Chemical addition (phosphorus, nitrogen, peroxide, and alum) was divided into three

stages. The first stage was initiated once the filters were deemed to have reached steady-state DOC

removal rates and lasted for 69 days. The second was 24 days and third was 28 days. Chemical

dosing conditions were changed once DOC removal had reached steady-state and sampling was

completed. Hydrogen peroxide (H2O2) was supplemented to provide an influent concentration of

0.1, 0.5 and 1 mg/L for the first, second, and third experimental stages. Peroxide dosages were

chosen to cover a similar range as reported in previous studies (Lauderdale et al., 2012; Urfer and

Huck, 1997). Alum was added directly to the biofilter influent at a concentration of 0.1, 0.5 and

0.25 mg/L, respectively, for the three consecutive stages. An alum dose of < 0.5 mg/L (0.0225 mg

Al3+/L) was selected based on recent work done by Wray et al. (2014) where 0.1 mg/L alum (0.005

mg Al3+/L) was shown to be the point of diminishing return for biopolymer removal. The ambient

raw water carbon:nitrogen:phosphorus stoichiometric ratio (C:N:P) was 100:30:0. Nutrient

addition (nitrogen and phosphorus) was achieved by supplementing phosphoric acid (H2PO4) and

ammonium chloride (NH4Cl) to the biofilter influent. As such, phosphorus was dosed at 0.025

mg/L PO4-P for the first and second stages and at 0.25 mg/L PO4-P for the third to adjust the C:P

ratio to 100:2 and 100:20, respectively. The nitrogen dose was 0 mg/L NH4-N for the first stage

and 0.1 mg/L for the second and third stages in order to adjust the carbon to nitrogen (C:N) ratio

to 100:30 and 100:40, respectively. One large and small biofilters were used as controls (Table

4.2).

Table 4.2: Expermintal design for biofiltration pilot

Filter/Stage Passive

Monitoring

Stage 1 (24-July-

13 to 30-Sept-13)

Stage 2 (30-Sept-

13 to 23-Oct-13)

Stage 3 (23-Oct-

13 to 21-Nov-13)

Control large

(BF1) Used as a control for the large filters (installed: October 2012)

Nutrient

enhancement

(BF2)

C:N:P

100:30:0

C:N:P

100:30:2

C:N:P

100:40:2

C:N:P

100:40:20

Control small

(BF3) Used as a control for the small filters (installed: February 2013)

Peroxide

addition (BF4) 0 mg/L 0.1 mg/L 0.5 mg/L 1 mg/L

Alum addition

(BF5) 0 mg/L 0.1 mg/L 0.5 mg/L 0.25 mg/L

GAC (BF6) GAC column (installed: June 2013)

58

4.2.5 Analytical Methods

Biofilter influent (raw water) and effluents were collected 4 hours after backwash during

the monitoring stage and first three weeks of stage 1 then at 26 hours after backwash to ensure

steady-state DOC removal (Figure 3.6). Temperature, pH, turbidity, dissolved organic carbon

(DOC), UVA at 254 nm (UV254), liquid chromatography-organic carbon detection (LC-OCD), 24

hour disinfection by-product (DBPs) formation potential (DBPFP) (trihalomethanes (THMs),

haloacetic acids (HAAs), and absorbable organic halogens (AOX)) were measured. In addition,

biofilter media was assayed for adenosine triphosphate (ATP), polysaccharides (PS) and proteins

(Pr). Media samples were obtained at the end of a given filter run (approximately 48 hours) and

collected from the surface of the large biofilters and 5 cm below the surface from the small

biofilters.

Raw water temperature was measured using a Siemens Milltronics Airanger DPL

thermometer (Peterborough, ON); a change in temperature was not observed through the filters.

Turbidity was measured using a Hach 2100N Turbidimeter (Mississauga, ON). pH was measured

using a Thermo Scientific Orion Star A111 pH Meter (Mississauga, ON).

DOC was measured using a wet oxidation method as described in standard method 5310

D (APHA, 2012). An O-I Corporation Model 1010 Analytical TOC Analyzer (College Station,

Texas, USA) was used to conduct the analysis. All samples were collected in amber glass bottles,

stored at 4ºC and analyzed within 7 days of collection. A new calibration curve was prepared for

each analysis. UV254 was measured using a CE 3055 Single Beam Cecil UV/Visible

Spectrophotometer (Cambridge, England) equipped with a 1 cm quartz cell (Hewlett Packard,

Mississauga, ON). LC-OCD analysis was conducted according to the method described Huber et

al. (2010) at the University of Waterloo (Waterloo, ON). Samples were filtered through 0.45 um

filter, stored in 40 ml amber glass vials at 4ºC, and shipped for analysis. Periodic calibration curves

and check standards were prepared at the University of Waterloo.

ATP analyses were conducted using a Luminultra Deposit Surface Analysis kit (DSA-

100C, Fredericton, NB) following manufacturer’s instructions. Media samples were stored at 4ºC

and analyzed within 8 hours of collection. Protein and polysaccharide analyses were carried out

using a method adapted from Papineau et al. (2012) and DuBois et al. (1956), respectively. CE

59

3055 Single Beam Cecil UV/Visible Spectrophotometer (Cambridge, England) and a Hach

Odyssey DR/2500 Scanning Spectrophotometer (Mississauga, ON) were used for protein and

polysaccharide analysis, respectively. Samples were extracted and analyzed within 24 hours of

collection.

DBP precursor removal was examined by bench-chlorination of raw water and biofilter

effluents. The 24 hours chlorine (Cl2) demand was determined by dosing samples with 7 mg/L Cl2

in 125 ml chlorine-demand-free amber bottles and the free chlorine residual was measured after

24 hours as described in Standard Method 4500-CI G (APHA, 2012). The free chlorine dosage

was calculated to provide a residual of 1.5±0.5 mg/L after 24 hours. Samples were chlorinated in

500 ml amber bottles with no head space, incubated at 20±2˚C for 24 hours, after which the

chlorine residuals were measured and were quenched using ascorbic acid (100 mg/L).

THMs (chloroform (trichloromethane, TCM), bromodichloromethane (BDCM),

dibromochloromethane (DBCM), and bromoform (tribromomethane, TBM)) were analyzed using

a liquid-liquid extraction gas chromatographic method based on Standard Method 6232 B (APHA,

2012). HAAs (monochloroacetic acid (MCAA), monobromoacetic acid (MBAA), dichloroacetic

acid (DCAA), trichloroacetic acid (TCAA), bromochloroacetic acid (BCAA), dibromoacetic acid

(DBAA), bromodichloroacetic acid (BDCAA), dibromochloroacetic acid (DBCAA), and

tribromoacetic acid (TBAA)) analyses were conducted using a liquid-liquid extraction gas

chromatographic method based on Standard Method 6251 B (APHA, 2012). A Hewlett Packard

5890 Series II Plus Gas Chromatograph (Mississauga, ON) equipped with an electron capture

detector (GC-ECD) and a DB 5.625 capillary column (Agilent Technologies Canada Inc.,

Mississauga, ON) was used for THM and HAA analysis. Check standards for THM and HAA

were analyzed after every tenth sample as per Standard Method 1020 (APHA, 2012); QA/QC

charts are shown in Appendix C. AOX analysis was conducted using a titration method based on

Standard Method 5320 (APHA, 2012) using a Trace Element Instruments Xplorer Organic

Halogens Analyzer (Delft, Netherlands).

4.2.6 Statistical Analysis

A paired t-test was used to evaluate the impact of the nutrient (nitrogen and/or phosphorus),

peroxide, and alum additions on DOC removal, turbidity reduction and headloss. A student’s t-test

60

was used to determine impacts on DBP precursor removal and UV254 reduction. All statistical

analyses were performed at the 95% confidence level.

4.2.7 UF Fouling Quantification

Membrane fouling was quantified by calculating the increase in membrane resistance over

the first 60 cycles (30 min each) for each UF experiment (if not available, the maximum number

of cycles was used). The reversible fouling rate for each cycle was calculated as the difference

between the final and starting resistance of two consecutive cycles; while the irreversible fouling

was calculated by subtracting the initial resistance of these cycles. Statistical analysis was not

performed since the UF experiments were not duplicated.

4.3 Results and Discussion

4.3.1 Biofilter Acclimation and Activity

Biological activity was determined by measuring ATP levels on biofilter media whereas

acclimation was determined by measuring DOC removal. Following installation of the large

biofilters in October 2012, DOC removals by biofiltration were monitored. After 2 months of

operations, the filters had a high ATP concentrations (approximately 800 ng/g), the DOC removal

remained below 2%. As the raw water temperature increased, the DOC removal increased and

reached steady-state removal (7±4%) once water temperature exceeded 20˚C. The small

anthracite/sand filters were installed in February 2013, and ATP levels monitored from June 2013.

The ATP concentrations were significantly lower than that of large filters (125±80 ng/g compared

to 635±5 ng/g). Yet, by the end of July, the small filters had similar ATP concentrations to the

large ones (800±107 ng/g compared to 541±78 ng/g). The average DOC removal for the small

filters was 4±3% in June and July. The GAC biofilter media, which has been in service for 8 years,

was installed on the June, 4th 2013. ATP concentration and DOC removal were approximately

1350 ng/g and 5±4%, respectively. The filters were deemed to be fully acclimated by the end of

July 2013, at which point, the DOC removals of all filters were found to be statistically similar

using a paired t-test (α= 0.05). This shows that the control filters are the same and by which the

first stage of enhacment was intiated.

61

Figure 4.2: ATP concentrations of biofilter media during the experimental period, test error <10%

based on manufacturer manual

4.3.2 Impact of Engineered Biofiltration on Turbidity Removal

The large and small biofilters removed 64±10% and 50±20% of the raw water turbidity

(0.333-0.842 NTU), respectively (data in Appendix B). There was no statistical difference between

the large and small control biofilters (α= 0.05, n= 6). Nutrient enhancement, alum and peroxide

addition did not impact on turbidity removal (α= 0.05, n= 5). In all instances, the biofilter effluent

had a turbidity level < 0.3 NTU as required by Health Canada for chemically assisted filtration

(Health Canada, 2009), except for one recording of < 0.4 NTU on the 12th August 2013 for the

small biofilters. The goal of this study was not to evaluate biofiltration as a single treatment process

but as a pre-treatment for ultrafiltration which itself can provide high turbidity removals, but it is

reassuring that biofiltration can perform to regulatory guidelines.

1

10

100

1000

10000

4-Jun-13 24-Jul-13 30-Sep-13 23-Oct-13 18-Nov-13

AT

P (

ng/g

med

ia)

Logar

ithm

ic-s

cale

Control Large (BF1) Nutrient Enhancement (BF2)

Control Small (BF3) Peroxide Addition (BF4)

Alum Addition (BF5) GAC (BF6)

C:N:P 100:30:2

H2O2 0.1 mg/L

Alum 0.1 mg/L

C:N:P 100:40:2

H2O2 0.5 mg/L

Alum 0.5 mg/L

C:N:P 100:40:20

H2O2 1 mg/L

Alum 0.25 mg/L

C:N:P 100:30:0

H2O2 0 mg/L

Alum 0 mg/L

N

ot

inst

alle

d

62

4.3.3 Impact of Engineered Biofiltration on NOM Removal

The DOC removal achieved by biofiltration across the control filters (one large and one

small) was consistently 5±2% (n= 18) throughout the experiment as shown in Figure 4.3. Direct

biofiltration studies conducted by Hallé et al. (2009) reported removals of 11% on average; also,

Peldszus et al. (2012) reported removals <15%. DOC removal was heavily influenced by water

temperature (Figure 4.2), where lower temperatures resulted in lower DOC removals which is

consistent with results reported by Emelko et al. (2006) and Moll et al. (1999) where they reported

a reduction in DOC and TOC removals at cold (<5˚C) temperatures.

DOC removal was not impacted (α= 0.05) by nitrogen, phosphorus, and alum addition in

this study, irrespective of the chemical dose (Figure 4.3), which was unexpected. The

ineffectiveness of nutrient enhancement is in contrast with the literature. A possible explanation is

a biodegradable carbon limitation in water, whereby the biofilters consumed all of the easily

biodegradable organic material. This hypothesis is supported by the fact that the supplemented

phosphorus or nitrogen were not utilized or consumed across the filter (no change was observed

between the influent and effluent phosphorus and nitrogen concentrations). Further studies could

examine biodegradable DOC (BDOC) or assimilable organic carbon (AOC) and impact of ozone.

As expected, hydrogen peroxide addition did not impact DOC removal as supported by the

literature (Urfer and Huck, 1997). Theoretically, an alum dose of 0.1-0.5 mg/L is sufficient to

reduce the biopolymer fraction (biopolymers were <10% of DOC), yet no variation in DOC was

observed in the biofilter effluent (α= 0.05). Hallé et al. (2009) also reported similar results,

suggesting that DOC measurements are not sensitive enough to detect biopolymer removal. GAC

was the most effective filter for DOC removal and removed an additional 0.1 mg/L of DOC on

average in comparison to control (α= 0.05). As previously discussed, the ability of GAC to remove

DOC might be attributed to higher surface area and/or the bio-regeneration at adsorption sites.

NOM characterization by LC-OCD was used to calculate the biopolymer removal by

biofiltration (Table 4.3). The control biofilters readily removed 20±4% (large column) and 16±7%

(small column) of raw water biopolymers (initial concentration (C0)= 0.41-0.53 mg/L). This trend

was not reflected in the DOC removals, demonstrating that DOC is not an effective tool for

predicting BP removal. Phosphorus addition (at 0.025 mg/L PO4-P to adjust C:N:P 100:30:2)

increased the biopolymer removal by 4%; however, the removal decreased by 25% when the C:N:P

63

Figure 4.3: DOC removal (%) by biofiltration (x-axis not to scale)

0

5

10

15

20

25

30

-2

0

2

4

6

8

10

12

14

T (

˚C)

DO

C R

emoval

(%

)Control Large (BF1) Nutrient Enhancement (BF2) Control Small (BF3)

Peroxide Addition (BF4) Alum Addition (BF5) GAC (BF6)

T (˚C)

C:N:P 100:40:2

H2O2 0.5 mg/L

Alum 0.5 mg/L

C:N:P 100:40:20

H2O2 1 mg/L

Alum 0.25 mg/L

C:N:P 100:30:2

H2O2 0.1 mg/L

Alum 0.1 mg/L

C:N:P 100:30:0

H2O2 0 mg/L

Alum 0 mg/L

64

ratio was adjusted to 100:40:2 by adding 0.25 mg/L PO4-P and 0.1 mg/L NH4-N. This indicates

that phosphorous dosing has a point of diminishing and even detrimental returns. Applying 0.1

mg/L of peroxide increased biopolymer removal by 4% when compared to control; however, when

the dose was increased to 1 mg/L, the biopolymer removal decreased by 5%. This could reflect the

peroxide-induced degradation of biofilm or EPS where approximately 50% less EPS was measured

in comparison to control. As expected, applying 0.1 and 0.25 mg/L of alum removed an additional

2% and 4% of biopolymers, respectively. This translates into an overall removal of 0.01 and 0.02

mg/L of biopolymers across the filter. In terms of media type, GAC did not provide an additional

advantage when compared to anthracite as the biopolymer removal by GAC was similar to the

anthracite control (α= 0.05) at 15±10%. It has been documented that biopolymers are relatively

large in terms of size (molecular weight) which may prevent them from accessing GAC absorptive

sites (Gibert et al., 2013; Velten et al., 2011b); thus, biopolymer removal by GAC was attributed

to solely to biological activity.

Table 4.3: Biopolymer and humic substance removals (%) by biofiltration, where BP=

biopolymers, HS= humic substances, and NS = not sampled.

Sample Date 16-Jun-13 12-Aug-13 24-Sep-13 12-Nov-13

Filter/Compound BP HS BP HS BP HS BP HS

Control Large (BF1) 22% 3% 21% -1% 21% 0% 15% 0%

Nutrient Enhancement

(BF2) -13% -2% 27% 4% 22% 3% 11% -5%

C:N:P 100:30:0 100:30:2 100:40:20

Control Small (BF3) 8% 4% 9% -4% 23% 1% 15% 1%

Peroxide Addition

(BF4) NS NS 13% -4% 24% 7% 9% 0%

Dose H2O2 0 mg/L H2O2 0.1 mg/L H2O2 1 mg/L

Alum Addition (BF5) 5% 5% 7% -4% 26% 5% 19% 1%

Dose Alum 0 mg/L Alum 0.1 mg/L Alum 0.25 mg/L

GAC (BF6) 17% 4% 0% 0% 25% 2% 17% -2%

As identified by LC-OCD, biofiltration was ineffective at removing humic substances

(Table 4.3). The large and small control biofilters removed 0±2% and 0±4% (n= 4) of the raw

waters humic substances (HS) (C0= 3.0-3.3 mg/L), respectively. Hallé et al. (2010) reported

preferential biopolymer removal (61±22%) and low (6±4%) HS removal by biofiltration. Unlike

65

biopolymers, phosphorus and nitrogen addition did not have an impact on HS. However, by the

end of the enhancement period, application of 0.1 mg/L of peroxide and alum increased the humic

substance removal from 1% in control to 7% and 5%, respectively. No improvement was observed

when the peroxide concentration was increased to 1 mg/L or when alum was increased to 0.25

mg/L. These results highlight the importance of dosing at optimal levels, rather than in excess.

GAC media did not have an effect on HS removal (α= 0.05) when compared to the anthracite

control.

4.3.4 Impact of Enhancement Strategies on Headloss (LOH)

The addition of 0.1, 0.5, and 1 mg/L hydrogen peroxide had a positive effect on

performance by decreasing headloss by 9, 48, and 40%, respectively (Table 4.4). In the first stage,

where 0.1 mg/L was added, the reduction was not significant, but in the second and third stages,

the reduction proved to be statistically significant (α= 0.05). The peroxide reacted immediately

with raw water resulting in a 0.1-0.3 mg/L demand prior to the filter influent, and no residual was

detected in the filter effluent (MDL= 0.1 mg/L). Headloss reduction is presumed to be a direct

result of EPS degradation. EPS is composed mainly of proteins and polysaccharides; it was found

that both constituents were approximately 30% higher when compared to control during the first

stage when 0.1 mg/L peroxide was applied; additionally, they were 6% and 52% lower,

respectively, when applying 0.5 and 1 mg/L peroxide (Figure 4.4). Lauderdale et al. (2011) also

reported lower EPS in peroxide-enhanced biofilter (1 mg/L). The effect of alum on headloss was

found to be dose-dependent, where 0.1 mg/L was the optimal dose resulting in a 40% decrease in

LOH. When the dose was increased to 0.25 mg/L, the filter performed similarly to the control (α=

0.05); however, a significant increase in LOH was noted when the dose increased to 0.5 mg/L

alum (α= 0.05). Typically, direct filtration coagulant dosages are lower than 20 mg/L, and are

expected to increase headloss, except during period of low turbidity and TOC conditions. This is

because aluminum hydroxide ions can precipitate and clog the filter (Edzwald et al., 1987). Yet in

this study, 0.1 and 0.25 mg/L of alum did not have an adverse effect on LOH. As such, low

coagulant dosages (<0.5 mg/L) have the potential for use in high turbidity (0.3-0.9 NTU) and DOC

(5-7 mg/L) water without significant impact on headloss. GAC filter had similar headloss to that

of the anthracite filter.

66

Table 4.4: Average hourly headloss for biofilters during the period of chemical addition, bold=

statistically significant difference at the 95% confidence level when compared to control

Filter/Condition

24-July-13 to

30-Sept-13

30-Sept-13 to

23-Oct-13

23-Oct-13 to

12-Nov-13

Control large (BF1) 0.015±0.012 0.005±0.003 0.012±0.014

Nutrient addition (BF2)

(C:N:P)

0.013±0.007

(100:30:2)

0.009±0.004

(100:40:2)

0.012±0.012

(100:40:20)

Control small (BF3) 0.018±0.012 0.006±0.004 0.021±0.01

Peroxide addition (BF4)

(Dose mg/L)

0.017±0.013

(0.1 mg/L H2O2) 0.003±0.004

(0.5 mg/L H2O2)

0.013±0.009

(1 mg/L H2O2)

Alum addition (BF5)

(Dose mg/L)

0.013±0.008

(0.1 mg/L alum) 0.009±0.004

(0.5 mg/L alum)

0.019±0.01

(0.25 mg/L alum)

GAC (BF6) 0.02±0.012 0.007±0.005 0.018±0.009

Figure 4.4: Protein and polysaccharide concentrations on media (PS= Polysaccharides, Pr=

Proteins), single sample analyzed in duplicate and triplicate for PS and Pr, respectively. Test

error <10%

0

100

200

300

400

500

600

700

PS Pr PS Pr PS Pr

24-Jul-13 30-Sep-13 18-Nov-13

Pro

tein

and P

oly

sacc

har

ide

Conce

ntr

atio

n (

μg/g

med

ia)

Control Large (BF1) Nutrient Enhancement (BF2)

Control Small (BF3) Peroxide Addition (BF4)

Alum Addition (BF5) GAC (BF6)

C:N:P 100:40:20

H2O2 1 mg/L Alum 0.25 mg/L

C:N:P 100:30:0

H2O2 0 mg/L Alum 0 mg/L

C:N:P 100:30:2

H2O2 0.1 mg/L Alum 0.1 mg/L

67

As shown in Figure 4.5, ATP concentrations were weakly correlated with protein

concentration (R2= 0.50), but no correlation was detected with polysaccharides (R2= 0.29),

suggesting EPS constituents are not a good surrogate for biological activity.

Figure 4.5: Relationship between ATP, proteins and polysaccharides

4.3.5 DBP Precursor Reduction

Biofiltration was an effective mean of reducing DBP formation potential (DBPFP) for

THMs, HAAs, and AOX (Figure 4.6). THMs consisted primarily of TCM and BDCM whereas

HAAs consisted of BCAA and TCAA. THMs and HAAs comprised 55±7% of the total AOX,

leaving 45% as unknown AOX. THM formation potentials (C0= 145-202 μg/L in raw water) were

significantly reduced by 20±11% by the large filter and by 18±9% across the small biofilter (α=

0.05). Similarly, HAAs (C0= 60-132 μg/L in raw water) decreased by 14±10% and 12±10% in the

large and small control filters, respectively. However, this decrease was not statistically

significant. Influent AOX (C0= 475-557 μg/L) was reduced by 6±11% and 11±10%, respectively,

however these too were not significant. These DBP precursor removals are similar to those

observed by others (Chaiket et al., 2002; Wang et al., 1995). Chaiket et al. (2002) reported 1-25%

removals for THMs and 6-25% for HAAs.

y = 0.13x + 7.5

R² = 0.27

y = 0.41x + 109

R² = 0.50

0

100

200

300

400

500

600

700

800

0 200 400 600 800 1000 1200

Pro

tein

and P

oly

sacc

har

ide

conce

ntr

atio

n (

ug/g

med

ia)

ATP concentration (ng/g media)

Proteins Polysaccharides

68

Figure 4.6: DBP formation potentional. Single samples analyzed in duplicate, vertical bars represent one standard deviation.

0

100

200

300

400

500

600

TH

M, H

AA

and A

OX

Conce

trat

ion (

μg/L

as

Cl- )

THMs HAAs Unknown AOX

C:N:P 100:30:2

H2O2 0.1 mg/L Alum 0.1 mg/L 24-Sept-13

C:N:P 100:40:2

H2O2 0.5 mg/L Alum 0.5 mg/L 23-Oct-13

C:N:P 100:40:20

H2O2 1 mg/L Alum 0.25 mg/L 21-Nov-13

C:N:P 100:30:0

H2O2 0 mg/L Alum 0 mg/L 24-July-13

69

Nutrient enhancement (P and N), peroxide, and alum addition had no significant impact on

DBPFP (Figure 4.6). Differences in THM, HAA, and AOX concentrations in the biofilter effluent

with C:N:P ratios of 100:30:2, 100:40:2, and 100:40:20 were not statically significant (α= 0.05)

when compared to control. Applying 0.1 mg/L alum only removed an additional 5% of THMs (11

μg/L) and 3% of HAAs (6 μg/L) when compared to the small control. Addition of 0.25 mg/L alum

did not impact THM removal but increased biofilter effluent HAA and AOX concentrations by 8

μg/L and 51 μg/L when compared to the control. Alum concentration of 0.5 mg/L had a detrimental

impact on DBP removals, where the THM removal was 7% less compared to the control and no

HAA removal was observed. These findings along with headloss results and UF fouling mitigation

observations, indicate that low alum dosages (<0.5 mg/L) may be beneficial at temperatures >

15˚C but not at colder temperatures < 15˚C. Doses of 0.1 and 0.5 mg/L of peroxide did not have

an adverse effect on the DBP precursor removal; however, at a dose of 1 mg/L, THM, HAA, and

AOX removals decreased by 2-12% compared to the control. When compared to anthracite, GAC

removed an additional 4% of THMs and 6% of AOX, but the trend did not extend to HAAs where

anthracite outperformed GAC by removing an additional 4% on average of HAAs (n= 4). While

no correlation was found between THM and biopolymer removal, both were selectively removed

by biofiltration which could suggest that biopolymers may serve as THM precursors (Figure 4.7).

Figure 4.7: DBP removal as a function of biopolymer removal

R² = 0.3158

R² = 0.5462

R² = 0.2943

-10%

-5%

0%

5%

10%

15%

20%

25%

30%

35%

0% 5% 10% 15% 20% 25% 30%

DB

P (

%)

Rem

oval

Biopolymer (%) Removal

THMs HAAs AOX

70

4.3.6 Engineered Biofiltration for Ultrafiltration Fouling Mitigation

To evaluate whether biofiltration can reduce UF fouling, two ultrafiltration units were run

in parallel; one with raw water and the other with effluent from the control biofilter (large) for 28

hours. Biofiltration reduced the irreversible fouling resistance rate from (7.4±29)×1010 m-1/cycle

(m-1/c) to (2.7±25)×1010 m-1/c (Figure 4.8), which translates into a 60% improvement. No impact

was observed on the reversible fouling as it remained at an average of 1.1×1012 m-1/c for both UF

units. Reduction in the irreversible UF fouling was attributed to biopolymer removal (20±4%) and

particulate matter removal (turbidity removal of 64±10%). Further investigation to quantify the

biopolymer, protein, and polysaccharide rejection by the membrane would clarify which

constituent is responsible for the irreversible fouling. The control biofilter was connected to the

UF immediately following a backwash. This resulted in an initial increase in membrane resistance

by approximately 3×1012 m-1 over 4 cycles (approximately 2 hours) (Figure 4.8). This filter

ripening effect was transient in nature and did not adversely affect the overall results, however in

future studies, it is recommended that filters should be given 3 hours of run time post-backwash

prior to connection with the UF units.

Figure 4.8: Normalized UF resistance profile comparing raw water vs large control biofilter

effluent (October, 30th 2013, T= 11˚C)

Both large and small biofilters had similar irreversible fouling rates (6.5±48)×1010 m-1/c

and (6.9±35)×1010 m-1/c, respectively, therefore biofilter size did not influence fouling profiles.

0

2

4

6

8

0 5 10 15 20 25 30

Norm

aliz

ed m

embra

ne

resi

stan

ce @

20

˚C (

×1

01

2 m

-1)

Permeation Time (hr)

Control Large (BF1)

Raw Water

71

Phosphorus and nitrogen addition (C:N:P 100:40:20 or stage 3) did not improve

irreversible fouling rates (7.8±22)×1010 m-1/c and (7.7±57)×1010 m-1/c, for nutrient enhance and

control, respectively (Figure 4.9). The reversible fouling rate, however, increased by

approximately 50% and was attributed to the lower biopolymer removals where nutrient enhanced

removed 11% compared to 15% in control. Turbidity removal followed a similar trend with

removals of (49% compared to 55%), for the nutrient enhanced and control (data not shown).

Figure 4.9: Normalized UF resistance profile comparing nutrient enahnced vs large control

biofilter effluent (BF1: November 18th,2013, T= 7ºC. BF2: November 15th, 2013, T= 6ºC)

In-line coagulation prior to biofiltration at dosages of 0.1, 0.25, and 0.5 mg/L was evaluated

in terms of UF fouling control (Figures 4.10, 4.11, and 4.12). When compared to the control,

application of 0.1 and 0.25 mg/L alum decreased irreversible fouling by 60% and 35%,

respectively. This improvement might be attributed to the minimal increase biopolymer removal

(3-4% more than the control), which underlines the importance of biopolymer removal for

irreversible UF fouling control. From an operational standpoint, irreversible fouling mitigation is

a key consideration for minimizing costs. These improvements in irreversible fouling rates did not

translate into improvements in reversible fouling, which increased by approximately 30% and 55%

compared to the control for the respective runs. It is well documented that alum performance

decreases with water temperature (Van Benschoten and Edzwald, 1990) and during the study,

temperatures steadily decreased which could have resulted in poor biopolymer removal, despite

the increase in dose. The increase in alum dose may also result in an increase in aluminum residuals

0

2

4

6

8

0 5 10 15 20 25 30

Norm

aliz

ed m

emb

ran

e

resi

stan

ce @

20

˚C (

×10

12

m-1

)

Permeation Time (hr)

Control Large (BF1)

Nutrient Enhanced (BF2, C:N:P 100:40:20)

72

(not measured), which can foul the UF membrane (Zheng et al., 2012). This was more pronounced

when applying 0.5 mg/L where no irreversible fouling improvement was observed. Conversely, it

resulted in lower reversible fouling rates of 35% when compared to control. This improvement can

be attributed to the formation of cake layer that prevents further attachment of foulants, as

discussed by Paar et al. (2011) and Zheng et al (2012). LC-OCD data was not available for this

testing condition; therefore biopolymer removal could not be analytically quantified or correlated.

Figure 4.10: Normalized UF resistance profiles comparing 0.1 mg/L in-line alum vs large

control biofilter effluent (Septmeber, 16th 2013, 18˚C)

Figure 4.11: Normalized UF resistance profile showing the impact of 0.25 mg/L in-line alum vs

small control biofilter effluent (BF3: Novermber 20th, 2013, T= 6.5ºC. BF5: November 22th,

2013, T= 5.5ºC)

0

2

4

6

8

0 5 10 15 20 25 30

Norm

aliz

ed m

emb

ran

e

resi

stan

ce @

20

˚C (

×10

12

m-1

)

Permeation Time (hr)

Control Large (BF1)

Alum Addition (BF5, 0.1 mg/L)

0

2

4

6

8

0 5 10 15 20 25 30

No

rmal

ized

mem

bra

ne

resi

stan

ce @

20

˚C (

×10

12

m-1

)

Permeation Time (hr)

Control Small (BF3)

Alum Addition (BF5, 0.25 mg/L)

73

Figure 4.12: Normalized UF resistance profile the impact of 0.5 mg/L in-line alum vs small

control biofilter effluent (October 7th, 2013, T= 18ºC)

Peroxide addition (at 1 mg/L) increased the reversible fouling by 45% and had little impact

on irreversible, with a nominal increase of 5% as compared to control (Figure 4.13). Peroxide can

degrade EPS causing the release of proteins and polysaccharides in biofilter effluent. This is

supported by the 60% reduction of EPS constituent in the biofilter media when 1 mg/L peroxide

was added. LC-OCD corroborates the peroxide-induced degradation of EPS since the biopolymer

concentration actually increased (by 0.03 mg/L) in the filter effluent, in comparison to control. The

improvement in headloss with peroxide did not translate into improvements in UF fouling, in fact,

peroxide adversely affected UF performance. This novel finding merits further investigation of

peroxide doses < 1 mg/L.

GAC resulted in a similar irreversible fouling as compared to anthracite, but 29% lower

reversible fouling (Figure 4.14) and is reflected by a slight increase in biopolymer removal (2%),

again demonstrating the importance of biopolymers when discussing UF fouling. As was the case

for NOM, the improvement in GAC performance was attributed to higher surface area and

potential for reactivation of adsorption sites. Furthermore, this study shows the potential of for the

use of exhausted filter media as a method for eliminating acclimation time.

0

2

4

6

8

0 5 10 15 20 25 30

No

rmal

ized

mem

bra

ne

resi

stan

ce @

20˚C

10

12

m-1

)

Permeation Time (hr)

Control Small (BF3)

Alum Addition (BF5, 0.5 mg/L)

74

Figure 4.13: Normalized UF resistance profile showing the impact of 1 mg/L of peroxide vs the

small control biofilter effluent (November 11th, 2013 T= 8ºC)

Figure 4.14: Normalized UF resistance profiles depict the impact of filter media type ( GAC vs

anthrcite) (BF3: Novermber 20th, 2013, T= 6.5ºC. BF6: November 25th, 2013, T= 4ºC)

4.4 Summary

Biofiltration was found to be effective for turbidity, biopolymer, and DBP precursor

reduction as well as a pre-treatment strategy to reduce irreversible UF fouling. Adjustment of the

C:N:P ratio in the filter influent (to 100:30:2, 100:40:2, and 100:40:20) did not improve

performance. Hydrogen peroxide addition at dosages < 1 mg/L decreased headloss, THM

formation potential, and AOX; however, it adversely affected UF fouling mitigation and HAA

precursor removal. In-line coagulation (< 0.5 mg/L) prior to biofiltration has the potential to

0

2

4

6

8

0 5 10 15 20 25 30

No

rmal

ized

mem

bra

ne

resi

stan

ce @

20

˚C (

×1

01

2 m

-1)

Permeation Time (hr)

Control Small (BF3)Peroxide Addition (BF4, 1 mg/L)

0

2

4

6

8

0 5 10 15 20 25 30

Norm

aliz

ed m

embra

ne

resi

stan

ce @

20˚C

10

12

m-1

)

Permeation Time (hr)

Control Small (BF3) GAC (BF6)

75

improve UF fouling and reduce DBP precursors without impacting headloss by increasing

biopolymer removal. GAC outperformed anthracite in terms of NOM and DBP precursor removal,

and reduced UF reversible fouling; however no impact on irreversible fouling was observed.

Tables 4.5 and 4.6 summarize the findings of the study.

Table 4.5: Impact of filter media and biofiltration conditions on performance parameters (↑=

positive impact, ↓= adverse impact, ●= no impact, = significant and = not statistically

significant at the 95% confidence. Where symbols not present= statistical analysis was not

possible, NS= Not Sampled, NA= Not applicable)

Condition DOC BP HS THMs HAAs AOX LOH

Biofiltration NA ↑ ↑ ● ↑ ↑ ↑ NA

Filter Size NA ● ● ● ● ● ● NA

Nutrient

Enhancement

(C:N:P)

100:30:2 ● ↑ ● ● ● ● ●

100:40:2 ● NS NS ● ● ● ●

100:40:20 ● ↓ ● ● ● ● ●

Peroxide

Addition

(mg/L)

0.1 ● ↑ ↑ ● ● ● ↑

0.5 ● NS NS ● ● ● ↑

1 ● ↓ ● ↓ ↓ ↓ ↑

Alum

Addition

(mg/L)

0.1 ● ↑ ↑ ↑ ↑ ● ↑

0.5 ● NS NS ↓ ↓ ↑ ↓

0.25 ● ↑ ● ● ↓ ↓ ●

Filter Media GAC ↑ ● ● ↑ ↓ ↑ ●

Table 4.6: Impact of biofiltration, filter media, and chemical addition on reversible and

irreversible UF fouling (↑= positive impact, ↓= adverse impact, ●= no impact)

Condition Irreversible UF

fouling control

Reversible UF

fouling control

Biofiltration NA ↑ ●

Filter Size NA ● ↓

Nutrient Enhancement

(C:N:P) 100:30:2 ● ↓

Peroxide Addition (mg/L) 1 ↓ ↓

Alum Addition (mg/L)

0.1 ↑ ↓

0.5 ● ↑

0.25 ↑ ↓

Filter Media GAC ● ↑

76

5. Comparison between Conventional Treatment and

Biofiltration for DBP Precursor Removal and UF

Fouling Control

5.1 Introduction

Application of low pressure membranes such as ultrafiltration (UF) has gained momentum

in recent years because it can provide high quality drinking water at a competitive cost when

compared to conventional treatment (Neubrand et al., 2010; Wang and Wang, 2006). However,

fouling is a major limitation because it can increase operation costs and can damage membrane

surfaces (Howe and Clark, 2002; Singh, 2006). Natural organic matter (NOM) and particulates

have been identified as major foulants (Wang and Wang, 2006); recent studies have shown that

biopolymers and colloids may represent the major sources of irreversible organic fouling (Hallè,

2010; Neubrand et al., 2010; Zheng et al., 2010; Zheng et al., 2012). Coagulation may result in

severe irreversible fouling if improper dosing occurs (Malievialle et al., 1996; Neubrand et al.,

2010; Wang and Wang, 2006). Biofiltration presents a cost effective and chemical-free process for

fouling control (Hallé et al., 2009; Huck et al., 2011). Yet, a parallel comparison between

biofiltration and coagulation has not been performed to date (Peldszus et al., 2012).

Coagulation enables the destabilization of particles by different mechanisms (e.g. charge

neutralization, sweep flocculation) to facilitate the formation of larger particles that can be

removed by settling or filtration (Crittenden et al., 2012). Coagulation is effective for removing

NOM and DBP precursors (Matilainen et al., 2010). Different coagulation configurations have

been applied as UF pre-treatment including coagulation/flocculation with/without sedimentation

(Crittenden et al., 2012), and in-line coagulation (Neubrand et al., 2010; Wang and Wang, 2006;

Zheng et al., 2012). Some studies reported improvement in UF fouling mitigation when applying

coagulation upstream (Choi and Dempsey, 2004; Li et al., 2013), while others did not (Howe and

Clark, 2006; Neubrand et al., 2010). Howe and Clark (2006) observed that coagulation was

effective for UF fouling control at an enhanced dose but not at lower doses. Wray et al. (2014) did

not observe an improvement when applying coagulation (< 15 mg/L alum, below enhanced

dosages) as a pre-treatment for UF membranes and found the point of diminishing return for

biopolymer removal to be 0.1 mg/L.

77

Applying biofiltration as a UF pre-treatment to reduce organic fouling has been explored

by others (Basu and Huck, 2004; Hallé et al., 2009; Huck et al., 2011). Hallé et al. (2009) reported

an increase of 9 psi at 2000 L/m2 in UF transmembrane pressure (TMP) when using raw water.

However, when using biofilter effluent with EBCT of 5 and 14 minutes, a decrease in TMP rates

of 4 and 1 psi at a permeate volume of 7500 L/m2, respectively, were reported. Peldszus et al.

(2011) found that pilot-scale biofilters operated at 5, 10 and 15 minutes were effective for fouling

control even at low temperatures (2-14°C). In contrast, Geismar et al. (2012) reported the impact

of post ozonation lab-scale BAC (EBCT= 11 min) on the unified membrane fouling index (UMFI)

to be statically insignificant (α= 0.05).

The objective of this pilot-scale study was to compare biofiltration to coagulation/

flocculation/sedimentation with/without filtration in terms of NOM removal, DBP precursor

reduction, and ultrafiltration (UF) fouling mitigation. Two parallel pilot-scale conventional

treatment trains using aluminum sulfate (alum) and polyaluminum chloride (PACl) and a

biofiltration train were used for this research. In addition, full-scale plant effluent samples were

collected for comparison.

5.2 Materials and Methods

5.2.1 Source Water Quality

The study was conducted at Peterborough water treatment plant, where the Otonabee River

(Ontario, Canada) was used as source water. The water was pre-chlorinated for zebra mussel

control when the water temperature was > 12˚C. Prior to biofiltration, chlorine (< 0.5 mg/L) was

quenched by applying sodium thiosulfate at 3.5 mass ratio. Raw water DOC ranged from 5.8-7.3

mg/L with turbidities between 0.3 and 1.5 NTU; temperature ranged from 3 to 25˚C during the

study period.

5.2.2 Pilot- and Full-scale Plant Configuration

The pilot-plant consisted of two parallel conventional treatment trains (using alum and

PACl as coagulants) and a biofiltration column. The coagulation train consisted of rapid mix,

tapered flocculation, parallel plate settlers, and a glass dual media filter (Diameter (D)= 15.24 cm,

empty bed contact time (EBCT)= 12-36 min). The biofilter (D= 15.24 cm) was operated at an

78

EBCT of 11 minutes. A general schematic of the pilot-plant configuration is shown in Figure 5.1.

The filters (biological and conventional) contained 50 cm of anthracite (effective size (d10)= 0.85,

uniformity coefficient (UC)= 1.8) on top of 50 cm of sand (d10= 0.5, UC= 1.8); a wedge wire plate

was used to support the media. The water level above each filter was maintained at 90 cm from

the top of the filter media (constant head mode). The conventional filters were backwashed using

chlorinated full-scale plant effluent when the required flow could not be maintained or a maximum

of twice a week. The biofilter was backwashed using its own effluent when the required flow could

not be maintained, or a maximum of three times per week (on Monday, Wednesday, and Friday).

The backwash procedure included a slow backwash with air scour for 2 minutes (30% bed

fluidization) followed by an 8 minutes of fast backwash at collapse pulsing conditions (50% bed

fluidization), and a slow backwash for 4 minutes. The full-scale plant configuration was similar to

that of the pilot; however, the filters contained 37 cm of anthracite on top of 40 cm of sand, and

23 cm of gravel was used for support. Media and non-chlorinated plant effluent samples were

collected from full-scale filter #11. The average full-scale plant alum dosage was applied at the

pilot and was adjusted every two weeks. A PACl dose which provided an equivalent DOC, when

compared to alum, was used as well (Appendix D, Figure 8.13). The conventional pilot treatment

train was off-line from 28th June until 18th October 2013; a transition study took place between

November 8th and 31st, 2013, where coagulants were switched between the pilot-plant conventional

treatment trains. The biofilter’s EBCT varied between 11 and 30 minutes during the month of June

2013, otherwise it was stable at 11 minutes for the rest of the study.

5.2.3 Ultrafiltration Unit and Fouling Quantification

Two automated ultrafiltration pilot systems were adapted for this research work as

discussed in section 4.2.3. The reversible and irreversible fouling was calculated as described in

section 4.2.7.

5.2.4 Analytical Methods

The biofiltration and the two conventional treatment trains were equipped with on-line

water quality instruments and connected to the full-scale plant sophisticated supervisory control

and data acquisition (SCADA) system. These included pH (Hach DPD1P1, Mississauga, ON),

79

temperature (Siemens Milltronics Airanger DPL, Peterborough, ON), turbidity (Hach, 1720E,

Mississauga, ON), UV absorbance at 254 (UV254) (Real Tech, UV254 M3000, Whitby, ON),

Figure 5.1: Pilot-plant schematic (BF= Biofilter, CF= Conventional Filter, and = Sampling

location) (Not to scale)

80

dissolved organic carbon (DOC) (General Electric, 5310 C On-Line TOC analyzer, Colorado,

USA), and differential headloss meters (Siemens, Sitrans P XP/DIP, Peterborough, ON).

The following parameters were measured for raw water, full- and pilot-plant conventional

treatment (alum and PACl), and biofilter effluents (biofiltration samples were obtained 4 hours

after a backwash before Aug 12th, 2013, and 26 hours after a backwash onwards as it was discussed

in section 3.3.3): liquid chromatography-organic carbon detection (LC-OCD), 24 hours

disinfection by-product (DBPs) formation potentials (DBPFP) (trihalomethanes (THMs),

haloacetic acids (HAAs), and absorbable organic halogens (AOX)) as described in section 4.2.5.

Adenosine triphosphate (ATP), polysaccharides (PS) and proteins (Pr) were also measured on filter

media obtained at approximately 48 hours of a given filter run as described in section 4.2.5. Media

samples were collected from the surface of the filters.

DBPs were formed by chlorinating raw water and biofilter effluents. 24 hour chlorine (Cl2)

demands were estimated by dosing the conventional treatment and biofilter samples with 5 and 7

mg/L Cl2, respectively, as discussed in section 4.2.5.

5.2.5 Statistical Analysis

A paired t-test was used to compare DOC, UV254, DBP precursor, turbidity, biopolymer,

and humic substance removals and headloss between the treatment options. All statistical analyses

were performed at the 95% confidence level.

5.3 Results and Discussion

5.3.1 Biofilter Acclimation and Activity

The anthracite/sand filter was installed in October 2012 at the Peterborough Water

Treatment Plant and fed raw unchlorinated water to encourage biological growth. Adenosine tri-

phosphate (ATP) measurements serve as a reliable surrogate for estimating the microbial content

present in filter media (Magic-Knezev and van der Kooij, 2004; Velten et al., 2007). While

biological activity can be inferred from ATP measurements, performance of the biomass was

determined by biodegradation of organic material or DOC removal across the filter. DOC removals

were low (< 2%) during the winter months despite the high ATP concentrations (> 800 ng/g

media), as measured in December 2012 (Table 5.1). DOC removal increased as a function of water

81

temperatures (from 3 to 20°C) and reached 7±2% by June 2013. Biofilter media ATP

concentrations (706±217 ng/g media) were similar to values reported by Lauderdale et al. (2012)

of approximately 800 ng/g media, and were 1-2 orders of magnitude higher than the non-

biological, conventional treatment filters (α= 0.05) (Table 5.1). ATP concentrations on

conventional filters remained below 100 ng/g throughout the study. Pilot- and full-scale filter

media had similar ATP concentrations (α= 0.05).

Table 5.1: ATP concentrations (ng/g media) on biofilter and conventional filters media, NS=

Not sampled. a Transition study: coagulants were switched between trains

Date/Filter Biofilter Full-scale Plant

Filter #11

Pilot-plant filter

(Alum)

Pilot-plant filter

(PACl)

22-Dec-12 889 94 42 10

20-Feb-13 1115 48 35 26

4-Jun-13 708 28 8 7

24-Jul-13 814 76 NS NS

30-Sep-13 493 67 NS NS

23-Oct-13 259 52 NS NS

18-Nov-13a 669 68 8 58

5.3.2 Turbidity Removal

The biofilter was effective (71±11%) at removing raw water turbidity (0.29-1.46 NTU)

(Figure 5.2). Pilot conventional treatment trains (enhanced coagulation/filtration) using both alum

and PACl achieved 94±2% and 93±4% turbidity removal respectively, and maintained filter

effluent values below 0.1 NTU (Figure 5.2). At temperatures > 15ºC, coagulation contributed to

approximately 66-75% of the turbidity removal; however < 15ºC, coagulation increased or did not

significantly (α= 0.05) impact the turbidity.

5.3.3 NOM Removal

During the period of June to November 2013, biofiltration removed 6±2% of the raw water

DOC (initial concentration (C0)= 6.2±0.1 mg/L) as shown in Figure 5.3 (α= 0.05), consistent with

Hallé et al. (2009). DOC removal was directly impacted by temperature; where removal was 6-

10% when the temperature was > 10˚C and decreased to 2-4% below this value. These results are

consistent with Liu et al. (2001) and Emelko et al. (2006) findings.

82

Figure 5.2: Average monthly turbidity trends. NS= Not sampled. a Transition study: coagulants

were switched between trains. Vertical bars present one standard deviation

Conventional treatment with alum and PACl consistently removed 46±1% and 45±1% of

the raw water DOC; full-scale treatment removed 46±1%. DOC removal was not influenced by

temperature as the coagulant dose was adjusted to maximize DOC removal on a bi-weekly basis

(Appendix D, Figure 8.13). In all cases, DOC removal was attributed to coagulation and not

filtration (Table 5.1) since filtration only removed an additional 0-4% of the raw water DOC,

consistent with Volk and Lechevallier (2002). Hence, the application of biofiltration post

coagulation could provide an additional 5-10% DOC removal. Finally, DOC removal in

conventional treatment was 5-6 times higher than biofiltration at temperatures > 10˚C and

approximately an order of magnitude higher at temperatures < 10˚C. Despite the low DOC removal

in comparison to conventional treatment, it is important to keep in mind that biofiltration is not

meant to be applied as a single process but combined with UF and/or pre-ozonation.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

June July Aug Sept Oct 1-7 Nov 7-30 Nov

Turb

idit

y (

NT

U)

Raw water BiofilterPost Full-scale Plant Filter #11 Pre-filter (Alum)Post-filter (Alum) Pre-filter (PACl)Post-filter (PACl)

Transition

study a

NS

NS

NS

NS

NS

NS

83

Figure 5.3: Average monthly DOC concentrations. NS= Not sampled. a Transition study: coagulants were switched between trains.

Vertical bars represent one standard deviation

0

5

10

15

20

25

30

35

0

1

2

3

4

5

6

7

Jun-13 Jul-13 Aug-13 Sep-13 Oct-13 1-7 Nov-13 7-30 Nov-13

Raw

wat

er t

empra

ture

(˚C

)

DO

C (

mg/L

)

Raw Water Biofilter Post Full-scale Plant Filter #11

Post-filter (Alum) Post-filter (PACl) Tempreture

Transition

study a

NS

NS

NS

NS

NS

NS

84

Table 5.2: DOC removal pre/post pilot conventional filters with alum and PACl, and after

biofiltration, SL= Sample lost during transportation

Sample 4-Jun-13 24-Jun-13 23-Oct-13 12-Nov-13

Post Full-scale Plant Filter #11 SL SL 44% 47%

Post Biofiltration 11% 5% 3% 3%

Pre-filter (Alum) 48% 39% 44% 44%

Post-filter (Alum) 50% 43% 45% 44%

Pre-filter (PACl) 45% 38% 39% SL

Post-filter (PACl) 49% 40% 41% 44%

Raw water UV254 was approximately 0.16±0.2 cm-1 for the months from June to November

2013 (Figure 5.4). Average UV254 monthly removal via biofiltration was 2±4% (α= 0.05).

Conventional treatment with alum (full- and pilot-scale) and PACl removed 55±2% and 55±3%

of the raw water UV254. UV254 measures the aromatic fraction of NOM (Singer, 1999); hence,

coagulation appears to be more effective than biofiltration at removing these aromatic species.

UV254 and DOC removals in biofiltration were correlated (R2= 0.6, n= 91) as shown in Figure 5.5.

This suggets that UV254 could be used as a surogate for DOC removal in biofiltration studies.

Figure 5.4: Average monthly UV254 values. NS= Not sampled. a Transition study: coagulants were

switched between trains. Vertical bars represent one standard deviation

0

0.05

0.1

0.15

0.2

0.25

Jun-13 Jul-13 Aug-13 Sep-13 Oct-13 1-7 Nov-137-30 Nov-13

UV

25

4(c

m-1

)

Raw Water Biofilter

Post Full-scale Plant Filter #11 Post-filter (Alum)Transition

study a

NS

NS

NS

NS

NS

NS

85

Figure 5.5: UV254 removal as a function of DOC removal

The biofilter consistently removed 20±4% of raw water biopolymers (BP) (C0= 0.41-0.53

mg/L) and 0±2% of humic substances (HS) (C0= 3.0-3.3 mg/L) (Figure 5.6). Biopolymer removals

were lower than values (approximately 50%) reported by Halle et al. (2009) and Peldszus et al

(2012) under the similar temperature conditions. That said, low HS removals (< 2%) were

consistent with Hallé et al. (2009) findings. Conventional treatment (pilot- and full-scale) was

effective for biopolymer (58±9%) and HS (58±7%) removal. Coagulation alone removed the

majority of the BP (42-49%) and HS (52-71%). These removals are consistent with Diemert and

Andrews (2013) where they observed 50-75% and 43-63%, for the same water matrix used in the

present study. The ability of conventional treatment to remove HS when compared to biofiltration

explains the higher DOC removals observed in this study, since HS comprise 54±10% of the raw

water DOC. Higher DBP precursor reductions are expected when applying conventional treatment

as opposed to biofiltration, since humic substances have been identified as major precursors for

DBPs (Wassink et al. 2011). Unlike PACl, coagulation with alum followed by filtration actually

added HS to the effluent by (6-10% lower). This might be attributed to the release of humic

substances via alum floc breakage during filtration (Jarvis et al., 2005; Yu et al., 2009).

y = 0.74x - 0.0044

R² = 0.60

0%

1%

2%

3%

4%

5%

6%

7%

8%

9%

0% 2% 4% 6% 8% 10% 12%

UV

25

4R

emo

val

DOC removal

86

Figure 5.6: Biopolymer (BP) and humic substance (HS) removals. NS= Not sampled. SL=Sample

lost during transportation. UFS= UF experiment samples. a Transition study: coagulants were

switched between trains.

5.3.4 Headloss (LOH)

Biofiltration headloss was 75-90% higher when compared to pilot- and full-scale alum

treatment, but similar (α= 0.05) to the PACl filter at temperatures > 20˚C (Figure 5.7 and 5.8a).

When temperatures were < 20˚C, headloss through the biofilter and alum filter were the same (α=

0.05) but slightly higher than PACl (Figure 5.8b). PACl was shown to be superior to alum at

temperatures < 15˚C as can be observed from the steady turbidity removals (Figure 5.3) which

explains the lower headloss when compared to alum. Headloss for biofiltration showed high

variability (relative standard deviation= 52-143%) which was attributed to variability in raw water

turbidity (0.29-1.46 NTU), as described earlier. Microbial activity in biofilters is thought to

increase filter headloss by clogging the filter pores and voids (Kim et al., 2009). The increase in

headloss might also be attributed to EPS excreted by microorganisms (Lauderdale et al., 2012).

EPS (proteins and polysaccharides) in the biofilters were significantly higher when compared to

conventional treatment as a result of the biological activity (Figure 5.9). EPS concentrations

decreased as a function of temperature, similar to ATP.

-10%

0%

10%

20%

30%

40%

50%

60%

70%

80%

BP HS BP HS BP HS BP HS BP HS

16-Jun-13 12-Aug-13 24-Sep-13 12-Nov-13 29-Nov-13

Hum

ic a

nd B

ipoly

mer

Rem

oval

Biofilter Post Full-scale Plant Filter #11Pre-filter (Alum) Post-filter (Alum)Pre-filter (PACl) Post-filter (PACl)

Transition study a

NS

NS

NS

NS

NS

NS

NS

NS

SL

SL

UF

S

UF

S

UF

S

UF

S

87

Figure 5.7: Average hourly filter headloss. NS= Not sampled. a Transition study: coagulants were

switched between trains. Vertical bars represent one standard deviation

Figure 5.8: Typical headloss curves for the biofilter and pilot-plant (alum and PACl) filters during

summer 2003 (a: 24th June 2013, T=22˚C) and winter 2013 (b: 25th October 2013, T=12˚C)

-0.01

0

0.01

0.02

0.03

0.04

0.05

0.06

June July Aug Sept Oct 1-7 Nov 7-30 Nov

Avar

age

hourl

y L

OH

(m

/hour)

Biofilter Full-scale Plant Filter #11

Pilot-plant Filter (Alum) Pilot-plant Filter (PACl)

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 50 100

LO

H (

m)

Time following backwash (hr)

Pilot-plant Filter (Alum)Pilot-plant Filter (PACl)Biofilter

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

0 50 100

LO

H (

m)

Time following backwash (hr)

Pilot-plant Filter (Alum)Pilot-plant Filter (PACl)Biofilter

Transition

study a

NS

NS

NS

NS

NS

NS

a) b)

88

Figure 5.9: Protein and polysaccharide concentrations for biofilter and conventional filter media

(PS= Polysaccharides, Pr= Proteins), single sample analyzed duplicate and triplicate for PS and

Pr, respectively. Test error < 10%. a Transition study: coagulants were switched between trains

5.3.5 DBP Precursor Removal

Biofiltration reduced DBPFP, as measured by THMs, HAAs, and AOX (Figure 5.9). THMs

consisted of TCM and BDCM whereas HAAs comprised of BCAA and TCAA. THMs and HAAs

comprised 49±10% of the total AOX. Raw water THM formation potential (C0=179-255 µg/L)

was reduced by 13±11% through biofiltration (α= 0.05), and HAAs (C0=110-202 µg/L) decreased

by 9±13% (not significant at α= 0.05). Similarly, raw water AOX (C0= 425-551 ug/L) was reduced

by 2±5% by biofiltration (not significant). These removals are similar but slightly lower than

removals reported by others (Chaiket et al., 2002; Wang et al., 1995). The biofilter effluent THM

and HAA concentrations are higher than the EPA’s guidelines (EPA, 2009), and they are not

expected to be significantly decreased by UF because it is incapable of removing humic

substances, which are the major DOC constituents (data not shown). Therefore, water with high

humic substances concentration (> 3 mg/L) is not expected to conform to DBP guidelines when

treated with biofiltration/ultrafiltration without applying coagulation.

Conventional treatment reduced THM, HAA, and AOX by 57-72%, 49-76%, and 48-73%,

respectively. Alum and PACl provided similar DBP precursor removal (α= 0.05). Filtration

contributed to THM, HAA, and AOX removal by 0-2%, 6-8% and 3-19%, respectively, when the

0

100

200

300

400

500

600

700

800

PS Pr PS Pr PS Pr PS Pr

4-Jun-13 24-Jul-13 30-Sep-13 18-Nov-13

Pro

tein

and P

oly

sacc

hri

de

Conce

ntr

atio

n (

µg/g

med

ia)

Biofilter Full-scale plant filter 11

Pilot-plant Filter (Alum) Pilot-plant Filter (PACl)

a Transition

study

NS

NS

NS

NS

NS

NS

NS

NS

89

Figure 5.10: 24 hour DBP formation potentionals. NS= Not sampled. SL=Sample lost during transportation. a Transition study:

coagulants were switched between trains. Vertical bars represent one standard deviation

0

100

200

300

400

500

600T

HM

, H

AA

and A

OX

Conce

ntr

atio

n (

µg/L

as

Cl- )

THMs HAAs Unknown AOX

24-Sept-13 4-June-13 23-Oct-13 a Transition study

12-Nov-13

NS

NS

NS

NS

SL

90

temperature was > 10˚C, and by 22%, 30%, and 34%, respectively when temperate was <

10˚C. The lower removal by filtration at temperatures > 10˚C was attributed to the increase

of HS following filtration as discussed in section 5.3.3. The higher DBP precursor control

by conventional treatment as compared to biofiltration was attributed to the higher DOC,

biopolymer, and humic substance removal. These parameters have been previously

correlated with DBPs in the same water matrix (Wassink et al., 2011). Also, the lower pH

(approximately 7.5 for coagulation and 8 for biofiltration) as a result of coagulant addition

may have contributed to the lower DBP formation potentials.

5.3.6 Ultrafiltration Fouling Mitigation

As discussed in section 4.3.6, biofiltration was capable of reducing UF irreversible

fouling by approximately 60% when compared to raw water. An evaluation of biofiltration

vs. coagulation with/without filtration was performed to compare both strategies for

fouling control. Coagulation (29.1-30.1 mg/L alum) resulted in approximately 25% higher

UF irreversible fouling mitigation, and approximately an order of magnitude higher

reversible fouling control when compared to biofiltration (Figure 5.11). Superior

biopolymer removal by coagulation (56% compared to 10% by biofiltration) was thought

to be responsible for reducing irreversible UF fouling. However, the increased turbidity, as

discussed earlier, may have adversely impacted irreversible fouling. Therefore, turbidity-

biopolymer interactions are a critical parameter for reduction UF fouling. These findings

are consistent with Peldszus et al. (2011) where irreversible fouling was attributed to

interaction between biopolymer and particulate matter. Also, fouling due to aluminum ions

may have masked the impact of biopolymer removal (not measured).

Prendiville and Chung (1992) reported that turbidity and alum carry-over could

impact membrane fouling. To isolate this impact, dual media filter effluent following

coagulation (30.0 mg/L alum) was compared to biofiltration as a UF pre-treatment strategy

(Figure 5.12). The filtered effluent resulted in approximately an order of magnitude lower

reversible and irreversible fouling when compared to biofiltration. The reduction of

turbidity and alum residual post filtration decreased reversible and irreversible fouling

confirming the importance of the biopolymer and turbidity interaction. Both coagulation

91

and biofiltration have the potential to be applied as pre-treatments for UF with preference

given to coagulation because of its ability to decrease DBP precursors.

Figure 5.11: Normalized UF resistance profile comparing coagulation (29.1-30.1 mg/L

alum) to biofiltration (Biofiltration: November 27th,2013, T= 3ºC. Pre-filter alum:

December 2nd, 2013, T= 4.5 ºC)

Figure 5.12: Normalized UF resistance profile comparing coagulation (30.0 mg/L alum)

with filtration to biofiltration (Biofiltration: November 8th,2013, T= 9 ºC. Post-filter

alum: November 8th, 2013, T= 9.5 ºC)

5.4 Summary

Biofiltration as a stand-alone treatment was found to be effective for removal of

turbidity (71±11%), biopolymers (20±4%), as well as THM (13±11%) and HAA (9±13%)

0

2

4

6

8

0 5 10 15 20 25 30

No

rmal

ized

mem

bra

ne

resi

stan

ce @

20

˚C (

×1

01

2 m

-1)

Permeation Time (hr)

Pre-filter (Alum) Biofiltration

0

2

4

6

8

0 5 10 15 20 25 30

Norm

aliz

ed m

embra

ne

resi

stan

ce @

20˚C

10

12

m-1

)

Permeation Time (hours)

Post-filter (Alum) Biofiltration

92

precursor reduction, and to provide irreversible UF fouling control. However, it fell short

when compared to coagulation where significantly higher DOC (up to 50%), biopolymer

(42-72%), and DBP precursor removals (48-76%) were observed. Biofiltration provided

similar irreversible fouling control when compared to coagulation, however unlike

biofiltration coagulation was also able to control reversible fouling. Conventional

treatment prior to ultrafiltration provided superior irreversible and reversible fouling

control, which emphasizes the importance of turbidity along with biopolymers when

considering UF fouling control strategies. From this study, it is evident that by allowing

conventional filters to work in a biological mode, significant improvements in NOM

removal could be achieved as well as UF fouling control. Table 5.3 summarizes the study

findings.

Table 5.3: Impact of biofiltration and conventional treatment on the removal of the

variables in this study (↑=positive impact, ↓=adverse impact, ●=no impact, =

statistically significant, = not statistically significant difference at the 95% confidence

level, Where symbols not present= statistical analysis was not possible, NS= Not

Sampled)

Biofilter

Post Full-scale

Plant Filter

#11

Pre-

filter

(Alum)

Post-

filter

(Alum)

Pre-

filter

(PACl)

Post-

filter

(PACl)

Turbidity ↑ ↑ ↓ ↑ ↓ ↑

DOC ↑ ↑ ↑ ↑ ↑ ↑

BP ↑ ↑ ↑ ↑ ↑ ↑

HS ● ↑ ↑ ↓ ↑ ↑

THMs ↑ ↑ ↑ ↑ ↑ ●

HAAs ↑ ↑ ↑ ↑ ↑ ↑ AOX ↑ ↑ ↑ ↑ ↑ ↑ Irreversible

UF fouling

↑ NS ↑ ↑ NS NS

Reversible

UF fouling

● NS ↑ ↑ NS NS

93

6. Overall Summary, Conclusions and

Recommendations

6.1 Summary

The impact of phosphorus, nitrogen, peroxide, and alum addition on biofiltration to

control NOM (DOC, UV254, and LC-OCD), DBP precursors (THMs, HAAs, and AOX),

and UF fouling was investigated via a pilot-scale experiment. In addition, the impact of

media type (anthracite vs. GAC) was examined. The study was conducted using parallel

biofiltration treatment trains with two large (D= 15.24 cm) glass, and four small (D= 7.62

cm) acrylic biofilters which were installed at the Peterborough Water Treatment Plant.

Nutrients (phosphorus and nitrogen), alum, and hydrogen peroxide were incrementally

added to each biofilter influent during three stages, each ranging from 24 to 69 days.

Furthermore, a parallel comparison to conventional treatment (alum and PACl) was

conducted.

6.2 Conclusions

Results indicated that biofiltration was capable of reducing turbidity, biopolymers

(major UF foulant), and DBP precursors, as well as UF irreversible fouling. Nutrient

addition (phosphorus and nitrogen at C:N:P ratios of 100:30:2, 100:40:2, and 100:40:20)

did not improve biofilter performance in terms of the measured parameters. Peroxide (at

0.1 and 0.5) increased biofilter runtime without adversely affecting removal of NOM or

DBP precursors. However, Peroxide addition at 1 mg/L decreased THM and AOX

formation potentials, and resulted in higher UF irreversible fouling when compared to a

control. The addition of low alum dosages (< 0.5 mg/L) was found to improve biopolymer

removal and DBP precursor reduction. It also resulted in lower UF irreversible fouling,

with a higher impact at low doses (0.1 and 0.25 mg/L) and at temperatures > 15˚C.

Exhausted GAC provided higher DOC, and DBP precursor removal when compared to

anthracite.

When compared to biofiltration, conventional treatment resulted in a significant

reduction in NOM, DBP precursors and UF reversible and irreversible fouling. However,

94

coagulation without filtration provided similar UF irreversible fouling control when

compared to biofiltration. Furthermore, biofiltration was observed to improve biopolymer

and DBP precursor removal when compared to conventional filtration which shows its

potential to be applied post coagulation. Biofiltration was more effective at temperatures

>10˚C, but even at low temperatures (3-10˚C), provided NOM and turbidity reduction as

well as UF fouling mitigation.

6.3 Recommendations

The results of this work warrant further investigation as follows:

1- While this work has shown the potential ability of alum addition (< 0.5 mg/L) to

improve UF fouling control, and the adverse impact of peroxide addition (1 mg/L)

on irreversible fouling, duplicate UF runs could not be performed. Therefore,

further trials are required to validate these findings. In addition, more dosages and

the use of other cationic and polymeric coagulants should be explored.

2- This study only explored the application of biofiltration as a single process in

comparison to conventional treatment, while the application of biofiltration as a

part of treatment process such as ozone/biofiltration/UF or coagulation/biofiltration

is of interest.

3- The study was conducted at temperatures ranging from 3-25˚C; hence, further

investigation during winter conditions (< 3˚C) to evaluate biofilter performance

year-round. Also, the effectiveness of coagulation as a pre-treatment for UF at these

temperatures should be directly compared to biofiltration.

4- The focus of this work was biofiltration optimization; however, further

characterization of membrane foulants, their interaction with each other and with

the membrane surface, and their removal during biofiltration is required.

5- The combination of the enhancement methods applied in this study along with pH

adjustment should be studied.

59

7. References

Ahmad, R., A. Amirtharajah, A. Al-Shawwa, and P. M. Huck. 1998. Effects of backwashing on

biological filters. Journal / American Water Works Association, 90(12), 62-73

Allen, M. J., S. C. Edberg, and D. J. Reasoner. 2004. Heterotrophic plate count bacteria - What is

their significance in drinking water? International Journal of Food Microbiology, 92(3),

265-274

Andersson, A., P. Laurent, A. Kihn, M. Prévost Michèle, and P. Servais. 2001. Impact of

temperature on nitrification in biological activated carbon (BAC) filters used for drinking

water treatment. Water Research, 35(12), 2923-2934

Baghoth, S. A., M. Dignum, A. Grefte, J. Kroesbergen, and G. L. Amy. 2009. Characterization

of NOM in a drinking water treatment process train with no disinfectant residual. Water

Science and Technology: Water Supply, 9(4), 379-386

Barrett, S. E., S. W. Krasner, and G. L. Amy. 2000. Natural organic matter and disinfection by-

products: Characterization and control in drinking water - An overview. ACS Symposium

Series, 761, 2-14

Basu, O. D. and P. M. Huck. 2004. Integrated biofilter-immersed membrane system for the

treatment of humic waters. Water Research, 38(3), 655-662

Boon, N., B. F. G. Pycke, M. Marzorati, and F. Hammes. 2011. Nutrient gradients in a granular

activated carbon biofilter drives bacterial community organization and dynamics. Water

Research, 45(19), 6355-6361

Brinkman, B. M. and R. M. Hozalski. 2011. Temporal variation of NOM and its effects on

membrane treatment. Journal / American Water Works Association, 103(2), 98-106

Carlson, K. H. and G. L. Amy. 2001. Ozone and biofiltration optimization for multiple

objectives. Journal / American Water Works Association, 93(1), 88-98

96

Chaiket, T., P. C. Singer, A. Miles, M. Moran, and C. Pallotta. 2002. Effectiveness of

coagulation, ozonation, and biofiltration in controlling DBPs. Journal / American Water

Works Association, 94(12), 81-95

Chellam, S., J. G. Jacangelo, T. P. Bonacquisti, and B. A. Schauer. 1997. Effect of pretreatment

on surface water nanofiltration. Journal / American Water Works Association, 89(10), 77-89

Chen, C., X. Zhang, L. Zhu, W. He, and H. Han. 2011. Changes in different organic matter

fractions during conventional treatment and advanced treatment. Journal of Environmental

Sciences, 23(4), 582-586

Choi, K. Y. and B. A. Dempsey. 2004. In-line coagulation with low-pressure membrane

filtration. Water research, 38(19), 4271-4281

Chow, C. W. K., R. Fabris, and M. Drikas. 2004. A rapid fractionation technique to characterise

natural organic matter for the optimisation of water treatment processes. Journal of Water

Supply: Research and Technology - AQUA, 53(2), 85-92

Crittenden, J. C., R. R. Trussell, D. W. Hand, K. J. Howe, and G. Tchobanoglous. 2012. Water

treatment: principles and design.

Croft, J., 2012. Natural Organic Matter Characterization of Different Source and Treated Waters;

Implications for Membrane Fouling Control. University of Waterloo

Diemert, S. and R. C. Andrews. 2013. The impact of alum coagulation on pharmaceutically

active compounds, endocrine disrupting compounds and natural organic matter. Water

Science and Technology: Water Supply, 13(5), 1348-1357

Drioli, E., 2009. Membrane Operations: Innovative Separations and Transformations. WILEY-

VCH Verlag GmbH & Co. KGaA

Dussert, B. W. and W. G. Tramposch. 1997. Impact of support media on the biological treatment

of ozonated drinking water. Ozone: Science and Engineering, 19(2), 97-108

97

Edzwald, J. K., W. C. Becker, and S. J. Tambini. 1987. Organics, polymers and performance in

direct filtration. Journal of Environmental Engineering, 113(1), 167-185

Edzwald, J. K. and J. E. Tobiason. 1999. Enhanced coagulation: US requirements and a broader

view. Water Science and Technology, 40(9), 63-70

Egashira, K., K. Ito, and Y. Yoshiy. 1992. Removal of musty odor compound in drinking water

by biological filter. Water Science and Technology, 25(2), 307-314

Elhadi, S. L. N., P. M. Huck, and R. M. Slawson. 2006. Factors affecting the removal of geosmin

and MIB in drinking water biofilters. Journal / American Water Works Association, 98(8),

108-119

Elhadi, S., P. Huck, and R. Slawson. 2004. Removal of geosmin and 2-methylisoborneol by

biological filtration. Water Science and Technology, 49(9), 273-280

Ellis, K. V., 1985. Slow sand filtration. Critical Reviews in Environmental Control, 15(4), 315-

354

Emelko, M. B., P. M. Huck, B. M. Coffey, and E. F. Smith. 2006. Effects of media, backwash,

and temperature on full-scale biological filtration. Journal / American Water Works

Association, 98(12), 61-73

EPA, 2009. National Primary Drinking Water Regulations. US EPA, 2013, 6

Escobar, I. C. and A. A. Randall. 2001. Assimilable organic carbon (AOC) and biodegradable

dissolved organic carbon (BDOC): Complementary measurements. Water Research, 35(18),

4444-4454

Escobar, I. C., S. Hong, and A. A. Randall. 2000. Removal of assimilable organic carbon and

biodegradable dissolved organic carbon by reverse osmosis and nanofiltration membranes.

Journal of Membrane Science, 175(1), 1-17

98

Gao, W., H. Liang, J. Ma, M. Han, Z. L. Chen, Z. S. Han, et al. 2011. Membrane fouling control

in ultrafiltration technology for drinking water production: A review. Desalination, 272(1-

3), 1-8

Geismar, N., P. R. Bérubé, and B. Barbeau. 2012. Variability and limits of the unified membrane

fouling index: Application to the reduction of low-pressure membrane fouling by ozonation

and biofiltration. Desalination and Water Treatment, 43(1-3), 91-101

Gibert, O., B. Lefèvre, M. Fernández, X. Bernat, M. Paraira, and M. Pons. 2013. Fractionation

and removal of dissolved organic carbon in a full-scale granular activated carbon filter used

for drinking water production. Water Research, 47(8), 2821-2829

Hallè, C., 2010. Biofiltration in drinking water treatment: reduction of membrane fouling and

biodegradation of organic trace contaminants.

Hallé, C., P. M. Huck, S. Peldszus, J. Haberkamp, and M. Jekel. 2009. Assessing the

performance of biological filtration as pretreatment to low pressure membranes for drinking

water. Environmental Science and Technology, 43(10), 3878-3884

Hammes, F., M. Berney, Y. Wang, M. Vital, O. Koster, and T. Egli. 2008. Flow-cytometric total

bacterial cell counts as a descriptive microbiological parameter for drinking water treatment

processes. Water Research, 42(1-2), 269-277

Hammes, F., S. Velten, T. Egli, and T. Juhna. 2011. Biotreatment of drinking water.(6.4), 517-

530

Health Canada, 2009. Guidelines for Canadian Drinking Water Quality: Guideline Technical

Document: Trihalomethanes., 2013(December)

Henderson, R. K., N. Subhi, A. Antony, S. J. Khan, K. R. Murphy, G. L. Leslie, et al. 2011.

Evaluation of effluent organic matter fouling in ultrafiltration treatment using advanced

organic characterisation techniques. Journal of Membrane Science, 382(1-2), 50-59

Howe, K. J. and M. M. Clark. 2006. Effect of coagulation pretreatment on membrane filtration

performance. Journal / American Water Works Association, 98(4), 133-146+12

99

Howe, K. J. and M. M. Clark. 2002. Fouling of microfiltration and ultrafiltration membranes by

natural waters. Environmental Science and Technology, 36(16), 3571-3576

Hozalski, R. M., E. J. Bouwer, and S. Goel. 1999. Removal of natural organic matter (NOM)

from drinking water supplies by ozone-biofiltration. Water Science and Technology, 40(9),

157-163

Hozalski, R. M., S. Goel, and E. J. Bouwer. 1995. TOC removal in biological filters. Journal /

American Water Works Association, 87(12), 40-54

Huang, G., F. Meng, X. Zheng, Y. Wang, Z. Wang, H. Liu, et al. 2011. Biodegradation behavior

of natural organic matter (NOM) in a biological aerated filter (BAF) as a pretreatment for

ultrafiltration (UF) of river water. Applied Microbiology and Biotechnology, 90(5), 1795-

1803

Huang, H., N. Lee, T. Young, A. Gary, J. C. Lozier, and J. G. Jacangelo. 2007. Natural organic

matter fouling of low-pressure, hollow-fiber membranes: Effects of NOM source and

hydrodynamic conditions. Water Research, 41(17), 3823-3832

Huang, H., K. Schwab, and J. G. Jacangelo. 2009. Pretreatment for low pressure membranes in

water treatment: A review. Environmental Science and Technology, 43(9), 3011-3019

Huber, S. A., A. Balz, and M. Abert. 2011a. New method for urea analysis in surface and tap

waters with LC-OCD-OND (liquid chromatography-organic carbon detection-organic

nitrogen detection). Journal of Water Supply: Research and Technology - AQUA, 60(3),

159-166

Huber, S. A., A. Balz, M. Abert, and W. Pronk. 2011b. Characterisation of aquatic humic and

non-humic matter with size-exclusion chromatography – organic carbon detection – organic

nitrogen detection (LC-OCD-OND). Water Research, 45(2), 879-885

Huber, S. A. and F. H. Frimmel. 1994. Direct gel chromatographic characterization and

quantification of marine dissolved organic carbon using high-sensitivity DOC detection.

Environmental Science and Technology, 28(6), 1194-1197

100

Huck, P. M., S. Peldszus, C. Hallé, H. Ruiz, X. Jin, M. Van Dyke, et al. 2011. Pilot scale

evaluation of biofiltration as an innovative pre-treatment for ultrafiltration membranes for

drinking water treatment. Water Science and Technology: Water Supply, 11(1), 23-29

Huck, P. M. and M. M. Sozański. 2008. Biological filtration for membrane pre-treatment and

other applications: Towards the development of a practically-oriented performance

parameter. Journal of Water Supply: Research and Technology - AQUA, 57(4), 203-224

Jacangelo, J. G., J. DeMarco, D. M. Owen, and S. J. Randtke. 1995. Selected processes for

removing NOM: An overview. Journal / American Water Works Association, 87(1), 64-77

Jarvis, P., B. Jefferson, and S. A. Parsons. 2004. Characterising natural organic matter flocs.

Water Science and Technology: Water Supply, 4(4), 79-87

Jarvis, P., B. Jefferson, and S. A. Parsons. 2005. Breakage, regrowth, and fractal nature of

natural organic matter flocs. Environmental Science and Technology, 39(7), 2307-2314

Kennedy, M. D., H. K. Chun, V. A. Quintanilla Yangali, B. G. J. Heijman, and J. C. Schippers.

2005. Natural organic matter (NOM) fouling of ultrafiltration membranes: fractionation of

NOM in surface water and characterisation by LC-OCD. Desalination, 178(1-3), 73-83

Kim, J. S., M. J. Kim, J. Vilagos, B. Pickard, P. Lowe, J. Gordy, et al. 2009. Chloramine auto

decomposition and stability of BAC filtered water. Water Quality Technology Conference

and Exposition 2009, 3360-3380

Klymenko, N. A., I. P. Kozyatnyk, and L. A. Savchyna. 2010. Removing of fulvic acids by

ozonation and biological active carbon filtration. Water Research, 44(18), 5316-5322

Lauderdale, C., P. Chadik, M. Kirisits, and J. Brown. 2012. Engineered biofiltration: enhanced

biofilter performancethrough nutrient and peroxideaddition. Journal American Water Works

Association, 104(73), E289-E309

Lee, N., G. Amy, J. P. Croue, and H. Buisson. 2004. Identification and understanding of fouling

in low-pressure membrane (MF/UF) filtration by natural organic matter (NOM). Water

Research, 38(20), 4511-4523

101

Leenheer, J. A. and J. P. Croué. 2003. Characterizing aquatic dissolved organic matter.

Environmental Science and Technology, 37(1), 18A-26A

Leenheer, J. A., J. P. Croué, M. Benjamin, G. V. Korshin, C. J. Hwang, A. Bruchet, et al. 2000.

Comprehensive isolation of natural organic matter from water for spectral characterizations

and reactivity testing. ACS Symposium Series, 761, 68-83

Li, K., H. Liang, T. J. Ye, W. X. Luo, R. -. Lai, X. Z. Lin, et al. 2013. Effect of in-line

coagulation on treated water quality and membrane fouling for immersed ultrafiltration.

Beijing Gongye Daxue Xuebao/Journal of Beijing University of Technology, 39(2), 287-291

Liu, X., P. M. Huck, and R. M. Slawson. 2001. Factors affecting drinking water biofiltration.

Journal / American Water Works Association, 93(12), 90-101

Madrid, R. E. and C. J. Felice. 2005. Microbial biomass estimation. Critical reviews in

biotechnology, 25(3), 97-112

Magic-Knezev, A. and D. van der Kooij. 2004. Optimisation and significance of ATP analysis

for measuring active biomass in granular activated carbon filters used in water treatment.

Water Research, 38(18), 3971-3979

Malievialle, J., P. Odendaal, and M. Wiesner. 1996. Water treatment membrane processes.

Matilainen, A., E. T. Gjessing, T. Lahtinen, L. Hed, A. Bhatnagar, and M. Sillanpää. 2011. An

overview of the methods used in the characterisation of natural organic matter (NOM) in

relation to drinking water treatment. Chemosphere, 83(11), 1431-1442

Matilainen, A., M. Vepsäläinen, and M. Sillanpää. 2010. Natural organic matter removal by

coagulation during drinking water treatment: A review. Advances in Colloid and Interface

Science, 159(2), 189-197

McDowall, B., D. Hoefel, G. Newcombe, C. P. Saint, and L. Ho. 2009. Enhancing the

biofiltration of geosmin by seeding sand filter columns with a consortium of geosmin-

degrading bacteria. Water Research, 43(2), 433-440

102

Miltner, R. J., R. S. Summers, and J. Z. Wang. 1995. Biofiltration performance: part 2, effect of

backwashing. Journal / American Water Works Association, 87(12), [d]64-70

Moll, D. M., R. S. Summers, and A. Breen. 1998. Microbial characterization of biological filters

used for drinking water treatment. Applied and Environmental Microbiology, 64(7), 2755-

2759

Moll, D. M., R. S. Summers, A. C. Fonseca, and W. Matheis. 1999. Impact of temperature on

drinking water biofilter performance and microbial community structure. Environmental

Science and Technology, 33(14), 2377-2382

Mosqueda-Jimenez, D. B., P. M. Huck, and O. D. Basu. 2008. Fouling characteristics of an

ultrafiltration membrane used in drinking water treatment. Desalination, 230(1-3), 79-91

Neubrand, W., S. Vogler, M. Ernst, and M. Jekel. 2010. Lab and pilot scale investigations on

membrane fouling during the ultrafiltration of surface water. Desalination, 250(3), 968-972

Nguyen, S. T. and F. A. Roddick. 2010. Effects of ozonation and biological activated carbon

filtration on membrane fouling in ultrafiltration of an activated sludge effluent. Journal of

Membrane Science, 363(1-2), 271-277

Okei, I., Y. Akagami, K. shimizu, Q. Xue, C. Sato, M. Utsumi, et al. 2009. Removal

characteristics of musty odor "geosmin" by biofiltration using various carriers.

Paar, H., J. Benecke, M. Ernst, and M. Jekel. 2011. Pre-coagulation and ultrafiltration of effluent

impaired surface water for phosphorus removal and fouling control. Water Science and

Technology: Water Supply, 11(2), 211-218

Peiris, R. H., C. Hallé, H. Budman, C. Moresoli, S. Peldszus, P. M. Huck, et al. 2010a.

Identifying fouling events in a membrane-based drinking water treatment process using

principal component analysis of fluorescence excitation-emission matrices. Water Research,

44(1), 185-194

Peiris, R. H., H. Budman, C. Moresoli, and R. L. Legge. 2010b. Understanding fouling

behaviour of ultrafiltration membrane processes and natural water using principal

103

component analysis of fluorescence excitation-emission matrices. Journal of Membrane

Science, 357(1-2), 62-72

Peldszus, S., J. Benecke, M. Jekel, and P. M. Huck. 2012. Direct biofiltration pretreatment for

fouling control of ultrafiltration membranes. Journal - American Water Works Association,

104(7), E430-E445

Peldszus, S., C. Hallé, R. H. Peiris, M. Hamouda, X. Jin, R. L. Legge, et al. 2011. Reversible and

irreversible low-pressure membrane foulants in drinking water treatment: Identification by

principal component analysis of fluorescence EEM and mitigation by biofiltration

pretreatment. Water Research, 45(16), 5161-5170

Penru, Y., X. Simon, A. Guastalli, S. Esplugas, J. Llorens, and S. Baig. 2011. Characterization of

natural organic matter from Mediterranean coastal seawater. IWA Specialty Conference on

Natural Organic Matter, CA, USA

Persson, F., G. Heinicke, W. Uhl, T. Hedberg, and M. Hermansson. 2006. Performance of direct

biofiltration of surface water for reduction of biodegradable organic matter and biofilm

formation potential. Environmental Technology, 27(9), 1037-1045

Plakas, K. V. and A. J. Karabelas. 2012. Removal of pesticides from water by NF and RO

membranes - A review. Desalination, 287, 255-265

Prendiville, P. W. and P. Y. Chung. 1992. The reduction of aluminium carryover into

distribution systems by the careful control of coagulation and filtration. Water Supply,

10(4), 77-81

Qian, J. and K. Mopper. 1996. Automated high-performance, high-temperature combustion total

organic carbon analyzer. Analytical Chemistry, 68(18), 3090-3097

Rana, D., B. Scheier, R. M. Narbaitz, T. Matsuura, S. Tabe, S. Y. Jasim, et al. 2012. Comparison

of cellulose acetate (CA) membrane and novel CA membranes containing surface

modifying macromolecules to remove pharmaceutical and personal care product

micropollutants from drinking water. Journal of Membrane Science, 409-410, 346-354

104

Richardson, M. L. and J. M. Bowron. 1985. The fate of pharmaceutical chemicals in the aquatic

environment. Journal of Pharmacy and Pharmacology, 37(1), 1-12

Richardson, S. D., M. J. Plewa, E. D. Wagner, R. Schoeny, and D. M. DeMarini. 2007.

Occurrence, genotoxicity, and carcinogenicity of regulated and emerging disinfection by-

products in drinking water: A review and roadmap for research. Mutation Research-Reviews

in Mutation Research, 636(1-3), 178-242

Richardson, S. D., 2009. Water analysis: emerging contaminants and current issues. Analytical

Chemistry, 81(12), 4645-4677

Richardson, S. D. and T. A. Ternes. 2011. Water analysis: emerging contaminants and current

issues. Analytical Chemistry, 83(12), 4614-4648

Rittmann, B. E., D. Stilwell, and A. Ohashi. 2002. The transient-state, multiple-species biofilm

model for biofiltration processes. Water Research, 36(9), 2342-2356

Rittmann, B. E., 2001. Environmental biotechnology : principles and applications.

Rizzo, L., V. Belgiorno, and S. Meriç. 2004. Organic THMs precursors removal from surface

water with low TOC and high alkalinity by enhanced coagulation. Water Science and

Technology: Water Supply, 4(5-6), 103-111

Sang, J., X. Zhang, L. Li, and Z. Wang. 2003. Improvement of organics removal by bio-ceramic

filtration of raw water with addition of phosphorus. Water Research, 37(19), 4711-4718

Siebel, E., Y. Wang, T. Egli, and F. Hammes. 2008. Correlations between total cell

concentration, total adenosine tri-phosphate concentration and heterotrophic plate counts

during microbial monitoring of drinking water. Drinking Water Engineering and Science,

1(1), 1-6

Singer, P. C., 1999. Humic Substances as Precursors for Potentially Harmful Disinfection By-

Products. Water Science and Technology, 40(9), 25-30

105

Singh, R., 2006. Hybrid membrane systems for water purification : technology, systems design

and operation.

Srinivasan, R. and G. A. Sorial. 2011. Treatment of taste and odor causing compounds 2-methyl

isoborneol and geosmin in drinking water: A critical review. Journal of Environmental

Sciences-China, 23(1), 1-13

Stedmon, C. A., B. Seredyńska-Sobecka, R. Boe-Hansen, N. Le Tallec, C. K. Waul, and E.

Arvin. 2011. A potential approach for monitoring drinking water quality from groundwater

systems using organic matter fluorescence as an early warning for contamination events.

Water Research, 45(18), 6030-6038

Tian, J. Y., Z. L. Chen, H. Liang, X. Li, Z. Z. Wang, and G. B. Li. 2009. Comparison of

biological activated carbon (BAC) and membrane bioreactor (MBR) for pollutants removal

in drinking water treatment. Water Science and Technology, 60(6), 1515-1523

Tihomirova, K., J. Rubulis, and T. Juhna. 2010. Changes of NOM fractions during conventional

drinking water treatment process in Riga, Latvia. Water Science and Technology: Water

Supply, 10(2), 157-163

Tsujimoto, W., H. Kimura, T. Izu, and T. Irie. 1998. Membrane filtration and pre-treatment by

GAC. Desalination, 119(1-3), 323-326

Tufenkji, N., J. N. Ryan, and M. Elimelech. 2002. Bank filtration. Environmental Science and

Technology, 36(21), 422A-428A

Urfer, D. and P. M. Huck. 1997. Effects of hydrogen peroxide residuals on biologically active

filters. Ozone: Science and Engineering, 19(4), 371-386

Urfer, D. and P. M. Huck. 2000. A study of the impacts of periodic ozone residuals on

biologically active filters. Ozone: Science and Engineering, 22(1), 77-97

Urfer, D., 1998. Effects of Oxidants on drinking water biofilters. University of Waterloo

106

Urfer, D. and P. M. Huck. 2001. Measurement of biomass activity in drinking water biofilters

using a respirometric method. Water Research, 35(6), 1469-1477

Vahala, R., V. Moramarco, R. M. Niemi, J. Rintala, and R. Laukkanen. 1998. The effects of

nutrients on natural organic matter (NOM) removal in biological activated carbon (BAC)

filtration. Acta Hydrochimica et Hydrobiologica, 26(3), 196-199

Van Benschoten, J. E. and J. K. Edzwald. 1990. Chemical aspects of coagulation using

aluminum salts - I. Hydrolytic reactions of alum and polyaluminum chloride. Water

research, 24(12), 1519-1526

Van der kooij, D., 1992. Assimilable organic-carbon as an indicator of bacterial regrowth.

Journal American Water Works Association, 84(2), 57-65

Velten, S., M. Boller, O. Köster, J. Helbing, H. -. Weilenmann, and F. Hammes. 2011a.

Development of biomass in a drinking water granular active carbon (GAC) filter. Water

Research, 45(19), 6347-6354

Velten, S., F. Hammes, M. Boller, and T. Egli. 2007. Rapid and direct estimation of active

biomass on granular activated carbon through adenosine tri-phosphate (ATP) determination.

Water Research, 41(9), 1973-1983

Velten, S., D. R. U. Knappe, J. Traber, H. Kaiser, U. von Gunten, M. Boller, et al. 2011b.

Characterization of natural organic matter adsorption in granular activated carbon adsorbers.

Water research, 45(13), 3951-3959

Volk, C. J. and M. W. Lechevallier. 2002. Effects of conventional treatment on AOC and BDOC

levels. Journal / American Water Works Association, 94(6), 112-123

Wahman, D. G., L. E. Katz, and G. E. Speitel. 2011. Performance and biofilm activity of

nitrifying biofilters removing trihalomethanes. Water Research, 45(4), 1669-1680

Wang, J. and X. C. Wang. 2006. Ultrafiltration with in-line coagulation for the removal of

natural humic acid and membrane fouling mechanism. Journal of Environmental Sciences

(China), 18(5), 880-884

107

Wang, J. Z., R. S. Summers, and R. J. Miltner. 1995. Biofiltration performance: part 1,

relationship to biomass. Journal / American Water Works Association, 87(12), [d]55-63

Wang, Y., F. Hammes, K. De Roy, W. Verstraete, and N. Boon. 2010. Past, present and future

applications of flow cytometry in aquatic microbiology. Trends in Biotechnology, 28(8),

416-424

Wassink, J. K., R. C. Andrews, R. H. Peiris, and R. L. Legge. 2011. Evaluation of fluorescence

excitation-emission and LC-OCD as methods of detecting removal of NOM and DBP

precursors by enhanced coagulation. Water Science and Technology: Water Supply, 11(5),

621-630

Wert, E. C., J. J. Neemann, D. J. Rexing, and R. E. Zegers. 2008. Biofiltration for removal of

BOM and residual ammonia following control of bromate formation. Water Research, 42(1-

2), 372-378

Wray, H. E., R. C. Andrews, and P. R. and Bérubé. 2014. Ultrafiltration organic fouling control:

comparison of air sparging and coagulation. Journal - American Water Works Association

Wu, F. C., R. D. Evans, and P. J. Dillon. 2003. Separation and characterization of NOM by high-

performance liquid chromatography and on-line three-dimensional excitation emission

matrix fluorescence detection. Environmental Science and Technology, 37(16), 3687-3693

Wu, S. H., B. Z. Dong, T. J. Qiao, and J. S. Zhang. 2008. Effect of a biological activated carbon

filter on particle counts. Journal of Zhejiang University: Science A, 9(11), 1576-1581

Yang, B. M., J. K. Liu, C. C. Chien, R. Y. Surampalli, and C. M. Kao. 2011. Variations in AOC

and microbial diversity in an advanced water treatment plant. Journal of Hydrology, 409(1-

2), 225-235

Yavich, A. A., K. Lee, K. Chen, L. Pape, and S. J. Masten. 2004. Evaluation of biodegradability

of NOM after ozonation. Water Research, 38(12), 2839-2846

Yu, W., G. Li, Y. Xu, and X. Yang. 2009. Breakage and re-growth of flocs formed by alum and

PACl. Powder Technology, 189(3), 439-443

108

Zhang, D., W. Li, K. Wang, L. Zhang, and H. Gong. 2010. Bacterial community dynamics and

its effects during biological activited carbon filter process for drinking water treatment. 2nd

International Conference on Chemical, Biological and Environmental Engineering,

Proceedings, 137-142

Zhang, D., W. Li, S. Zhang, M. Liu, X. Zhao, and X. Zhang. 2011a. Bacterial community and

function of biological activated carbon filter in drinking water treatment. Biomedical and

Environmental Sciences, 24(2), 122-131

Zhang, W., X. Zhang, C. Chen, J. Wang, and Y. Liu. 2011b. Pilot study on hybrid biological

filter-coagulation-submerged ultrafiltration membrane for drinking water treatment. 5th

International Conference on Bioinformatics and Biomedical Engineering, iCBBE 2011

Zheng, L., N. Gao, Y. Deng, E. Du, M. Sui, and S. Liu. 2011. The effect of backwashing in the

structure of microbial community on biological activated carbon (BAC) in a water treatment

plant. Fresenius Environmental Bulletin, 20(7A), 1741-1748

Zheng, X., M. Ernst, P. M. Huck, and M. Jekel. 2010. Biopolymer fouling in dead-end

ultrafiltration of treated domestic wastewater. Water Research, 44(18), 5212-5221

Zheng, X., S. Plume, M. Ernst, J. -. Croué, and M. Jekel. 2012. In-line coagulation prior to UF of

treated domestic wastewater - foulants removal, fouling control and phosphorus removal.

Journal of Membrane Science, 403-404, 129-139

Zhu, I. X., T. Getting, and D. Bruce. 2010. Review of biologically active filters in drinking water

applications. Journal / American Water Works Association, 102(12), 67-77

109

8. Appendices

8.1 Appendix A (Calibration Curves)

Figure 8.1: Sample DOC calibration curve

Figure 8.2: Sample THM calibration curve (June 2013)

y = 4954.3x + 6133.7

R² = 0.9997

0

10000

20000

30000

40000

50000

60000

0 2 4 6 8 10 12

Are

a C

ounts

DOC (mg/L)

y = 0.3634x + 2.7418

R² = 0.9964

y = 2.8648x - 5.9451

R² = 0.9945

y = 3.3806x - 5.3124

R² = 0.9967

y = 1.5847x + 0.146

R² = 0.9992

-50

0

50

100

150

200

250

0 10 20 30 40 50 60 70

Are

a C

ounts

Concentration (μg/L)

TCM BDCM CDBM TBM

221

Figure 8.3: HAA calibration curves (June-September 2013)

y = 0.0045x - 0.0011

R² = 0.9965

y = 0.0412x + 0.0353

R² = 0.9992

y = 0.0456x + 0.0545

R² = 0.9986

y = 0.0917x + 0.2212

R² = 0.9896

y = 0.084x + 0.1187

R² = 0.9969

y = 0.0649x + 0.2813

R² = 0.9732

y = 0.059x - 0.1198

R² = 0.9953

y = 0.0316x - 0.0806

R² = 0.9888

y = 0.0167x - 0.0687

R² = 0.931

0

1

2

3

4

5

6

7

0 10 20 30 40 50 60 70

Are

a C

ounts

Concentration (ug/L)

MCAA MBAA DCAA TCAA BCAA DBAA BDCAA CDBAA TBAA

222

Figure 8.4: Sample PS calibration curve (June 2013)

Figure 8.5: Sample Pr. calibration curve (June 2013)

y = 0.0125x + 0.014

R² = 0.9993

0.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 20 40 60 80 100 120

Abso

rban

ce (

OD

) at

492nm

Glucose concentration (ug/mL)

y = 0.0035x + 0.0014

R² = 0.9991

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1.0

0 50 100 150 200 250 300

Ab

sorb

ance

(O

D)

at 7

50 n

m

BSA concentration in ug/mL

221

8.2 Appendix B (Raw Data)

Table 8.1: Water quality and DOC results (Raw) (- indicate that the data is not available)

Sample Date pH Turbidity (NTU) UV254 (cm-1) DOC (mg/L)

Alum CF 4-Jun-13 7.3 0.150 0.053 3.5

Alum CF 4-Jun-13 - 0.205 - 3.5

PACl CF 4-Jun-13 7.7 0.120 0.055 3.6

PACl CF 4-Jun-13 - 0.170 - 3.6

FSP 4-Jun-13 7.9 0.128 0.051 -

FSP 4-Jun-13 - 0.157 - -

Alum SW 4-Jun-13 7.1 0.196 0.053 3.7

Alum SW 4-Jun-13 - 0.192 - 3.7

PACl SW 4-Jun-13 7.9 0.225 0.057 3.9

PACl SW 4-Jun-13 - 0.247 - 3.9

BF1 4-Jun-13 7.8 0.240 0.141 6.2

BF1 4-Jun-13 - 0.253 - 6.3

BF2 4-Jun-13 7.8 0.226 0.140 6.2

BF2 4-Jun-13 - 0.222 - 6.1

BF3 4-Jun-13 7.8 0.384 0.143 6.5

BF3 4-Jun-13 - 0.384 - 6.5

BF4 4-Jun-13 7.9 0.535 0.151 6.4

BF4 4-Jun-13 - 0.578 - 6.5

BF5 4-Jun-13 8.2 0.383 0.151 6.5

BF5 4-Jun-13 - 0.400 - 6.5

RW 4-Jun-13 8.2 0.535 0.141 7.0

RW 4-Jun-13 - 0.592 - 7.0

Alum CF 24-Jun-13 - - - 3.1

Alum CF 24-Jun-13 - - - 3.1

PACl CF 24-Jun-13 - - - 3.3

PACl CF 24-Jun-13 - - - 3.3

FSP 24-Jun-13 - - - -

FSP 24-Jun-13 - - - -

Alum SW 24-Jun-13 - - - 3.4

Alum SW 24-Jun-13 - - - 3.3

PACl SW 24-Jun-13 - - - 3.4

PACl SW 24-Jun-13 - - - 3.4

BF1 24-Jun-13 - - - 5.2

BF1 24-Jun-13 - - - 5.3

BF2 24-Jun-13 - - - 5.3

113

Table 8.1: Water Quality and DOC results (cont.) (- indicate that the data is not available)

Sample Date pH Turbidity (NTU) UV254 (cm-1) DOC (mg/L)

BF2 24-Jun-13 - - - 5.4

BF3 24-Jun-13 - - - 5.4

BF3 24-Jun-13 - - - 5.3

BF4 24-Jun-13 - - - 5.5

BF4 24-Jun-13 - - - 5.3

BF5 24-Jun-13 - - - 5.4

BF5 24-Jun-13 - - - 5.5

BF6 24-Jun-13 - - - 5.4

BF6 24-Jun-13 - - - 5.4

RW 24-Jun-13 - - - 5.4

RW 24-Jun-13 - - - 5.6

FSP 24-Jul-13 7.5 0.105 0.040 2.9

FSP 24-Jul-13 7.5 0.115 0.050 2.9

BF1 24-Jul-13 8.0 0.507 0.120 5.2

BF1 24-Jul-13 8.0 0.506 0.120 5.2

BF2 24-Jul-13 8.1 0.455 0.112 5.1

BF2 24-Jul-13 8.2 0.456 0.112 5.2

BF3 24-Jul-13 8.1 0.878 0.112 5.3

BF3 24-Jul-13 8.1 0.877 0.110 5.3

BF4 24-Jul-13 8.1 0.647 0.109 5.3

BF4 24-Jul-13 8.2 0.668 0.110 5.3

BF5 24-Jul-13 8.2 0.681 0.111 5.4

BF5 24-Jul-13 8.2 0.682 0.111 5.3

BF6 24-Jul-13 8.0 1.070 0.105 5.1

BF6 24-Jul-13 8.8 1.040 0.104 5.1

RW 24-Jul-13 8.7 1.530 0.127 5.5

RW 24-Jul-13 8.8 1.520 0.128 5.5

FSP 12-Aug-13 8.1 0.095 0.050 2.9

FSP 12-Aug-13 8.1 0.090 0.049 2.8

BF1 12-Aug-13 7.6 0.222 0.126 4.9

BF1 12-Aug-13 7.7 0.226 0.128 4.9

BF2 12-Aug-13 7.7 0.241 0.126 4.9

BF2 12-Aug-13 7.7 0.245 0.124 5.0

BF3 12-Aug-13 7.8 0.405 0.124 4.9

BF3 12-Aug-13 7.8 0.391 0.126 4.9

BF4 12-Aug-13 7.8 0.385 0.124 4.9

BF4 12-Aug-13 7.8 0.384 0.124 4.9

114

Table 8.1: Water Quality and DOC results (cont.) (- indicate that the data is not available)

Sample Date pH Turbidity (NTU) UV254 (cm-1) DOC (mg/L)

BF5 12-Aug-13 7.8 0.374 0.125 4.9

BF5 12-Aug-13 7.8 0.395 0.124 5.0

BF6 12-Aug-13 7.6 0.367 0.120 4.9

BF6 12-Aug-13 7.6 0.370 0.118 4.0

RW 12-Aug-13 7.0 0.894 0.127 4.9

RW 12-Aug-13 6.9 0.790 0.127 5.0

BF1 24-Sept-13 - - - 4.9

BF1 24-Sept-13 - - - 4.9

BF2 24-Sept-13 - - - 4.9

BF2 24-Sept-13 - - - 4.9

BF3 24-Sept-13 - - - 4.9

BF3 24-Sept-13 - - - 4.8

BF4 24-Sept-13 - - - 4.9

BF4 24-Sept-13 - - - 5.1

BF5 24-Sept-13 - - - 5.0

BF5 24-Sept-13 - - - 4.9

BF6 24-Sept-13 - - - 4.8

BF6 24-Sept-13 - - - 4.8

RW 24-Sept-13 - - - 5.2

RW 24-Sept-13 - - - 5.2

Alum SW 23-Oct-13 6.7 0.139 0.046 3.1

Alum SW 23-Oct-13 6.7 0.125 0.048 3.1

Alum CF 23-Oct-13 6.8 0.063 0.046 3.1

Alum CF 23-Oct-13 6.8 0.061 0.048 3.1

PACl SW 23-Oct-13 7.5 0.188 0.052 3.3

PACl SW 23-Oct-13 7.5 0.193 0.054 3.4

PACl CF 23-Oct-13 7.5 0.058 0.053 3.3

PACl CF 23-Oct-13 7.5 0.06 0.054 3.3

FSP 23-Oct-13 6.8 0.062 0.047 3.1

FSP 23-Oct-13 6.8 0.055 0.051 3.1

BF1 23-Oct-13 7.5 0.178 0.113 5.5

BF1 23-Oct-13 7.6 0.17 0.109 5.5

BF2 23-Oct-13 7.5 0.153 0.110 5.2

BF2 23-Oct-13 7.6 0.158 0.105 5.2

BF3 23-Oct-13 7.5 0.162 0.107 5.2

BF3 23-Oct-13 7.6 0.164 0.109 5.2

BF4 23-Oct-13 7.5 0.17 0.105 5.1

115

Table 8.1: Water Quality and DOC results (cont.) (- indicate that the data is not available)

Sample Date pH Turbidity (NTU) UV254 (cm-1) DOC (mg/L)

BF4 23-Oct-13 7.6 0.167 0.110 5.2

BF5 23-Oct-13 7.5 0.168 0.106 5.1

BF5 23-Oct-13 7.5 0.168 0.110 5.1

BF6 23-Oct-13 7.5 0.166 0.102 5.1

BF6 23-Oct-13 7.5 0.16 0.106 5.1

RW 23-Oct-13 7.7 0.542 0.119 5.4

RW 23-Oct-13 7.8 0.566 0.121 5.3

Alum SW 12-Nov-13 7.1 0.143 0.064 3.0

Alum SW 12-Nov-13 7.1 0.142 0.064 3.1

Alum CF 12-Nov-13 7.2 0.063 0.063 3.1

Alum CF 12-Nov-13 7.2 0.065 0.063 3.1

PACl CF 12-Nov-13 7.9 0.067 0.063 3.1

PACl CF 12-Nov-13 7.8 0.067 0.063 3.1

FSP 12-Nov-13 7.0 0.071 0.058 2.9

FSP 12-Nov-13 7.0 0.07 0.058 2.9

BF1 12-Nov-13 7.6 0.2 0.142 5.3

BF1 12-Nov-13 7.6 0.192 0.145 5.3

BF2 12-Nov-13 7.6 0.219 0.143 5.3

BF2 12-Nov-13 7.6 0.215 0.144 5.2

BF3 12-Nov-13 8.0 0.256 0.144 5.2

BF3 12-Nov-13 8.0 0.26 0.145 5.2

BF4 12-Nov-13 8.0 0.249 0.146 5.2

BF4 12-Nov-13 8.0 0.246 0.147 5.2

BF5 12-Nov-13 7.9 0.269 0.145 5.2

BF5 12-Nov-13 7.9 0.271 0.146 5.3

BF6 12-Nov-13 7.8 0.286 - -

BF6 12-Nov-13 - - - -

RW 12-Nov-13 8.1 0.383 0.148 5.4

RW 12-Nov-13 8.1 0.375 0.148 5.5

More DOC results are shown in Table 8.8

116

Table 8.2: THM and AOX results (Raw) (- indicate that the data is not available)

Sample Date

TCM

(μg/L)

BDCM

(μg/L)

CDBM

(μg/L)

TBM

(μg/L)

THMs

(μg/L)

AOX

(μg/L)

Alum CF 4-Jun-13 54 14 ND ND 68 143

Alum CF 4-Jun-13 54 14 ND ND 68 153

PACl CF 4-Jun-13 73 17 ND ND 90 212

PACl CF 4-Jun-13 70 16 ND ND 86 158

FSP 4-Jun-13 50 13 ND ND 63 172

FSP 4-Jun-13 49 13 ND ND 62 155

Alum SW 4-Jun-13 58 15 ND ND 73 251

Alum SW 4-Jun-13 59 16 ND ND 75 254

PACl SW 4-Jun-13 80 19 ND ND 99 254

PACl SW 4-Jun-13 83 19 ND ND 102 268

BF1 4-Jun-13 153 29 ND ND 182 488

BF1 4-Jun-13 158 29 ND ND 187 508

BF2 4-Jun-13 157 29 ND ND 186 485

BF2 4-Jun-13 158 29 ND ND 187 495

BF5 4-Jun-13 154 29 ND ND 183 518

BF5 4-Jun-13 147 30 ND ND 177 529

RW 4-Jun-13 160 33 ND ND 195 550

RW 4-Jun-13 169 34 ND ND 204 551

Alum CF 24-Jun-13 69 17 ND ND 85 -

Alum CF 24-Jun-13 68 16 ND ND 84 -

PACl CF 24-Jun-13 103 24 ND ND 127 -

PACl CF 24-Jun-13 101 23 ND ND 124 -

Alum SW 24-Jun-13 75 26 ND ND 101 -

Alum SW 24-Jun-13 71 25 ND ND 96 -

PACl SW 24-Jun-13 99 26 ND ND 125 -

PACl SW 24-Jun-13 114 30 ND ND 144 -

BF1 24-Jun-13 180 40 ND ND 220 -

BF1 24-Jun-13 140 31 ND ND 171 -

BF2 24-Jun-13 169 40 ND ND 209 -

BF2 24-Jun-13 155 36 ND ND 191 -

BF3 24-Jun-13 190 42 ND ND 232 -

BF3 24-Jun-13 175 38 ND ND 213 -

BF4 24-Jun-13 183 40 ND ND 224 -

BF4 24-Jun-13 202 43 ND ND 245 -

BF5 24-Jun-13 193 41 ND ND 234 -

BF5 24-Jun-13 207 43 ND ND 250 -

117

Table 8.2: THM and AOX results (Raw) (cont.) (- indicate that the data is not available)

Sample Date

TCM

(μg/L)

BDCM

(μg/L)

CDBM

(μg/L)

TBM

(μg/L)

THMs

(μg/L)

AOX

(μg/L)

BF6 24-Jun-13 163 35 ND ND 199 -

BF6 24-Jun-13 165 35 ND ND 201 -

RW 24-Jun-13 207 45 ND ND 252 -

RW 24-Jun-13 201 43 ND ND 244 -

BF1 24-Jul-13 143 27 ND ND 171 407

BF1 24-Jul-13 135 26 ND ND 160 412

BF2 24-Jul-13 153 29 ND ND 182 400

BF2 24-Jul-13 163 32 ND ND 195 400

BF3 24-Jul-13 166 33 ND ND 199 361

BF3 24-Jul-13 152 29 ND ND 180 411

BF4 24-Jul-13 180 33 ND ND 213 423

BF4 24-Jul-13 170 32 ND ND 202 416

BF5 24-Jul-13 152 32 ND ND 184 375

BF5 24-Jul-13 144 32 ND ND 175 406

BF6 24-Jul-13 140 30 ND ND 171 365

BF6 24-Jul-13 136 30 ND ND 165 365

RW 24-Jul-13 192 42 ND ND 234 500

RW 24-Jul-13 198 44 ND ND 242 538

FSP 24-Sept-13 86 18 ND ND 104 237

FSP 24-Sept-13 80 16 ND ND 96 237

BF1 24-Sept-13 153 28 ND ND 181 398

BF1 24-Sept-13 136 25 ND ND 161 395

BF2 24-Sept-13 150 28 ND ND 178 394

BF2 24-Sept-13 151 26 ND ND 177 388

BF3 24-Sept-13 189 36 ND ND 225 397

BF3 24-Sept-13 165 31 ND ND 196 378

BF4 24-Sept-13 177 33 ND ND 210 374

BF4 24-Sept-13 175 32 ND ND 207 378

BF5 24-Sept-13 173 31 ND ND 203 399

BF5 24-Sept-13 168 30 ND ND 197 375

BF6 24-Sept-13 171 32 ND ND 204 380

BF6 24-Sept-13 161 33 ND ND 195 379

RW 24-Sept-13 193 36 ND ND 230 439

RW 24-Sept-13 203 38 ND ND 240 411

BF1 23-Oct-13 155 35 ND ND 190 484

BF1 23-Oct-13 152 34 ND ND 186 465

118

Table 8.2: THM and AOX results (Raw) (cont.)

Sample Date

TCM

(μg/L)

BDCM

(μg/L)

CDBM

(μg/L)

TBM

(μg/L)

THMs

(μg/L)

AOX

(μg/L)

BF2 23-Oct-13 145 33 ND ND 178 454

BF2 23-Oct-13 136 30 ND ND 167 451

BF3 23-Oct-13 128 29 ND ND 156 477

BF3 23-Oct-13 133 29 ND ND 162 447

BF4 23-Oct-13 133 30 ND ND 163 433

BF4 23-Oct-13 125 28 ND ND 153 441

BF5 23-Oct-13 142 32 ND ND 174 458

BF5 23-Oct-13 142 31 ND ND 173 435

BF6 23-Oct-13 124 28 ND ND 152 415

BF6 23-Oct-13 121 28 ND ND 149 424

RW 23-Oct-13 186 43 ND ND 229 507

RW 23-Oct-13 181 43 ND ND 225 444

FSP 23-Oct-13 55 15 ND ND 70 244

FSP 23-Oct-13 55 15 ND ND 70 241

Alum SW 23-Oct-13 60 17 ND ND 77 246

Alum SW 23-Oct-13 55 15 ND ND 70 253

Alum CF 23-Oct-13 54 16 ND ND 70 232

Alum CF 23-Oct-13 58 17 ND ND 75 233

PACl SW 23-Oct-13 78 20 ND ND 98 296

PACl SW 23-Oct-13 80 21 ND ND 101 287

PACl CF 23-Oct-13 78 20 ND ND 98 247

PACl CF 23-Oct-13 76 20 ND ND 96 248

BF1 12-Nov-13 139 25 ND ND 164 513

BF1 12-Nov-13 129 24 ND ND 152 537

BF2 12-Nov-13 130 21 ND ND 151 489

BF2 12-Nov-13 139 22 ND ND 161 464

BF3 12-Nov-13 131 21 ND ND 152 475

BF3 12-Nov-13 131 21 ND ND 152 475

BF4 12-Nov-13 139 23 ND ND 162 535

BF4 12-Nov-13 134 22 ND ND 156 533

BF5 12-Nov-13 131 22 ND ND 154 533

BF5 12-Nov-13 133 23 ND ND 519 156

BF6 12-Nov-13 127 22 ND ND 404 148

BF6 12-Nov-13 131 22 ND ND 430 152

RW 12-Nov-13 143 25 ND ND 506 168

RW 12-Nov-13 146 24 ND ND 505 171

FSP 12-Nov-13 45 8 ND ND 53 218

119

Table 8.2: THM and AOX results (Raw) (cont.)

Sample Date

TCM

(μg/L)

BDCM

(μg/L)

CDBM

(μg/L)

TBM

(μg/L)

THMs

(μg/L)

AOX

(μg/L)

FSP 12-Nov-13 44 8 ND ND 52 229

Alum SW 12-Nov-13 86 17 ND ND 103 419

Alum SW 12-Nov-13 78 15 ND ND 93 437

Alum CF 12-Nov-13 51 10 ND ND 61 260

Alum CF 12-Nov-13 50 10 ND ND 60 253

PACl CF 12-Nov-13 61 11 ND ND 72 251

PACl CF 12-Nov-13 62 11 ND ND 73 251

211

Table 8.3: HAA results (Raw) (ND=Not Detected)

Sample Date MCAA

(μg/L)

MBAA

(μg/L)

DCAA

(μg/L)

TCAA

(μg/L)

BCAA

(μg/L)

DBAA

(μg/L)

BDCAA

(μg/L)

CDBAA

(μg/L)

TBAA

(μg/L)

HAAs

(μg/L)

Alum CF 4-Jun-13 ND ND 20 23 ND ND ND ND ND 43

Alum CF 4-Jun-13 ND ND 23 25 ND ND ND ND ND 48

PACl CF 4-Jun-13 ND ND 23 27 ND ND ND ND ND 50

PACl CF 4-Jun-13 ND ND 22 27 ND ND ND ND ND 49

FSP 4-Jun-13 ND ND 20 24 ND ND ND ND ND 44

FSP 4-Jun-13 ND ND 19 23 ND ND ND ND ND 42

Alum SW 4-Jun-13 ND ND 25 30 ND ND ND ND ND 55

Alum SW 4-Jun-13 ND ND 22 29 ND ND ND ND ND 51

PACl SW 4-Jun-13 ND ND 29 36 ND ND ND ND ND 65

PACl SW 4-Jun-13 ND ND 26 35 ND ND ND ND ND 61

BF1 4-Jun-13 ND ND 56 66 ND ND ND ND ND 122

BF1 4-Jun-13 ND ND 51 61 ND ND ND ND ND 117

BF2 4-Jun-13 ND ND 45 59 ND ND ND ND ND 104

BF2 4-Jun-13 ND ND 43 58 ND ND ND ND ND 101

BF5 4-Jun-13 ND ND 41 57 ND ND ND ND ND 98

BF5 4-Jun-13 ND ND 44 58 ND ND ND ND ND 102

RW 4-Jun-13 ND ND 52 59 ND ND ND ND ND 110

RW 4-Jun-13 ND ND 51 59 ND ND ND ND ND 110

Alum CF 24-Jun-13 ND ND 20 15 ND ND ND ND ND 35

Alum CF 24-Jun-13 ND ND 21 16 ND ND ND ND ND 37

PACl CF 24-Jun-13 ND ND 22 19 ND ND ND ND ND 41

PACl CF 24-Jun-13 ND ND 23 20 ND ND ND ND ND 43

Alum SW 24-Jun-13 ND ND 28 28 ND ND ND ND ND 56

Alum SW 24-Jun-13 ND ND 28 29 ND ND ND ND ND 57

PACl SW 24-Jun-13 ND ND 28 32 ND ND ND ND ND 60

121

Table 8.3: HAA results (Raw) (ND=Not Detected) (cont.)

Sample Date MCAA

(μg/L)

MBAA

(μg/L)

DCAA

(μg/L)

TCAA

(μg/L)

BCAA

(μg/L)

DBAA

(μg/L)

BDCAA

(μg/L)

CDBAA

(μg/L)

TBAA

(μg/L)

HAAs

(μg/L)

PACl SW 24-Jun-13 ND ND 27 31 ND ND ND ND ND 58

BF1 24-Jun-13 ND ND 46 69 ND ND ND ND ND 115

BF1 24-Jun-13 ND ND 43 63 ND ND ND ND ND 106

BF2 24-Jun-13 ND ND 44 64 ND ND ND ND ND 108

BF2 24-Jun-13 ND ND 43 62 ND ND ND ND ND 105

BF3 24-Jun-13 ND ND 47 66 ND ND ND ND ND 113

BF4 24-Jun-13 ND ND 48 68 ND ND ND ND ND 116

BF4 24-Jun-13 ND ND 49 70 ND ND ND ND ND 119

BF5 24-Jun-13 ND ND 47 66 ND ND ND ND ND 113

BF5 24-Jun-13 ND ND 47 67 ND ND ND ND ND 114

BF6 24-Jun-13 ND ND 46 65 ND ND ND ND ND 111

BF6 24-Jun-13 ND ND 45 66 ND ND ND ND ND 111

RW 24-Jun-13 ND ND 46 64 ND ND ND ND ND 110

RW 24-Jun-13 ND ND 46 68 ND ND ND ND ND 114

BF1 24-Jul-13 ND ND 35 56 ND ND ND ND ND 91

BF1 24-Jul-13 ND ND 35 57 ND ND ND ND ND 92

BF2 24-Jul-13 ND ND 36 58 ND ND ND ND ND 94

BF2 24-Jul-13 ND ND 36 55 ND ND ND ND ND 91

BF3 24-Jul-13 ND ND 37 60 ND ND ND ND ND 97

BF3 24-Jul-13 ND ND 34 54 ND ND ND ND ND 88

BF4 24-Jul-13 ND ND 39 59 ND ND ND ND ND 98

BF4 24-Jul-13 ND ND 35 55 ND ND ND ND ND 90

BF5 24-Jul-13 ND ND 40 60 ND ND ND ND ND 100

BF5 24-Jul-13 ND ND 40 61 ND ND ND ND ND 101

BF6 24-Jul-13 ND ND 38 59 ND ND ND ND ND 97

BF6 24-Jul-13 ND ND 36 56 ND ND ND ND ND 92

122

Table 8.3: HAA results (Raw) (ND=Not Detected) (cont.) (- indicate that the data is not available)

Sample Date MCAA

(μg/L)

MBAA

(μg/L)

DCAA

(μg/L)

TCAA

(μg/L)

BCAA

(μg/L)

DBAA

(μg/L)

BDCAA

(μg/L)

CDBAA

(μg/L)

TBAA

(μg/L)

HAAs

(μg/L)

RW 24-Jul-13 ND ND 45 59 ND ND ND ND ND 104

RW 24-Jul-13 ND ND 43 58 ND ND ND ND ND 101

BF1 24-Sept-13 ND ND 31 85 ND ND ND ND ND 116

BF1 24-Sept-13 ND ND 34 93 ND ND ND ND ND 127

BF2 24-Sept-13 ND ND 30 88 ND ND ND ND ND 118

BF2 24-Sept-13 ND ND 35 95 ND ND ND ND ND 131

BF3 24-Sept-13 ND ND 31 85 ND ND ND ND ND 117

BF3 24-Sept-13 ND ND 33 86 ND ND ND ND ND 119

BF4 24-Sept-13 ND ND 33 80 ND ND ND ND ND 113

BF4 24-Sept-13 ND ND - - ND ND ND ND ND -

BF5 24-Sept-13 ND ND 25 91 ND ND ND ND ND 115

BF5 24-Sept-13 ND ND 23 85 ND ND ND ND ND 108

BF6 24-Sept-13 ND ND 38 101 ND ND ND ND ND 139

BF6 24-Sept-13 ND ND 35 96 ND ND ND ND ND 131

RW 24-Sept-13 ND ND 46 118 ND ND ND ND ND 163

RW 24-Sept-13 ND ND 44 114 ND ND ND ND ND 158

BF1 23-Oct-13 ND ND 56 76 ND ND ND ND ND 132

BF1 23-Oct-13 ND ND 57 78 ND ND ND ND ND 135

BF2 23-Oct-13 ND ND 55 76 ND ND ND ND ND 131

BF2 23-Oct-13 ND ND 54 76 ND ND ND ND ND 130

BF3 23-Oct-13 ND ND 56 77 ND ND ND ND ND 133

BF3 23-Oct-13 ND ND 55 77 ND ND ND ND ND 132

BF4 23-Oct-13 ND ND 57 80 ND ND ND ND ND 137

BF4 23-Oct-13 ND ND 57 81 ND ND ND ND ND 138

123

Table 8.3: HAA results (Raw) (ND=Not Detected) (cont.)

Sample Date MCAA

(μg/L)

MBAA

(μg/L)

DCAA

(μg/L)

TCAA

(μg/L)

BCAA

(μg/L)

DBAA

(μg/L)

BDCAA

(μg/L)

CDBAA

(μg/L)

TBAA

(μg/L)

HAAs

(μg/L)

BF5 23-Oct-13 ND ND 56 79 ND ND ND ND ND 136

BF5 23-Oct-13 ND ND 57 80 ND ND ND ND ND 137

BF6 23-Oct-13 ND ND 51 75 ND ND ND ND ND 127

BF6 23-Oct-13 ND ND 53 78 ND ND ND ND ND 131

RW 23-Oct-13 ND ND 59 78 ND ND ND ND ND 136

RW 23-Oct-13 ND ND 58 78 ND ND ND ND ND 136

FSP 23-Oct-13 ND ND 55 15 ND ND ND ND ND 70

FSP 23-Oct-13 ND ND 55 15 ND ND ND ND ND 70

Alum SW 23-Oct-13 ND ND 60 17 ND ND ND ND ND 77

Alum SW 23-Oct-13 ND ND 55 15 ND ND ND ND ND 70

Alum CF 23-Oct-13 ND ND 54 16 ND ND ND ND ND 70

Alum CF 23-Oct-13 ND ND 58 17 ND ND ND ND ND 75

PACl SW 23-Oct-13 ND ND 78 20 ND ND ND ND ND 98

PACl SW 23-Oct-13 ND ND 80 21 ND ND ND ND ND 101

PACl CF 23-Oct-13 ND ND 78 20 ND ND ND ND ND 98

PACl CF 23-Oct-13 ND ND 76 20 ND ND ND ND ND 96

BF1 12-Nov-13 ND ND 68 98 ND ND ND ND ND 165

BF1 12-Nov-13 ND ND 68 100 ND ND ND ND ND 167

BF2 12-Nov-13 ND ND 73 111 ND ND ND ND ND 184

BF2 12-Nov-13 ND ND 73 109 ND ND ND ND ND 182

BF3 12-Nov-13 ND ND 69 103 ND ND ND ND ND 172

BF3 12-Nov-13 ND ND 78 117 ND ND ND ND ND 195

BF4 12-Nov-13 ND ND 74 111 ND ND ND ND ND 185

BF4 12-Nov-13 ND ND 76 114 ND ND ND ND ND 190

BF5 12-Nov-13 ND ND 78 115 ND ND ND ND ND 192

BF5 12-Nov-13 ND ND 76 115 ND ND ND ND ND 191

124

Table 8.5: HAA results (Raw) (ND=Not Detected) (cont.)

Sample Date MCAA

(μg/L)

MBAA

(μg/L)

DCAA

(μg/L)

TCAA

(μg/L)

BCAA

(μg/L)

DBAA

(μg/L)

BDCAA

(μg/L)

CDBAA

(μg/L)

TBAA

(μg/L)

HAAs

(μg/L)

BF6 12-Nov-13 ND ND 81 119 ND ND ND ND ND 201

BF6 12-Nov-13 ND ND 78 116 ND ND ND ND ND 194

RW 12-Nov-13 ND ND 81 123 ND ND ND ND ND 203

RW 12-Nov-13 ND ND 79 121 ND ND ND ND ND 200

FSP 12-Nov-13 ND ND 20 29 ND ND ND ND ND 49

FSP 12-Nov-13 ND ND 20 29 ND ND ND ND ND 49

Alum SW 12-Nov-13 ND ND 59 67 ND ND ND ND ND 126

Alum SW 12-Nov-13 ND ND 58 65 ND ND ND ND ND 123

Alum CF 12-Nov-13 ND ND 24 36 ND ND ND ND ND 60

Alum CF 12-Nov-13 ND ND 26 39 ND ND ND ND ND 65

PACl CF 12-Nov-13 ND ND 32 37 ND ND ND ND ND 69

PACl CF 12-Nov-13 ND ND 35 40 ND ND ND ND ND 75

219

Table 8.4: Raw ATP and EPS data

Date ATP (ng/g media) Pr (μg/g media) PS (μg/g media)

Alum CF 4-Jun-13 8 19 6

PACl CF 4-Jun-13 7 0 0

FSP 4-Jun-13 28 52 14

BF1 4-Jun-13 708 741 60

BF2 4-Jun-13 757 734 73

BF3 4-Jun-13 103 156 1

BF4 4-Jun-13 156 359 13

BF5 4-Jun-13 124 224 3

FSP 24-Jul-13 76 12 7

BF1 24-Jul-13 814 465 54

BF2 24-Jul-13 693 567 80

BF3 24-Jul-13 1040 497 65

BF4 24-Jul-13 828 547 73

BF5 24-Jul-13 1071 421 45

BF6 24-Jul-13 2087 461 55

FSP 30-Sept-13 67 12 6

BF1 30-Sept-13 493 426 32

BF2 30-Sept-13 446 463 37

BF3 30-Sept-13 567 280 24

BF4 30-Sept-13 867 380 32

BF5 30-Sept-13 480 258 23

BF6 30-Sept-13 423 399 21

FSP 23-Oct-13 52 - -

BF1 23-Oct-13 259 - -

BF2 23-Oct-13 302 - -

BF3 23-Oct-13 510 - -

BF4 23-Oct-13 303 - -

BF5 23-Oct-13 269 - -

BF6 23-Oct-13 153 - -

BF1 18-Nov-13 669 287 32

BF2 18-Nov-13 956 323 34

BF3 18-Nov-13 775 329 27

BF4 18-Nov-13 672 156 25

BF5 18-Nov-13 780 191 22

BF6 18-Nov-13 410 439 36

FSP 18-Nov-13 68 207 71

Alum CF 18-Nov-13 58 105 29

PACl CF 18-Nov-13 8 64 15

126

Table 8.5: LC-OCD results (Raw)

Date BP

(μg/L)

HS

(μg/L)

BB (μg/L) LMWN

(μg/L)

LMWA

(μg/L)

Alum CF 16-Jun-13 207 1294 832 299 137

PACl CF 16-Jun-13 177 1310 711 456 142

Alum SW 16-Jun-13 230 953 281 377 151

PACl SW 16-Jun-13 154 1584 753 372 142

BF1 16-Jun-13 316 3217 863 564 145

BF2 16-Jun-13 460 3373 767 923 125

BF3 16-Jun-13 377 3184 771 490 135

BF5 16-Jun-13 386 3161 731 545 147

BF6 16-Jun-13 337 3180 709 384 136

RW 16-Jun-13 408 3315 777 576 151

FSP 12-Aug-13 121 1654 259 390 259

BF1 12-Aug-13 343 3058 727 493 146

BF2 12-Aug-13 318 2898 892 415 161

BF3 12-Aug-13 396 3135 716 428 159

BF4 12-Aug-13 378 3142 647 462 166

BF5 12-Aug-13 406 3146 679 409 149

BF6 12-Aug-13 434 3014 738 413 143

RW 12-Aug-13 436 3013 781 452 137

FSP 24-Sept-13 219 1310 742 433 174

BF1 24-Sept-13 420 3190 833 400 157

BF2 24-Sept-13 415 3094 855 404 155

BF3 24-Sept-13 409 3150 826 416 150

BF4 24-Sept-13 404 2968 787 414 151

BF5 24-Sept-13 397 3046 909 423 154

BF6 24-Sept-13 403 3118 803 420 149

RW 24-Sept-13 534 3191 861 382 144

BF1 18-Nov-13 418 3263 1112 381 127

BF2 18-Nov-13 435 3448 889 336 137

BF3 18-Nov-13 416 3241 1093 337 128

BF4 18-Nov-13 444 3287 1056 343 130

BF5 18-Nov-13 398 3247 1092 357 130

BF6 18-Nov-13 408 3338 1031 329 133

RW 18-Nov-13 489 3277 1122 310 140

FSP 18-Nov-13 192 1051 902 306 125

Alum SW 18-Nov-13 283 1228 908 295 125

Alum CF 18-Nov-13 251 1256 887 328 140

PACl CF 18-Nov-13 206 1366 695 393 184

127

8.3 Appendix C (QA/QC)

8.3.1 THMs

Table 8.6: Sample THMs control chart warning and control limits

TCM (μg/L) BDCM (μg/L) CDBM (μg/L) TBM (μg/L)

Mean (n = 7) 20.9 21.1 25.1 25.3

SD (n = 7) 1.6 1.5 19.4 1.4

UWL 24 24 25.1 28.1

LWL 17.7 18.2 19.4 22.5

UCL 25.6 25.5 26.5 29.5

LCL 16.1 16.7 17.9 21.0

Figure 8.6: TCM control chart (June 2013)

Figure 8.7: TCM control chart (July-Nov 2013)

5

10

15

20

25

30

0 2 4 6 8 10 12 14

TC

M C

once

ntr

atio

n (

ug/L

)

5

10

15

20

25

30

35

0 2 4 6 8 10 12 14

TC

M C

once

ntr

atio

n

(μg/L

)

June 2013

Mean

UWL

LWL LCL

UCL

LWL

LCL

UCL

Mean

UWL

July 2013 Nov 2013 Sept 2013

128

Figure 8.8: BDCM control chart (June 2013)

Figure 8.9: BDCM control chart (July-Nov 2013)

10

12

14

16

18

20

22

24

26

28

0 2 4 6 8 10 12 14

BD

CM

Conce

ntr

atio

n (

μg/L

)

10

12

14

16

18

20

22

24

26

28

0 2 4 6 8 10 12 14

BD

CM

Conce

ntr

atio

n (

μg/L

)

LWL

LCL

UCL

Mean

UWL

June 2013

LWL

LCL

UCL

Mean

UWL

July 2013 Nov 2013 Sept 2013

129

8.3.2 HAAs

Table 8.7: HAA control chart warning and control limits

MCAA MBAA DCAA TCAA BCAA DBAA BDCAA CDBAA TBAA

Mean

(n=8) 38.51 39.54 40.23 44.71 42.12 46.69 52.47 61.97 73.10

Standard

deviation

(n=8) 1.74 1.89 1.86 1.72 1.85 2.05 3.77 6.47 8.36

UWL 41.99 43.33 43.95 48.15 45.82 50.78 60.01 74.92 89.81

LWL 35.03 35.75 36.51 41.26 38.41 42.60 44.93 49.02 56.39

UCL 43.73 45.22 45.81 49.87 47.68 52.82 63.78 81.39 98.17

LCL 33.28 33.85 34.65 39.54 36.55 40.55 41.16 42.55 48.03

Figure 8.10: DCAA control chart (June-Nov 2013)

Figure 8.11: TCAA control chart (June-Nov 2013)

30

32

34

36

38

40

42

44

46

48

0 2 4 6 8 10 12 14 16

DC

AA

g/L

)

35

37

39

41

43

45

47

49

51

0 2 4 6 8 10 12 14 16

TC

AA

g/L

)

LWL

LCL

UCL

Mean

UWL

LWL

LCL

UCL

Mean

UWL

July 2013 Nov 2013 Sept 2013 Oct 2013 June 2013

July 2013 Nov 2013 Sept 2013 Oct 2013 June 2013

130

Figure 8.12: AOX control chart (Sept-Nov 2013)

8.4 Appendix D (Additional Information)

Figure 8.13: Daily coagulant dose during the study

90

95

100

105

110

115

0 2 4 6 8 10 12 14

AO

X C

once

ntr

atio

n (

ug/L

as

Cl-

)

20

25

30

35

40

45

50

55

01-Jun-13 21-Jul-13 09-Sep-13 29-Oct-13

Coag

ula

nt

Dose

(m

g/L

)

Full-scale Plant Pilot-Plant Alum Pilot-Plant PACl

Transition

study

Sept 2013 Oct 2013 Nov 2013

LWL

LCL

UCL

Mean

UWL

131

Figure 8.14: Avarage daily raw water tempreture at the peterborough water treatment

plant

Figure 8.15: Average daily DOC concentration in raw water, conventional, biofilter

effluent

0

5

10

15

20

25

30

35

5/6/2013 6/25/2013 8/14/2013 10/3/2013 11/22/2013 1/11/2014

Avar

age

Dai

ly R

aw W

ater

Tem

pre

ture

(˚C

)

0

1

2

3

4

5

6

7

6-May-13 25-Jun-13 14-Aug-13 3-Oct-13 22-Nov-13 11-Jan-14

DO

C (

mg/L

)

Raw Water Full-scale plant filter 11

Post-filter Al Post-filter PACl

Biofilter (large control) BF2 (Nutrient Enhanced)

132

Figure 8.16: Average daily UV254 in raw water, conventional, biofilter effluent

Figure 8.17: Normalized UF resistance profile comparing small vs large control biofilter

effluent (BF1: September 16th,2013, T = 18 ºC. BF3: September 30th, 2013, T = 18 ºC)

0

0.05

0.1

0.15

0.2

0.25

6-May-13 25-Jun-13 14-Aug-13 3-Oct-13 22-Nov-13 11-Jan-14

UV

25

4(c

m-1

)Raw Water Full-scale plant filter 11

Post-filter Al Post-filter PACl

Biofilter (large control) BF2 (Nutrient Enhanced)

0

2

4

6

8

0 5 10 15 20 25 30

Norm

aliz

ed m

embra

ne

resi

stan

ce @

20˚C

10

12

m-1

)

Permeation Time (hr)

BF1 (Large Control)

BF3 (Small Control)

133

Table 8.8: DOC concentrations in raw water and biofilter effluent (- indicate that the data is not available)

BF1 (Control

Large)

BF2 (Nutrient

Enhancement)

BF3 (Control

Small)

BF4 (Peroxide

Addition)

BF5 (Alum

Addition)

BF6

(GAC)

Raw

Water

4-Jun-13 6.2 6.2 6.5 6.5 6.5 - 7.0

24-Jun-13 5.2 5.3 5.4 5.4 5.4 5.4 5.5

24-Jul-13 5.2 5.2 5.3 5.3 5.3 5.1 5.5

12-Aug-13 4.9 4.9 4.9 4.9 5.0 4.9 5.0

27-Aug-13 5.0 5.0 5.1 5.1 5.1 5.1 5.4

24-Sep-13 4.9 4.9 4.9 5.0 4.9 4.8 5.2

26-Sep-13 5.6 5.6 5.5 5.4 5.5 5.5 5.9

2-Oct-13 5.4 5.4 - 5.4 5.3 5.2 5.7

4-Oct-13 5.5 5.3 5.3 5.4 5.5 5.3 5.8

8-Oct-13 5.2 5.2 5.3 5.5 5.4 5.3 5.7

10-Oct-13 5.3 5.2 5.1 5.2 5.2 5.0 5.6

16-Oct-13 5.2 5.3 5.2 5.3 5.3 5.1 5.6

23-Oct-13 5.4 5.2 5.2 5.1 5.1 5.1 5.5

24-Oct-13 5.3 5.4 5.3 5.3 5.3 5.3 5.6

29-Oct-13 5.3 5.4 5.3 5.5 5.2 5.2 5.5

5-Nov-13 5.2 5.3 5.2 5.3 5.2 5.2 5.3

7-Nov-13 5.2 5.5 5.3 5.4 5.3 5.3 5.4

12-Nov-13 5.3 5.2 5.2 5.2 5.3 - 5.5