adsorption and biological filtration of microcystins

125
ADSORPTION AND BIOLOGICAL FILTRATION OF MICROCYSTINS By Haixiang Wang School of Chemical Engineering The University of Adelaide Australia A thesis submitted to the University of Adelaide for the degree of Masters in Engineering Science - May 2006 –

Upload: others

Post on 03-Oct-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Adsorption and biological filtration of microcystins

ADSORPTION AND BIOLOGICAL FILTRATION

OF

MICROCYSTINS

By

Haixiang Wang

School of Chemical Engineering

The University of Adelaide

Australia

A thesis submitted to the University of Adelaide for the degree of Masters in Engineering Science

- May 2006 –

Page 2: Adsorption and biological filtration of microcystins

ii

Abstract For removal of microcystins (liver tumor promoting cyanotoxins) from drinking water

granular activated carbon (GAC) filtration has been shown to be a very promising

treatment option. It is not only an efficient adsorbent for microcystins but also can be

operated as a biological reactor in which bacterial degradation of the toxins can occur

extending the life-time of this application.

However, the competitive adsorption of coexistent natural organic matter (NOM) in all

natural water sources would cause early breakthrough of microcystins and a lag-phase

with uncertain length occurs prior to initiation of biodegradation. This project aimed to

investigate the individual microcystin removal abilities by adsorption and

biodegradation during GAC filtration in order to better understand the overall efficiency

of this application. The simultaneous elimination of NOM in GAC filters was also

investigated. In addition, to facilitate the biological removal of microcystins, the

research aimed to identify the potential effect of key operational conditions on the

degradation efficiency.

In this study microcystin removal in GAC filtration was divided into the adsorption and

biodegradation phases. Effective adsorption of the toxins lasted only a short term in

virgin GAC filters and the breakthrough behaviour was able to be modelled by the

homogenous surface diffusion model (HSDM). The presence of biofilm on the surfaces

of GAC resulted in a lower mass transfer coefficient (Kf) and lower adsorption kinetics

of microcystins. In the biodegradation phase enhanced or complete removal was evident

that was mainly due to biological metabolism. However, the highly efficient

biodegradation of the toxins was difficult to predict in GAC filtration. Biodegradation

was found to be easily affected by many operational factors.

NOM removal in GAC filters demonstrated specific features in this study due to the

relatively high organic content in the Australian water. The adsorption efficiency

decreased rapidly as the adsorption capacity became saturated.

Page 3: Adsorption and biological filtration of microcystins

iii

Statement of Originality This work contains no material which has been accepted for the award of any other

degree or diploma in any university or other tertiary institution and, to the best of my

knowledge and belief, contains no material previously published or written by another

person, except where due reference has been made in the text.

I give consent to this copy of my thesis being made available in the University Library.

The author acknowledges that copyright of published works contained within this thesis

(as listed below) resides with the copyright holder/s of those works.

Signed: Date:

Page 4: Adsorption and biological filtration of microcystins

iv

Acknowledgments Throughout this two-year project, many people have helped me. Firstly, I would like to

offer my heartfelt appreciation to all my supervisors, David Lewis, Keith King, Justin

Brookes and Gayle Newcombe, for their encouragement, enthusiasm, instrumental

suggestions and invaluable support. My principal supervisor, David Lewis, has provided

many precious hours and physical effort to transport weekly 150 litres of water samples

from Myponga water treatment plant which is more than 100 km away to the laboratory

for the duration of the experiment. Gayle Newcombe and Justin Brookes, my very

helpful co-supervisors at Australian Water Quality Centre (AWQC), have provided the

most vital instrumental contribution in development of this project and offered many

hours of effort in reviewing all my paper work. My sincerest appreciation also gives to

Lionel Ho at AWQC for all his detailed and patient supervision in the laboratory and the

very helpful discussions. There were also many other people who have facilitated completion of this project and

the preparation of my thesis. I will forever grateful to Najwa Slyman, Edith Kozlik,

Miriam Nedic, Rolando Fabris and Mary Drikas for their assistance and care at the

Water Treatment Unit, AWQC. The School of Chemical Engineering at The University

of Adelaide has also been a great place for me to do research, and the very helpful and

kind staff there made my time during the two years of study stimulating and enjoyable.

Last but not least I would like to give my deepest thank to my parents who financially

supported and encouraged my study overseas and made all of this possible.

Page 5: Adsorption and biological filtration of microcystins

v

List of Publications Haixiang W., Lewis D. M., Brookes J. B., Newcombe G. and Lionel H. (2005)

Combined adsorption and biodegradation of microcystins (Poster Presentation). The

10th International Conference in Applied Phycology. July 2005.

Haixiang W., Brookes J. B., Lewis D. M., Newcombe G. and Lionel H. (2006)

Adsorption and biodegradation of microcystins in GAC filtration. In Proceedings of the

1st Australian Young Water Professionals Conference, Sydney, Australia. February 2006.

Haixiang W., Lewis D. M., Brookes J. B. and Newcombe G. (2006) Removal of

microcystins in GAC filtration. In Proceedings of the Fifth Biennial Postgraduate

Student Conference of the CRC for Water Quality and Treatment, Melbourne, Australia.

July 2006.

Haixiang W., Lewis D. M., Brookes J. B., Newcombe G. and Lionel H. (2006)

Separated adsorption and bacterial degradation of microcystins in GAC filtration. In

Proceedings of the CHEMECA Conference, Auckland, New Zealand. September 2006.

In press.

Page 6: Adsorption and biological filtration of microcystins

vi

Table of Contents Abstract................................................................................................................................. ii

Statement of Originality ..................................................................................................... iii

Acknowledgments ............................................................................................................... iv

List of Publications ...............................................................................................................v

Table of Contents ................................................................................................................ vi

List of Figures........................................................................................................................x

List of Tables...................................................................................................................... xiii

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

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

1.1.1 Cyanobacteria and cyanotoxins ...................................................................................................1

1.1.2 What are microcystins?.................................................................................................................4

1.1.3 Toxicity of microcystins ...............................................................................................................5

1.1.4 Microcystin persistence in drinking water treatment plants ....................................................7

1.2 Available treatment technologies in removing or reducing microcystins in drinking water .................. 8

1.2.1 Oxidation.........................................................................................................................................9

Chlorination ............................................................................................................................. 9

Ozonation............................................................................................................................... 10

Potassium permanganate ....................................................................................................... 10

Ultraviolet (UV) light ............................................................................................................ 11

Hydrogen peroxide ................................................................................................................ 11

Titanium dioxide.................................................................................................................... 12

1.2.2 Membrane filtration.....................................................................................................................13

1.2.3 Biodegradation .............................................................................................................................14

1.2.4 Disadvantages from these treatment solutions ........................................................................15

1.3 Activated carbon........................................................................................................................................... 16

1.3.1 Powdered activated carbon.........................................................................................................17

1.3.2 Granular activated carbon filtration ..........................................................................................18

1.4 Objectives of this research........................................................................................................................... 20

Page 7: Adsorption and biological filtration of microcystins

vii

2. EXPERIMENTAL METHODS AND MATERIALS ................................................................................... 22

2.1 Experimental methods ................................................................................................................................. 22

2.1.1 Experiment setup .........................................................................................................................22

2.1.2 Selection of the experimental design ........................................................................................24

2.1.3 Operation of the columns ...........................................................................................................25

2.1.4 Backwashing methods.................................................................................................................26

2.1.5 Scanning electron microscopy (SEM) study............................................................................26

2.2 Reagents and materials ................................................................................................................................ 27

2.2.1 Reagents ........................................................................................................................................27

2.2.2 Microcystin...................................................................................................................................27

2.2.3 Granular activated carbon (GAC) .............................................................................................28

2.2.4 Sand ...............................................................................................................................................29

2.2.5 Water sample ................................................................................................................................29

2.3 Analytical methods ....................................................................................................................................... 30

2.3.1 Microcystin analysis....................................................................................................................30

2.3.2 Dissolved organic carbon and UV254 absorbance....................................................................30

2.3.3 High performance size exclusion chromatography.................................................................31

2.3.4 Scanning electron microscopy ...................................................................................................31

2.4 Validation of the experimental design - side effects from the sterilisation method of autoclaving ... 32

2.4.1 Effects on the water quality........................................................................................................32

2.4.2 Loss of organic carbon from GAC structure and apparatus ..................................................33

2.4.3 Effects on adsorption capacity of the GAC..............................................................................34

2.4.4 Effects on desorption...................................................................................................................34

3. LABORATORY-SCALE COLUMN EXPERIMENT: MICROCYSTIN REMOVAL.............. 35

3.1 Introduction .....................................................................................................................35

3.2 Results from the first set of experiments.................................................................................................... 36

3.2.1 The non-sterile GAC column .....................................................................................................37

3.2.2 The sterile GAC column .............................................................................................................38

3.2.3 The sand column ..........................................................................................................................40

3.3 Results from the repeat experiment............................................................................................................ 44

Page 8: Adsorption and biological filtration of microcystins

viii

3.4 Combined results .......................................................................................................................................... 47

3.5 Phases of microcystin removal in GAC filters........................................................................................... 48

3.5.1 Phase I - Adsorption removal and biofilm effect.....................................................................50

3.5.2 Phase II - Biodegradation removal and lag-phase...................................................................52

3.6 Conclusions ................................................................................................................................................... 54

4. REMOVAL OF NATURAL ORGANIC MATTER IN GAC AND SAND COLUMNS .........56

4.1 Introduction .................................................................................................................................................. 56

4.2 Methods ......................................................................................................................................................... 58

4.3 Results and Discussion ................................................................................................................................. 59

4.3.1 Non-sterile GAC column............................................................................................................60

4.3.2 Sterile GAC column ....................................................................................................................62

4.3.3 Sand column .................................................................................................................................64

4.4 Conclusions ................................................................................................................................................... 66

5. EVALUATING THE VALIDITY OF THE EXPERIMENTAL DESIGN..............................67

5.1 Introduction .................................................................................................................................................. 67

5.2 Experimental methods ................................................................................................................................. 68

5.3 Results and Discussion ................................................................................................................................. 68

5.3.1 Effects on water quality ..............................................................................................................68

5.3.2 Loss of organic carbon from GAC structure and apparatus ..................................................69

5.3.3 Effects on the adsorption efficiency of GAC...........................................................................72

5.3.4 Effects on desorption during autoclaving.................................................................................73

5.4 Conclusion..................................................................................................................................................... 76

6. BIODEGRADATION OF MICROCYSTINS AND THE OPERATIONAL CONDITIONS…78

6.1 Introduction .................................................................................................................................................. 78

6.2 Methods ......................................................................................................................................................... 79

6.2.1 Microcystin biodegradation under varied temperature, light irradiation and water

quality conditions ..................................................................................................................................79

Page 9: Adsorption and biological filtration of microcystins

ix

6.2.2 Microcystin biodegradation in cultures of different bacterial concentration ......................81

6.3 Results and Discussion ................................................................................................................................. 81

6.3.1 Effects from temperature ............................................................................................................81

6.3.2 Effects from light condition .......................................................................................................82

6.3.3 Effects from water quality ..........................................................................................................83

6.3.4 Effects from bacterial cell concentration..................................................................................85

6.4 Conclusions ................................................................................................................................................... 86

7. CONCLUSIONS ................................................................................................................87

7.1 Conclusions ................................................................................................................................................... 87

7.2 Future work .................................................................................................................................................. 89

8. REFERENCES ..................................................................................................................90

APPENDIX A BACTERIAL COLONY COUNT..........................................................105

APPENDIX B CALCULATION OF MICROCYSTIN ADSORPTION RATE IN

THE NON-STERILE GAC COLUMN ..........................................................................110

Page 10: Adsorption and biological filtration of microcystins

x

List of Figures

Figure 1.1 Microscopic photo of the toxin producing Microcystis aeruginosa (cell coenobium)

(Murphy et al., 2003) ......................................................................................................................... 4

Figure 1.2 Representation of the molecular structure of microcystins (taken from “The Toxic

Cyanobacteria” website http://lurac.latrobe.edu.au/~botbml/mictox.html) ....................................... 5

Figure 2.1 Schematic diagram of the experiment setup ................................................................... 23

Figure 2.2 Demonstration of the backwashing technique ................................................................ 26

Figure 2.3 Testing for carbon leaching from tubing and connectors belonging to the sterile

GAC system during autoclaving ....................................................................................................... 34

Figure 3.1 Influent and effluent concentrations of m-LR (A) and m–LA (B) of the columns

during the first set of experiment ...................................................................................................... 37

Figure 3.2 Percent removal of m-LR and m–LA in the non-sterile GAC column (the dotted

circle indicates the reoccurred toxin breakthrough) .......................................................................... 38

Figure 3.3 m-LR and m–LA concentration in the effluent of the sterile and the non-sterile GAC

column from days 66 to100 (the dotted circle indicates the lower toxin breakthrough in the

sterile GAC column) ......................................................................................................................... 40

Figure 3.4 Impact of autoclaving on the effluent concentration of both m-LR (upper) and m-LA

(lower) from the sterile GAC column – enhanced toxin adsorption after autoclaving ..................... 41

Figure 3.5 Removal of m-LR in the sand column (the dotted rectangular indicates the sharp

decrease in the toxin outlet concentration when biodegradation commenced in the sand column).. 42

Figure 3.6 Scanning electron microscopic images showing the biofilm that grew on the surface

of sand (left) and the GAC (right) after approximately 3 months (top) and 8 months (bottom)....... 43

Figure 3.7 Scanning electron microscopic images of the different surface characteristics of the

sand (left) and the GAC A6 (right) particles ..................................................................................... 44

Figure 3.8 Scanning electron microscopic images showing the bacterial growth in the internal

pores of the GAC .............................................................................................................................. 44

Figure 3.9 m-LR and m–LA concentration for the inflow and outflow from the columns in the

repeated set of experiment ............................................................................................................... 46

Figure 3.10 Scanning electron microscopic images showing the size of main bacteria species

that were isolated from the non-sterile GAC column ....................................................................... 47

Figure 3.11 Final results of m-LR (top) and m-LA (bottom) breakthrough profiles in the three

columns ............................................................................................................................................. 49

Figure 3.12 A) Normalised percent breakthrough curves of microcystins in the sterile GAC, the

non-sterile GAC and the sand column in the column experiment. Microcystin removal in GAC

Page 11: Adsorption and biological filtration of microcystins

xi

filtration is divided into Phase I (adsorption) and Phase II (biodegradation)

B) Conceptual model describing microcystin removal mechanisms by GAC before and after

biofilm development. The corresponding removal rate by adsorption and biodegradation

calculated from this study is also given ............................................................................................ 50

Figure 3.13 HSDM fit of the sterile and non-sterile GAC experimental data at different mass

transfer coefficient Kf values ............................................................................................................. 52

Figure 4.1 DOC concentration of the influent and effluent of the non-sterile GAC column (the

dotted circle indicates the initial NOM breakthrough)...................................................................... 61

Figure 4.2 UV254 absorbance of the influent and effluent of the three columns containing the

sterile and non-sterile GAC and the sand column (the dotted circle indicates the initial NOM

breakthrough) .................................................................................................................................... 61

Figure 4.3 HPSEC chromatograph showing NOM adsorption in new GAC (day 5)....................... 62

Figure 4.4 DOC concentration of the influent and effluent of the sterile GAC column .................. 63

Figure 4.5 HPSE chromatograph showing NOM adsorption in the preloaded sterile GAC (5.5

months) ............................................................................................................................................. 64

Figure 4.6 DOC characterised NOM measured on the influent and effluent of the sand column.... 66

Figure 4.7 HPSEC chromatograph showing NOM removal in the sand column (~ 300 days)........ 66

Figure 5.1 Molecular weight (MW) distributions from Myponga treated water before and after

autoclaving determined by high performance size exclusion chromatography ................................ 70

Figure 5.2 Carbon leached from the original GAC structure during six sets of autoclaving as

determined by DOC and UV abs @ 254 nm..................................................................................... 71

Figure 5.3 Comparison of the carbon released from the GAC structure during autoclaving and

the carbon content of the source water. Determined by high performance size exclusion

chromatography ................................................................................................................................ 72

Figure 5.4 Concentration of carbon leached from the autoclaved tubing compared with the

usual DOC concentration in the source water ................................................................................... 73

Figure 5.5 Comparison of the adsorption rates of PAC A6 before and after autoclaving as

determined from microcystins (m-LR and m-LA) and NOM (UVA254 and DOC) in the Myponga

source water ...................................................................................................................................... 74

Figure 5.6 Variation of DOC and UVA254 in the effluent of the autoclaved GAC column as a

function of the volume (mL) of treated water ................................................................................... 75

Figure 5.7 MW distribution of the UV sensitive organic carbons that were desorbed from GAC

during autoclaving and that exist in the source water (determined by HPSEC) ............................... 76

Figure 5.8 Specific UV absorbance (SUVA) of the released organic carbons from GAC during

autoclaving, and the SUVA of the source water ................................................................................ 77

Figure 6.1 Biodegradation of m-LR in bioreactors operated at ambient temperature (22 °C), 25

Page 12: Adsorption and biological filtration of microcystins

xii

°C, 33 °C and 40 °C, respectively ..................................................................................................... 83

Figure 6.2 Biodegradation of m-LR in bioreactors with and without irradiation of natural light .... 84

Figure 6.3 Biodegradation of m-LR in bioreactors of undiluted (100% NOM), 50% (NOM)

diluted and 10% (NOM) diluted Myponga treated water ................................................................. 85

Figure 6.4 Biodegradation of m-LR in bioreactors of different bacterial cell concentration ........... 86

Page 13: Adsorption and biological filtration of microcystins

xiii

List of Tables

Table 1.1 Reported intoxication cases by microcystins...................................................................... 7

Table 2.1 Operating parameters for the three columns (Figure 2.1)................................................. 24

Table 2.2 Physico-chemical properties of GAC A6.......................................................................... 28

Table 3.1 HSDM input parameters (Kf, K & 1/n) describing m-LR adsorption onto the sterile

and non-sterile GAC in the present study, and MIB and m-LA adsorption parameters cited in

Ho’s (2004) study.............................................................................................................................. 52

Page 14: Adsorption and biological filtration of microcystins

xiv

Glossary of Abbreviations and Acronyms

AC

Adda

BGAC

DOC

EBCT

GAC

HPLC

HPSEC

HSDM Kf

m-LA

m-LR

MW

NOM

PAC

SEM

UV

UVA254

WTP

TCE

THMs

Activated carbon

3-amino-9-methoxy-2,6,8-trimethyl-10

phenyldeca -4,6-dienoic acid

Biological granular activated carbon

Dissolved organic carbon

Empty bed contact time

Granular activated carbon

High performance liquid chromatography

High performance size exclusion

chromatography

Homogenous surface diffusion model

Film mass transfer coefficient

Microcystin-LA

Microcystin-LR

Molecular weight

Natural organic material

Powdered activated carbon

Scanning electron microscopy

Ultraviolet

Ultraviolet absorbance at 254nm wavelength

Water treatment plant

Trichloroethylene

Trihalomethanes

Page 15: Adsorption and biological filtration of microcystins

1

CHAPTER 1

INTRODUCTION

The aim of this research was to separate the mechanism of adsorption from

biodegradation of microcystins in granular activated carbon (GAC) filtration. This

enabled a comprehensive evaluation of the relative contributions of the physical and

biological mechanisms for microcystin removal in GAC filters.

1.1 Background

1.1.1 Cyanobacteria and cyanotoxins

“Cyanobacteria” or “cyanotoxins” may not be as publicly well known as “algal blooms”,

“green tides”, “red tides”, “inedible shellfish” or “fish kills” etc, but they have a

common link–eutrophication.

Eutrophication is the enrichment of lakes, rivers and reservoirs with chemical nutrients

(typically phosphorus, nitrogen or both) (APIS, 2005). This changing status of the

aquatic environment, resulting in excessive growth of algal population, high levels of

organic matter, oxygen depletion and specific unpleasant smells etc, causes specific

health risks for communities who must rely on eutrophic waters as drinking water

sources. One of the most severe health threads is derived from cyanobacteria, because

approximately 75% of cyanobacterial blooms are accompanied by toxic metabolites,

cyanotoxins, which have been implicated in both animal and human poisoning [Fott,

1971; World Health Organisation (WHO), 1998].

Page 16: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

2

Cyanobacteria, also called cyanophyte or blue/green algae, are organisms capable of

plant-type photosynthesis and have close relationship with both bacteria and algae

(Eutrophication and Health, 2002). They belong to an ancient group of organisms that

are approximately three billion years old (Yoo et al., 1995), and are perhaps the most

primitive group of organisms living on Earth for they posses significant ecological

advantages over other planktonic species (Haider et al., 2003). For example, many

species of cyanobacteria possess gas vacuoles that allow them to adjust their position

(depth) in the water column under varied nutrient conditions and they can also benefit

from the nitrogen fixing capability in low nitrogen waters that increases their natural

productivity.

Cyanobacteria are able to grow under a wide range of environmental conditions

throughout the world. Recent studies showed that their habitats range from thermal

springs to the cold oceans of Antarctica, and some species subsist in terrestrial

environments, growing well in soil and on rocks (Yoo et al., 1995). Cyanobacteria can

grow prolifically under favorable conditions that include ample sunlight; moderate to

high concentrations of essential nutrients (phosphorus, ammonia, nitrate), which can be

received from household, industrial and agricultural wastes; suitable water temperature

(generally between 15 °C to 30 °C); and pH >6 (Skulberg et al., 1984). Therefore,

cyanobacterial blooms typically occur in late summer or early autumn in eutrophic or

hypereutrophic water bodies.

Since the first report of toxic cyanobacteria resulting in animal stock death was made in

Australia in 1878 (Francis, 1878), these potentially hazardous compounds have been

detected in fresh water in 27 countries and on all continents (Newcombe, 2002). With

the development of more sensitive analytical techniques for the analysis of algal toxins

over the past decade, the detection of cyanotoxins in drinking water has increased

internationally.

Page 17: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

3

During normal growth conditions, a number of species of cyanobacteria can produce

one or more types of toxin, which are called cyanotoxins. Cyanotoxins are characterized

by their existence state, either intracellular (cell bound) or extracellular (dissolved). In

healthy cyanobacterial blooms, more than 90% of the total amount is bounded within

the cells (Sivonen and Jones, 1999). However, when the population ages, the bound

compounds will be released into the surrounding water and high levels of extracellular

(dissolved) toxins will be generated (Kaebernick et al., 2000).

According to the current taxonomy, 150 genera with about 2000 species of

cyanobacterium have been recognized, and at least 40 of them are known to be

toxicogenic (Skulberg et al., 1993). Although the toxins produced are diverse, both in

structure and function, they are generally separated into four distinct categories:

1. Neurotoxins (such as: saxitoxins, anatoxins)

Characteristics: less common, potent nerve poisons; can cause muscle spasms, affect

respiration, result in death, but no apparent chronic effects.

2. Hepatotoxins (such as: microcystins, nodularins)

Characteristics: the most frequently observed cyanotoxins, inhibit specific enzyme

systems and cause severe damage to the gut and liver, can lead to death within a few

hours at acute dose exposure.

3. General Cytotoxins (such as: cylindrospermopsin)

Characteristics: implicated in a number of diseases such as gastroenteritis, kidney

damage, liver malfunction, and in particular, are carcinogenic.

4. Endotoxins (such as: lipopolysaccharides)

Characteristics: may result in gastrointestinal upsets and skin irritations.

Among these known cyanotoxins, the hepatotoxic cyclic peptides (microcystins) are

considered the most common (Sivonen and Jones, 1999), and investigations conducted

on water treatment for the removal of cyanotoxins has been primarily focused on

Page 18: Adsorption and biological filtration of microcystins

Chenfer I Tnfrorfircfion

microcystins (Donati et al., 1993;Yoo et al.,1995; Newcombe et aL.2002)

1.1.2 What are microcystins?

Microcystins are the largest family in the group of hepatotoxins (liver toxins) (WHO,

1998). They are monocyclic heptapeptide toxins commonly produced by the freshwater

cyanobacterium, Microcystis aerugínosa (cell coenobium, Figure 1.1), from which the

toxins derived the name. They can also be produced from other geÍLera, namely,

Anabaena, Nostoc and PlanWothrix (Carmichael, 1992; Codd, 1995), and also a

terrestrial (soil) species Haphalosiphon hibernicu, implying its widespread potential

existence in the environment.

tr'igure 1.1 Microscopic photo of the toxin prodtcingMicrocystis aeruginosa (cell coenobium)

(Murphy et a1.,2003)

Microcystins have more than 60 structural variants, each with different toxicity

(Hyenstrand et a1.,2001). They all possess seven amino acids, with five non-protein

amino acids (D-alanine, D-erythro-ß-methyl aspartic acid, D-glutamic acid,

N-methyldehydroalanine and Adda (3-amino-9-methoxy-2,6,8-trimethyl-10-phenyldeca

-4,6-dienoic acid)) and two variable L-amino acids (protein amino acids) at position X

andZ respectively (Figure 1.2). Different structures in microcystins result in different

lipophilicites and polarities, which could in turn affect toxicity (WHO, 1998).

Microcystin-LR (m-LR) and -LA (m-LA), which were investigated in this study, have

leucine (L), arginine (R) for m-LR and alanine (A) for m-LA in variable positions. They

4

Page 19: Adsorption and biological filtration of microcystins

Chanter 1 Introduction

represent two of the best characterised and the most toxic analogues of microcystin

(Oudra et a|.,2002; Oberholster et aL.,2003).

1

H

z 2

1 - D-Alanine2 - Yariable L-amino acid3 - D-Methylaspartic acid4 - Variable L-amino acid5 - 3-amino-g-methoxy-2,6,8-trimethyl

-1 O-phenyldeca-4,6-dienoic acid (Adda)6 - D-Glutamic acid7 - N-Methyldehydroalanine

tr'igure 1.2 Representation of the molecular structure of microcystins ("The Toxic

Cyanobacteria" website )

The unusual Adda amino acid is an essential side chain used to study properties of

microcystins. Variation of the Adda structure could potentially influence microcystin

toxicity, as it is important in the binding to protein phosphatases where microcystins

exhibit toxic activities (John and Charles, 1999; Mankiewicz et al., 2002). The

stereochemistry and extent of methylation of Adda also greatly influences toxicity.

Furthermore, the absorbance characteristics of Adda at wavelength 238nm are also

utilized as a method for microcystin analysis.

1.1.3 Toxicity of microcystins

Microcystins can severely compromise human health. Although the mechanisms that

mediate toxicity have not been fully understood, it is believed that uptake into

76

5

X

4 3

Page 20: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

6

hepatocytes followed by the inhibition of serine/threonine protein phosphatases type 1

and 2A (PP1, PP2A) results in protein phosphorylation imbalance. The imbalance

further causes disruption of the liver cytoskeleton, which leads to cytoskeletal damage,

necrosis and pooling of blood in the liver, with a consequent large increase in liver

weight and massive hepatic haemorrhage that leads to death (Honkanen et al., 1996;

Eriksson et al., 1990; Romanowska- Duda et al., 2002). Microcystin-LR has an LD50 (the amount of a material, in a single dose, which causes

the death of 50% of a group of test animals) value of 36-122 μg/kg in mice and rats by

intraperitoneal injection or by inhalation (Stoner et al., 1989), which is comparable to

the toxicity of chemical organophosphate nerve agents (Oberholster et al., 2003). Acute

(short-term) exposure to high concentrations of microcystins can result in death from

liver haemorrhage or liver failure, while chronic (long-term) exposure to low doses may

encourage the growth of liver, kidney and other tumors (Jones and Orr, 1994; Codd and

Bell, 1996; Jones and Negri, 1997). The observed symptoms, associated with

microcystins, include diarrhea, vomiting, piloerection, weakness and pallor (Bell and

Codd, 1994).

The global reports of the intoxication of human (Table 1.1) and many animals by

microcystins at variable times and locations make it very difficult to investigate the

health threat at which microcystin concentration is of concern. Although 1 μg/L has

been recommended by the World Health Organization (WHO) as the maximum

allowable microcystin-LR concentration in drinking water (WHO, 1998), a proposed

value of 0.01 μg/L was suggested by Ueno et al. (1996) based on the occurrence of

primary liver cancer and the presence of microcystins in the surface waters in certain

regions of China.

Page 21: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

7

Table 1.1 Reported intoxication cases by microcystins

1.1.4 Microcystin persistence in drinking water treatment plants

Microcystins have been detected in surface waters used for potable water sources

internationally (Sivonen and Jones, 1999; Hitzfeld et al., 2000; Kabzinski et al., 2000),

where the main hazardous effect on humans is through consumption of drinking water

(Gupta, 1998). It is, therefore, vital for drinking water authorities world-wide to develop

reliable water treatment strategies for the removal of cyanobacterial cells and associated

toxins. The traditional water treatment technologies (coagulation / sedimentation / filtration)

have been documented to be insufficient at producing safe drinking water from

microcystin contaminated sources (Hoffman, 1976; Keijola et al., 1988; Anselme et al.,

1988; Lahti and Hiisvirta, 1989; Himberg et al., 1989). Microcystins, at varying

concentrations, have been detected in product drinking water in a number of countries,

including China (Wu et al., 2005), Australia (Bourke et al., 1983; Falconer et al., 1983),

Finland (Lahti et al., 1997) and Latvia (Eynard et al., 2000). The presence of

microcystins in drinking water is a major concern for water authorities. The main

reasons for the microcystins recalcitrance to conventional treatment are as follows:

Location and date Symptoms Consequences Reference

USA, 1931 Gastro-enteritis No data Tisdale et al.,

1931

Canada, 1959 Gastro-enteritis, headaches,

nausea, muscular pains 30 people affected

Australia, 1981 Gastro-enteritis, Liver

injury No data

Brazil, 1988 Gastro-enteritis 2000 people affected, 88

deaths

Chorus and Bartram, 1999

UK, 1989 Gastro-enteritis, vomiting,

sore throats 20 people affected, 2 hospitalizations

Turner et al., 1990

Brazil, 1996 Acute hepatic failure 50 dialysis patients died Jochimsen et al.,

1998

Page 22: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

8

First, the commonly applied conventional flocculation and filtration procedures are

generally identified to be effective only in reducing microcystins that remain in healthy

cyanobacteria, but do not remove the dissolved toxins that are known to be relatively

stable compounds. When cyanobacterial cells lyse, either from natural decay or from

artificial treatment, the intracellular toxins will be released into the water column. The

hepatotoxins are “small” molecules with a molecular weight ranging from 800-1000

Daltons (Botes et al., 1982a, b) and most congeners are hydrophilic, so they are

relatively soluble and stable in water. Moreover, they are non-volatile and resist

extremes in pH (Wannemacher, 1989). Microcystins have been reported to withstand

many hours of boiling and may persist for many years when stored dry at room

temperature (Linda, 1999). Hence, once released to the water column, microcystins

provide all the challenges of removing a water-soluble contaminant from water.

Additionally, the conventional treatment methods may even cause release of

intracellular toxins by system turbulence and pressure gradients during flocculation, as

well as application of pre-oxidation (Chow et al., 1998) and consequently pose an even

bigger challenge for microcystin removal. Therefore, a broad range of alternative

treatment methods (Section 1.2) that could provide more efficient and reliable

microcystin removal have been investigated.

1.2 Available treatment technologies in removing or reducing

microcystins in drinking water

There are several remedial technologies to compensation for the inadequate microcystin

removal during conventional water treatment. These methods mainly include oxidation

(chlorination, ozonation, potassium permanganate, UV, hydrogen peroxide and titanium

dioxide); dissolved air flotation (DAF); membrane filtration, biodegradation as well as

activated carbon adsorption.

Page 23: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

9

1.2.1 Oxidation

Chlorination

Chlorine has been used as a disinfection reagent for water treatment since the 19th

century (Lawton and Robertson, 1999) and it still is one of the most common chemicals

used in modern water treatment facilities. However, its efficiency has been found to

largely depend on the pH of the water. Chlorine gas dissolved in water forms

hypochlorous acid (HOCl) following the reaction

Cl2 + H2O → HOCl + H+ + Cl¯ (Equ. 1)

Because HOCl is weak acid, under pH 5, chlorine solution is primarily hypochlorous

acid which is an effective disinfecting agent. While above pH 5, hypochlorous acid

starts to dissociate and form hypochlorite ions that are comparatively less effective,

according to the reaction

HOCl ↔ OCl¯ + H+ (pH ≤ 5) (Equ. 2)

At pH values above 10, 100% dissociation of hypochlorous acid will occur as

HOCl → OCl¯ + H+ (pH > 5) (Equ. 3)

Application of chlorination for the removal of microcystins was carried out by

Hoffmann (1976), Keijola et al. (1988), and Himberg et al. (1989), but all indicated the

process was ineffective. This may be due to the fact that the pH at which the work was

performed was unfavorable for the formation of hypochlorous acid (Lawton and

Robertson, 1999). Contrary to the above results, chlorine has been demonstrated as an

effective oxidant for the destruction of microcystins (Nicholson et al., 1994; Tsuji et al.,

1997; Senogles-Derham et al., 2003). Nicholson et al. (1994) showed that

microcystin-LR was effectively destroyed provided there was a chlorine residual of at

least 0.5 mg/L with contact time of 30 minutes. Likewise, Tsuji et al. (1997) reported

that 99% of microcystin-LR was removed with a chlorine dose of 2.8 mg/L after a

contact time of 30 minutes. As in the results of Hoffmann (1976), the effect of pH on

chlorination of microcystins was also shown by Nicholson et al. (1994), who reported

Page 24: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

10

that the most efficient microcystin destruction occurred at pH values between 5 and 8

with a decreasing efficiency above pH 8.

In general, chlorination does appear to be an effective method for removing

microcystins from drinking water, although this is largely dependent on the pH, dose

level and maintenance of the residual chorine in the water. Moreover, the products and

the mechanism of microcystin decomposition still remain to be characterised; therefore

harmful by-products may be as yet unknown.

Ozonation

Ozone, one of the most powerful oxidizing reagents, has been widely investigated for

cyanotoxin removal in the last 10 years. Early studies conducted by Keijola et al. (1988)

showed 100% removal of up to 60 μg/L microcystins was achieved with 1 mg/L ozone.

Likewise, a 99% removal of microcystins after 15 seconds when treated with 0.05 mg/L

ozone was found by Rositano et al. (1998). This extremely fast destruction of

microcystins was corroborated in other Australian studies (Nicholson et al., 1993;

Rositano and Nicholson, 1994; Rositano et al., 1998), which identified that ozone was

more effective than chlorine, hydrogen peroxide and potassium permanganate for the

destruction of microcystins provided residual ozone is guaranteed.

Use of ozone is sometimes unpredictable and its reliability to remove microcystins from

water is highly dependent upon the quality of the water, which is noted to be the most

significant parameter in the determination of the ozone dose required for the removal of

the cyanotoxins. In particular, the nature and concentration of the dissolved organic

material in the treated water body have been found to affect the oxidation reaction, and

may result in incomplete toxin removal (Hart and Stott, 1993; Carlile, 1994; Rositano,

1996; Mouchet and Bonnelye, 1998; Rositano et al., 1998, 2001; Shawwa and Smith,

2001).

Page 25: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

11

Potassium permanganate

Potassium permanganate, a powerful oxidizing agent capable of destroying organic

compounds and microorganisms, has been widely applied since the 1960s. It is

commonly used by organic chemists for the hydroxylation of alkenes for diol formation.

Therefore, it should be effective in removing microcystins by attacking the unsaturated

bonds.

Several studies have shown that permanganate was effective in removing microcystins.

Carlile (1994) reported a removal of 76% microcystin-LR with a low dose of 0.7 mg/L

permanganate, a higher removal of 88% with a dose of 1 mg/L, and a nearly 100%

removal with doses greater than 1 mg/L. Similarly, Fawell et al. (1993) found that

permanganate is efficient for microcystin removal in both raw and treated waters. They

reported that a dose of 2 mg/L could destroy microcystins to below detection limit.

Rositano et al. (1998) found that more than 90% of microcystins at 1 mg/L could be

oxidized by a dose of 2 mg/L permanganate.

Potassium permanganate appears to show much promise as an oxidising agent, but

preoxidation with permanganate followed by coagulation may cause algal cells to lyse,

releasing additional amount of toxins, and also results in manganese being present in the

treated water (Lam et al., 1995; Pietsch et al., 2002). Moreover, little is known about the

character of the by-products from this decomposition.

Ultraviolet (UV) light

For the removal of some cyanotoxins, photo-oxidation has been shown to be a

remarkably effective method. For example, 45-56% of microcystins, which are

chemically and physically stable compounds (Tsuji et al., 1994), were destroyed after 8

hours of natural sunlight irradiation (maximum cumulative UV-radiation was about 30

W/m2). However, the dose requirement of this method often exceeds the range that is

practical for water treatment application. In Carlile’s (1994) study, it was shown that a

Page 26: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

12

dose of 22.5 W/cm2, which is two orders of magnitude greater than that needed for

disinfection, was required to achieve 90% removal of anatoxins, and to remove 91% of

microcystins-LR, a dose of 24 W/cm2 was required.

Hydrogen peroxide

In general, the use of hydrogen peroxide alone in water treatment is very limited. It has

only been applied to the oxidation of phenolic wastewater and to the treatment of paper

mill effluent, drilling muds and other types of organic wastewater (Lawton and

Robertson, 1999).

Hydrogen peroxide is relatively ineffective in degrading microcystins. Only 17%

cyanotoxin removal was obtained after 60-minute contact time with a dose of 20 mg/L

hydrogen peroxide (Rositano and Nicholson, 1994). Researchers who investigated the

effect of a 2 mg/L hydrogen peroxide solution on 1 mg/L microcystin-LR observed no

toxin removal after 10 minutes. Although the oxidizing effectiveness of hydrogen

peroxide seems to be enhanced by irradiating with UV or ozone, where all toxins are

removed within 30 seconds (Lawton and Robertson, 1999), earlier research with UV

alone showed similar destruction, suggesting a negligible contribution from the

hydrogen peroxide. Therefore, this reagent appears not to be promising as an effective

treatment method for microcystins.

UV photocatalysis Titanium dioxide

Titanium dioxide under UV photolysis has been found to be efficient for

cylindrospermopsin removal at pH 9, where there would be greater production of

hydroxyl radicals (Senogles et al. 2001). In contrast, Feitz et al. (1999) reported that the

TiO2 catalyst rapidly degraded microcystin-LR at low pH 3.5, while at higher pH values,

a distinct lag was observed before the toxin started to be degraded. This may be due to

the adsorption of microcystin-LR to TiO2 at high pH, causing degradation by persistent

organic radicals (Feitz et al., 1999).

Page 27: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

13

No toxic by-products have been found during the TiO2 induced photocatalytic oxidation

of microcystin-LR, because the major mechanism of photocatalysis was to isomerize,

substitute and cleave the Adda conjugated diene that is generally associated with the

toxicity of microcystin (Liu et al., 2002). Recent investigations also found that the

photocatalytic oxidation of microcystin-LR by TiO2 and UV was enhanced with the

addition of H2O2 (Cornish et al., 2000; Liu et al., 2002). They reported that the

combination of H2O2/ TiO2/UV resulted in no detectable by-products, with a bioassay

indicating that the toxicity of the treated water had been removed. However, in a study

by Senogles et al. (2001), the efficiency of the photocatalytic degradation was greatly

reduced at elevated concentration of dissolved organic carbon (DOC) between 15-32.5

mg/L. The authors attributed this to the decreased transmission caused by higher levels

of DOC.

Although titanium dioxide seems to be an effective technology to remove microcystins,

its removal efficiency can be easily affected by environmental conditions (such as pH

and elevated concentrations of DOC).

1.2.2 Membrane filtration

Many types of membrane filtration have been reported to be effective in microcystin

removal. Microfiltration (MF) and ultrafiltration (UF), for example, are both emerging

membrane filtration technologies, which have been found very effective (>98%) in

removing whole cells of microcystis (Chow et al., 1997). An 82-99% removal of

dissolved microcystin-LR has also been reported by Simpson and Macleod (2002), as

well as by Smith et al. (2002) who trialed eight different nanofiltration (NF) membranes.

Similarly, three different reverse osmosis (RO) membranes have been shown to be able

to efficiently remove m-LR and m-RR from potable water, with average removal rates

of 96.7% and 99.9%, respectively (Neumann and Weckesser, 1998).

Page 28: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

14

Generally, low pressure MF and UF membranes are better for removal of intracellular

cyanotoxins, while high pressure membranes (NF and RO) may have the ability to

remove extracellular dissolved cyanotoxins. However, when considering filtration, an

important point is the lysis of cells. In the case of the above-cited studies, some damage

to cells was observed (Chow et al., 1997).

1.2.3 Biodegradation

Biological treatment of contaminants in drinking water has been utilised by water

suppliers worldwide for many years. It generally possesses many advantages; for

example, 1) requires relatively little maintenance technology, 2) relates to relatively low

infrastructure and running costs, 3) it does not require any additional treatment and only

involves processes that remove contaminants without the addition of chemicals that in

themselves may have potential health effects (Newcombe et al., 2001).

In recent years, specific attention has been given for the application of biodegradation

for cyanotoxin removal, especially microcystin removal (Watanabe et al., 1992; Jones

and Orr, 1994; Lam et al., 1995; Cousins et al., 1996; Tsuji et al., 1996, Christoffersen

et al. 2002). Christofferesen et al. (2002) reported that a great number of microcystins

in the environment are degraded by microorganisms, though they generally need a lag

phase for adaptation. In the sediment infiltration study conducted by Holst et al. (2003),

microcystins were detoxified by indigenous microorganisms in the sediment located in a

water recharge facility, specifically at anoxic (<0.3% O2) conditions.

It was believed that the high efficiency of biodegradation is due to specific aquatic

microorganisms that effectively prey on cyanobacterial cells (Jones, 1994; Ho, 2004).

Up to date, several strains of bacteria, such as Pseudomonas aeruginosa and

Sphingomonas sp., have been identified as microcystin degraders (Takenaka and

Watanabe, 1997; Lahti et al., 1998; Park et al., 2001; Christoffersen et al., 2002; Saito

et al., 2003; Harada et al., 2004 and Ishii et al., 2004).

Page 29: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

15

The efficiency of microcystin removal by biodegradation is also affected by the

selection of the media on which bacteria colonise. For example, more than 90% of

microcystins were biologically degraded using a sand filter (Grützmacher et al., 2002;

Ho et al., 2006). In a batch scale bank filtration experiment, complete microcystin

removal took 10 days using soil containing high organic and low sand content, but 16

days using soil with low organic and high sand content (Miller and Hallowfield, 2001).

The authors indicate that the difference could be due to insufficient nutrient availability

for microorganism growth in low organics.

Successful removal of cyanotoxins employing biological filtration also depends on

many other operational conditions, such as the presence and concentration of certain

types of bacteria, the quality of the water being treated, the contact time and hydraulic

loading of the filter. Furthermore, the delay (days to months) for biological degradation

to occur, which is often referred to as the lag phase, is still a major hindrance for the

confident application of biological filtration technology (Newcombe et al., 2002).

Researches in the United Kingdom indicated a lag period of less than a week when

m-LR at 10μg/L was used in reservoir water (Cousins et al., 1996). In the study

conducted by Codd and Bell (1996), a lag phase of about 1 week was observed. In

contrast, a lag phase of up to 21 days was observed by Jones and Orr (1994) when

studies were carried out after cyanobacterial bloom treatment with algicide (i.e. copper

sulfate).

1.2.4 Disadvantages from these treatment solutions

Although the previously discussed remedial treatment techniques can reduce

microcystins relatively effectively, they usually exhibit many limitations that usually

make them outside commonly adopted protocols. For example, oxidation by chlorine is

a well established and effective one-step post-treatment technology, but for toxin

removal a multi-barrier approach is desirable at the end of water treatment process.

Page 30: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

16

Ozonation is a very rapid oxidation option but the high cost may prove to be prohibitive,

particularly since the appearance of microcystins in water sources is typically seasonal

and unpredictable. Ultraviolet (UV) light treatment appears to be promising, for it is a

simple and clean technology, but required UV doses for the purpose of microcystin

removal are significantly higher than those generally used for disinfection, also this

technology is relatively new and thus remains to be seen how well it will perform.

Moreover, further research must be carried out to examine the potential hazardous

by-products associated with these oxidation methods. As for the option of filtration, in

particular membrane filtration, the release of free toxins and the increase in the

operational costs due to the high pressure drops make this type of treatment unsuitable

for large-scale applications.

Therefore, it is vitally important to utilise an effective microcystin removal method,

which should be cost-effective and sustainable, relatively easy to operate, and with

minimal chance for generation of harmful by-products. Possessing all these traits, the

use of activated carbon (Section 1.3), especially granular activated carbon filtration

(Section 1.3.2), appears to be a very promising option.

1.3 Activated carbon Activated carbon (AC) adsorption has been widely used in drinking water, wastewater

treatment, as well as food, beverage, pharmaceutical and chemical industries. It has

been recommended as one of the best available environmental control technologies by

the US Environmental Protection Agency (EPA).

The initial application of AC by the water industry was to remove taste and odour

problems. Recently, many of the studies relating to AC have been conducted on the

removal of microcystins (Donati et al., 1993, 1994; Fawell et al., 1993; Craig and

Bailey, 1995; Lambert et al., 1996; Cook and Newcombe, 2002). Results from most of

Page 31: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

17

these studies indicate that the removal efficiency of AC filters is greatly dependant on

the type of AC being used, and suggest that chemically-activated coal and wood-based

ACs are the best options for microcystin removal (Donati et al., 1994; Hart and Stott,

1993; Lambert et al., 1996) due to their large volume of mesopores (2-50 nm), which

are the appropriate size for the microcystin molecules (Donati et al., 1994).

There exist two forms of AC, namely, powdered activated carbon (PAC) and granular

activated carbon (GAC).

1.3.1 Powdered activated carbon

PAC is commonly added directly to water prior to coagulation or filtration for

microcystin removal. Although such application has been revealed to be effective by

many studies, it has some disadvantages.

Firstly, its effective dosing rates are always very high. For example, in the investigation

conducted by Lawton and Robertson (1999), a PAC dose of 800 g/m3 was required to

remove both the microcystin-LR and microcystin-RR. Moreover, Falconer et al. (1989)

found that only very small percentages of the toxins could be removed by many types of

the PACs in their investigation at a dosing level of 1 kg/m3.

Secondly, the removal efficiency of PAC is dependent upon several operation conditions.

One condition is the microcystin concentration in water being treated. At a full-scale

Canadian treatment plant, a PAC supplement at 30 mg/L was monitored over 6 weeks

(Lambert et al. 1996). The authors found that the average removal of all active

microcystins ranged from 82% when the raw water microcystin concentrations were

above 0.5 µg/L to only 31% when the concentrations were below this level.

A particular condition that can affect the removal rate of PAC is the concentration of

natural organic material (NOM) in the source water. A study comparing PAC removal

Page 32: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

18

efficiency for different water qualities was conducted (Lawton and Robertson, 1999),

using both high purity waters and raw waters. The results showed that the toxin

adsorption to PAC was greatly decreased by the high NOM concentration in raw water,

due to competition for binding sites on the carbon.

Therefore, the efficient PAC dose rate for microcystin removal is related to the type of

PAC being used, and the concentration of both the microcystins being investigated and

the natural organic carbon in the water being treated. In contrast, the use of GAC

filtration (Section 1.3.2), that has less of the disadvantages listed above, can be a very

efficient process.

1.3.2 Granular activated carbon filtration

Granular activated carbon (GAC) is typically utilised in flow-through column reactors

in water treatment processes to control various micropollutants such as taste and odor

compounds, heavy metals and pesticide residue etc. It has also been shown to be one of

the most effective options for removal of the cyanotoxins (Falconer et al., 1989; Lahti

and Hiisvirta, 1989; Jones et al., 1993; Donati et al., 1993; Craig and Bailey, 1995;

Lambert et al., 1996; Cook and Newcombe, 2002; Ho, 2004). The high toxin removal

efficiency of GAC is attributable to the co-existence of adsorption and biodegradation

mechanisms, both of which are significant contributing factors as microcystins are both

adsorbable (Section 1.3) and biodegradable (Section 1.2.3). Of particular interest,

biodegradation of microcystins could considerably extend the service life of GAC filters,

which would reduce the cost of this technology and facilitate its practical application.

Firstly, GAC is an effective microcystin adsorbent. Previous studies with bench-scale

treatment processes (including alum coagulation, sand filtration, GAC adsorption and

chlorination) found that toxins from freeze-dried Microcytis and Oscillatoria blooms

were essentially eliminated at feed concentrations between 30 and 56 µg/L (Keijola et

al., 1988 and Himberg et al., 1989). Since the conventional treatment processes have

Page 33: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

19

been reported to have negligible removal of dissolved microcystins, most of the

achieved removal could be attributed to GAC adsorption.

In the above GAC studies by Keijola et al. (1988) and Himberg et al. (1989), the treated

volume of the polluted water was only 10 bed volumes, thus the GAC capacity for toxin

adsorption could not be seriously challenged and therefore complete removal of

microcystins was observed. However, a study by Newcombe (2002) showed

breakthrough of microcystins after a GAC filter was operated for 6 months with water

that had a relatively high NOM concentration. The author pointed out that it may be due

to the high loading of NOM in the water sample that reduced the adsorption capacity of

the GAC filter. Similarly, research conducted on a full-scale GAC adsorber by Lambert

et al. (1996) found that NOM may expend the adsorption capacity in GAC, which

would reduce the removal efficiency for microcystins. Again, tests conducted on GAC

that had been used in a water treatment plant for 5 months before evaluating its

adsorption capacity for microcystins, demonstrated that 88% of the adsorption capacity

of the GAC was reduced (Lawton and Robertson, 1999).

The findings presented above indicate that although GAC adsorption filters have been

increasingly employed as an effective process for microcystin removal in modern

treatment works, they may be easily saturated with NOM in normal water treatment

circumstances resulting in toxin breakthrough after a relatively short life-time ranging

from hours to hundreds of days (Falconer et al., 1983, 1989). This time uncertainty due

to the adsorption exhaustion not only results in very high operating cost due to the

frequent thermal regeneration required, but also makes it is difficult to estimate how

long a GAC filter could last as an effective adsorbent to assure microcystin-safe water.

It is, therefore, important to determine the microcystin adsorption capacity of GAC or

the toxin breakthrough into the product drinking water.

Importantly, GAC packed columns could be used as a microcystin biodegradation

Page 34: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

20

reactor. As documented previously (Section 1.2.3), microcystins can be decomposed by

a wide variety of microorganisms in the environment, but the efficiency of this process

will be affected by the selection of the substrate on which bacteria can attach. With

GAC filters, given there is no disinfectant residual left in the treated water, a biological

layer (biofilm) can be very quickly cultivated on the rough and porous surface of GAC

particles (Miklas and Robert, 1998). Therefore, the overall toxin removal by the GAC

column would be greatly enhanced and the overall lifetime of the GAC filter would also

be extended.

However, the biodegradation process can only occur after a lag-phase of several days up

to a few months (Jones and Orr, 1994; Lam et al., 1995; Cousins et al., 1996; Ho, 2004).

The lag-phase is required for the microorganisms to acclimatize to the toxins as a

carbon source (Jones and Orr, 1994). The length of the lag time depends on many

operational conditions, such as the presence and concentration of certain types of

bacteria, the quality of the water being treated, the contact time and hydraulic loading of

the filter. Thus, it is uncertain when the biodegradation mechanism will occur under

normal operating conditions for a wide range of treated water qualities. The variability

of the lag-phase before onset of biodegradation, combined with the unknown lifetime of

the filters functioning in adsorption mode, results in a serious uncertainty in the

application of GAC for the removal of microcystins.

At any time during the use of GAC filters, it is unknown which mechanism, adsorption

or biodegradation, takes the primary responsibility for microcystin reduction in GAC

filters. Additionally, there is a lack of knowledge of the extent to which the remaining

adsorption capacity contributes to toxin elimination after biodegradation of microcystins

is triggered on in GAC filters.

Therefore, a comprehensive study that can separate the mechanism of adsorption from

biodegradation of the toxins in GAC filters would be extremely beneficial for further

Page 35: Adsorption and biological filtration of microcystins

Chapter 1 Introduction

21

understanding of the removal principles during the filtration and would also make it

possible to answer the above-mentioned problems.

1.4 Objectives of this research

The primary objective of this project was to determine the relative contributions of the

physical adsorption and the biological process for the removal of microcystins in GAC

filtration. This knowledge was obtained by designing and carrying out detailed GAC

experiments that separated the two mechanisms.

In addition, the GAC experiment was also used to investigate simultaneous natural

organic material (NOM) removal during GAC filtration, the disinfection by-products of

which have been recognized as potential public health threats since the 1950s

(Middleton and Rosen, 1956). As an effective NOM adsorber and potential

biodegradation reactor, GAC filters have been extensively studied for effective

elimination of NOM. However, these studies have rarely been conducted with the

unique Australian waters that have high concentration of NOM, and did not separate the

removal process regarding adsorption and biodegradation.

Finally, to facilitate the practical application of microcystin biodegradation in GAC filters,

this study also aimed to identify the potential effects of different operational conditions on

the behavior of the biological processes.

Page 36: Adsorption and biological filtration of microcystins

- 22 -

CHAPTER 2

EXPERIMENTAL METHODS AND MATERIALS

2.1 Experimental methods

The lab-scale, column experiment was designed to discriminate between the adsorption

and biodegradation mechanisms for microcystin removal, which were expected to occur

in GAC filters simultaneously.

2.1.1 Experiment setup

To separate the adsorption and biodegradation mechanisms during GAC filtration, three

lab-scale glass columns (25 mm I.D. and 15 cm bed depth) were constructed and

operated in parallel (Figure 2.1). These three columns included a non-sterile GAC

column, a sterile GAC column and a sand column to simulate the combination of

adsorption and biodegradation, solely adsorption and solely biodegradation, respectively.

The sterile GAC was from the same source as the non-sterile GAC, but was maintained

under sterile conditions, which was achieved by regular (weekly) autoclaving of the

associated experimental apparatus (at 121 ºC, 500 kPa for 20 minutes) for the duration

of the experiment. Thus, biological activity and/or biodegradation was minimised in this

column with only the adsorption mechanism occurring. The sand in the third column

was non-absorbing and thus any microcystin removal in this column could be only

attributed to biodegradation. A brief summary of the three columns is shown in Table

2.1.

Page 37: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

23

filter 3

collect effluent water

1

20L

3

2

autoclavable plastic tank

20Lgo to drainage

GACfilter 2

filter 1Sand

1

2

1

2

taking sample

magnetic stirrer

glass tank

pump

3

Figure 2.1 schematic diagram of the experiment setup

The Sterile GAC filter system

Page 38: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

24

Table 2.1 Operating parameters for the three columns (Figure 2.1)

Component Sterile GAC column Sand column Non-sterile GAC

column

Packing material Sterile virgin GAC I.D. 1.0 – 1.4 mm

Non-porous fresh sand I.D. 1.0 -1.4 mm

Sterile virgin GAC I.D. 1.0 – 1.4 mm

Operating condition Aseptic Normal (non-aseptic) Normal (non-aseptic)

Microcystin removal mechanism(s)

Adsorption Biodegradation Adsorption and biodegradation

2.1.2 Selection of the experimental design

Investigations concerning the separation of the adsorption and biodegradation

mechanisms in GAC filters using sterile and non-sterile carbon beds have been

previously carried out by Laat et al. (1985). The authors added silver sulfate (at Ag+

concentration of 0.5-1.0 mg/L) to maintain sterile condition in one of the filters.

However, the silver sulfate adsorption on GAC (Jia and Demopoulos, 2002) could affect

the adsorption efficiency of microcystins that are typically trace-level (micro-gram per

litre) compounds. Therefore, the innovative experimental design (Figure 2.1)

concerning autoclaving of the sterile GAC system was employed in this study.

There are a number of advantages in selecting non-porous sand as the packing material

for assessing biodegradation rather than using spent GAC. Firstly, it could not be

guaranteed that the saturated GAC would not retain any adsorption capacity for any of

the adsorbable DOC constituents present in the water. The equilibrium of adsorption is a

kinetic balance which can change in response to variable factors in the water such as pH

and DOC. The non-porous sand, on the other hand, possesses no adsorbability itself, so

it could exclusively simulate the biodegradation in GAC filtration. Secondly, a

bio-regeneration effect, which could refresh the exhausted adsorption sites to a certain

Page 39: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

25

degree, may occur once biological activities become significant on the GAC surface. In

that case, the adsorption ability in the GAC would be re-exhibited.

2.1.3 Operation of the columns

During the experiment (Figure 2.1) microcystin (m-LR and m-LA) spiked water was

transferred to a 20 L tank and mixed with a magnetic stirrer. The water was continually

pumped through the three columns via a three-port peristaltic pump (Adelab, Australia)

at an empty bed contact time (EBCT) of 15 minutes. Samples (250 mL) before and after

each column were collected at regular time intervals for subsequent DOC and UVA254

measurement, as well as SPE extraction and HPLC analysis of the toxin concentration.

In order to minimise biological contamination of the sterile GAC column, a separate

water reservoir and apparatus for collecting samples were constructed for the system;

these are highlighted in dashed rectangles in the schematic diagram (Figure 2.1). The

separate reservoir was a 20 L polypropylene autoclavable carboy with spigot (Southern

Cross Science Pty Ltd, Australia). The water was sterilised in the carboy by autoclaving

(at 121 ºC, 500 kPa for 20 minutes) prior to use. To minimise any bacteria in the toxin

spiking solution, the toxins were injected into the carboy via a 0.22 µm sterile filter disc

(Gelman Sciences, Australia), which was then rinsed with 30 mL autoclaved Milli Q

water to wash out the residual toxins into the carboy. The spiking procedure was

conducted following aseptic techniques in a laminar flow cupboard.

To assess the sterility of the above procedures heterotrophic plate counts were employed.

This method employed R2A agar plates (Oxoid Limited Corporate, UK) on which 100

μL of the effluent sample (taken from the system at daily intervals) was spread

following aseptic methods. After incubation at 25 °C for 7 days, the number of bacterial

colonies (count/mL) obtained on the plate was recorded.

Page 40: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

26

2.1.4 Backwashing methods

The columns were backwashed weekly using distilled water for approximately three

minutes at a 10% bed expansion (Figure 2.2). This was conducted to simulate WTP

filters which are routinely backwashed to remove any algal growth and air bubbles that

are trapped in the columns and to redistribute the media. Cleaning of the entire

apparatus including tubing, connectors and feed tanks was also undertaken weekly to

remove any algal or particle deposits.

Figure 2.2 Demonstration of the backwashing technique

The weekly backwashing procedure was also carried out for the sterile GAC column

with non-sterile distilled water; however, autoclaving (at 121 ºC, 500 kPa for 20

minutes) of the sterile GAC was performed after backwashing to maintain sterile

conditions.

2.1.5 Scanning electron microscopy (SEM) study

To evaluate the development of biofilm on the surface of the GAC and sand particles,

samples were taken from the columns for SEM analysis (Section 2.3.4) at two, four and

six monthly intervals during the experiment. Samples containing approximately 20

particles were collected from the top portion of each column after backwashing.

Page 41: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

27

2.1.6 Bacterial Enumeration

To quantify the bacteria concentration in the GAC and the sand columns when

microcystin biodegradation was evident in both of the systems (day 220), GAC and

sand samples of the same volume (5 cm3) were withdrawn from the corresponding

columns after backwashing. To detach the biofilm, the GAC and sand samples were

separately vortexed in 20 mL glass test tubes containing sterilised Myponga water for 2

minutes. The resultant supernatant with bacteria was immediately decanted and

collected to avoid any bacteria reattachment. The above procedures were repeated for

five times for each of the samples. The collected bacterial solutions were mixed

thoroughly using a magnetic stirrer. Bacterial concentration in these solutions was then

analysed using flow cytometry (Becton Dickinson Biosciences, Australia), details of

this method were presented by Ho (2004).

2.2 Reagents and materials

2.2.1 Reagents

All reagents used were high performance liquid chromatography (HPLC) grade or

analytical grade. Aqueous solutions were made using deionised water (DOC ≤ 0.07

mg/L, 18 MΩ conductivity; Millipore Pty Ltd, USA) produced using a Milli-Q system.

2.2.2 Microcystin

Microcystin LR standard (purity > 98.0% by HPLC) was obtained from a commercial

supplier (Sapphire Bioscience, Australia). The spiking microcystin solution which

contained both m-LR and m-LA was extracted from Microcystis aeruginosa blooms that

occurred in the Gippsland Lakes, Victoria, Australia. The isolation and purification of

the toxins were conducted at the Australia Water Quality Centre. This procedure (Ho,

2004) involved freeze-thawing the bloom material in methanol and water three times in

Page 42: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

28

order to lyse the algal cells and release the intracellular toxins. The resulting solution

was sonicated for 30 minutes (FX 10 ultrasonicator from Unisonics Pty Ltd, Australia)

and then centrifuged (Allegra 6 Centrifuge from Beckman Instruments, USA) for one

hour. The separated toxin supernatant was decanted and filtered through Whatman GF/C

filter membranes (Adelab Scientific, Australia).

The solid pellet obtained from centrifuging was resuspended in 75% methanol solution

and treated following the same procedures as described above to extract any residual

toxins. All the collected toxin solutions were combined and rotary-evaporated (Buchi

Rotavap System from Medos Company Pty Led, Australia) to remove the methanol

constituent. The final purification involved passing the toxin solution through

preparative reverse phase flash chromatography and preparative HPLC. The received

spiking solution of microcystins was stored in a freezer (~ 4 ºC) prior to use.

2.2.3 Granular activated carbon (GAC)

A commercially available GAC, A6 (PICA, Australia) was selected for this study. It is a

coal based, steam-activated GAC that has a relatively high micro- and meso-pore range.

Additionally, the hardness of the GAC makes it suitable for regeneration. The

physico-chemical properties of the GAC are listed in Table 2.2.

Table 2.2 Physico-chemical properties of GAC A6

The desired GAC particle diameter of 1.0-1.4 mm was achieved by screening between

1.0 mm and 1.4 mm opening size sieves. The sieve tower was shaken for 15 minutes at

Page 43: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

29

an angle of 15° to 20° and rotated 90° every 3 minutes. To minimize the potential

effects from the water soluble contaminants, the sieved GAC was then rinsed

thoroughly using deionised water and transferred to a 100 °C oven to dry overnight.

After the GAC reached a constant weight, it was then stored in a desiccator prior to use.

2.2.4 Sand

New sand (River Sands Pty. Ltd. Australia) of the same particle size range as the GAC

was selected for this study. The sand was non-porous, therefore, it was assumed to have

minimal adsorption capacity. Pre-treatment for the sand was performed following the

same procedure as the GAC.

2.2.5 Water samples

The water used in this study was obtained weekly from the Myponga water treatment

plant (WTP) (South Australia) and was transported to the laboratory in 25 L water

containers. The collected water had been treated via the processes of alum coagulation,

dissolved air flotation and rapid sand filtration. The water samples were taken after

these processes prior to disinfection with chlorine. The water had a comparatively high

DOC concentration of approximately 6-7 mg/L.

2.3 Analytical methods

2.3.1 Microcystin analysis

Water samples (250 mL) containing microcystins were concentrated by solid phase

extraction (SPE) using 500 mg Sep-Pak Vac 3cc C18 cartridges (Waters Pty Ltd,

Australia) following the methods given by Nicholson et al. (1994). Water samples were

filtered through 0.45 µm cellulose nitrate membrane filters (Schleicher and Schuell,

Germany) to remove any particulate matter prior to extraction by the cartridges, which

Page 44: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

30

were pre-conditioned with 10mL methanol and 10 mL deionised water. During

extraction, water samples of one in every five were spiked with 250 μL of 5 mg/L m-LR

standard as an internal efficiency control. Samples were then forced through the

cartridges by application of a vacuum chamber (Alltech Associates, Australia) at a flow

rate of less than 10 mL/min. The cartridges were washed sequentially with 10mL of

deionised water, 10% methanol and 20% methanol. After vacuum drying (for

approximately 2 minutes), the concentrated toxins in the cartridges were eluted into 10

mL glass test tubes with 10 mL 100% methanol. The eluants were then dried in a Ratek

Dry Block Heater (Adelab Scientific, Australia) at 40 °C under a nitrogen stream. The

residues were then taken up with 500 μL of methanol and 500 μL of deionised water.

The obtained 1mL concentrated extracts were finally filtered through 0.45 μm

polyvinylidene fluoride (PVDF) filters (Gelman Sciences, Australia).

The high performance liquid chromatography (HPLC) methods described by Ho (2004)

were used for the analysis of microcystins in this study. The HPLC was equipped with a

Waters 717plus autosampler, which sequentially injected 50 μL sample from the 1 mL

concentrated extracts. The flow rate was controlled by a Waters 600 pump controller at

a flow rate of 1 mL/min; a Waters 996 photodiode array detector (Waters Pty Ltd,

Australia) was set to the wavelength of 238 nm. The HPLC system was controlled by a

computer equipped with the Waters Millenium 32 software, which also collected and

stored the analytical data. The chromatography separation was carried out on a 150 x

4.6 mm I.D. column packed with 5 μm pore size Luna C18 (Phenomenex, Australia).

The mobile phase consisted of a gradient mixture of solvent A (30% acetonitrile +

0.05% trifluoroacetic acid) and solvent B (55% acetonitrile + 0.05% trifluoroacetic

acid).

2.3.2 Dissolved organic carbon and UV254 absorbance

Dissolved organic carbon (DOC) and UV254 absorbance (UVA254) measurements were

conducted using an 820 Total Organic Carbon Analyser (Sievers Instruments Inc, USA)

Page 45: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

31

and a UV/VIS 918 Spectrophotometer (GBC Scientific Equipment Pty Ltd, Australia) at

the wavelength of 254 nm. Samples were initially filtered through 0.45 μm cellulose

nitrate membrane filters (Schleicher and Schuell, Germany). UVA254 was measured in 1

cm quartz cells using deionised water (Millipore Pty Ltd, USA) as a reference.

2.3.3 High performance size exclusion chromatography

Prior to analysis samples were filtered through 0.2 μm cellulose nitrate membrane filters

(Schleicher and Schuell, Germany) and stored at -18 °C until measurement. The HPSEC

system consisted of a Waters 2690 Alliance System and built-in pumps, column heater

and mobile phase degasser as well as a Waters 996 photodiode array detector

UV-detection system (Waters Pty Ltd, Australia). The column was a Protein KW-802.5

column (diameter = 8 mm, length = 300 mm; Waters Pty Ltd, Australia). The mobile

phase solution [920 mM sodium di-hydrogen phosphate (NaH2PO4) at pH 6.8] and

samples were injected at a flow rate of 1 mL/min. Results were analysed using

polystyrene sulphonates (35000, 18000, 8000 and 4600 Daltons) and acetone as

standards (Chin et al., 1994).

2.3.4 Scanning electron microscopy

SEM was used in this study for investigation and comparison of the biological growth

on surfaces of both the GAC and the sand. The microscope was a Philips XL 30 field

emission scanning electron microscope with an operating voltage of 10 kV. GAC

samples taken from the column were immediately transferred into a 4 %

paraformaldehyde/1.25 % glutaraldehyde solution in lead sulphide (PBS) + 4 % sucrose

at pH 7.2 to fix overnight. The fixed samples were twice washed with PBS + 4 %

sucrose solution followed by 30 minutes post-fix in 1% OsO4 in PBS. The GAC

particles were subsequently dehydrated with gradient ethanol (70%, 90%, 95% and

100%) solutions for 10 minutes, twice for each step, except the 100% solution which

was used three times. A Balzers CPD 030 critical point dryer was then employed to dry

the samples. Finally, the prepared GAC particles were mounted on stubs and coated

Page 46: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

32

with carbon before SEM imaging (Ho, 2004).

2.4 Validation of the experimental design - side effects from

the sterilisation method of autoclaving The autoclave process, which is defined as “an apparatus using superheated steam under

high pressure in order to decontaminate materials or render them sterile”

(Merriam-Webster Dictionary) was used in this study to sterilise the influent water, the

GAC and the apparatus of the sterile GAC system. It was possible that the original

characteristics of the autoclaved items could change and the adsorption capacity of

GAC could be varied to some extent (Salvador and Jiménez, 1996; Shende and

Mahajani, 2002). Thus, to ensure identical operating conditions between the sterile and

the non-sterile GAC systems, the following procedures were performed to assess any

side effects introduced during autoclaving.

2.4.1 Effects on the water quality

To determine any changes in the water quality and thus the adsorbability of the NOM

due to autoclaving, duplicate water samples before and after this procedure were

collected and analysed simultaneously in terms of DOC, UVA254 and HPSEC.

2.4.2 Loss of organic carbon from GAC structure and apparatus

It has been documented that up to 10% of carbon could be lost from the original GAC

structure at 650-850 °C during one set of thermal regeneration (Salvador and Jiménez,

1996). Although the autoclave temperature (121 °C) in this study was much lower, it

was carried out frequently (weekly) compared to the regeneration process, which is

normally undertaken every few months. Thus autoclaving of GAC may have led to

carbon loss. Moreover, at the lab-scale study even a very small percentage loss of

carbon would significantly affect the results. Therefore, it was necessary to quantify the

Page 47: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

33

possible loss of carbon from the autoclaved GAC.

To assess the carbon loss, a new GAC sample from the same stock of GAC used in the

column study was tested. It was autoclaved (at 121 ºC, 500 kPa for 20 minutes) for six

times in 150 mL of fresh deionised water, which was retained after each autoclave run

for UVA254, DOC and HPSEC measurements.

The apparatus of the sterile GAC system, such as the tubing and connectors could

potentially leach carbon during the autoclaving process. Thus, it was important to

determine the extent to which the leaching could occur. To test for carbon leaching, all

the apparatus were disconnected and separately autoclaved in individual Myponga

water samples (Figure 2.3). One control sample containing just the water was also

autoclaved. After autoclaving, the water in each sample was analysed for DOC

concentration.

Figure 2.3 Testing for carbon leaching from tubing and connectors belonging to the sterile GAC system during autoclaving

2.4.3 Effects on adsorption capacity of the GAC

The adsorption capacity of the sterile GAC was examined before and after autoclaving.

During this set of experiments, PAC A6 rather than the GAC A6 was examined. This

was because PAC possessed more specific surface area than GAC and is thus more

sensitive to any change. For the procedure, two identical slurry samples (10% w/v) of

Page 48: Adsorption and biological filtration of microcystins

Chapter 2 Experimental Methods and Materials

34

50 mg PAC A6 were prepared in distilled water. One sample was autoclaved (121 °C,

500 kPa for 20 minutes) while the other was untreated. After the autoclaved sample was

allowed to cool to room temperature, the two PAC slurries were individually added into

Gator jars, each of which contained 2 L of Myponga WTP water spiked with

microcystins. The stirring speed of the jar test was 100 rpm. Sub-samples of 100 mL

were collected at time intervals of 0, 5, 10, 20, 30, 60 and 90 minutes for analysis of

microcystins, DOC and UVA254.

2.4.4 Effects on desorption

Studies on the effect of autoclaving upon the pre-adsorbed substances in the sterile GAC

were carried out. It was considered important as under the high temperature of

autoclaving, various processes or physical-chemical reactions could occur and

potentially affect the compounds adsorbed on the GAC. For example, some volatile

compounds were found to desorb from pre-loaded GAC at temperatures exceeding 100

°C (Salvador and Jiménez, 1996). Furthermore, the frequent (weekly) autoclave

treatment in this study could have compounded the desorption effect to an extent where

the inherent adsorption capacity for the trace levels of microcystins could be increased.

To determine the amount of organic carbon that could be desorbed during autoclaving,

effluent samples (15 mL) were periodically collected from the sterile GAC column

shortly after autoclaving. Effluent of the column prior to autoclaving was also taken for

comparison. Collected samples were immediately filtered through 0.45 μm cellulose

nitrate membrane filters (Schleicher and Schuell, Germany) and then analysed for DOC

and UVA254.

To determine any desorption of microcystins from the sterile GAC due to autoclaving,

microcystin analysis was carried out on the water sample in which the GAC was

autoclaved (121 °C, 500 kPa for 20 minutes). These tests were carried out after the

GAC had been operated for 2.5 months.

Page 49: Adsorption and biological filtration of microcystins

35

CHAPTER 3

LABORATORY-SCALE COLUMN

EXPERIMENT: MICROCYSTIN REMOVAL 3.1 Introduction Although a few studies on microcystin removal using GAC filters have been undertaken

(Hart et al., 1998; Bernezeau, 1994; Hart and Stott, 1993, Jones et al., 1993; Craig and

Bailey, 1995), most of them put emphasis on its adsorption effect and not

biodegradation. In the study of Ho (2004), biological degradation of microcystins in

GAC filter was observed and considered a promising treatment option because of its

high removal efficiency and ability to extend the service life of the otherwise very

costly GAC filters. However, it was still unknown how long a newly established GAC

bed could be used for effective microcystin removal or its efficiency as a function of the

operation time. There is currently no information available on efficiencies of the two

individual mechanisms, adsorption and biodegradation, which are likely to occur

simultaneously in a GAC filter for microcystin removal.

The aim of the laboratory-scale experiment was to discriminate the relative importance

of adsorption from biodegradation of microcystins in GAC filters. As shown in Section

2.1, three laboratory-scale columns, i.e. the sterile GAC, the sand and the non-sterile

GAC columns, were operated in parallel for the replication of adsorption,

biodegradation and the combination of both, respectively, in GAC filtration.

3.2 Results from the first set of experiments Microcystin concentration in the influent and effluent of the three parallel columns

(Section 2.1) was monitored for approximately 230 days during this study. Results

Page 50: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

36

obtained from day 1 to day 65 were considered unreliable as an unsuitable SPE

cartridge was used, and this part of the experiment was repeated (Section 3.3). Results

after day 65 are presented in Figure 3.1 where concentration profiles of m-LR and

m-LA in both of the inlet and outlet of the three columns (the sterile GAC, the sand and

the non-sterile GAC column) are plotted against operation time. Two feed tanks

supplying the toxin spiked water were used for this study to separately supply feed

water to the sterile and the non-sterile GAC and the sand columns. Microcystin

concentration in the two reservoirs was similar (Figure 3.1). The degradation of the two

microcystin analogues (m-LR and m–LA) presented similar trends in the columns.

Details of the degradation are described as follows regarding each of the three columns.

Figure 3.1 Influent and effluent concentrations of m-LR (A) and m–LA (B) of the columns

during the first set of experiment

Page 51: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

37

3.2.1 The non-sterile GAC column A significant breakthrough of microcystins was evident in the effluent of the non-sterile

GAC column between days 66 to 100 (Figure 3.2). Approximately 40% m-LR and 60%

m-LA was detected in the column outlet. The breakthrough was not surprising as the

GAC column had been operating for two months (days 1 to 65) with water containing

relatively high concentration of NOM. Therefore, the GAC toxin adsorption capacity

would have been partially depleted by the presence of NOM. Similar breakthrough

results were also reported by earlier researchers including Craig and Bailey (1995) and

Ho (2004) who studied adsorption of microcystins using two-month old GAC.

0%

20%

40%

60%

80%

100%

66 86 106 126 146 166

Time (days)

mic

rocy

stin

rem

oval

per

cent

age

m-LR m-LA

Figure 3.2 Percent removal of m-LR and m–LA in the non-sterile GAC column (the dotted

circle indicates the reoccurred toxin breakthrough)

The adsorbability of m-LR was stronger than m-LA when GAC A6 was used for this

study, as considerably higher percentages of m-LR were adsorbed compared to m-LA

during the breakthrough period between days 66 to 100 (Figure 3.2). Similar

observations were reported from research by Cook and Newcombe (2002), Newcombe

et al. (2003) and Ho (2004), who investigated different types of GAC. However, this

outcome was unexpected because m-LA is of lower molecular weight (MW) and higher

hydrophobicity compared to m-LR, both of which are characteristics that would lead to

greater adsorption. It is possible that the adsorbed m-LA molecules on the GAC surface

presented strong electrostatic repulsion, which would negatively affect further

Page 52: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

38

adsorption (Newcombe et al., 2003) and lead to the lower adsorption of m-LA.

After breakthrough, toxin removal rapidly increased to 100% at approximately day 100.

This was attributed to biological degradation, as there was no significant change in the

operation conditions and it was very unlikely that microcystin adsorbability of the GAC

could be improved with time (Ho, 2004). It has also been shown that adsorption by dead

cells or organic materials would not be significant (Lahti et al., 1997; Holst et al., 2003).

The initial period of time prior to the occurrence of rapid biodegradation is known as

the lag-phase. This phase is commonly observed in the application of the biological

treatment technology and is a natural requirement as bacteria have to acclimatise to new

carbon sources. Further discussion on the lag-phase phenomenon is given in Section

3.5.2 and Chapter 6.

Unexpectedly, toxin breakthrough reappeared (as indicated by the dotted line in Figure

3.2) after the initial biodegradation commenced at approximately day 100. Similar

trends were not observed in Ho’s study (2004) even when the influent toxin

concentration was nearly doubled. Explanations for the second breakthrough were

unkonwn, although it could be attributed to the newly developed and less robust biofilm

on the GAC surface, which could be easily affected by any variation of the operation

conditions (temperature, water quality etc). Stable biological metabolism was reached

from day 130 as no further toxin breakthrough was detected from then on (Figure 3.2).

3.2.2 The sterile GAC column Biological activity was absent in the sterile GAC column during the experiment as

indicated by results of the regular agar assays (Appendix A). Therefore, adsorption was

the sole mechanism for microcystin removal. Toxin breakthrough in the sterile GAC

column was detected, which is presented in Figure 3.3. The data indicates that the

adsorption capability for microcystins gradually decreased as the adsorption sites were

depleted with time by NOM and microcystins.

Page 53: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

39

0

1

2

3

4

5

6

7

8

9

10

66 91 116 141 166 191 216Time (Days)

Mic

rocy

stin

con

cent

ratio

n (μ

g/L)

m-LR (sterile GAC) m-LA (sterile GAC)

m-LR (non-sterile GAC) m-LA (non-sterile GAC)

Figure 3.3 m-LR and m–LA concentration in the effluent of the sterile and the non-sterile GAC column from days 66 to100 (the dotted circle indicates the lower toxin breakthrough in the

sterile GAC column)

The sterile GAC, which simulated only adsorption, should present the same toxin

removal capacity as the non-sterile GAC before day 100 when adsorption was the

predominant function (indicated in Figure 3.3). If there was any biological activity

contributing to the removal, the non-sterile GAC column should show higher removal

than the sterile GAC column. However, considerably lower removal rates were recorded

for the non-sterile GAC (highlighted in Figure 3.3). Approximately 30% less of m-LR

and 40% less of m-LA was adsorbed by the non-sterile GAC during the adsorption

breakthrough period between days 66 to 100. Reasons for the different adsorption

efficiency were unclear, although it was suspected that the biofilm that grew on the

surface of the non-sterile GAC would have restricted the external pores thus decrease

the diffusion efficiency or alternatively the weekly autoclaving treatment for the sterile

GAC could affect the inherent adsorption capacity for microcystins. More details on

these aspects are given in Section 3.5 and Chapter 5.

Compared with the microcystin removal in the non-sterile GAC column (Figure 3.2),

the removal in the sterile GAC was highly variable. One reason for the variability could

be the weekly autoclaving treatment where the GAC was sterilised. The microcystin

adsorption capacity of the GAC was assumed unaffected by autoclaving at 121 ºC, 500

kPa for 20 minutes as the carbon was activated at temperatures greater than 1000 ºC.

Page 54: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

40

However, outcomes from a trial, during which microcystin adsorption in the sterile

GAC was measured daily between days 192 to 202 (Figure 3.4), indicate that adsorption

was higher in GAC shortly after autoclaving on days 192 and 200 (underlined points in

Figure 3.4). It was, therefore, proposed that the weekly autoclaving could be

contributing to the fluctuation of toxin removal in the sterile GAC column.

0

1

2

3

4

5

6

7

8

66 91 116 141 166 191 216

Time (Days)

m-L

R co

ncen

tratio

n (μ

g/L) m-LR

m-LA

Figure 3.4 Impact of autoclaving on the effluent concentration of both m-LR (upper) and m-LA (lower) from the sterile GAC column – enhanced toxin adsorption after autoclaving

Reasons for the temporarily enhanced adsorption in the autoclaved sterile GAC column

could be attributed to the partial desorption of pre-adsorbed substances (described in

Section 5.3.4). Desorption from some of the pre-occupied adsorption sites could

improve the adsorption capacity for microcystins. Importantly, it was shown that the

amount of organic material released due to weekly autoclaving was not significant (<5%)

relative to the amount that was adsorbed (Section 5.3.4). Hence, the interference caused

by the autoclave treatment was a temporary phenomenon, and most of the data obtained

for this study truly represented the adsorption trend of the sterile GAC column.

3.2.3 The sand column Microcystin (m-LR) removal in the sand column is presented as toxin concentration in

the influent and effluent water for this column (Figure 3.5). Microcystins were not

eliminated in the sand column prior to day 211 as similar levels of the toxin were

evident in the column inflow and outflow. However, a sharp decrease in the outlet toxin

Page 55: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

41

concentration was observed from days 211 to 215 (highlighted in Figure 3.5), which

indicates that biodegradation of microcystins was initiated during that time. The

initiation of biodegradation in the sand column shows similarity with the degradation in

the non-sterile GAC column at about day 100, which suggests that day 100 was the start

of biodegradation in the non-sterile GAC, and the removal mechanism before this time

was primarily adsorption.

Figure 3.5 Removal of m-LR in the sand column (the dotted rectangular indicates the sharp

decrease in the toxin outlet concentration when biodegradation commenced in the sand column)

The initiation time prior to biodegradation of microcystins was significantly different in

the non-sterile GAC (approximately 3 months, Figure 3.1 or 3.2) and the sand column

(greater than 7 months, Figure 3.5). This may be attributed to different biological

growth patterns on the two substrates as indicated by SEM images (Figure 3.6). On the

surface of the sand (image A), only sparse biological attachment and growth was

evident after the column had been operated for approximately 3 months. In contrast, on

the surface of the GAC (image B) where microcystin biodegradation had just been

initiated at the time of imaging, a flourishing microbial community was well developed.

After approximately 8 months (image C and D), the biofilm on both surfaces had

matured. The sand surface (image C) still had much less biomass than the surface of

GAC (image D), but it was comparable to the 3-months old GAC biofilm.

Page 56: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

42

Figure 3.6 Scanning electron microscopic images showing the biofilm that grew on the surface of sand (left) and the GAC (right) after approximately 3 months (top) and 8 months (bottom)

Biofilm growth (shown in Figure 3.6) occurred much faster on the virgin GAC particles

than on the new sand. Reasons for the different biofilm growth rates were not obvious

although it could be attributed to the different surface characteristics of the two

substrate materials as shown in the SEM images in Figure 3.7. The surface of the sand

(left) is rather smooth and nonporous, while the GAC (right) shows a much rougher

surface with random widely distributed crevasses and ridges which provides additional

surface area. Such structures of GAC could help protect newly attached bacteria from

shear forces which could have been a major hindrance for biofilm development.

A. Sand - 3months B. GAC - 3 months

C. Sand - 8 months D. GAC - 8 months

Page 57: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

43

Figure 3.7 Scanning electron microscopic images of the different surface characteristics of the

sand (left) and the GAC A6 (right) particles

In addition to favorable surface characteristics of GAC for bacteria attachment, it is also

possible that the external macropores or crevasses of the GAC, which lead to internal

macropores, could result in further bacterial growth by providing inner surface area. To

examine this possibility, SEM was used to view GAC granules that had been sliced in

half to uncover the internal macropore surfaces of the GAC (Figure 3.8). It was evident

that a negligible amount of bacteria grew in these pores as no biofilm was evident.

Therefore, it is proposed that the biofilm only covered the outer surface of the GAC

particles and no considerable extra area was provided by the open macropores or

crevasses.

Figure 3.8 Scanning electron microscopic images showing the bacterial growth in the internal pores of the GAC

Sand GAC

Page 58: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

44

3.3 Results from the repeat experiment

Microcystin removal rates prior to day 66 were obtained from the repeat experiment that

was conducted shortly after the initial study (Section 3.1). The employed apparatus and

operational parameters were kept the same for the two experiments, except the sand

column that was not re-commissioned as it was presumed that no biodegradation of

microcystins occurred during the initial period of time because no removal was

observed for 7 months (Figure 3.5).

Toxin concentration for the inflow and outflow from the columns was plotted against

operation time (Figure 3.9). Toxin m-LR was completely adsorbed in both of the two

GAC columns in the initial 9 days. Slight breakthrough was evident from the non-sterile

GAC after day 10, but similar breakthrough was not observed in the sterile GAC

column until day 35. From day 10, m-LR breakthrough from the non-sterile GAC

column gradually developed and then stabilised at about 10% from day 18. After day 34

this decreased and after a further 4 days, no additional breakthrough was recorded.

Therefore, biodegradation of microcystins was proposed to have started from day 34 in

the non-sterile GAC column. In the sterile GAC column toxin breakthrough started

from day 35 and then gradually increased.

Page 59: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

45

Figure 3.9 m-LR and m–LA concentration for the inflow and outflow from the columns in the

repeated set of experiment

Similar toxin removal behaviour was observed in the two experiments with respect to

the difference in microcystin adsorption efficiency between the sterile and the

non-sterile GAC column. Toxins were shown to be more readily adsorbed by the sterile

GAC (no biofilm growth) than the non-sterile GAC (with biofilm growth) in both of the

experiments. Further discussion regarding biofilm effects is given in Section 3.5.1.

Disagreement between the two studies was also evident. The length of the lag-phase

prior to biodegradation of microcystins in the non-sterile GAC column was much longer

in the first experiment (approximately 100 days) compared with the repeated work (only

34 days). This could be due to pre-filtration of the water source in the initial two months

of the first study, which was not performed in the repeat work. A 0.2/0.8 µm capsule

filter (Pall Life Sciences, USA) was used for pre-filtration of the water sample prior to

use. This procedure had been proposed for removal of some of the microorganisms in

the water, which may have caused pre-degradation of the toxins in the water reservoir

before it was pumped to the columns as documented by a previous study also using

Myponga water (Ho, 2004).

The pre-filtration through the 0.2/0.8 μm pore size filter potentially removed all bacteria

that were responsible for microcystin metabolism. SEM images (Figure 3.10) were

taken from the primary bacterial colonies that were extracted from the GAC column

where biodegradation of microcystins was present. Five different types of bacteria

Page 60: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

46

(unidentified) were dominant in respect to their morphology. It was evident that only

two types (A and D) with short rod shapes were small enough to pass through the filter

and be present in the filtered water, while the rest were relatively larger (B, C and E)

and were likely to be trapped in the filter. Hence, there might be significantly less

bacteria in the feed water that was supplied to the non-sterile GAC column in the initial

two months of the first study, so this was probably the reason for the longer lag-phase

before biodegradation occurred in the original non-sterile GAC column.

Figure 3.10 Scanning electron microscopic images showing the size of main bacteria species that were isolated from the non-sterile GAC column

A B

C D

E

Page 61: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

47

3.4 Combined results The dimensionless concentration of m-LR and m-LA in the three columns containing

the sterile, the non-sterile GAC column and the sand column was determined (Figure

3.11) based on the first (Section 3.1) and the repeated (Section 3.2) experiments. Toxin

breakthrough in the two experiments was well integrated with each other except for the

different lag-phase periods prior to commencement of biodegradation in the non-sterile

GAC column. It was considered acceptable to obtain different periods of initiation time

as the GAC columns were studied at different times, and most importantly, the different

lag-phase period does not affect the resultant toxin removal trend in the column.

Therefore, the lag-phase for the repeated experiment is presented for discussion of the

final results.

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

0 25 50 75 100 125 150 175 200 225

Time (days)

Dim

ensi

onle

ss c

once

ntra

tion

of m

-LR,

C/

Co

Non-sterile GAC

Sterile GAC

Sand

Page 62: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

48

0.00

0.20

0.40

0.60

0.80

1.00

1.20

1.40

0 25 50 75 100 125 150 175 200 225

Time (days)

Dim

issi

onle

ss c

once

ntra

tion

of m

-LA

, C

/Co

Non-sterile GAC

Sterile GAC

Sand

Figure 3.11 Final results of m-LR (top) and m-LA (bottom) breakthrough profiles in the three

columns

Each of the three columns (Figure 3.11) represented different microcystin removal

mechanisms, i.e. adsorption (the sterile GAC column), biodegradation (the sand column)

and the combination of the mechanisms (the non-sterile GAC column). Thus it is

proposed that the partitioning of the relative importance of the two microcystin removal

mechanisms in GAC filtration was successfully separated. The separated data provides

essential information for further investigating and facilitating microcystin removal using

GAC filtration. A mathematical model, which is able to predict the toxin removal

efficiency in GAC filtration, could be developed by separately modelling the two

removal mechanisms (adsorption and biodegradation) and then combining the results to

calculate the overall efficiency.

3.5 Phases of microcystin removal in GAC filters Microcystin breakthrough observed in the columns in this study is summarised in

Figure 3.12 (A) based on the combined experimental results discussed previously and

presented in Figure 3.11. The whole removal process in GAC filtration could be divided

into two phases, i.e. Phase I (principally adsorption) and Phase II (principally

biodegradation). A corresponding conceptual model describing the toxin elimination

mechanisms in the two phases is also presented in Figure 3.12 (B).

Page 63: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

49

Figure 3.12 A) Normalised percent breakthrough curves of microcystins in the sterile GAC, the non-sterile GAC and the sand column in the column experiment. Microcystin removal in

GAC filtration is divided into Phase I (adsorption) and Phase II (biodegradation);

B) Conceptual model describing microcystin removal mechanisms by GAC before and after biofilm development. The corresponding removal rate by adsorption and biodegradation

calculated from this study is also given.

Dim

ensi

onle

ss c

once

ntra

tion,

C/C

o

1.0

0 Phase I Phase II

Operation Time

Non-sterile GAC Sterile GAC Sand

Adsorption Biodegradation

Before (GAC) After (GAC + Biofilm)

Adsorption Biodegradation

Rate: ≥ 1.25 μg/m2/d (m-LR)

Rate: ≥ 1.06 μg/m2/d (m-LR)

Adsorption

Unknown rate

Kf Kf’ = Kf +Kf,b (biofilm effect)

A)

B)

Page 64: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

50

3.5.1 Phase I - Adsorption removal and biofilm effect Adsorption is the primary mechanism for microcystin removal in a newly established

GAC filter (indicated as Phase I in Figure 3.12) as the initiation of toxin biodegradation

requires a period of time for development of biofilm in the GAC bed. For the present

study, the highest microcystin (m-LR) adsorption rate in this phase was no less than

1.25 μg/m2/d (Appendix B).

The adsorption process is essentially the mass transfer process of toxin molecules from

a liquid phase into the GAC porous structure followed by adsorption on the surface. The

first transfer operation is usually described by mass transfer coefficient parameter Kf.

Analytical models such as the plug flow Homogeneous Surface Diffusion Model

(HSDM), is available from the scientific literature. This model is used widely to predict

adsorption kinetics of organic compounds by activated carbon (Rosen, 1952; Weber and

Liu, 1980; Yuasa, 1982; Sontheimer et al., 1988; Smith et al., 1987; Traegner et al.,

1989; Roy et al., 1993 and Schmidt, 1994). The adsorption equilibria of HSDM can be

described by the Freundlich isotherm equation (Freundlich, 1906):

q = Kf C1/n (Equ. 6)

Where q is the equilibrium solid phase concentration (mg, adsorbate/g, adsorbent), Kf =

Freundlich isotherm constant (cm/s), C = equilibrium liquid phase concentration (μg/L),

and 1/n = Freundlich isotherm constant. The application of HSDM for modelling GAC

adsorption of trace organic contaminates including microcystins from natural water has

been previously validated (Ho, 2004). In the present study, the adsorption behaviour in

the sterile and the non-sterile GAC columns was modelled using the HSDM (Figure

3.13).

Page 65: Adsorption and biological filtration of microcystins

Chanter 3 T le Cnlrmn F.xnerimenf: Mi R cmnr¡cl

1 + HSDM Fit at Kf =3.0E-3 (sterile GAC)

+ HSDM Fit at Kf =1 .2E-3 (non-sterile GAC)

+ Eperimental data (sterile GAC)

+ Experimental data (non-sterile GAC)

oooE'fo

-v.EollÉ.

I

E

0.8

0.6

0.4

0.2

0

0 50 100

Time (days)150 200

tr'igure 3.13 HSDM fit of the sterile and non-sterile GAC experimental data at different mass

transfer coeffrcient KT values

The isotherm parameters (Ç K and I/n) applied in the HSDM model runs were

obtained by minimizingthe error between the experimenta| data and the curve fit. The

values of these parameters (presented in Table 3.1) are reasonable when compared with

those for other organic compounds cited in the literature, for example,

2-methylisoborneol (MIB) and m-LA adsorption using virgin P1100 and Picazine

carbon in Ho's (2004) study (shown in Table 3.1).

Table 3.1 HSDM input parameters (K¡ K & 1/n) describing m-LR adsorption onto the sterile

and non-sterile GAC in the present study, and MIB and m-LA adsorption parameters cited inHo's (2004) study

Present study Ho's (2004) studyParameters

Sterile GAC Non-sterile GAG MIB M.LA

K¡(cm/s) 3.0 x 10-3 l.2xlo-3 4'90x10-a 4'36x10-a

I( (nglmg)(n7lL¡-tt" 6.202 6.202 t.257 0.016

l/n 0.805

5l

0.805 0.2t5 1.t44

Page 66: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

52

For the best fit of the non-sterile GAC adsorption data, the Kf value needed to be lower

than the value for the sterile GAC (Table 3.1). As mentioned previously (Section 3.2

and 3.3), this could be due to the biofilm that grew on the GAC surface causing

reduction in surface diffusion of microcystins (Kf, b). Therefore, the toxin transfer

efficient onto the GAC after development of biofilm (Kf’) is given by considering the

interference of biofilm (conceptual model in Figure 3.12 B). Observation of the affected

mass transfer rate by biological accumulation has also been documented by many other

authors (Lowry and Burkhead, 1980; Schultz and Keinath, 1984; Bishop et al., 1995;

Zhao et al., 1999; Lahav and green, 2001 and Song et al., 2006). For example, in

Olmstead’s (1989) study the mass transfer coefficient (Kf) of trichloroethylene was

found to decrease with increasing biomass growth on the GAC, but the Freundlich

parameter (1/n), an indicator of the adsorption capacity of the adsorbent/adsorbate

system, did not change significantly. 3.5.2 Phase II - Biodegradation removal and lag-phase The transition from Phase I adsorption to Phase II biodegradation is very quick via a

sharp decrease in the toxin breakthrough (Figure 3.12 A). This rapid transition from

adsorption to biodegradation implies that the highly efficient biological removal of

microcystins commenced in the non-sterile GAC column immediately after the bacteria

acclimatised to the toxins. Therefore, biodegradation, as opposed to adsorption, became

the primary mechanism for microcystin elimination in the biologically activated GAC

column. Additionally, this finding also indicates that microcystins are highly

biodegradable given appropriate operating conditions. In a previous GAC study

conducted by Ho (2004) similar phenomenon of a sudden increase in microcystin

removal was detected. The author attributed this to biological activities in the GAC

filter, because significant breakthrough of the toxins reappeared in the filter outlet after

the biofilm on the GAC was inactivated by sterilisation.

During Phase II, microcystin removal would be largely enhanced due to biodegradation.

According to this study, the m-LR removal rate was shown increased from 380 to no

less than 1.06 μg/m2/d after biodegradation commenced in the non-sterile GAC column.

Page 67: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

53

A test (Section 2.1.6) was undertaken to quantify the bacterial concentration in the

non-sterile GAC and the parallel sand column, both of which were then in

biodegradation phase. Results (60 x 107 cell / cm3 in the GAC and 1.37 x 107 cell / cm3

in the sand) from this study showed that the biofilm that grew in the GAC and the sand

column had very similar bacterial abundance. It thus indicates that the biofilm in the

non-sterile GAC column was able to completely remove the microcystins under the

employed operational conditions. Additionally, biodegradation occurs before adsorption

as biomass is located on the surface of the carbon (Figure 3.12 B), and mass transfer

through this biological layer would occur before adsorptive diffusion into the GAC solid

occurs. Therefore, the removal of microcystins in Phase II was mainly due to

biodegradation rather than adsorption (Figure 3.12 B).

Although biological GAC (BGAC) filtration seems to be a highly efficient option for

microcystin treatment, the prerequisite lag-phase (Section 3.2.1) which may take days or

months (depending upon the operational conditions) is a major obstacle for the practical

application of this technology. Similar lag periods of varying length have been widely

reported from a large spectrum of biological studies. It has been found to be very difficult

to predict the lag-phase, as its occurrence and length relates to many factors, such as

previous exposure to the substances (Mort and Deanross, 1994; Mirgain et al., 1995;

Gotvajn and Zagorckoncan, 1996; Mort et al., 1999; Guieysse et al., 2001); concentration

of the substances (Gotvajn and Zagorckoncan, 1996; Thomas et al., 2002); existence of

enzyme inducers (Yassir et al., 1998); reactions between genes involved in degradation

(Churchill et al., 1999); the volume of test medium (Ingerslev et al., 2000); and the

amount and concentration of specific degrading microorganisms (Ingerslev et al., 2000;

Preuss et al., 2002; Bending et al., 2001).

The lag-phase period prior to microcystin biodegradation is hard to predict. In Ho’s

(2004) GAC column study, the lag period for microcystin degradation decreased and in

one instance disappeared upon intermittent re-dosing of microcystin. Observations from

the batch tests conducted in the present study (Chapter 6) indicate that the length of delay

could also be affected by a diverse range of operation conditions, such as the temperature,

the quality of water, the concentration of bacterial cells and the aeration level in the

Page 68: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

54

solution. Since microcystin biodegradation is extremely important for GAC filtration

processes, further investigation to minimise the lag-phase is necessary.

3.6 Conclusions

The results from this long-term column experiment demonstrate that the relative

contribution of adsorption and biodegradation for the removal of microcystins in GAC

filtration have been successfully separated. Relevant information on the toxin

adsorption, biodegradation and the combination of the two mechanisms is represented

by the results of the sterile GAC column, the sand column and the non-sterile GAC

column, respectively (Figure 3.11). The whole microcystin removal process in GAC

filtration could be divided into two phases, i.e. Phase I (mainly adsorption) and Phase II

(mainly biodegradation).

The GAC adsorption capacity for microcystin would be gradually depleted by natural

organic matters in the source water and the toxins themselves. Generally, adsorption of

the two analogues of microcystin (m-LR and m-LA) both follows a typical

breakthrough curve, but they are of different adsorbability. Higher adsorbability was

evident for m-LR in this study (Figure 3.3). The proposed reason for this difference was

that the adsorbed m-LA molecules (-2 charge) which have higher negative charge than

m-LR (-1 charge), may present stronger electrostatic repulsion and thus inhibit

adsorption due to hydrophobic interaction with the surface (Newcombe, 2003).

The behavior of microcystin adsorption in the sterile and the non-sterile GAC column

could be modelled using the HSDM (Figure 3.13). The value of the mass transfer

coefficient, Kf, was found to be lower for the non-sterile GAC than it for the sterile

GAC as the adsorption of microcystins was higher in the sterile GAC compared with

the non-sterile GAC (Figure 3.3). This is proposed to be attributable to the biofilm that

blocked the external pores of the GAC thus decreasing the surface diffusion of

microcystins into the internal adsorption sites (Figure 3.12 B).

Transition from Phase I adsorption to Phase II biodegradation is very quick (Figure 3.12

Page 69: Adsorption and biological filtration of microcystins

Chapter 3 Laboratory-Scale Column Experiment: Microcystin Removal

55

A) and therefore difficult to predict. In Phase II, microcystin removal would be

enhanced due to the initiation of biodegradation, which also becomes the primary

removal mechanism rather than adsorption as in Phase I. The high removal efficiency in

the biodegradation phase is preferred but it usually occurs after an unpredictable

lag-phase period. For confident practical application of this technology, further

investigation aiming to eliminate this delay period should be a priority.

Finally, biodegradation of the toxins occurred much earlier in the GAC column than in

the parallel sand column. The possibility that GAC supplied extra surface area that was

provided by the open macropores or crevasses in GAC was excluded as the biofilm was

found to exist mainly on the outer surface of the GAC particles (Figure 3.8). The reason

for the difference in rate of biodegradation between the columns is attributed to the

distinctive surface characteristics of the two materials (Figure 3.7), where GAC appears

to be a better substrate for biofilm attachment and development than the sand in this

study (Section 3.2.3).

Page 70: Adsorption and biological filtration of microcystins

56

CHAPTER 4

REMOVAL OF NATURAL ORGANIC MATTER IN GAC AND SAND COLUMNS

4.1 Introduction

Although the major aim of this project was to identify the relative contributions of

adsorption and biodegradation for the removal of microcystins in GAC filtration, an

additional aspect for the application of these processes is the simultaneous removal of

Natural Organic Material (NOM). NOM in drinking water is usually a complex mixture

of organic compounds that are formed from natural decomposition of animal and plant

materials; industry, agriculture and domestic effluents as well as urban and rural rainfall

runoff. The composition of NOM typically ranges from small hydrophilic acids,

proteins and amino acids to relatively larger humic and fulvic acids (apparent molecular

weight usually between approximately 500 and 30,000) which is the predominant

constituent of the comparatively high NOM level waters in Australia (Choudry, 1984).

Health problems associated with NOM have been emphasised since the discovery of its

disinfection by-products (DBP), which have been recognized as potential threats for the

public wellbeing since the 1950s (Middleton and Rosen, 1956). For example, there is

particular concern with chlorine DBPs, such as trihalomethanes (THMs), some of which

are carcinogenic and may cause series of ailments varying in character and intensity

(Sontheimer, 1988). High levels of NOM can also impact human health by promoting

intensive biological regrowth in drinking water distribution systems (Yavich et al.,

2004). Therefore, water treatment processes that are able to decrease the NOM

Page 71: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

57

concentration to a reasonable extent are highly desired.

Adsorption, which concentrates organic substances on the inner surfaces of solid

adsorbents, is the most effective process for removal of NOM in water treatment plants

(Sontheimer, 1988). Activated carbon has large internal surface areas (1,000,000 m2/kg)

that are non-polar, and is the most common adsorbent used by water authorities

(Orshansky and Narkis, 1997). Activated carbon can remove a wide range of organic

compounds from water including some health-concerning chemicals such as

trihalomethanes, pesticides, industrial solvents, polychlorinated biphenyls, and

polycyclic aromatic hydrocarbons.

GAC filtration, which has been extensively used for water treatment processes since the

1930’s offers an effective option for NOM removal due to its high adsorption capacity

and rapid adsorption kinetics (Guerine and Boyd, 1997; Rangel-Mendez and Streat,

2002; Song et al., 2006). To optimise the design and the practical operation of this

application, extensive studies concerning characterisation of NOM adsorption have

been carried out. However, different water sources have different NOM constituents and

thus exhibit different adsorbability during GAC filtration (Aiken and Cotsaris, 1985;

Najm et al., 1990; Summers, 1989). Consequently, it is difficult to predict the

adsorption effect of NOM in water treatment plants. To address this problem, many

on-site studies have been conducted for specific water environments (Vahala et al., 1999;

Wietlik et al., 2002), but research has been restricted in Australian situations, which

usually exhibit high NOM concentrations due to the unique climate and soil conditions.

Application of GAC filters in water treatment systems can also stimulate biodegradation

of various target NOM compounds (Mollah and Robinson, 1996; Xu et al., 1997; Jones

et al., 1998 and Jung et al., 2001) given that an appropriate biofilm is developed on the

GAC grains. This biological process is considered important as it can potentially

accelerate the NOM removal and prolong the life-time of GAC filters compared to

Page 72: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

58

solely adsorptive GAC units (Jonge et al., 1996). However, there is no literature

available relating to the separation of adsorption and biodegradation for removal of

NOM in long-term GAC filters.

The objectives of the research reported in this chapter were to characterise the NOM

removal profiles in GAC filtration based on Australian reservoir water, and to separate

the adsorption and the biological reduction of these organic compounds.

4.2 Methods

The experimental design as presented previously for the microcystin study (Chapter 2

and Chapter 3), which consisted of three columns, i.e. the sterile and the non-sterile

GAC and the sand column (Section 2.2), was also used for the NOM investigation.

Separated adsorption (the sterile GAC column) and biodegradation removal (the sand

column) and the combination of the two mechanisms (the non-sterile GAC) are

represented by the three columns. The NOM breakthrough profile was chosen as the

measurement criteria for the performance of GAC columns in this study. Influent and

effluent samples of the three columns were regularly sampled for NOM analysis. The

analytical instruments used to obtain experimental data were: a dissolved organic

carbon (DOC) analyser, a UV spectrophotometry and a High Performance Size

Exclusion Chromatography (HPSEC) unit.

DOC is the most commonly applied parameter used to quantify NOM. It represents the

fraction of total organic carbon (TOC) that passes through a 0.45 μm pore size filter,

and is a heterogeneous mixture of various complex organic materials ranging from

macromolecular humic substances to small molecular weight hydrophilic acids and

various hydrocarbons (Thurman, 1985). Breakthrough of substrates that are precursors

of the carcinogenic THMs in GAC units is often similar to that of the DOC (Sontheimer

et al., 1988). Samples were pre-filtered through 0.45 μm cellulose nitrate membrane

filters (Schleicher and Schuell, Germany) and measured using an 820 Total Organic

Page 73: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

59

Carbon Analyser (Sievers Instruments Inc, USA). The determined DOC value from the

analytical instrument is based on the oxidation of the organic matters to CO2, which is

measured and converted to the amount of DOC.

UV absorbance, as a surrogate for DOC measurement, is usually applied as it is quick,

straightforward and uses only spectrophotometric instrumentation. Many organic

compounds in raw water sources possess the property of absorbing light in the

ultraviolet region (10-400 nm) and the visible region (400-800 nm) (Sontheimer, 1988).

Absorption at 254 nm (UVA254) correlates well with DOC although it tends to reflect

only the more complex NOM structures especially the aromatic and unsaturated carbon

bonds (–C=C–) (Traina et al., 1990), including humic substances, aromatic compounds,

tannins and lignins. Pre-filtered (0.45 μm) water samples were measured using a

UV/VIS 918 Spectrophotometer (GBC Scientific Equipment Pty Ltd, Australia) at 254

nm in 1 cm quartz cells using Milli-Q water as a reference.

HPSEC is a reliable and informative technique. The use of HPSEC allows

characterisation of UV absorbing NOM constituents according to their molecular

weight, so it is useful in determining what range of size molecules are removed from the

columns during this study. HPSEC analysis was conducted to assist with interpretation

of the obtained DOC and UV254 data. HPSEC samples were filtered through 0.22 μm

cellulose nitrate membrane filters (Schleicher and Schuell, Germany) and analysed

using a Waters 2690 Alliance System (Section 2.3.3).

4.3 Results and Discussion

The removal profile of NOM in the three columns (the sterile and non-sterile GAC, and

the sand column) is discussed below.

Page 74: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

60

4.3.1 Non-sterile GAC column

0

1

2

3

4

5

6

7

8

0 25 50 75 100 125 150 175 200 225 250 275 300 325

Time (days)

DO

C (m

g/L)

non-sterile GAC influent non-sterile GAC effluent

Figure 4.1 DOC concentration of the influent and effluent of the non-sterile GAC column (the

dotted circle indicates the initial NOM breakthrough)

0

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0 50 100 150 200 250 300

Time (days)

UV

abso

rptio

n @

254

nm

GAC/Sand Inf GAC eff Sand eff

Sterile GAC Inf Sterile GAC eff

Figure 4.2 UV254 absorbance of the influent and effluent of the three columns containing the sterile and non-sterile GAC and the sand column (the dotted circle indicates the initial NOM

breakthrough)

NOM removal in the non-sterile GAC column for the duration of this study was

determined by analysis of DOC and UVA254 in the influent and effluent of the column

(Figure 4.1 and 4.2). Generally, the removal of NOM in this column followed a typical

breakthrough curve (Sontheimer et al., 1988). During the initial period, approximately 1

week, a low concentration of NOM appeared in the effluent (as indicated by the dotted

line in Figure 4.1 and 4.2), which could be attributed to the less adsorbable or

non-adsorbable proportion of NOM in the water. HPSEC analysis was carried out to

Page 75: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

61

determine the molecular weight distribution of UV absorbing NOM in the influent and

effluent of the non-sterile GAC column (at day 5) (Figure 4.3). It is evident that almost

all the UV absorbing organic matter was removed in the initial period of operation of

the GAC column except a small proportion of NOM that may be of extremely low or no

adsorbability (highlighted in Figure 4.3) (Roberts and Summers, 1982; Roberts et al.,

1984 and Sontheimer et al., 1988). This immediate breakthrough was found to be

strongly related to the concentration of the non-adsorbable NOM in the water being

treated and the EBCT condition used in the operation (Sontheimer et al., 1988).

-0.002

0.000

0.002

0.004

0.006

0.008

0.010

100 1000 10000 100000

Apparent Molecular Weight

UV

Abs

@ 2

60 n

m

The influent

GAC effluent (day 5)

Figure 4.3 HPSEC chromatograph showing NOM adsorption in new GAC (day 5)

After the initial period (approximately 7 days), NOM breakthrough in the effluent of the

non-sterile GAC column immediately increased at a high rate as determined by the

DOC concentration (Figure 4.1) and UVA254 (Figure 4.2). This part of the breakthrough

profile corresponded to the rapid decrease in the available adsorption sites as the

adsorption zone reached the end of the column, and the extent of this rapid

breakthrough largely depended on the EBCT of the filter (Hyde, 1980) and the

adsorption capacity of the GAC for strongly adsorbing compounds (Sontheimer et al.,

Page 76: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

62

1988). Regarding the DOC concentration in this study, NOM adsorption continued until

day 60 when the breakthrough rate reached up to 80%, while after approximately 5

months (150 days), only a low removal (~10%) was evident (Figure 4.1). The reason for

the final minor removal was attributed to the mechanism of biodegradation in the filters.

In practise, operation of GAC units is often terminated before the low adsorption phase

is reached (Sontheimer et al., 1988). Conclusively, the breakthrough profile observed

herein was in good agreement with the literature (Roberts et al., 1984) in relation to

DOC removal in drinking water plants.

4.3.2 Sterile GAC column

Removal of NOM in the sterile GAC column was also determined in terms of UVA254

(Figure 4.2) and DOC concentration (Figure 4.4), and the resultant NOM breakthrough

curve was compared with that of the non-sterile GAC column. It was observed that the

removal of NOM in the sterile column appeared highly variable. A similar fluctuation

was evident in the microcystin concentration (Chapter 3), and it was attributed to the

weekly autoclaving of the sterile GAC column, which was proposed as a possible cause

of the observed variability (Section 3.3.2). Further studies on this aspect are presented

in Chapter 5. Ignoring the fluctuation, DOC removal in the sterile column (Figure 4.4)

exhibited very similar behaviour with the non-sterile GAC column throughout the

experiment, details of which were presented in Section 4.3.1.

0

1

2

3

4

5

6

7

8

0 25 50 75 100 125 150 175 200 225

Time (days)

DO

C (m

g/L)

Sterile GAC influent Sterile GAC effluent

non-sterile GAC effluent

Figure 4.4 DOC concentration of the influent and effluent of the sterile GAC column

Page 77: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

63

After operation of 3 to 6 months there was still up to 20% adsorption of DOC occurring

in the sterile GAC column (Figure 4.4). Part removal of the UV absorbing NOM

constituents was also observed with the HPSEC analysis (Figure 4.5) which was

conducted on both of the influent and effluent water samples of the sterile GAC column

after 5.5 months operation. It was evident that some adsorption capacity remained in the

5.5 months preloaded GAC especially for the NOM consisting of low range of

molecular weight. From this evidence it is proposed that the low percentage (~10%)

removal of NOM that is usually shown in GAC filters after several months’ operation

can contribute in part to residual adsorption capacity rather than biodegradation as

usually assumed (Sontheimer et al., 1988).

-0.002

0.000

0.002

0.004

0.006

0.008

0.010

0.012

100 1000 10000 100000

Apparent Molecular Weight

UV

Abs

@ 2

60 n

m

Sterile GAC influent (day 165)

Sterile GAC effluent (day 165)

Figure 4.5 HPSE chromatograph showing NOM adsorption in the preloaded sterile GAC (5.5

months)

Adsorption of the UVA254 characterised NOM in the sterile GAC column (Figure 4.2)

was higher in comparison with the DOC characterised NOM (Figure 4.4) as there was

still approximately 30% UVA254 removal after 6 months when the adsorption capacity

for DOC was shown to be almost saturated. Similar preferential adsorption of UV

Page 78: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

64

absorbing organic matters was reported by other researchers (Sontheimer et al., 1988;

Åsa et al., 1997; Li et al., 2003). Reasons for this phenomenon are unknown although it

could be attributed to the relatively small MW (<1000 Daltons) of the UV absorbing

NOM in the Myponga water (Figure 4.3 and 4.5), because smaller molecules are more

readily adsorbable than large MW compounds. Åsa et al. (1997) reported the same

conclusion when interpreting a similar adsorption preference that was observed in a

pilot-plant GAC study. Another possible reason for the observed higher adsorption rate

of the UV absorbing NOM could be due to the different analytical methods used. As

mentioned previously (Section 4.2), the DOC analyser measures organic compounds

that can be oxidised to CO2 during analysis, the amount of which is consequently

correlated to the amount of DOC. Since it can not be ensured that all the carbon

constituents in the water sample are oxidisable, this approach could be less sensitive

than the analysis by UV absorbance. Therefore, the significantly decreased removal

efficiency of NOM at the end of the adsorption was difficult to be concluded from the

DOC determination (Figure 4.4) compared with the more sensitive UV absorbance

analysis (Figure 4.2).

4.3.3 Sand column

Removal of NOM in the sand column was determined as UVA254 (Figure 4.2) and DOC

concentration (Figure 4.6) in the influent and effluent of the column. Generally, there

was no observable removal of NOM in this column throughout its operation for

approximately 11 months (Figure 4.2 and 4.6). It is, therefore, concluded that the

adsorption of NOM was absent in the sand column.

Page 79: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

65

0

1

2

3

4

5

6

7

8

0 25 50 75 100 125 150 175 200 225 250 275 300 325

Time (days)

DO

C (m

g/L)

Sand influent Sand effluent

Figure 4.6 DOC characterised NOM measured on the influent and effluent of the sand column

Observable biological removal of NOM was also not evident in the sand column. A

HPSEC analysis (Figure 4.7) conducted on water samples prior to and after the sand

column that was operated for approximately 10 months, showed very similar MW

distribution of the organic matters in the two samples. Reasons for the absence of

biodegradation of NOM in the sand column could be attributed to the EBCT (15

minutes, to be consistent with the parallel GAC column) employed in this study, which

is low compared to the EBCT of 30–60 minutes that is usually used in a biological sand

column.

-0.001

0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

100 1000 10000

Apparent Molecular Weight

UV

Abs

@ 2

60 n

m

The influent

Sand effluent (10 months)

Figure 4.7 HPSEC chromatograph showing NOM removal in the sand column (~300 days)

Page 80: Adsorption and biological filtration of microcystins

Chapter 4 Removal of Natural Organic Matter in GAC and Sand Columns

66

4.4 Conclusions

The removal behaviour of NOM in GAC (A6) filtration was studied using Australian

reservoir water. Because of the relatively high NOM concentration in the water, some

specific features were observed from the DOC breakthrough curve. First, the adsorption

efficiency of the GAC decreased rapidly after the initial breakthrough, and the removal

capacity was shown to be almost saturated after a very short time (about two months).

Moreover, the preferred high rate, steady-state adsorption phase (a time span of nearly

constant effluent concentration) that usually occurs after the initial breakthrough was

not observed in this case. All these features could be due to the high load of the organic

compounds in the water, which exhausted the adsorption sites rapidly.

Low percentage (~10%) removal of NOM that is usually observed in commercial GAC

filters after several months operation was also observed in the experimental GAC

columns during this study. Since any removal of DOC in the sterile GAC column was

due to adsorption, the small reduction rate during this time period could be contributed

in part to the residual adsorption capacity of the GAC rather than biodegradation, as is

usually assumed (Sontheimer et al., 1988).

Finally, the favoured adsorption of UV absorbing NOM compounds compared to the

DOC constituents during GAC filtration, which was previously reported by other

investigators, was also observed in the sterile GAC column in this study. Two probable

reasons could explain this phenomenon. One is the relatively small MW (MW<1000

Daltons) of the UV absorbing NOM in the water, which could be more easily adsorbed

on to GAC. Secondly, it was assumed that the DOC analyser used in this study was less

sensitive than the UV analyser due to the possible non-oxidising nature of some NOM

compounds.

Page 81: Adsorption and biological filtration of microcystins

67

CHAPTER 5

EVALUATING THE VALIDITY OF THE EXPERIMENTAL DESIGN

5.1 Introduction

To quantify the relative importance of the toxin removal by adsorption and

biodegradation in a GAC filter, three columns were commissioned: the sterile and the

non-sterile GAC column and the sand column. The toxin removal properties of each

column were adsorption only for the sterile GAC, adsorption and biodegradation for the

non-sterile GAC and biodegradation only for the sand column. To enable the

comparison between the columns, it was essential to ensure that the operation

conditions for the columns were consistent throughout the experiment. However, the

sterilisation treatment (autoclaving) that was conducted prior to application of the sterile

(GAC) column may induce a potential difference in operational conditions between the

two GAC columns.

The sterilisation treatment, described in Chapter 2, was achieved by autoclaving (121

°C, 500 kPa for 20 minutes). It is a high temperature and pressure process during which

various physical or chemical reactions could be promoted in the water and GAC sample,

or even in the apparatus involved in the sterile system. Relevant information on this

aspect was not available in the literature. Moreover, removal of the microcystins (m-LR

and m-LA), which were at trace levels (μg/L), could easily be affected by any change in

Page 82: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

68

the water and GAC characteristics. It was proposed that the fluctuation in the results

obtained from the sterile GAC column (Figure 3.4) could be attributed to the weekly

autoclaving.

It was, therefore, considered essential to validate the experimental design for this study.

In particular, it was necessary to determine the possible side-impacts that could result

from autoclaving the sterile GAC system.

5.2 Experimental methods

To investigate possible impacts from autoclaving, a series of experiments (Section 2.4)

were carried out on the equipment items involved in the sterile column system that were

regularly autoclaved including source water, GAC and associated apparatus. These

autoclave studies consisted of the following four experiments: 1) water quality before

and after autoclaving were measured for change in terms of DOC, UVA254 and HPSEC;

2) increase of the carbon concentration in the water due to carbon leaching from the

GAC structure and tubing during autoclaving was quantified by autoclaving the items

separately in deionised water, which was consequently subjected to carbon analysis; 3)

variation of GAC adsorption efficiency due to autoclaving was examined in a jar test by

comparing toxin adsorption kinetics of a PAC sample (the identical type of carbon A6 as

the GAC) before and after autoclaving; and 4) desorption of the pre-adsorbed organic

carbon from the GAC during autoclaving was quantified by periodically analysing the

carbon in the effluent stream from the GAC column that had been autoclaved.

5.3 Results and Discussion

5.3.1 Effects on water quality

HPSEC results from the treated water before and after autoclaving are shown in Figure

Page 83: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

69

5.1. No variation was observed in molecular weight distribution of the NOM

constituents in the water after autoclaving. This was supported by the results from DOC

and UVA254 analysis (indicated in Figure 5.1), which were additional parameters used

for determination of the water quality. Thus, there was no observed effect on water

quality during autoclaving in regard to the HSPEC analysis.

Figure 5.1 Molecular weight (MW) distributions, DOC and UV254 results of Myponga treated water before and after autoclaving determined by high performance size exclusion

chromatography

5.3.2 Loss of organic carbon from GAC structure and apparatus

A virgin GAC A6 sample, which was the same amount as used in the column

experiments, was repeatedly autoclaved in deionised water (DOC ≤0.07 mg/L) to

quantify carbon that may be released from GAC during autoclaving. Results sh1own in

Figure 5.2 indicate the amount of carbon lost during autoclaving characterised by DOC

and UVA254. Approximately 0.8 mg DOC (or 0.002 cm-1 UV absorption) was released

from the GAC structure during a single autoclaving run, which accounted for

approximately 0.4% of the carbon treated by the GAC column (based on a flow rate of

Page 84: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

70

4.9 mL/min and DOC of the water at ~6 mg/L).

0

1

2

3

4

5

6

7

1st 2nd 3rd 4th 5th 6th

Autoclaving set

Leac

hing

of D

OC

(mg)

0

0.002

0.004

0.006

0.008

0.01

0.012

Leac

hing

of U

V ab

s ca

rbon

@

254

nm

DOC leached

UV leached

Figure 5.2 Carbon leached from the original GAC structure during six sets of autoclaving as determined by DOC and UV abs @ 254 nm

HPSEC results plotted in Figure 5.3 show the MW distribution of organic carbon in the

water, which was autoclaved with the virgin GAC. The HPSEC and DOC/UVA254

results were consistent as a negligible amount of carbon was released from the GAC

after autoclaving (pink curve) compared to the source water (blue curve). Therefore, it

was concluded that the weekly sterilisation treatment via autoclaving did not result in

significant carbon loss from the GAC structure compared to the total amount of carbon

treated through the GAC column.

Page 85: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

71

-0.001

0.000

0.001

0.002

0.003

0.004

0.005

0.006

0.007

100 1000 10000

Apparent Molecular Weight (Daltons)

UV

Abs

@ 2

60 n

m

The source water

Milli-Q water autoclavedwith GAC A6

Figure 5.3 Comparison of the carbon released from the GAC structure during autoclaving and

the carbon content of the source water. Determined by high performance size exclusion chromatography

In addition to the GAC structure, carbon release could occur from autoclaving the

experimental apparatus, especially the silica tubing and connectors. The data plotted in

Figure 5.4 show the amount of carbon released from the apparatus after autoclaving.

The results are based on the area of each tube that was in contact with the water.

Although higher concentrations of carbon were evident in water samples that were

autoclaved with tubing, minimal carbon, approximately 3 mg in DOC, was released

during one autoclaving run, which accounted for less than 1% of the carbon treated by

the system in one week. It was, therefore, concluded that the possible confounding

effects from autoclaving experimental apparatus of the sterile GAC system was

assumed to be insignificant, although to fully quantify this observation repeated

experiment would be required to give statistic significant results.

Page 86: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

72

0

1

2

3

4

5

6

7

Tube 1 Tube 2 Tube 3 Tube 4 Tube 5 Tube 6 Tube 7 Tube 8

Tubing

DO

C (m

g/L)

Figure 5.4 Concentration of carbon leached from the autoclaved tubing compared with the

usual DOC concentration in the source water

5.3.3 Effects on GAC adsorption efficiency

Alteration of GAC (A6) adsorption efficiency after autoclaving was examined using the

methods described in Section 2.4.3. The adsorption behaviour of PAC A6 was analysed

from a sample before and after autoclaving. The adsorption rate was measured with

respect to removal of microcystins (m-LR and m-LA) and NOM (DOC and UVA254)

from the Myponga treated water (Figure 5.5). The adsorption rate for both the toxin and

NOM constituents were identical in the autoclaved and non-treated PAC samples. Thus,

it was concluded that autoclaving GAC A6 did not significantly affect adsorption

capacity for both microcystins and NOM constituents in the source water.

DOC concentration in the source water

Carbon leached from the autoclaved tubing

Page 87: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

73

Figure 5.5 Comparison of the adsorption behaviour of PAC A6 before and after autoclaving as determined from microcystins (m-LR and m-LA) and NOM (UVA254 and DOC) in the Myponga

source water

5.3.4 Effects on desorption during autoclaving

The pre-adsorbed NOM and microcystins could be desorbed from the sterile GAC

during autoclaving (Section 2.4.4), and consequently result in greater adsorption

capacity for microcystins.

Page 88: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

74

y = 0.1113x-0.0819

R2 = 0.9048

0

0.02

0.04

0.06

0.08

0.1

0.12

0 1000 2000 3000 4000 5000 6000 7000

Water treated after autoclaving (mL)

UV a

bs @

254

nm

InfluentEffluent after autoclavingEffluent prior to autoclavingTrendline (Effluent after autoclaving)

Figure 5.6 Variation of DOC and UVA254 in the effluent of the autoclaved GAC column as a

function of the volume (mL) of treated water

The desorption of NOM from the autoclaved GAC was measured using DOC

concentration and UV254 absorption (Figure 5.6). It was evident that desorption of the

pre-adsorbed carbon occurred from the autoclaved GAC as the concentration of NOM

(characterised by DOC and UVA254) was elevated in the effluent water of the column.

Approximately 4 times higher DOC concentration was shown in the effluent than in the

influent water when the column was operated immediately after autoclaving.

Subsequently the desorption of DOC decreased rapidly and ceased after approximately

y = 139.34x-0.5662

R2 = 0.9866

0

4

8

12

16

20

0 1000 2000 3000 4000 5000 6000 7000

Water treated after autoclave (mL)

DOC

(mg/

L)InfluentEffluent prior to autoclavingEffluent after autoclavingTrendline (Effluent after autoclaving)

Water passed through the filter after autoclaving (mL)

Water passed through the filter after autoclaving (mL)

Page 89: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

75

500 mL of treated water had passed through the column. The UV254 absorbance from

the outlet of the GAC column was almost doubled immediately after autoclaving. This

elevated absorbance then gradually decreased and vanished after about 5 L water had

flowed through the column (Figure 5.6).

-0.002

0.003

0.008

0.013

0.018

100 1000 10000 100000

Apparent Molecular Weight

UV

Abs

@ 2

60 n

m

Supernatant of the autoclavedSterile GAC

The Myponga treated water

Figure 5.7 MW distribution of the UV sensitive organic carbons that were desorbed from GAC during autoclaving and that were present in the source water (determined by HPSEC)

The UV sensitive organic carbons that were desorbed from the GAC were analysed

using HPSEC (Figure 5.7). The MW distribution of the desorbed organic carbon

indicated that the desorption did not peak at the MW range that would potentially affect

the adsorption of microcystins (>1000 Daltons). Rather, the desorbed UV sensitive

organics showed a similar MW distribution to the organics in the source water except

with the organics at MW of approximately 500 Daltons (indicated in Figure 5.7).

Moreover, the desorbed organics possessed relatively lower aromaticity than the source

water as shown by the specific UV absorbance (SUVA = UVA254 (m-1) / DOC (mg/L))

parameter (Figure 5.8), which indicates that aromatic compounds are less likely to be

desorbed from the GAC.

Page 90: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

76

0

0.5

1

1.5

2

2.5

0 1000 2000 3000 4000 5000 6000 7000

Water passed through the column after autoclaving (mL)

SUV

A, L

/(mg.

m)

Influent source water

Released during autoclaving

Figure 5.8 Specific UV absorbance (SUVA) of the released organic carbons from GAC during autoclaving, and the SUVA of the source water

Quantitatively, however, the amount of NOM desorbed was not significant compared to

the amount adsorbed. The total amount of desorbed organic carbon could be calculated

using the trendline equations of the desorption results (indicated in Figure 5.6). The

calculation showed that approximately 2 mg of DOC or 0.004 of UV254 absorbing

carbon was desorbed from the GAC during one autoclave run, while during one week,

approximately 50 mg of DOC or 1.75 of UV254 absorbing carbon was adsorbed by the

GAC at the time when the test was conducted (2.5 months from start up). Additionally,

microcystins were not detected in the water in which the preloaded GAC was

autoclaved. Thus desorption of the pre-adsorbed microcystins was less likely to occur

during autoclaving of GAC. The results demonstrated that the sterilisation treatment via

autoclaving did not result in considerable desorption of the pre-adsorbed substances

from the sterile GAC.

5.4 Conclusion

The possible impact from autoclaving the sterile GAC system was quantified by a series

Page 91: Adsorption and biological filtration of microcystins

Chapter 5 Evaluating The Validity of The Experimental Design

77

of experiments that covered various aspects including water quality, GAC adsorption

efficiency and organic detachment from GAC during autoclaving. It was concluded that

no significant impact occurred due to the regular autoclaving treatment of the GAC. The

operation conditions for the sterile and the non-sterile GAC columns were verified and

thus, the experimental method used in this project was considered valid.

Page 92: Adsorption and biological filtration of microcystins

78

CHAPTER 6

BIODEGRADATION OF MICROCYSTINS AND THE OPERATIONAL CONDITIONS

6.1 Introduction

Granular activated carbon (GAC) has been shown to be a successful option for mitigation

of microcystins, which are recalcitrant compounds to conventional water treatment.

Adsorption of GAC plays a critical role in microcystin removal in a freshly established

filter, but the adsorption capacity and therefore the removal efficiency would be quickly

reduced by the excess loading of NOM. The subsequent biodegradation of microcystins

could be expected in GAC filters, which is able to enhance the removal efficiency.

Observations in the long-term column experiment in this study (Chapter 3) concluded that

biodegradation was the primary microcystin removal mechanism once it is initiated in

GAC columns.

To date, many investigations have been carried out on microcystin biodegradation, but

most focus on identification and isolation of the degrading bacteria, or on determination of

the detoxification process. Only a few biodegradation studies have been conducted to

identify possible impacts from the operational conditions. Christoffersen et al. (2002)

investigated the microcystin biodegradation behaviour in cultures of different DOC

properties (by adding lysed algal materials that were collected from different algal bloom

sources). The biodegradation process, in this study, was found to be not affected by the

DOC properties as almost the same reduction rates were evident in the studied algal

Page 93: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

79

cultures. In the study by Holst et al. (2003) microcystins were biodegraded in the sediment

of a water recharge facility under oxic, microaerophilic or anoxic conditions. The authors

reported that the degradation was significantly accelerated under the last condition,

suggesting that the process was affected by the abundance of electron acceptors in the

reaction solution. For the impact of other operational conditions on the microcystin

biodegradation process, however, there was no available information in the literature.

To facilitate the application of microcystin biodegradation in GAC filters, it was required

to identify the potential effects from different operational conditions. In this study, an

attempt was made to record the degradation behaviour under varied temperature, light

irradiation and water quality conditions as well as different initial bacterial concentrations.

6.2 Methods

The bacterial source used for this biodegradation study was derived from the non-sterile

GAC column that was operated for about 6 months (when biodegradation of microcystins

was evident in the column) in the long-term column experiment (Chapter 3). The

biological GAC column was subjected firstly to backwashing for removal of any

interfering contaminants such as algal growth and particle deposits, then, 5 g of the GAC

sample was withdrawn from the column. Bacteria on the GAC particles was detached from

the carbon surface by repeating the following step 5 times: the GAC was vortexed with

deionised water in a 20 mL glass test tube for 2 minutes and the supernatant was

immediately decanted and collected. The collected bacterial solutions were then mixed

thoroughly using a magnetic stirrer, and kept well mixed prior to the experiments, thus it

was assumed that the bacteria were evenly distributed within the culture.

6.2.1 Microcystin biodegradation under varied temperature, light

irradiation and water quality conditions

Seven sterilised glass bioreactors (2 L) were each filled with 1.5 L of sterilised water and

Page 94: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

80

continuously mixed with sterilised magnetic stirrers. The water contained in all five

reactors was Myponga treated water. In the other two reactors Myponga treated water was

diluted with deionised water at volume ratio of 1:1 and 1:9 resulting in 50% and 10% of

the original NOM concentration. Microcystins were added into these reactors at a target

concentration of 5 μg/L m-LR. After allowing complete mixing of the toxins in the reactors,

bacteria were introduced by adding 5 mL of the prepared bacterial solutions into each of

these bioreactors.

For determination of temperature effects, four of the bioreactors containing non-diluted

Myponga water were covered by aluminium foil (no light irradiation) and put under

different temperature conditions. Three of them were placed into incubators with

temperature settings of 25 °C, 33 °C and 40 °C respectively, and the fourth reactor was run

at ambient temperature of 22 °C.

To identify the effect of light irradiation, the fifth reactor containing non-diluted Myponga

water was also placed under ambient temperature (22 °C) conditions but without covering

(i.e. natural light irradiation allowed). The degradation profile in this reactor was compared

with the covered fourth bioreactor.

The last two reactors filled with diluted Myponga water, were used to determine the effect

of water characteristics. They were studied under ambient temperature (22 °C) without

covering. Therefore the biodegradation behaviour in these two reactors could be compared

with the fourth bioreactor as described above.

Microcystin degradation in these bioreactors was monitored for approximately 10 days.

Subsamples of 150 mL were taken from each of the containers at regular time intervals.

Microcystin concentration in these samples was determined.

Page 95: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

81

6.2.2 Microcystin biodegradation in cultures of different bacterial

concentration

In this test, another seven sterilised 2 L glass bioreactors were each filled with 1.5 L of

sterilised Myponga treated water. Microcystin solutions were then added into these reactors

with a target m-LR concentration of 10 μg/L, and the reactors were then mixed

continuously using magnetic stirrers. Different volumes (5 mL, 2.5 mL, 1.0 mL, 0.5 mL,

0.2 mL and 0.1 mL) of the prepared bacterial solution [corresponding to 75.5 x 106, 38.0 x

106, 15.1 x 106, 7.6 x 106, 3.0 x 106 and 1.5 x 106 cell/mL determined by flow cytometric

analysis (Becton Dickinson Biosciences, Australia)] were added into six of the bioreactors

respectively. One reactor was used as a control where bacteria were not introduced.

Samples (150 mL) were taken from the bioreactors at regular intervals and analysed for

microcystin concentration using HPLC.

6.3 Results and Discussion

6.3.1 Effects from temperature

The concentration of m-LR in the bioreactors that were placed under different temperature

conditions (Section 6.2.1) was monitored (Figure 6.1). Biodegradation at 25 °C and 33 °C

was very similar and presented the highest degradation rate after the initial lag period of

about 50 hours. Complete removal of the toxin occurred after approximately 75 hours in

these two bioreactors. The reactor at ambient temperature (22 °C) had the same lag-phase

(~50 hours) but the subsequent removal rate was much lower than observed in the reactors

at 25 °C and 33 °C. Toxin could not be detected after 150 hours in this bioreactor. In

contrast, decrease of m-LR was not evident in the 40 °C reactor during the experiment

(~10 days), and the bacteria died as observed during a subsequent agar plate assay.

Page 96: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

82

0

1

2

3

4

5

6

0 50 100 150 200 250

Time (hours)

m-L

R co

ncen

trat

ion

(μg/

L)

Ambient Temp (22 °C)

25 °C

33 °C

40 °C

Figure 6.1 Biodegradation of m-LR in bioreactors operated at ambient temperature (22 °C), 25 °C, 33 °C and 40 °C, respectively

It is indicated from this test (Figure 6.1) that the temperature under which microcystins are

biodegraded had a considerable effect on the degradation rate. Temperatures ranging from

25 °C to 33 °C favour this particular biological process, whilst high temperature conditions

(40 °C) did not suit the growth of bacteria and the biodegradation of microcystins ceased.

Similar observations of these temperature effects were also evident during biodegradation

of many other organic contaminants (Dibble and Bartha, 1979; Cho et al., 1994; Cadiz,

2002; Timothy, 2003 and Altaf et al., 2005).

6.3.2 Effects from light condition

Microcystin LR degradation in both of the coated (no light) and the non-coated (natural

light) bioreactors (Section 6.2.1) is shown in Figure 6.2. The lag period prior to the rapid

biodegradation process was similar in the two reactors. However the reduction rate was

more than doubled in the reactor with natural light irradiation compared with the one

without light. Total toxin removal occurred after approximately 100 hours (natural light)

and 150 hours (no light) in the reactors.

Page 97: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

83

0

1

2

3

4

5

0 50 100 150 200 250

time (hours)

m-L

R co

ncen

tratio

n (μ

g/L)

Natural light

no Light

Figure 6.2 Biodegradation of m-LR in bioreactors with and without irradiation by natural light

Therefore, for this particular biological study (Figure 6.2) the lag-phase prior to

microcystin biodegradation was not considerably affected by the light irradiation condition,

but the subsequent biological removal process was significantly accelerated by it. The

impact of exposure to light has been previously investigated on biodegradation of various

pollutant compounds, but different findings were reported. For example, Jie and Wang

(2005) has investigated biodegradation of dibutyl phthalate under illumination conditions,

and found the degradation rate slightly decreased. The author pointed out that illumination

could affect the proliferation of algae and bacteria, so retarded the biodegradation. On the

contrary, a research conducted by Hung and Liao (1996) found that organophosphate was

degraded more rapidly when UV irradiation was applied, while growth rate of biomass was

reported reduced. According to the authors, this outcome was most likely attributed to an

increase in membrane permeability to the substrate, the process of which was found to be

the rate-limiting step in the biodegradation. For the current microcystin biodegradation

study, reasons that were attributable to the light enhanced degradation rate are still

unknown, and further investigation is needed to quantify the effects.

6.3.3 Effects from water quality

Degradation of m-LR in water samples of different NOM concentrations (100%, 50% and

Page 98: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

84

10%) was determined (Figure 6.3). As described above, a lag-phase of about 50 hours was

observed in the undiluted (100% NOM) Myponga water. In contrast, no observable

lag-phase occurred in the diluted 50% NOM sample (Figure 6.3). However, in the highly

diluted water, i.e. the 10% NOM, the 50 hour lag-phase reappeared. So, the water sample

featured with medium NOM abundance in this study had significantly decreased the lag

period prior to microcystin degradation.

0

1

2

3

4

5

6

0 50 100 150 200 250 300

time (hours)

m-L

R c

once

ntra

tion

(μg/

L)

100% NOM

50% NOM

10% NOM

Figure 6.3 Biodegradation of m-LR in bioreactors of undiluted (100% NOM), 50% (NOM) diluted

and 10% (NOM) diluted Myponga treated water

NOM concentration of the water sample also showed impact during the toxin

biodegradation phase. The 10% NOM sample presented a similar degradation rate (~0.1

μg/hour) with that in the undiluted (100% NOM) water sample, whilst in the 50% diluted

solution the degradation was at a much lower rate (~0.04 μg/hour, Figure 6.3). In

extremely low NOM water, the activity of bacteria was restricted by nutrient supply

shortage and thus affected the metabolism efficiency of microcystins. The microorganisms

in the 10% NOM water died at approximately 75 hours as shown in a subsequent agar plate

assay, and consequently the degradation of microcystins halted at that time (Figure 6.3).

This is in agreement with the observation from the study by Bengtsson and Zerhouni

Page 99: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

85

(2003). The originally slow biodegradation process of polycyclic aromatic hydrocarbons

(PAHs) in low organic-content soil cultures was found to be accelerated by increasing the

NOM in the cultures. The authors attributed this phenomenon to the restricted

microorganic activity by the initial limited nutrient supply.

The finding of restricted biodegradation of microcystins under extremely low NOM

concentrations is essential for practical application of biological GAC filtration. It

indicates that if extremely low nutrient levels are present in the water that is being treated

to eliminate microcystins, alternative treatment profiles, rather than biodegradation must be

considered.

6.3.4 Effects from bacterial cell concentration

Microcystin LR degradation in bioreactors with a series of different bacterial

concentrations (Section 6.2.2) is shown in Figure 6.4. The degradation behaviour,

especially the lag-phase in each of the reactors was monitored. As expected, the higher the

concentration of the initial bacterial cells, the shorter the lag period and the quicker the

complete degradation. A lag time of approximately 100–150 hours was taken for the first

three highest cell concentration reactors, but of more than 200 hours was required for the

other three low bacterial concentration reactors (Figure 6.4).

0

2

4

6

8

10

0 50 100 150 200 250

Time (hours)

m-L

R c

once

ntra

tion

(μg/

L)

Control

75.5x10 6 cell/mL

38.0x10 6 cell/mL

15.1x10 6 cell/mL7.55x10 6 cell/mL

3.02x10 6 cell/mL

1.51x10 6 cell/mL

Figure 6.4 Biodegradation of m-LR in bioreactors of different bacterial cell concentration

Page 100: Adsorption and biological filtration of microcystins

Chapter 6 Biodegradation of Microcystins and the Operational Conditions

86

Although different lag period was evident (Figure 6.4), the biodegradation rate of m-LR

was shown to be very similar in all bioreactors. Total removal of m-LR was realised after

approximately 50 hours in all biological reactors after the lag-phase. It was thus

demonstrated that the initial bacterial concentration would have limited effect on the

biodegradation rate after the initial lag period.

6.4 Conclusions

During these batch tests, an attempt was made to identify the effect of selected operational

conditions on the biodegradation of microcystins. It was concluded from the outcome of

these studies that the toxin biodegradation process was easily affected by the tested

parameters (temperature, light irradiation, NOM water characteristics and the initial

bacterial concentration) in respect to the lag-phase and the degradation rate. Based upon

the observations reported, the lag period varied under different water characteristics (NOM

concentration) and different bacterial concentrations; while the operational conditions of

temperature, light irradiation and NOM affected the rate of biodegradation that occurred

after the lag period.

The results observed from this study were reasonably predictable, and for a full

understanding of the effect of varying these operational conditions further work is

necessary. Variation of operating parameters is unavoidable in practice and reasonably

unpredictable in large-scale water treatment facilities, so it would be difficult to model and

estimate the application of microcystin biodegradation under real world operating

conditions.

Page 101: Adsorption and biological filtration of microcystins

87

CHAPTER 7

CONCLUSIONS

7.1 Conclusions

Treatment of cyanobacterial toxins, in particular microcystins, in contaminated water

has become an ever-increasing problem for drinking water utilities internationally due

to the increasing occurrence of these potentially hazardous compounds. Granular

Activated Carbon (GAC) filtration, which has been readily adopted in many developed

countries for removal of problematic micropollutants, is a promising option for

treatment of microcystins due to the ability of GAC to accommodate mechanisms of

both adsorption and biodegradation. However, insight into the individual toxin removal

from the two mechanisms that determine the effective lifetime of a GAC filter is

required to optimise the application of this technology. Valuable information is provided

from this study in which the partitioning of the relative importance of adsorption and

biodegradation of microcystins in GAC filtration was successfully demonstrated.

The two microcystin reduction mechanisms in GAC filtration were successfully

separated in a lab-scale, column experiment, which was validated (Chapter 5) and

operated continuously for 11 months. In this study, GAC A6 columns with an empty bed

contact time (EBCT) of 15 minutes were used for treatment of m-LR and m-LA at a

total inflow concentration of 10-20 μg/L in Myponga treated water. The whole

time-course of GAC operation was divided into adsorption and biodegradation phases.

In the initial adsorption phase, competition from the natural organic matter in the source

water resulted in a very short lifetime (7-10 days or 900–1400 bed volumes) of the

Page 102: Adsorption and biological filtration of microcystins

Chapter 7 Conclusions

88

virgin GAC bed for effective microcystin removal (<1 μg/L m-LR in effluent).

Adsorption of both m-LR and m-LA followed a typical breakthrough curve, but higher

adsorbability was evident for m-LR during this study. The rate of adsorption during this

period decreased from ≥ 520 to 280 μg/m2/d (m-LR). The plug flow homogeneous

surface diffusion model was used for modelling the microcystin adsorption in the sterile

and the non-sterile GAC columns, where the mass transfer coefficient, Kf, was shown to

be lower for the latter column. This was attributed to the biofilm that blocked the

external pores and decreased the surface diffusion of microcystins into the internal

adsorption sites.

Transition from the adsorption phase to the biodegradation phase occurred quickly and

was difficult to predict. In the later phase, biological removal of microcystins greatly

enhanced the behaviour of the GAC columns and biodegradation became the primary

mechanism. The dosed toxins were completely biodegraded, which led to a high

elimination rate of 440 μg/m2/d (m-LR). In addition, GAC appeared to be more

favorable as a substrate for biofilm attachment and development than the sand with the

same particle size used during this study. This was attributed to the distinctive surface

characteristics of the two materials and the ability of microorganisms to colonise on

them.

The highly efficient biodegradation of microcystins in GAC filtration was difficult to

predict, because this process could be easily affected by a series of operational

conditions such as temperature, light irradiation, NOM water characteristics and

bacterial concentration. For development of a mathematic model to estimate this

biological application in water treatment facilities, further studies are necessary to fully

understand the effect of these operational variances.

Natural organic matter removal in the GAC columns was also investigated. Several

specific features were observed from their breakthrough curves due to the relatively

Page 103: Adsorption and biological filtration of microcystins

Chapter 7 Conclusions

89

high NOM concentration in the water. The adsorption efficiency was shown to decrease

rapidly and the adsorption capacity was almost saturated in a very short time (about two

months). The preferred high efficiency steady-state adsorption phase was not observed

because the NOM adsorption zone rapidly reached the end of the column.

Low percentage (~10%) NOM removal in preloaded GAC filters (several months) is

frequently observed in practical applications. The mechanisms that led to this removal

were unknown and usually assumed to be biodegradation. During this study similar

removal was evident in the adsorption GAC column (the sterile GAC), which indicated

that the minimal NOM reduction rate could be contributed, in part, to the residual

adsorption capacity of the GAC rather than biodegradation as is generally assumed in

the literature.

7.2 Future work It was concluded from this study that biodegradation is most likely to be the primary

function for maintaining efficient microcystin reduction in GAC filters, thus, it is of

vital importance to facilitate this process. However, the lag-phase for the growth of

microorganisms during acclimitisation is unavoidable and currently there is limited

understanding on how to shorten or remove the lag-phase prior to the initiation of

biodegradation, which is hindering the implementation of this technology. Further

investigation is required to identify optimal growth conditions to minimise the

lag-phase.

Additionally, a validated mathematical model that could predict the efficiency of a GAC

filter for microcystin removal as a function of operation time would be of tremendous

help and enable practical application of this technology. The model could be developed

based on the two functioning mechanisms, adsorption and biodegradation, the removal

efficiency of which was successfully obtained from this study.

Page 104: Adsorption and biological filtration of microcystins

Chapter 8 References

90

CHAPTER 8

REFERENCES Aiken G. and Cotsaris E. (1995) Soil and Hydrology: their effect on NOM. Journal of

American Water Works Association, 87(1), 36–45. Altaf H. W., Deborah R. F. and Jeffrey L. D. (2005) RDX Biodegradation Column

Study: Extent of RDX Mineralization and Influence of Temperature on Rate of RDX Biotransformation. Environmental Engineering Science 22(3), 310-323.

Anselme C., Suffet I. H. and Mallevialle J. (1988) Effects of ozonation on tastes and

odors. Journal of the American Water Works Association 80, 45-51.

APIS (2005) Website: Air Pollution Information System Eutrophication. Accessed at 11:30am, 18/11/2005.

Åsa E. and Teknisk M. (1997) Herbicide adsorption to activated carbon in drinking water treatment. http://www.miljo.chalmers.se/forskarutbildning/abstract/Edell _A.pdf. Accessed at 04:30pm, 18/12/2004.

Bell S. G. and Codd G. A. (1994) Cyanobacterial toxins and human health. Reviews in Medical Microbiology 5, 256-264.

Bending G. D. Shaw E. and Walker A. (2001) Spatial heterogeneity in the metabolism

and dynamics of isoproturon degrading microbial communities in soil. Biology & Fertility of Soils 33(6), 484-489.

Bengtsson G. and Zerhouni P. (2003) Effects of carbon substrate enrichment and DOC

concentration on biodegradation of PAHs in soil. Journal of Applied Microbiology 94, 608–617.

Bernezeau F. (1994) Can microcystins enter drinking water distribution systems, In:

D.A. Steffensen and B.C. Nicholson [Eds] Toxic Cyanobacteria, Current Status of Research and Management. Proceedings of an International Workshop, Adelaide, Australia, American Water Works Association Research Foundation, Australian Centre for Water Quality Research, Centre for Water Research, Belgium, 115-118.

Page 105: Adsorption and biological filtration of microcystins

Chapter 8 References

91

Bishop P. L., Zhang T. C. and Fu Y. C. (1995) Effects of biofilm structure, microbial

distributions and mass transport on biodegradation processes. Water Science and Technology 31(1), 143-152.

Botes D. P., Kruger H. and Viljoen C. C. (1982a) Isolation and characterization of four

toxins from the blue green alga Microcystis aeruginosa. Toxicon 20, 945-954. Botes D. P., Viljoen C. C., Kruger H., Wessels P. L. and Williams D. H. (1982b)

Configuration assignments of the amino acid residues and the presence of N-methy dehydroalanine in toxins from the blue-green alga Microcystis aeruginosa. Journal of the American Chemical Society 1, 2745-2748.

Bourke A. T. C., Hawes R. B., Neilson A. and Stallman N. D. (1983) An outbreak of

hepatoenteritis (the Palm Island mystery disease) possibly caused by algal intoxication. Toxicon 3, 45-48.

Bruchet A., Bernazeau F., Baudin I. and Pieronne P. (1998) Algal toxins in surface

waters: Analysis and treatment. Water Supply 16(1/2), 619-623. Cadiz (2002) Marine Biomediation Technologies Screening Matrix and Reference

Guide. http://europa.eu.int/comm/environment/civil/marin/reports_publications/AECIPE_FLOWCHART_95.pps#1 Accessed at 11.45am, 24/03/2005.

Carlile P. R. (1994) Further studies to investigate microcystin-LR and anotoxin-a

removal from water. Foundation for Water Research Report, FR 0458, Swindon, UK.

Carmichael W. W. (1992) Cyanobacterial Secondary Metabolites-the Cyanotoxin.

Journal of Applied Bacteriology 72, 445-459. Carmichael W. W. (1994) The toxins of cyanobacteria. American Scientist 270, 78-86. Chin Y. P., Aiken G. and O’Loughlin E. (1994) Molecular weight, polydispersity, and

spectroscopic projperties of aquatic humic substances. Environmental Science & Technology 28(11), 1853-1858.

Cho K. J., Kazuto T. and Mitsumasa O. (1994) Fate of tricresyl phosphate isomers in

Kurose River (Japan). Water Science and Technology 30(10), 189–197. Chorus J. and Bartram J. (1999) Toxic Cyanobacteria in Water: A Guide to Public

Health Significance, Monitoring and Management Für WHO durch E & FN Spon/Chapman & Hall, London.

Page 106: Adsorption and biological filtration of microcystins

Chapter 8 References

92

Choudry G. G. (1984) Humic substances-structural, photo-physical, photochemical and

free radical aspects of interactions with environmental chemicals. Gordon and Breach, New York.

Chow C. W. K., Drikas M., House J. and Burch M. D. (1997) Removal of intact

cyanobacterial cells by water treatment. Urban Water Research Association of Australia, Research Report No. 134.

Chow C. W. K., Fabris R. and Drikas M. (2004) A rapid fractionation technique to

characterise natural organic matter for the optimization of water treatment processes. Aquatic 53, 85-92.

Chow C. W. K., House J., Velzeboer R. M. A., Drikas M., Burch M. D. and Steffensen

D. A. (1998) The effect of ferric chloride flocculation on cyanobacterial cells. Water Research 32(3), 808-814.

Christoffersen K., Lyck S. and Winding A. (2002) Microbial activity and bacterial

community structure during degradation of microcystins. Aquatic Microbial Ecology 27, 125-136.

Churchill S. A., Harper J. P. and Churchill P. F. (1999) Isolation and characterization of

a Mycobacterium species capable of degrading three- and four-ring aromatic and aliphatic hydrocarbons. Applied & Environmental Microbiology 65(2), 549-552.

Codd G. A. (1995) Cyanobacterial toxins: Occurrence, properties and biological

significance. Water Science & Technology 32(4), 149-156. Codd G. A. and Bell S. G. (1996) The occurrence and fate of blue-green algal toxins in

freshwaters. London, Her Majesty’s Stationery Office (National Rivers Authority R and D Report 29).

Coquet S. (2001) Évaluation des risques lies à la présence de cyanobactéries dans les

eaux destinées à la consommation humaine, rapport AFSSA, Paris. Cook D. and Newcombe G. (2002) Removal of microcystin variants with powdered

activated carbon. Water Science & Technology: Water Supply 2(5/6), 201-207. Cornish B. J. P. A., Linda A. L. and Peter K. J. R. (2000) Hydrogen peroxide enhanced

photocatalytic oxidation of microcystin-LR using titanium dioxide. Applied Catalysis B: Environmental 25, 59-67.

Cousins I. T, Bealing D. J., James H. A. and Sutton A. (1996) Biodegradation of

Page 107: Adsorption and biological filtration of microcystins

Chapter 8 References

93

microcystin-LR by indigenous mixed bacterial populations. Water Research 30(2), 481-485.

Craig K. and Bailey D. (1995) Cyanobacterial toxin microcystin-LR removal using

activated carbon – Hunter Water Corporation Experience. In Proceedings of the 16th Federal AWWA Convention, April 2-6, 1995, Sydney, Australia, 579-586.

Dibble J. T. and Bartha R. (1979) Effect of environmental parameters on the

biodegradation of oil sludge. Applied Environmental Microbiology 37(4), 729–739.

Donati C., Drikas M., Hayes R. and Newcombe G. (1993) Adsorption of microcystin-

LR by powdered activated carbon. Water 20(3), 25-28. Donati C., Drikas M., Hayes R. and Newcombe G. (1994) Microcystin-LR adsorption

by powdered activated carbon. Water Research 28(8), 1735-1742. Eriksson J. E., Toivola D., Meriluoto J. A. O., Karaki H., Han Y. G. and Hartshorne D.

(1990) Hepatocyte deformation induced by cyanobacterial toxins reflects inhibition of protein phosphatases. Biochemical and Biophysical Research Communications 173, 1347-1353.

Eynard F, Mez K, Walther J. L. (2000) Risk of cyanobacterial toxins in Riga waters

(Latvia). Water Research 34, 2979-2988. Falconer I. R., Beresford A. M. and Runnegar M. T. C. (1983) Evidence of Liver

Damage by Toxin. From a Bloom of the Blue-Green Algae, Microcystis aeruginosa. Medical Journal of Australia 1, 511-514.

Falconer I. R., Runnegar M. T. C., Buckley T., Huyn Y. L. and Bradshaw P. (1989) Using

activated carbon to remove toxicity from drinking water containing cyanobacterial blooms. Journal of AWWA 102-105.

Falconer I. R., Runnegar M. T. C. and Huynh V. L. (1983) Effectiveness of activated

carbon in the removal of algal toxin from potable water supplies: a pilot plant investigation. Proceedings of the 10th Australian Water and Wastewater Association Federal Convention, (26)1-(26)8. Australian Water and Wastewater Association Incorporation., Sydney, Australia.

Fawell J. K., Hart J., James H. A. and Parr W. (1993) Blue-green algae and their toxins-

Analysis, toxicity, treatment and environmental control. Water Supply 11(3/4), 109-121.

Page 108: Adsorption and biological filtration of microcystins

Chapter 8 References

94

Feitz A. J., Waite T. D., Jones G. J. Bayden B. H. and Orr P. T. (1999) Photocatalytic degradation of the blue green algal toxin microcystin-LR in a natural organic-aqueous matrix. Environmental Science & Technology 33, 243-249.

Fott B. (1971) Algenkunde, 2. Stuttgart: Aufl. Gustav Fischer Verlag. Francis G. (1878) Poisonous Australian Lake. Nature 18, 11-12. Freundlich H. (1906) Über die adsorption in lösungen. Zeitschrift für physikalische

Chemie 57, 385-470. Gotvajn A. Z. and Zagorckoncan J. (1996) Comparison of biodegradability assessments

for chemical substances in water. Water Science & Technology 33(6), 207-212. Grützmacher G., Böttcher G., Chorus I. and Bartel H. (2002) Removal of microcystins

by slow sand filtration. Environmental Toxicology 17(4), 386-394. Guan X., Shang C. and Chen G. H. (2005) Competitive adsorption of organic matter

with phosphate on aluminum hydroxide. Journal of Colloid and Interface Science 296(1), 51-58.

Guerin W. F. and Boyd S. A. (1997) Bioavailability of naphthalene associated with

natural and synthetic sorbents. Water Research 31, 1504-1512.

Guieysse B., Wikstrom P., Forsman M., and Mattiasson B. (2001) Biomonitoring of continuous microbial community adaptation towards more efficient phenol-degradation in a fed-batch bioreactor. Applied Microbiology & Biotechnology. 56(5-6), 780-787.

Gupta S. (1998) Cyanobacterial toxins: Microcystin-LR. Guidelines for drinking-water

quality. Health criteria and other supporting information 2, 95-110. Haider S., Naithani V., Viswanathan P. N. and Kakkar P. (2003) Cyanobacterial toxins: a

growing environmental concern. Chemosphere 52, 1-21. Harada K. I., Imanishi S., Kato H., Masayoshi M., Ito E. and Tsuji K. (2004) Isolation

of Adda from microcystin-LR by microbial degradation. Toxicon 44 (1), 107–109.

Hart J. and Stott P. (1993) Microcystin-LR removal from water. Foundation for Water

Research Report, FR 0367, Swindon, UK. Himberg K., Keijola A. M., Hiisvirta L., Pyysalo H. and Sivonen K. (1989) The effect

Page 109: Adsorption and biological filtration of microcystins

Chapter 8 References

95

of water treatment processes on the removal of hepatotoxins from Microcystins and Oscillatoria cyanobacteria; A laboratory study. Water Research 23(8), 979-984.

Hitzfeld B. C., Hoger S. J., Dietrich R. (2000) Cyanobacterial toxins: removal during

drinking water treatment, and human risk assessment. Environmental Health Prospect 10, 113-122.

Ho S. W. L. (2004) The removal of cyanobacterial metabolites from drinking water

using ozone and granular activated carbon. PhD Thesis. The University of South Australia.

Ho L., Meyn T, Keegan A, Hoefel D., Brookes J., Saint C. P. and Newcombe G. (2006)

Bacterial degradation of microcystin toxins within a biologically active sand filter. Water Research 40(4) 768-774.

Hoffman J. R. H. (1976) Removal of Microcystis toxins in water purification processes.

Water SA 2(2), 58-60. Holst T., Jorgensen N. O. G., Jorgensen C., and Johansen A. (2003) Degradation of

microcystin in sediments at oxic and anoxic denitrifying conditions. Water Research 37, 4748-4760.

Hung S. C. and Liao J. C. (1996) Applied Biochemical Biotechnology 56 (1), 37-47. Hyenstrand P., Metcalf J. S., Beattie K. A. and Codd G. A. (2001) Losses of the

cyanobacterial toxin microcystin-LR from aqueous solution by adsorption during laboratory manipulations. Toxicon 39, 589-594.

Ingerslev F. Torang L. and Nyholm N. (2000) Importance of the test volume on the

lagphase in biodegradation studies. Environmental Toxicology & Chemistry 19(10), 2443-2447.

Ishii H., Nishijima M. and Abe T. (2004) Characterization of degradation process of

cyanobacterial hepatotoxins by a Gram-negative aerobic bacterium. Water Research 38 (11), 2667–2676.

Jia Y. F. and Demopoulos G. P. (2002) Adsorption of Silver onto Activated Carbon from Acidic Media: Nitrate and Sulfate Media. Industry and Engineering Chemical Research 42 (1), 72 -79.

Jie C. and Wang Z. K. (2005) Environmental Protection Science 31(6), 8-10. Jochimsen E. M., Carmichael W. W., Cardo D. M., Cookson S. T., Holmes C. E. M.,

Page 110: Adsorption and biological filtration of microcystins

Chapter 8 References

96

Antunes B. C., Filho D. A. M., Lyra M. D., Barreto V. S., Azevedo S. M. F. O. and Jarvis W. R. (1998) Liver failure and death after exposure to microcystins at a haemodialysis center in Brazil. New England Journal of Medicine 338, 873-878.

Jones G. J. and Orr P. T. (1994) Release and degradation of microcystin following

algicide treatment of a Microcystis aeruginosa bloom in a recreational lake, as determined by HPLC and protein phosphatase inhibition assay. Water research 28, 871-876.

Jones G. J. and Negri A. P. (1997) Persistence and degradation of cyanobacterial

paralytic shellfish poisons (PSPs) in freshwaters. Water Research 31(3) 525-533.

Jones L. R., Owen S. A., Horrel P. and Burns R. G. (1998) Bacterial inoculation of

granular activated carbon filters for the removal of atrazine from surface water. Water Research 32, 2542-2549.

Jones G., Minatol W., Craig K. and Naylor R. (1993) Removal of low level

cyanobacterial peptide toxins from drinking water using powdered and granular activated carbon and chlorination – Results of laboratory and pilot plant studies. In Proceedings of the 15th Federal AWWA Convention, April 18-23, 1993, Gold Coast, Queensland, Australia, 579-586.

Jonge R. J., Breure A. M., and Andel J. G. (1996) Bioregeneration of powdered

activated carbon (PAC) loaded with aromatic compounds. Water Research 30, 875-882.

Jung M. W., Ahn K. H., Lee Y., Kim K. P., Rhee J. S., Park J. T. and Paeng K. (2001)

Adsorption characteristics of phenol and chlorophenols on granular activated carbon (GAC). Microchemistry Journal 70, 123-131.

Kabzinski A. K. M., Juszczak R., Miekos E., Tarczynska M, Sivonen K. and Rapala J. (2000) Polish Journal of Environmental Studies 9, 171.

Kaebernick M., Neilan B. A., Borner T. and Dittmann E. (2000) Light and the transcriptional response of the microcystin biosynthesis gene cluster. Applied and Environmental Microbiology 66, 3387-3392.

Keijola A. M., Himberg K., Esala A. L., Sivonen K. and Hiisvirta L. (1988) Removal of cyanobacterial toxins in water treatment processes: laboratory and pilot-scale experiments. Toxic Assess 3, 643.

Laat J. D., Bouanga F., Dore M., and Mallevialle J. (1985) Influence of bacterial growth

Page 111: Adsorption and biological filtration of microcystins

Chapter 8 References

97

in granular activated carbon filters on the removal of biodegradable and of non-biodegradable organic compounds. Water Research 19 (12), 1565-1578.

Lahav O. and Green M. (2001) Bioregenerated ion-exchange process: the effect of the

biofilm on the ion-exchange capacity and kinetics. Water SA 26, 51-57. Lahti K. and Hiisvirta L. (1989) Removal of cyanobacterial toxins in water treatment

processes: Review of studies conducted in Finland. Water Supply 7, 149-154. Lahti K., Niemi M. R., Rapala J. and Sivonen K. (1998) Biodegradation of

cyanobacterial hepatoxins—characterization of toxin degrading bacteria. Harmful algae 363-365. Xunta de Galicia and Intergovernmental Oceanographic Commission of UNESCO 1998. Santiago de Compostela, Spain.

Lahti K., Rapala J., Kivimaki A. L., Kukkonen J., Niemela M. and Sivonen K. (1997)

Occurrence of microcystins in raw water sources and treated drinking water of Finnish waterworks. Water Science Technology 43, 225-228.

Lam A. K. Y., Prepas E. E., Spink D. and Hrudey S. E. (1995) Chemical control of

hepatotoxic phytoplankton blooms: implications for human health. Water Research 29, 1845-1854.

Lambert T. W., Holmes C. F. B. and Hrudey S. W. (1996) adsorption of microcystin-LR

by activated carbon and removal in full scale water treatment. Water Research 30(6), 1411-1422.

Lawton L. A. and Robertson P. K. J. (1999) Physico-chemical treatment methods for the

removal of microcystins from potable waters. Chemical Society Reviews 28, 217-224.

Li D. http://www.watercampws.uiuc.edu/pdf_documents/research_posters/3_ding_final.p

df Accessed at 4:00pm, 03/04/2005. Li F., Yuasa A., Chiharada H. and Matsui Y. (2003) Polydisperse adsorbability

composition of several natural and synthetic organic matrices. Journal of Colloid Interface Science 265(2), 265-275.

Liu I., Lawton L. A., Cornish B. and Robertson P. K. J. (2002) Mechanistic studies of

the photocatalytic oxidation of microcystin-LR: An investigation of byproducts of the decomposition process. Environmental Science & Technology 37, 3214-3219.

Lowry J. D. and Burkhead C. E. (1980) The role of adsorption in biologically extended

Page 112: Adsorption and biological filtration of microcystins

Chapter 8 References

98

activated carbon columns. Journal of Water Pollution. Control Federation. 52, 389-398.

Mankiewicz J., Walter Z., Tarczynska M., Palyvoda O, Staniaszczyk M. W. and

Zalewski M. (2002) Genotoxicity of cyanobacterial extracts containing microcystins from Polish water reservoirs as determined by SOS chromotest and comet assay. Environmental Toxicology 17(4), 341-350.

Middleton F. M. and Rosen A. A. (1956) Organic contaminants affecting the quality of

water. Public Health Reports 71, 1125-1133. Miklas S. and Robert J. M. (1998) Control of Bio-regernerated Granular Activated

Carbon by Spreadsheet Modelling. Journal of Chemical Technology and Biotechnology 71, 253-261.

Miller M. J. and Hallowfield H. J. (2001) Degradation of cyanobacterial hepatotoxins in

batch experiments. Water Science & Technology 43 (12), 229-232. Mirgain I., Green G. and Monteil H. (1995) Biodegradation of the herbicide atrazine in

groundwater under laboratory conditions. Environmental Technology 16(10), 967-976.

Mollah A. H. and Robinson C. W. (1996) Pentachlorophenol adsorption and desorption

characteristics of granular activated carbon – I. Isotherms. Water Research 30(12), 2901-2906.

Mort S. L. and Deanross D. (1994) Biodegradation of phenolic compounds by sulfate-

reducing bacteria from contaminated sediments. Microbial Ecology 28(1), 67-77.

Mort S. L., Deanross D., Churchill S. A., Harper J. P. and Churchill P. F. (1999)

Isolation and characterization of a Mycobacterium species capable of degrading three- and four-ring aromatic and aliphatic hydrocarbons. Applied and Environmental Microbiology 65(2), 549-552.

Mouchet P. and Bonnelye V. (1998) Solving algae problems: French expertise and

world-wide applications. Water Supply 47(3), 125-141. Murphy R. J., Crawford C. M. and Barmuta L. (2003) Estuarine health in Tasmania,

status and indicators: water quality, Technical Report Series Number 16, Tasmanian Aquaculture and Fisheries Institute, Hobart.

Page 113: Adsorption and biological filtration of microcystins

Chapter 8 References

99

Najm I. N., Snoeyink V. L., Suidan M. T., Lee C. H. and Richard Y. (1990). Effect of particle size and background natural organics on the adsorption efficiency of PAC. Journal of American Water Work Association 82, 65-72.

Neumann U. and Weckesser J. (1998) Elimination of microcystin peptide toxins from

water by reverse osmosis. Toxicology and Water Quality 13(2), 143-157. Newcombe, G., Morrison, J. and Hepplewhite C. (2001) Simultaneous adsorption of

MIB and NOM onto activated carbon: I Characterisation of the system and NOM adsorption. Carbon 40(12), 2135-2146.

Newcombe G. (2002) Removal of Algal Toxins from Drinking Water Using Ozone and

GAC, AWWA Research Foundation and American Water Works Association, USA.

Nicholson B. C., Rositano J., Humpage A. R. and Burch M. D. (1993) Proceedings of

the 15th Australian Water and Waste Water Association Federal Convention, Australia. 2, 327-331.

Nicholson B. C., Rositano J. and Burch M. D. (1994) Destruction of cyanobacterial

peptide hepatotoxins by chlorine and chloramines. Water Research 28(6), 1297-1303.

Oberholster P. J., Botha A. M. and Grobbelaar J. U. (2003) Microcystis aeruginosa:

source of toxic microcystins in drinking water. African Journal of Biotechnology 3(3), 159-168.

Olmstead K. P. (1989) Microbial interference with the adsorption of target organic

contaminants by granular activated carbon. Universtiy of Michigan. Orshansky F. and Narkis N. (1997) Characteristics of organics removal by PATC

simultaneous adsorption and biodegradation. Water Research 31, 391-398. Oudra B, Loudiki M., Sabour B., Martins R., Amori A. and Vasconcelos V. (2002)

Detection and variation of microcystin contents of Microcystis blooms in eutrophic Lalla Takerkoust Lake, Morocco. Lakes & Reservoir: Research and Management 7(1), 35-44.

Park H. D., Sasaki Y., Maruyama T., Yanagisawa E., Hiraishi A. and Kato K. (2001)

Degradation of cyanobacterial hepatotoxin microcystin by a new bacterium isolated from a hypertrophic lake. Environmental Toxicology and Pharmacology 16 (4), 337–343.

Page 114: Adsorption and biological filtration of microcystins

Chapter 8 References

100

Pietsch J., Bornmann K. and Schmidt W. (2002) Relevance of intra- and extracellular cyanotoxins for drinking water treatment. Acta hydrochimica et hydrobiologica 30(1), 7-15.

Preuss G. Willme U. and Zullei-Seibert N. (2002) Behaviour of some pharmaceuticals during artificial groundwater recharge - Elimination and effects on microbiology. Acta Hydrochimica et Hydrobiologica 29(5), 269-277.

Rangel-Mendez J. R. and Streat M. (2002) Adsorption of cadmium by activated carbon

cloth: influence of surface oxidation and solution pH. Water Research 36, 1244-1252.

Roberts P. V. and Summers R. S. (1982) Performance of granular activated carbon for

total organic carbon removal. Journal of American Water Works Association. 74(2), 113-118.

Roberts P. V., Summers R. S. and Regli S. (1984) Adsorption techniques in drinking

water treatment, EPA-570/9-84-005, U.S. Environmental Protection Agency, ODW, Washington, D.C.

Romanowska-Duda Z., Mankiewicz J., Tarczynska M., Walter Z. and Zalewski M.

(2002) The effect of toxic cyanobacteria (blue-green algae) on water plants and animal cells. Polish Journal of Environmental Studies 11, 561-566.

Rosen J. B. (1952) Kinetics of a fixed bed system for solid diffusion into spherical

particles. Journal of Chemical Physics 20(3), 387-394. Rositano J. (1996) The destruction of cyanobacterial peptide toxins by oxidants used in

water treatment. Urban Water Research Association of Australia Report No.110, Melbourne, Australia.

Rositano J., Nicholson B. C. and Pieronne P. (1998) Destruction of cyanobacterial

toxins by ozone. Ozone: Science & Engineering 20, 223-238. Rositano J. and Nicholson B. C. (1994) Australian Centre for Water Quality Research,

Adelaide. Report 2/94. Roy D., Wang G. T. and Adrian D. D. (1993) A simplified solution technique for

carbon adsorption model. Water Research 27 (6), 1033-1040. Saito T., Okana K., Park H. D., Itayama T., Inamori Y., Neilan B. A., Burns B. P. and

Sugiura N. (2003) Detection and sequencing of the microcystin LR-degrading

Page 115: Adsorption and biological filtration of microcystins

Chapter 8 References

101

gene, mlrA, from new bacteria isolated from Japanese lakes. FEMS Microbiology Letters 229 (2), 271–276.

Salvador F. and Jiménez S. C. (1996) A new method for regenerating activated carbon

by thermal desorption with liquid water under subcritical conditions. Carbon 34(4), 511.

Schmidt K. J. (1994) Prediction of GAC column performance using bench scale

techniques. Masters Dissertation, University of Illinois, Illinois, USA. Shawwa A. R. and Smith D. W. (2001) Kinetics of microcystin-LR oxidation by ozone.

Ozone: Science & Engineering 23(2), 161-170. Shende R. V., Mahajani V. V. (2002) Wet oxidative regeneration of activated carbon

loaded with reactive dye. Waste Management 22(1), 73-83.

Senogles-Derham P. J., Seawright A. (2003) Toxicological aspects of treatment to remove cyanobacteria toxins from drinking water determined using the heterozygous P53 transgenic mouse model. Toxicology 41, 979-988.

Senogles P. J., Scott J. A., Shaw G. and Stratton H. (2001) Photocatalytic degradation of the cyanotoxin cylindrospermopsin, using titanium dioxide and UV irradiation. Water Research 35(5), 1245-1255.

Shultz J. R. and Keinath T. M. (1984) Powdered activated carbon treatment process mechanisms. Journal of Water Pollution. Control Federation. 56, 143-151.

Skulberg O. M., Carmichael W. W., Cood G. A. and Skulberg R. (1993) Toxigenic cyanophytes - identification and taxonomy. Tn: Algal toxins in seafood and drinking water. Academic Press, London.

Skulberg O. M., Codd G. A., and Carmichael W. W. (1984) Toxic blue-green algal blooms in Europe: a growing problem. Ambil 13, 244-247.

Smith E. H., Tseng S. K. and Weber J. W. J. (1987) Modelling the adsorption of target compounds by GAC in the presence of background dissolved organic matter.

Environmental Progress 6(2), 18-25. Song Z., Suzanne R. E. and Richard G. B (2006) Treatment of naphthalene-2-sulfonic

acid from tannery wastewater by a granular activated carbon fixed bed inoculated with bacterial isolates Arthrobacter globiformis and Comamonas testosterone. Water Research 40, 495-506.

Sontheimer H., Crittenden J. C. and Summers R. S. (1988) Activated Carbon for Water

Page 116: Adsorption and biological filtration of microcystins

Chapter 8 References

102

Treatment, Second Edition, AWWA Research Fundation, USA.

Sivonen K. and Jones G. (1999) Cyanobacterial toxins. In: Toxic cyanobacteria in water. A guide to their public health consequences, monitoring and management. Routledge, London.

Summers R. S. (1989) The influence of background organic matter on GAC adsorption. Journal of American Water Work Association 89(5), 66-74.

Takenaka S. and Watanabe M. F. (1997) Microcystin-LR degradation by Pseudomonas

aeruginosa alkaline protease. Chemosphere 34, 749–757. Thomas S. Sarfaraz S. Mishra L. C. and Iyengar L. (2002) Degradation of phenol and

phenolic compounds by a defined denitrifying bacterial culture. World Journal of Microbiology & Biotechnology 18(1), 57-63.

Thurman E. M. (1985) Organic Geochemistry of Natural Waters, Martinus Nijho-Dr.

Junk Publishers, Dordrecht, The Netherlands. Timothy E. (2003) Determination of Biodegradation Rates for Ethanol Production

Wastewater. http://www.ccee.iastate.edu/research/detail.cfm?projectID=-456381360

Accessed at 5.30pm, 26/05/2005.

Tisdale E. S. and Charleston W. (1931) Epidemic of intestinal disorders in VA, occurring simultaneously with unprecedented water supply conditions. American Journal of Public Health 21, 198-200.

Traegner U. K. and Suidan M. T. (1989) Evaluation of surface and film diffusion coefficients for carbon adsorption. Water Research 23, 267-273.

Traina S. J., Novak J. and Smeck N. E. (1990) An Ultraviolet Absorbance Method of

Estimating the Percent Aromatic Carbon Content in Humic Acids", Journal of Environmental Quality 19, 151-153.

Tsuji K., Naito S., Kondo F., Ishikawa N., Watanabe M. F., Suzuki M. and Harada K. I.

(1994) Stability of microcystins from cyanobacteria: Effect of light on decomposition and isomerisation. Environmental Science & Technology 28, 173-177.

Tsuji K., Watanuki T., Kondo F., Watanabe M. F., Nakazawa H., Suzuki M., Uchida H.

and Harada K.-I. (1997) Stability of microcystins from cyanobacteria – IV. Effectct of chlorination on decoposition. Toxicon 35(7), 1033-1041.

Page 117: Adsorption and biological filtration of microcystins

Chapter 8 References

103

Turner P., Gammie A., Hoolinrake K. and Codd G. (1990) Pneumonia associated with contact with cyanobacteria. British Medical Journal 300, 1165-1175.

Ueno Y., Nagata S., Tsutsumi T., Hasegawa A., Watanabe M. F., Park H. D., Chen G. C.,

Chen G. and Yu S. Z. (1996) Detection of microcystins, a blue-green algal hepatotoxin, in drinking water sampled in Haimen and Fusui, endemic area of primary liver cancer in China, by highly sensitive immunoassay, Carcinogenesis 17, 1317-1321.

Vahala R., Långvik V. A. and Laukkanen R. (1999) Controlling adsorbable organic

halogens (AOX) and trihalomethanes (THM) formation by ozonation and two-step granule activated carbon (GAC) filtration. Water Science and Technology 40(9), 249-256.

Wannemacher R. W. (1989) Chemical stability and laboratory safety of naturally

occurring toxins. Fort Detrick, Frederick, MD, US Army Medical Research Institute of Infectious Disease.

Watanabe M. F., Tsuji K., Watanabe Y., Harada K. and Suzuki M. (1992) Release of

heptapeptide toxin (microcystin) during decomposition process of Microcystis aeruginosa. Natural Toxins 1, 48-53.

Weber W. J. and Liu K. T. (1980) Determination of mass transport parameters for fixed

bed adsorbers. Chemical Engineering Communications 6, 49-60. WHO (World Health Organisation), Guidelines for drinking-water quality, 2nd edition

(1998) Addendum to Volume 2. Health criteria and other supporting information. Geneva, 95-110.

Wietlik J., Raczyk-Stanis U., Awiak S. B., Ozor W. I. and Nawrcki J. (2002) Adsorption of natural organic matter oxidized with ClO2 on granular activated carbon. Water Research 36(9), 2328-2336.

Wu H., Zheng L., Su J., and Shi W. (2005) Survey on the contamination of microcystin-LR in water supply of Shanghai city. Journal of Hygiene Research 34 (2), 152-154.

Xu Z. Y., Zhang Q. W., Wu C. L. and Wang L. S. (1997) Adsorption of naphthalene

derivatives on different macroporous polymeric adsorbents. Chemosphere 35(10), 2269-2276.

Yavich A. A., Lee K. H., Chen K. C., Pape L. and Masten S. J. (2004) Evaluation of

biodegradability of NOM after ozonation. Water Research 38, 2839-2846.

Page 118: Adsorption and biological filtration of microcystins

Chapter 8 References

104

Yassir A., Rieu C. and Soulas G. (1998) Microbial n-dealkylation of atrazine effect of exogeneous organic substrates and behaviour of the soil microflora. Pesticide Science 54(1), 75-82.

Yoo R. S., Carmichael W. W., Hoehn R. C. and Hrudey S. E. (1995) Cyanobacterial

(Blue-Green Algal) Toxins: A Resource Guide. AWWA Research Foundation, USA.

Yuasa A. (1982) A kinetic study of activated carbon adsorption processes. PhD

Dissertation, Hokkaido University, Sapporo, Japan. Zhao X., Hickey R. F. and Voice T. C. (1999) Long-term evaluation of adsorption

capacity in a biological activated carbon fluidized bed reactor system. Water Research 33, 2983-2991.

Page 119: Adsorption and biological filtration of microcystins

105

Appendix A

Bacterial Colony Count

date day R2A agar plate A R2A agar plate B (duplicate) 24/11/2004 66 0 0 25/11/2004 67 0 0 26/11/2004 68 0 0 27/11/2004 69 0 0 28/11/2004 70 0 0 29/11/2004 71 0 0 30/11/2004 72 0 0 1/12/2004 73 0 0 2/12/2004 74 0 0 3/12/2004 75 0 0 4/12/2004 76 0 0 5/12/2004 77 0 0 6/12/2004 78 0 0 7/12/2004 79 0 0 8/12/2004 80 0 0 9/12/2004 81 0 0 10/12/2004 82 0 0 11/12/2004 83 0 0 12/12/2004 84 0 0 13/12/2004 85 0 0 14/12/2004 86 0 0 15/12/2004 87 0 0 16/12/2004 88 0 0 17/12/2004 89 0 0 18/12/2004 90 0 0 19/12/2004 91 0 0 20/12/2004 92 0 0 21/12/2004 93 0 0 22/12/2004 94 0 0 23/12/2004 95 0 0 24/12/2004 96 0 0 25/12/2004 97 0 0 26/12/2004 98 0 0 27/12/2004 99 0 0 28/12/2004 100 0 0 29/12/2004 101 0 0 30/12/2004 102 0 0 31/12/2004 103 0 0 1/01/2005 104 0 0 2/01/2005 105 0 0 3/01/2005 106 0 0 4/01/2005 107 0 0 5/01/2005 108 0 0

Page 120: Adsorption and biological filtration of microcystins

106

6/01/2005 109 0 0 7/01/2005 110 0 0 8/01/2005 111 0 0 9/01/2005 112 0 0 10/01/2005 113 0 0 11/01/2005 114 0 0 12/01/2005 115 0 0 13/01/2005 116 0 0 14/01/2005 117 0 0 15/01/2005 118 0 0 16/01/2005 119 0 0 17/01/2005 120 0 0 18/01/2005 121 0 0 19/01/2005 122 0 0 20/01/2005 123 0 0 21/01/2005 124 0 0 22/01/2005 125 0 0 23/01/2005 126 0 0 24/01/2005 127 0 0 25/01/2005 128 0 0 26/01/2005 129 0 0 27/01/2005 130 0 0 28/01/2005 131 0 0 29/01/2005 132 0 0 30/01/2005 133 0 0 31/01/2005 134 0 0 1/02/2005 135 0 0 2/02/2005 136 0 0 3/02/2005 137 0 0 4/02/2005 138 0 0 5/02/2005 139 0 0 6/02/2005 140 0 0 7/02/2005 141 0 0 8/02/2005 142 0 0 9/02/2005 143 0 0 10/02/2005 144 0 0 11/02/2005 145 0 0 12/02/2005 146 0 0 13/02/2005 147 0 0 14/02/2005 148 0 0 15/02/2005 149 0 0 16/02/2005 150 0 0 17/02/2005 151 0 0 18/02/2005 152 0 0 19/02/2005 153 0 0 20/02/2005 154 0 0 21/02/2005 155 0 0 22/02/2005 156 0 0 23/02/2005 157 0 0 24/02/2005 158 0 0 25/02/2005 159 0 0 26/02/2005 160 0 0 27/02/2005 161 0 0 28/02/2005 162 0 0

Page 121: Adsorption and biological filtration of microcystins

107

1/03/2005 163 0 0 2/03/2005 164 0 0 3/03/2005 165 0 0 4/03/2005 166 0 0 5/03/2005 167 0 0 6/03/2005 168 0 0 7/03/2005 169 0 0 8/03/2005 170 0 0 9/03/2005 171 0 0 10/03/2005 172 0 0 11/03/2005 173 0 0 12/03/2005 174 0 0 13/03/2005 175 0 0 14/03/2005 176 0 0 15/03/2005 177 0 0 16/03/2005 178 0 0 17/03/2005 179 0 0 18/03/2005 180 0 0 19/03/2005 181 0 0 20/03/2005 182 0 0 21/03/2005 183 0 0 22/03/2005 184 0 0 23/03/2005 185 0 0 24/03/2005 186 0 0 25/03/2005 187 0 0 26/03/2005 188 0 0 27/03/2005 189 0 0 28/03/2005 190 0 0 29/03/2005 191 0 0 30/03/2005 192 0 0 31/03/2005 193 0 0 1/04/2005 194 0 0 2/04/2005 195 0 0 3/04/2005 196 0 0 4/04/2005 197 0 0 5/04/2005 198 0 0 6/04/2005 199 0 0 7/04/2005 200 0 0 8/04/2005 201 0 0 9/04/2005 202 0 0 10/04/2005 203 0 0 11/04/2005 204 0 0 12/04/2005 205 0 0 13/04/2005 206 0 0 14/04/2005 207 0 0 15/04/2005 208 0 0 16/04/2005 209 0 0 17/04/2005 210 0 0 18/04/2005 211 0 0 19/04/2005 212 0 0 20/04/2005 213 0 0 21/04/2005 214 0 0 22/04/2005 215 0 0

Page 122: Adsorption and biological filtration of microcystins

108

The repeated set of experiment

date day R2A agar plate A R2A agar plate B (duplicate) 25/04/2005 1 0 0 26/04/2005 2 0 0 27/04/2005 3 0 0 28/04/2005 4 0 0 29/04/2005 5 0 0 30/04/2005 6 0 0 1/05/2005 7 0 0 2/05/2005 8 0 0 3/05/2005 9 0 0 4/05/2005 10 0 0 5/05/2005 11 0 0 6/05/2005 12 0 0 7/05/2005 13 0 0 8/05/2005 14 0 0 9/05/2005 15 0 0 10/05/2005 16 0 0 11/05/2005 17 0 0 12/05/2005 18 0 0 13/05/2005 19 0 0 14/05/2005 20 0 0 15/05/2005 21 0 0 16/05/2005 22 0 0 17/05/2005 23 0 0 18/05/2005 24 0 0 19/05/2005 25 0 0 20/05/2005 26 0 0 21/05/2005 27 0 0 22/05/2005 28 0 0 23/05/2005 29 0 0 24/05/2005 30 0 0 25/05/2005 31 0 0 26/05/2005 32 0 0 27/05/2005 33 0 0 28/05/2005 34 0 0 29/05/2005 35 0 0 30/05/2005 36 0 0 31/05/2005 37 0 0 1/06/2005 38 0 0 2/06/2005 39 0 0 3/06/2005 40 0 0 4/06/2005 41 0 0 5/06/2005 42 0 0 6/06/2005 43 0 0 7/06/2005 44 0 0 8/06/2005 45 0 0 9/06/2005 46 0 0 10/06/2005 47 0 0 11/06/2005 48 0 0 12/06/2005 49 0 0 13/06/2005 50 0 0

Page 123: Adsorption and biological filtration of microcystins

109

14/06/2005 51 0 0 15/06/2005 52 0 0 16/06/2005 53 0 0 17/06/2005 54 0 0 18/06/2005 55 0 0 19/06/2005 56 0 0 20/06/2005 57 0 0 21/06/2005 58 0 0 22/06/2005 59 0 0 23/06/2005 60 0 0 24/06/2005 61 0 0 25/06/2005 62 0 0 26/06/2005 63 0 0 27/06/2005 64 0 0 28/06/2005 65 0 0

Page 124: Adsorption and biological filtration of microcystins

110

Appendix B

Calculation of Microcystin Adsorption Rate in the Non-Sterile

GAC Column

Assuming the GAC particles used in this study were spherical thus total surface area, S

(m2), that was presented in the non-sterile GAC column can be calculated as following,

V = (п/4) * D2 * H (1)

v = (п/6) * d3 (2)

s = п * d2 (3)

S = V * s / v (4)

Where:

V – volume of the GAC bed, m3;

D – diameter of the GAC column, m;

H – height of the GAC column, m;

v – volume of one GAC particle, m3;

d – diameter of the GAC particle, m;

s – surface area of one GAC particle, m2;

S – total surface area in the GAC column, m2.

Since the highest adsorption rate normally occurs at the beginning of operation with a

virgin GAC filter, so in this study, the highest microcystin adsorption rate in the non-

sterile GAC column could be calculated based on the toxin removal results during the

initial period of time. This calculation only applies to the microcystin dosage in this study,

the highest of which was approximately 6.5 μg/L m-LR in the beginning of the repeated

Page 125: Adsorption and biological filtration of microcystins

111

set of experiment (Figure 3.9). Therefore, the detected microcystin (m-LR) adsorption

rate in the non-sterile GAC column, R (μg/m2/d), was

R = Cm-LR * υ * 60min * 24hr * 1000 / S (5)

Where:

R – the detected m-LR adsorption rate in the non-sterile GAC column;

Cm-LR – the highest m-LR concentration that used in the initial complete

adsorption period, 6.5μg/L;

υ – water flow rate of the non-sterile GAC column, mL/min.

Substitute the following operating parameters to the equations (1) – (5),

D = 2.5 * 10-2 m;

H = 15 * 10-2 m;

d = 1.2 * 10-3 m (the average value of the GAC particle size, 1.0 – 1.4 mm);

υ = 4.9 mL/min;

the observed adsorption rate is calculated,

R = 1.25 μg/m2/day.

Note: The maximum adsorption rate of microcystins was not measured in this study.