127
ENHANCING IN SITU PAH BIODEGRADATION The Effects of Amendments on Bench-Scale Bioremediation Systems by Avery Gottfried A thesis submitted in partial fulfilment of the requirements for the degree of Master of Engineering, Department of Civil and Environmental Engineering. The University of Auckland, 2009.

Avery Gottfried - ME thesis 2009

Embed Size (px)

Citation preview

Page 1: Avery Gottfried - ME thesis 2009

ENHANCING IN SITU PAH BIODEGRADATION

The Effects of Amendments on Bench-Scale Bioremediation

Systems

by

Avery Gottfried

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

Master of Engineering,

Department of Civil and Environmental Engineering.

The University of Auckland, 2009.

Page 2: Avery Gottfried - ME thesis 2009

ii

ABSTRACT

Current research in the field of bioremediation is uncovering a growing number of

microorganisms with the metabolic potential to degrade PAHs in soil and water. In situ

bioremediation is based on encouraging the growth of microorganisms, either indigenous or

introduced, to improve the degradation of contaminants without excavating or transporting

the soil. The majority of PAHs sorb strongly to soil organic matter posing a complex barrier

to biodegradation. Biosurfactants can increase soil-sorbed PAHs desorption, solubilisation,

and dissolution into the aqueous phase, which increases the bioavailability of PAHs for

microbial metabolism. In this study, biosurfactants, carbon sources, metabolic pathway

inducers, and oxygen were tested as stimulators of microorganism degradation.

Phenanthrene served as a model PAH and Pseudomonas Putida ATCC 17484 was used as the

naphthalene and phenanthrene degrading microorganism for the liquid solutions, soil

slurries and column systems used in this investigation. Bench-scale trials demonstrated that

the addition of rhamnolipid biosurfactant increases the apparent aqueous solubility of

phenanthrene, and overall degradation by at least 20% when combined with salicylate and

glucose. In soil slurries containing salicylate, the effects of biosurfactant additions were

negligible as there was greater than 90% removal, regardless of the biosurfactant

concentration. An in situ enhancement strategy for phenanthrene degradation could focus

on providing additional carbon substrates to induce metabolic pathway catabolic enzyme

production, if degradation pathway intermediates are known. The results of experiments

performed in this study provide further evidence that future studies should focus on

enhancing the metabolic processes responsible for successful in situ bioremediation.

Page 3: Avery Gottfried - ME thesis 2009

iii

ACKNOWLEDGEMENTS

I would like to thank the Commonwealth Scholarship and Fellowship Plan which awarded

me the funding to come to New Zealand and pursue my research interests in this unique

part of the World. While here I was lucky enough to be part of a diverse Environmental

Engineering research group headed by Dr. Naresh Singhal, who also provided supervision for

this work. Special thanks to Abel Francis, our laboratory technician, for providing training,

maintaining equipment, and helping with experimental setup; Dr. Simon Swift in the

Molecular Medicine and Pathology Department for guidance, laboratory training, and

laboratory facilities to develop the biological aspects of this research; and Roy Elliot for

sharing his microbiology expertise in designing experiments and assisting with many

laboratory techniques over the research period.

Page 4: Avery Gottfried - ME thesis 2009

iv

TABLE OF CONTENTS

ABSTRACT ........................................................................................................................ II

ACKNOWLEDGEMENTS .................................................................................................... III

LIST OF FIGURES .............................................................................................................. VI

LIST OF TABLES .............................................................................................................. VIII

ABBREVIATIONS ........................................................................................................... IX

CHAPTER 1 INTRODUCTION AND THESIS OBJECTIVES ........................................................ 1

1.1 INTRODUCTION .................................................................................................. 1

1.2 THESIS OBJECTIVES ............................................................................................. 3

1.3 ORGANIZATION OF THE THESIS ........................................................................... 4

CHAPTER 2 LITERATURE REVIEW ....................................................................................... 5

2.1 CONTAMINATED SITES IN NEW ZEALAND ............................................................ 5

2.2 BIOREMEDIATION ............................................................................................... 6

2.3 POLYCYCLIC AROMATIC HYDROCARBONS ........................................................... 8

2.4 IN SITU BIOREMEDIATION ................................................................................. 11

2.5 BACTERIAL DEGRADATION ................................................................................ 13

2.5.1 PHENANTHRENE METABOLISM ......................................................................... 13

2.6 LIMITING FACTORS AND STRATEGIES FOR BIOREMEDIATION ............................ 18

2.6.1 BIOAVAILABILITY OF PAH CONTAMINANTS ....................................................... 20

2.6.2 BIOSTIMULATION AND BIOAUGMENTATION .................................................... 21

2.6.3 SURFACTANTS AND BIOSURFACTANTS .............................................................. 22

2.6.4 SORPTION AND DESORPTION ............................................................................ 27

2.6.5 IONIC STRENGTH AND PH EFFECTS ON BIOSURFACTANTS ............................... 33

2.6.6 BIOSURFACTANT MICROBUBBLE DISPERSIONS ................................................. 34

2.6.7 METABOLIC PATHWAY INDUCERS ..................................................................... 37

2.7 DETERMINING TRANSPORT PARAMATERS ........................................................ 41

CHAPTER 3 MATERIALS AND METHODS .......................................................................... 45

3.1 LABORATORY SUPPLIES AND MICROOGANIM STORAGE .................................... 45

3.1.1 CHEMICALS ......................................................................................................... 45

3.1.2 BIOSURFACTANT ................................................................................................ 45

3.1.3 MICROORGANISMS ............................................................................................ 45

3.1.4 MEDIA AND NUTRIENT SUPPLY ......................................................................... 46

3.2 CELL CULTURING ............................................................................................... 48

3.2.1 AGAR PLATES ..................................................................................................... 48

3.2.2 INOCULANT PREPARATION AND HARVESTING .................................................. 48

3.2.3 PLATE COUNTS ................................................................................................... 49

3.2.4 OPTICAL DENSITY ............................................................................................... 50

Page 5: Avery Gottfried - ME thesis 2009

v

3.3 SOIL METHODS ................................................................................................. 51

3.3.1 SOIL PROPERTIES ................................................................................................ 51

3.3.2 SOIL CONTAMINATION ...................................................................................... 52

3.3.3 CONTAMINANT EXTRACTION ............................................................................ 53

3.4 EXPERIMENTAL PROCEDURES ........................................................................... 55

3.4.1 OBJECTIVE 1: LIQUID MEDIUM TESTS ................................................................ 55

3.4.1.1 Inoculant culture growth ............................................................................ 55

3.4.1.2 Phenanthrene dissolution ........................................................................... 56

3.4.1.3 Degradation trials ....................................................................................... 56

3.4.2 OBJECTIVE 2: SOIL SLURRY TESTS ...................................................................... 58

3.4.2.1 Phenanthrene Desorption in the Presence of Surfactants .......................... 58

3.4.2.2 Degradation Trials ...................................................................................... 59

3.4.3 OBJECTIVE 3: COLUMN TESTS ............................................................................ 61

3.4.3.1 Experimental Apparatus ............................................................................. 61

3.4.3.2 Pressure Measurement ............................................................................... 62

3.4.3.3 Micro-foam Generation and Stability ......................................................... 62

3.4.3.4 Column Packing and Unpacking ................................................................. 63

3.4.3.5 Experimental Operating Conditions............................................................ 64

3.5 ANALITICAL METHODS ...................................................................................... 66

3.5.1 PAH DETECTION ................................................................................................. 66

3.5.2 BIOSURFACTANT DETECTION ............................................................................ 67

3.5.3 CHLORIDE ANION ............................................................................................... 67

CHAPTER 4 RESULTS AND DISCUSSION ............................................................................ 68

4.1 OBJECTIVE 1: LIQUID CULTURES ........................................................................ 68

4.1.1 BACTERIA GROWTH ON VARIOUS SUBSTRATES ................................................ 68

4.1.2 EFFECT OF INOCULANT ACCLIMATIZATION AND PRE-TREATMENT .................. 70

4.1.3 BIOSURFACTANT SOLUBILITY ENHANCEMENT .................................................. 73

4.1.4 PHENANTHRENE DEGRADATION ....................................................................... 75

4.2 OBJECTIVE 2: SOIL SLURRIES .............................................................................. 78

4.2.1 PHENANTHRENE DISTRIBUTION IN THE PRESENCE OF BIOSURFACTANTS ....... 78

4.2.2 SOIL DEGRADATION ........................................................................................... 83

4.3 OBJECTIVE 3: COLUMN TESTS ............................................................................ 91

4.3.1 TRACER AND BIOSURFACTANT BREAKTHROUGH CURVE FITTING .................... 91

4.3.1.1 Tracer Breakthrough Curves ....................................................................... 91

4.3.1.2 Biosurfactant breakthrough ....................................................................... 92

4.3.2 PRESSURE DROP ASSOCIATED WITH MICROBUBBLE DISPERSION PUMPING ... 97

4.3.3 BIODEGRADING TRIALS IN CONTINUOUS FLOW SYSTEMS .............................. 100

CHAPTER 5 CONCLUSIONS AND RECOMENDATIONS ...................................................... 106

5.1 RECOMMENDATION FOR FUTURE WORK ........................................................ 109

CHAPTER 6 WORKS CITED ............................................................................................. 112

Page 6: Avery Gottfried - ME thesis 2009

vi

LIST OF FIGURES

Figure 2.1 Factors that influence biodegradation systems in bioremediation. Adapted

from Singh and Ward (2004) ............................................................................................ 7

Figure 2.2 The diagrammatic structures of 16 U.S EPA priory pollutant PAH compounds.

Modified from Rogers et al. (2002) .................................................................................. 9

Figure 2.3 Bay, K and L regions of PAHs involved in the formation of metabolically active

epoxides. Adapted from Chauhan et al. (2008) ............................................................. 14

Figure 2.4 Illustration of common steps in the upper pathway for aerobic metabolism of

phenanthrene (Samanta, Chakraborti et al. 1999; Chauhan, Fazlurrahman et al.

2008; Jun 2008) .............................................................................................................. 16

Figure 2.5 Illustration of common steps in aerobic metabolism of naphthalene and one of

the lower pathways for aerobic metabolism of phenanthrene (Samanta,

Chakraborti et al. 1999; Chauhan, Fazlurrahman et al. 2008; Jun 2008) ....................... 17

Figure 2.6 Schematic diagram of physical changes that occur due to surfactant addition

above the CMC. Adapted from (Mulligan, Yong et al. 2001) ......................................... 24

Figure 2.7 Examples of typical glycolipid biosurfactants produced by Pseudomonas

aeruginosa ...................................................................................................................... 25

Figure 2.8 Schematic diagram of abiotic processes in a soil-aqueous-surfactant system

containing a non-ionic surfactant, phenanthrene and soil organic matter. Adapted

from Edwards et al. (1994). ............................................................................................ 30

Figure 2.9 Plasmid-encoded naphthalene (upper pathway) and salicylate (lower pathway)

degradation genes of NAH7 catabolic plasmid for Pseudomonas sp. Genes nahA-D

encode the upper pathway operon which encodes enzymes for the degradation of

naphthalene to salicylate and genes nahG-M encode the lower pathway operon,

where salicylate is further degraded to pyruvate and acetylaldehyde.The product

from nahR (a trans-acting positive control regulator) is the positive regulator for

both operons and is induced by salicylate. The location of each respective operon

promoter is shown and locations of genes encoding the naphthalene dioxygenase

complex are indicated. ................................................................................................... 40

Figure 3.1 Schematic diagram for serial dilutions to determine cfu/mL of solution .............. 50

Figure 3.2 Particle size distribution.......................................................................................... 52

Figure 3.3 Soil column setup for uplflow pumping experiments ............................................ 61

Figure 4.1 Partial growth curves for P.Putida until early stationary phase in four growth

medias of glucose (2g/L); naphthalene (0.5g/L); salicylate (0.5g/L); and naphthalene

(0.5g/L) + biosurfactant (1g/L) ....................................................................................... 69

Figure 4.2 Naphthalene degradation and cell growth in liquid cultures containing

different bacteria inoculant seeds which were pre-grown in seven different

solutions (s1 BHB+naphthalene+glucose grown for 1 week; s2 BHB+glucose; s3

BHB+naphthalene+glucose; s4 BHB+salicylic acid+glucose; s5 LB; s6

LB+naphthalene; s7 LB+salicylic acid; s2 - s7 grown overnight approximately 20

hours growth) ................................................................................................................. 71

Figure 4.3 Phenanthrene solubility enhancement as a function of biosurfactant

concentration. The equation refers to the fit of data above the CMC and � � �� ��� ................................................................................................................................... 73

Page 7: Avery Gottfried - ME thesis 2009

vii

Figure 4.4 Phenanthrene degradation in liquid cultures containing BHB and/or

biosurfactant (1000mg/L), salicylate (100mg/L), and glucose (100mg/L). Data

presented is the average of triplicate measurements taken at 22 and 46 hours after

inoculation. ..................................................................................................................... 75

Figure 4.5 Phenanthrene desorption from soil in the presence of biosurfactant.

Desorption partitioning coefficient Kd calculated from the linear regression

trendline for each series of data. ................................................................................... 78

Figure 4.6 Phenanthrene desorption from contaminated soil (50, 100, 250, 500 mg/kg)

into aqueous solution in the presence of biosurfactant over a 48 hour period. ........... 80

Figure 4.7 Total phenanthrene concentration in solution and suspended/dissolved

organic matter in soil slurries containing BHB and/or biosurfactant (0.25, 1, 5 g/L),

salicylate (100mg/L), and glucose (100mg/L) over a 10 day period .............................. 84

Figure 4.8 Total remaining phenanthrene in soil slurries after 10 days of bioremediation,

results presented as phenanthrene remaining in mg/kg of dry soil. ............................. 86

Figure 4.9 Live cell counts (cfu/mL) taken from soil slurry solution over 10 days. Results

presented are averages from duplicate or triplicate plate counts. ............................... 88

Figure 4.10 Observed breakthrough curves and fitted breakthrough curve models using

CXTFIT inverse parameter estimation for (a) chloride with v = 1.54cm/h (b) chloride

with v = 77.87cm/h and (c) biosurfactant with v = 1.54 cm/h ....................................... 93

Figure 4.11 Microbubble dispersion breakthrough curve with a conservative tracer in the

liquid fraction. ................................................................................................................ 96

Figure 4.12 Pressure distribution across the length of the soil column during biosurfactant

(1 g/L) solution pumping. Data presented corresponds to depth in the column with

the highest pressure at the inlet, and the lowest pressure at -17cm assuming the

outlet is the datum. ........................................................................................................ 98

Figure 4.13 Pressure distribution across the length of the soil column during biosurfactant

microfoam (1 g/L) pumping. Data presented corresponds to depth in the column

with the highest pressure at the inlet, and the lowest pressure at -17cm assuming

the outlet is the datum. .................................................................................................. 99

Figure 4.14 Pressure distribution across the length of the soil column during biosurfactant

microfoam ( 5g/L) pumping. Data presented corresponds to depth in the column

with the highest pressure at the inlet, and the lowest pressure at -17cm assuming

the outlet is the datum. .................................................................................................. 99

Figure 4.15 Trial 1 phenanthrene distribution in soil column after 10 days continuous

upflow at 0.2 mL/min , phenanthrene in column effluent over 10 days. Column 1

influent solution biosurfactant 1g/L + salicylate 100 mg/L; Column 2 influent

solution biosurfactant 1 g/L. Soil distribution assuming effluent (0cm) is the top of

the column and influent (-37 cm) is the bottom of the column. ................................. 100

Figure 4.16 Trial 2 phenanthrene distribution in soil column after 10 days continuous

upflow with BHB broth at 0.5mL/min; phenanthrene in column effluent over 10

days. Column 1 influent pulse solution biosurfactant microfoam 1 g/L + salicylate

100mg/L; Column 2 influent pulse solution biosurfactant 1 g/L. Soil distribution

assuming effluent (0cm) is the top of the column and influent (-37 cm) is the

bottom of the column. ................................................................................................. 101

Page 8: Avery Gottfried - ME thesis 2009

viii

LIST OF TABLES

Table 2.1 Chemical characteristics of EPA PAH priority pollutants. Adapted from data

sources (Aitken, Stringfellow et al. 1998; World Health Organization 1998; Rogers,

Ong et al. 2002) .............................................................................................................. 10

Table 2.2 Advantages and disadvantages for selecting in situ bioremediation. Adapted

from The Interstate Technology & Regulatory Council (2005) ...................................... 12

Table 3.1 BHB marine salts broth approximate formula per litre of prepared media ............ 47

Table 3.2 Soil properties .......................................................................................................... 51

Table 3.3 Soil slurry media constituents .................................................................................. 60

Table 3.4 Column trial experimental design ............................................................................ 65

Table 4.1 Rate of phenanthrene degradation in liquid cultures expressed as mg of

phenanthrene degraded / hour ..................................................................................... 77

Table 4.2 Calculated phenanthrene soil partitioning coefficient Kd, and phenanthrene

partitioning onto soil sorbed surfactant coefficient Ks .................................................. 81

Table 4.3 Total percentage removal of phenanthrene due to soil flushing and

biodegradation in soil column tests after 10 days continuous flow. ........................... 103

Page 9: Avery Gottfried - ME thesis 2009

ix

ABBREVIATIONS

PAHs polycyclic aromatic hydrocarbon/s SOM soil organic matter IRZs in situ reactive zones CMC critical micelle concentration HOCs hydrophobic organic compounds BHB Bushnell-Hass marine salts Broth LB Lysogeny Broth PV pore volume HPLC high performance liquid chromatography TOC total organic carbon OD600 optical density at 600nm

Page 10: Avery Gottfried - ME thesis 2009

1

CHAPTER 1 INTRODUCTION AND THESIS OBJECTIVES

1.1 INTRODUCTION

Interest in the bioremediation processes of soil contaminated with Polycyclic Aromatic

Hydrocarbons (PAH) has been growing over the past two decades. This interest stems from

the identification of microbes with the ability to degrade toxic xenobiotic compounds in soil

and water. Although bioremediation primarily relies on the catalytic roles of soil

microorganisms to break down contaminants into innocuous by-products, the

understanding of the microbial communities’ operation and behaviour in complex soil

systems remains limited. Bioremediation occurs in the natural environment where most

organisms are uncharacterized, and each site is unique in terms of its soil, microbes, and

contamination. These variable site characteristics create numerous challenges to

understanding the interactions taking place which actually contribute to the desired

decrease in harmful contamination. The biodegradation process is mostly treated as a

unknown ‘black box’ process where soil amendments are made and desired contaminant

removal is achieved without fully understanding the microbial processes that were

enhanced to bring about contaminant mineralization (Singh and Ward 2004). Recent

research has focused on the biochemical and physiological aspects of the bioremediation

process with an emphasis on determining key parameters that make the process more

efficient and reliable (Samanta, Singh et al. 2002). This includes improving the

bioavailability of the contaminants and understanding the metabolic pathways and the

enzymatic reactions that are used in contaminant breakdown, with the goal of identifying

the rate-limiting steps. Ultimately obtaining this knowledge will enable scientist to engineer

better bioremediation processes. Biotechnology and advanced molecular techniques are

Page 11: Avery Gottfried - ME thesis 2009

2

now providing researchers with the tools to advance understanding in these areas. There is

tremendous potential for engineered bioremediation to make microorganisms more

effective and efficient in removing contaminants, accelerating the remediation process.

It is difficult to replicate the complexity of real PAH contaminated sites in constructed lab

scale systems. However, there is certainly a need to determine optimal treatment

conditions and degradative capabilities in a single bacteria strain in order to unravel the

underlying interactions. Simple systems, where most variables can be controlled and

monitored, show insight into the microbial response to specific variables that are altered,

and will offer advances for understanding the underlying complexities of in situ

bioremediation (Pieper and Reineke 2000). This research was carried out to determine why

specific soil amendments, including the addition of biosurfactants, have been shown to

increase (or decrease) overall contaminant degradation. Results from these trials provide

further evidence to the processes responsible for successful in situ bioremediation

treatment and contribute to the understanding and capabilities of these processes in PAH

field contaminated sites.

Page 12: Avery Gottfried - ME thesis 2009

3

1.2 THESIS OBJECTIVES

The focus of this research was to enhance the degradation of phenanthrene in soil by the

microbe Pseudomonas putida ATCC 17484 (P.putida) with amendments that included co-

substrates, electron acceptors, and metabolic pathway inducers to the system.

Amendments were designed to increase contaminant bioavailability, enhance microbial

degrading activity, and increase the amount of contaminant degradation. To fully

understand the interactions between biodegradation, amendments, and soil, each process

was isolated and independently evaluated in order to focus on specific characteristics that

enhanced in situ biodegradation. The experiments were designed in stages to achieve each

specific objective:

OBJECTIVE 1: To study the effect of co-substrates, metabolic pathway inducers, and

inoculant pre-treatment on the degradation of phenanthrene and naphthalene by P.putida

in liquid cultures.

Task A: studied growth characteristics of P.putida in various substrates

Task B: determined changes in phenanthrene solubility in the presence of rhamnolipid

biosurfactant

Task C: determined contaminant degradation rates in liquid cultures with added

biosurfactant, glucose, and salicylate

OBJECTIVE 2: To evaluate the effects and monitor the changes in the degradation of

phenanthrene by P.putida due to various soil amendments to contaminated soil slurry.

Task A: determined soil characteristics and the contaminant desorption characteristics

in the presence of rhamnolipid biosurfactant

Task B: determined contaminant degradation rates in soil slurries with amendments

Page 13: Avery Gottfried - ME thesis 2009

4

OBJECTIVE 3: To design a continuous-flow bench-scale micro-environment to model in situ

remediation and observe the degradation of phenanthrene by P.putida in a saturated

contaminated soil. This system was used to study the effects and flow of microfoam through

the system, and analyze the substrate transport parameters in a soil column.

Task A: determined transport parameters in soil columns using non reactive tracers

Task B: determined microfoam characteristics and evaluate pressure build up in the

soil during the injection of microfoam for various flow rates and microfoam

qualities

Task C: evaluated the efficiency of microfoam and various soil amendments on the

overall removal of phenanthrene from contaminated soil

1.3 ORGANIZATION OF THE THESIS

This thesis consists of five chapters:

Chapter One gives a brief introduction to bioremediation and gives an overview of research

objectives and thesis setup. Chapter Two defines the nature of the problem and discusses

the most important areas of investigation. It also provides an overview of the topic and

highlights key knowledge gained in similar areas, which are relevant to the work presented

in this thesis. Chapter Three presents all of the methods that were used in this study.

Experimental designs are presented in detail for each of the experiments that were

necessary to complete the three main objectives. Chapter Four summarizes the results

obtained and offers an interpretation and discussion of them. Chapter Five forms the

conclusion, and provides recommendations for future research.

Page 14: Avery Gottfried - ME thesis 2009

5

CHAPTER 2 LITERATURE REVIEW

2.1 CONTAMINATED SITES IN NEW ZEALAND

As modern economies move to enhance environmental protection, more effective testing

methods, increased legislation, and stricter monitoring guidelines have been developed. The

result of this has been the location and acknowledgement of numerous sites of soil

contamination (Doyle, Muckian et al. 2008). Many strategies have been proposed, including

physical, chemical, and biological methods to restore contaminated soil sites. Polycyclic

Aromatic Hydrocarbons (PAHs) are present in many contaminated soil sites, stemming

primarily from the use of oil and petroleum-derivatives; including potentially hazardous,

carcinogenic, and toxic hydrocarbons. Sites with high PAH concentration can act as sources

because contaminants mobilize and leach offsite posing extra risks to groundwater, soil

fertility, and living systems (Singh and Ward 2004). PAHs significantly accumulate in surface

and subsurface soils and an increased concentration can result in a highly toxic

environmental site, necessitating cleanup. Depending on the site location and the level of

groundwater contamination such contamination can pose serious human health risks. In

New Zealand, surface and subsurface soil contamination has been linked to historical land

uses which include agricultural and horticultural activities, gas works, landfills, petrol

station, dry cleaners, sheep dips, and timber treatment sites. The New Zealand Ministry of

the Environment (2007) reports 1,238 contaminated sites resulting from industrial inputs

deemed as ‘Hazardous Activities and Industries List’ (HAIL). However, this number could be

over 50,000 when sites not currently on the HAIL list are considered. These include a large

number of urban sites contaminated unknowingly by fill materials—and such sites are

slowly being discovered (Auckland City Council 2007).

Page 15: Avery Gottfried - ME thesis 2009

6

2.2 BIOREMEDIATION

The term bioremediation can be applied to any biological process that uses enzymes in

microorganisms, fungi, or green plants to break down undesired contaminants and

contribute to the restoration of the environment to its original condition. Biodegradation is

defined as the breakdown of organic compounds to less complex metabolites, or the

complete breakdown through mineralization into the inorganic minerals H20, C02, or CH4.

Understanding the bioremediation process requires the examination and interpretation of

both biochemical and physiological aspects. Knowledge of these processes will allow key

parameters to be manipulated and bioremediation optimized (Singh and Ward 2004).

Generally, the environmental conditions in which microbial processes are occurring must be

altered to encourage the desired outcomes. With bioremediation, a variety of factors

(Figure 2.1) can influence microbial growth and bioactivity which ultimately increase the

microbial, physiological and biochemical activity and enhance the biodegradation of

contaminants. Even if these factors are optimized, PAH degradation can remain slow unless

the mass transfer rates and the bioavailability of the PAHs for microbial metabolism are

increased (Cerniglia 1993). To effectively accelerate the removal of PAHs from contaminated

soils a greater understanding of not only the physical processes involved but the

physiology, biochemistry, molecular genetics and microbial ecology of the degrading strains

of microorganisms is required (Chauhan, Fazlurrahman et al. 2008). Consideration of the

biotic factors such as the production of toxic or dead-end metabolites, metabolic repression,

presence of preferred substrates and lack of co-metabolites or inducer substrates are also

important in optimizing the overall efficiency of the bioremediation process (Chauhan,

Fazlurrahman et al. 2008).

Page 16: Avery Gottfried - ME thesis 2009

Figure 2.1 Factors that influence

7

Factors that influence biodegradation systems in bioremediation. Adapted from

Singh and Ward (2004)

in bioremediation. Adapted from

Page 17: Avery Gottfried - ME thesis 2009

8

2.3 POLYCYCLIC AROMATIC HYDROCARBONS

PAHs are identified as a class of chemicals with two or more fused aromatic rings containing

solely carbon and hydrogen (Figure 2.2). They are formed as products in the incomplete

combustion or pyrolysis of fossil fuels and organic matter (Harvey 1991). PAHs are natural

components of fossil fuels such as petroleum or coal, but leakage and accidental spill of

these products can cause accumulation in the environment (Doyle, Muckian et al. 2008).

The World Health Organization (1998) reports that the PAHs contained in various

environmental wastes, including coal combustion residues, motor vehicle exhaust, used

motor lubricating oil, and tobacco smoke “are mainly responsible for their carcinogenic

potential.” The largest and most important PAH contaminated sites occur near large

industrial sources where individual PAH levels of up to 1g/kg of soil have been found. These

originate from concentrated emissions of combustion residues and storage and handling of

coal, coke, fly ash or liquid petroleum reserves. Soils present around crude-oil refineries,

fuel storage depots, petrol stations, gas works, landfills, incinerators and wood preservation

facilities are the most common areas where significant PAH accumulation has been

detected. PAH sources such as automobile exhaust have been shown to cause

contamination next to busy roadways in the range of 2-5mg/kg of soil, whereas background

levels of PAHs are 5-100µg/kg soil deriving from natural sources of atmospheric deposition

such as forest fires and volcanic eruptions (World Health Organization 1998).

The environmental fate of PAH contaminants is governed by the number of aromatic rings

present, and the nature of the linkage between the rings (Doyle, Muckian et al. 2008). PAHs

with low molecular weight are typically defined as those containing up to three aromatic

rings and tend to be more soluble and volatile. High molecular weight PAHs are generally

Page 18: Avery Gottfried - ME thesis 2009

9

defined as those containing four or more aromatic rings and tend to be less soluble, less

volatile and have a tendency to accumulate in the environment as they sorb strongly to soil

organic matter (SOM) (World Health Organization 1998).

Figure 2.2 The diagrammatic structures of 16 U.S EPA priory pollutant PAH compounds.

Modified from Rogers et al. (2002)

The Environmental Protection Agency (EPA) Priority Pollutant List of 126 pollutants includes

16 PAH compounds (Figure 2.2), the important details of their chemical properties include

thier aqueous solubility (Table 2.1). The contaminants included on this list are regulated,

and the EPA has developed analytical testing methods for accurate detection in the

environment (U.S Environmental Protection Agency 2008). This frequently referenced list

originated from the 1972 Clean Water Act and the 1977 Clean Water Act Amendment, and

Page 19: Avery Gottfried - ME thesis 2009

10

the only modifications to this list were the removal of 3 compounds from the list in 1981

showing the long recognition of PAHs as toxic compounds (Hendricks 2006; U.S

Environmental Protection Agency 2008).

Table 2.1 Chemical characteristics of EPA PAH priority pollutants. Adapted from data

sources (Aitken, Stringfellow et al. 1998; World Health Organization 1998; Rogers, Ong et

al. 2002)

U.S EPA PAH priority

compound

Chemical

Formula

Aqueous

solubility

Csat

(mg/L)

Molecular

weight

(g/mol)

n-Octanol

water

partition

coefficient

(log Kow)

Organic

carbon

partition

coefficient

(log Koc)

Naphthalene C10H8 31 128.17 3.37 3.11

Acenaphthylene C12H8 3.4 152.19 4.0 3.4

Acenaphthene C12H10 3.8 154.21 3.94 3.65

Fluorene C13H10 1.9 166.22 4.18 3.86

Phenanthrene C14H10 1.1 178.23 4.57 4.15

Anthracene C14H10 0.045 178.23 4.54 4.15

Fluoranthene C16H10 0.26 202.25 5.22 4.58

Pyrene C16H10 0.132 202.25 5.18 4.58

Chrysene C18H12 0.002 228.29 5.65 5.3

Benz[a]anthracene C18H12 0.011 228.29 5.91 6.14

Benzo[b]fluoranthene C20H12 0.0015 252.31 5.8 5.74

Benzo[k]fluoranthene C20H12 0.0008 252.31 6.0 5.74

Benzo[a]pyrene C20H12 0.0038 252.31 6.04 6.74

Dibenz[a,h]anthracene C22H14 0.0006 278.35 6.75 6.52

Indeno[1,2,3-c,d]pyrene C22H12 0.062 276.33 7.66 6.2

Benzo[g,h,i]perylene C22H12 0.00026 276.33 6.5 6.2

Page 20: Avery Gottfried - ME thesis 2009

11

2.4 IN SITU BIOREMEDIATION

Conventional remediation techniques, including ex situ treatment, and removing the soil for

treatment and disposal at another, safer location are difficult to apply to many PAH

contaminated sites. The slow mobilization of contaminants, and leaching into ground water

tables have lead to situations which contaminated ground water supplies and contaminant

plumes migrate below city infrastructure and developed areas, make ex situ treatment and

soil removal impossible. Sites of PAH contaminations are varied, including: developed urban

and industrial areas, parks and natural environments, and spread over large geographical

areas. Many contaminated sites were discovered after industrial activities ceased, as PAHs

are essentially recalcitrant and persistent (Gómez, Alcántara et al. In Press). The increasing

number and widespread distribution of contaminated sites has encouraged the

development of in situ technologies such as heat based injection, air sparging, soil vapour

extraction, soil washing or flushing, bioremediation, enhanced bioremediation, microbial

filters, and others (Warith, Fernandes et al. 1999). In situ technologies offer an advantage

over ex situ techniques as they are designed to be implemented in place, without

transporting or disturbing the soil (Error! Reference source not found.). Furthermore, in situ

technologies have now been proven to be much cheaper alternatives to traditional

methods, saving time and resulting in a less invasive remediation design that can

complement the natural attenuation process (Suthersan and Payne 2005). In situ

bioremediation is a technology based on stimulating the growth of indigenous or introduced

microorganisms to improve the degradation of contaminants without excavating or

transporting the soil to other locations for treatment. In situ bioremediation can include

augmenting the natural microbial population with the addition of specific microbes that can

metabolize and grow on specific compounds.

Page 21: Avery Gottfried - ME thesis 2009

12

Table 2.2 Advantages and disadvantages for selecting in situ bioremediation. Adapted

from The Interstate Technology & Regulatory Council (2005)

In situ reactive zones (IRZs) are designed to manipulate oxidation and reduction reactions,

and other biogeochemical processes to affect the mobility, transport, and fate of inorganic

and organic contaminants in the subsurface. Successful designs of IRZs need to account for

all the variables presented in Figure 2.1 to optimize the reactions required for the

biodegradation of target contaminants. Each contaminated site has different characteristics

and naturally variable conditions that need to be taken into account in an engineered

remediation solution to create an effective microbial reactive zone.

ADVANTAGES DISADVANTAGES It is usually less expensive than other

remediation options.

Complete contaminant destruction is not

achieved in some cases, leaving the risk of a

residual toxic intermediate. It is almost always faster than baseline

pump-and-treat.

Some contaminants are resistant to

biodegradation.

It may be possible to completely destroy

the contaminant, leaving only harmless

metabolic by-products (no ex situ waste

created).

Some contaminants (or their co-

contaminants) are toxic to the

microorganisms and prevent complete

metabolism and site restoration.

It can be designed with minimal

disturbance to the site and facility

operations, and also can be incorporated

into ongoing site development activities.

Biodegradation of organic species can

sometimes cause mobilization of naturally

occurring toxic inorganic species such as

manganese or arsenic.

It is not limited to a fixed area, typical of

chemical flushing or heating technologies,

because it can move with the contaminant

plume.

Alteration of groundwater redox conditions or

substrate supply can reduce the down

gradient effectiveness of natural

bioattenuation processes. It can treat both dissolved and sorbed

contaminants.

Uncontrolled proliferation of the

microorganism may clog the subsurface.

The processes usually use reagents that are

easily accepted by regulators and the

public.

The hydrogeology of the site may not be

conducive to enhancing the microbial

population.

Page 22: Avery Gottfried - ME thesis 2009

13

2.5 BACTERIAL DEGRADATION

The biochemical pathways and enzymes responsible for the initial transformation stages are

usually specific to particular contaminants, but bacteria have the capacity to evolve new

catabolic pathways when exposed for long periods of time to specific contaminants. Due to

the complex mixture of low and high molecular weight PAHs present in some contaminated

sites, there tends to be incomplete bioremediation of higher weight PAHs even if aggressive

approaches are used to enhance the process (Singh and Ward 2004). Due to the

recalcitrance of high molecular weight it is difficult for any single microbial organism to use

them as sole energy and carbon growth, but they are more likely to be oxidized in a series of

steps by consortia of microbes (Perry 1979). Cerniglia (1993) stated that “a better

understanding of the metabolism, enzyme mechanisms, and genetics of polycyclic aromatic

hydrocarbon-degrading microorganisms is critical for the optimization of these

bioremediation processes” and this fact holds true 15 years later and remains an effective

motivation for research today.

2.5.1 PHENANTHRENE METABOLISM

Phenanthrene is a three-ring PAH with low aqueous solubility and is commonly used in

laboratory research as an ideal PAH contaminant for the study of various aspects of

microbial metabolism and physiology (Woo, Lee et al. 2004; Labana, Manisha et al. 2007).

Phenanthrene is the smallest PAH which has both a low aqueous solubility and contains an

“L-region” “bay-region,” and a “K-region” which is common in many higher ringed PAHs

(Figure 2.3). Bay-regions are locations where there is a terminal ring on one side of the bay

region (the terminal ring is also termed the A region), K-regions are areas of high electron

density in all resonance structures and L-regions are sites between two ring fusion points

Page 23: Avery Gottfried - ME thesis 2009

14

(Yan 1985). The understanding of phenanthrene metabolism can be correlated to studies

on higher-ringed PAHs such as benzo[a]pyrene, benzo[a]anthrancene and chrysene. The

metabolism of bay-region and K-region is believed to be important in understanding the

degradation of both higher and lower ringed PAH compounds and phenanthrene serves as

the example (Xiang, Xian-min et al. 2006). The Bay-region dihydrodiol epoxides are believed

to be the main carcinogenic species and in benzo[a]pyrene these metabolites are cytotoxic,

cause DNA strand breaks and are also mutagenic (World Health Organization 1998). The

Bay, K, and L PAH regions (Figure 2.3) are involved in the formation of metabolically active

and highly reactive epoxides. PAH epoxides arise via metabolism of the parent PAH and

occur whenever oxygen atoms are added across double bonds, a process that can be

catalyzed by the action of enzymes or by an uncatalyzed oxidation process (Josephy and

Mannervik 2006).

There are several bacteria strains that are capable of degrading phenanthrene aerobically

and the more commonly identified strains are Pseudomonas sp, Rhodococcus sp.,

Mycobacterium flavescens, Mycobacterium sp., Flavobacterium sp., and Beijerinckia sp.,

which are capable of using phenanthrene as the sole carbon source and growth substrate

(Cerniglia 1993; Samanta, Chakraborti et al. 1999; Chauhan, Fazlurrahman et al. 2008 ).

Figure 2.3 Bay, K and L regions of PAHs involved in the formation of

metabolically active epoxides. Adapted from Chauhan et al. (2008)

Page 24: Avery Gottfried - ME thesis 2009

15

Phenanthrene has two potential degradation pathways that are established based on the

bacteria present. These pathways take advantage of the biologically and chemically active

bay and K-region epoxides, which can be formed metabolically by enzymes present in

phenanthrene degrading bacteria (Samanta, Chakraborti et al. 1999). Both pathways share

the same common upper route (Figure 2.4) and are initiated by the double hydroxylation of

a phenanthrene ring by a dioxygenase enzyme to yield cis-3,4-dihydroxy-3,4

dihydrophenanthrene, which then undergoes enzymatic dehydrogenation to 3,4-

dihydroxyphenanthrene. From here the diol is cleaved and metabolized, and 1-hydroxy-2-

naphthoic acid remains and is degraded by one of the two routes termed the lower

pathways (Prabhu and Phale 2003).

The lower pathways consist of two separate routes for degradation depending on the

enzymes that are present in the organisms. In route one (Figure 2.5) 1-hydroxy-2-naphthoic

acid is degraded via the naphthalene pathway to salicylate and then further metabolized via

the formation of catechol or gentisic acid, while route two uses the phthalate pathway. Both

naphthalene and phenanthrene share a common upper metabolic pathway and organisms

that degrade phenanthrene via route one have the ability to degrade naphthalene,

salicylate and catechol (Kiyohara, Torigoe et al. 1994; Samanta, Chakraborti et al. 1999;

Prabhu and Phale 2003). Both oxygen and water are consumed during metabolism and H+

ions are produced, which can affect the pH of the environment if enough degradation

activity is occurring. Understanding the metabolic processes that are involved in the

degradation of phenanthrene are important when determining how additions such as

oxygen or salicylate will influence microbial activity, or determining why changes in pH or a

build up of intermediate metabolites is occurring.

Page 25: Avery Gottfried - ME thesis 2009

16

Fig

ure

2.4

Ill

ust

rati

on

of

com

mo

n s

tep

s in

th

e u

pp

er

pa

thw

ay

fo

r a

ero

bic

me

tab

oli

sm o

f p

he

na

nth

ren

e

(Sa

ma

nta

, C

ha

kra

bo

rti

et

al.

19

99

; C

ha

uh

an

, F

azl

urr

ah

ma

n e

t a

l. 2

00

8;

Jun

20

08

)

Page 26: Avery Gottfried - ME thesis 2009

17

Fig

ure

2.5

Ill

ust

rati

on

of

com

mo

n s

tep

s in

ae

rob

ic m

eta

bo

lism

of

na

ph

tha

len

e a

nd

on

e o

f th

e l

ow

er

pa

thw

ay

s fo

r a

ero

bic

me

tab

oli

sm o

f p

he

na

nth

ren

e (

Sa

ma

nta

, C

ha

kra

bo

rti

et

al.

19

99

; C

ha

uh

an

, F

azl

urr

ah

ma

n e

t a

l. 2

00

8;

Jun

20

08

)

Page 27: Avery Gottfried - ME thesis 2009

18

2.6 LIMITING FACTORS AND STRATEGIES FOR BIOREMEDIATION

It is commonly assumed that bacteria can only access PAHs in the aqueous phase, and the

relatively low bioavailability of PAHs in this phase limits their consumption by the microbial

biomass. As this is a limiting factor, increasing the bioavailability of PAHs in the aqueous

phase by increasing mass transfer rates of PAHs from the soil into solution is essential in

furthering research in this area (Wick, Colangelo et al. 2001). Recent review papers

(Cerniglia 1993; Samanta, Chakraborti et al. 1999; Chauhan, Fazlurrahman et al. 2008 )

conclude that virtually all of the PAHs of concern are biodegradable, and that organisms

capable of degrading PAHs are ubiquitous in the natural environment. PAHs also have

strong hydrophobicity and associate with nonaqueous phases in soil and natural organic

matter where they are not bioavailable, meaning they are in a location where they are not

able to be adsorbed and metabolised by microorganisms. Because of these tendencies,

bioremediation of PAHs in the environment is usually incomplete, even when soil

amendments attempt to enhance the system (Singh and Ward 2004).

Looking specifically at the microbial kinetics, there are several methods possible to enhance

the rate of biodegradation of a PAH. The simplest way to determine what factors influence

this rate in a simple solution is to look at Monod growth kinetics (Equation 2.1). The Monod

growth kinetics take the same form as Michaelis-Menten kinetics with the assumption that

a certain number of new cells grow per unit mass of chemical transformed (Hemond and

Fechner-Levy 2000). This equation is the most commonly used in modelling growth kinetics

associated with PAH degradation.

Page 28: Avery Gottfried - ME thesis 2009

19

Monod Growth Kinetics

� � �� · �� ��

Equation 2.1

Where:

• u : specific growth rate [T-1

]

• umax : maximum specific growth rate [T-1

]

• C : concentration of dissolved chemical [M/L3]

• Ks : half-saturation constant [M/L3]

The biodegradation rate depends on µmax, Ks and substrate concentration C. Therefore,

increasing µmax, decreasing Ks, increasing microbial cell density, or increasing contaminant

concentration will be sufficient strategies to enhance the biodegradation process. At low

contaminant concentrations, the rate at which bacteria can degrade the substrate can also

depend on the specific affinity for the substrate (Johnsen, Wick et al. 2005). Specific affinity

refers to the ratio of the maximal rate of substrate uptake and the half saturation constant,

and high affinities lead towards efficient contaminant removal at low concentrations due to

steeper concentration gradients and higher transfer rates between the substrate and the

cell (Johnsen, Wick et al. 2005). Enhancements in microbial growth kinetics can only occur if

no chemicals other than the contaminants are limiting the microbial community.

Specifically, oxygen and mineral nutrients must be in excess. Understanding the basic

parameters that influence the degradation rate of contaminants, highlights the importance

of increasing the bioavailability of contaminants for effective in situ bioremediation.

Page 29: Avery Gottfried - ME thesis 2009

20

2.6.1 BIOAVAILABILITY OF PAH CONTAMINANTS

Significant PAH accumulation in the environment occurs in subsurface organic soil matter

due to the hydrophobicity and low aqueous solubility of PAHs. The majority of PAHs are

difficult to remove because they sorb strongly to soil organic matter. Long term

contamination of soil, which is commonly referred to as aging or weathering, is the result of

chemical oxidation reactions and slow chemical diffusion into small pores, both of which

decrease PAH bioavailability over time (Singh and Ward 2004). The degradation process

therefore involves the transfer of contaminants from the soil to the enzymes in the

microorganisms which begin the mineralization of the contaminant (Noordman 1999).

Contaminant characteristics such as molecular structure, solubility, and the octanal/water

partitioning coefficient (Kow) are relevant for substance sorption in or onto soil and can be

used to indicate the availability of the contaminant to bacteria in the soil.

It is the influence of the dissolution and desorption process of PAHs in soil that are often

cited as the rate limiting step in the degradation process. The slow transport of PAHs from

the soil matrix to bacteria is the slowest process and limits degradation (Mulder, Breure et

al. 2001; Prabhu and Phale 2003; Johnsen, Wick et al. 2005; Doyle, Muckian et al. 2008).

However, contaminant bioavailability can be species-specific, with different bacteria strains

able to access different contaminant pools in the soil-water system. Understanding the

interactions amongst bacteria will also provide further opportunities for enhanced

degradation of bioavailable contaminants (Dean, Jin et al. 2001). For instance, some

organisms have the ability to affect sorption kinetics on their own, through the production

of surface active agents termed biosurfactants. These can increase the apparent solubility of

PAHs in the aqueous phase and concomitantly increase the concentration gradient, allowing

Page 30: Avery Gottfried - ME thesis 2009

21

improved mass transfer of contaminants from the soil to the aqueous phase (Pignatello and

Donald 1999). Dean et al. (2001) demonstrated that some of the sorbed phase

phenanthrene was bioavailable to certain Pseudomonas bacteria, and called into question

the frequently used assumption that only bulk aqueous phase contaminant is available for

degradation. Woo et al. (2001) included a term for sorbed phase biodegradation of

phenanthrene when modelling the process to account for the rapid degradation that

occurred in soil slurry tests. Kwok and Loh (2003) also proposed that bacteria which have

attached themselves directly to soil particles can utilize the nutrients sorbed at that

location. Several techniques have been developed to effectively enhance the bioavailability

of contaminants.

2.6.2 BIOSTIMULATION AND BIOAUGMENTATION

Enhanced biodegradation is usually accomplished through biostimulation and

bioaugmentation. Biostimulation refers to the modification of the environment via the

addition of oxygen, nutrients, other electron donors or acceptors, and surfactants. These

additions stimulate the existing bacteria and increase the number or rate at which the

organisms are degrading a contaminant. Biostimulation relies on making the natural

environment more favourable to the metabolic capacities of the indigenous microbial

populations, whereas bioaugmentation describes the addition of adapted microorganisms

to the environment that are capable of degrading contaminants that are present.

Depending on the characteristics of the contaminated site, either biostimulation or

bioaugmentation may be needed to achieve the desired outcomes. Ruberto (2006) found

that a combination of both techniques using fish meal for nutrient supply and surfactant

Brij700 with bioaugmentation using a psychrotolerant PAH degrading bacterial consortium

Page 31: Avery Gottfried - ME thesis 2009

22

caused significant removal (46.6%) of phenanthrene whereas when each technique was

applied separately, insignificant reduction was observed (Ruberto, Vazquez et al. 2006). It is

commonly reported that either the availability of electron acceptors, or nutrient limitations,

are the cause of slow biodegradation processes at contaminated sites (Institute for Ecology

of Industrial Areas 1999). Laboratory studies often report high rates of biodegradation

compared to results actually achieved in the field with similar soil and bacteria types, which

can be due to the optimization of many variables such as temperature, mixing, nutrient

balances and nutrient delivery. These variable are sometimes impossible to replicate in the

field (Institute for Ecology of Industrial Areas 1999).

2.6.3 SURFACTANTS AND BIOSURFACTANTS

Surfactants are used to describe surface-active agents that lower the surface tension of a

liquid (Riser-Roberts 1998). Surfactants have both a hydrophilic group and a hydrophobic

group and can be described as either anionic or cationic depending on whether they release

an anion or a cation when dissociating in water. They are termed non-ionic if no net charge

is dissociated. Therefore an anionic surfactant has an anionic hydrophilic group at its head,

whereas a non-ionic surfactant has no net charge groups at its head. Anionic and non-ionic

surfactants tend to be the best solubilizers and are relatively non-toxic compared to cationic

surfactants (Oostrom, Dane et al. 2006). Jin (2007) ranked the toxicity of the studied

surfactants to bacterial activity in soil and determined the order of toxicity towards bacteria

as follows: non-ionic surfactants (Tween 80, Brij30, 10LE and Brij35) < anionic surfactants

(LAS) < cationic surfactants (TDTMA) (Jin, Jiang et al. 2007).

Page 32: Avery Gottfried - ME thesis 2009

23

Suitable co-solvents or surfactants must be selected according to solution chemistry, proven

ability to solubilise PAH compounds, and compatibility with the remediation technique. In

addition, they must not be toxic or a threat to human health or the environment (Gómez,

Alcántara et al. In Press). The presence of surfactants in the bulk phase causes an increase in

the free energy of the system. In order to lower the free energy, the surfactant molecules or

monomers are concentrated at the surface and interface, and the surface tension is lowered

increasing the solubility of hydrophobic contaminants (Myers 1988). The surface tension will

decrease to a given value, known as the critical micelle concentration (CMC) beyond which

point it will remain constant (Figure 2.6). Once the concentration of surfactants is above the

CMC, the surfactants begin to aggregate to form micelles, vesicles, and lamellae. Surfactant

micelles increase the apparent aqueous solubility of hydrophobic particles by reducing the

interfacial tension between the oil phase and the aqueous phase. In contaminated systems

this results in PAHs partitioning within the hydrophobic micellar core of the micelles. This

creates higher apparent aqueous solubility as PAHs are dissolved both in aqueous solution

and inside surfactant micells which are present in the bulk aqueous phase (Error! Reference

source not found.) (Noordman, Ji et al. 1998; Cameotra and Bollag 2003; Makkar and

Rockne 2003).

Surfactants can also be produced by bacteria or yeasts from growth on various substrates

including sugars, oils, hydrocarbons and agricultural wastes. These are termed

biosurfactants (Lin 1996). In terms of surface activity, heat and pH stability, many

biosurfactants are comparable to synthetic surfactants (Lin 1996). Biosurfactants are

receiving increasing attention as they have lower toxicity and higher biodegradability

compared to their chemical counterparts (Rosenberg and Ron 1999). Specifically,

Page 33: Avery Gottfried - ME thesis 2009

rhamnolipid biosurfactants produced by

extensively as they have excellent emulsifying power with a variety of hydrocarbon

vegetable oils (Wang, Fang et al. 2007)

glycolipidic surface-active molecules that are produced in mixtures of one or two rhamnoses

attached to β–hydroxyalkanoic acid

resulting in lengths of 8, 10, 12 and 14 carbons

Fang et al. 2007). The in situ

them potentially more cost effective while also using natural resources instead of chemical

inputs.

Figure 2.6 Schematic diagram of physical changes that occur due to surfactant addition

above the CMC. Adapted from

Surfactant

monomer

24

rhamnolipid biosurfactants produced by Pseudomonas aeruginosa have been studied

excellent emulsifying power with a variety of hydrocarbon

(Wang, Fang et al. 2007). Rhamnolipid (Figure 2.7) is the name given to the

active molecules that are produced in mixtures of one or two rhamnoses

hydroxyalkanoic acid. The length of the fatty acid chains can va

12 and 14 carbons (Soberón-Chávez, Lépine et al. 2005; Wang,

in situ production of biosurfactants at contaminated sites renders

ntially more cost effective while also using natural resources instead of chemical

Schematic diagram of physical changes that occur due to surfactant addition

above the CMC. Adapted from (Mulligan, Yong et al. 2001)

Surfactant

Micelle

have been studied

excellent emulsifying power with a variety of hydrocarbons and

the name given to the

active molecules that are produced in mixtures of one or two rhamnoses

an vary significantly,

Chávez, Lépine et al. 2005; Wang,

production of biosurfactants at contaminated sites renders

ntially more cost effective while also using natural resources instead of chemical

Schematic diagram of physical changes that occur due to surfactant addition

g et al. 2001)

Page 34: Avery Gottfried - ME thesis 2009

25

Figure 2.7 Examples of typical glycolipid biosurfactants produced by Pseudomonas

aeruginosa

Research into the addition of surfactants and biosurfactants have produced mixed results,

from greatly enhanced rates of PAH degradation to the inhibition of PAH degradation

(Pieper and Reineke 2000; Makkar and Rockne 2003; Avramova, Sotirova et al. 2008). There

are several hypotheses explaining the mixed results. Beneficial results may be due to the

facilitation of bioremediation through increases in desorption, solubilisation, and dissolution

of PAHs from soil sorbed or solid phase contaminant into the aqueous phase, which results

in increased bioavailability of PAHs for microbial metabolism (Mulligan, Yong et al. 2001;

Makkar and Rockne 2003; Shin, Kim et al. 2004; Avramova, Sotirova et al. 2008). Negative

result led to an assortment of conclusions including:

• the preferential use of surfactants as a growth substrate by degrading

microorganisms;

• the toxicity of the applied surfactants preventing increased microbial growth;

Page 35: Avery Gottfried - ME thesis 2009

26

• the toxicity of the PAHs resulting from the increased bioavailability that is caused by

the surfactant solubilization of PAHs;

• the reduction of PAH bioavailability due to the uptake into surfactant micelle which

then could not be available for bacteria;

• the sorption of surfactant into the soil blocking access to PAHs that could have been

further absorbed into the soil or causing PAH sorption into soil sorbed surfactants

(Garcia-Junco, Gomez-Lahoz et al. 2003; Shin, Kim et al. 2004; Avramova, Sotirova et

al. 2008).

An example of these mixed results was meaningfully demonstrated by Allen et al. (1999)

with the use of titron X-100 with Pseudomonas sp. strain 9816/11 and Sphingomonas

yanoikuyae B8/36. Triton X-100 increased the rate of oxidation of phenanthrene with strain

9816/11. Conversely, the surfactant inhibited the biotransformation of both naphthalene

and phenanthrene with strain B8/36 under the same conditions (Allen, Boyd et al. 1999).

These observations show an important knowledge gap in how surfactants truly alter the

biodegradation process and interact with bacteria. Considering that a non-ionic surfactant

could have contrasting effects on the ability to degrade PAHs by different bacteria, there is a

requirement for additional research relating to surfactants, including all stages of soil-water-

surfactant-bacteria interactions.

There is a recurring assumption that the remediation of PAHs in soil or soil-water systems

depends strongly on the desorption rates of the PAHs from the soil into the aqueous phase

(Jin, Jiang et al. 2007). It is assumed that once PAHs are in the bulk aqueous phase, it is

possible to use engineering treatment steps to enhance the remediation process and create

Page 36: Avery Gottfried - ME thesis 2009

27

an effective bioremediation strategy. However, there are an increasing number of studies

that have demonstrated that bacteria can attach to soil particles and use the nutrients

sorbed to the soil surface (Dean, Jin et al. 2001; Wick, Colangelo et al. 2001). This could

explain why the addition of surfactants to some systems does not predictably enhance

contaminant biodegradation. As the natural role of biosurfactant is to increase the

bioavailability of contaminants by decreasing surface tension, there can be a reduction in

direct adhesion of bacteria to the desired contaminants of interest due to the decrease in

surface tension (Pieper and Reineke 2000).

The mixed effects of surfactant on biodegradation show the complex interactions between

the PAH, surfactant, microorganism, soil, and water in the environment. Due to variable that

are important in the bioremediation process all researchers have to provide caveats in the

conclusions section to isolate results to the unique system of bacteria, soil type,

contaminant, and test conditions that was studied.

2.6.4 SORPTION AND DESORPTION

The partitioning and transport processes (sorption, desorption, and dissolution) between

the soil and water phases of both contaminants and surfactants affect the overall

degradation of contaminants (Schlebaum, Schraa et al. 1999; Kraaij, Ciarelli et al. 2001;

Mulder, Breure et al. 2001; Zhou and Zhu 2005; Zhou and Zhu 2007; Wang and Keller 2008;

Zhu and Zhou 2008; Laha, Tansel et al. In Press). Soil organic matter and natural organic

matter is not homogeneous and PAHs strongly absorb to soot carbon, and more slowly

partition into humic matter (Jonsson, Persson et al. 2007). As PAHs adsorb onto the surface

Page 37: Avery Gottfried - ME thesis 2009

28

of soil organic mater they slowly begin to penetrate further into cavities and diffuse into the

organic fraction over time. Landrum et al (1992) observed a continuous increase in the

partition coefficient of phenanthrene and pyrene into soil over a period of six months, after

in-lab contamination of the soil. The length of this process makes it impractical for the

determination of single sorption or desorption coefficients to model the process over the

long term. Schlebaum et al (1999) successfully modelled the sorption of hydrophobic

organic compounds (HOCs) from the soil matrix with a kinetic model using two separate

compartments. A Freundlich isotherm represented high affinity sites, and a linear sorption

isotherm and first order kinetics represented low affinity sites. Even if the amount of

organic matter is low, PAHs can still become trapped in pores and voids and these variables

will affect the efficiency and success of any remediation process.

It is not just the average aqueous concentration of the target contaminant that determines

its availability. The rate of mass transfer to microbial cells relative to the intrinsic substrate

utilization capacity of the microbial cells must also be considered because it determines the

bioavailability of the contaminant (Wick, Colangelo et al. 2001). As a result, limited

bioavailability occurs when the environment is unable to deliver the substrate at the rate

consumable by the microbial biomass. The biodegradation rate in the subsurface is often

reported as first-order even when total contaminant concentrations are high. Wick et al.

(2001) provided an explanation for these observations by considering that the mass

transport processes are slow for hydrophobic organic soil pollutants which cause the same

degradation rates to be obtained even when the substrate concentration has changed.

Page 38: Avery Gottfried - ME thesis 2009

29

When enhancing bioremediation, it is important to consider the effect surfactants can have

on the desorption and dissolution of contaminants from the soil. When surfactants exceed

their CMC , it is well established that there is an increase in desorption of PAHs from the soil

(Noordman 1999). However, when surfactants adsorb to soil they increase the overall

organic content of the soil and provide additional sorption capacity. This can enhance

sorption of hydrophobic compounds onto soil sorbed surfactants. This influences the

amount of PAHs present in the aqueous phase, accessible for biodegradation (Edwards,

Adeel et al. 1994). Conversely, the micelles present in the bulk aqueous phase can greatly

enhance the solubilisation of the PAHs, causing increased desorption from soil. The

efficiency of surfactants at enhancing PAH desorption shows a strong dependence on the

soil composition, surfactant structure and concentration, and PAH properties as concluded

by Zhou and Zhu (2005).

The process of surfactant adsorption to the soil has been described as a three stage process

by Torrens et al.(1998). The first stage is controlled by electrostatic attraction between

surfactants and the soil surface. As the surfactant concentration increases, there is a

tendency for self-association of surfactant ions due to the electrostatic and hydrophobic

forces. This is analogous to the micelle formation but it leads to the formation of

hemimicelles which is the second stage. This stage is more rapid than the first stage, and

results in neutralization of the particle surface, causing the sorption process to slow. After

the second stage, micelle formation begins, which results in the reversal of the surface

charge. This greatly reduces surfactant sorption due to charge repulsion. The third stage is a

plateau region, and additional surfactant will be present in solution (Torrens, Herman et al.

1998). Figure 2.8 encapsulates the interactions that are believed to occur between the soil-

Page 39: Avery Gottfried - ME thesis 2009

30

water-surfactant systems. The stages of sorption also result in the wetting of soil grains

which enables the washing out of the hydrophobic substances from the soil pores. For most

hydrophobic contaminants they can be assumed to be un-wetted with water, and

surfactants increase the wettablity of the hydrophobic surfaces through attachment and

sorption to the soil surface (Pastewski, Hallmann et al. 2006).

Figure 2.8 Schematic diagram of abiotic processes in a soil-aqueous-surfactant system

containing a non-ionic surfactant, phenanthrene and soil organic matter. Adapted from

Edwards et al. (1994).

Page 40: Avery Gottfried - ME thesis 2009

31

The distribution of contaminants between the soil fraction and the aqueous phase is

generally described by the partitioning coefficient Kd. Kd refers to the ratio of the

concentration of contaminant in the soil fraction to the concentration in the aqueous phase.

In the simplest form, the equation Cs = KdCw where Cs is the concentration of the

contaminant sorbed by the soil fraction and Cw in the concentration of the contaminant in

the aqueous phase respectively. Kmc is another commonly used partitioning coefficient,

which defines the amount distributed between the aqueous phase and the surfactant

micelle phase. The total amount is commonly referred to as the apparent aqueous solubility

as there is more contaminant in the aqueous phase, although it is located inside the

surfactant micelle. There are many different theoretical models that are used to determine

the partitioning coefficient Kd, taking into account surfactant adsorption modelled by the

Langmuir isotherm, Kow, and the fraction of organic carbon, and PAH sorption (Huang and

Cha 2001). The following equation appears to be the most commonly used to describe PAH

partitioning within a soil-water-surfactant system (Zhu, Chen et al. 2003; Zhou and Zhu

2007; Wang and Keller 2008; Zhu and Zhou 2008).

��� � �� ����

� ��� ��� ��� ��� Equation 2.2

Where:

• ��� : ratio of sorbed PAH to mobile PAH in the aqueous solution (L/kg);

• Kd : PAH sorption coefficient with the soil in the absence of surfactant (L/kg);

• Qs : quantity of surfactant sorbed to the soil;

• Ks : solute distribution coefficient with the soil-sorbed surfactant (L/kg);

• Xmm and Xmc : surfactant monomer and micellar concentration in water (g/L);

• Kmm and Kmx : PAH partitioning coefficients with the surfactant monomer and

micellar phases (L/kg).

Page 41: Avery Gottfried - ME thesis 2009

32

The overall factors that effect ��� are the partitioning of PAH to soil due to the presence of

sorbed surfactants (terms in the numerator in equation 2.2), and decreased PAH

partitioning to soil by the enhanced aqueous solubility of the PAH in the presence of

surfactant monomers and micelles (denominator in equation 2.2).

Depending on the quantity of surfactant added to the system, the majority may be in the

soil sorbed-phase (Laha, Tansel et al. In Press). The result of this is increased partitioning of

PAHs onto soil until the solubilisation by micellar phase surfactant is at a high enough

concentration to compete with the increased PAH sorption on the surfactant sorbed soil

(Laha, Tansel et al. In Press). However, the cation exchange capacity of the soil can

significantly affect the sorption of surfactants (Ks), and well as the ionic strength or pH of the

system.

Page 42: Avery Gottfried - ME thesis 2009

33

2.6.5 IONIC STRENGTH AND PH EFFECTS ON BIOSURFACTANTS

Anionic surfactants are strongly affected by the presence of electrolytes in solution as they

can influence the solubilization capacity, cause precipitation of the surfactant from the

aqueous phase, and increase the adsorption to subsurface porous media (Stellner and

Scamehorn 1989; Jafvert and Heath 1991; Guiyun, Brusseau et al. 1998). Torrens et al (1998)

saw 67% rhamnolipid sorption to soil at low K+ concentrations (10mM) but this increased to

98% in the presence of 20mM K+ in solution. The ionic strength and presence of cations in

solution has been shown to further enhance the solubility of hydrophobic organic

contaminants in rhamnolipid solution. Guivun (1998) reported that both Na+ and Mg

2+

enhanced the solubility of PAHs as there was an increase in the interior volume of

rhamnolipid micelles in the presence of cations, and Mg2+

, being a divalent cation, had a

stronger affection on reducing the repulsion forces between anionic head groups (Guiyun,

Brusseau et al. 1998). However, Ca2+

had little affect on solubility, due to competing effects

between rhamnolipid precipitation and enhanced contaminant solubility. The presence of

cations also reduced the interfacial tension between rhamnolipid solutions and hexadecane

from 2.2 to 0.89 dyn cm-1

(Guiyun, Brusseau et al. 1998). A decrease in pH from 7 to 6 was

seen to have the same qualitative effect to the interfacial tension as the increase in Na+

concentration. The carboxyl group in the rhamnolipid head group has a pKa of 5.6, causing it

to become more protonated as the pH decreases, thereby reducing repulsion between the

head groups. A similar effect was seen by Shin (2004) as the apparent solubility of

phenanthrene was 3.8 times greater at a pH of 5.5 when compared with a pH of 7 in the

presence of 240 mg/L rhamnolipid. In another study, more rhamnolipid molecules were lost

by sorption to sand particles at a pH 4 than at both higher and lower pH values, explaining

why a dramatic decrease in apparent aqueous solubility of phenanthrene was seen at that

Page 43: Avery Gottfried - ME thesis 2009

34

pH (Shin, Kim et al. 2008). These findings are particularly important in soil remediation as

subsurface matrix solutions contain electrolytes such as Ca2+

, Mg2+

, Na+, K

+, and Al

3+ which

can have affect the surfactant performance (Guiyun, Brusseau et al. 1998).

2.6.6 BIOSURFACTANT MICROBUBBLE DISPERSIONS

Microbubble dispersions, also known as colloidal gas aphrons (CGA) or microfoam, are a

series of micro-bubbles that were first investigated by Sebba (1971). Microfoam displays

colloidal properties because of its micron-sized bubbles (typically 0.7-100µm) and its unique

bubble structure which consists of multiple layers of surfactant monomers surrounding the

surface of the microbubble. In contrast, standard foam consists of just one layer of

surfactant monomers (Jauregi and Varley 1999; Wan, Veerapaneni et al. 2001; Larmignat,

Vanderpool et al. 2008). Microbubble dispersions can flow like water, and can be pumped

easily without collapse (Jauregi and Varley 1999). Surfactant microfoam technology is a

relatively new approach for enhancing in situ bioremediation, showing promising

advantages over air sparging or surfactant solution application. Foam can flow in a plug flow

manner, delivering oxygen or air uniformly (Wang and Mulligan 2004). Microbubble

dispersion flow is also capable of overcoming heterogeneity in porous media, enhancing

bacterial transport, and delivering oxygen and nutrients to the subsurface (Wan,

Veerapaneni et al. 2001; Choi, Park et al. 2008; Park, Choi et al. In Press). Foam and

microfoam technology is designed either to remove contaminants and/or act

simultaneously as an augmentation for existing technologies such as pump-and-treat and

bioremediation. It is designed to enhance the process and improve removal efficiencies and

cost effectiveness (Wang and Mulligan 2004). Foam stability reflects the ability of the

Page 44: Avery Gottfried - ME thesis 2009

35

suspension to resist bubble collapse, and is typically measured as the time required for half

of the foam to collapse. The half-life for microfoam can range from minutes to days,

depending on the generation method, surfactant, and additions such as nutrients, bacteria

or soil particles.

Microbubble dispersions can facilitate mobilisation and transport of contaminants trapped

in porous media, and can take less pore volumes to achieve high contaminant removal when

compared to surfactant solutions (Wang and Mulligan 2004). Couto et al (In Press) saw 96%

removal in sandy soils using microfoam in soil flushing to remove diesel oil, versus 88%

removal with regular foam and 35% removal with surfactant solution. Park et al. (In Press)

saw a 2.2-fold increase in phenanthrene degradation when 3 pore volumes of microbubbles

were injected instead of 1 pore volume. There are several studies demonstrating the ability

of conventional foam to enhanced remediation of PAH contaminated soils, and the

beneficial transport mechanisms of foam (Chowdiah, Misra et al. 1998; Rothmel, Peters et

al. 1998). Microfoam appears to have an added advantage over conventional foam as

dispersion can be generated that contain less gas (60-70% versus up to 99% with

conventional foam) in smaller sized bubbles, making them easier to pump through the

subsurface (Roy, Kommalapati et al. 1995; Jauregi and Varley 1999).

Microbubble injection systems have been shown to be efficient oxygen delivery systems in

pilot scale tests that used microbubble generators that were encapsulated in pressurized

chambers that contained oxygen and biosurfactant solution. Leigh et al. (1997)

demonstrated that microbubbles generated using this method persistent in the subsurface

for longer periods of time and have different migration characteristics compared to air

Page 45: Avery Gottfried - ME thesis 2009

36

bubbles injected in by typical air sparging. Using this generation method and a mixture of

anionic and non ionic surfactants, Wan et al. (2001) was able to generate microbubbles that

were still present in solution up to six weeks after generation.

Subsurface foam and microfoam flow is typically accompanied by a pressure drop due to the

flow characteristics. Higher-viscosity foams flow forward and fill up larger channels and pore

spaces. When the pressure drop builds up in the channel, the foam flows into less accessible

spill areas. This pressure dependent “clogging” process means that channelling, or poor

sweep, should not occur with the microbubble scouring as compared with surfactant

flushing (Riser-Roberts 1998). However, applications could be limited by the pressure drop

required to pump microbubbles into soil with low permeability (Riser-Roberts 1998; Choi,

Park et al. 2008; Park, Choi et al. In Press).

Page 46: Avery Gottfried - ME thesis 2009

37

2.6.7 METABOLIC PATHWAY INDUCERS

Another biostimulation strategy that can enhance the intrinsic biodegradation rate of target

compounds is the addition of one or more known pathway intermediate catabolite. These

are usually produced by the bacteria when mineralizing a contaminant and they stimulate

growth, enzymatic expression, and ultimately increase the biodegradation of PAHs

(Ogunseitan and Olson 1993; Cho, Seung et al. 2006). This process is defined as co-

metabolism, where bacteria may co-utilize various substrates that compete with the

structurally similar primary substrate for the enzyme’s active sites (Mohan, Kisa et al. 2006).

The introduction of carbon sources that are metabolic pathway inducers into the soil can

enhance in situ bioremediation by stimulating the growth of specific indigenous micrograms

that are capable of degrading organic contaminants. Unfortunately, additional carbon

sources can also be used preferentially by soil bacteria causing diauxic growth which can

have a negative effect on the degradation process (Lee, Park et al. 2003).

A number of studies have used salicylate as a pathway inducer to enhance initial rates of

naphthalene and phenanthrene removal (Chen and Aitken 1999; Lee, Park et al. 2003; Woo,

Jeon et al. 2004; Lee, Lee et al. 2005; Powell, Singleton et al. 2008; Basu, Das et al. In Press).

Salicylate is the third intermediate formed in the degradation of naphthalene (Figure 2.5)

and it is also an intermediate formed in the degradation of phenanthrene for bacteria that

degrade phenanthrene via the naphthalene pathway. Most information about PAH

metabolism has been derived from the study of naphthalene catabolic plasmids in

Pseudomonas putida G7 (Yen and Serdar 1988). In the plasmid there are genes which

encode the pathway for naphthalene degradation (Figure 2.9) In the first operon, there are

genes which encode the pathway for conversion of naphthalene to salicylate, and in the

Page 47: Avery Gottfried - ME thesis 2009

38

second operon are the genes which code for the conversion of salicylate via catechol meta-

cleavage to acetaldehyde and pyruvate (Eaton and Chapman 1992; Platt, Shingler et al.

1995). The regulatory mechanism for both operons is encoded in a third operon which acts

as the regulatory protein and positively regulates the two operons by the increased

presence of salicylate (Schell and Wender 1986; Atlas and Philip 2005). The principle

mechanism for the aerobic bacterial metabolism of naphthalene is via the oxidative action

of the naphthalene dioxygenase enzyme; that introduces molecular oxygen into the

aromatic ring. The naphthalene (upper pathway) and salicylate (lower pathway) degradation

genes located in the NAH7 catabolic plasmid from Pseudomonas sp. are regulated by

salicylate induction to both operons (Figure 2.9). Chen and Aitken (1999) showed that

salicylate greatly enhanced removal of fluoranthene, pyrene, benz[a]anthracene, chrysene,

and benzo[a]pyrene, all of which are high molecular weight PAHs which the strain

Pseudomonas saccharophila P15 could not use as a sole carbon for growth. This showed

that high-molecular weight PAH metabolism by this organism is induced by salicylate. Lee et

al. (2005) saw phenanthrene degradation rates 3.5-fold higher with Burkholderia cepacia

PM07 compared to the rates achieved without salicylate addition in aqueous solutions. They

also saw a decrease in phenanthrene removal with the addition of glucose (Lee, Lee et al.

2005). Basu et al. (In Press) determined Pseudomonas Putida CSV86 preferentially utilized

aromatics over glucose and co-metabolized them with organic acids, indicating that

intermediate metabolites enhance the mineralization rate of PAHs more effectively than

additional carbon sources (Basu, Das et al. In Press).

In other studies Cho et al. (2006) saw up to 12 times increase in the degradation of target

chemicals per equivalent cell mass with the addition of various intermediate metabolites

Page 48: Avery Gottfried - ME thesis 2009

39

into solution. In this experiment all phenanthrene was soluble due to the addition of 1%wt

Triton X-100. Woo et al. (2004) saw up to a 3-fold increase in phenanthrene degradation

using salicylate in soil water systems, however addition of triton X-100 saw inhibitory effects

towards total phenanthrene mineralization. Other substances such as 1-hudroxy-2-

naphthoate, catechol, and pyruvate have also shown their potential as effective pathway

inducers to enhance in situ bioremediation (Cho, Seung et al. 2006; Basu, Das et al. In Press).

Chemotaxis is another strategy that can be used to enhance the degradation of

contaminants in the environment. Chemotaxis is “a complex process [in] which bacterial

cells detect temporal changes in the concentrations of specific chemicals, respond

behaviourally to theses changes and then adapt to the new concentration of the chemical

stimuli” (Samanta, Singh et al. 2002). It is not clear if it is the metabolism of the substrate or

if it is the binding of the substrates to the chemoreceptors that is the crucial inducer of

chemotaxic behaviour. The NAH7 plasmid in Pseudomonas putida, which encodes the

enzymes for the degradation of naphthalene and salicylate, also encodes the chemotaxis

towards these compounds (Samanta, Singh et al. 2002). This chemotaxis in Pseudomonas

putida was found to be homologous to chemotaxis, flagellar and mobility genes from other

known E.coli bacteria. The ability to foster chemotaxis phenomenon via metabolic

influences could be important to enhance in situ bioremediation.

Page 49: Avery Gottfried - ME thesis 2009

40

Fig

ure

2.9

Pla

smid

-en

cod

ed

na

ph

tha

len

e (

up

pe

r p

ath

wa

y)

an

d s

ali

cyla

te (

low

er

pa

thw

ay

) d

eg

rad

ati

on

ge

ne

s o

f N

AH

7 c

ata

bo

lic

pla

smid

for

Pse

ud

om

on

as

sp.

Ge

ne

s n

ah

A-D

en

cod

e t

he

up

pe

r p

ath

wa

y o

pe

ron

wh

ich

en

cod

es

en

zym

es

for

the

de

gra

da

tio

n o

f n

ap

hth

ale

ne

to

sali

cyla

te a

nd

ge

ne

s n

ah

G-M

en

cod

e t

he

lo

we

r p

ath

wa

y o

pe

ron

, w

he

re s

ali

cyla

te i

s fu

rth

er

de

gra

de

d t

o p

yru

va

te a

nd

ace

tyla

lde

hy

de

.Th

e p

rod

uct

fro

m n

ah

R (

a t

ran

s-a

ctin

g p

osi

tiv

e c

on

tro

l re

gu

lato

r) i

s th

e p

osi

tiv

e r

eg

ula

tor

for

bo

th o

pe

ron

s a

nd

is

ind

uce

d b

y s

ali

cyla

te.

Th

e l

oca

tio

n o

f e

ach

re

spe

ctiv

e o

pe

ron

pro

mo

ter

is s

ho

wn

an

d l

oca

tio

ns

of

ge

ne

s e

nco

din

g t

he

na

ph

tha

len

e

dio

xy

ge

na

se c

om

ple

x a

re i

nd

ica

ted

.

Page 50: Avery Gottfried - ME thesis 2009

41

2.7 DETERMINING TRANSPORT PARAMATERS

Laboratory tracer experiments are useful for flow characteristics in soil. The fundamental

mass balance of the system uses:

Inputs + Production – Outputs – Losses = Accumulation Equation 2.3

All contaminant transport and biodegradations models use this fundamental principal to

derive equations when approximating parameters such as convection and dispersion. These

assumptions allow the derivation of the partial differential equation referred to as the

convection-dispersion (Equation 2.4). The CDE is a mathematical model used in

quantitatively simulating the transport of solutes in porous media. The CDE is derived by

assuming the change in chemical flux into and out of a control volume is controlled by an

advection component (which is controlled by the velocity of the chemical), and dispersion

(which can be through of as mimicking diffusion in the sense that the dispersive flux appears

to be driven by concentration gradients) (Toride, Leij et al. 1995). The CDE for one-

dimensional transport of reactive solutes subject to adsorption, first-order degradation, and

zero-order production, in homogenous soil is given as: (Toride, Leij et al. 1993):

� ���� � � ����

��� � ����� � μ� " #$% Equation 2.4

The initial boundary conditions used to solve this equation assume that there is a fixed

known concentration of solute added to the system. This is expressed as the following:

• &'&( $∞, +% � 0 (exit condition where the concentration = 0)

• C(x,t) = 0 for x = 0

• C(x,0) = Ci (where the concentration of influent tracer is constant)

• C(0,t) = -./0 0 / 12 3 24

2 5 24 (tracer on/off after time t)

Page 51: Avery Gottfried - ME thesis 2009

42

Where:

• C : dissolved aqueous chemical concentration;

• x and t : dimensionless space and time variable respectively;

• 6 � 1 " 89 :;< (Where => is the bulk density; �� is a partitining coefficient; and n is

the porosity);

• D : dispersion coefficient;

• ? : pore water velocity;

• @ : first order decay rate constant;

• A : zero order production rate constant.

When decay appears in such a system it can be due to strong sorption, chemical and

biological activity, or other physicochemical interactions of the solutes in the porous media.

The inclusion of decay represents the change in dispersive flux out and assumes that decay

affects the mass inside the controlled volume. Another common form of equation 2.4 uses

a term called the Peclet number (P) which is a dimensionless number relating the amount of

advection to dispersion (P=v/D).

The transport of solutes in soil and groundwater systems includes a large number of

complicated physical, chemical, and microbiological processes (Toride, Leij et al. 1993).

Variations to the standard CDE presented in equation 2.3 have been added to account for

the simultaneous effect of sorption (including zero and first order), convective transport,

molecular diffusion, hydrodynamic dispersion, zero-order production, and first order decay

(Chen, Wang et al. 2006). Equilibrium transport processes refer to exchange reactions that

are perceived as instantaneous and are commonly described by equilibrium isotherms

including linear, Freundlich or Langmuir type. However, these equilibrium models appear to

fail in situations where chemical transport processes are not at equilibrium (Nielsen, Van

Genuchten et al. 1986). This has led to the development of non-equilibrium transport

Page 52: Avery Gottfried - ME thesis 2009

43

models that incorporate first order reactions and various chemical, kinetic and diffusion

limited rate laws to describe the non equilibrium transport. A familiar chemical non-

equilibrium model (Equations 2.5 & 2.6) includes one-site and two-site sorption. This means

sorption onto one site can be considered to be instantaneous (equilibrium) while sorption

onto the second site will be rate limited by first order kinetics (non-equilibrium) (Toride, Leij

et al. 1993). The two-site model presented is also used for physical non-equilibrium and is

termed a two-region (dual-porosity) type formulation which contains two distinct liquid

regions, one being mobile (flowing) and the other being immobile and the rate constants

refer to the mass transfer between the two regions which is modelled as a first-order

process.

B� ����� � �

C ������� � � ����

� � � D$E� � E�% � �� " #�$% Equation 2.5

$� � B%� ����� � D$E� � E�% � �� " #�$% Equation 2.6

Where:

• subscripts 1 and 2 refer to equilibrium and non equilibrium sites respectively;

• β : partitioning coefficient of adsorption sites that equilibrates with instantaneous

and kinetic adsorption sites or mobile and immobile liquid phase;

• ω : dimensionless mass transfer coefficient (Toride, Leij et al. 1995).

The two-site equilibrium and non-equilibrium equation is incorporated into a software

package called CXTFIT developed by Torride et al. (1995), which permits one to fit a variety

of analytical solutions to the concentration distributions observed in laboratory and field

tracer studies as a function of time and/or distance. The transport of PAHs in the presence

of surfactants (Linear Alkylbenzene Sulfonate) has been successfully modelled using the

two-site model presented in equations 2.5 and 2.6 using this CXTFIT software (Chen, Wang

Page 53: Avery Gottfried - ME thesis 2009

44

et al. 2006). Noordman et al. (1998) were also successful in predicting the removal of

phenanthrene from soil using a rhamnolipid biosurfactant using a similar two site

formulation which accounted for both micellar solubilisation and admicellar sorption due to

the presence of the biosurfactant.

Page 54: Avery Gottfried - ME thesis 2009

45

CHAPTER 3 MATERIALS AND METHODS

3.1 LABORATORY SUPPLIES AND MICROOGANIM STORAGE

The important characteristics and handling procedures for the chemicals, biosurfactants,

microorganisms, and nutrient broth media purchased for this study are detailed below.

3.1.1 CHEMICALS

HPLC grade acetonitrile, hexane, acetone, and dichloromethane were purchased through

Biolab New Zealand and supplied by Mallinckrodt Baker. Naphthalene technical crystals

were from B.D.H., London, England and from Ajax Finechem, Auckland, New Zealand.

Salicylic acid crystals were from B.D.H., London, England. Phenanthrene crystals were >96%

purity HPLC grade and were supplied by Sigma Aldrich.

3.1.2 BIOSURFACTANT

Biosurfactant was purchased from Jeneil Biosurfactant Co., LLC Saukville, Wisconsin, USA.

The biosurfactant is a glycolipid produced by Pseudomonas aeruginosa with the trademark

name JBR 425. Biosurfactant stock solution contained a 25% solution of Rhamnolipids Rha-

C10-C10 (termed R1 or RLL) with the molecular formula C26H48O9 and Rha-Rha-C10-C10 (termed

R2 or RRLL) with the molecular formula C32H58O13. The rhamnolipids are an anionic

surfactant with a pKa=5.6.

3.1.3 MICROORGANISMS

The microorganisms used in this study were Pseudomonas putida ATCC 17484 (P.putida),

obtained from the American Type Culture Collection and purchased through Cryosite

Distribution in Australia. This isolate is chemoheterotropic, and is from biotype B which is

cited to degrade naphthalene. P.putida are a gram-negative rod-shaped flagellated

Page 55: Avery Gottfried - ME thesis 2009

46

bacterium which stains a pink colour when a gram stain test is performed for identification

under a microscope. P.putida are aerobic bacteria with an optimum growth temperature

between 25-30°C in a neutral pH environment. They are easily isolated from environmental

samples, and are found in most aerobic soil environments which makes it a good

representative isolate of the soil microbial consortium. For shorter storage periods, nutrient

agar plates inoculated with bacteria cultures were kept at 4°C and were re-streaked onto

fresh agar every two weeks from a single colony. For long term storage a stock of frozen

isolates, consisting of 0.5 mL of fresh overnight culture (approximately 1 x 108 – 1 x 10

9

cfu/mL) added to an equal volume of sterile 50% glycerol in a 1 mL sterile plastic tube, was

constructed and stored at -80˚C.

3.1.4 MEDIA AND NUTRIENT SUPPLY

All experiments, unless otherwise noted, were carried out using DifcoTM

Bushnell-Hass Broth

(BHB) as the nutrient supply. BHB is designed for study of microbial utilization of

hydrocarbons. It contains no carbon source, but provides all the trace elements necessary

for bacterial growth. It provides the monopotassium and diammonium hydrogen phosphate

to buffer the growth media, in an initial pH of 7.0 at 25°C. BHB was mixed at 3.27g/L and

autoclaved for 15 minutes at 121°C according to manufacturer’s instructions in 0.5 or 1L

increments (Table 3.1). Nutrient agar plates and Pseudomonas isolation agar plates, along

with a gram stain set with stabilized gram iodine, were purchased from Fort Richard

Laboratories, Auckland, New Zealand. Nutrient agar plates were also made in house by

adding 1.5% DifcoTM

agar to Lysogeny broth (LB). LB was purchased from USB corporation,

Cleveland, USA, and was made in a 20g/L solution which contains 10g/L casein peptone,

5g/L yeast extract, and 5g/L of sodium chloride. BHB and Agar were purchased from Becton

Page 56: Avery Gottfried - ME thesis 2009

47

Dickinson and Company, Sparks, USA. Other solutes that were used include glucose (Glucosa

1-hidrato) and sodium chloride reagent grade, both supplied by Panreac and Scharlav,

Spain.

Table 3.1 BHB marine salts broth approximate formula per litre of prepared media

Approximate Formula Chemical Formula Concentration (g/L)

Magnesium Sulfate MgSO4 0.2

Calcium Chloride CaCl2 0.02

Monopotassium Phosphate KH2PO4 1.0

Diammonium Hydrogen Phosphate (NH4)2HPO4 1.0

Potassium Nitrate KNO3 1.0

Ferric Chloride F2Cl3 0.05

Page 57: Avery Gottfried - ME thesis 2009

48

3.2 CELL CULTURING

This section details the bacteria storage and inoculant growth methods. Bacteria were

stored on agar plates, and single colonies were used to grow inoculants used in

experiments. Serial dilution plate counts and optical density at 600nm (OD600) methods are

presented to quantify the bacteria used in this study.

3.2.1 AGAR PLATES

Initially, bacteria were streaked out from a freeze-dried culture and allowed to grow at 28°C

until colonies were visible. For shorter storage periods, nutrient agar plates were inoculated

with bacteria cultures and re-streaked from a single colony onto fresh agar every two

weeks. They were grown for 24-48 hours at 28°C until colonies were large, and then stored

at 4°C to be used as inoculant for the liquid cultures. All agar plates were checked to ensure

colonies were of uniform shape, size, colour, and consistency. Periodic tests using

Pseudomonas selective agar and the gram-stain tests were employed to ensure everything

was aseptic and the only culture was indeed a gram-negative rod shaped Pseudomonas.

3.2.2 INOCULANT PREPARATION AND HARVESTING

The liquid cultures used to inoculate all aqueous phases, soil slurries, and soil column tests,

were prepared by transferring one loop from a single bacteria colony from a previously

cultivated plate, to 125 mL or 250 mL Erlenmeyer flasks with 50-100 mL of sterilized

medium. Flasks were then placed on a rotary shaker at 200 rpm and maintained at 25-30°C

overnight (12-20 hours) until the bacteria reached an OD600 of 1.0 to 2.0. This indicates that

they have reached their late exponential growth phase. Standard inoculant growth media

Page 58: Avery Gottfried - ME thesis 2009

49

(used for all tests other than those outlined in section 3.4.1.1) were a BHB broth with

glucose as per minimal media of 2g/L. After overnight growth, bacteria were transferred to

sterile centrifuge tubes and spun at 4000g for 5 minutes, the supernatant was then poured

off, and cells were re-suspended in 0.85% saline (w/v) at room temperature. This process

was repeated before trials were performed to remove residual broth or carbon sources.

Finally, cells were concentrated to an OD600 of 1.5 to 2.0 depending on the aqueous phase

or soil slurry phase experiment and used as the inoculant for the degradation trials.

3.2.3 PLATE COUNTS

The plate count method was used routinely in all experiments as a means of enumeration of

the viable bacteria present. This method involved performing serial dilutions and plating the

dilution series to obtain a dilution of 10 – 100 colony forming units (cfu), which then can be

accurately enumerated visually. Dilutions were done in either 1 mL culture tubes or in 96-

well micro plates, depending on the number of samples necessary. Serial dilutions in 96-well

plates were performed by adding 10 µL of neat solution to 90 µL of saline solution, while

avoiding immersing the pipette tips in the saline solution. New sterile pipette tips were used

to mix the contents in the well and add 10 µL of the mixture to the next well, until a 1x10-7

dilution series was complete. In the 1 mL culture tubes, the same technique was used,

however 100 µL of neat solution was added to 900 µL of saline solution and 100 µL was

transferred for each dilution. Ten microlitres from each dilution were then transferred onto

nutrient agar plates, and placed in a 28°C incubator overnight, or until colonies were large

enough to enumerate (Figure 3.1). All serial dilutions were done in duplicate or triplicate.

The amount of 10 µL was chosen as the transfer volume to the plate because it evaporated

and soaked into the agar within a few minutes and did not allow bacteria to become

Page 59: Avery Gottfried - ME thesis 2009

50

detached from the growing colonies. This method also allowed an entire dilution series to

be plated on a single agar plate. Final calculations for cfu/mL follow the formula:

cfu/mL = (colonies on the plate) * 10 (dilution# + 2)

Equation 3.1

3.2.4 OPTICAL DENSITY

Optical density (OD) at 600nm was also used as an indicator of cell density in solution. A

standard curve for OD600 and the density of cells per mL was constructed to calibrate the

absorbance value and obtain comparisons between readings. OD was measured by

transferring 1 mL of solution to a disposable UV-cuvette. The same spectrophotometer. The

spectrophotometer was used for all readings, and was zeroed by using the sample matrix

from a sterile stock solution.

CFU/ML ??

100/10 µL 100/10 µL 100/10 µL 100/10 µL 100/10 µL 100/10 µL 100/10

90/900 µL saline

1 2 3 4 5 6 7

x2 duplicate

10 µL of sample

Incubated the plate at 28ºC until colonies

visible and counted dilutions with

number of colonies between 10 and 100

cfu/mL = (colonies on the plate) * 10 (dilution# + 2)

Figure 3.1 Schematic diagram for serial dilutions to determine cfu/mL of solution

Page 60: Avery Gottfried - ME thesis 2009

51

3.3 SOIL METHODS

Soil was obtained from a laboratory supply (original location unknown) of mixed sand and

silt, and the properties are outlined in the following section. The artificial soil contamination

and contaminant extraction techniques are also presented, along with the efficiency of the

contaminant extraction method.

3.3.1 SOIL PROPERTIES

The particle size distribution of the soil used in all experiments was determined using

American Society for Testing and Materials (ASTM) D6913 standard testing methods for

gradation of soils, using sieve analysis (Figure 3.2). Any material larger the 2mm opening

was discarded to obtain a homogeneous mixture that would be appropriate for bench-scale

testing. According to the ASTM standards, sand is 2mm to 0.05mm; silt is 0.05mm to

0.002mm; and clay is less than 0.002mm. The soil used can therefore be classified as loamy

sand using the soil texture triangle. The loss on ignition test method ASTM D2974-87 was

used to determine the organic content of the soil (Table 3.2). Soil pH was determined in a

1:1 soil slurry with distilled water. Porosity and density were determined gravimetrically

when soil column were packed (Table 3.2 Soil properties

Table 3.2 Soil properties

Soil Parameter Symbol/Units Value

Organic Content % 2.15

pH --- 5.4

Dry Density ρd /(g/cm3) 1.75

Unit Weight γd (kN/m3) 17.2

Specific Gravity Gs 2.8

Average porosity n 0.38

Page 61: Avery Gottfried - ME thesis 2009

52

Figure 3.2 Particle size distribution

3.3.2 SOIL CONTAMINATION

Soil was sterilized by autoclaving it at 120°C in 100g increments three times, after which the

soil samples were plated on nutrient agar plates to ensure continued sterility. Sterile dry soil

was placed in 1 L shot bottles and spiked with phenanthrene dissolved in acetone. It was

then shaken vigorously for 5 minutes to promote homogeneous distribution of

phenanthrene in the soil. The amount of acetone added was sufficient to completely

saturate the soil, without producing excess liquid after shaking. Acetone was then

evaporated by allowing the sample to rest for 3 days at 30°C under a fume hood. The

contaminated soil was aged between 2 weeks and 1 month before each experiment. After

contamination, soil was re-autoclaved (as it was proven not to be sterile) and the

0.00

20.00

40.00

60.00

80.00

100.00

0.010.1110

Pe

rce

nt

Pa

ssin

g

Particle Size (µm)

Page 62: Avery Gottfried - ME thesis 2009

53

phenanthrene concentration was analyzed before and after autoclaving. There was no

change in the concentration of phenanthrene in the soil, ensuring that no phenanthrene

volatilized during the autoclaving process.

3.3.3 CONTAMINANT EXTRACTION

The remaining phenanthrene in the soil samples was extracted ultrasonically using a

modified EPA method 3550B (U.S Environmental Protection Agency 1996). Several

researchers, including Lee et al. (1999), Song et al. (2002), and Son et al. (2003 ), have

applied a modified EPA 3550B method and have assessed the efficiency in comparison to:

Soxhlet extraction, Microwave-assisted extraction, Alkaline saponification, and Direct

solvent-extraction. Final recommendations indicate that at low contamination levels, all

methods produce similar results, although further study is needed since each soil sample is

unique. There are also abnormalities in the efficiency of the commonly preferred but less

rigorous ultrasonic and shaking methods due to lower solvent consumption, quick

experimental time, and less complex equipment (Song, Jing et al. 2002; Buco, Moragues et

al. 2004). These shortcomings have now been included in the recently updated EPA method

3550C and the results are comparable to those from the modified 3550B. U.S EPA (2007)

states that samples should be extracted using a solvent system that gives optimum,

reproducible recovery of the analytes from the sample matrix, at the concentrations of

interest. For the soil in this experiment, acetone was selected as the optimum extraction

solvent, as recoveries averaged 96% (range of 85-105%) with RSD of 6%. Dichloromethante,

acetonitrile, and hexane were tested in this study, but the recovery percentage and RSD

were less favourable than acetone.

Page 63: Avery Gottfried - ME thesis 2009

54

Two grams of soil were placed with 10 mL of solvent [soil : solvent ratio of 1:5 (w/v)] in a 10

mL glass tube with Teflon-lined screw cap. All assays were conducted in triplicates. Acetone

was added to the sample which was disrupted by vortexing for 2 minutes, and

phenanthrene extracted for one hour in a Sonic Bath from Sonicator Instrument

Corporation N.Y Model SC-120. The contents in the tube were mixed for 24 hours on a

horizontal shaker table at room temperature between 22 - 27°C. Samples were centrifuged

for 10 minutes at 4000 rpm, and 1mL of supernatant collected with a glass syringe. This

liquid was filtered through a 0.2 μm poly(tetrafluoroethene) (PTFE) filter into 2mL amber

vials with PTFE-lined screw-caps. These were analyzed for phenanthrene concentration by

High Performance Liquid Chromatography (HPLC). Two grams of the same soil were weighed

and placed into a furnace at 104°C for one hour to determine the dry weight. Results

presented are on a dry weight basis. Control samples from the stock of contaminated soil

were run concurrently with the samples to ensure accuracy of the extraction procedure.

Moisture content in the soil did not appear to change the extraction results.

Page 64: Avery Gottfried - ME thesis 2009

55

3.4 EXPERIMENTAL PROCEDURES

The following sections define the experimental procedures that were used for this research.

The first objective of the investigation was to determine how bacteria respond in aqueous

solutions to a combination of biosurfactant, glucose, and salicylate in order to enhance the

biodegradation of phenanthrene by measuring changes in cell growth and biodegradation

rates. The second objective aimed to use knowledge gained from these experiments to

determine how well each amendment would work in soil slurry environments. The third

objective was to create a more realistic in situ micro environment to conduct flow-through

experiments and to determine the effect of amendments on degradation rates and flow

characteristics.

3.4.1 OBJECTIVE 1: LIQUID MEDIUM TESTS

Liquid medium tests were performed to determine both cell numbers and phenanthrene

concentration in liquid media after various amendments. The following sections outline the

procedures followed to grow inoculant cultures in various types of media, to determine the

dissolution of phenanthrene in the presence of biosurfactant, and to determine the

degradation rate of phenanthrene due to amendments.

3.4.1.1 Inoculant culture growth

Amendments were made to seed growth media to evaluate their effect on seed culture

growth stages and the degradation of phenanthrene and naphthalene in the liquid culture

trials. Naphthalene, phenanthrene, and salicylic acid were first dissolved in acetone, and

then added to each sterile flask. The acetone was allowed to evaporate before liquid growth

media was added. These cultures were then used as different types of inoculants, and the

degradation trial method was followed, as outlined in section 3.4.1.3.

Page 65: Avery Gottfried - ME thesis 2009

56

3.4.1.2 Phenanthrene dissolution

Batch dissolution experiments were conducted with phenanthrene at various biosurfactant

concentrations which ranged from 0 to 2,500 mg/L. All trials were conducted in triplicate.

Phenanthrene dissolved in acetone was added to a series of 50 mL glass tubes at a

concentration of 500 mg/L, well above the solubility limit of 1.3 mg/L. The acetone was

allowed to evaporate. Then 10 mL of biosurfactant solution at various concentrations was

added and the tubes placed in a shaker at 200rpm for 48 hours. This allowed them to reach

equilibrium conditions. The mixtures were then centrifuged at 2,500 rpm for 10 minutes to

separate the undissolved phenanthrene crystals from the supernatant. Samples were

filtered through a 0.2µm PTFE filters using a glass syringe, and transferred into 2mL amber

glass vials for analysis on HPLC. Samples were diluted 50% with acetonitrile prior to analysis

in order to obtain the same matrix solution as the prepared calibration standards for the

HPLC.

3.4.1.3 Degradation trials

Degradation trials in liquid media were conducted in 10 mL sterilized glass tubes with Teflon

lined screw caps. Naphthalene, phenanthrene and salicylic acid were first dissolved in

acetone, added to the sterile glass tubes, and the acetone was allowed to evaporate prior to

the addition of the liquid media. All tests were conducted in triplicate and 3 tubes served as

uninoculated controls during each sampling period. This was done to ensure that losses of

contaminate due to vapourization, total recovery of contaminant, and sterility was

accounted for and monitored throughout the experimental procedure. The following steps

summarize the procedure for determining cell concentration and phenanthrene removal in

batch degradation experiments.

1. Starting inoculant cultures were prepared and harvested as per section 3.2.2

Page 66: Avery Gottfried - ME thesis 2009

57

2. Naphthalene, phenanthrene and salicylic acid were added, depending on the test

conditions and acetone was evaporated.

3. Two milliliters of sterile liquid media containing BHB as the nutrient source and

glucose or biosurfactant (depending on the test conditions) were added and the

tubes left to equilibrate overnight.

4. Tubes were then inoculated with a 1 to 100 dilution (20µL in 2mL) from the

harvested cell suspensions

5. Three tubes for each variable served as initial controls and were examined

immediately. The remaining samples were incubated at 28°C on a shaker table at

200rpm, and every 24 hours three tubes for each trial were removed and examined

as follows:

6. Live cell numbers were determination by drawing off 10µL from each sample and

using the serial dilution plate count method as outlined in section 3.2.3.

7. Phenanthrene or naphthalene concentration was determined by solvent extraction

through the addition of 2mL of hexane to each tube and vortexing for 2 minutes.

Samples were then allowed to stand for one hour until the water and hexane phases

separated and then samples were centrifuged at 2000 rpm for 10 minutes. One mL

of the supernatant was drawn off and placed in an amber 2 mL vial with PTFE lined

screw cap for determination of total phenanthrene or naphthalene concentration by

HPLC analysis.

Page 67: Avery Gottfried - ME thesis 2009

58

3.4.2 OBJECTIVE 2: SOIL SLURRY TESTS

Soil slurry tests were performed to both determine cell number and the remaining

phenanthrene concentration after various amendments were made to the system. Batch

tests methods to determine the phenanthrene partitioning coefficient in the presence of

biosurfactant are outlined. The procedure followed to determine degradation rates of

phenanthrene in soil slurries due to amendments is explained and summarized.

3.4.2.1 Phenanthrene Desorption in the Presence of Surfactants

Batch desorption experiments were conducted with phenanthrene contaminated soil at

various biosurfactant concentrations, ranging from 0 to 1000mg/L. The soil samples used

were pre-contaminated with phenanthrene concentrations of 50, 100, 250 and 500 mg/kg.

Two grams of contaminated soil were added to a series of 50 mL glass tubes and 10 mL of

biosurfactant solution at concentrations of 0, 250, 500, and 1000 mg/L. The tubes were

placed on a shaker table at 200rpm. Tubes were removed at 0.5, 1, 4 ,8 ,24 , and 48 hours

for determination of phenanthrene in solution. The mixtures were centrifuge at 2,500 rpm

for 10 minutes to separate the soil particles from the supernatant. Samples were then

filtered through a 0.2µm PTFE filter using a glass syringe and transferred into 2 mL amber

glass vials for analysis on HPLC. Samples were diluted 50% with acetonitrile prior to analysis

in order to obtain the same matrix solution as the prepared calibration standards for the

HPLC.

Page 68: Avery Gottfried - ME thesis 2009

59

3.4.2.2 Degradation Trials

Degradation trials in soil slurries were conducted in 125 mL Erlenmeyer flasks which were

capped with tinfoil during the experiments. Five grams of 500 mg/kg phenanthrene

contaminated soil were added to the flasks along with 50 mL of liquid media for a soil to

water ration of 1:10. All tests were conducted in triplicate and various amendments made

to the liquid media (Table 3.3). Four uninoculated controls were also run with the

experiment to determine the recovery efficiently of the soil extraction, to determine the

amount of phenanthrene in the solution phase during each sampling period, to ensure

sterility throughout the testing period, and to serve as a bases of comparison for the results.

The following steps summarize the procedure for determining cell concentration and

phenanthrene removal in batch degradation soil slurry experiments.

1. Starting innoculant cultures were prepared.

2. Five grams of soil were added to each tube, along with the liquid media

3. Tubes were inoculated with a 1 to 100 dilution (0.5mL in 50mL) from the harvested

cell suspensions to achieve a cell density of 1x107 cells/mL of liquid or 1x10

8 cells/g

of soil.

4. At set time periods on days 4, 7, and 10, samples were taken from the flasks to

determine the amount of phenanthrene in solution and number of active bacteria.

5. Determination of cell number by drawing off 10 µL and using the serial dilution plate

count method.

6. Determination of phenanthrene in suspension by drawing off 1 mL of liquid from

each flask into 2 mL amber HPLC vials. One mL of acetone was added to stop

degradation activity, and bring all phenanthrene into the soluble phase for detection

by HPLC. Everything was completed in triplicate.

Page 69: Avery Gottfried - ME thesis 2009

60

7. On day 10, in addition to step four, the remaining liquid in the flasks was drawn off

and 10 mL of acetone was added to each flask, placed on the shaker table overnight

to enable the extraction of the remaining phenanthrene from the soil.

Table 3.3 Soil slurry media constituents

Biosurfactant Media and Amendments

0 g/L BHB (uninoculated control)

0 g/L BHB

0 g/L BHB + Salicylate 100mg/L

0 g/L BHB + Glucose 100 mg/L

0.25 g/L BHB (uninoculated control)

0.25 g/L BHB

0.25 g/L BHB + Salicylate 100mg/L

0.25 g/L BHB + Glucose 100 mg/L

1 g/L BHB (uninoculated control)

1 g/L BHB

1 g/L BHB + Salicylate 100mg/L

1 g/L BHB + Glucose 100 mg/L

5 g/L BHB (uninoculated control)

5 g/L BHB

5 g/L BHB + Salicylate 100mg/L

5 g/L BHB + Glucose 100 mg/L

Page 70: Avery Gottfried - ME thesis 2009

3.4.3 OBJECTIVE 3: COLUMN TESTS

Continuous bench-scale column tests were

conducted using the same contaminated

soil as the slurry tests in order

enhanced bioremediation procedures in a

system replicating in situ treatment.

3.4.3.1 Experimental Apparatus

Column experiments were conducted using

two glass columns, set up identically

conduct parallel experiments comparing

variables in each column

Columns measured 37cm in length

in diameter with PTFE endplates and screw

caps purchased from Omnifit

Valve INC in Boonton, New Jersey

and 30 cm along the length

ceramic cups on the inside of the column to prevent soil clogging and washout during

sampling. The other ends of the ports were connected to PTFE 3

ports could be closed and pressure sensor equip

conducted in an up-flow mane

Parmer EW-07553-85 Masterflex

13 & 14 pharmed tubing. The tubing was

UNF fittings attached to column ports and end plates

61

3: COLUMN TESTS

scale column tests were

me contaminated

in order to optimize

enhanced bioremediation procedures in a

treatment.

Experimental Apparatus

xperiments were conducted using

up identically, to

conduct parallel experiments comparing

(Figure 3.3).

Columns measured 37cm in length and 5cm

with PTFE endplates and screw

Omnifit Bio-Chem

ersey, USA . Sampling ports were installed at 2.5, 5.0, 7.5, 10, 20

(Figure 3.3). PTFE sampling ports were covered with porous

ceramic cups on the inside of the column to prevent soil clogging and washout during

of the ports were connected to PTFE 3-way v

be closed and pressure sensor equipment attached. Experiments were

flow maner to maintain complete saturation in the column using

Masterflex L/S variable-speed modular drive at 1 to 100 rpm

. The tubing was connected to 1/16" O.D. PTFE tubing with 1/4"

attached to column ports and end plates.

Figure 3.3 Soil column setup

pumping experiments

. Sampling ports were installed at 2.5, 5.0, 7.5, 10, 20

were covered with porous

ceramic cups on the inside of the column to prevent soil clogging and washout during

way valves so that the

ment attached. Experiments were

column using a Cole-

1 to 100 rpm, with L/S

/16" O.D. PTFE tubing with 1/4"-28

Soil column setup for uplflow

pumping experiments

Page 71: Avery Gottfried - ME thesis 2009

62

3.4.3.2 Pressure Measurement

Pressure changes that occurred along the length of the column were measured in five

separate locations using -100 to +200 kPa pressure transducers supplied by ICT International

in Armidale, NSW, Australia. A resolution of 0.1 kPa (1 cm H2O) could be attained for the

pressure transducers by using smart interfaces to maintain a stable reference voltage. Data

was recorded on a 16-bit Plug & Play Smart Logger supplied by ICT international which could

store up to 500,000 date-and-time stamped readings.

3.4.3.3 Micro-foam Generation and Stability

Micro-foam was generated based on a method proposed by Sebba (1985) using a high-

speed spinning disk method. One litre of surfactant solution was poured into a baffled

container and stirred at 8,000 rpm for three minutes. The disk was slowed down to 6,000

rpm for the duration of pumping experiments to maintain microfoam quality. The

temperature of the solution slowly increased over the course of the experiment starting at

25°C and rising to 30°C after approximately four hours of operation. Microfoam stability

measurements used the half-life method. After three minutes of intensive stirring, 100 mL

of microfoam was transferred to a 100 mL measuring cylinder. The height of the clear liquid

interface below the dispersion was recorded at various times, and after approximately 4

hours dispersion was complete and the final volume of the drained liquid recorded. The

stability of the micro-foam is reported as the time needed for draining half of the liquid

from the dispersion, based on a plot of time versus percentage drained.

Page 72: Avery Gottfried - ME thesis 2009

63

3.4.3.4 Column Packing and Unpacking

All column tests were dry-packed using contaminated soil. Columns were first flushed with a

90% ethanol solution and left for 48 hours to sterilize all parts of the column system.

Afterwards, columns were flushed continuously with sterile miliQ water to remove any

remaining ethanol. Thirty grams of coarse sand was added to the column at the inlet to

prevent finer particles from settling into the inlet tubing. Contaminated soil was pre-

inoculated with P.putida prior to soil column packing. Harvested cells were concentrated to

an OD600 of 2.0 and 1mL per 100g of soil of the concentrated cell solution was added,

corresponding to a cell density of approximately 1x108 cells/g of soil. The soil was vigorously

shaken in sterile lab bottles prior to column packing to distribute the bacteria evenly

throughout the soil. Columns were then dry-packed with soil in 200g increments. After each

increment was added, the column was tapped and a sterile rod was used to pack the soil by

prodding the added increment 5 – 7 times. A 90µm PTFE frit was placed in the top cap of the

column to prevent any washout of soil particles. Columns were then purged with CO2 gas for

approximately 30 minutes to displace any oxygen. They were then saturated with sterile

BHB nutrient broth. Column unpacking was done in 7 equally spaced layers from top to

bottom. At each layer triplicate extractions were performed on the soil to determine

remaining phenanthrene concentration. Columns were repacked for each experiment that

was carried out using the same mass of soil and amount of packing.

Page 73: Avery Gottfried - ME thesis 2009

64

3.4.3.5 Experimental Operating Conditions

In un-inoculated soil, column tracer breakthrough curve tests were conducted using a

potassium chloride solution in BHB broth at a flow rate of 0.2mL/min. A biosurfactant

breakthrough curve was attained using the same flow rate to determine the amount of

adsorption that occurs during biosurfactant addition. A flow rate of 0.2mL/min corresponds

to a flow velocity of 0.026 cm/min or 24 hours to pump one pore volume of solution

through the column.

Column tracer tests were then performed using a potassium chloride solution in BHB broth

with biosurfactant added at a concentration of 1000 g/L and 5000 g/L at a flow rate of 10

mL/min. The same solution was also used to conduct tracer tests using microfoam instead

of aqueous solution to determine the liquid fraction break through curve when using

microfoam, and determine pressure changes that occur during microfoam pumping across

the length of the column.

Three series of degradation trials were conducted with the solution amendments and flow

rates as outlined below in Table 3.4. Degradation trials were run for up to 10 days and on

each day samples were taken from the inlet, various sampling ports along the column

length, and the column outlet to determine phenanthrene and biosurfactant concentration

in solution, cfu/mL, and dissolved oxygen levels (using a MI-730 dip-type 02 microelectrode

connected to an OM-4 Oxygen meter from Microelectrodes INC, Bedford, New Hampshire,

USA). Influent solutions were all sterilized and were placed on magnetic stir plates to

maintain saturated levels of oxygen in the influent solution. Depending on the testing

conditions as indicated in Table 3.4, the influent was switched to a second solution for the

Page 74: Avery Gottfried - ME thesis 2009

65

pulse duration simply by using a 3-way PTFE valve on the influent line. Special care was

taken to sterilize all tubing and equipment throughout the testing period. At the end of 10

days the column was unpacked and total remaining phenanthrene in soil was determined.

Table 3.4 Column trial experimental design

Column

number

Flow

conditions

Total flow Standard

influent

solution

Pulse

Influent

solution

Pulse

duration

Initial soil

contamination

TRIAL 1

1 0.2mL/min

10 days

2,880mL

(10.1 PV)

BHB

biosurfactant 1g/L

salicylate 100mg/L

n/a n/a 450 mg/kg

2 0.2mL/min

10 days

2,880mL

(10.1 PV)

BHB

biosurfactant 1g/l

n/a n/a 450 mg/kg

TRIAL 2

1 0.5mL/min

10 days

6,900mL

(24.3 PV)

BHB biosurfactant

1g/L salicylite

100mg/L as

microfoam

4 hours

on days

2,4,6 &

8

450 mg/kg

2 0.5mL/min

10 days

6,900mL

(24.3 PV)

BHB biosurfactant

1g/L salicylite

100mg/L as

liquid solution

4 hours

on days

2,4,6 &

8

450 mg/kg

Page 75: Avery Gottfried - ME thesis 2009

66

3.5 ANALITICAL METHODS

3.5.1 PAH DETECTION

Independent stock solutions of naphthalene and phenanthrene were prepared by dissolving

0.01g of naphthalene or phenanthrene crystals in a small amount of hexane or acetonitrile,

then bulking up to 10 mL with the same solvent. This method was used to create stock

solutions of 1000 mg/L, stored in the dark at 4°C, and last up to three months. From the

stock solutions independent solutions ranging from 0.1 to 100mg/L were prepared to

construct 5 point calibration lines for each sample sequence that was run; determination

coefficients obtained were higher than 0.99 to insure good linearity and reproducibility over

the sampling range. To account for various matrix compositions multiple standards were

made in various solvent solutions. This insured the relative recovery was near 100% -- the

calibration standards along with given test samples were processed and run using the HPLC

in the same way (García-Falcón, Cancho-Grande et al. 2004). HPLC was equipped with a

Dionex P580 pump, Dionex ASI-100 Automated Sample Injector, Dionex UVD340S UV

detector and a Phenomenex C18 Column (150x4.6mm), used for the detection of

naphthalene and phenanthrene. HPLC analysis was performed isocratically with a mobile

phase of 30% water and 70% acetonitrile at a flow rate of 1.5 ml/min to elute the column,

with the sample injection volume being 10µl. The optimal UV detection wavelengths were

found to be 220nm and 250nm for Naphthalene and Phenanthrene respectively; these

wavelengths gave the greatest response height and area, and allowed for detection limits

below 0.01 mg/L for phenanthrene and 0.1 mg/L for naphthalene using this method.

Chromeleon Client Version 6.80 by Dionex Corporation was used for data collection and

analysis.

Page 76: Avery Gottfried - ME thesis 2009

67

3.5.2 BIOSURFACTANT DETECTION

Biosurfactant was quantified using two different methods, the first using a total organic

carbon (TOC) analyzer to determine the relative amount of carbon present and relate back

to the concentration of rhamnolipid, and the second using the HPLC to quantify both the R1

and R2 rhamnolipids present in solution. TOC samples were collected in 40-mL pre-cleaned

glass vials were measured immediately after sapling using a TOC-V-CSH analyzer (Shimadzu

Corporation, Kyoto, Japan).

Biosurfactant was detected using a HPLC/ELSD following the method outlined by (Wang,

Fang et al. 2007). The HPLC gradient was obtained by the following method: starting at 8%

solvent B and holding for 1 minute, ramping to 75% solvent B in 20 minutes, holding 75%

solvent B for 10 minutes, backing to 8% solvent B in 1 minute and holding at 8% solvent B

for 5 minutes. Solvent A was 98:2 (v/v) water : acetonitrile with 0.1% acetic acid, and solvent

B is 10:90 water : acetonitrile with 0.1 % acetic acid. Rhamnolipid samples were filtered

through a 0.22um syringe filter prior to analysis using a 50µl sample injection volume.

3.5.3 CHLORIDE ANION

Chloride used as a tracer in the column tests was detected using a Dionex DX-120 Ion

Chromatograph (IC) with an AS9-HC column. The effluent was a Na2CO3 solution. A

potassium chloride stock solution was made to prepare calibration standards, and using this

method a detection limit of 0.1ppm was obtained.

Page 77: Avery Gottfried - ME thesis 2009

68

CHAPTER 4 RESULTS AND DISCUSSION

4.1 OBJECTIVE 1: LIQUID CULTURES

4.1.1 BACTERIA GROWTH ON VARIOUS SUBSTRATES

Influences of glucose, naphthalene, salicylate, and biosurfactant on bacteria growth were

investigated to evaluate the rate and extent of growth that could be obtained from the

various carbon sources. Bacteria growing in suspended, shaken, cultures with crystalline

PAH above the aqueous solubility limit as the sole carbon source of energy, exhibit

characteristic growth curves involving four phases. The phases are: (i) lag phase, (ii)

exponential phase, (iii) a subsequent phase with pseudo-linear growth, and (iv) pseudo-

stationary phase (Volkering, Breure et al. 1992; Volkering, Breure et al. 1993; Wick,

Colangelo et al. 2001; Johnsen, Wick et al. 2005). The partial growth curves (Figure 4.1) for

P.putida in four different medias of glucose (2g/L), naphthalene (0.5g/L), salicylate (0.5g/L)

and naphthalene (0.5g/L) + biosurfactant (1g/L) showed P.putida was not just capable of

using naphthalene as a sole growth substrate, but was able to obtain the same amount of

growth as with highly water-soluble substrates such as glucose. Adding biosurfactant to

increase the solubility of naphthalene had no effect on increasing the growth rate beyond

that obtained without biosurfactant addition or with glucose, indicating that bioavailability

of naphthalene is not limiting the growth of P.putida in liquid cultures. Naphthalene has a

higher water solubility (31.4mg/L) compared to most other PAHs (<1mg/L), therefore the

dissolution of naphthalene crystals into solution occurred at a rate high to support the

organisms’ exponential growth. In a batch study by Wick et al. (2001), only a few anthracene

crystals (<2g/L) resulted in pseudo-linear growth due to low dissolution fluxes, whereas

exponential growth was only obtained when high amounts of solid anthracene (30g/L) were

Page 78: Avery Gottfried - ME thesis 2009

69

provided. As the solubility of naphthalene is greater than 600-fold higher than anthracene it

could be assumed that the addition of 500mg/L of naphthalene crystals would have

provided dissolution fluxes that would produce enough substrate to support exponential

growth.

With this specific Pseudomonas putida ATCC 17484 strain, the biodegradation rate appears

to be controlled by the metabolic activity of the bacteria. Salicylate did not produce the

same rate of growth as glucose or naphthalene indicating a slightly lower amount of

metabolic activity occurring when salicylate is the sole carbon source. Cells grown in 500

mg/L and 1000 mg/L biosurfactant were also tested; after 24 hours the growth became

stationary with OD measuring 0.131 and OD measuring 0.235, respectively. The increased

biosurfactant concentration produced roughly twice the cell density, but the low amounts of

biomass produced overall indicates that biosurfactant is not a preferential carbon source.

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

1.1

0 2 4 6 8 10 12 14

OD

60

0n

m

Time (hours)

Glucose Naphthalene Salicylate Naphthalene + Biosurfactant

Figure 4.1 Partial growth curves for P.Putida until early stationary phase in four growth medias

of glucose (2g/L); naphthalene (0.5g/L); salicylate (0.5g/L); and naphthalene (0.5g/L) +

biosurfactant (1g/L)

Page 79: Avery Gottfried - ME thesis 2009

70

4.1.2 EFFECT OF INOCULANT ACCLIMATIZATION AND PRE-TREATMENT

Different acclimatization or pre-treatments of the seed inoculants were tested to determine

what affect the seed growth conditions have on the removal of naphthalene and

phenanthrene in subsequent batch studies. Bacteria for lab studies are often grown in the

presence of target contaminants to maintain the degradative phenotype, and to increase

the enzyme activity that is responsible for the degradation of the target compound. This is

done to improve removal efficiency and decrease lag time in cell growth when bacterial

seeds are added to the system where the contaminant is to be degraded. The affect of

seven inoculant pre-treatments were tested (Figure 4.2) on the subsequent degradation of

naphthalene over a four day period. All cultures obtained the same amount of cell growth

(between 1x108 and 1x10

9 cfu/mL) in the first 24 hours. This concentration was maintained

over the 4 day testing period (Figure 4.2). The results for the bacteria growth in each

separate system were proven to be statistically similar to each other and are presented as

an average from all the tested inoculant systems (single factor ANOVA F=0.88 p=0.51). The

total amount of degradation was similar in all systems after 96 hours. After 74 hours the

samples where BHB was the nutrient source during seed growth showed virtually total

mineralization of all the naphthalene, whereas the samples which had LB as the nutrient

source during seed growth required 96 hours before all the naphthalene was mineralized.

Growth of the seed for one week prior to inoculation showed a lag phase of approximately

24 hours before significant naphthalene removal was measured. After 24 hours

naphthalene removal was similar in all the systems regardless of the growth conditions of

the inoculant seeds. Guerin and Boyd (1995) obtained similar observations where cultures

of P.putida maintained a high degree of capacity to degrade naphthalene for several days

following entry into stationary phase. Naphthalene degradation activity was present

Page 80: Avery Gottfried - ME thesis 2009

71

constitutively at low levels under all growth conditions and was rapidly induced to high

levels upon exposure to naphthalene. Naphthalene loss due to volatilization was substantial

over the 5 day experimental period as indicated by a 50% loss in the control samples.

Volatilization adds uncertainty to the results; therefore, it is difficult to draw further

conclusions from the study of aerobic naphthalene degradation as accounting for

naphthalene loss.

Figure 4.2 Naphthalene degradation and cell growth in liquid cultures containing different

bacteria inoculant seeds which were pre-grown in seven different solutions (s1

BHB+naphthalene+glucose grown for 1 week; s2 BHB+glucose; s3 BHB+naphthalene+glucose; s4

BHB+salicylic acid+glucose; s5 LB; s6 LB+naphthalene; s7 LB+salicylic acid; s2 - s7 grown overnight

approximately 20 hours growth)

1.E+02

1.E+03

1.E+04

1.E+05

1.E+06

1.E+07

1.E+08

1.E+09

0

1

10

100

1000

0 20 40 60 80 100

Ba

cte

ria

Co

nce

ntr

ait

on

(cf

u/m

L)

To

tal

Na

ph

tha

len

e R

em

ain

ing

(m

g/L

)

Time (hours)

s1 s2 s3 s4 s5 s6 s7 control Viable Cells

Page 81: Avery Gottfried - ME thesis 2009

72

Different pre-treatments were also tested with regards to phenanthrene degradation and

showed little change in the degradation ability of the bacteria (data not shown). The only

added benefit was seen in pure phenanthrene solution with no amendments, where an

increase in total phenanthrene degradation occurred when the inoculant was cultured in a

naphthalene and BHB solution overnight and spiked with glucose 3 hours prior to

inoculation. However, there was no statistical difference between inoculants that were

added to phenanthrene solutions that contained amendments of salicylate, biosurfactant

and glucose. This indicates that pre-treatment and acclimatization of the bacteria had

minimal effect compared to amendments to the culture during degradation as discussed in

section 4.1.4.

Studies with a P.putida have shown the strongest expression of nah genes (genes

responsible for naphthalene degradation) in late-log phase growth, where the catabolic

operons were poorly induced in the early-exponential growth phase but strongly induced in

the late-exponential-growth phase (Hugouvieux-Cotte-Pattat, Kohler et al. 1990; Guerin and

Boyd 1995). The results obtained in this study support these observations and indicate that

harvesting the seed cultures in the late-exponential or early stationary phase will have the

best ability to enhance the degradation process, and it is less important to acclimatize

P.putida bacteria seeds prior to inoculation.

Page 82: Avery Gottfried - ME thesis 2009

73

4.1.3 BIOSURFACTANT SOLUBILITY ENHANCEMENT

The apparent aqueous solubility enhancement of phenanthrene in a biosurfactant solution

can be expressed as a function of both the interaction with the surfactant monomers and

with the surfactant micelles once the concentrations exceeds the respective CMC (Zhu,

Chen et al. 2003; Wang and Keller 2008).

���

��� � " ������ " ������ (Equation 4.1)

• FG� is the apparent phenanthrene solubility at a total stoichiometric surfactant

concentration of Xmn and Xmc (g/L)

• FG is the intrinsic phenanthrene solubility in water in the absence of the surfactant

(mg/L)

• HI<and HIJ are the surfactant monomer and micelle concentration in water (g/L)

• �I<and �IJ are the phenanthrene portioning coefficient with the surfactant

monomer and micellar phases (L/g)

0

5

10

15

20

25

30

35

40

0 500 1000 1500 2000 2500

Ph

en

an

thre

ne

Co

nce

ntr

ati

on

(S

*w

/Sw

)

Biosurfactant Concentration (mg/L)

S*w/Sw = 1

Figure 4.3 Phenanthrene solubility enhancement as a function of biosurfactant

concentration. The equation refers to the fit of data above the CMC and

� � ��� ��⁄

Page 83: Avery Gottfried - ME thesis 2009

74

FG� and FGwere measured directly on the HPLC and the results of the Phenanthrene

solubility enhancement are presented ( ). Linear regression was conducted on the solubility

data for values of FG� FG L 1⁄ , using the average of triplicate measurements of

phenanthrene solubility at various biosurfactant concentrations above the CMC. Similar

linear relationship for surfactant enhanced solubility curves above the CHC have been

determined by a number of researchers (Kim, Park et al. 2001; Shin, Kim et al. 2004; Yu, Zhu

et al. 2007). It should be noted that the value for Sw (phenanthrene solubility in water)

obtained in this experiment was 1.12 mg/L and this value was used in subsequent

calculations. There appears to be no consensus for the phenanthrene solubility in water as

results are reported in the range of 0.4-1.6 mg/L for temperatures ranging from 8.5-30°C;

there is no consensus for a single solubility value in water at 25°C (Verschueren 1983). It is

assumed there is a lack of interaction between surfactant monomers and phenanthrene, as

phenanthrene Kow = 104.57

and according to literature only extremely hydrophobic organic

compounds such as DDT with at Kow of 106.36

are known to associate with surfactant

monomers in the aqueous phase, allowing Kmn to equal 0 in Equation 4.1. This allows

FG� FG � 1⁄ before the CMC is reached, and after the CMC has been exceeded there is a

linear increase in solubility with increasing surfactant concentration, with R2 = 0.9979 for the

data presented (Figure 4.3). Using equation 4.1 the CMC can be solved for by equating

FG� FG L 1⁄ with the linear regression line and this gives a CMC of 63 mg/L. This value

correlates well with Shin et al. (2008), who reports a CMC with a similar rhamnolipid of 0.1

mmol/L at a pH of 7. This is 56 mg/L assuming that the biosurfactant product supplied was

an equal mix of mono to di-rhamnolipied by weight. It was also shown by (Shin, Kim et al.

2008) that pH can have a significant effect on the CMC of the rhamnolipid and can cause

considerable changes in the solubility of phenanthrene in the presence of rhamnolipid.

Page 84: Avery Gottfried - ME thesis 2009

75

4.1.4 PHENANTHRENE DEGRADATION

Phenanthrene degradation took place in all systems over a 46 hour period, with the addition

of BHB broth and/or biosurfactant (1.0 g/L), salicylate (100 mg/L), and glucose (100 mg/L)

(Figure 4.4). Increased removal of phenanthrene occurred in systems which contained

salicylate, showing a 3.5 fold increase in removal compared to no amendment addition. The

addition of biosurfactant enhanced the amount of degradation in all systems with 1.5 to 2.6-

fold increase in phenanthrene removal versus equivalent systems with no biosurfactant

addition.

Figure 4.4 Phenanthrene degradation in liquid cultures containing BHB and/or biosurfactant

(1000mg/L), salicylate (100mg/L), and glucose (100mg/L). Data presented is the average of

triplicate measurements taken at 22 and 46 hours after inoculation.

0

10

20

30

40

50

60

70

80

90

100

BHB BHB + Salicylic

Acid

BHB+ Glucose BHB +

Biosurfactant

BHB +

Biosurfactant

+ Salicylic Acid

BHB +

Biosurfactant

+ Glucose

% P

he

na

nth

ren

e R

em

ov

al

After 22 hours After 46 hours

Page 85: Avery Gottfried - ME thesis 2009

76

Bacterial uptake of solubilized compounds is hypothesized to be influenced by

biosurfactants, as the uptake of biosurfactant-solubilized molecules has been found to be

faster than the uptake of dissolved (i.e monodispered) molecules (Johnsen, Wick et al.

2005). Results here indicate P.putida degradation occurs faster in the presence of

biosurfactant. However, salicylate addition, with no biosurfactant addition, increased the

total degradation of phenanthrene 30% more than system with only biosurfactant addition.

Glucose was shown to improve phenanthrene degradation which is attributed to the

presence of an additional carbon source that increased bacteria growth in the system.

Phenanthrene removal was nearly 2-fold more due to salicylate addition compared to the

glucose system; therefore it can be assumed that it is not just the increased bacterial growth

due to the presence of an additional carbon source that is responsible for the increased

removal in the salicylate systems. This indicates a greater amount of phenanthrene

degradation could be achieved through amendments that specifically cause metabolic

pathway induction, when compared to increased solubility brought about by the addition of

biosurfactant, or increased growth brought about by additional carbon substrates.

Salicylate enhanced the rate of degradation in the first 22 hours, although this rate was not

sustained, and decreased considerably from 2.2 mg/hr to 0.88 mg/hr in the following 24

hours (Table 4.1). The addition of biosurfactant increased initial degradation rates, and

sustained relatively higher degradation rates over the 46 hour period compared to systems

without biosurfactant. Determining the length of time apparent enhancement strategies

influence the system for has important implications for the design of in situ treatment

strategies. Powell et al. (2008) determined that the rate at which a carbon source is made

available or the instantaneous concentration in the medium (spike addition versus

Page 86: Avery Gottfried - ME thesis 2009

77

continuous addition) considerably changes the outcomes achieved. Therefore it is important

to determine what method works best for a particular system when evaluating remediation

strategies. Salicylate offers the best initial and total phenanthrene removal, and

biosurfactant augmented these results and could offer benefit over a longer period of time.

Table 4.1 Rate of phenanthrene degradation in liquid cultures expressed as mg of

phenanthrene degraded / hour

Media Constituents Degradation rate

after 22 hours

Degradation rate

22 hours - 46 hours

Average over

46 hours

BHB 0.47 0.43 0.45

BHB + salicylate 2.20 0.88 1.51

BHB + glucose 1.11 0.52 0.80

BHB + biosurfactant 1.45 0.89 1.16

BHB + biosurfactant + salicylate 2.60 1.40 1.98

BHB + biosurfactant + glucose 1.89 1.33 1.59

Page 87: Avery Gottfried - ME thesis 2009

78

4.2 OBJECTIVE 2: SOIL SLURRIES

4.2.1 PHENANTHRENE DISTRIBUTION IN THE PRESENCE OF BIOSURFACTANTS

Desorption experiments were conducted to determine how increasing rhamnolipid

concentrations affected the amount of phenanthrene in the aqueous phase. Samples were

removed at succeeding time intervals of (1, 4, 8, 24, and 48 hours) to determine the change

in aqueous phenanthrene concentration over time, as well as the final equilibrium

phenanthrene concentration after 48 hours. Equilibrium phenanthrene concentration for

increasing surfactant concentrations of 0, 250, 500 and 1000 mg/L (Figure 4.5) indicate that

equilibrium desorption isotherms can be fitted with a linear equation and described with a

single soil-water partitioning coefficient Kd. Similar linear desorption isotherms were found

by Haung and Cha (2001) and were used to calculate the desorption of PAHs in the

0

100

200

300

400

500

600

0 1 2 3 4 5

Ph

en

an

thre

ne

in

so

il u

g/g

Phenanthrene in solution mg/L

0 mg/l 250 mg/l 500 mg/l 1000 mg/lKd = 829 Kd = 520 Kd = 217 Kd = 96

Biosurfactant Concentration

Figure 4.5 Phenanthrene desorption from soil in the presence of biosurfactant.

Desorption partitioning coefficient Kd calculated from the linear regression trendline for

each series of data.

Page 88: Avery Gottfried - ME thesis 2009

79

presences of a rhamnolipid biosurfactant. Results indicate that phenanthrene interaction

with the biosurfactant solution is independent of the phenanthrene concentration in the

soil in the range of soil contamination studied (50 to 500 mg/kg), as aqueous solubility limits

were not reached. The partitioning coefficient Kd decreased with increasing biosurfactant

concentration, with Kd decreasing by a multiple of 8.5, from 829 to 96, due to the addition

of 1000 mg/L of biosurfactant.

The sorption of surfactant onto soils is known to have a significant effect on the

performance of surfactant enhanced desorption of contaminants (Zhou and Zhu 2007; Zhu

and Zhou 2008; Laha, Tansel et al. In Press). Although biosurfactants form a mobile micellar

pseudophase, they are also adsorbed by the soil matrix and can lead to phenanthrene

partitioning into the adsorbed biosurfactants (Zhou and Zhu 2007). This in turn enhances

the sorption of phenanthrene on the soil. The process of increased phenanthrene

adsorption to soil appears to be responsible for the trend observed over a 48 hour period

(Figure 4.6). The initial amount of phenanthrene in solution increased for approximately 4 to

8 hours, and then began to decrease over the remainder of the experiment. This is believed

to be caused by the sorption of biosurfactant to the soil surface, which in turn enhanced the

sorption of phenanthrene back onto the soil. The sorption sites on the soil would become

‘available’ after phenanthrene initially desorbs into solution, leaving sites available for

biosurfactants to sorb to the soil surface. This would increase the overall organic content of

the soil, causing sorption of phenanthrene back onto the soil before an equilibrium

concentration is reached. Garcia-Junco et al. (2003), saw a similar occurrence where

biosurfactant may have promoted sorption of PAH to the soil by modifying the surface

hydrophobicity due to the orientation of the hydrophobic moieties of the first

Page 89: Avery Gottfried - ME thesis 2009

80

Figure 4.6 Phenanthrene desorption from contaminated soil (50, 100, 250, 500 mg/kg) into

aqueous solution in the presence of biosurfactant over a 48 hour period.

0

1

2

3

4

5

6

Ph

en

an

thre

ne

in

solu

tio

nm

g/L

50 mg/kg 100 mg/kg 250 mg/kg 500 mg/kg

Biosurfactant 1000 mg/L

0

1

2

3

4

5

Ph

en

an

thre

ne

in

solu

tio

n m

g/L

Biosurfactant 500 mg/L

0

0.5

1

1.5

2

2.5

3

Ph

en

an

thre

ne

in

solu

tio

n m

g/L

Biosurfactant 250 mg/L

0

0.2

0.4

0.6

0.8

0 10 20 30 40 50

Ph

en

an

thre

ne

in

solu

tio

n m

g/L

Time Hours

Biosurfactant 0 mg/L

Page 90: Avery Gottfried - ME thesis 2009

81

biosurfactant layer that adsorbs to the surface in the aqueous phase. PAHs may also

partition into the hemimicelles that are formed as a layer on the soil surface by sorbed

biosurfactants. The time dependent phenanthrene sorption does not appear to occur in the

1000 mg/L biosurfactant solution (Figure 4.6). At lower biosurfactant concentrations the

amount of surfactant sorption to the soil accounts for a majority of the added surfactant;

whereas at high biosurfactant concentrations (1000 mg/L) there is still a significant amount

of biosurfactant in the aqueous phase which can act to increase the overall aqueous

solubility of phenanthrene.

Rhamnolipid adsorption to soil types with similar amounts of organic matter have been

reported as high as 2,500 mg/kg, with a rhamnolipid partitioning coefficient of 34 L/kg

(Huang and Cha 2001). However, if there is enough biosurfactant remaining in solution (as

with the 1000 mg/L solution), it appears the amount of rhamnolipid sorption to the soil has

no significant effect on the overall PAH desorption. Low levels of biosurfactant increase the

affinity of the solids for aqueous phenanthrene due to biosurfactant sorption, whereas at

levels well above the CMC, the Kd values were lower due to competition between micellar

and sorbed biosurfactants for phenanthrene partitioning.

Table 4.2 Calculated phenanthrene soil partitioning coefficient Kd, and phenanthrene

partitioning onto soil sorbed surfactant coefficient Ks

Biosurfactant Concentration (mg/L) Kmc (L/kg) Kd* (L/kg) Log Ks

250 8,894 520 4.40

500 11,535 217 4.51

1,000 12,585 96 4.52

Page 91: Avery Gottfried - ME thesis 2009

82

Using equations 2.2 and 4.1, the distribution coefficient Ks (the solute distribution

coefficient with the soil-sorbed surfactant) which represents the quantity of phenanthrene

sorbed to the soil-sorbed biosurfactant can be solved using the Kd values determined, and

assuming Qs equals 34 (L/kg) from (Huang and Cha 2001). The calculated Ks values for

phenanthrene partitioning onto soil sorbed surfactant (Table 4.2) follow a similar trend as

reported by Zhu and Zhou (2008). Ks values ranging from 3.75 to 4.75 on the log scale were

reported, depending on the concentration and type of surfactant added. The trend Zhu and

Zhou (2008) determined showed when 0 to 200 mg/L of titron X-100 were added, values of

Ks increased from 4.0 to 4.25. Surfactant addition above 200 mg/L did not increased Ks any

further and values were constant for the addition of 200 mg/L to 2000 mg/L of surfactant.

This is similar to the trend (Table 4.2) were Ks is 4.4 at 250 mg/L and increases to a constant

value near 4.5 at 500 mg/L and 1000 mg/L. Ks has a strong relation to the amount of

surfactant sorbed to the soil, and accounts for a large amount of increased PAH sorption to

soil that occurs due to the addition of surfactant. This can be attributed to the three stage

surfactant adsorption process that was described by Torrens et al. (1998). At low surfactant

concentrations, the sorbed surfactant molecules are spread out over the soil surface. As the

surfactant concentration increases the monomers form a monolayer or hemimicelles, which

is considered as a weak partition phase for PAHs (Zhu and Zhou 2008). As the surfactant

concentration increases the soil surface reaches a point where the entire solid surface is

covered by a hemimicelles and begins to form surface micelles (admicelle) which are

bilayers on the soil surface that have greater sorption capacity for PAHs. Finally surfactant

sorption plateaus as micelles begin to return to solution and the soil has research surfactant

sorption capacity and equilibrium conditions exist.

Page 92: Avery Gottfried - ME thesis 2009

83

4.2.2 SOIL DEGRADATION

Phenanthrene degradation occurred in the combined aqueous phase and suspended

organic matter in soil slurries containing biosurfactant (0, 0.25, 0.5, 1.0 g/L), salicylate (100

mg/L), and glucose (100 mg/L) inoculated with P.putida (Figure 4.7). Both natural organic

mater and biosurfactants appear to increase the apparent aqueous solubility of

phenanthrene as the concentration in the control flasks averaged 32 mg/L with no

biosurfactant present. This increased to 38 mg/L in the presence of 1 g/L biosurfactant. The

results presented (Figure 4.7) represent both the aqueous solubility and the phenanthrene

concentration in the suspended organic matter; as samples were taken from the liquid in

the soil slurries but were not filtered to remove suspended soil particles prior to the

addition of acetone to extract the phenanthrene. Filtered samples were taken during the

first sampling period from the control flasks and there was 1.24, 3.01, 6.27, 6.88 mg/L of

phenanthrene for 0, 0.5, 1.0, 5.0 g/L biosurfactant concentrations respectively. This

correlated with the expected values from the phenanthrene desorption results present in

section 4.2.1. The total phenanthrene concentrations (Figure 4.7) are a result of the

enhanced solubility due to the addition of biosurfactant, in addition to approximately 32

mg/L of phenanthrene that was present in the suspended organic matter and dissolved

organic matter. Similar results were obtained by Cho et al. (2002) where apparent solubility

of phenanthrene and other PAHs was accounted for by the sum of the phenanthrene

solubility in natural organic matter solution plus the solubility in a non-ionic surfactant

solution. Roskam and Comans (In Press) also observed PAH levels in solution that were up to

an order of magnitude higher in batch tests versus column tests due to the large dissolved

organic carbon molecules that were generated by more vigorous mixing.

Page 93: Avery Gottfried - ME thesis 2009

84

The amount of phenanthrene present in solution was constant over sampling periods of 4,

7, and 10 days. Equilibrium existed between concentration of phenanthrene being degraded

by the bacteria, and the rate of phenanthrene desorption from the soil to the aqueous

phase. There was 1 to 4-fold less phenanthrene present in solution in flasks which contained

salicylate, regardless of the amount of biosurfactant present. There were 5.6, 2.3, 1.5, and

1.7 mg/L of phenanthrene present in the systems which contained biosurfactant at 0, 0.25,

0.1

1

10

Co

ntr

ol

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

Co

ntr

ol

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

Co

ntr

ol

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

Co

ntr

ol

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

No Biosurfactant Biosurfactant 0.25 g/L Biosurfactant 1 g/L Biosurfactant 5 g/L

Ph

en

an

thre

ne

(m

g/L

)

Day 4 Day 7 Day 10

Figure 4.7 Total phenanthrene concentration in solution and suspended/dissolved organic

matter in soil slurries containing BHB and/or biosurfactant (0.25, 1, 5 g/L), salicylate

(100mg/L), and glucose (100mg/L) over a 10 day period

Page 94: Avery Gottfried - ME thesis 2009

85

1, and 5 g/L respectively. As the amount of biosurfactant increased, the amount of

phenanthrene in solution decreased, until a point between 1 g/L and 5 g/L biosurfactant

concentration. This signifies increased microbial degradation due to the addition of

biosurfactant until a point between 1 g/L and 5 g/L where the additional biosurfactant did

not improved degradation. Biosurfactant also appeared to enhance degradation in systems

with glucose, as the average concentration of phenanthrene in solution decreased from 2.5

mg/L with no biosurfactant present, to 0.5 mg/L in 5 g/L biosurfactant solution.

Biosurfactant had little effect on systems with salicylate as average amounts of

phenanthrene were 0.4 mg/L in all salicylate systems. There was a 4-fold increase in the

concentration of phenanthrene in the 5 g/L biosurfactant and salicylate system, compared

to the system with no biosurfactant.

Mass balance calculations were performed to determine total phenanthrene remaining

including both phenanthrene in the soil and aqueous phase after 10 days of incubation.

Results are presented as total phenanthrene in mg/kg of dry soil (Figure 4.8). The greatest

phenanthrene removal was in systems that contained salicylate with greater than 90%

phenanthrene removal achieved in all instances. There was also a significant decrease in the

concentration of phenanthrene remaining due to the addition of biosurfactant with 86, 90

and 91 percent removal due to the addition of 0.25, 1.0 and 5.0 g/L of biosurfactant

respectively compared to 68% removal when no biosurfactant was added. Biosurfactant

also improved the amount of removal when glucose was present, but the effects were less

significant when compared to the system with no biosurfactant and only glucose, with

removals increasing by 1.6, 1.9 and 1.6-fold with the addition of 0.25, 1.0 and 5.0 g/L of

biosurfactant respectively. With salicylate the effects of biosurfactant were almost non

Page 95: Avery Gottfried - ME thesis 2009

86

existent as the total removal was 1.2-fold more due to the presence of 0.25 g/L

biosurfactant, and there was no improvement and a decrease in total removal when 1.0 and

5.0 g/L biosurfactant was added. From the results obtained it appears that almost complete

removal was obtained in most systems, and the concentration of phenanthrene in solution

changed very little over 10 days.

Total live cell counts over the 10 day experimental period (Figure 4.9) represent both the

cell concentration in the aqueous phase and cells that could have been attached to the

suspended organic matter, as solution was drawn off from each flask and not filtered to

remove suspended soil matter prior to serial dilutions. The same level of active growth near

0

20

40

60

80

100

120

140

160

180

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

BH

B

BH

B +

Sa

licy

lite

BH

B +

Glu

cose

No Biosurfactant Biosurfactant 0.25 g/L Biosurfactant 1 g/L Biosurfactant 5 g/L

To

tal

ph

en

an

thre

ne

re

ma

inin

g (

mg

/kg

)

Figure 4.8 Total remaining phenanthrene in soil slurries after 10 days of bioremediation, results

presented as phenanthrene remaining in mg/kg of dry soil.

Page 96: Avery Gottfried - ME thesis 2009

87

4x107 cfu/mL occurred in all samples with no biosurfactant. This suggests salicylate

improved degradation through means other than increased quantities of live cells. All live

cell counts were in the range of 1x107 to 1x10

9 cfu/mL with biosurfactant generating

positive effects on the total cfu/mL when glucose was present. Biosurfactant caused no

significant change on the amount of cfu/mL in systems with salicylate. Theses results

support a hypothesis by Johnsen et al. (2005) that PAH degrading population in soil are

mostly not growing, and cells are in a pseudo-stationary phase where transient growth only

replaces decaying and washed out cells until the habitat’s mass transfer-controlled carrying

capacity is reached again .

The growth characteristics correspond to the trends observed in the concentration of

phenanthrene remaining in the aqueous phase (Figure 4.7). There appeared to be more

growth present in glucose systems as biosurfactant concentrations increased from 0 to 5

g/L, and this corresponds to the lower amount of phenanthrene present in aqueous solution

for the same increase in biosurfactant concentration. There was roughly the equivalent

phenanthrene concentration in all systems with salicylate regardless of the amount of

biosurfactant present, and this corresponds to the equivalent live cell concentration

between 5x107 and 1x10

8. Powell et al. (2008) observed a similar trend where salicylate

degrading bacteria increased in abundance substantially after enrichment by continuous

addition of salicylate in batch cultures but did not increase in abundance in response to the

spike addition. However Powell et al. (2008) suggested that enrichment with salicylate can

select for naphthalene-degrading bacteria, but does not select for organisms responsible for

degrading PAHs of higher molecular weight, as phenanthrene and benzo[a]pyrene

degradation where not enhanced. This result actually depends on the strain of bacteria

Page 97: Avery Gottfried - ME thesis 2009

88

present and the metabolic pathway that exists in the bacteria, as it has been observed in

this research that salicylate can enhance phenanthrene degradation in cultures that are

capable of both phenanthrene and naphthalene degradation. The ability to induce the

metabolism or cometabolism of a target compound can be a sufficient bioremediation

strategy. However, information about the metabolic pathways intermediates that induce

the catabolic enzymes being utilized by the bacteria are required.

Figure 4.9 Live cell counts (cfu/mL) taken from soil slurry solution over 10 days. Results

presented are averages from duplicate or triplicate plate counts.

1.00E+06

1.00E+07

1.00E+08

1.00E+09

0 2 4 6 8 10

Pse

ud

om

on

as

Pu

tid

a 1

74

84

cfu

/mL

Time Days

BHB BHB SALICYLITE BHB GLUCOSE

No Biosurfactant

0 2 4 6 8 10Time Days

5 g/L Biosurfactant

1.00E+06

1.00E+07

1.00E+08

1.00E+09

0 2 4 6 8 10

Pse

ud

om

on

as

Pu

tid

a 1

74

84

cfu

/mL

Time Days

BHB BHB SALICYLITE BHB GLUCOSE

1g/L Biosurfactant

0 2 4 6 8 10Time Days

0.25 g/L Biosurfactant

Page 98: Avery Gottfried - ME thesis 2009

89

Research by Grimm and Harwood (1997) with Pseudomonas Putida G7 and strain NCIB

9816-4 indicated that salicylate, the compound that directly induces naphthalene

degradation, was also an inducer of naphthalene chemotaxis. This gene along with the

naphthalene degrading genes is encoded in the NAH7 plasmid. Chemotaxis enhances the

ability of motile bacteria, such as Pseudomonas Putida ATCC 17484, to locate and degrade

organic compounds. It is probable that this process can be used by bacteria to move

towards phenanthrene, as a common metabolic pathway exists for the degradation of both

naphthalene and phenanthrene in P.Putida. The improved phenanthrene removal observed

in systems with salicylate did not show a relative increase in live cell concentration (cfu/mL),

however increases in phenanthrene degradation were observed. This can be explained by

the induction of chemotactic behaviour due to salicylate addition. Chemotactic responses

require a concentration gradient of the attractant for a response to occur, and the greater

the concentration gradient the more effective a chemotactic response could be in

enhancing the degradation of the contaminant (Samanta, Singh et al. 2002). Since

biosurfactant did not enhance the degradation of phenanthrene in the presence of

salicylate, this could be due to the increased mass transfer of phenanthrene into the

aqueous phase which decreased the concentration gradient. This would negatively impact

the advantage of chemotactic responses as concentration gradients in the systems were

decreased due to biosurfactant addition. The result of chemotactic attraction can lead to an

increase in phenanthrene bioavailability due to bacteria migrating towards high

concentrations and result in the observed biodegradation rate increasing. The strain of

P.Putida used in this research is capable of attaching to solid surfaces such as soil or solid

phase PAHs and using the nutrients and contaminants directly, indicating that solubility

enhancement of phenanthrene is only one of the mechanisms available to enhance the

Page 99: Avery Gottfried - ME thesis 2009

90

degradation rate (Dean, Jin et al. 2001). Similar results demonstrating the chemotactic

response have been obtained in diffusion limited systems with naphthalene (Marx and

Aitken 2000; Ortega-Calvo, Marchenko et al. 2003). The positive chemotaxis of P.putida

towards naphthalene due to the metabolic induction by salicylate is most likely extended to

the degradation of phenanthrene, as this may account for the increased removal that is

achieved with salicylate addition.

Results from soil slurry tests are similar to results obtained in aqueous tests (section 4.1.4).

A greater amount of phenanthrene degradation could be achieved through amendments

that specifically induce metabolic pathway induction, when compared to increased solubility

brought about by the addition of biosurfactant, or increased growth brought about by

additional carbon substrates. Biosurfactant addition showed improved phenanthrene

removal when compared to the system with no amendments. Increasing biosurfactant

addition from 0.25 g/L to 5 g/L did not produce a large benefit for the 20-fold increase in

quantity applied. The overall results after 10 days showed nearly complete phenanthrene

removal in most systems; therefore it is not possible to determine the rate at which specific

amendments enhanced the biodegradation in soil slurries. This information would be

important to asses the overall efficiency of each amendment, as some systems could have

achieved the quantity of phenanthrene removal observed in a shorter time frame than the

10 day results studied.

Page 100: Avery Gottfried - ME thesis 2009

91

4.3 OBJECTIVE 3: COLUMN TESTS

4.3.1 TRACER AND BIOSURFACTANT BREAKTHROUGH CURVE FITTING

The steady-state breakthrough curves for conservative tracer Cl- were analyzed with local

equilibrium convective-dispersive transport model to determine the dispersion (D) and

retardation (R) transport parameters using the computer program STANMOD (Simunek, M.

Th. van Genuchten et al. 2003). STANMOND is a windows based user platform that

incorporates the CXTFIT 2.0 mathematical models to evaluate solute transport in porous

media using analytical solutions to the convection-dispersion solute transport equation

(Toride, Leij et al. 1995). The retardation factor for the rhamnolipid biosurfactant was

determined from the frontal limb of the rhamnolipid breakthrough curve following a

method presented by Noordman et al.(1998).

4.3.1.1 Tracer Breakthrough Curves

Chloride breakthrough curves at a steady-state flow of 0.2 mL/min (pore water velocity =

1.542 cm/h) were fitted using the deterministic equilibrium CDE to determine the simple

hydrodynamic characteristics of the soil columns (Figure 4.10a). The symmetrical

breakthrough

curves obtained with the conservative tracer showed no

physical non-

equilibrium and were used to estimate dispersivity (D), with an average fitted value of 34

cm2/h, with R equal to 1. The chloride breakthrough curves fitted well with the CDE model

parameters, indicating the column system was stable and the hydrodynamic conditions

were similar in separately packed columns.

Chloride breakthrough curves at a high flow of 10 mL/min (pore water velocity = 77.87

cm/h) were fitted using the same deterministic equilibrium CDE to determine the

Page 101: Avery Gottfried - ME thesis 2009

92

hydrodynamic characteristics of the soil columns (Figure 4.10b). Breakthrough curves

obtained with the conservative tracer showed slight tailing, and the average fitted

dispersion (D) was 749 cm2/h and R equal to 0.9. The curve could not be fitted if R was equal

to 1, and R equal to 0.9 was the best parameter estimate to fit the observed data. This

indicates at high flow rates, some short circuiting of the tracer flow might have occurred at

the column inlet reducing the effective flow length in the soil column. The dispersion to

velocity difference increased the Peclet number to 3.84, compared to 1.68 for the lower

flow rate indicating more advective transport, and less dispersive transport is responsible

for the tracer movement at a high flow rate.

4.3.1.2 Biosurfactant breakthrough

Compared to chloride breakthrough curves, biosurfactant breakthrough produced an

asymmetrical curve. After 10 PV of pumping the biosurfactant concentration in the effluent

was 70% of that in the influent (Figure 4.10c). It is evident from the gradual rising limb of the

biosurfactant breakthrough curve that significant adsorption had occurred, causing

retardation of the biosurfactant flow. Equilibrium CDE models were not able to simulate the

data. A non-equilibrium CDE model assuming one-site chemical adsorption obtained the

best fit to the rising limb of the biosurfactant breakthrough curve. The rapid drop in the

tailing limb could not be fitted to the non-equilibrium model, and data was not collected for

the extent of tailing that occurred for the breakthrough curve, therefore the tailing limb was

not modelled and information about the entire transport process could not be obtained.

Using this method D equalled 34 cm2/h, R equalled 6.59, ω equalled 0.067, and β equalled

0.15.

Page 102: Avery Gottfried - ME thesis 2009

93

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 1 2 3 4 5 6

C/C

o

Pore Volume

Cobs(x,t) High Flow

Cfitted(x,t) High Flow

v = 77.87 cm/h

D = 749 cm2/h

R = 0.9

(b)

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 5 10 15 20

C/C

o

Pore Volume

Cobs(x,t) Biosurfactant

Cfitted(x,t) Biosurfactant

v = 1.54 cm/h

D = 34 cm2/h

R = 6.59

(c)

Figure 4.10 Observed breakthrough curves and fitted breakthrough curve models

using CXTFIT inverse parameter estimation for (a) chloride with v = 1.54cm/h (b)

chloride with v = 77.87cm/h and (c) biosurfactant with v = 1.54 cm/h

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 1 2 3 4 5 6

C/C

o

Pore Volume

Cobs(x,t) Low Flow

Cfitted(x,t) Low Flow

v = 1.54 cm/h

D = 34 cm2/h

R = 1.0

(a)

Page 103: Avery Gottfried - ME thesis 2009

94

4.3.2 MICROBUBBLE DISPERSION FLOW CHARACTERISTICS

Stability of the microbubble dispersions improved with increasing biosurfactant

concentrations from 1 g/L to 5 g/L. The increased stability can be identified by the longer

half drainage times (data not shown) and the increased gas hold-up, which describes the

fraction of the microbubble dispersion that is in gas phase versus liquid phase. Based on the

collection of effluent from the column during the microbubble pumping experiments there

was an average of 30% liquid, and 70% gas breakthrough from the 1 g/L biosurfactant

microbubble dispersion, and 24% liquid, and 76% gas breakthrough from the 5g/L

biosurfactant microbubble dispersion.

Visual observations of the microbubble dispersion saw separation of the dispersion into

liquid and gas phase directly after injection into the column. There were only a few short

instances over the four hours of pumping that visible microbubbles could be observed a few

centimetres beyond the inlet of the column. This could be due to a variety of factors

including: microbubble stability; low flow rate of the microbubble dispersion versus half-life

of the suspension; interaction of the microbubble dispersion with dissolved organic matter;

and the sorption of surfactant onto the soil. In a separate test microbubble dispersions were

pumped at a high flow rate (1PV in 3 minutes) through a small (volume = 10 cm3) column

containing clean washed medium grain sand. The microbubble dispersion transported

through the column without separating into liquid and gas phases and microbubbles were

collected in the effluent. Oliveira et al. (2004) determined the presence of fine particles and

hydrophobic fine particles interfere with the foam breaking process, and with the solution

foaming ability. Sorption of biosurfactant also leads to a reduction of the surfactant

concentration which in turn would decrease the foam stability (Oliveira, Oliveira et al. 2004).

Page 104: Avery Gottfried - ME thesis 2009

95

Increasing surfactant concentration during continuous biosurfactant pumping initially

increases the adsorption of biosurfactant on the particle surfaces rendering them less

hydrophobic and would decrease the foam breaking ability over time

The effluent liquid flow rate occurred at a constant rate throughout the experiment. The

liquid front was observed to be moving slower than the gas phase, as gas breakthrough in

the effluent would occur within minutes of microbubble injection. To determine the liquid

flow characteristics a conservative tracer (chloride) was added to the microbubble

dispersion during generation and effluent was collected to determine the breakthrough

curve for the liquid fraction of the microbubble dispersion. The pumping flow rate was set

to 10cm3/min per minute, which describes the flow rate of the microbubble dispersion

including both the gas and liquid phases. The equivalent liquid flow rate based on the

quantity of liquid present in the microbubble dispersion was 2.37 mL/min for the 5 g/L

microbubble dispersion. Using the liquid fraction flow rate, data was modelled for the 5g/L

microbubble dispersion with the equilibrium CDE and R equall to 1.37, and D equall to 32.1

cm2/h producing the best model (Figure 4.11).

Visualisation of microbubble dispersion flow patterns by Choi et al. (2008) in sand media

showed chemical surfactant sodium dodecyl sulphate microbubble suspensions separated

into liquid and gas phase directly after injection. The liquid phase showed faster movement,

whereas the gas from of the microbubble suspension flowed in a plug-flow manner (Park,

Choi et al. In Press). The liquid fraction did not advance ahead of the gas phase in these

experiments, but the flow was in a low dispersion (D = 32.1 cm2/h) plug-flow manner. The

gas fraction was considered an immobile fraction (decreasing the relative permeability) and

Page 105: Avery Gottfried - ME thesis 2009

96

characteristic break through curves were obtained (Figure 4.11). This behaviour is consistent

with foam drainage in porous media where flow of the liquid fraction is through the

interstices of the immobile foam (Zitha, Nguyen et al. 2006). Zitha et al. (2006) concluded

the role of trapped gas during foam flow was not as important as it might have been

believed previously by most foam flow researchers. The microbubble dispersions in this

research acted more like standard foam where the process of breakage of the foam and

regeneration caused a relatively large decrease in liquid mobility due to foam development

(Chowdiah, Misra et al. 1998).

0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

1

0 0.5 1 1.5 2 2.5 3 3.5 4 4.5 5

C/C

o

Pore Volume

Cobserved(x,t) microfoam 1 g/L

Cobserved(x,t) microfoam 5 g/L

Cfitted(x,t) microbublle dispersion 5g/L

v = 18.4 cm/h

D = 32 cm2/h

R = 1.37

Figure 4.11 Microbubble dispersion breakthrough curve with a conservative tracer in the liquid

fraction.

Page 106: Avery Gottfried - ME thesis 2009

97

4.3.2 PRESSURE DROP ASSOCIATED WITH MICROBUBBLE DISPERSION PUMPING

The flow propagation of the microbubble dispersion was accompanied by a large pressure

drop greater than 50 kPa after 10 pore volumes pumping with 5 g/L biosurfactant

microfoam (Figure 4.12, Figure 4.13, & Figure 4.14). The results from the pumping show the

increase in biosurfactant concentration from 1g/L to 5g/L cause pressure build up in the soil

after 10 PV to be nearly 40% higher at the inlet. This is possibly due to the increased stability

and stronger films that make up foam as the biosurfactant concentration increased. This

creates more effective displacing or blocking agents in the porous media which eventually

creates pressure build up in the soil column. A variety of measures including foam quality,

flow rate, and surfactant type have shown to greatly influence the magnitude of the

pressure drop that is associated with foam flow. With standard foam it is assumed the

presence of the gas phase makes foams more compressible. Higher quality foam will have

larger bubbles and thinner liquid films which causes the pressure gradients in soil to

decrease as the foam quality is increased (Chowdiah, Misra et al. 1998).

Pressure profiles (Figure 4.13; Figure 4.14) demonstrate the pressure gradient due to the

presence of foam at the inlet was significantly higher than the pressure in the remainder of

the column. The pressure drop was significantly lower from 17cm depth to the surface, and

this small pressure drop corresponds to single phase water flow, versus larger pressure

drops creted by foam flow (Apaydin and Kovscek 2001). The retarded flow of the

microbubble dispersion resulted in larger pressure drops which increased over the duration

of the experiment. When the pressure drop builds up in a channel, the foam flows into less

accessible spill areas. This pressure dependent “clogging” process indicates that channelling

or poor sweep efficiency should not occur, allowing microbubbles to overcome subsurface

Page 107: Avery Gottfried - ME thesis 2009

98

heterogeneity. This process also causes increases in pressure that can be undesirable if to

large (Riser-Roberts 1998). It is considered necessary to maintain the pressure drop below

22.6 kPa/m to prevent mechanisms such as soil heaving occurring due to the injection of

foam at high pressure (Chowdiah, Misra et al. 1998).

0

10

20

30

40

50

60

0 2 4 6 8 10 12

Pre

ssu

re (

kP

a)

Pore Volumes

Biosurfactant 1g/L solution: Flow = 10 mL/min

Inlet -37cm

-34.5cm

-32cm

-29.5cm

-27cm

-17cm

Figure 4.12 Pressure distribution across the length of the soil column during biosurfactant (1

g/L) solution pumping. Data presented corresponds to depth in the column with the highest

pressure at the inlet, and the lowest pressure at -17cm assuming the outlet is the datum.

Page 108: Avery Gottfried - ME thesis 2009

99

0

10

20

30

40

50

60

0 1 2 3 4 5 6 7 8 9 10 11 12

Pre

ssu

re k

Pa

Pore Volumes

Biosurfactant 5 g/L microfoam: Flow =

10cm3/minInlet -37cm

-34.5cm

-32cm

-29.5cm

-27cm

-17cm

0

10

20

30

40

50

60

0 1 2 3 4 5 6 7 8 9 10 11 12

Pre

ssu

re k

Pa

Pore Volumes

Biosurfactant 1 g/L microfoam: Flow =

10cm3/min

Inlet -37cm

-34.5cm

-32cm

-29.5cm

-27cm

-17cm

Figure 4.14 Pressure distribution across the length of the soil column during biosurfactant

microfoam ( 5g/L) pumping. Data presented corresponds to depth in the column with the highest

pressure at the inlet, and the lowest pressure at -17cm assuming the outlet is the datum.

Figure 4.13 Pressure distribution across the length of the soil column during biosurfactant

microfoam (1 g/L) pumping. Data presented corresponds to depth in the column with the highest

pressure at the inlet, and the lowest pressure at -17cm assuming the outlet is the datum.

Page 109: Avery Gottfried - ME thesis 2009

100

4.3.3 BIODEGRADING TRIALS IN CONTINUOUS FLOW SYSTEMS

Biodegradation trials in pre-inoculated soil columns were operated as a continuous flow

system over a 10 day period. Soil profiles of the remaining phenanthrene in soil over the

column depth indicate phenanthrene mobilization and transport due to the upward flow of

biosurfactant solution (Figure 4.15; Figure 4.16). Inlet concentrations of phenanthrene were

up to 2-fold lower than the concentration remaining in the upper half of the soil column.

Significant dissolved organic matter was present in the effluent as TOC readings averaged

near 50 mg/L, and there was a visible brown colour to the effluent samples.

-40

-35

-30

-25

-20

-15

-10

-5

0

0 50 100 150 200 250 300 350 400 450 500

De

pth

in

co

lum

n (

cm)

Phenanthrene remaining in soil (mg/kg)

Column 1

Column 2

0

5

10

15

20

25

30

35

40

45

0 50 100 150 200 250

Ph

en

an

the

re m

g/L

Time Hours

Phenanthrene in Efluent

Figure 4.15 Trial 1 phenanthrene distribution in soil column after 10 days continuous upflow at

0.2 mL/min , phenanthrene in column effluent over 10 days. Column 1 influent solution

biosurfactant 1g/L + salicylate 100 mg/L; Column 2 influent solution biosurfactant 1 g/L. Soil

distribution assuming effluent (0cm) is the top of the column and influent (-37 cm) is the bottom

of the column.

Page 110: Avery Gottfried - ME thesis 2009

101

Four pore volumes of continuous biosurfactant pumping occurred before notable amounts

of phenanthrene were detected in the effluent solution (Figure 4.15). After 4 days the

amount of phenanthrene detected in the effluent significantly increased to over 30 mg/L.

This indicates that biosurfactant sorption to the soil was occurring for the first 4 PVs and not

contributing to an increase in the apparent aqueous phase phenanthrene concentration.

Depending on the quantity of biosurfactant added to the system, the sorbed-phase

biosurfactant can account for the majority of the added biosurfactant (Laha, Tansel et al. In

Press). The result of this process indicates there is increased partitioning of phenanthrene

-40

-35

-30

-25

-20

-15

-10

-5

0

0 50 100 150 200 250 300 350 400 450

De

pth

in

co

lum

n (

cm)

Phenanthrene remaining in soil (mg/kg)

Column 1

Column 2

Figure 4.16 Trial 2 phenanthrene distribution in soil column after 10 days continuous upflow

with BHB broth at 0.5mL/min; phenanthrene in column effluent over 10 days. Column 1

influent pulse solution biosurfactant microfoam 1 g/L + salicylate 100mg/L; Column 2 influent

pulse solution biosurfactant 1 g/L. Soil distribution assuming effluent (0cm) is the top of the

column and influent (-37 cm) is the bottom of the column.

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 48 96 144 192 240

Ph

en

an

the

re m

g/L

Time Hours

Phenanthrene in Efluent

Pulse injections

Page 111: Avery Gottfried - ME thesis 2009

102

onto soil, until the biosurfactant is at a high enough concentration above the CMC in the

aqueous phase to promote phenanthrene desorption from the surfactant sorbed soil (Laha,

Tansel et al. In Press). The amount of biosurfactant sorption occurring for the first 4 days

most likely decreased the overall biosurfactant concentration below the CMC. After this

period the rate of sorption to the soil decreased and the concentration became high enough

in solution to exceed the CMC. The amount of phenanthrene in the effluent samples from

continuous pumping without biosurfactant (Figure 4.16) maintained a relatively constant

aqueous phenanthrene concentration near 1mg/L with some noted increases to 1.2mg/L

after pulse injection of both the microbubble dispersion and biosurfactant solution.

It has been demonstrated that the application of surfactant enhanced soil washing or

flushing of PAH contaminants generally does occur until after the sorption of surfactant on

the soil reaches saturation (Zhou and Zhu 2007; Zhu and Zhou 2008). No advantageous

results are demonstrated until after the sorption of surfactant on the soil reaches

saturation. This effect was evident in the biosurfactant pumping trials, indicating that higher

flow rate would be beneficial in the initial stages of a biosurfactant enhanced remediation

process to speed up the sorption process. The adsorption and desorption process appears

to be a non-equilibrium process where two-site sorption models can account for the

sorption behaviour of both PAHs and surfactants (Noordman 1999; Chen, Wang et al. 2006).

In surfactant flushing experiments, after biosurfactant sorption reaches saturation, the

removal of fast desorbing phenanthrene fractions can occur quickly, and the slow desorbing

fractions could be efficiently removed at very low purging rates (Schlebaum, Schraa et al.

1999). Overall, results of this study demonstrate that surfactant sorption to the solid phase

can lead to increases in phenanthrene retardation when biosurfactant concentrations are

Page 112: Avery Gottfried - ME thesis 2009

103

low. This effect would be desirable if the treatment objective was to immobilize

phenanthrene; however, the effect is undesirable in surfactant enhanced phenanthrene

removal applications (Ko, Schlautman et al. 1998).

Table 4.3 Total percentage removal of phenanthrene due to soil flushing and

biodegradation in soil column tests after 10 days continuous flow.

Trial 1: Continuous Biosurfactant Flushing (0.2mL/min)

Column 1: biosurfactant and salicylate Column 2: biosurfactant only

Soil Flushing 10.8% ± 0.1% 7.9% ± 0.1%

Biodegradation -0.7% ± 7.4 % 10.5% ± 6.9%

Biodegradation/hr --- 0.20 mg phenanthrene/hr

Total Removal 10.1% ± 5.4% 18.4% ± 4.8%

Trial 2: Biosurfactant and Salicylate Pulse Injection (0.5mL/min)

Column 1: microfoam pulse Column 2: liquid pulse

Soil Flushing 1.4% ± 0.1% 1.4% ± 0.1%

Biodegradation 27.8% ± 7.7% 21.6% ± 7.2%

Biodegradation/hr 0.51 mg phenanthrene/hr 0.40 mg phenanthrene/hr

Total Removal 29.3% ± 5.8% 23.0% ± 5.2%

Mass balance calculations were performed to determine the amount of biodegradation that

occurred (Table 4.3). The total amount of biodegradation was calculated by taking the total

initial phenanthrene amount, subtracted by the total remaining phenanthrene after 10 days

and the total amount of phenanthrene that was present in the effluent due to soil flushing.

With an average of 5% error on soil extractions (both to obtain the initial soil contamination,

and to obtain the total remaining concentration) the error becomes over 7% when the mass

balance calculations are performed. This shows accuracy of the soil extraction method

strongly influences the overall accuracy of the results obtained. However flushing data

obtained on the HPLC had very little error as phenanthrene is already in the aqueous phase

and easily quantified. This highlights the importance in determining accurate soil extraction

methods to obtain statistically significant results, and the importance conducting tests in

triplicate or greater in order to report experimental results with certainty.

Page 113: Avery Gottfried - ME thesis 2009

104

The results from the second column trial showed cell elution from column 1 averaging an

order of magnitude larger than column 2 with an average of 5x107, versus 5x10

6 cfu/mL

respectively, for samples taken from the effluent at the same time as phenanthrene

measurements. The final amount of cells in the soil when unpacking was completed

followed the same trend, where there was approximately one order of magnitude more

cells present in column 1. Significant cell elution was observed by Yolcubal et al.(2002) in

sand column tests which increased in the presence of substrate in the influent solution.

These results suggest that experiments which exhibited the greatest degradation were

associated with the production of new cells both in the soil and in the aqueous phase.

Degradation rates in the soil columns were much lower than phenanthrene degradation in

soil slurry batch tests indicating that contaminant mass transfer and quantity of available

oxygen would be responsible for the differences observed. Park et al. (2001) determined

that batch systems degradation followed zero order kinetics and was independent of the

concentration range, while column tests exhibited decreased rates at concentrations less

than 100 mg/L of naphthalene. Half-saturation constant Ks, as described by Michaelis-

Mention and Monod growth kinetics (Equation 2.1) were elevated in column tests due to

length dependent transfer of substrate to cell surfaces in column tests. Whereas kinetic

parameters in well mixed batched systems create shorter and more uniform mass transfer

distances (Park, Zhao et al. 2001). This can also be explained using non-equilibrium

conditions present in diffusion control sorption process. Vigorous shaking allows diffusion to

generally be eliminated as a rate-limiting step in batch studies (Nielsen, Van Genuchten et

al. 1986). In soils with flowing water, the sorption rate can be limited by the rate at which

ions are transported to the exchange sites, and can be determined by physical non-

Page 114: Avery Gottfried - ME thesis 2009

105

equilibrium models (Nielsen, Van Genuchten et al. 1986). Physical non-equilibrium models

use the same formulation as the chemical non-equilibrium models presented, (Equation 2.4

& 2.5) but two-site partitioning in soil water phases are defined as mobile (flowing) and

stagnant (immobile) phases (Nielsen, Van Genuchten et al. 1986). This assumes the

convective dispersive transport is confined to the faction of the liquid filed mobile pores,

and the stagnant non moving liquid decreases the overall mass transfer rate (Nielsen, Van

Genuchten et al. 1986).

Oxygen appears to have been a rate limiting amendment in soil columns tests. The

increased flow rate and microfoam pulse increased the total amount of degradation (Table

4.3). The system with microfoam pulse injection saw 27.8% degradation versus 21.6% in the

same setup with a liquid pulse instead of a microfoam pulse injection every 48 hours. The

low flow rate saw only 10.5 % degradation with continuous biosurfactant supply and no

degradation with biosurfactant and salicylate addition. These results are contrary to what

were expected based on preliminary batch soil slurry tests results (Section 4.2.2). The

explanation for the observed biodegradation rates can be explained assuming oxygen was

the limiting amendment. Added salicylate at a low flow rate in the aqueous phase would

have been preferentially used by the bacteria as an available carbon source. This amount of

added bacterial activity due to the salicylate degradation would have decreased the oxygen

concentration prior to any phenanthrene degradation, resulting in no phenanthrene

degradation due to oxygen limitations. The increased flow rate in the second trial, delivered

increased oxygen to the system which resulted in higher total amounts of degradation, and

microfoam addition further increased the added oxygen to the system resulting in 27.8%

degradation.

Page 115: Avery Gottfried - ME thesis 2009

106

CHAPTER 5 CONCLUSIONS AND RECOMENDATIONS

The focus of this research was to enhance in situ biodegradation of phenanthrene in soil. In

order to determine why specific physical variations have been shown to increase (or

decrease) overall contaminant biodegradation, bench-scale systems were created to

quantify the effects of rhamnolipid biosurfactant, salicylate, glucose, and biosurfactant

microbubble dispersions on the overall biodegradation process. The conclusions of these

experiments are briefly described below.

Liquid Cultures

• Higher levels of phenanthrene degradation can be achieved through amendments

that target metabolic pathway induction, as phenanthrene removal was doubled

upon addition of salicylate, as compared to the addition of glucose.

• The addition of rhamnolipid biosurfactant increases the apparent aqueous solubility

of phenanthrene, and augmented the amount of degradation in all systems in

combination with added salicylate and glucose.

Soil Slurries

• Biosurfactant addition caused competition between micellar and sorbed

biosurfactants for phenanthrene partitioning. The addition of rhamnolipid

biosurfactant above the CMC enhanced the desorption of phenanthrene from soil,

decreasing Kd over 8-fold from the addition of 1 g/L biosurfactant.

• Biosurfactant additions improved biodegradation by reducing both the aqueous

phase phenanthrene, and total remaining phenanthrene in soil slurries with at least

86% total phenanthrene removal in systems containing 0.25g/L biosurfactant versus

67% removal without biosurfactant .

Page 116: Avery Gottfried - ME thesis 2009

107

• With salicylate present, the effects of biosurfactant were almost nonexistent as

greater than 90% removal occurred in all systems containing salicylate regardless of

the biosurfactant concentration. Although bioavailability is commonly perceived as

the rate limiting step, an in situ enhancement strategy does not need to increase

bioavailability via increased aqueous solubility by adding a biosurfactant (Shin, Kim

et al. 2004).

• The positive chemotaxis of P.putida towards naphthalene, due to the metabolic

induction by salicylate, is most likely extended to the degradation of phenanthrene,

as this accounts for the increased removal achieved with salicylate present.

• Chemotactic attractive behaviour of P.putida towards phenanthrene leads to an

increase in phenanthrene bioavailability. As a result, the observed biodegradation

rates increased.

Column Systems

• Non-equilibrium CDE models assuming one-site chemical adsorption obtained the

best fit to the rising limb of the biosurfactant breakthrough curve. Adsorption and

desorption appears to be a non-equilibrium process where two site sorption models

in CXTFIT can account for the sorption behaviour of both PAHs and surfactants

(Noordman 1999; Chen, Wang et al. 2006).

• Degradation rates in the soil columns were much lower than phenanthrene

degradation in soil slurry batch tests. Enhancement strategies that increased delivery

of oxygen to the system result in higher total amounts of degradation. Other

amendments have little effect when oxygen is limiting the system.

Page 117: Avery Gottfried - ME thesis 2009

108

Understanding the metabolic pathways and the enzymatic reactions utilized by

Pseudomonas putida ATCC 17484 during contaminant breakdown yields information about

the induction ability of salicylate. Liquid culture experiments were relatively simple and less

time consuming, and provided good preliminary data to asses the viability of amendments.

Soil organic matter affected the efficiency of biosurfactant addition due to partitioning

processes, demonstrating the importance of maintaining a biosurfactant concentration in

solution above the CMC to enhance the overall phenanthrene desorption from the soil.

Results from these trials provide further evidence to the importance of determining the

metabolic processes that are responsible for successful in situ bioremediation, as salicylate

proved to be the best amendment in non-oxygen limiting environments.

Page 118: Avery Gottfried - ME thesis 2009

109

5.1 RECOMMENDATION FOR FUTURE WORK

Due to the observations of microbubble dispersion separation into liquid and gas phases

upon injection, it is apparent that microbubble dispersion flow is dependent on the

conditions of the soil and the stability of the microfoam generated. Microbubble stability,

low flow rates of dispersion, half life of the microbubble suspension, interactions between

the microbubble dispersion and the dissolved organic matter, as well as sorption of

surfactant to soil could negatively impact the effectiveness of microbubble dispersions as

amendments to in situ bioremediation. Further studies detailing the microbubble dispersion

flow in simple sand, versus systems with higher amounts of organic materials and fine

particulates, would help determine the efficiency of microbubble dispersions.

A limitation to the studies performed in this research was the continuous operation of the

microbubble generating system. The spinning disk motor could not be run for longer than

four hours without overheating, which limited the scope of the research trials significantly.

Improved methods for generating the microbubble dispersion, such as methods that

encapsulate the spinning disk apparatus inside a pressure vessel have been shown to create

microfoam that is more stable and can be used in continuous operation (Wan, Veerapaneni

et al. 2001). Microbubble generation and injection under pressure have been shown to

minimize microbubble loss due to gas dissolution (Wan, Veerapaneni et al. 2001). A

comparison of Generation methods of sonication versus mechanical agitation showed that

the generation method greatly affected the size of microbubbles. Therefore the use of a

suitable generation method should be an important consideration for future work (Xu,

Nakajima et al. 2008).

Page 119: Avery Gottfried - ME thesis 2009

110

The length of time required for effective biodegradation experiments in soil (and application

in the field) is difficult to determine. Although batch tests showed steady phenanthrene

removal, scaling up to column tests requires longer project timelines depending on the

amendments made. This constrains the number of trials that can be run, the number of

variables that can be monitored, and therefore the amount of data that can be collected

obtained. Developing or using existing modelling software is an important tool to simulate

transport parameters, estimate biodegradation rates, and determine effective treatment

strategies. It would also be beneficial to characterize microbubble flow based on the

predicted pressure build up in the soil. This would determine oxygen delivery, and could be

used to determine the amount of oxygen available for uptake by bacteria.

The overall results in soil slurries after 10 days of biodegradation showed nearly complete

phenanthrene removal in most systems. Therefore it is not possible to determine the rate

at which each amendment enhanced the biodegradation in soil slurries. This info is essential

to asses the overall efficiency of each amendment, as some systems may have achieved the

same quantity of phenanthrene removal in shorter time frames. The recent development of

an HPLC method to accurately determine rhamnolipid biosurfactant concentration will also

be an important tool in future research, allowing the quantification of biosurfactant

sorption and its utilization by microbial communities.

Contaminant bioavailability can be species-specific, with different microbial strains capable

of accessing different contaminant pools in the soil-water system. Thus, understanding the

interactions between microbes will provide further opportunities for enhanced

biodegradation of bioavailable contaminants (Dean, Jin et al. 2001). The recent acquisition

Page 120: Avery Gottfried - ME thesis 2009

111

of a variety of PAH degrading bacterial strains (graciously donated by Landcare Research)

which were characterized from contaminated sites in New Zealand, will allow for the

development of more complex systems. This will enable the study of interactions between a

variety of bacteria and their model environments. Research with different strains to

determining metabolic influences will be an important tool to use, and gain an

understanding of how to control subsurface environments with amendments that target

known metabolic pathway induction.

Work is underway to develop fluorescent protein reporter organisms capable of degrading

target contaminants. This will allow population growth dynamics to be monitored in situ

with a fiberoptic spectroscopic probe system with fluorescence linked to nah operon

promoter activity. This technique will offer unique opportunities to monitor bacterial

growth characteristics in a non-destructive manner which is linked to salicylate

mineralization.

The results of this research indicate that a sound understanding of the microbial processes

that occur, aid in determining the most efficient strategy to enhance the in situ

biodegradation process. A better understanding of metabolic pathways and inducers for

degradation of higher-ringed PAHs for more bacterial strains in the environment will result

in the identification of rate-limiting steps in the process. This will help future researchers

target specific enhancement strategies to make in situ biodegradation more reliable and

efficient. Microbubble dispersions show promise for effective oxygen delivery to the

subsurface and will help to extend the feasibility of in situ bioremediation to more complex

systems.

Page 121: Avery Gottfried - ME thesis 2009

112

CHAPTER 6 WORKS CITED

Aitken, M. D., W. T. Stringfellow, et al. (1998). "Characteristics of phenanthrene-degrading bacteria

isolated from soils contaminated with polycyclic aromatic hydrocarbons." Canadian Journal

of Microbiology 44(8): 743-752.

Allen, C. C. R., D. R. Boyd, et al. (1999). "Contrasting Effects of a Nonionic Surfactant on the

Biotransformation of Polycyclic Aromatic Hydrocarbons to cis-Dihydrodiols by Soil Bacteria."

Appl. Environ. Microbiol. 65(3): 1335-1339.

Apaydin, O. G. and A. R. Kovscek (2001). "Surfactant Concentration and End Effects on Foam Flow in

Porous Media." Transport in Porous Media 43(3): 511-536.

Atlas, R. and J. Philip (2005). Bioremediation - applied microbial solutions for real-world

environmental cleanup, ASM press.

Auckland City Council. (2007). "Urban Soils Review." Retrieved December 21, 2008, from

http://www.aucklandcity.govt.nz/council/documents/soilquality/default.asp.

Avramova, T., A. Sotirova, et al. (2008). "Effect of Triton X-100 and rhamnolipid PS-17 on the

mineralization of phenanthrene by Pseudomonas sp. cells." International Biodeterioration &

Biodegradation 62(4): 415-420.

Basu, A., D. Das, et al. (In Press). "Sequential utilization of substrates by Pseudomonas putida CSV86:

Signatures of intermediate metabolites and online measurements." Microbiological

Research In Press, Corrected Proof.

Buco, S., M. Moragues, et al. (2004). "Analysis of polycyclic aromatic hydrocarbons in contaminated

soil by Curie point pyrolysis coupled to gas chromatography-mass spectrometry, an

alternative to conventional methods." Journal of Chromatography A 1026(1-2): 223-229.

Cameotra, S. S. and J.-M. Bollag (2003). "Biosurfactant-Enhanced Bioremediation of Polycyclic

Aromatic Hydrocarbons." Critical Reviews in Environmental Science and Technology 30(2):

111-126.

Cerniglia, C. E. (1993). "Biodegradation of polycyclic aromatic hydrocarbons." Current Opinion in

Biotechnology 4(3): 331-338.

Chauhan, A., Fazlurrahman, et al. (2008). "Bacterial metabolism of polycyclic aromatic hydrocarbons:

strategies for bioremediation." Indian Journal of Microbiology 48(1): 95-113.

Chen, J., X. J. Wang, et al. (2006). "Modeling surfactant las influenced PAHs migration in soil

column." Water, Air, and Soil Pollution 176(1-4): 217-232.

Chen, S.-H. and M. D. Aitken (1999). "Salicylate Stimulates the Degradation of High-Molecular

Weight Polycyclic Aromatic Hydrocarbons by Pseudomonas saccharophila P15."

Environmental Science & Technology 33(3): 435-439.

Cho, H.-H., J. Choi, et al. (2002). "Combined Effect of Natural Organic Matter and Surfactants on the

Apparent Solubility of Polycyclic Aromatic Hydrocarbons." J Environ Qual 31(1): 275-280.

Cho, H. Y., H. W. Seung, et al. (2006). "Effects of Intermediate Metabolites on Phenanthrene

Biodegradation." Journal of Microbiology and Biotechnology 16(6): 969-973.

Choi, Y. J., J. Y. Park, et al. (2008). "Flow Characteristics of Microbubble Suspensions in Porous Media

as an Oxygen Carrier." CLEAN - Soil, Air, Water 36(1): 59-65.

Chowdiah, P., B. R. Misra, et al. (1998). "Foam propagation through soils for enhanced in-situ

remediation." Journal of Hazardous Materials 62(3): 265-280.

Couto, H. J. B., G. Massarani, et al. (In Press). "Remediation of sandy soils using surfactant solutions

and foams." Journal of Hazardous Materials In Press, Corrected Proof.

Dean, S. M., Y. Jin, et al. (2001). "Phenanthrene Degradation in Soils Co-Inoculated with

Phenanthrene-Degrading and Biosurfactant-Producing Bacteria." J Environ Qual 30(4): 1126-

1133.

Doyle, E., L. Muckian, et al. (2008). Chapter 2 Microbial PAH Degradation. Advances in Applied

Microbiology, Academic Press. Volume 65: 27-66.

Page 122: Avery Gottfried - ME thesis 2009

113

Eaton, R. W. and P. J. Chapman (1992). "Bacterial metabolism of naphthalene: Construction and use

of recombinant bacteria to study ring cleavage of 1,2-dihydroxynaphthalene and subsequent

reactions." Journal of Bacteriology 174(23): 7542-7554.

Edwards, D. A., Z. Adeel, et al. (1994). "Distribution of Nonionic Surfactant and Phenanthrene in a

Sediment/Aqueous System." Environmental Science & Technology 28(8): 1550-1560.

García-Falcón, M. S., B. Cancho-Grande, et al. (2004). "Stirring bar sorptive extraction in the

determination of PAHs in drinking waters." Water Research 38(7): 1679-1684.

Garcia-Junco, M., C. Gomez-Lahoz, et al. (2003). "Biosurfactant- and biodegradation-enhanced

partitioning of polycyclic aromatic hydrocarbons from nonaqueous-phase liquids."

Environmental Science and Technology 37(13): 2988-2996.

Gómez, J., M. T. Alcántara, et al. (In Press). "A two-stage process using electrokinetic remediation

and electrochemical degradation for treating benzo[a]pyrene spiked kaolin." Chemosphere

In Press, Corrected Proof.

Grimm, A. C. and C. S. Harwood (1997). "Chemotaxis of Pseudomonas spp. to the polyaromatic

hydrocarbon naphthalene." Appl. Environ. Microbiol. 63(10): 4111-4115.

Guerin, W. F. and S. A. Boyd (1995). "Maintenance and induction of naphthalene degradation

activity in Pseudomonas putida and an Alcaligenes sp. under different culture conditions."

Appl. Environ. Microbiol. 61(11): 4061-4068.

Guiyun, B., M. L. Brusseau, et al. (1998). "Influence of cation type, ionic strength, and pH on

solubilization and mobilization of residual hydrocarbon by a biosurfactant." Journal of

Contaminant Hydrology 30(3-4): 265-279.

Harvey, R. G. (1991). Polycyclic Aromatic Hydrocarbons: Chemistry and Carcinogenicity. Cambridge,

Cambridge University Press.

Hemond, H., F and E. Fechner-Levy, J (2000). Chemical Fate and Transport in the Environment.

London, Academic Press.

Hendricks, D. W. (2006). Water Treatment Unit Processes: Physical and Chemical, CRC Press.

Huang, C. P. and D. Cha (2001). Desorption of PAHs in the Presence of Surfactant. Electrochemical

Processes for in-situ Treatment of Contaminated Soils. Newark, US Department of Energy.

Hugouvieux-Cotte-Pattat, N., T. Kohler, et al. (1990). "Growth-phase-dependent expression of the

Pseudomonas putida TOL plasmid pWW0 catabolic genes." J. Bacteriol. 172(12): 6651-6660.

Institute for Ecology of Industrial Areas (1999). Bioremediation of Petroleum Hydrocarbon-

contaminated Soils. Department of Energy USA. Katowice, Poland.

Jafvert, C. T. and J. K. Heath (1991). "Sediment- and saturated-soil-associated reactions involving an

anionic surfactant (dodecylsulfate). 1. Precipitation and micelle formation." Environmental

Science & Technology 25(6): 1031-1038.

Jauregi, P. and J. Varley (1999). "Colloidal gas aphrons: potential applications in biotechnology."

17(10): 389-395.

Jin, D., X. Jiang, et al. (2007). "Effects of concentration, head group, and structure of surfactants on

the degradation of phenanthrene." Journal of Hazardous Materials 144(1-2): 215-221.

Johnsen, A. R., L. Y. Wick, et al. (2005). "Principles of microbial PAH-degradation in soil."

Environmental Pollution 133(1): 71-84.

Jonsson, S., Y. Persson, et al. (2007). "Degradation of polycyclic aromatic hydrocarbons (PAHs) in

contaminated soils by Fenton's reagent: A multivariate evaluation of the importance of soil

characteristics and PAH properties." Journal of Hazardous Materials 149(1): 86-96.

Josephy, D. and B. Mannervik (2006). Chapter18 Polycyclic Aromatic Hydrocarbon Carcinogens.

Molecular Toxicology, Oxford University Press: 589.

Jun, O. (2008). Phenanthrene Degradation. Biocatalysis/Biodegradation Database, University of

Minnestoa.

Kim, I. S., J.-S. Park, et al. (2001). "Enhanced biodegradation of polycyclic aromatic hydrocarbons

using nonionic surfactants in soil slurry." Applied Geochemistry 16(11-12): 1419-1428.

Page 123: Avery Gottfried - ME thesis 2009

114

Kiyohara, H., S. Torigoe, et al. (1994). "Cloning and characterization of a chromosomal gene cluster,

pah, that encodes the upper pathway for phenanthrene and naphthalene utilization by

Pseudomonas putida OUS82." J. Bacteriol. 176(8): 2439-2443.

Ko, S.-O., M. A. Schlautman, et al. (1998). "Partitioning of Hydrophobic Organic Compounds to

Sorbed Surfactants. 1. Experimental Studies." Environmental Science & Technology 32(18):

2769-2775.

Kraaij, R. H., S. Ciarelli, et al. (2001). "Bioavailability of lab-contaminated and native polycyclic

aromatic hydrocarbons to the amphipod Corophium volutator relates to chemical

desorption." Environmental Toxicology and Chemistry 20(8): 1716-1724.

Kwok, C. K. and K. C. Loh (2003). "Effects of Singapore soil type on bioavailability of nutrients in soil

bioremediation." Advances in Environmental Research 7: 889-900.

Labana, S., K. Manisha, et al. (2007). Diversity, Biodegradation and Bioremediation of Polycyclic

Aromatic Hydrocarbons. Environmental Bioremediation Technologies. S. N. Singh and R. D.

Tripathi. Berlin, Springer Press: 409-443.

Laha, S., B. Tansel, et al. (In Press). "Surfactant-soil interactions during surfactant-amended

remediation of contaminated soils by hydrophobic organic compounds: A review." Journal of

Environmental Management In Press, Corrected Proof.

Landrum P, Eadie B, et al. (1992). "Variation in the bioavailability of polycyclic aromatic

hydrocarbons to the amphipod Diporeia spp. with sediment aging." Environmental

Toxicology and Chemistry 11(8): 1197-1208.

Larmignat, S., D. Vanderpool, et al. (2008). "Rheology of colloidal gas aphrons (microfoams)."

Colloids and Surfaces A: Physicochemical and Engineering Aspects 322(1-3): 199-210.

Lee, D. S., M. W. Lee, et al. (2005). "Effects of salicylate and glucose on biodegradation of

phenanthrene by Burkholderia cepacia PM07." Journal of Microbiology and Biotechnology

15(4): 859-865.

Lee, K., J.-W. Park, et al. (2003). "Effect of additional carbon source on naphthalene biodegradation

by Pseudomonas putida G7." Journal of Hazardous Materials 105(1-3): 157-167.

Lee, P.-H., S. K. Ong, et al. (1999). "Extraction Method for Analysis of PAHs in Coal-Tar-Contaminated

Soils." Practice Periodical of Hazardous, Toxic, and Radioactive Waste Management 3(4):

155-162.

Leigh, D., D. Jerger, et al. (1997). Injetion of Surfactant/Ozygen Microbubbles as a Delivery System

for In Situ Bioremediation. 4th International in Situ and on-Site Bioremediation Symposium.

B. Alleman, C and A. Leeson. New Orleans, LA, Battle Press. 2: 603 - 608.

Lin, S.-C. (1996). "Biosurfactants: Recent advances." Journal of Chemical Technology & Biotechnology

66(2): 109-120.

Makkar, R. S. and K. J. Rockne (2003). "Comparison of synthetic surfactants and biosurfactants in

enhancing biodegradation of polycyclic aromatic hydrocarbons." Environmental Toxicology

and Chemistry 22(10): 2280-2292.

Marx, R. B. and M. D. Aitken (2000). "Bacterial Chemotaxis Enhances Naphthalene Degradation in a

Heterogeneous Aqueous System." Environmental Science & Technology 34(16): 3379-3383.

Mohan, S., T. Kisa, et al. (2006). "Bioremediation technologies for treatment of PAH-contaminated

soil and strategies to enhance process efficiency." Reviews in Environmental Science and

Biotechnology 5(4): 347-374.

Mulder, H., A. M. Breure, et al. (2001). "Application of a mechanistic desorption-biodegradation

model to describe the behavior of polycyclic aromatic hydrocarbons in peat soil aggregates."

Chemosphere 42(3): 285-299.

Mulligan, C. N., R. N. Yong, et al. (2001). "Surfactant-enhanced remediation of contaminated soil: a

review." Engineering Geology 60(1-4): 371-380.

Myers, D. (1988). Surfactant Science and Technology. New York, VCH Publishers Inc.

New Zealand Ministry for the Environment (2007). Environment New Zealand 2007. Wellington.

Page 124: Avery Gottfried - ME thesis 2009

115

Nielsen, D. R., M. T. Van Genuchten, et al. (1986). "Water Flow and Solute Transport Processes in the

Unsaturated Zone." Water Resources Research 22(9S): 89S - 108 S.

Noordman, W., W. Ji, et al. (1998). "Effects of Rhamnolipid Biosurfactants on Removal of

Phenanthrene from Soil " Environ. Sci. Technol. 32(12): 1806 -1812.

Noordman, W. H. (1999). Biosurfactant-enhanced soil remediation : effect of rhamnolipid on the

bioavailability and microbial uptake of hydrophobic compounds. science, University of

Groningen. Phd: 116.

Ogunseitan, O. A. and B. H. Olson (1993). "Effect of 2-hydroxybenzoate on the rate of naphthalene

mineralization in soil." Applied Microbiology and Biotechnology 38(6): 799-807.

Oliveira, R. C. G., J. F. Oliveira, et al. (2004). The effect of hydrophobic fine particles on the foam

flushing remediation process. Surface and Colloid Science: 293-297.

Oostrom, M., J. H. Dane, et al. (2006). "A Review of Multidimensional, Multifluid Intermediate-Scale

Experiments: Nonaqueous Phase Liquid Dissolution and Enhanced Remediation." Vadose

Zone J 5(2): 570-598.

Ortega-Calvo, J. J., A. I. Marchenko, et al. (2003). "Chemotaxis in polycyclic aromatic hydrocarbon-

degrading bacteria isolated from coal-tar- and oil-polluted rhizospheres." FEMS Microbiology

Ecology 44(3): 373-381.

Park, J. H., X. Zhao, et al. (2001). "Comparison of biodegradation kinetic parameters for naphthalene

in batch and sand column systems by Pseudomonas putida." Environmental Progress 20(2):

93-102.

Park, J. Y., Y. J. Choi, et al. (In Press). "Microbubble suspension as a carrier of oxygen and acclimated

bacteria for phenanthrene biodegradation." Journal of Hazardous Materials In Press,

Corrected Proof.

Pastewski, S., E. Hallmann, et al. (2006). "Physicochemical aspects of the application of surfactants

and biosurfactants in soil remediation." Environmental Engineering Science 23(4): 579-588.

Perry, J. J. (1979). "Microbial cooxidations involving hydrocarbons." Microbiol. Mol. Biol. Rev. 43(1):

59-72.

Pieper, D. H. and W. Reineke (2000). "Engineering bacteria for bioremediation." Current Opinion in

Biotechnology 11(3): 262-270.

Pignatello, J. J. and L. S. Donald (1999). The Measurement and Interpretation of Sorption and

Desorption Rates for Organic Compounds in Soil Media. Advances in Agronomy, Academic

Press. Volume 69: 1-73.

Platt, A., V. Shingler, et al. (1995). "The 4-hydroxy-2-oxovalerate aldolase and acetaldehyde

dehydrogenase (acylating) encoded by the nahM and nahO genes of the naphthalene

catabolic plasmid pWW60-22 provide further evidence of conservation of meta-cleavage

pathway gene sequences." Microbiology 141(9): 2223-2233.

Powell, S. N., D. R. Singleton, et al. (2008). "Effects of Enrichment with Salicylate on Bacterial

Selection and PAH Mineralization in a Microbial Community from a Bioreactor Treating

Contaminated Soil." Environ. Sci. Technol. 42(11): 4099-4105.

Prabhu, Y. and P. S. Phale (2003). "Biodegradation of phenanthrene by Pseudomonas sp. strain PP2:

novel metabolic pathway, role of biosurfactant and cell surface hydrophobicity in

hydrocarbon assimilation." Applied Microbiology and Biotechnology 61(4): 342-351.

Riser-Roberts, E. (1998). Remediation of Petroleum Contaminated Soils: Biological, Physical, and

Chemical Processes. Boca Raton, Florida, Lewis Publishers.

Rogers, S. W., S. K. Ong, et al. (2002). "Natural Attenuation of Polycyclic Aromatic Hydrocarbon-

Contaminated Sites: Review." Practice Periodical of Hazardous, Toxic, and Radioactive Waste

Management 6(3): 141-155.

Rosenberg, E. and E. Z. Ron (1999). "High- and low-molecular-mass microbial surfactants." Applied

Microbiology and Biotechnology 52(2): 154-162.

Page 125: Avery Gottfried - ME thesis 2009

116

Roskam, G. D. and R. N. J. Comans (In Press). "Availability and leaching of polycyclic aromatic

hydrocarbons: Controlling processes and comparison of testing methods." Waste

Management In Press, Corrected Proof.

Rothmel, R. K., R. W. Peters, et al. (1998). "Surfactant foam/bioaugmentation technology for in situ

treatment of TCE-DNAPLs." Environmental Science and Technology 32(11): 1667-1675.

Roy, D., R. R. Kommalapati, et al. (1995). "Soil flushing of residual transmission fluid: Application of

Colloidal Gas Aphron suspensions and conventional surfactant solutions." Water Research

29(2): 589-595.

Ruberto, L. A. M., S. C. Vazquez, et al. (2006). "Phenanthrene Biodegradation in Soils Using an

Antarctic Bacterial Consortium." Bioremediation Journal 10(4): 191 - 201.

Samanta, S. K., A. K. Chakraborti, et al. (1999). "Degradation of phenanthrene by different bacteria:

evidence for novel transformation sequences involving the formation of 1-naphthol."

Applied Microbiology and Biotechnology 53(1): 98-107.

Samanta, S. K., O. V. Singh, et al. (2002). "Polycyclic aromatic hydrocarbons: environmental pollution

and bioremediation." Trends in Biotechnology 20(6): 243-248.

Schell, M. A. and P. E. Wender (1986). "Identification of the nahR gene product and nucleotide

sequences required for its activation of the sal operon." Journal of Bacteriology 166(1): 9-14.

Schlebaum, W., G. Schraa, et al. (1999). "Influence of Nonlinear Sorption Kinetics on the Slow-

Desorbing Organic Contaminant Fraction in Soil." Environ. Sci. Technol. 33(9): 1413-1417.

Sebba, F. (1971). "Microfoams--an unexploited colloid system." Journal of Colloid and Interface

Science 35(4): 643-646.

Sebba, F. (1985). "Separations Using Aphrons." Separation & Purification Reviews 14(1): 127-148.

Shin, K.-H., K.-W. Kim, et al. (2008). "Rhamnolipid Morphology and Phenanthrene Solubility at

Different pH Values." J Environ Qual 37(2): 509-514.

Shin, K. H., K. W. Kim, et al. (2004). "Combined effects of pH and biosurfactant addition on

solubilization and biodegradation of phenanthrene." Applied Microbiology and

Biotechnology 65(3): 336-343.

Simunek, J., M. Th. van Genuchten, et al. (2003). Studio of Analytical Models for Solving the

Convection-Dispersion Equation (STANMOD) I. G. W. M. Center. Golden, Colorado School of

Mines: CXTFIT 2; CHIAN; CFITM; 3DADE; N3DADE; SCREEN.

Singh, A. and O. P. Ward, Eds. (2004). Biodegradation and Bioremediation Soil Biology. Berlin

Heidelberg, Springer-Verlag.

Soberón-Chávez, G., F. Lépine, et al. (2005). "Production of rhamnolipids by Pseudomonas

aeruginosa." Applied Microbiology and Biotechnology 68(6): 718-725.

Son, A.-J., J.-U. Lee, et al. (2003 ). "Chemical and Ecotoxicity Assessment of PAH-Contaminated Soils

Remediated by Enhanced Soil Flushing." Environmental Engineering Science 20(3): 197-206.

Song, Y. F., X. Jing, et al. (2002). "Comparative study of extraction methods for the determination of

PAHs from contaminated soils and sediments." Chemosphere 48(9): 993-1001.

Stellner, K. L. and J. F. Scamehorn (1989). "Hardness tolerance of anionic surfactant solutions. 1.

Anionic surfactant with added monovalent electrolyte." Langmuir 5(1): 70-77.

Suthersan, S. S. and C. F. Payne (2005). In Situ Remediation Engineering. London, CRC Press.

The Interstate Technology & Regulatory Council (2005). Overview of In Situ Bioremediation of

Chlorinated Ethene DNAPL Source Zones. E. R. I. o. t. S. (ERIS). Washington DC, Interstate

Technology & Regulatory Council, Bioremediation of Dense Nonaqueous Phase Liquids (Bio

DNAPL) Team.

Toride, N., F. J. Leij, et al. (1993). "A Comprehensive Set of Analytical Solutions for Nonequilibrium

Solute Transport With First-Order Decay and Zero-Order Production." Water Resources

Research 27(7): 2167–2182.

Toride, N., F. J. Leij, et al. (1995). CXTFIT Model: Estimates parameters in several equilibrium and

non-equilibrium transport models from laboratory or field tracer experiments USDA

Agricultural Research Service

Page 126: Avery Gottfried - ME thesis 2009

117

Torrens, J. L., D. C. Herman, et al. (1998). "Biosurfactant (Rhamnolipid) Sorption and the Impact on

Rhamnolipid-Facilitated Removal of Cadmium from Various Soils under Saturated Flow

Conditions." Environ. Sci. Technol. 32(6): 776-781.

U.S Environmental Protection Agency. (1996). "Test methods for evaluating solid wastes

physical/chemical methods: Method 3550B Ultrasonic Extraction." Retrieved July 9, 2008,

2008, from http://www.epa.gov/sw-846/pdfs/3550b.pdf.

U.S Environmental Protection Agency. (2007). "Test methods for evaluating solid wastes

physical/chemical methods: Method 3550C Ultrasonic Extraction." Retrieved July 9, 2008,

2008, from http://www.epa.gov/SW-846/pdfs/3550c.pdf.

U.S Environmental Protection Agency. (2008, October 18, 2008). "Toxic and Priority Pollutants: Clean

Water Act Analytical Test Methods." Retrieved December 21, 2008, from

http://www.epa.gov/waterscience/methods/pollutants-background.htm.

Verschueren, K. (1983). Handbook of environmental data on organic chemicals. New York, Van

Nostrand Reinhold.

Volkering, F., A. M. Breure, et al. (1993). "Effect of micro-organisms on the bioavailability and

biodegradation of crystalline naphthalene." Applied Microbiology and Biotechnology 40(4):

535-540.

Volkering, F., A. M. Breure, et al. (1992). "Microbial degradation of polycyclic aromatic

hydrocarbons: effect of substrate availability on bacterial growth kinetics." Applied

Microbiology and Biotechnology 36(4): 548-552.

Wan, J., S. Veerapaneni, et al. (2001). "Generation of stable microbubbles and their transport

through porous media." Water Resources Research 37(5).

Wang, P. and A. A. Keller (2008). "Partitioning of hydrophobic organic compounds within soil-water-

surfactant systems." Water Research 42(8-9): 2093-2101.

Wang, Q., X. Fang, et al. (2007). "Engineering bacteria for production of rhamnolipid as an agent for

enhanced oil recovery." Biotechnology and Bioengineering 98(4): 842-853.

Wang, S. and C. N. Mulligan (2004). "An evaluation of surfactant foam technology in remediation of

contaminated soil." Chemosphere 57(9): 1079-1089.

Warith, M., L. Fernandes, et al. (1999). "Design of in-situ microbial filter for the remediation of

naphthalene." Waste Management 19(1): 9-25.

Wick, L. Y., T. Colangelo, et al. (2001). "Kinetics of Mass Transfer-Limited Bacterial Growth on Solid

PAHs." Environmental Science & Technology 35(2): 354-361.

Woo, S., C. Jeon, et al. (2004). "phenanthrene biodegradation in soil slurry systems: Influence of

salicylate and triton X-100." Korean Journal of Chemical Engineering 21(2): 412-418.

Woo, S. H., M. W. Lee, et al. (2004). "Biodegradation of phenanthrene in soil-slurry systems with

different mass transfer regimes and soil contents." Journal of Biotechnology 110(3): 235-

250.

Woo, S. H., J. M. Park, et al. (2001). "Evaluation of the interaction between biodegradation and

sorption of phenanthrene in soil-slurry systems." Biotechnology and Bioengineering 73(1):

12-24.

World Health Organization (1998) "Environmental Criteria 202. Selected Non-Heterocyclic Polycyclic

Aromatic

Hydrocarbons."

Xiang, L., M. Xian-min, et al. (2006). "Experiments and modelling of phenanthrene biodegradation in

the aqueous phas by a mixed culture." Journal of Environmental Sciences 18(1): 147-153.

Xu, Q., M. Nakajima, et al. (2008). "A comparative study of microbubble generation by mechanical

agitation and sonication." Innovative Food Science & Emerging Technologies 9(4): 489-494.

Yan, L. (1985). "Study of the carcinogenic mechanism for polycyclic aromatic hydrocarbons--

extended bay region theory and its quantitative model." Carcinogenesis 6(1): 1-6.

Yen, K. M. and C. M. Serdar (1988). "Genetics of naphthalene catabolism in pseudomonads." Critical

reviews in microbiology 15(3): 247-268.

Page 127: Avery Gottfried - ME thesis 2009

118

Yolcubal, I., S. A. Pierce, et al. (2002). "Biodegradation during Contaminant Transport in Porous

Media: V. The Influence of Growth and Cell Elution on Microbial Distribution." J Environ Qual

31(6): 1824-1830.

Yu, H., L. Zhu, et al. (2007). "Enhanced desorption and biodegradation of phenanthrene in soil-water

systems with the presence of anionic-nonionic mixed surfactants." Journal of Hazardous

Materials 142(1-2): 354-361.

Zhou, W. and L. Zhu (2005). "Distribution of polycyclic aromatic hydrocarbons in soil-water system

containing a nonionic surfactant." Chemosphere 60(9): 1237-1245.

Zhou, W. and L. Zhu (2007). "Efficiency of surfactant-enhanced desorption for contaminated soils

depending on the component characteristics of soil-surfactant-PAHs system." Environmental

Pollution 147(1): 66-73.

Zhu, L., B. Chen, et al. (2003). "Interactions of Organic Contaminants with Mineral-Adsorbed

Surfactants." Environmental Science & Technology 37(17): 4001-4006.

Zhu, L. and W. Zhou (2008). "Partitioning of polycyclic aromatic hydrocarbons to solid-sorbed

nonionic surfactants." Environmental Pollution 152(1): 130-137.

Zitha, P., Q. Nguyen, et al. (2006). "Coupling of Foam Drainage and Viscous Fingering in Porous

Media Revealed by X-ray Computed Tomography." Transport in Porous Media 64(3): 301-

313.