You are on page 1of 189

Biodegradation of Polycyclic Aromatic Hydrocarbons

by Soil Bacteria

By

Fazal ur Rehman

Department of Microbiology
Quaid-i-Azam University
Islamabad
2011

Biodegradation of Polycyclic Aromatic Hydrocarbons


by Soil Bacteria

A thesis submitted in partial fulfillment of the requirements


for the degree of
DOCTOR OF PHILOSOPHY
IN
MICROBIOLOGY

By

Fazal ur Rehman
Department of Microbiology
Quaid-i-Azam University
Islamabad
2011

DECLARATION
The material contained in this thesis is my original work and I have not presented any
part of this thesis/work elsewhere for any other degree.

Fazal ur Rehman

DEDICATED
TO
TO MY LOVING FAMILY MEMBERS

Certificate
The Department of Microbiology, Quaid-i-Azam University, Islamabad,
Pakistan accepts this thesis by Mr. Fazal ur Rehman in its present form as
satisfying the thesis requirements for the Ph.D Degree in Microbiology.

Internal Examiner: ____________________


Dr. Safia Ahmed

External Examiner: _____________________


Dr. Allah Nawaz

External Examiner: _____________________


Dr. Ghazala Kaukab

Chairman:

_____________________
Dr. Safia Ahmed

Dated: May 27, 2011

Table of Contents

Contents

Page

List of Abbreviations

List of Tables

iii

List of Figures

iv

Acknowledgements

ix

Abstract

Chapter 1: Introduction

Chapter 2: Review of Literature

13

Chapter 3: Materials and Methods

32

3.1 CHEMICALS

32

3.2 SELECTION OF PAH TRANSFORMING BACTERIA

32

3.2.1 Culture condition and media

32

3.2.2 Screening of PAHs degrading bacteria

33

3.2.3 Identification of selected bacterial isolates

33

3.2.3.1 Morphological and biochemical characterization

33

3.2.3.2 Molecular characterization based on 16S rRNA

34

gene
3.3 BIODEGRADATION STUDIES OF PAHs
3.3.1 Biotransformation of anthracene and naphthalene using B.

34
34

cereus KWS2 and P. stutzeri 10-1


3.3.2 Resting cell biotransformation of PAHs by Bacillus cereus

35

KWS2 and B. cereus KWS4


3.3.3 Biotransformation of PAHs by Mycobacterium PY146

35

3.3.4 Biodegradation of PAHs in Soil

36

3.3.5 Analytical methods

36

3.3.5.1 Viable Cell Counts

36

3.3.5.2 HPLC Analysis

37

3.3.5.3 Extraction and Analysis of PAHs and their

37

products
3.3.5.4 GC/MS Analysis

37

3.4 MINERALIZATION STUDIES OF PAHs

38

3.4.1 14C-PAH mineralization assay

38

3.4.2 14C-Pyrene and phenanthrene mineralization using

39

different cells densities


3.5 EXPRESSION OF PYRENE DIOXYGENASE GENE (nidAB)

40

FROM MYCOBACTERIUM PY146 into E. coli


3.5.1 Isolation of genomic DNA

40

3.5.2 Amplification of nidAB

40

3.5.3 Cloning of nidAB gene in E. coli

41

3.5.3.1 Cloning of nidAB into pCR 2.1 TOPO vector

41

3.5.3.2 Transformation of E. coli with recombinant pCR

41

2.1 TOPO vector


3.5.3.3 Isolation of recombinant pCR 2.1 TOPO vector

41

from transformed E. coli cells


3.5.3.4 Restriction enzyme analysis

41

3.5.3.5 Addition of restriction sites on the ends of nidAB

42

gene
3.5.3.6 Cloning of nidAB gene in Expression Vector

42

pK223-3
3.5.3.6.1 Dephosphorylation Reaction

43

3.5.3.6.2 Gel Purification of fragments

43

3.5.3.6.3 Ligation of EcoRInidAB HindIII and

43

pkk223-3 fragment
3.5.3.6.4 Transformation of E. coli with recombinant

44

pkk223-3
3.5.3.6.5 Isolation of a recombinant pkk223-3
plasmid containing EcoRInidAB HindIII

44

3.5.3.6.6 Restriction enzyme analysis of

44

recombinant pkk223-3 vector


3.5.3.6.7 Sequencing of the recombinant pk223-3

45

plasmid contained EcoRInidAB HindIII


3.5.3.6.8 Direct Colony PCR to verify fragment

45

insert size
3.5.4 Biotransformation of PAHs by E. coli with cloned nidAB

45

into pkk223-3
3.5.5 Invitro RNA synthesis of nidAB transcripts

46

3.5.5.1 Linearizing the nidAB DNA template

46

3.5.5.2 Synthesis of large quantities of RNA

46

3.5.5.3 Isolation of RNA transcript of nidAB

47

3.5.6 Quantification of DNA and RNA copy number of nidAB


3.5.6.1 Isolation of DNA and RNA from Mycobacterium

47
47

PY146
51

Chapter 4: Results
4.1 SELECTION OF PAH TRANSFORMING BACTERIA

51

4.1.1 Identification of selected bacterial isolates

51

4.1.1.1 Morphological and biochemical characterization

51

4.1.1.2 Molecular characterization based on 16S rRNA gene

52

4.2 BIODEGRADATION STUDIES OF PAHS


4.2.1 Biotransformation of anthracene and naphthalene using B.

59
59

cereus KWS2 and P. stutzeri 10-1


4.2.2 Resting cell biotransformation of PAHs by Bacillus cereus

60

KWS2 and KWS4


4.2.3 Biotransformation of PAHs by Mycobacterium PY146

67

4.2.4 Biodegradation of PAHs in soil

73

4.3 MINERALIZATION STUDIES OF PAHs

75

4.3.1

14

C-PAH mineralization by Bacillus cereus KWS2 and

75

Mycobacterium PY146
14

4.3.2

C-Pyrene and phenanthrene mineralization using

75

different cells densities


4.4 EXPRESSION OF PYRENE DIOXYGENASE GENE (nidAB)

80

FROM MYCOBACTERIUM PY146


4.4.1 Cloning of nidAB gene

80

4.4.2 Biotransformation of PAHs by E. coli with cloned nidAB

81

into pkk223-3
4.4.3 Invitro RNA synthesis of nidAB transcripts
4.4.3.1 Quantification of RNA transcripts of nidAB from E.

81
81

coli
4.4.4 Comparison of nidAB DNA and RNA copy number using

81

real-time PCR with PAH mineralization in Mycobacterium


PY146

Chapter 5: Discussion

88

Conclusions

101

Future Prospects

102

References

103

Appendices I-V

132

List of Abbreviations
BaA

Benzo(a)anthracene

BaP

Benzo(a)pyrene

BSM

Basal salt medium

BSTFA

N,O-Bis(trimethylsilyl)triflouroacetamide

DGGE

Denaturing gradient gel electrophoresis

DNAPLs

Dense nonaqueous phase liquids

ER

Estrogen receptors

GC/MS

Gas chromatography/mass spectrometry

GEM

Genetically engineered microorganisms

HMW

High molecular weight

HPLC

High performance liquid chromatography

IPTG

Isopropyl--D-thiogalactopyranoside

LB

Luria-Bertani

LNAPLs

Light nonaqueous phase liquids

LP

Lignin peroxidase

MnP

Manganese peroxidase

MR-VP

Methyl red-Voges Proskauer

NADPH

Nicotinamide adenine dinucleotide phosphate

NDO

Naphthalene 1,2-dioxygenase

PCBs

Polychlorinated biphenyls

PCR

Polymerase chain reaction

PAHs

Polycyclic aromatic hydrocarbons

PHE

Phenanthrene

pht

Phthalate degrading operon

PNR

Mineral salt medium

PNRG

PNR with 5 mM Glucose

POPs

Persistent organic pollutants

ppm

Parts per million

QC-PCR

Quantitative competitive polymerase chain reaction

rpm

Rounds per minute

RT-QPCR

Real time quantitative polymerase chain reaction


i

SPM

Suspended particulate matter

TBE buffer

Tris/Borate/EDTA buffer

TCA

Tricarboxylic acid

TPH

Total petroleum hydrocarbons

TSAP

Thermosensitive alkaline phosphatase

TSP

Total suspended particulates

US EPA

United States Environmental Protection Agency

WHO

World Health Organization

ii

List of Tables
Table No.
1.1

4.1
4.2
4.3
4.4
4.5
4.6
4.7

4.8

4.9

4.10

Caption
Structures, Formulas, MW and solubilities of some
priority polycyclic aromatic hydrocarbons identified
by the US EPA
Growth of bacterial isolates on PNR agar plates
containing anthracene
Growth of bacterial isolates on PNR agar plates
containing Pyrene
Morphological characteristics of bacterial isolates
Biochemical characterization of selected bacterial
isolates
Sugar fermentation tests of the bacterial isolate
Identification of the bacterial isolates on the basis of
16S rRNA
GC mass spectral data of metabolites from
naphthalene,
phenanthrene
and
anthracene
degradation by B. cereus KWS2
GC mass spectral data of metabolites from
naphthalene and phenanthrene degradation by B.
cereus KWS4
GC mass spectral data of metabolites from
phenanthrene, anthracene and pyrene degradation by
resting cells of Mycobacterium PY146
GC mass spectral data of metabolites from
phenanthrene, anthracene and pyrene degradation by
growing cells of Mycobacterium PY146

Page No.
3

53
54
55
56
57
58
65

66

71

72

iii

List of Figures
Figure No.
1.1

1.2
3.1
3.2
4.1

4.2

4.3

4.4

4.5
4.6
4.7
4.8
4.9
4.10

4.11

4.12
4.13
4.14
4.15

Caption
Mechanisms of action of aromatic dioxygenases. A)
Aromatic ring hydroxylating dioxygenase; B) extadiol
ring cleavage dioxygenase; and C) intradiol ring
cleavage dioxygenase.
Proposed phenanthrene degradation pathway in M.
vanbaalenii PYR-1
Map of plasmid vector pCR 2.1-TOPO used for
nidAB cloning
Map of vector pKK223-3 used for nidAB expression
Percentage reduction of anthracene concentration (10
ppm) in PNRG medium at different pH values by
Bacillus cereus KWS2 in shake flask experiment.
Growth (CFU/mL) of Bacillus cereus KWS2 in
PNRG medium with 10 ppm anthracene at different
pH values.
Percentage reduction of anthracene concentration (10
ppm) in PNRG medium at different temperature by
Bacillus cereus KWS2 in shake flask experiment.
Growth (CFU/mL) of Bacillus cereus KWS2 in
PNRG medium with 10 ppm anthracene at different
temperature.
Percentage reduction of naphthalene concentration by
Pseudomonas stutzeri 10-1
Percentage reduction of anthracene concentration by
Pseudomonas stutzeri 10-1.
Growth of different Mycobacterium strains on 10 ppm
pyrene in BSM medium.
Effect of different temperature on the growth of
Mycobacterium PY146
Effect of different pH on the growth of
Mycobacterium PY146
Biodegradation of PAHs (phenanthrene, anthracene
and pyrene) in soil by Mycobacterium PY146 without
addition of yeast extract.
Biodegradation of PAHs (phenanthrene, anthracene
and pyrene) in soil with addition of yeast extract by
Mycobacterium PY146.
14
C-phenanthrene mineralization by Mycobacterium
PY146
14
C-pyrene mineralization by Mycobacterium PY146
14
C-benzo(a)pyrene mineralization by Mycobacterium
PY146.
Mineralization of 14C-pyrene by cells of
Mycobacterium PY146 along different incubation
time

Page No.
10

11
49
50
62

62

63

63

64
64
69
70
71
74

74

77
77
78
78

iv

4.16

Mineralization of pyrene with different cell densities


79
by Mycobacterium PY146
4.17
Mineralization of 14C-phenanthrene with different cell
79
densities by Mycobacterium PY146.
4.18
Ethidium bromide stained agarose gel showing the
83
PCR product of nidAB gene from Mycobacterium
PY146. Lane L: 1KB DNA ladder; Lanes 1: nidAB.
The arrow indicates the location of the nidAB
product.
4.19
Ethidium bromide stained 1% agarose gel showing the
83
location of the plasmid (top arrow) and nidAB gene
(bottom arrow) after digestion with EcoR1. Lane L:
1KB DNA ladder; Lanes 1, 2 and 5: have both the
plasmid and nidAB.
4.20
Ethidium bromide stained 1 % agarose gel showing
84
the plasmid and nidAB after digestion with EcoR1 and
HindIII. Lane L: 1KB DNA ladder; Lanes 1 and 2
have only plasmid; Lanes 3-8: have both the plasmid
and nidAB.
84
4.21
Ethidium bromide stained 1 % agarose gel showing
the presence of cloned nidAB gene in transformed E.
coli. Lane L: 1KB DNA ladder; Lanes 1 to 6: all have
the cloned nidAB.
85
4.22
Copies of nidAB RNA transcripts in E. coli clones
grown in LB containing ampicillin with and without
IPTG.
4.23
Standard curves of DNA and RNA/ml of bacteria
86
from different cell densities from 0.1 to 0.5 O.D.
4.24
RNA to DNA ratio at different cell densities from 0.1
86
to 0.5 O.D
4.25
Comparison of mineralization of phenanthrene from
87
different cell densities and the nidAB RNA/DNA ratio
in Mycobacterium PYR146.
4.26
Effect of added un-labled pyrene on the mineralization
87
14
of C- labled pyrene at different cell densities of
Mycobacterium PYR146 and the nidAB RNA/DNA
ratio.
5.1
Proposed pathway for the degradation of naphthalene
99
and phenanthrene by B. cereus. Compounds in
brackets are from other reported studies.
5.2
Metabolites detected from the biotransformation of
100
phenanthrene,
anthracene
and
pyrene
by
Mycobacterium PY146
Figures in Appendix
Ia
Phylogenetic tree of bacterial isolate
132
Ib
Percent identity of bacterial isolates
133
Ic
nidAB gene sequence cloned from Mycobacterium
134-135
PY146 and containing the EcoRI and HindIII
restriction enzyme adaptor sequences

II a

II b

II c

II d

II e

II f.

II g

II h

II i

II j

II k

III a

Mass spectra obtained by GC/MS of 2-Naphthalnol


produced by Bacillus cereus KWS2 grown with
naphthalene (A) and from NIST mass spectral
library (B).
Mass spectra obtained by GC/MS of benzeneacetic
acid produced by Bacillus cereus KWS2 grown
with naphthalene (A) and from NIST mass spectral
library (B).
Mass
spectra
obtained
by
GC/MS
of
Benzenecarboxylic acid produced by Bacillus
cereus KWS2 grown with naphthalene (A) and
from NIST mass spectral library (B).
Mass spectra obtained by GC/MS of Bezaldehyde
produced by Bacillus cereus KWS2 grown with
naphthalene (A) and from NIST mass spectral
library (B).
Mass spectra obtained by GC/MS of 9phenanthrenol produced by Bacillus cereus KWS2
grown in phenanthrene (A) and from NIST mass
spectral library (B).
Mass spectra obtained by GC/MS of 9, 10-dihydro9, 10-dihydroxyphenanthrene produced by Bacillus
cereus KWS2 grown in phenanthrene (A) and from
NIST mass spectral library (B).
Mass spectra obtained by GC/MS of benzeneacetic
acid produced by Bacillus cereus KWS2 grown in
phenanthrene (A) and from NIST mass spectral
library (B).
Mass spectra obtained by GC/MS of 4-hydroxy
benzeneacetic acid produced by Bacillus cereus
KWS2 grown in phenanthrene (A) and from NIST
mass spectral library (B).
Mass
spectra
obtained
by
GC/MS
of
benzenecarboxylic acid produced by Bacillus
cereus KWS2 grown in anthracene (A) and from
NIST mass spectral library (B).
Mass spectra obtained by GC/MS of benzeneacetic
acid produced by Bacillus cereus KWS2 grown in
anthracene (A) and from NIST mass spectral library
(B).
Mass spectra obtained by GC/MS of 4-hydroxy
benzeneacetic acid produced by Bacillus cereus
KWS2 grown in anthracene (A) and from NIST
mass spectral library (B).
Mass spectra obtained by GC/MS of 2-naphthalenol
produced by Bacillus cereus KWS4 grown with
naphthalene (A) and from NIST mass spectral
library (B).

136

137

138

139

140

141

142

143

144

145

146

147

vi

III b

III c

III d

III f

III g

III h

III i

IV a

IV b

IV c

IV d

IV e

Mass spectra obtained by GC/MS of benzeneacetic


acid produced by Bacillus cereus KWS4 grown
with naphthalene (A) and from NIST mass spectral
library (B).
Mass spectra obtained by GC/MS of Bezaldehyde
produced by Bacillus cereus KWS4 grown with
naphthalene (A) and from NIST mass spectral
library (B).
Mass spectra obtained by GC/MS of 9phenanthrenol produced by Bacillus cereus KWS4
grown in phenanthrene (A) and from NIST mass
spectral library (B).
Mass spectra obtained by GC/MS of 9,10-dihyro9,10-dihydroxyphenanthrene produced by Bacillus
cereus KWS4 grown in phenanthrene (A) and from
NIST mass spectral library (B).
Mass spectra obtained by GC/MS of benzeneacetic
acid produced by Bacillus cereus KWS4 grown in
phenanthrene (A) and from NIST mass spectral
library (B).
Mass spectra obtained by GC/MS of 3-hydroxy
benzeneacetic acid produced by Bacillus cereus
KWS4 grown in phenanthrene (A) and from NIST
mass spectral library (B).
Mass spectra obtained by GC/MS of 2-hydroxy
benzoic acid produced by Bacillus cereus KWS4
grown in phenanthrene (A) and from NIST mass
spectral library (B).
Mass spectra obtained by GC/MS of 1, 2benzenecarboxylic acid produced by resting cells of
Mycobacterium PY146 with phenanthrene (A) and
from NIST mass spectral library (B).
Mass spectra obtained by GC/MS of 1,2benzenecarboxylic acid produced by resting cells of
Mycobacterium PY146 with anthracene (A) and
from NIST mass spectral library (B).
Mass spectra obtained by GC/MS of 1,2benzenecarboxylic acid produced by resting cells of
Mycobacterium PY146 with anthracene (A) and
from NIST mass spectral library (B).
Mass spectra obtained by GC/MS of 1,2benzenecarboxylic acid produced by resting cells of
Mycobacterium PY146 with pyrene (A) and from
NIST mass spectral library (B).
Mass spectra obtained by GC/MS of 1hydroxypyrene produced by resting cells of
Mycobacterium PY146 with pyrene (A) and from
NIST mass spectral library (B).

148

149

150

151

152

153

154

155

156

157

158

159

vii

Va

Vb

Vc

Vd

Ve

Vf

Vg

Vh

Vi

Mass spectra obtained by GC/MS of 1,2benzenedicarboxylic


acid
produced
by
Mycobacterium PY146 grown in phenanthrene (A)
and from NIST mass spectral library (B).
Mass spectra obtained by GC/MS of benzeneacetic
acid produced by Mycobacterium PY146 grown in
phenanthrene (A) and from NIST mass spectral
library (B).
Mass spectra obtained by GC/MS of 9,10anthracenedione produced by Mycobacterium
PY146 grown in anthracene (A) and from NIST
mass spectral library (B).
Mass
spectra
obtained
by
GC/MS
of
benzenecarboxylic
acid
produced
by
Mycobacterium PY146 grown in anthracene (A)
and from NIST mass spectral library (B).
Mass spectra obtained by GC/MS of benzeneacetic
acid produced by Mycobacterium PY146 grown in
anthracene (A) and from NIST mass spectral library
(B).
Mass spectra obtained by GC/MS of 1,2benzenedicarboxylic
acid
produced
by
Mycobacterium PY146 grown in anthracene (A)
and from NIST mass spectral library (B).
Mass
spectra
obtained
by
GC/MS
of
benzenecarboxylic
acid
produced
by
Mycobacterium PY146 grown in pyrene (A) and
from NIST mass spectral library (B).
Mass spectra obtained by GC/MS of benzeneacetic
acid produced by Mycobacterium PY146 grown in
pyrene (A) and from NIST mass spectral library
(B).
Mass spectra obtained by GC/MS of 1,2benzenedicarboxylic
acid
produced
by
Mycobacterium PY146 grown in pyrene (A) and
from NIST mass spectral library (B).

160

161

162

163

164

165

166

167

168

viii

Acknowledgments
I am grateful to Almighty Allah, the most merciful and beneficent; His blessings enabled me to achieve my
goals. Tremulous veneration is for His Holy Prophet Hazrat Muhammad (PBUHS) who is forever torch of
guidance and knowledge for humanity.
I have great reverence & admiration for my research supervisor Dr. Safia Ahmed. I am highly indebted for
her patience, moral support, sincere criticism and valuable guidance throughout the study.
I am also grateful to Prof. Dr. Abdul Hameed Dean Faculty of Biological Sciences, Quaid-i-Azam
University Islamabad, for providing research facilities at the Department.
Sincere thanks to Prof. Dr. Gary S. Sayler, Director Center for Environmental Biotechnology, University of
Tennessee, USA for providing me an opportunity to work under his kind supervision and valuable guidance.
Thanks to all the CEB fellows for their support and help specially Dr. Fu, Dr. Alice, Dr. Jennifer, Jim and
Jack .
Special thanks are to Dr. Fariha Hasan, Dr. Aamer Ali Shah and Dr. Naeem Ali for suggestions & guidance
that helped me towards the completion of this work.
I highly appreciated the efforts, guidance, and help of Dr. M. Ayub, Dr. Noor Wazir, Masroor Hussain,
Dr.Khalid Mahmood, Khalid Mahsood, Dr. Tatheer, Murad and Zia Ullah for their cooperation and
guidance.
A great depth of loving thanks to my lab fellows Lubna Tahir, Sadia, Naima, Nida, Munzer and Ali for
their help, care and enjoyable company throughout my study, which made me relax during the whole period. I
sincerely acknowledge the friendly attitude & assistance of technical staff of the lab and the office staff of
the department throughout my research work.
I also want to acknowledge Higher Education Commission, Pakistan who provided me the financial support
through indigenous scholarship for my studies and an opportunity for six months training abroad under
IRSIP program.
A non-payable debt to my family members, their wish motivated me in striving for higher education; they
prayed for me, shared the burden and made sure that I sailed through smoothly. Last but not the least; my
endless thanks are to my parents whose motivation is sprinkling added in my way.

Fazal ur Rehman
ix

Abstract
Biodegradation of organic pollutants such as polycyclic aromatic hydrocarbons using
soil microorganisms where they use these compounds as carbon and energy source, is
the safe, cheap and environment friendly way.
In the present study bacterial samples were isolated from crude oil contaminated soil.
A total of 52 isolates were initially screened, fifteen of these isolates were checked for
growth on PNR medium containing up to 1200 and 800 ppm, anthracene and pyrene,
respectively. Five bacterial isolates were selected on the basis of their best on PAHs
growth and identified as Bacillus cereus KWS2, Bacillus cereus KWS4, Bacillus
licheniformis DW3, Bacillus licheniformis Sol-10, and Pseudomonas stutzeri 10-1.
Biodegradation studies of anthracene were carried out by Bacillus cereus KWS2 at
different pH, temperature and incubation time. The optimum pH and temperature
were 7 and 30C respectively, for degradation and growth where 45 % of reduction in
anthracene concentration was observed after seven days of incubation. Growth was
maximum after 2 days of incubation (12 x 107 CFU/mL). Pseudomonas stutzeri 10-1
was also checked for anthracene and naphthalene degradation capability. 33.81 %
reduction in naphthalene concentration was observed after one week, while 33.69 and
45.35 % reduction after two and three weeks time, respectively. In case of anthracene,
87.37 % reduction was observed after first week of incubation whereas 92.98 and
95.56 % was observed after two and three weeks, respectively.
Resting cell biotransformation studies of naphthalene, phenanthrene and anthracene
were carried out by Bacillus cereus KWS2 and KWS4, analyzed by GC/MS. From
naphthalene biotransformation four metabolites, 2-naphthol, benzeneacetic acid,
benzoic acid and benzaldehyde were detected and identified. From phenanthrene
biotransformation,

9-phenanthrenol,

9,10-Dihydro,9,10-dihydroxyphenanthrene,

bezeneacetic acid, 4-hydroxy-benzeneacetic acid and 2-hydroxy benzoic acid were


detected and identified. Benzenecarboxylic acid, benzeneacetic acid and 4-hydroxybenzeneacetic acid were detected and identified from anthracene biotransformation by
B. cereus KWS2. The detection of both mono and dihydroxylated compounds
x

suggested that B. cereus strains KWS2 & KWS4 harbor both mono and dioxygenase
genes.
Biotransformation potential of a selected bacterial isolate, Mycobacterium PY146 for
PAHs was studied. Resting and growing cell biotransformation of PAHs including
phenanthrene, anthracene and pyrene by Mycobacterium PY146 one metabolite i.e,
1,2-benzenedicarboxylic acid (phthalic acid) from phenanthrene and two metabolites,
1,2-benzenedicarboxylic acid and 9,10-anthracendione from the anthracene
biotransformation, where as two metabolites, 1,2-benzenedicarboxylic acid and 1hydroxypyrene were detected and identified from pyrene biotransformation. While in
growing cell biotransformation of phenanthrene, 1,2-benzenedicarboxylic acid and
benzeneacetic acid, from anthracene 9,10-anthracenedione, benzenecarboxylic acid,
benzeneacetic

acid

and

1,2-benzenedicarboxylic

acid

and

from

pyrene

benzenecarboxylic acid, benzeneacetic acid and 1,2-benzenedicarboxylic acid were


detected and identified.
Biodegradation of PAHs including phenanthrene, anthracene and pyrene were
checked in soil by Mycobacterium PY146 without and with yeast extract. After one
week of incubation the detectable concentration of PAHs without yeast extract,
phenanthrene (98 g), anthracene (60 g) and pyrene (56 g) were decreased to 35,
18 and 27 g, and after 2 weeks of incubation all of the detectable PAHs were 0.14,
0.45 and 0.19 g respectively. There was no significant effect of yeast extract on
degradation except phenanthrene, which decreased from 98 to 27 g after one week.
Mineralization studies of different 14C-PAHs (naphthalene, phenanthrene, pyrene and
benzo(a)pyrene) were carried out by Bacillus cereus KWS2 and Mycobacterium
PY146. In case of naphthalene and phenanthrene mineralization by Bacillus cereus
KWS2, no mineralization was observed. Mineralization of phenanthrene, pyrene and
benzo(a)pyrene by Mycobacterium PY146 showed that 45.92, 39.89 and 4.82 % of
14

CO2 was produced and detected in the form of Na214CO3.

The mineralization of

14

C pyrene and phenanthrene with the different cell densities

from OD600 0.1 to 0.5 of Mycobacterium PY146 cultures showed that the maximum
mineralization of pyrene was observed in the OD600 0.2 culture, while in the case of
xi

phenanthrene there is increase in mineralization between the OD600 0.1 to 0.2 culture.
There was subsequent decrease in both pyrene and phenanthrene mineralization as the
cell density increases.
The expression of pyrene dioxygenase gene (nidAB) from Mycobacterium PY146
was checked in E. coli by cloning the nidAB in pKK223-3 expression vector and then
transformed the E. coli. The biotransformation of naphthalene, phenanthrene,
anthracene and pyrene was checked using transformed E. coli cells containing nidAB.
Neither any hydroxy nor any derivatized metabolite were detected of any PAH after
derivatization

with

N,O-Bis(trimethylsilyl)trifluoroacetamide

with

%Trimethylechlorosilane (BSTFA).
Quantification of nidAB RNA transcripts in E. coli was done by RT-QPCR which
revealed an ~ 2- fold increase in RNA transcripts in one sample grown in the presence
of IPTG. RNA to DNA ratio of nidAB calculated by RT-QPCR was increased
between OD600 0.1 to 0.4 and then declined at and OD600 of 0.5. Relation of
phenanthrene and pyrene mineralization to RNA/DNA ratio with the OD600 from 0.1
to 0.5 showed the positive relationship between phenanthrene mineralization and the
RNA/ DNA ratio while in case of pyrene only the OD600 of 0.3 to 0.5 corresponds to
the ratio.
From the present study, it is concluded that soil is the rich source of microbes having
the degradative capability of pollutants like PAHs. Some microbes have the ability to
degrade and mineralize the PAHs efficiently while others can only transform but cant
mineralize. The expression and the RNA transcript of pyrene dioxygenase (nidAB)
copy number and RNA/DNA ratio correspond to the mineralization of PAHs.

xii

Chapter1Introduction

Polycyclic Aromatic Hydrocarbons (PAHs)


Extensive urbanization and rapid industrialization has resulted in the accumulation of
a large number of xenobiotic chemicals into the natural environment. The use of
numerous aromatic compounds in explosives, dyestuffs, pharmaceuticals and
pesticides, and the growth of industrialization have contributed very much in the
environmental pollution and have attracted considerable attention continuously. Many
aromatic hydrocarbons, polycyclic aromatic hydrocarbons, polychlorinated biphenyls,
nitroaromatic compounds, dioxins and their derivatives are highly mutagenic, toxic
and/or carcinogenic to naturally occurring microflora as well as to higher organisms
including humans (Singh and Jain, 2003).
Polycyclic aromatic hydrocarbons (PAHs) are organic substances composed of carbon
and hydrogen atoms grouped into at least two condensed aromatic ring structures.
These compounds have low water solubility, as the number of benzene ring and
molecular weight increases the solubility and degradability decreases (Table 1.1). In
optimum laboratory conditions, half life values of approximately less than three days
for two ring naphthalene have been seen in the soil and water mixture in aerobic
conditions. Half life values for high molecular weight PAHs varies from 30 to 300
days (Al-Turki, 2009).
Sources and Occurrence of PAHs in the environment
Polycyclic aromatic hydrocarbons are produced naturally and by human activities,
therefore they are found commonly everywhere in the environment (Mueller et al.,
1996). Fossil fuels like crude oil and coal are the natural sources, while the partial
burning of the organic matters like vegetation, wood, coal and mineral oils are best
studied examples of PAHs production to the environment by humans (Freeman and
Catell, 1990; Lim et al., 1999). Automobiles are the major contributors to produce
PAHs. The quantity of these contaminants is dependent on the types of reactions and
the fuels used. The concentration of PAHs emission is higher in case of diesel
combustion as compared to petrol or their blended mixture with methanol (Stenberg et
al., 1983). The US EPA include 16 PAHs in their pollutant list, out of these 10 PAHs
are of high molecular weight. These are benzo(a)pyrene, benzo(a)anthracene,
benzo(b)fluoranthene,

benzo(k)fluoranthene,

benzo(g,h,i)perylene,
1

Chapter1Introduction

dibenz(a,h)anthracene, chrysene, fluoranthene, pyrene, and indeno(1,2,3-cd)pyrene


(Yan et al., 2004).
Crude oil spills and the industrial wastes generated during coal gasification and
liquefactions for the production of coke are the point sources of PAHs origin (Husian,
2008). Coal tar and creosotes (contain about 85 percent of PAHs), produced during
the coke production have the large amount of PAHs. Burnt food and cigarette smokes
are also the cause of PAHs generation in smaller volumes (Bamforth and Singleton,
2005).
The existence of contaminants like PAHs in soil and sediments pose considerable
hazards to that soil due to the toxicity, mutagenicity and, in some cases,
carcinogenicity associated with the PAHs (Cerniglia and Heitkamp, 1989; Patnaik,
1992). Chemical pollutants present in different parts of the environment slowly
change into more or less harmful products due to natural processes. The process of
degradation depends on various factors particular to the environment of pollutants and
susceptibility of these compounds to degradation (Dabrowska et al., 2005).

Chapter1Introduction

Table 1.1: Structures, Formulas, MW and solubilities of some priority polycyclic aromatic
hydrocarbons

identified

by

the

US

EPA

(Keith

and

Telliard,

1979).

(http://chemfinder.cambridgesoft.com/)
PAH

Structure

Formula

MW

Aqueous
Solubility(mg/L
H2O)

Naphthalene

C10H8

128.1732

31.7

Phenanthrene

C14H10

178.233

1.18

Anthracene

C14H10

178.233

0. 043

Pyrene

C16H10

202.255

0.013

Benzo(a)

C18H12

228.2928

0. 014

C20H12

252.3148

0. 0038

C22H12

276.3368

0. 00026

Anthracene

Benzo(a) pyrene

Benzo(g,h,i)
perylene

Chapter1Introduction

Environmental Hazards of PAHs


Many PAHs and their oxidative metabolites are highly toxic, mutagenic and
carcinogenic to microorganisms as well as to higher organisms including humans
(Samanta et al., 2002). PAHs have the tendency to accumulate in the organisms at the
end of the aquatic food chain. PAHs and their oxidative metabolite are reported in
many types of samples from different sources (municipal and industrial waste water,
drinking water and rainwater) (Sabate et al., 2001; Maier et al., 2000; Jamroz et al.,
2003). More angular ring structures of the high molecular weight PAHs results in
increased hydrophobicity (Zander, 1983; Harvey, 1997). In general, HMW PAH
molecules are more hydrophobic, toxic and persist longer in the environment
(Cerniglia, 1992).
PAHs are soluble in lipids and easily absorbed from the gastrointestinal tract of
mammals (Cerniglia, 1984). PAHs are thought to be genotoxic only when they are
activated by mammalian cytochrome P450 monooxygenase enzyme, which oxidizes
the aromatic ring to form reactive intermediates such as epoxide and diol epoxide
(Harvey, 1996). The resultant epoxides demonstrated carcinogenicity (Goldman et al.,
2001). Naphthalene can cause acute poisoning in humans may lead to haemolytic
anaemia and nephrotoxicity. Occupational exposure to naphthalene cause dermal and
some ophthalmological changes and have been observed in workers. Phenanthrene is
known to be a photosensitizer of human skin, a mild allergen and also mutagenic to
bacterial systems under specific conditions (Mastrangela et al., 1996). Little
information is available for other PAHs such as acenaphthene, fluranthene and
flourene with respect to their toxicity in mammals. However, the toxicity of
benzo(b)fluoranthene, benzo(a)anthracene, benzo(a)pyrene, benzo(k)fluranthene,
dibenz(a,h)anthracene and indeno(1,2,3-c,d)pyrene has been studied and there is
sufficient experimental evidence to show that all these are carcinogenic (Liu et al.,
2001; Mastrangela et al., 1996).
Fate of PAHs in Environment
The photo oxidation, adsorption to organic matters and microbial degradation are the
probable routes of PAHs in the environment. Beside this they may be deposit in
adipose tissues (Harvey, 1996).
4

Chapter1Introduction

Once PAHs are in the soil, they are subjected to a variety of natural weathering
processes. Photooxidation is one of these processes but is of limited value because
once the source of PAHs spilled in the soil, it usually seeps into the soil away from
sunlight (Cofield et al., 2007). Another process that may take place is dissolution.
This process has low significant because much of PAHs have only minimal solubility
in water. Another mechanism that may determine the fate is the adsorption of PAHs
to soil particles. PAHs have characteristic hydrophobic nature; as a result, these
compounds have a tendency to be adsorbed on soil particles, especially on the organic
fraction of solids (Al-Turki, 2009). This type of binding can make the PAHs
unavailable to biological systems hence there is reduction in toxicity but inhibition in
biodegradation. The reason for this inhibition is that they are partially insoluble in the
medium and are not available for degradation, since the bacteria are well known to
degrade those chemical compounds which are soluble in water (Al-Turki and Dick,
2003). The sequestrations are aging stages in which hydrophobic adsorbed substances
show a declining availability to both remediation and chemical extraction (Brion and
Pelletier, 2005).
The primary compounds of PAHs are converted to different metabolites by the
process of degradation before the process reaches an end (Dabrowska et al., 2005).
There are some reported examples of these intermediate metabolite of PAHs which
stimulate the formation of nucleic acid adducts and result in hindering the proper
regulation of genetic material and cause mutations. After entering the body, these
compounds are oxidized in the liver and transform the no polar PAHs to hydroxylated
ones. The dihydrodiols produced during naphthalene biotransformation by microbes
are more toxic than its parent compound because they are more soluble in water
which increases their availability (Wison et al., 1996).
Photodegradation
Photodegradation is a process which occurs in the presence of light and actively takes
place in environment (air and on water surface and soil), but is inefficient deeply in
soil and sediments.
In aquatic environment, chemical and photo oxidation are important routes for abiotic
degradation of PAHs. High molecular weight PAHs having condensed and complex
5

Chapter1Introduction

ring structure are usually oxidized by photolysis easily and more rapidly than low
molecular weight compounds. The angularity of the PAHs also affects the photo
oxidation rate, as the linear PAHs such as anthracene are more photo reactive than
phenanthrene (Kochany and Maguire, 1994). Photo oxidation is thought to be the
initial reaction to oxidize the PAHs partially which facilitate the microbes to degrade
it further. The metabolites produced by photo oxidation are thought to be more
biodegradable (Letho et al., 2000). 9, 10 anthracenedione a photo oxidation product
of anthracene is easily degraded by bacteria (Papadoyannis et al., 2002).
Biodegradation
Polycyclic aromatic hydrocarbons oxidizing microbes are present in the environment
everywhere, which includes soil, sediments and lignin containing wood material (Tam
et al., 2002). During the past decade, novel degradation pathways for polycyclic
aromatic hydrocarbons have been proposed by different isolated microbes from the
environment and characterized for their mineralization ability. Some biochemical
pathways for PAHs biodegradation by bacterial isolates have been investigated and
reported, such as for naphthalene (Resnick et al., 1996), phenanthrene (Kiyohara et
al., 1994) and for anthracene and acenaphthene (Dean-Ross et al., 2001; Pinyakong et
al., 2004).
The new approaches of advancements in molecular biology can be used for the
detection of PAH degrading microbes from different environmental samples
(Watanabe, 2001). Technique of DNA-DNA hybridization has been used to identify
and monitor the important crucial populations recovered from the soil environment
(Sayler et al., 1985; Guo et al., 1997). Laurie and Jones, (2000) detected two different
PAH catabolic genotypes from the contaminated soil with PAHs by using quantitative
competitive PCR (QC-PCR). A microbial consortium from soil having the ability to
efficiently mineralize benzo(a)pyrene was analyzed using denaturing gradient gel
electrophoresis (DGGE) profiling of PCR amplified 16S rDNA fragments (Kanaly et
al., 2002a,b).
In the biodegradation mechanisms molecular oxygen play an important role to start
oxidation of the benzene ring of PAHs by enzymatic action. In the very first step,
dioxygenase catalyzed oxidation of arenes normally occurs in aerobic conditions by
6

Chapter1Introduction

bacterial multicomponent enzymes to produce cis dihydrodiols. The resulted


hydroxylated compounds are further cleaved by extradiol or intradiol ring cleaving
dioxygenase enzymes to ortho or meta cleavage pathway, which leads the PAHs
metabolites to further conversion to the intermediate compounds of the TCA cycle
(Cerniglia, 1992; Gibson and Parales, 2000; Figure 1.1). The degradation of a
particular PAH by bacteria from a mixture of compounds is often directed by the
differences in their bioavailability. The bioavailable compound will be degraded first
by bacteria. The microbes isolated from environment such as bacteria normally
metabolize a very limited number of PAHs. Bacteria degrade selectively low
molecular weight PAHs first from the mixture. High molecular weight PAHs having
more than four benzene ring like benzo(a)pyrene metabolisms by bacteria is
comparatively less studied. The oxidation of two and three rings PAHs (like
naphthalene, anthracene and phenanthrene) by dioxygenase enzymes of bacteria is
well studied (Peng et al., 2008). Kim et al., (2004) explained the degradation pathway
of phenanthrene which is initiated by dioxygenase (NidAB) (Figure 1.2).
Microbial Metabolism of High Molecular Weight PAHs
The presence of pollutant chemicals in the environment along with microbes triggers
specific genes to be induced which produce enzymes enhance the biodegradation
process (Chen and Aitken, 1999). To overcome the low solubility of hydrophobic
compounds like PAHs, some microbes produce surfactants having both hydrophilic
and hydrophobic parts to enhance the solubility of PAHs (Georgiou et al., 1992).
These surfactants are classified as anionic, cationic, nonionic and zwitter ionic on the
basis of charge on hydrophilic portion (Kosaric, 2001; Christofi and Ivshina, 2002).
The function of surfactants depending on their structure affects the biodegradation by
different ways. As high molecular weight surfactants normally bind strongly to the
different surfaces where as the reduction of surface tension is achieved by the
surfactants having low molecular weight (Ron and Rosenberg, 2001). The presence of
some PAHs mixture also affects the bacterial metabolism by enhancing or inhibiting
its rate (Juhasz et al., 2004) and some PAHs degradative metabolites even inhibit the
growth of some bacteria (Said et al., 2000). The biodegradation of PAHs and its
metabolism by individual bacterial strain or consortium can also be affected the use of
different substrate and bacterial strains (Sutherland et al., 1995).
7

Chapter1Introduction

Microbial metabolism of PAHs occurs through oxidation/reduction reactions in which


transfer of electrons take place to the terminal electron acceptors from the substrates
which are highly reduced. In PAHs metabolism, different mechanisms occur in which
these compounds are completely mineralized, co metabolized or partially
transformed. When the PAHs are completely converted to inorganic molecules like
CO2 and H2O, the process is known as mineralization. Fungi mostly transformed the
PAHs to their hydroxylated aromatic acids. Some microbes use one for their energy
and carbon needs for metabolism and transforming another compound like PAHs to
reduce its size without using as source of carbon and energy, the mechanism known
as co metabolism (Alexander, 1999).
Biodegradation of PAHs and their metabolism by microbes leads to the reduction in
concentration which can be analyzed and quantify by different techniques like High
performance liquid chromatography and Gas chromatography coupled with mass
spectrometry. The presence of different microbial enzymes in the medium like
dioxygenase (Cenci and Caldini, 1997), mono oxygenase (Bezalel et al., 1997) and
laccases by fungi (Srinivasan et al., 1995) also indicate the PAHs metabolism.
Microbes or their genes can serve as biomarkers for the detection and metabolism of
different contaminants like PAHs, like Luciferase (lux) gene of several bacterial
species, Vibrio fischeri and Photobacterium, while the presence of lipase activities in
soil is an indication of oil degradation (King et al., 1990; Margesin et al., 1999).
Biodegradation of organic pollutants and its rate of metabolism can be inhibited or
decreased by several factors including the concentration of the substrate, deficient
electron acceptor like O2 and N (Loehr, 1992). Adsorption to the particles of soil and
the less soluble nature of PAHs also affect the degradation rate because of their non
bioavailability (Mackay, 1997; Karimi- Loftabad et al., 1996). Nutrients, water
content and pH values are some critical factors for polluted soil remediation (Kim et
al., 2005; Rolling et al., 2002). The slow release of nonaqueous phase liquids
(NAPLs) phase PAHs to aqueous phase greatly affect biodegradation rate in
contaminated soil (Arora et al., 2009).

Chapter1Introduction

Anaerobic PAH Biodegradation


Polycyclic aromatic hydrocarbons are also present in anaerobic environments like
aquifers and marine sediments (Meckenstock et al., 2000; Ohkouchi et al., 1999). The
presence of PAHs in light NAPLs and dense NAPLs ultimately transfer and sink it to
deep in the soil where O2 is not available and needs anaerobic biodegradation. The
anaerobic degradation of PAHs is not completely known yet, but it is clear that
compounds having additional functional groups or substituted PAHs are more
biodegradable than non-substituted compounds. The additional electron donating
groups on substituted compound facilitate the initial aromatic ring oxidation to
provide O2 by splitting H2O molecule (Wilson and Bouwer, 1997; Hutchins, 1991).
In contrast to the metabolism of single aromatic ring hydrocarbons by anaerobic
biodegradation, less data is available of PAHs. However, anaerobic degradation has
observed mostly in microbes isolated from the sediment of marine environment. Little
is known about the initial activation reactions of these PAHs. This is another area of
intense current research activity (Al-Turki, 2009).
The ability of anaerobic biodegradation of PAHs by microbes in the absence of O2 is
now recognized and the naphthalene biodegradation pathway is proposed (Rockne
and Strand, 1998; Rockne et al., 2000). Now several studies reported that many
classes of PAHs were transformed and degraded in the absence of molecular oxygen
by using other electron acceptors like ferrous irons, nitrate or sulphate and also under
methanogenesis conditions (Widdle and Rabus, 2001). Biodegradation process can be
slower in exposed drier soils on hillocks than in sheltered moist area. Degradation is
slower in cold climates than in more temperate or tropical zones and the degradation
rates generally decrease with depth through soil profile (Al-Turki and Dick, 2003).

Chapter1Introduction

Figure 1.1: Mechanisms of action of aromatic dioxygenases. A) Aromatic ring hydroxylating


dioxygenase; B) extadiol ring cleavage dioxygenase; and C) intradiol ring cleavage
dioxygenase.

10

Chapter1Introduction

Figure 1.2: Phenanthrene biodegradation pathway in M. vanbaalenii PYR-1 proposed


by Kim et al., 2004.
11

Aim&Objectives

Aim & Objectives


The aim of the project was to study the biodegradation of polycyclic aromatic
hydrocarbons, was achieved by following objectives;
To select PAHs degrading isolates from crude oil contaminated soil and to
characterize them
To study the effect of pH and temperature on the biodegradation of anthracene
To study the growing and resting cell biotransformation of PAHs (naphthalene,
phenanthrene, anthracene and pyrene) by the selected microbes
To determine the mineralization rate of

14

C radiolabled PAHs (naphthalene,

phenanthrene, pyrene and benzo(a)pyrene) by the bacterial isolates


To investigate and identify the genes encoding enzymes for PAH degradation
To isolate, clone and express the pyrene dioxygenase gene of Mycobacterium
PY146 into E.coli
To investigate the relationship between the pyrene and phenanthrene
mineralization and expression of nidA gene of Mycobacterium PY146

12

Chapter2 LiteratureReview
Polycyclic aromatic hydrocarbons contamination
The industrial revolution in late 1700s started the increasing use and dependency of
mankind on fossil fuels to generate electricity, industrial processes and power
transportation. As a result the oil wastes and products can be found in different
compartments of the environment (Vilanova et al., 2001).
The wide spread occurrence of PAHs in the environment is mainly because of their
production and release from the partial combustion of all the organic materials. Once
the compound release to the air by any process, they transported to the far areas and
may deposit on soil particles, plants and in water bodies. The PAHs presence in all the
compartments of the environment creates a major risk to humans as well as to all the
living organisms (Van et al., 1997). The data of total emission of 16 PAHs to the
environment from different countries showed in 2004 that 520 giga grams per year
from biofuel, wild fire and consumer usage (56.7, 17.0 and 6.9 %, respectively) as the
major sources. China, India and United States were the top three with highest PAH
emissions amongst the studied countries (Zhang and Tao, 2009).
On March 24, 1989, the oil tanker Exxon valdez spilled around 11 million gallon of
crude oil in Alaska, the oil spread by tides, winds currents and contaminated
approximately 1200 miles shoreline and affect the sea mammals, birds and other
marine life by killing or injuring (U.S. coast guard, 1993). Peterson, (2001) reported
the affect of oil spill on the habitat and species, they observed the decrease in number
of crabs and sea aster, reduced eelgrass density, increase scavenging birds mortality
and reduction in the fishes population. The MT Tasman Spirit, which was carrying
approximately 62,000 metric tons of crude oil for the Pakistan Oil Refinery, was
ruptured in July 2003 and the spilled oil spread to Clifton beach, Karachi. Blood
profiles from the nearby residents a month after of the spill showed slightly increased
levels of lymphocytes and eosinophils, while the creatinine level in serum and urea
level in blood were in normal ranges. Some skin rash cases in children were observed
(Khurshid et al., 2008). The spill of Tasman Spirit affected badly the marine life and
heavier component of the crude oil such as PAHs were detected in Fishes, Crabs and
Shrimps, Crustaceans, Mollusks and Echinoderms along with the passing time. In
fishes out of 16 PAHs 15 were found and detected (naphthalene, acenaphthylene,
13

Chapter2 LiteratureReview
acenaphthene,

fluorene,

phenanthrene,

benzo(a)anthracene, chrysene,

anthracene,

fluoranthene,

pyrene,

benzo(b)fluoranthene, benzo(k)fluoranthene and

benzo(a)pyrene benzo(g,h,i)perylene and indeno(1,2,3-c,d)pyrene) (Siddiqi et al.,


2009). The oil spill also affected the marine plants (primary producers) in the coastal
area of Karachi, the most affected group were phytoplankton which is reduced in
abundance as well as in species diversity (Saifullah and Chaghtai, 2005). The acute
health effects were observed in the people exposed to the oil spill area, including
increased frequency of sore throat, sore eyes, headache, breathing difficulties and
nausea/vomiting (Janjua et al., 2006).
World Health Organization (WHO) identified the high concentrations of total
suspended particulates (TSP) in Lahore, Pakistan to be amongst the highest in the
world. The high concentration of TSP is because of the high intensity of polluting
sources (from the many small and large industrial activities, and a diversity of poorly
controlled combustion sources) as well as a lack of effective pollution control. Also,
the arid climate aggravates the high pollution levels, resulting in enormous levels of
suspended particulate matter (SPM). Hi-vol air sampling for yearly mean pollutant
concentrations like metals, ammonium, elemental and organic carbon, various anions
and PAHs associated with particles obtained in Lahore were found greater than one
order of magnitude when compared with those from Birmingham, U.K. These levels
are similar with findings in Karachi, Bombay and Calcutta. Vapours to particle ratios
of PAH in the hotter climate of Lahore were thought to be much higher (Smith et al.,
1996).
Even worse is the situation in cities like Peshawar. Air pollution is one of the major
problems of the city. The increasing population and their transportation means, buses,
cars and trucks produces black soot in very huge amount into the environment which
is the main source of PAHs. Roadside soils are heavily polluted by traffic activities. In
comparison with the PAHs concentration level of roadside soils in other countries
(Jordan, France and Czech Republic), the roadside soil of Peshawar is highly
contaminated with PAHs. Soil samples collected from different roadsides were
analyzed and the individual PAH concentration were ranged from 0.1 to 18.0 mg/kg.
Pyrene and phenanthrene were the most abundant amongst the other PAH
14

Chapter2 LiteratureReview
compounds. The highest concentration of PAHs was found where the traffic density
was highest; suggesting that the principal source of PAH in the roadside soils is
vehicle emissions (Khan et al., 2008).
The presence and detection of PAHs in food items like fishes, meat, milk, vegetables
and fruits increasing the potential for dietary exposure of human to these compounds.
The sources of compounds are predominantly from environmental pollution and some
food processing. The presence of PAHs on food, mainly like fruits and vegetables due
to contaminated soil, air and water mainly by processes of PAHs adsorption, their
deposition and bioaccumulation (Martens et al., 1997; Chung et al., 2008). In 2003,
Camargo and Toledo, reported the presence of different PAHs in the fruits and
vegetables including; cabbage, lettuce, tomato, grape, pear and apple. Among
different PAHs, benzo(a)anthracene was the most abundant being found in 89% of the
samples. There was no effect on locale in PAHs concentration in fruit while in case of
vegetables, more PAHs detected in roadway sites grown than in the rural area. Shen
and Zhu, (2007) observed the presence of 15 different PAHs in vegetable species and
demonstrated that highest PAHs burden was in leafy vegetables. Bishoni et al., (2006)
reported the presence of PAHs in different vegetables; spinach, cauliflower and potato
in the range of 59.6-194.3 g/kg and fruits; grapes, apple, pineapple, papaya in the
range of 25.5-51.7 g/kg. Ciemniak and Chrachol, (2008) identified 16 PAHs in
different cereal products i.e corn, musli, oats and barley and found the low levels of
PAHs in all the samples. Benzo(a)pyrene was detected in all the tested samples in a
range of 0.0216 g/kg. Salinas et al., (2010) observed different PAHs on the surface
of apples in Mexico and suggested that this deposition was from the contaminated air.
PAHs were also detected in water samples (Tap, river, well and sea water) and fruit
juices in China (Zhao et al., 2009).
DouAbul et al., (1997) reported the presence of 40 different un-substituted and alkyle
substituted PAHs in the fishes from Red sea. The concentration of the total PAHs was
422.1 ng/g of dry weight edible muscles of collected fish samples. The considerable
amounts of different PAHs were detected in the gall bladder and liver of the two
different species of fish (Pointet and Milliet, 2000). Reinik et al., (2007) studied

15

Chapter2 LiteratureReview
presence of different PAHs including benzo(a)pyrene in meat and meat products and
reported the high concentration of these compounds in smoky and grilled cook meats.
After exposure and absorption of the PAHs by the body, they excreted largely in urine
or feces as hydroxylated metabolites. In addition, because of their lipophilicity, some
portions of PAHs accumulate in the adipose tissue which then may be excreted in
human milk. The breast fed infants are more sensitive to these contaminants than
adults; even the slight concentration may affect the infants health. In a study by a
group in Japan observed and detected the PAHs in commercial and human milk in the
range of 0.232.01 and 0.192.15 mg/kg, respectively (Kishikawa et al., 2003).
Different PAHs in the human breast milk of patients in maternity home after a short
period of delivery and in commercial milk and in their products were detected
(Aguinaga et al., 2007; Hunter et al., 2010).
Toxicity of Poly Aromatic Hydrocarbons
The first recognized chemical carcinogens were the combustion products of coal. In
1775, Percival Pott reported the cancer in human from the coal soot in chimney
sweeps followed by studies in animals in the 1920s and the discovery of carcinogenic
polycyclic aromatic hydrocarbons (PAH), e.g., benzo(a)pyrene (BaP), in the 1930s
(Brown and Thornton, 1957; Searle, 1976). Several PAHs including BaP are reperted
and known to be mutagens, carcinogens, and developmental toxicants in humans.
PAHs, such as BaP suggested to be a carcinogen and involved in cancer development
in humans. The activation of estrogen receptors by PAHs are also reported in several
studies which produce estrogenic metabolites. The activation and induction of ER and
metabolites by PAHs could stimulate the cells propagation in estrogen sensitive cells.
Some PAHs (BaP and benzo(a)anthracene) at concentration of 100 nM or higher
could stimulate the proliferation of carcinoma in breast cells of human (Pliskova et
al., 2005).
Some PAHs are transplacental carcinogens in experimental bioassays, producing
tumors in the liver, lung, lymphatic tissues, and nervous system of the offspring
(Bulay and Wittenberg, 1971). The adduct formation of PAHs with DNA is a
biomarker and indication of increased cancer risk and used as measure for genetic
16

Chapter2 LiteratureReview
damage. The fetus is more susceptible to PAHs associated damages of DNA (10 fold)
than mother and the exposure to PAHs inutero increases the risks. There should be a
preventive measure to minimize the exposure of PAHs to women with pregnancy and
children (Perera et al., 2005). Human macrophages exposure to PAHs affects their
functional properties like adhesion and endocytosis, reduces the surface integrins and
triggers the apoptosis of macrophages (Grevenynghe et al., 2004). Mane et al., (1990)
studied the effects of PAHs on DNA synthesis of human and rat mammary epithelial
cells and observed the reduce activity. Liu et al., (2010) reported the effect of
occupational exposure of PAHs on foundry workers and found the increased amount
of 1-hydroxypyrene in urine sample and DNA strand breakage in the individuals.
The crude oil and their components have phytotoxic effects on plants (red beans and
corn) growth. The aliphatic hydrocarbons have non-significant effects, while the
PAHs such as naphthalene, pyrene and phenanthrene have the phytotoxic effects and
it increases with the increase in benzene rings (Baek et al., 2004). Low molecular
weight hydrocarbons, benzene, toluene, xylene, indene, styrene and naphthalene
inhibited the germination and growth of different plants (Henner et al., 1999). PAHs
and their N-heterocyclic derivatives were studied and found that PAHs derivatives
have more phytotoxic effect than the parent compounds against plant species Sinapis
alba, T. aestivum and P. vulgaris (Paskova et al., 2006).
Control of PAHs pollution by Biodegradation
The PAHs pollution and associated hazards to the environment could be neutralized
with some conventional strategies including isolation, alteration or removal of
pollutants. The process involves, for example excavation of polluted site soil or
incineration of contaminant compounds. In these processes, expensive technologies
are involved and also the transfer of pollutant from one phase to the other. While in
bioremediation, the use of biological organisms to control the pollution by
transforming the hazardous or toxic compounds to less or non hazardous form with
less energy, time and chemical inputs (Ward et al., 2003).
Microbes like fungi and bacteria degrade PAHs partially, oxidize co metabolically or
mineralized completely. Several bacterial species are reported having the ability to
17

Chapter2 LiteratureReview
utilize PAHs as the only source for energy and carbon. These bacteria degrade and
transform the PAHs compounds into simple molecules which can be used in central
metabolic pathways (Cerniglia, 1992). The hydrophobic nature of PAHs tends it to
bind with the soil organic matter and non aqueous phases in the environment which
results in the unavailability of the compounds to microbial degradation.
Microorganisms especially bacteria have evolved several different mechanisms to
adapt to these hydrophobic compounds in the polluted area. Some microbes produce
biosurfactant to enhance the availability of less bioavailable compounds, while others
show the chemotaxis to reach the PAHs which improve the biodegradation. The
coexistence of fungi and bacteria ubiquitously in the soil environment and the mutual
catabolic activities of fugal and bacterial species may lead to enhance the degradation
process of PAHs in the environment (Peng et al., 2008).
Fungal biodegradation
A wide variety of fungal species have been isolated from different ecological sources
and shown to metabolize the several recalcitrant organic compounds including PAHs
from two to six benzene rings. Many filamentous fungi have been reported to colonize
the solid substrate and tolerate high concentration of toxic compounds (Cerniglia et
al., 1985; Robinovich et al., 2004). Fungi metabolize the PAHs mainly by the two
ways which are mediated by ligninolytic (white rot fungi) and non ligninolytic fungi
(Hammel et al., 1992).
Fungi also detoxify PAHs and other pollutants by the formation of bound residues by
the process known as humification rendering them unavailable to human and other
organisms (Cerniglia, 1997). Humification is the process by which fungi transform
the organic compounds and incorporate it to the humic substances to reduce their
availability and toxicity (Bollag, 1992). Humic substances play a role as binding
ligand for stable linkage with pollutants and to detoxify the environment. The
humification process is initiated by fungal enzymes, phenol oxidases and peroxidases
(Bollag and Myres, 1992). The detoxification of PAHs by the aid of biological
enzymes and its coupling to humic substances also facilitate the other soil microflora
for further transformation of the compounds (Gianfreda and Rao, 2004).
18

Chapter2 LiteratureReview
The initial reaction of non ligninolytic fungi in the PAHs metabolism is to oxidize the
benzene ring of the compound to produce arene oxides by monoxygenase enzyme,
cytochrome P450 (Sutherland et al., 1995). The metabolism of PAHs by mammals
also occurred in similar way. Dioxygenase enzymes of bacteria oxidize the aromatic
ring of PAHs to produce cis dihydrodiols, while the monoxygenase enzymes of fungi
incorporate single atom of oxygen to the aromatic ring and subsequently produce
trans dihydrodiols (Jerina, 1983).
Ligninolytic enzymes (laccases and peroxidases) which oxidize the lignin found in
different organic matters especially in wood are produced by the white rot fungi
(Mester and Tien, 2000). The oxidation of wood and other organic material occurs
through reactions based on non specific radicals by these extracellular enzymes (Kirk
and Farrell, 1987). Peroxidase enzymes can be classified to manganese peroxidase
(MnP) and lignin peroxidase (LP) on the basis of their reducing substrate, the LP and
MnP are able to oxidize different PAHs (Mester and Tien, 2000). Phenol oxidase
enyme, Laccases, are also able of oxidizing PAHs (Bamforth and Singleton, 2005).
Chulalaksananukul et al., (2006) isolated a filamentous fungus Fusarium sp. E033
from plant leaves exposed to traffic emissions which have the ability to degrade and
tolerate the high concentration of BaP. A strain of Absidia fusca was isolated from a
pesticide contaminated soil which can remove the fluoranthene and anthracene very
efficiently from the medium approximately 89 and 90 % after 24 hours. Another
strain of A. fusca from uncontaminated soil source also removed the fluoranthene and
anthracene (Villemain et al., 2006). Silva et al., (2009) studied the ability of different
fungal isolates from the rich humic substance soil in microaerobic and very low
oxygen condition for the degradation of PAHs. Naphthalene and phenanthrene were
degraded efficiently by Aspergillus sp, Fusarium oxysporum and Trichocladium
canadense while T. canadense, Aspergillus sp, Verticillium sp and Achremonium sp.
were efficient in degrading chrysene, perylene and naphthol(2,3-a)pyrene.
A white rot fungus, Polyporus sp. S133 isolated from soil contaminated with
petroleum, degrade phenanthrene and BaP efficiently in soil. Polyporu sp. S133
transform the phenanthrene and 9,10-phenanthrenequinone and 2,2-diphenic acid
were detected from the medium, while one metabolite, benzo(a)pyrene-6,12-quinone
19

Chapter2 LiteratureReview
was also detected from BaP degradation (Hadibarata and

Tachibana, 2009;

Hadibarata, 2009).
Chupungars et al., (2009) reported the fungal isolates Agrocybe sp. CU-43 and
Xylaria sp. CU-1 having the degradation activity of different PAHs. The fungus
Agrocybe sp. CU-43, degraded fluorene, phenanthrene, anthracene, fluoranthene and
pyrene at higher rate than Xylaria sp. CU-1. Agrocybe sp. CU-43 degraded different
concentration of fluorene efficiently upto 750 ppm. Two metabolites 9-fluorenol and
9-fluorenone were detected in fluorene degradation medium, the fungus was able to
degrade 9-fluorenone as well. Siripong et al., (2009) used five previously isolated
fungus from two ecologically different sites to degrade anthracene and 2,4-PCB.
More than 90% degradation was observed by all the isolates.
A white rot fungus, Pleurotus ostreatus has the ability to mineralize catechol,
phenanthrene, pyrene, BaP, anthracene and fluorene at slower rate but did not
mineralize fluoranthene (Bezalel et al., 1996). Bishnoi, et al., (2008) reported the
PAHs degradation capability of Phanerochaete chrysosporium isolated from the
contaminated soil of petroleum refinery which degraded acenaphthalene, anthracene,
phenanthrene, fluoranthene and pyrene in both sterile and unsterile soil.
Two fungal isolates Irpex lacteus and Pleurotus ostreatus were used for remediation
studies of two different soils contaminated with PAHs. Fungal isolates actively
degraded PAHs in both soil samples, but there was little or no effect of surfactants
addition on the overall PAHs degradation (Leonardi et al., 2007). Canas et al., (2007)
reported the effect of natural and synthetic mediators on transformation of different
PAHs by fungal laccase of P. cinnabarinus ss3. Laccase degrade anthracene
efficiently, whereas the natural mediators significantly increased the pyrene and BaP
removal.
Bacterial degradation of PAHs
A number of phenanthrene degrading bacterial isolates from different environments
are reported (Phale et al., 2007). Numerous pyrene metabolizing microorganisms are
also isolated and identified which include gram positive actinomycete, S.
xinjiangenesis that use pyrene as carbon and energy source (Ting et al., 2004). A
20

Chapter2 LiteratureReview
thermo and halotolerant Bacillus strain DHT which could degrade and utilize pyrene
was also reported by Kumar et al., in 2006. Mueller et al., in 1989, reported the PAHs
can be degraded by different bacterial strain isolated from the contaminated soil in the
form of consortium. This bacterial consortium was able to efficiently metabolize the
high molecular weight PAHs like fluoranthene.
Seo et al., (2007) isolated nineteen different bacterial strains from soil contaminated
with petroleum where different PAHs and their metabolite were detected. Out of
these, five bacterial strains degraded PAHs and also used some organophosphorous as
growth substrate. Daane et al., (2001) isolated different bacteria from the rhizosphere
which were able to degrade different PAHs. The identification and characterization
revealed three groups of bacteria, Pseudomonads, Nocardioforms and Paenibacillus.
The isolation of Paenibacillus which showed no homology with different probes
suggested the novel PAH degrader. They concluded the presence of PAHs degrading
microbes in rhizosphere associated with plants and these can be utilized for
contaminated soil bioremediation.
Fourteen different bacterial isolates were used for the degradation of both
phenanthrene and anthracene. After identification four isolates Thiobacter
subterraneus (EF105547), Escherichia coli (EF105548), Soil bacterium (EF105549)
and Alcaligenes sp. (EF105546) showed efficient degradation activity (Abd-Elsalam
et al., 2009). A microbial consortium YL isolated from sediment samples by
enrichment with pyrene degraded pyrene, fluorene, phenanthrene and fluoranthene
from 8795.9 % after three weeks. Two bacterial strains Bacillus cereus Py5 and
Bacillus megaterium Py6 from the consortium degraded 65.8 and 33.7 % of pyrene,
respectively, in three weeks. Escherichia coli cells transformed with the plasmids of
the enriched consortium degraded 85.7 % of pyrene in three weeks (Lin and Cai,
2008).
Tian et al., (2002) observed that Pseudomonas mendocina strain CGMCC 1.766 grew
well in the presence of different concentration of phenanthrene (20-200 mg/l). They
also detected the 1-hydroxy-2-naphthoic acid as the intermediate metabolite during
bacterial degradation of phenanthrene. Feijoo-Siota, et al., in 2008 studied the
naphthalene biodegradation in sea water by free and immobilized cells of
21

Chapter2 LiteratureReview
Pseudomonas stutzeri 19SMN4 and reported almost complete degradation of
naphthalene both by free and immobilized cells after six days of incubation at 30C.
Obayori et al., (2008) isolated three different Pseudomonas species from oil
contaminated soil in Nigeria, which utilized a broad range of substrates including
chlorobenzoates, petroleum fractions and different PAHs. All three isolates LP1, LP5
and LP6 degraded pyrene, and removed 67.8, 66.6 and 47.0 %, respectively, from the
medium in 30 days. Pathak et al., (2009) studied the naphthalene degrading potential
of Pseudomonas sp. HOB1, isolated from polluted sediments of Amlakadi canal,
India. The isolated bacterial strain degraded the 2000 ppm concentration of
naphthalene in 24 hours and tolerated high concentration of naphthalene up to 60,000
ppm. A bacterial strain from oil contaminated soil was isolated and identified as
Pseudomonas sp. degraded 74.8 % anthracene in 10 days. The degradation activity of
the isolate was diminished after plasmid curing with acridine orange (Kumar et al.,
2010).
Dean et al., (2001) investigated and comparatively studied the bioavailability of
phenanthrene to two degrading bacteria (Pseudomonas strain R and isolate P5-2) in
the presence of rhamnolipid biosurfactant and with the presence of P. aeruginosa
which produce biosurfactant. There is more mineralization of phenanthrene sorbed
with soil by Pseudomonas strain R than strain P5-2, but in aqueous cultures medium
the rate and extent of phenanthrene mineralization by P5-2 was higher than
Pseudomonas strain R. In sandy loam (which has high phenanthrene sorption
capacity) the addition of rhamnolipid biosurfactant increased the rate phenanthrene
mineralization by Pseudomonas strain R, while the mineralization activity of strain
P5-2 sandy loam minimal and there is no increased mineralization by biosurfactants
addition.
A biosurfactant producing bacteria was isolated from petrochemical contaminated
site, identified as Brevibacillus sp. PDM-3 had the ability to grow on the expense of
phenanthrene, anthracene and fluorene. For phenanthrene degradation after the
optimization of different growth parameters, bacteria showed 93 % degradation in six
days (Reddy et al., 2010). Mallick et al., (2007) reported Staphylococcus sp. strain
PN/Y, isolated from petroleum contaminated soil with the ability to utilize
22

Chapter2 LiteratureReview
phenanthrene as the only source of carbon and energy. Various metabolites were
identified from the phenanthrene degradation along 2-hydroxy-1-naphthoic acid
suggesting the meta cleavage. The catabolic plasmid, pPHN in strain PN/Y is
identified and confirmed by the loss of degradation activity in the plasmid cured strain
PN/Y and the degradative activity of S. aureus RN4220 transformed with pPHN
plasmid. Burkholderia cocovenenans (BU-3) isolated through enrichment technique
by Wong et al., (2002) showed the high degradation rate of phenanthrene and reduced
more than 95 % concentration of the compound from the medium.
From the data reported on PAHs degrading isolates from soil, it has been observed
that the majority of the degraders belogs to Sphingomonas, which are able to oxidize
and degrade a variety of xenobiotic and natural chemicals, such as diphenylether,
furan, dibenzo-p-dioxin, carbazole, estradiol, polyethylenglycols, chlorinated phenols
and different herbicides, pesticides and biphenyl, naphthalene, fluorene, phenanthrene
and pyrene (Basta et al., 2005). Sphingomonas sp. GY2B isolated from the enriched
consortium from PAH contaminated soil, utilize several PAHs including
phenanthrene, naphthalene, 2-naphthol and salicylic acid as the only source of carbon.
The isolated strain degraded 91 % of phenanthrene in two days, and the metabolites
from phenanthrene degradation identified as 1-hydroxy-2-naphthoic acid, 1-naphthol
and salicylic acid (Tao et al., 2007).
Kasai et al., (2002) studied the bacterial role in the degradation of petroleum and
PAHs in aerobic conditons. They enriched bacteria from sea water by using different
PAHs like 2-methyl naphthalene, phenanthrene, or anthracene as a carbon and energy
source and isolated numbers of genus Cycloclasticus, which could grow on crude oil,
naphthalene, dibenzothiohenes phenanthrene and fluorenes. That bacterial group was
propagated on grains coated with oil and immersed in sea water in beach simulating
tanks. PAHs degradation was observed and as well as the growth of Cycloclasticus
cells on the oil coated grains.
Several Mycobacterium species have been isolated and reported to oxidize and
mineralize the low as well as high molecular weight PAHs. Sepic et al., (1997)
studied the biodegradation of PAHs with different isolates of bacteria derived from an
activated sludge. Mycobacterium sp. prp-I showed an efficient biodegradation
23

Chapter2 LiteratureReview
potential for three and four ring PAHs. Sepic et al., (1998) studied the biodegradation
potential of two different bacteria; Pasteurella sp. IFA and Mycobacterium using
fluoranthene. After two weeks of incubation at room temperature in aqueous medium
resulted approximately 24% and 46% reduction in fluoranthene concentration was
observed. During that period the bacteria mineralized approximately two third and
four-fifth of biodegraded fluoranthene to CO2 while one third and one fifth of the
original fluoranthene remained as stable metabolic products.
A Mycobacterium sp. isolated from the petrochemical contaminated sediment has the
ability to mineralize pyrene (48 % of pyrene in 3 days) as well as other PAHs like
naphthalene, phenanthrene and fluoranthene. The Mycobacterium sp. also mineralize
the PAHs (2-methylnaphthalen, phenanthrene, pyrene and benzo(a)pyrene ) in
microcosm (Heitkamp et al., 1988a; Heitkamp et al., 1988b; Heitkamp and Cerniglia,
1989). Moody et al., (2001) reported the degradation of phenanthrene and anthracene
by Mycobacterium sp. PYR-1. The bacteria degrade 90 and 92 % of phenanthrene and
anthracene, respectively, in 14 days. Mycobacterium pyrenivorans sp, a pyrene
degrader having a distinctive mycolic acid profile in the cell wall containing epoxymycolates, alpha-mycolates and omega-carboxymycolates and 2-eicosanol (Derz et
al., 2004). The alkaliphilic Mycobacterium which is able to degrade and metabolize
PAHs like pyrene, and utilize it for carbon and energy source (Habe et al., 2004a).
Churchill et al., (1999) reported the mineralization activity of Mycobacterium sp.
strain CH1 isolated from PAHs contaminated sediment of fresh water which
mineralized pyrene, fluoranthene and phenanthrene and used these compounds as
carbon and energy source. Mycobacterium sp. strain A1-PYR degraded pyrene alone
and the degradation activity was increased significantly when fluoranthene or
phenanthrene was added along pyrene (Zhong et al., 2006). Alkaliphilic
Mycobacterium sp. MHP-1 degrades pyrene and Mycobacterium sp. SNP11 utilizes
pyrene, phenanthrene, fluorene and fluoranthene as carbon source (Habe, 2004b).
Enzymes involved in Biodegradation of PAHs
Microbial degradation of PAHs is mediated by oxygenase enzymes, which add one or
two (mono or dioxygenase) hydroxyl groups to the aromatic ring. Several oxygenase
genes were reported from a variety of bacterial isolates in different studies.
24

Chapter2 LiteratureReview
Monooxgenases are mixed function enzymes which incorporate a single oxygen atom
on to benzene ring of the PAHs (Leveau et al., 1998) and several monoxygenase
enzymes were reported in bacteria for pyrene metabolism which are salicylate
hydroxylase (Katagiri et al., 1996), salicyldehyde dehydrogenase (Schocken and
Gibson) and salicylate 5-hydroxylate (Grund et al., 1992).
Cytochrome P 450, the oxidative enzymes present in bacteria and fungi but produce
slightly different oxidative products symmetrically as in fungi the dihydrodiol
produced will be trans and by bacteria the resultant dihydrodiol will be in cis position.
(Cerniglia et al., 1984). The cis dihydrodiol formation is extremely stereoselective by
dehydrogenases which is NAD dependent (Harayama and Rekik, 1989). The
production of cytochrome P450 enzyme could be inhibited or decreased by some
other liver enzymes which have the ability of PAHs detoxification (Leahy and
Colwell, 1990; Chen et al., 2004).
Microbial dioxygenase enzymes add two atoms of oxygen to the benzene ring of the
PAHs and are classified as ring hydroxylating and ring cleaving enzymes. The PAH
ring hydroxylation enzymes are NADPH dependent and oxidize the PAH ring and
transform it to substituted aromatics by addition of hydroxyl groups. Pyrene and
naphthalene dioxygenase enzymes are the well studied examples (Harayama et al.,
1992).
The extra and intradiol dioxygenase enzymes cleave the aromatic ring of PAHs by
addition of two oxygen and subsequently produced metabolites through meta and
ortho cleavage pathway. Catechols are cleaved by extradiols having single ferrous ion
in active site of a subunit (Harayama and Rekik, 1989). Protocatechuates are normally
cleaved by intradiols and have two sbunits ( and ) and the active site contain
varying ferric ions. The example of intradiols is the salicylate 1,2-dioxygenase from
pseudomonas (Kita et al., 1999; Husain, 2008).
Bacteria having the capability of naphthalene metabolism are found in the
environment enormously which are isolated and from some of these the dioxygenase
enzyme for naphthalene were purified and studied (Peng et al., 2008). Naphthalene
1,2-dioxygenase (NDO), from the Pseudomonas sp. NCBI 9816-4, catalyze the
25

Chapter2 LiteratureReview
reaction by adding two oxygen to the naphthalene and transform to cis naphthalene
dihydrodiol, using the non heme iron center. NDO consist of two main subunits, and
. subunit of the NDO have six mixed sheets and not involved directly in substrate
specificity or catalysis, while the subunit have 13 helices and 25 strands. Two
distinct domains of subunit, a non heme ferrous iron catalytic and a Rieske domain
having the (2Fe-2S) center. The NDO have a wide range of substrate specificity and
oxidize biphenyl and anthracene to their dihydrodiols (Kauppi et al., 1998; Parales et
al., 2000).
Kasai et al., (2002) showed that phnA gene of Cycloclasticus sp. strain A5 was able to
degrade several the PAHs. They transformed E. coli cells with cloned phnA1A2A3A4
genes had wide substrate specificity and converted several polycyclic including
naphthalene,

phenanthrene,

methylnaphthalene

and

hetetocyclic

aromatic

hydrocarbons dibenzofuran and dibenzothiophene to their subsequent dihydroxylated


compounds. Sphingomonas sp. strain LB126 harbor dioxygenase gene (FlnA1FlnA2), which is responsible for initial attack on fluorene during microbial
degradation. FlnA1-FlnA2 was cloned in E. coli, and transformed cells were able to
oxidize fluorene, Phenathrene as well as some other PAHs (Schuler et al., 2008). The
phdABCD genes cluster encodes phenanthrene dioxygenase in which A encodes for
Large subunit , B for small subunit , C for ferrodoxin and D for ferrodoxin
reductase, from Nocardioides sp. were transferred into Streptomyces lividans. The
transformed cells of S. lividans efficiently coverted the phenanthrene and 1Methoxynaphthalene to their subsequent diol forms (Chun et al., 2001).
Shindo et al., (2001) reported the activity of phenanthrene, toluene biphenyl
dioxygenases genes cloned in E. coil. The transformed E. coil cells with phdABCD
from the Nocardioides sp. KP7 converted the phenanthrene and phenanthridine to
their hydroxylated product. On the other hand E. coli having toluene dioxygenase
(todC1C2BA) from P. putida F1 or biphenyl dioxygenase (bphA1A2A3A4) from P.
pseudoaligenes KF707 were not able to transform their substrates. When E. coli were
transformed with the hybrid of toluene and biphenyl dioxygase substituted subunits
(todC1:bphA2A3A4), the bacteria were able to oxidize dibenzofuran, fluorene and
dibenzothiophene. Pinyakong et al., (2004) reported degradation activity of
26

Chapter2 LiteratureReview
Sphingomonas sp. A4 that use only acenaphthylene and aceanphthene but not other
PAHs. Genes arhA1 and arhA2 which encode the terminal oxygenase of the ring
hydroxylating dioxygenase were transferred in E. coli cells having the electron
transport protein from Sphingobium sp. P2 having phenanthrene degradation ability.
The recombinant E. coli have specificity towards several PAHs like naphthalene,
phenanthrene, fluoranthene, acenaphthene and acenaphthylene.
In Mycobacterium, pyrene biodegradation is initiated by a pyrene ring-hydroxylating
dioxygenase (nidAB), which encodes and -subunit (Khan, et al. 2001). In
Mycobacterium, nidB is upstream of nidA. This arrangement is unique and
characteristic in the dioxygenase of PAHs degraders. nidAB from the Mycobacterium
vanbaalenii PYR-1 when cloned in E. coli, were functional and transformed the
pyrene to its dihydrodiol form. Since the reductase and ferrodoxin fragments are
required for the electron transfer in the dioxygenase system, dioxygenase might have
borrowed the redutase and ferrodoxin from elsewhere in the cloned E. coli. Brenza et
al., (2003) reported the presence of more than one copies of nidAB gene in other
mycobacterium species which have more than 98 % homology with published
sequence of nidAB of Mycobacterium vanbaalenii PYR-1.
Mycobacterium vanbaalenii PYR-1 possess several different dioxygenases and their
multiple copies besides nidAB. Another dioxygenase gene, NidA3B3 was identified
and isolated from Mycobacterium vanbaalenii PYR-1. The expression of this gene in
E. coli did not show any dioxygenase activity, unless coexpressed with ferrodoxin and
reductase. The recombinant E. coli transformed high molecular weight PAHs such as
pyrene, fluoranthene, benz(a)anthracene, anthracene and phenanthrene to their
dihydrodiols (Kim et al., 2006). Stingley et al., (2004a) identified a phthalate
degrading operon (pht) in Mycobacterium vanbaalenii PYR-1. The putative operon
was 12-19 kb from the nidA. Five different genes in the operon were cloned and
expressed in E. coli, efficiently degraded phthalate to dihydroxyphthalate. For the
identification of additional genes responsible for PAH degradation beside nidAB and
nidA3B3, genomic library of Mycobacterium vanbaalenii PYR-1 was constructed in
fosmid vector. A nidA positive fosmid clone had numerous genes which were
involved in polycyclic aromatic hydrocarbon degradation. The transformed E. coli
27

Chapter2 LiteratureReview
with fosmid vector converted the phenanthrene to phenanthrene cis-dihydrodiol
(Stingley et al., 2004b).
Sho et al., (2004) reported two distinct gene loci nid and pdo in Mycobacterium sp.
strain S65 which were homologous to nidA. The organizations of catabolic genes
were similar in both loci. In both loci, the nidA homologues have 89 and 99 %
homology with the nidA of Mycobacterium PYR-1. The pyrene catabolic genes of
both loci were expressed only in the presence of phenanthrene and pyrene. Strain S65
degradeg and mineralized the PAHs very efficiently and that activity was thought to
be because of two different set of catabolic genes. A pyrene degrading
Mycobacterium sp. strain CH-2 isolated by Churchill et al., (2008) had two pyrene
dioxygenase genes pdo2AB2 and nidAB and expressed only in the presence of PAHs.
The strain CH-2 mineralized several PAHs including fluoranthene, pyrene and
phenanthrene.
Klankeo et al., (2009) reported that Diaphorobacter sp. and Pseudoxanthomonas sp.
isolated from soil were efficient degraders of PAHs. These isolates had nidA gene on
the plasmid, which was the first report about gram negative bacteria having the nidA.
Beside different dioxygenases, the cytochrome P450 monoxygenase genes
responsible for addition of a single oxygen to an aromatic ring of the PAH in initial
oxidation in the degradation process was reported in Mycobacterium vanbaalenii
PYR-1 by Brezna et al., (2006). Three cytochrome P450 genes CYP151, CYP150 and
CYP51 were identified in strain PYR-1 as well as in other Mycobacterium isolates.
The expression studies in E. coli transformed with CYP151 and 150 showed the
monooxygenation of pyrene and other PAHs.

Bioremediation
Bioremediation is the process by which the hazardous and toxic compounds can be
removed or converted to nontoxic or less toxic forms by the use of biological systems,
such as microbes, which are known for their capabilities of catabolic processes. This
is an efficient and cost effective approach to reduce or completely remove the
problematic organic chemicals from the contaminated environment such as soil and
28

Chapter2 LiteratureReview
water by microbes. Different type of organic compounds could be removed, but
PAHs, some crude oil products and the derivatives of some organo chlorine can be
removed frequently by bioremediation. The bioremediation process will be successful
if the pollutants are removed or transformed to nonhazardous compounds. The
process lowers the cost and eliminates the need for treating the contaminated soil at
the dumpsite. Under many circumstances, bioremediation is the only economical and
available technology (Dabrowska et al., 2005).
In bioremediation process bacteria, fungi and algae are involved in degradation
through biotransformation and mineralization (into water and carbon dioxide). The
rate of bioremediation depends on the chemical structure, availability and nature of
the pollutant, environmental conditions such as temperature, pH and oxygen. The
abundance, activity and types of microbes also affect the rate of bioremediation
(Singh and Ward, 2004). Many PAHs and other organic compounds are ubiquitous in
the environment and persist longer because of their low solubility and thus are less
bioavailable to microbial degradation. The persistent nature and associated toxic
effects of PAHs result in extensive research on degradation capabilities of
microorganisms led to biodegradation pathways and the knowledge about presence of
PAHs degrading microbes in the environment. These studies increasd the interest to
overcome and remediate the PAHs polluted environment by the use of microbes
(Bamforth and Singleton, 2005).
Soil contaminated with polycyclic aromatic hydrocarbons and phenols can remediate
with biostimulation. Guerin, (1999) reported the removal of PAHs and phenol from
the contaminated soil by the process of stimulation which involves the aeration by soil
mixing and adding low amounts of fertilizers. The microbial population of the soil
increased with biostimulation and the total PAHs and phenol concentration were
significantly decreased in the soil, which confirmed that biostimulation was an
effective process for the remediation of PAHs related contaminated soil. In
bioremediation of PAHs contaminated soil, different factors such as moisture,
surfactants and nutrients addition increased the reduction of PAHs concentration
(Lilia et al., 2008).

29

Chapter2 LiteratureReview
Biostimulation and bioaugmentation processes were thought to be effective in
remediation. A comparative microcosm study of PAHs contaminated soil revealed
that the concentration of PAHs rapidly decreased in bioaugmented contaminated soil
with fungus Molinia sp. compared to biostimulated soil with ground corn cob (Wu et
al., 2008). Cunningham and Philp, (2000) also reported that bioremediation rate of
diesel contaminated soil was higher in bioaugmentation process with the inoculation
of enriched microbes than the biostimulation strategy by adding fertilizers. Yu et al.,
(2005) reported that bioaugmentation of contaminated sediment slurry with PAHs
degrading bacteria showed no degradation activity, while biostimulation with mineral
salt medium increased the degradation rate.
In bioremediation processes, nutrients are one of the limiting factors, but not in every
case. The addition of nutrients alone to the contaminated soil with coal tar PAHs did
not show any decrease in total PAHs in the lab and also in the field studies, while
addition of biodiesel and nutrients to the contaminated soil increased the removal rate
of PAHs and enhanced bioremediation process by solubilizing unavailable PAHs
(Taylor and Jones, 2001). Sediments contaminated with PAHs amended with nutrients
such as phosphorous and nitrogen, yeast extract, lactate and surfactant tween 80
individually increased the degradation of PAHs, while use of these agents altogether
except yeast extract increased the PAHs degradation rate up to 8 times (Bach et al.,
2005). Gennaro et al., (2008) also observed the addition of tween 80 and inoculum
increased the degradation rate in the bioremediation of PAHs in slurry phase
bioreactor.
A bacterial isolate, Paracoccus sp. strain HPD-2 having PAHs degradation ability
was introduced to the soil contaminated with PAHs in microcosm study. The higher
counts of bacteria, enzymatic activities and microbial biomass as well as reduced
concentration of total PAHs was observed in the bioaugmented microcosm (Teng et
al., 2010). The use of genetically engineered microorganism (GEM) for
bioremediation and to monitor the PAHs contaminants was first introduced by a group
in the University of Tennessee. The P. fluorescence strain HK44 transformed with
naphthalene catabolic plasmid containing lux gene fused in the promoter for catabolic
gene of naphthalene. During degradation of naphthalene lux gene gave the
30

Chapter2 LiteratureReview
luminescent signal, which could be used for in situ bioremediation monitoring and for
the detection of PAHs like naphthalene in the environment (Cox et al., 2000; Ripp et
al., 2000).

31

Chapter3MaterialsandMethods

3.1 CHEMICALS
Analytical grade pyrene (99 %), anthracene (99 %), naphthalene (99 %) and N,OBis(Trimethylsilyl) trifluroacetamide with 1 % trimethylchlorosilane, phenanthrene
14

(98+ %), indole, ampicillin and

C-labled PAHs (benzo(a)pyrene, pyrene,

naphthalene, phenanthrene and anthracene) were purchased from Sigma-Aldrich and


Fluka (St. Louis, MO, USA). Analytical grade Isopropyl -D-1-thiogalactopyranoside
(IPTG) was purchased from Anatrace (USA). Other chemical and media components
were of high purity and analytical grade from Oxide, Merck and BDH.

3.2 SELECTION OF PAH DEGRADING BACTERIA


3.2.1 Culture condition and media
Nutrient agar and Mineral salt medium was used for initial screening and
transformation experiments. The composition of PNR and PNRG (PNR + 5 mM
glucose) (Khan et al., 2006) is as follows;
Composition of mineral salt medium (PNR) per liter of distilled water
PN (20x) 50 mL used as 50 mL/L
KH2PO4

13.6 % (w/v)

(NH4)2SO4

2.4 % (w/v)

NaOH

2.5 % (w/v)

R salts used as 7ml/L


MgSO4.7H2O

8 % (w/v)

FeSO4.7H2O

0.2 % (w/v)

HCl

0.4 % (w/v)

Agar (1.5 %) was used as solidifying agent.


Basal salt medium (BSM) was used for transformation expreriments by resting
cells. The composition of BSM is as follows;

32

Chapter3MaterialsandMethods
Composition of BSM
NaNo3

4.0 g

KH2PO4

1.5 g

FeCl3

0.005 g

MgSO4

0.2 g

CaCl2

0.01 g

Na2HPO4

0.5 g

dH2O

1000 mL

Media and apparatus were sterilized in an autoclave at 121C and 15 psi for 20
minutes.

3.2.2 Screening of PAHs degrading bacteria


A total of 52 bacterial isolates from the crude oil polluted soil (Malik, 2009) were
grown on nutrient agar plates containing naphthalene (100 ppm) and anthracene (1001200 ppm).
Fifteen strains showing the best growth on nutrient agar plates with anthracene (800
ppm) were screened on PNRG (PNR + 5 mM glucose) and PNR media plates
containing anthracene and pyrene separately, (800-1200 ppm each respectively) as
sole source of carbon and energy. Five isolates were selected on the basis of their best
growth in solid medium with PAHs.

3.2.3 Identification of selected bacterial isolates


3.2.3.1 Morphological and biochemical characterization
Selected isolates were characterized by colony morphology on nutrient agar, gram
staining and morphological characteristics. Additional biochemical test were
performed according to Bergeys Manual of Determinative Bacteriology (Brenner et
al., 2005) for taxonomic characterization which included gelatine liquefaction, starch
hydrolysis, lipid hydrolysis, casein hydrolysis, H2S production, triple sugar iron,
nitrate reduction, catalase test, lipase test, oxidase test, methyl red, voges proskauer
(MR-VP), citrate utilization and sugar fermentation.

33

Chapter3MaterialsandMethods
3.2.3.2 Molecular characterization based on 16S rRNA gene
Genomic DNA was extracted from strains using the Wizard Genomic DNA kit
(Promega, Madison, WI, USA) using the gram positive bacteria protocol according to
the manufacturer instructions. 16S rRNA genes were amplified using universal
eubacterial primers 8F (5-AGAGTTTGATCMTGGCTCAG) and 1492R (5GGYTACCTTGTTACGACTT-3) (Lane et al., 1985). 25 L PCR reactions
consisted of PuRe Taq Ready-to-go PCR beads (GE Healthcare Buckinghamshire,
UK), 0.5 L of template genomic DNA, 400 nM of each primer. The thermocycler
program consisted of 95C for 10 minutes followed by 40 cycles of: 95C for 30
seconds, 48C for 30 seconds and 72C for 1 minute 30 seconds. The program ended
with a final extension step of 72C for 10 minuntes.
PCR products were column-purified using QIAquick PCR Purification Kit (Qiagen
Valencia, CA, USA). They were sequenced using the universal primer 519F (5CAGCAGCCGCGGTAATAC) on an ABI 3730 DNA Analyzer at the Molecular
Biology Resource Facility at the University of Tennessee, Knoxville, TN, USA.
High quality sequences (~950 bp) imported into the DNASTAR Lasergene v.7
software package, aligned using MegaAlign, and a phylogenetic tree was constructed
using ClustalW (Thompson et al., 2003) program contained within Lasergene.
Pairwise distances between sequences (used for phylogenetic tree construction) were
also exported in tabular format.

3.3 BIODEGRADATION STUDIES OF PAHs


3.3.1 Biotransformation of anthracene and naphthalene using B.
cereus KWS2 and P. stutzeri 10-1
Two selected bacterial isolates (KWS2 and 10-1) were evaluated for their ability to
transform naphthalene and anthracene. Inoculums were prepared in nutrient broth by
growing the cells at 30C overnight and used to inoculate the PNR and PNRG
medium (100 mL) containing PAHs with 10 % v/v. cultures were incubated at 30C at
150 rpm in shaking incubator. Effect of pH and temperature was determined on the
biodegradation experiment by conducting the transformation at different pH (4, 5, 6,
34

Chapter3MaterialsandMethods
7, 8 and 9) and temperature (30, 37, 45 and 50C) in PNRG medium. Samples were
collected from the flasks at time intervals, 0, 24, 48, 72, 96, 120, 144, 168 hrs and
monitored for growth and degradation by viable cell count and disappearance of
PAHs by HPLC analysis respectively.

3.3.2 Resting cell biotransformation of PAHs by Bacillus cereus


KWS2 and B. cereus KWS4
Resting cells biotransformation of PAHs (naphthalene, phenanthrene and anthracene)
by B. cereus KWS2 and B. cereus KWS4 was checked by growing the cells to 0.5
O.D (600 nm) in 1/4th conc. of LB with 10 ppm of selected PAH in 1000 mL
medium. The cells were harvested by centrifugation at 5000 rpm for 10 minutes and
washed twice with basal salt medium (BSM) and then used as an inoculum for 100
mL of BSM, with 50 ppm of PAHs (each in separate experiment) and incubated at
30C in a shaking incubator for 2 days. The PAHs were extracted from cultures and
analyzed by GC/MS.

3.3.3 Biotransformation of PAHs by Mycobacterium PY146


Seven different Mycobacterium strains which were previously isolated at the Center
for Environmental Biotechnology, University of Tennessee (Knoxville, TN, USA)
from PAHs contaminated sediment by enrichment culture technique with pyrene as
the sole source of carbon and energy (Debruyn, 2008) were screened for growth in
Basal Salt Medium (BSM) with 0.05 % yeast extract and 10 ppm pyrene.
From these seven isolates Mycobacterium PY146 was chosen for further study. The
optimal growth temperature for this strain was determined by growing it in LB broth
at different temperatures (25, 30, 37C) in a rotary shaker at 150 rpm. The growth was
monitored by measuring the absorbance at 600 nm using spectrophotometer. The
growth of Mycobacterium PY146 was also monitored in LB broth using varying pH
values of 5, 6, 7 and 8. Cultures were incubated at 30C on a rotary shaker at 150 rpm
and the growth was monitored spectrophotometrically at 600 nm.

35

Chapter3MaterialsandMethods
For resting cell biotransformation experiments, the Mycobacterium PY146 was grown
to 0.5 O.D in 1/4th conc. of LB with 10 ppm of selected PAHs in 1000 mL medium.
The cells were harvested by centrifugation and washed twice with BSM and then used
as an inoculum into 100 mL of BSM, with 50 ppm of individual PAHs and incubated
at 30C in a shaking incubator for 2 days. For growing cell biotransformation
experiments, the bacterial strains was grown overnight in 1/4th conc. of LB with 10
ppm of selected PAH in 500 mL flask. The cells were harvested by centrifugation and
washed twice with BSM. The cells were added to 100 mL BSM up to 0.05 O.D (600
nm), with 50 ppm of individual PAHs and incubated at 30C in a shaking incubator
for one week. The PAHs were extracted from these cultures and analyzed by GC/MS.

3.3.4 Biodegradation of PAHs in Soil


Biodegradation of phenanthrene, anthracene and pyrene using Mycobacterium PY146
was evaluated in soil. Clean uncontaminated soil (PCB Soil PR3-7) was collected,
dried, sieved and sterilized by autoclaving three times at temperature of 121C and
pressure of 1 atm. Clean soil (8 gram) and White quartz Sand (Sigma) (4 gram) were
mixed and placed in 40 ml EPA vials. Into triplicate vials for each sample, 120 g of
phenanthrene and 72 g of anthracene and pyrene each were added in and mixed
thoroughly. The samples were inoculated with 0.3 O.D (at 600 nm) of Mycobacterium
PY146 cells in six ml of BSM. The experiment was set up with and without 0.05 %
yeast extract. For each experiment, abiotic and biotic controls were prepared. The
abiotic control was same as the test samples, except without any added
Mycobacterium PY146 cells. The biotic control consisted of killed cells of
Mycobacterium PY146 with 1.0 mL of 2 N H2SO4. All vials were incubated at 30C
and at pH 7 in a horizontal shaker in dark. Whole sample contained in the individual
vials were extracted at 0 time, one week and two weeks.

3.3.5 Analytical methods


3.3.5.1 Viable Cell Counts
Viable cell counts were done to monitor bacterial growth by using serial dilutions and
spread plate technique.

36

Chapter3MaterialsandMethods
3.3.5.2 HPLC Analysis
Aliquots (1 mL) were centrifuged at 10000 rpm for 10 minutes and supernatant was
filtered by 0.2 m filter and kept for analysis. Analysis was performed by HPLC
(Waters) with a C-18 reverse phase column (thermo Hypersil 1504.6 mm 5
Hypersil). Anthracene was eluted in isocratic mode from column with acetonitrile and
deionized water (60:40) as eluent at a flow rate of 1.2 mL/min. The concentration of
naphthalene and anthracene was determined at 254 nm by comparison to standard
curve.
3.3.5.3 Extraction and Analysis of PAHs and their products
For extraction of PAHs and their metabolites, 100 mL of the cell suspension were
harvested. After centrifugation at 4C (5000 rpm, 15 min), the supernatant was
collected and extracted three times with equal volumes of ethyl acetate. All extracted
ethyl acetate were pooled together and generated neutral extract. The remaining
supernatant was further acidified to pH 2.5 with 1 M HCl to produce acidic extract in
the same manner. All of the collected ethyl acetate fractions were evaporated
separately at 40C using a rotary evaporator (Rotavapor RE121 Bchi, Switzwrland),
and the residue was dissolved in a 1 mL of acetone. The samples were dried in
vacuum and stored at 20C until used (Kim et al., 2005). Samples were analyzed by
GC/MS.
Phenanthrene, anthracene and pyrene were extracted from soil by adding an equal
volume of ethyl acetate and shaken for an hour on a horizontal shaker. This mixture
was allowed to settle for 10-15 minutes, and then 1 mL of organic phase was collected
from vial and stored at -20C till analysis. Analysis of biodegradation was performed
using GC/MS under SIM mode.
3.3.5.4 GC/MS Analysis
Samples were analyzed by GC/MS (Agilent 6890 Gas Chromatograph with 5973N
Mass Spectrometer with inert source) equipped with a DB5-MS column (30 m x 0.25
mm. i.d., 0.25 m film, (J&W Scientific) according to U.S. EPA Method 8270D.
The GC/MS operating conditions were as follows:
37

Chapter3MaterialsandMethods
Mass range:

35-500 amu

Scan time:

1 sec/scan

Initial temperature:

60C, hold for 4 min

Temperature program:

60-300C at 10C/min

Final temperature:

300C

Injector temperature:

250C

Transfer line temperature:

280C

Injector:

Splitless

Injection volume:

1 L

Carrier gas:

Helium

GC/MS were equipped with DB5-MS column (30 m x 0.25 mm i.d., 0.25 m film).

3.4 MINERALIZATION STUDIES OF PAHs


3.4.1 14C-PAH mineralization assay
Radiolabeled

14

C-PAHs mineralizations were carried out using naphthalene,

phenanthrene and anthracene for Bacillus cereus KWS2 and naphthalene,


phenanthrene, pyrene and benzo(a)pyrene for Mycobacterium PY146. The assay was
performed to determine the production of 14CO2 from 14C-PAHs. Cells were grown in
a shaking growth chamber (150 rpm) in of the LB medium to an O.D 600 of 0.8.
Cells were washed once with BSM and resuspended in BSM broth to an O.D 600 of
0.30. Five mL of the suspension was added to a 40-mL EPA vial. A CO2 trap (4-mL
vial) containing 1.0 mL of 0.5 M NaOH was placed inside the EPA vial. Before the
EPA vial was sealed with a Teflon-lined silicone septum, an appropriate amount of
14

C-PAH dissolved in acetone was added: 33,650 dpm benzo(a)pyrene, 60,550 dpm

pyrene, 78,210 dpm naphthalene, 81,100 dpm phenanthrene and 78,100, dpm
anthracene.
The assays were performed in triplicate with abiotic and biotic controls. Abiotic
controls consisted of BSM media amended with the same radioactive PAH as the test
samples and without any added bacterial cells. Biotic controls consisted of media
containing the radioactive PAH and cells killed with 1.0 mL of 2 N H2SO4. The vials
were incubated at 30C with shaking at 150 rpm for 7 days. The assays were stopped
by injecting 1.0 mL of 2 N H2SO4 through the septa. Then the vials were shaken for 1
38

Chapter3MaterialsandMethods
h to drive 14CO2 into the NaOH traps. The NaOH was mixed in 10 mL of ReadySafe
liquid scintillation cocktail (Beckman Coulter, Fullerton, CA, USA). Mass balances
were achieved by measuring radioactivity from

14

CO2, the liquid media, and the

biomass/residue portion of the mineralization assay. The

14

C-radioactivity was

measured with a TRI-CARB 2900TR liquid scintillation analyzer (PerkinElmer,


Shelton, CT 06484-4794 USA).

3.4.2 14C-Pyrene and 14C-phenanthrene mineralization using different


cells densities
The mineralization of

14

C-pyrene checked with different incubation times starting

with 0.25 O.D of Mycobacterium PY146. The mineralization of

14

C-pyrene and

phenanthrene was monitored using Mycobacterium PY146 cultures ranging in O.D at


600 nm from 0.1 to 0.5. In initial, culture were grown by inoculating the cells in of
the LB medium with 10 ppm of pyrene while shaking (150 rpm) at 30C to an O.D of
0.5 at 600 nm. That culture was transferred to different flasks containing of the LB
medium with 10 ppm of pyrene and phenanthrene separately. The individual flasks
were grown to an O.D of 0.1, 0.2, 0.3, 0.4 and 0.5 at 600 nm. The cells were harvested
with centrifugation and resuspended in the same volume of BSM broth with 0.05 %
yeast extract. Five mL of the suspension was added to a 40-mL EPA vial. A CO2 trap
(4-ml vial) containing 1.0 mL of 0.5 M NaOH was placed inside the EPA vial. Before
the EPA vial was sealed with a Teflon-lined silicone septum, an appropriate amount
of

14

C- 4,5,9,10-pyrene (128188 dpm) and

14

C-phenanthrene (70000 dpm) dissolved

in acetone was added.


The assays were performed in triplicate with abiotic and biotic controls. Abiotic
controls consisted of BSM media amended with the same radioactive PAH as the test
samples and without any added bacterial cells. Biotic controls consisted of cells killed
with 1.0 mL of 2 N H2SO4. All vials were incubated at 30oC with shaking at 150 rpm
for 20 hours. The assays were stopped by injecting 1.0 mL of 2 N H2SO4 through the
septa. Then vials were shaken for 1 h to drive

14

CO2 into the NaOH traps. The

14

C-

radioactivity was measured with a TRI-CARB 2900TR liquid scintillation analyzer.

39

Chapter3MaterialsandMethods

3.5 EXPRESSION OF PYRENE DIOXYGENASE GENE (nidAB)


FROM MYCOBACTERIUM PY146 into E. coli
3.5.1 Isolation of genomic DNA
The bacterium was grown in LB broth at 30C for 72 hours. Cells were pelleted by
centrifugation for 10 minutes at 1000 rpm. Genomic DNA was isolated according to
the protocol of FastDNA Spin Kit for soil (MP Biomedicals, Solon, OH, USA). The
isolated Genomic DNA was run on 1 % agarose gel to verify size.

3.5.2 Amplification of nidAB


The pyrene dioxygenase gene (nidAB) was amplified using a forward (5ATGAACGCGGTTGCGGTCGAT-3)

and

reverse

primers

(5-

TCAAGCACGCCCGCCGAATGC-3) and a 25 L PCR reactions consisting of:


Genomic DNA

1 L of template

dNTPs

1 L

Primers (Forward and reverse)

1 L (each)

Buffer A (10X)

2.5 L

Taq polymerase

1 L

dH2O

17.5 mL

The thermocycler program consisted of 94C for 5 minutes followed by 32 cycles of:
94C for 1.0 minute, 60C for 1.0 minute and 72C for 1 minute 30 seconds. After
PCR completion the amplified products were stored at 4C. To check the amplified
fragment size, PCR product and 1kb ladder was run on 1 % agarose gel in 1x TAE
buffer at 90 V.
Composition of 50x TAE buffer:
Tris

242 g/L

NaEDTA.2H2O

18.61 g/L

Glacial acetic acid

57 mL

40

Chapter3MaterialsandMethods

3.5.3 Cloning of nidAB gene in E.coli


3.5.3.1 Cloning of nidAB into pCR 2.1 TOPO vector
A polyA tail was added to both ends of the nidAB gene fragment obtained by PCR
using polyA beads and incubating the mixture at 72C for 30 minutes. The polyA
prepared PCR product (4 L) was then combined with 1 L of salt solution (1.2 M
NaCl, 0.06 M MgCl2) and 1 L of TOPO vector (having ampicillin resistant gene as
selectable marker) (Invitrogen, Carlsbad, CA, USA) (Figure 3.1) and incubated for 30
minutes at room temperature.
3.5.3.2 Transformation of E. coli with recombinant pCR 2.1 TOPO vector
For transformation 6 L of the recombinant nidAB- pCR 2.1 TOPO vector was added
to a vial of competent TOPO10 E. coli cells and kept on ice for 30 minutes. The
reaction vial was heat shocked at 42C for 30 seconds and placed back on ice for 2
minutes. One ml of nutrient rich YT 2x medium (yeast extract + tryptone) was added
to the cells and the cells were incubated at 37C for 1 hour in a shaking water bath.
Aliquots consisting of 50, 100 and 200 L of the transformed cells were spread on LB
agar plates (having ampicillin (100 g/mL) and incubated overnight at 37C. Only the
transformed cells containing the plasmid with ampicillin resistance were able to grow
on the plates.
3.5.3.3 Isolation of recombinant pCR 2.1 TOPO vector from transformed E. coli
cells
The LB broth (0.5 mL) (having ampicillin (100 g/mL)) was inoculated with
transformed E. coli from the agar plates. Samples were centrifuged at 10000 rpm for
10 minutes. The pellet was saved and the plasmid was isolated from the cell pellet
according to the protocol of Promega Wizard miniprep DNA purification System Kit
(Promega, Madison, WI, USA).
3.5.3.4 Restriction enzyme analysis
The isolated plasmid DNA was analyzed by restriction enzyme analysis. One L of
EcoR1, 2 L of Buffer H and 17 L of plasmid DNA were mixed together and
41

Chapter3MaterialsandMethods
incubated for 1 hour at 37C. The reaction products were checked by ruining on a 1 %
agarose gel.
3.5.3.5 Addition of restriction sites on the ends of nidAB gene
New

primer

sequences

were

CTGCAGAATTCGCCCTTATGAACG-3
AAGCTTTCAAGCACGCCCGCC-3

designed,
and

nidABEco
nidABHind

For

5-

Rev

5-

so that after PCR amplification the nidAB

gene would have EcoR1 and HindIII restriction sites on the both ends of the gene.
The PCR reaction consisted of:
Plasmid DNA

1 L

Primer (Forward)

1 L

Primer (Reverse)

1 L

dNTPs

1 L

Buffer A

(10x)

2.5 L

Tag polymerase

1 L

dH2O

17.5 L

The thermocycler program consisted of 94C for 5 minutes followed by 32 cycles of:
94C for 1.0 minute, 61C for 1.0 minute and 72C for 1 minute 30 seconds. The
amplified PCR product was kept at 4C until the PCR tube was removed from
thermocycler. To check the amplified fragment size, the PCR product was run on a 1
% agarose gel. The nidAB gene with the attached restriction sites (EcoR1 and
HindIII) was cloned into TOPO10 vector and transferred to competent E.coli cell. The
cells were grown and the recombinant plasmid was isolated from the cells according
to the protocol of Promega Wizard miniprep DNA purification System Kit. The
ligated restriction sites (EcoR1 and HindIII) at the ends of nidAB gene in this plasmid
were confirmed by sequencing of recombinant TOPO vector with M13 primers.
3.5.3.6 Cloning of nidAB gene in Expression Vector pK223-3
To clone the nidAB gene having EcoR1 and HindIII restriction sites into pkk223-3
expression vector (Amersham) (Figure 3.2), both the recombinant TOPO10 vector

42

Chapter3MaterialsandMethods
and pkk-233-3 expression vector were digested separately with the EcoR1 and
HindIII. The digestion reaction consisted of:
Plasmid DNA

20 L

EcoR1

2 L

HindIII

2 L

Buffer E

5 L

dH2O

21 L

Both the digestion reactions were incubated for 2 hours at 37C, then I L of each
enzyme was added to the reaction and incubated in the same condition for additional
one hour.
3.5.3.6.1 Dephosphorylation Reaction
After digestion with restriction enzymes (EcoR1 and HindIII), plasmid pkk233-3 was
dephosphorylated by adding 1 L of TSAP (Thermosensitive Alkaline Phosphatase)
(Promega USA) and incubated at 37C for 15 minutes. The restriction enzyme and
TSAP were inactivated by incubating at 74C for 15 minutes. Both the digestion
reactions were run on a 1.5 % agarose gel without ethidium bromide and DNA
fragments were excised from the gel.
3.5.3.6.2 Gel Purification of fragments
Bands of both fragments (EconidABHind and pkk233-3) were excised with sharp
blade and DNA was extracted and purified using Gene Clean Spin Kit (Q BioGene)
according to manufacturer's instructions.
3.5.3.6.3 Ligation of EcoRInidAB HindIII and pkk223-3 fragment
After digestion and purification the EconidABHind fragment was ligated with
pkk223-3 using DNA ligase. The reaction mixture was incubated at 16C overnight.
The ligation reaction consisted of:

43

Chapter3MaterialsandMethods
Plasmid vector (pk223-3)

1.5 L

EconidABHind fragment

5.75 L

Ligase buffer (10x)

2 L

Ligase

1 L

dH2O

9.75 L

3.5.3.6.4 Transformation of E. coli with recombinant pkk223-3


A -80C frozen culture of Escherichia coli TOPO10 was used for transformation
using the standard protocol. Cells were thawed on ice and 3 L of ligation reaction
was added to the cells and mixed by tapping. The vial was incubated on ice for 30
minutes. Then it was incubated for 30 seconds at 42C in water bath, and quickly
removed and placed on ice. Then 1 mL of YT medium was added in the vial, and
incubated in shaking water bath at 37C for 1 hour. Aliquots of 50, 100 and 200 L
were plated onto LB agar containing 100 g/mL ampicillin. The plates were inverted
and incubated at 37C overnight. Individual colonies were picked and streaked on
fresh LB agar plates containing 100 g/mL ampicillin and incubated overnight at
37C.
3.5.3.6.5 Isolation of a recombinant pkk223-3 plasmid containing EcoRInidAB
HindIII
A loop-full of each ampicillin resistant culture was taken from the agar plates using a
sterilized loop and mixed in an eppendorf tube with 0.5 mL of LB broth, having 100
g/mL ampicillin. The tubes were centrifuged at 10000 rpm for 10 minutes to pellet
the cells. The plasmid DNA was isolated from the pellet according to the protocol of
Promega Wizard miniprep DNA purification System Kit. The concentrations (ng/L)
of plasmid DNA having the EcoRINidABHindIII fragments were measured by Hoefer
DyNA Quant 200 Fluorometer (GE Healthcare Biosciences, Piscataway, NJ. USA).
3.5.3.6.6 Restriction enzyme analysis of recombinant pkk223-3 vector
The isolated plasmid DNA was analyzed by restriction enzyme analysis. The
restriction enzyme reactions consisted of:

44

Chapter3MaterialsandMethods
Plasmid DNA

15 L

EcoR1

1 L

HindIII

1 L

Buffer H

2 L

dH2O

1 L

The reaction mixture was incubated at 37C for 1 hour. Restriction enzyme products
were separated by gel electrophoresis to verify size on a 1 % agarose gel made with
1x TBE.
3.5.3.6.7 Sequencing of the recombinant pk223-3 plasmid contained EcoRInidAB
HindIII
Three plasmid DNAs (clones 6, 7 and 8) were sequenced using the M13 primers on an
ABI 3730 DNA Analyzer at the Molecular Biology Resource Facility at the
University of Tennessee, Knoxville, TN, USA.
3.5.3.6.8 Direct Colony PCR to verify fragment insert size
Direct colony PCR was done using a slight modification of the procedure from PerezPerez and Hanson 2002, to confirm that the right fragment size was cloned into the
pkk223-3 expression vector.

Briefly, the colony was touched with a disposable

pipette tip and immersed into a PCR tube containing 1 L of each primer (nidAB
Forward and Reverse), 1 L dNTPs, 2.5 L of Buffer A (10x), 1 L of Tag
polymerase and 17.5 L of dH2O. The PCR products and a 1kb ladder were run on 1
% agarose gel in 1x TBE.
3.5.4 Biotransformation of PAHs by E. coli with cloned nidAB into pkk223-3
E. coli having cloned NidAB fragment in plasmid pkk223-3 were grown in LB with
IPTG overnight at 37oC. The cells were washed with BSM twice.

Resting cell

assays using naphthalene, phenanthrene, anthracene and pyrene added individually to


the vials were performed as described in section 3.3.2. The samples were extracted
with an equal volume of Ethyl acetate three times and then analyzed on GC/MS as
described in section 3.3.5.3.

45

Chapter3MaterialsandMethods
In order to detect PAH metabolites 1 mL of the extracted samples were derivatized
with N,O-Bis(trimethylsilyl)trifluoroacetamide with 1 % Trimethylechlorosilane
(BSTFA) (Fluka, Switzerland) by adding 50 L of BSTFA to the samples and
incubated overnight at 50C as recommended by manufacturer. The samples were
analyzed on GC/MS as described above in section 3.3.5.3.

3.5.5 Invitro RNA synthesis of nidAB transcripts


3.5.5.1 Linearizing the nidAB DNA template
The nidAB fragment in the TOPO vector was linearized by digestion with BamH1
restriction enzyme. The digestion reaction consisted of:
Plasmid DNA

40 L

Buffer E

5 L

BamH1

2 L

dH2O

3 L

The digestion reaction was mixed and incubated for 3 hours at 37C, and then another
1 L of the BamH1 enzyme was added and the reaction was incubated again at the
same conditions.
3.5.5.2 Synthesis of large quantities of RNA
A T7 reaction were set up, and the reaction consisted of:
Linear DNA template

8 L

RiboMAXTM Express T7 2x Buffer

10 L

Enzyme Mix, T7 Express

2 L

The reaction was mixed gently and incubated at 37C for 30 minutes. After the
transcription reaction was complete, 2 L of Promega RQ1 RNase-free DNase (cat. #
M6101) was added to the reaction and the reaction was incubated for 1 hour at 37C,
to remove the DNA template. For additional clean-up of the in vitro transcription the
protocol of T7 RiboMAX Express Large Scale RNA Production System (Promega,
Madison, WI, USA) was followed. The RNA concentration was measured on
Nanodrop spectrophotometer (Thermo Fisher Scientific, Wilmington, DE, USA).
46

Chapter3MaterialsandMethods
3.5.5.3 Isolation of RNA transcript of nidAB
The TOPO10 E. coli cells (clone numbers 6, 7 & 8) containing the nidAB gene
fragment in the expression vector pkk233-3 which were grown in the LB broth
medium containing 100 g/mL of ampicillin with and without IPTG (0.8 mM) at
37C overnight. The RNA was isolated using FastRNA Pro Soil-Direct Kit
following the protocol.

3.5.6 Quantification of DNA and RNA copy number of nidAB


3.5.6.1 Isolation of DNA and RNA from Mycobacterium PY146
Total DNA and RNA was isolated and purified from Mycobacterium PY146 grown in
LB to different cell concentration ranging from 0.1, 0.2, 0.3, 0.4 and 0.5 O.D at 600
nm. The RNA was isolated using FastRNA Pro soil-Direct Kit, while DNA isolation
was done using FastDNA SPIN Kit for Soil (MP Biomedicals, Solon, OH, USA). A
primer and TaqMan probe set was designed for quantifying dioxygenase genes by
targeting the large () dioxygenase subunit (nidA) and was based on a multiple
sequence alignment of the nidA conserved regions from several PAH-degrading
Mycobacterium.

The

forward

and

reverse

primer

sequences

were

5-

TTCCCGAGTACGAGGGATAC-3, and 5-TCACGTTGATGAACGACAAA-3,


respectively.

The

TaqMan

probe

sequence

was

5d

FAM-

TCCTACCCGTCGCCGGTACA-BHQ-1 3.
The real-time PCR reactions consisted of:
2x Brilliant II SYBR QRT-PCR master mix

12.5 L

Forward primer

0.75 L

Reverse primer

0.75 L

TaqMan probe

0.5 L

dH2O

10 L

DNA or RNA Template

2.5 L

47

Chapter3MaterialsandMethods
The quantitative PCR reactions were performed on an MJ Opticon thermocycler (BioRad, Hercules, CA, USA) using the following protocol: 50C for 2 minutes, 95C for
15 minutes, then 40 cycles of denaturing at 94C for 15 seconds and annealing at
54C for 30 seconds. nidAB from Mycobacterium PY146 cloned into TOPO
(Invitrogen) vector was used to create an eight-point standard curve ranging from 101
to 108 gene copies per reaction.

48

Chapter3MaterialsandMethods

Figure 3.1: Map of plasmid vector pCR 2.1-TOPO used for nidAB cloning

49

Chapter3MaterialsandMethods

Figure 3.2: Map of vector pKK223-3 used for nidAB expression

50

Chapter4Results

4.1 SELECTION OF PAH TRANSFORMING BACTERIA


A total of 52 bacterial isolates were screened for their growth on PAHs from oil
contaminated sites. Fifteen strains showed the best growth on nutrient agar plates with
anthracene (800 ppm). These strains when tested for their growth in mineral salt
medium (PNR) with anthracene and pyrene as sole source of carbon and energy,
different levels of growth were observed. Majority of the isolates showed growth up
to 400 ppm concentration of the PAHs tested. Strain KWS2 showed rich growth up to
1200 ppm anthracene while other strains showing growth up to 1200 ppm include
KWS4, DW3, Sol-10 and 10-1 (Table 4.1). Growth tolerance in mineral salt medium
with pyrene was up to 800 ppm for the same five strains (Table 4.2). So these five
strains were selected as PAH degrading bacteria for further study.

4.1.1 Identification of selected bacterial isolates


4.1.1.1 Morphological and biochemical characterization
The five bacterial isolates KWS2, KWS4, DW3, Sol-10 and 10-1 were analyzed
taxonomically. The colony morphology of KWS2 and KWS4 were large, offwhite/creamy in color with irregular margins while DW3 and Sol-10 appeared
opaque. The morphology of 10-1 appeared as small to medium in their size, yellowish
in color with regular margins. All the isolates were gram positive except 10-1 which
was gram negative, and all were rod shape and motile. Spore staining revealed the
presence of spores in KWS2, KWS4, DW3 and Sol-10 (Table 4.3).
Biochemical tests revealed that all the isolates showed gelatin liquification and
catalase activity, and the starch hydrolysis, casein hydrolysis and nitrate reduction
were positive for all the isolates except 10-1. Indole production was negative for all
the isolates, while the methyl red and urease test were negative for all isolates except
KWS2. Simon citrate test was negative for all the isolates except 10-1. All the isolates
showed oxidase activity except Sol-10. Voges proskauer reaction was positive for
KWS2 and KWS4 while the other isolates showed negative reaction (Table 4.4).
Sugar fermentation tests showed that all the isolates were unable to ferment lactose
and fructose. Isolate KWS2 showed acid production from arabinose, sucrose, maltose,
mannose, glucose and inositol, KWS4 produced acid from maltose, mannose and
51

Chapter4Results

glucose, DW3 and Sol-10 were positive for acid production from arabinose, maltose
and mannose, 10-1 produces acid from arabinose, xylose, rhamminose, sucrose,
maltose and mannose fermentation. There was no gas production from any of the
above mentioned sugars by any isolate (Table 4.5).
4.1.1.2 Molecular characterization based on 16S rRNA gene
The five selected bacterial isolates were further identified on the basis of sequence
analysis of 16S rRNA. Based on the phylogenetic tree (Appendix I a) and the data in
tabular format which showed the percent identity (Appendix I b), isolates KWS2 and
KWS4 had more than 99 % (99.4 and 99.3 % respectively) homology in 16S rRNA
sequence to Bacillus cereus DQ 884352. Isolates DW3 and Sol-10 had 99.3 and 99.7
% homology to Bacillus licheniformis EU 373408, while isolate 10-1 had 99.3 %
homology to Pseudomonas stutzeri U 26262 (Table 4.6).

52

Chapter4Results

Table 4.1: Growth of bacterial isolates on PNR agar plates containing anthracene
Concentration of Anthracene (ppm)

Bacterial
Isolates

100

120

150

200

250

300

400

500

600

800

1000 1200

SS13

++

++

++

++

++

++

++

SS18

++

++

++

++

++

KWS3

++

++

++

++

++

++

++

++

++

KWS2

+++

+++

+++

+++

+++

++

++

++

+++ +++ ++

DW3

+++

++

++

++

++

++

++

+++ ++

++

++

EC1

+++

+++

+++

+++

+++

++

++

++

KWS 1S

+++

+++

++

++

++

++

++

++

KWS4

+++

+++

+++

+++

+++

+++

+++

+++ +++ +++ +++

++

KWS 5S

++

++

KWS 6

+++

+++

+++

+++

++

++

RED 18

+++

+++

+++

+++

++

++

Sol-10

+++

+++

+++

+++

+++

+++

+++

+++ +++ ++

++

++

10-1

++

++

+++

+++

++

++

++

++

++

++

KWS 5L

PC 1

++

++

++

++

++

++

++

PA 4

++

++

++

++

++

++

++

Slight Growth (+)

Good Growth (++)

Rich Growth (+++)

NO Growth (-)

53

+++

Chapter4Results

Table 4.2: Growth of bacterial isolates on PNR agar plates containing Pyrene
Bacterial

Concentration of Pyrene (ppm)

Isolates

20

50

75

100

200

300

400

600

700

800

SS13

++

++

++

SS18

++

++

++

++

++

KWS3

++

++

++

++

++

++

++

KWS2

+++

+++

+++

+++

+++

+++

+++

+++ ++

++

DW3

++

++

+++

++

++

++

++

++

++

EC1

+++

++

+++

+++

++

++

++

++

KWS1 S

+++

+++

++

++

++

++

++

KWS4

++

+++

+++

+++

+++

+++

+++

++

++

KWS5 l

++

++

KWS 6

+++

++

+++

++

++

++

RED 18

+++

+++

+++

+++

+++

Sol-10

+++

++

++

+++

++

++

++

++

10-1

++

+++

++

++

++

++

++

++

PA 4

++

++

++

++

++

++

++

KWS1 s

++

++

+++

++

KWS1 l

++

+++

++

++

++

KWS3 s

+++

KWS5 l

++

++

++

++

++

54

Chapter4Results

Table 4.3: Morphological characteristics of bacterial isolates


Strain

Gram

Shape

stain

Spore

Motility

stain

KWS2

Rod

KWS4

Rod

DW3

Rod

Sol-10

Rod

10-1

Rod

55

Chapter4Results

Table 4.4: Biochemical characterization of selected bacterial isolates

Isolates

Starch
Hydrolysis

KWS2

KWS4

DW3

Sol-10

10-1

Casein
hydrolysis

Simon citrate
test

Gelatin
Liquefaction

Catalase test

Indole
Production

Methyl Red

Voges Proskauer

Nitrate
Reduction

Urease

Oxidase

56

Chapter4Results

Table 4.5: Sugar fermentation tests of the bacterial isolates

Sugars Lactose Arabinose Fructose Xylose Rhamminose Sucrose Maltose Mannose Glucose Inositol
A

KWS2

KWS4

DW3

Sol-10

10-1

Strains

A = Acid Production
G = Gas Production

57

Chapter4Results

Table 4.6: Identification of the bacterial isolates on the basis of 16S rRNA
Isolates

%age homology

Accession No. of
sequence giving
homology

Identification

KWS2

99.4 %

DQ 884352

Bacillus cereus

KWS4

99.3 %

DQ 884352

Bacillus cereus

DW3

99.3 %

EU 373408

Bacillus licheniformis

Sol-10

99.7 %

EU 373408

Bacillus licheniformis

10-1

99.3 %

U 26262

Pseudomonas stutzeri

58

Chapter4Results

4.2 BIODEGRADATION STUDIES OF PAHS


4.2.1. Biotransformation of anthracene and naphthalene using B.
cereus KWS2 and P. stutzeri 10-1
Biotransformation studies of anthracene were carried out in growing cells shake flask
experiments in PNRG medium containing 10 ppm anthracene. The experiments were
designed to evaluate the effects of incubation time, pH and temperature on the growth
of Bacillus cereus KWS2 and biotransformation of anthracene.
Biotransformation of anthracene by Bacillus cereus KWS2 at different pH values
revealed that, the highest degradation was observed at pH 7 which was almost 40 %
after four days and 45 % after seven days of incubation at 30C. A 30 % degradation
was observed at pH 6 on day 7 while there was a significant decrease in the
concentration of anthracene at pH values 4, 5, 8 and 9 (Figure 4.1).
The growth study showed that there was progressive increase in growth in terms of
CFU/mL from the initial inoculation. The maximum growth was observed at pH 7
after two days of incubation (12 x107) further there was decline in growth as the
incubation time increased. The second highest growth was observed at pH 6 which
was 10.8 x107 CFU/mL after two days of incubation. There was a decrease in growth
below pH 6 and as well as above pH 7 (Figure 4.2).
The optimum temperature for the degradation of anthracene was observed to be 30C
where 45 % reduction in the anthracene concentration was observed after seven days
of incubation. As the temperature increased the percent degradation decreased. There
was 32, 27 and 20 % degradation at 37, 45 and 50C, respectively, after seven days of
incubation (Figure 4.3).
The effect of temperature on growth of Bacillus cereus KWS2 correlated with the
percentage degradation of anthracene and the highest levels of growth was observed
at 30C which was 12 x107 CFU/mL after two days of incubation. The second highest
growth was observed at 37C which was 7 x107. The growth was decreased
significantly with the increase in temperature (Figure 4.4).

59

Chapter4Results

Bacterial isolate Pseudomonas stutzeri 10-1 was also checked for growing cell
biotransformation of naphthalene and anthracene in PNRG medium separately. In
case of naphthalene, 32.81 % reduction was observed after one week of incubation
while 33.69 and 45.35 % reduction was observed after second and third week
respectively (Figure 4.5). In case of anthracene degradation, 87.37 % reduction was
observed after first week of incubation whereas 92.98 and 95. 56 % reduction was
observed in anthracene concentration after two and three weeks, respectively (Figure
4.6).

4.2.2 Resting cell biotransformation of PAHs by Bacillus cereus


KWS2 and B. cereus KWS4
Biotransformation experiment of PAHs (naphthalene, phenanthrene and anthracene)
was carried out by Bacillus cereus strains KWS2 and KWS4 in BSM medium
containing 50 ppm of each compound separately at 30C and pH 7.
Resting cell biotransformation of naphthalene by Bacillus cereus KWS2 was
evaluated after two days of incubation. Four metabolites were detected and identified
in GC-MS chromatograms from the ethyl acetate neutral extracts from naphthalene
grown culture after two days of incubation which includes; 2-naphthol {MS
fragments: 144(100), 145(12), 116(48),115(95), 89(14), 72(5), 63(10), 50(4), 39(4)},
benzeneacetic acid {MS fragments: 135.9(34), 92(19), 91(100), 65(14), 63(7), 51(5),
39(6)}, benzenecarboxylic acid {MS fragments: 121.9(96), 105(100), 76.9(81),
51(32), 50(25), 43(30)} and benzaldehyde {MS fragments: 106(100), 105(96), 78(20),
77(93), 74(13), 52(9), 51(42), 50(24), 39(6)} (Table 4.7; Appendix II a-d). All
metabolites were identified by comparing to the mass spectra of standards in NIST
mass spectral library (2005).
Biotransformation of phenanthrene was also evaluated after two days of incubations
by Bacillus cereus strains KWS2. A total of four metabolites were produced by
Bacillus cereus KWS2 which were detected and identified by GC/MS, three from the
neutral extracts of ethyl acetate including; 9-phenanthrenol {MS fragment: 194(100),
166(33), 165(90), 164(15), 163(16), 139(9), 82(11)}, 9,10-Dihydro, 9,10dihydroxyphenanthrene {MS fragment: 212(58), 193.9(34), 181(67), 166(51),
165(100), 153(13), 152(24), 138.9(4), 115(6), 82(9), 76.9(10)} where as bezeneacetic
60

Chapter4Results

acid, and 4-hydroxy-benzeneacetic acid, {MS fragment: 151.9(33), 108(8), 107(100),


77(23), 51(7), 39(4)} were detected in the acidic extract (Table 4.7; Appendix II e-h).
Degradation of anthracene by resting cells of B. cereus KWS2 after two days of
incubation gave three metabolites. These metabolites were detected in acidic extracts
which includes benzenecarboxylic acid, benzeneacetic acid and 4-hydroxybenzeneacetic acid (Table 4.7; Appendix II i-k).
Naphthalene biotransformation by Bacillus cereus KWS4 resulted the formation of
three metabolites which were detected and identified from the neutral extracts by
GC/MS. The detected metabolites were 2-naphthelenol, benzeneacetic acid and
benzaldehyde (Table 4.8; Appendix III a-c).
From phenanthrene biotransformation by Bacillus cereus KWS4 a total of five
metabolites were detected and identified, three from neutral extracts including; 9phenanthrenol, 9,10-Dihydro, 9,10-dihydroxyphenanthrene and bezeneacetic acid.
While metabolites form acidic extracts were detected and identified as 3-hydroxybenzeneacetic acid, {MS fragment: 152(34), 108(11), 107(100), 86.9(10), 77(22),
63(15), 51(11), 39(6)} and 2-hydroxy-benzoic acid {MS fragment: 137.9(56),
119.9(99), 92(100), 64(29), 63(22), 53(11), 43(23), 39(16)} (Table 4.8; Appendix III
d-h).

61

Chapter4Results

Reduction (%)

4 Days

7 Days

50
45
40
35
30
25
20
15
10
5
0
4

pH

Figure 4.1: Percentage reduction of anthracene concentration (10 ppm) in PNRG


medium at different pH values by Bacillus cereus KWS2 in shake flask experiment.

Figure 4.2: Growth (CFU/mL) of Bacillus cereus KWS2 in PNRG medium with 10
ppm anthracene at different pH values.

62

Chapter4Results

Reduction (%)

4 Days

7 Days

50
45
40
35
30
25
20
15
10
5
0
30

37
45
Temperature

50

Figure 4.3: Percentage reduction of anthracene concentration (10 ppm) in PNRG


medium at different temperature by Bacillus cereus KWS2 in shake flask experiment.

Figure 4.4: Growth (CFU/mL) of Bacillus cereus KWS2 in PNRG medium with 10
ppm anthracene at different temperature.

63

Chapter4Results

Figure 4.5: Percentage reduction of naphthalene concentration by Pseudomonas


stutzeri 10-1.

Figure 4.6: Percentage reduction of anthracene concentration by Pseudomonas


stutzeri 10-1.

64

Chapter4Results

Table 4.7: GC mass spectral data of metabolites from naphthalene, phenanthrene and
anthracene degradation by B. cereus KWS2

Anthracene

Phenanthrene

Naphthalene

Compound Metabolites

Retention
time (min)

m/z of fragment ions (%)


relative abundance

Benzaldehyde

10.24

106(100), 105(96), 78(20),


77(93), 74(13), 52(9),
51(42), 50(24), 39(6)

Benzenecarboxylic acid

13.69

121.9(96), 105(100),
76.9(81), 51(32), 50(25),
43(30)

Benzeneacetic acid

15.02

135.9(31), 92(19), 91(100),


89(5), 65(16), 63(8), 39(7)

2-naphthol

18.75

144(100), 145(12),
116(48),115(95), 89(14),
72(5), 63(10), 50(4), 39(4)

Bezeneacetic acid

15.08

135.9(31), 92(19), 91(100),


89(5), 65(16), 63(8), 39(7)

4-hydroxy-benzeneacetic
acid,

19.35

151.9(33), 108(8), 107(100),


77(23), 51(7), 39(4)

9,10-Dihydro, 9,10dihydroxyphenanthrene

25.18

212(58), 193.9(34), 181(67),


166(51), 165(100), 153(13),
152(24), 138.9(4), 115(6),
82(9), 76.9(10)

9-phenanthrenol

25.48

194(100), 166(33), 165(90),


164(15), 163(16), 139(9),
82(11)

4-hydroxy-benzeneacetic
acid

19.35

152(34), 107(100), 108(8),


77(20), 51(6), 39(3)

Benzeneacetic acid

15.25

136(36), 105(100), 101(24),


92(22), 75(9), 65(16), 39(7)

Benzenecarboxylic acid

14.04

122(85), 105(100), 77(62),


74(10), 50(16), 39(6)

65

Chapter4Results

Table 4.8: GC mass spectral data of metabolites from naphthalene and phenanthrene
degradation by B. cereus KWS4

Phenanthrene

Naphthalene

Compound

Metabolites

Retention
time (min)

m/z of fragment ions


(%) relative abundance

Benzaldehyde

10.23

106(100), 105(98),
78(22), 77(93), 74(16),
51(38), 50(27)

Benzeneacetic acid

15.02

135.9(34.5), 92(18),
91(100), 65(16.5), 51(5),
39(7.5)

2-naphthol

18.73

144(100), 116(45),
115(98), 89(12.5),
63(10), 50(3.5), 39(4)

Bezeneacetic acid

15.12

135.9(35), 92(19),
91(100), 65(16), 51(5),
39(8)

2-hydroxy-benzoic acid

15.94

137.9(56), 119.9(99),
92(100), 64(29), 63(22),
53(11), 43(23), 39(16)

3-hydroxy-benzeneacetic
acid,

19.27

152(34), 108(11),
107(100), 86.9(10),
77(22), 63(15), 51(11),
39(6)

9,10-Dihydro, 9,10dihydroxyphenanthrene

25.19

212(55), 194(34),
181(65), 166(47.5),
165(100), 15.9(26),
115(7.5), 82(8), 77(9)

9-phenanthrenol

25.48

194(100), 166(33),
165(99), 164(14),
163(15.5), 139(9), 82(11)

66

Chapter4Results

4.2.3 Biotransformation of PAHs by Mycobacterium PY146


Seven different Mycobacterium strains were screened for their growth in BSM
medium supplemented with 0.05% of yeast extract and 10 ppm of pyrene. The result
showed that the Mycobacterium PY146 had a maximum growth after 48 hours with
an O.D600 of 0.65 compared to other strains with O.D600 around 0.5 (Figure 4.7). The
Mycobacterium PY146 strain was selected for further biotransformation studies.
The effect of temperature on the growth of Mycobacterium PY146 was evaluated in
the LB broth. The results showed that the maximum growth was observed at 30C
with O.D600 of 1.12 after 120 hours, and growth decreased at both 25 and 37C
(Figure 4.8).
The effect of pH on growth of Mycobacterium PY146 in LB broth medium was also
evaluated. The maximum growth was observed at pH 7 with an O.D600 of 1.15 after
120 hours. Good growth was also observed at pH 6 with a maximum OD600 around
1.12. There was less growth at pH 5 and 8 (OD600s of 0.4 and 0.6), respectively
(Figure 4.9).
Resting cell biotransformation of PAHs including phenanthrene, anthracene and
pyrene by Mycobacterium PY146 were checked individually. Phenanthrene
biotransformation after 24 hours of incubation resulted in the detection of one
metabolite identified as 1,2-benzenedicarboxylic acid {Mass fragment: 148(23),
105(8), 104(100), 76(68), 74(15), 50(25), 38(6)} (Table 4.9; Appendix IV a) in the
acidic extract of ethyl acetate.
Two metabolites, 1,2-Benzenedicarboxylic acid from acidic extract and 9,10anthracendione {Mass fragment: 209(16), 208(100), 207(15), 181(12), 180(81),
152(61), 151(32), 150(14), 126(7), 76(20), 50(6)} (Table 4.9; Appendix IV b & c)
from neutral extract of ethyl acetate were detected and identified from the anthracene
biotransformation after 24 hours of incubation.
From pyrene biotransformation by Mycobacterium PY146, two metabolites, 1,2Benzenedicarboxylic acid from the acidic and 1-hydroxypyrene {Mass fragment:
219(19), 218(100), 217(12), 190(9), 189(41), 188(4), 187(6), 163(2), 94(3), 43(10)}
67

Chapter4Results

(Table 4.9; Appendix IV d & e) from the neutral extract of the ethyl acetate were
detected and identified.
Growing cell biotransformation of phenanthrene, anthracene and pyrene were also
checked individually by Mycobacterium PY146 in BSM medium with 0.05% yeast
extract in a shaking incubator at 30C.
From the growing cell biotransformation of phenanthrene by Mycobacterium PY146,
1,2-Benzenedicarboxylic acid was detected from the acidic extract of ethyl acetate
from both after 5 and 10 days of incubation, while benzene acetic acid (Table 4.10;
Appendix V a & b) was detected and identified in the acidic extract of ethyl acetate
after 10 days of incubation.
Four metabolites were detected from anthracene by Mycobacterium PY146 growing
cells biotransformation, in which 9,10-anthracenedione from the neutral extract and
benzenecarboxylic acid, benzeneacetic acid and 1,2-benzenedicarboxylic acid ( Table
4.10; Appendix V c - f) from the neutral extract of ethyl acetate after 10 days of
incubation.
Biotransformation of pyrene by growing cells of Mycobacterium PY146 resulted in
three metabolites including benzenecarboxylic acid, benzeneacetic acid and 1,2benzenedicarboxylic acid ( Table 4.10; Appendix V g-i) were detected and identified
from the acidic extract of the ethyl acetate after 7 days of incubation. After 14 days of
incubation one metabolite 1,2-benzenedicarboxylic acid was detected and identified
from the acidic extract of ethyl acetate.

68

Chapter4Results

Figure 4.7: Growth of different Mycobacterium strains on 10 ppm pyrene in BSM


medium.

69

Chapter4Results

Figure 4.8: Effect of different temperature on the growth of Mycobacterium PY146.

Figure 4.9: Effect of different pH on the growth of Mycobacterium PY146.

70

Chapter4Results

Table 4.9: GC mass spectral data of metabolites from phenanthrene, anthracene and
pyrene degradation by resting cells of Mycobacterium PY146

Pyrene

Anthracene

Phenanthrene

Compound

Metabolites

Retention m/z of fragment ions (%)


time (min) relative abundance

1,2-benzenedicarboxylic
acid

17.04

148(23), 105(8), 104(100),


76(68), 74(15), 50(25),
38(6)

1,2Benzenedicarboxylic
acid

17.05

148(23), 104(100), 105(8),


76(66), 74(10), 50(21),
27.2(5)

9,10-anthracendione

25.04

209(16), 208(100),
207(15), 181(12), 180(81),
152(61), 151(32), 150(14),
126(7), 76(20), 50(6)

1,2Benzenedicarboxylic
acid

17.03

148(22), 104(100), 105(8),


76(68), 74(16), 50(22),
43(14), 38.1(7)

1-hydroxypyrene

32.84

219(19), 218(100),
217(12), 190(9), 189(41),
188(4), 187(6), 163(2),
94(3), 43(10)

71

Chapter4Results

Table 4.10: GC mass spectral data of metabolites from phenanthrene, anthracene and
pyrene degradation by growing cells of Mycobacterium PY146

Pyrene

Anthracene

Phenanthrene

Compound

Metabolites

Retention m/z of fragment ions (%)


time (min) relative abundance

Benzene acetic acid

15.14

136(35), 92(20), 91(100),


65(15), 63(6), 51(4), 39(7)

1,2Benzenedicarboxylic
acid

16.04

148(21), 106(9), 104(100),


76(77), 75(11), 74(18),
50(33), 38(7)

Benzenecarboxylic acid

13.89

122(84), 105(100), 77(69),


74(11), 73(13), 51(32),
50(21), 43(17)

Benzeneacetic acid

15.2

136(35), 92(20), 91(100),


65(16), 63(9), 51(4), 39(7)

1,2-benzenedicarboxylic
acid

16.03

147.9(19), 105(8),
104(100), 76(70), 74(17),
50(30), 38(6)

9,10-Anthracenedione

23.92

208(99), 180(100),
152(71), 151(40), 150(19),
126(10), 76(29), 50(13)

Benzenecarboxylic acid

12.95

122(98), 105(100), 87(26),


77(72), 74(8), 73(17),
72(19), 51(28)

Benzeneacetic acid

14.37

136(34), 92(19), 91(100),


79(8), 65(14), 63(6), 51(5)

1,2Benzenedicarboxylic
acid

15.03

148(21), 105(9), 104(100),


76(74), 75(10), 74(18),
50.1(31)

72

Chapter4Results

4.2.4 Biodegradation of PAHs in Soil


Biodegradation of phenanthrene, anthracene and pyrene were checked in soil by
Mycobacterium PY146 with and without additional yeast extract. The extraction
efficiency of the method was 81.6 % for phenanthrene, 83.33 % for anthracene and
77.7 % for pyrene. Figure 4.10 shows the biodegradation of PAHs in soil without
yeast extract, after one week of incubation the detectable concentration of
phenanthrene (98 g), anthracene (60 g) and pyrene (56 g) were decreased to 35,
18 and 27 g, respectively. After 2 weeks of incubation all of the detectable PAHs
were 0.14, 0.45 and 0.19 g respectively.
The addition of yeast extract did not affect the biodegradation of PAHs, except
phenanthrene was detected at 27 g, while anthracene and pyrene were detected 15
and 24 g, respectively after one week of incubation. The detectable PAHs
concentrations of phenanthrene, anthracene and pyrene after 2 weeks of incubation
were 0.08, 0.13 and 0.32 g respectively (Figure 4.11).

73

Chapter4Results

Figure 4.10: Biodegradation of PAHs (phenanthrene, anthracene and pyrene) in soil


by Mycobacterium PY146 without addition of yeast extract.

Figure 4.11: Biodegradation of PAHs (phenanthrene, anthracene and pyrene) in soil


with addition of yeast extract by Mycobacterium PY146.

74

Chapter4Results

4.3 MINERALIZATION STUDIES OF PAHs


4.3.1 14C-PAH mineralization by Bacillus cereus KWS2 and
Mycobacterium PY146.
Mineralization studies of different

14

C-PAHs, including naphthalene, phenanthrene,

pyrene and benzo(a)pyrene were carried out using Bacillus cereus KWS2 and
Mycobacterium PY146 in BSM medium and incubated for one week at 30C.
In case of naphthalene and phenanthrene mineralization by Bacillus cereus KWS2, no
mineralization was observed. In contrast, Mycobacterium PY146 showed the ability to
mineralize the phenanthrene, pyrene and benzo(a)pyrene to CO2. In the case of
phenanthrene,a total of 77.32 % radiolabeled 14C was recovered, 45.92% of 14C was
detected in the form of Na214CO3, 23.19 % of the 14C remained in the supernatant and
8.21 % of the

14

C was detected in the biomass/residues (Figure 4.12). In the case of

pyrene mineralization, a total of 65.77 % radiolabeled 14C was recovered, 39.89 % of


14

C was detected in the form of Na214CO3, 15.18 % of the14C remained in the

supernatant and 10.7 % of the 14C was detected in the biomass/residues (Figure 4.13).
In the in case of benzo(a)pyrene mineralization, a total of 42.84% radiolabeled
was recovered, 4.82 % of

14

14

C was detected in the form of Na214CO3, 22.32 % of

the14C remained in the supernatant and 15.7 % of the

14

C was detected in the

biomass/residues (Figure 4.14).

4.3.2 14C-Pyrene and phenanthrene mineralization using with


different cells densities
The mineralization of pyrene by Mycobacterium PY146 was performed using
incubation times ranging from 4 to 48 hours. Little mineralization occurred in the first
5 hours of incubation but after 20 hours of incubation, 37.9 % of the pyrene was
mineralized and by 48 hours 100 % of the pyrene was mineralized (Figure 4.15). The
middle incubation time of 20 hours was selected for further studies.
The mineralization of pyrene with different cell densities of Mycobacterium PY146
was performed in 40 ml EPA vials. After 20 hours of incubation, the maximum
mineralization was observed in the OD600 0.2 of culture. The mineralization pattern
showed an increase in mineralization in the OD600 0.1 to 0.2 cultures with a

75

Chapter4Results

subsequent decrease in mineralization as the OD600 increased.

The minimum

mineralization was observed by the culture having a cell density of 0.5 (Figure 4.16).
Mineralization of phenanthrene was also evaluated using different Mycobacterium
PY146 cell densities ranging from an OD600 of 0.1 to 0.5. The mineralization pattern
also showed an increase in mineralization between the OD600 0.1 to 0.2 culture where
maximum mineralization was observed. There also was a subsequent decrease in
mineralization as the cell density increased (Figure 4.17).

76

Chapter4Results

Figure 4.12: 14C-phenanthrene mineralization by Mycobacterium PY146.

Figure 4.13: 14C-pyrene mineralization by Mycobacterium PY146.

77

Chapter4Results

Figure 4.14: 14C-benzo(a)pyrene mineralization by Mycobacterium PY146.

Figure 4.15: Mineralization of


different incubation time.

14

C-pyrene by cells of Mycobacterium PY146 at

78

Chapter4Results

60000

Mineralization of Pyrene (DPM)

50000
OD vs DPM
40000

30000

20000

10000

0.0

0.1

0.2

0.3

0.4

0.5

0.6

O.D (600 nm)

Figure 4.16: Mineralization of pyrene with different cell densities by Mycobacterium


PY146.

Mineralization of Phenanthrene (DPM)

70000

60000

OD vs DPM

50000

40000

30000

20000

10000
0.0

0.1

0.2

0.3

0.4

0.5

0.6

O.D (600 nm)

Figure 4.17: Mineralization of


Mycobacterium PY146.

14

C-phenanthrene with different cell densities by

79

Chapter4Results

4.4 EXPRESSION OF PYRENE DIOXYGENASE GENE (nidAB)


FROM MYCOBACTERIUM PY146
4.4.1 Cloning of nidAB gene
The strain Mycobacterium PY146 was grown in LB medium for 72 hours, the
genomic DNA was isolated and used as a template for the amplification of nidAB
gene. After PCR amplification with nidAB primers, an approximately 1.9 Kb
amplicon size of nidAB was detected when the PCR products were run on a 1 %
agarose gel (Figure 4.18).
After the nidAB PCR product was ligated into a pCR 2.1 TOPO vector, 6 L of
recombinant nidAB- pCR 2.1 TOPO vector was added to a vial of competent E.coli
cells for transformation. Transformed E. coli cells were grown on LB agar plates
containing 100 g/mL ampicillin. Single colonies were regrown in LB broth
containing ampicllin and the plasmid DNA was isolated. After digestion with EcoR1,
the isolated plasmid DNA from three samples had both the plasmid backbone
fragment (3.9 Kb) and cloned nidAB gene fragment (1.9 Kb) (Figure 4.19).
Restriction sites for EcoR1 and HindIII were incorporated into both ends of the
nidAB by amplification using the primers nidABEco forward and nidABHind reverse.
The nidAB gene with EcoR1 and HindIII restriction sites was cloned into the TOPO
vector and transformed into competent E. coli cells.

Following plasmid DNA

extraction, the plasmid was sequenced using the M13 forward and reverse primers to
confirm the presence of the EcoR1 and HindIII restriction sites at the ends of nidAB
gene (Appendix I c).
The TOPO cloning vector containing the EconidABHind product and pK233-3
expression vector were digested with EcoRI and HindIII and purified. Following
ligation of the EconidABHind with pK223-3, the plasmid products were transformed
into E. coli and E. coli containing functional plasmids was selected using ampicillin
antibiotics in the medium. Plasmid DNA was isolated from E. coli and restriction
enzyme analysis of isolated plasmid DNA with EcoR1 and HindIII revealed the
presence of both the fragments of pK223-3 vector and nidAB in 6 out of 8 samples

80

Chapter4Results

(Figure 4.20). Direct colony PCR from the transformed E. coli cells with nidAB
primers confirmed the presence of the cloned nidAB gene (Figure 4.21).

4.4.2 Biotransformation of PAHs by E. coli with cloned nidAB into pkk223-3


Resting cell biotransformation of naphthalene, phenanthrene, anthracene and pyrene
was performed using E. coli cells containing nidAB into pkk223-3 grown in LB
medium and with each chemical added individually in shake flask experiments. After
extraction with ethyl acetate and analysis on GC/MS, no hydroxy metabolites of any
of the PAH were detected.
In addition, no metabolites were detected for any of the PAH in GC/MS analysis after
derivatization

with

N,O-Bis(trimethylsilyl)trifluoroacetamide

with

Trimethylechlorosilane (BSTFA).

4.4.3 Invitro RNA synthesis of nidAB transcripts


RNA transcripts of nidAB were synthesized invitro using the T7 RiboMAX
Express Large Scale RNA Production System from the TOPO plasmid. Following
invitro transpcription, the concentration of nidAB RNA was 95.3 ng/L as measured
by the nanodrop spectrophotometer . This RNA was used as a standard for RT-QPCR.

4.4.3.1 Quantification of RNA transcripts of nidAB from E. coli


For the expression of nidAB gene, the transformed E. coli cells (clone numbers 6, 7
and 8 containing pKK223-3) were grown in LB medium containing ampicillin with
and without IPTG and the RNA were isolated. The RNA copy number/ml of bacterial
culture were quantified by RT-QPCR which revealed an ~ 2- fold increase in RNA
transcripts in one sample grown in the presence of IPTG (Figure 4.22).

4.4.4 Comparison of nidAB DNA and RNA copy number using realtime PCR with PAH mineralization in Mycobacterium PY146
Total DNA and RNA isolated and purified from Mycobacterium PY146 were
quantified by using real time PCR. Both nidAB genes (DNA) and transcripts (RNA)
increased with increasing OD600 (Figure 4.23). An RNA to DNA ratio also calculated
81

Chapter4Results

which showed that the RNA to DNA ratio increased between OD600 0.1 to 0.4 and
then declined at an OD600 of 0.5 (Figure 4.24). In order to relate mineralization
activity with nidAB RNA transcripts, phenanthrene mineralization was compared to
the RNA/DNA ratio. These results indicated that there was a positive relationship
between phenanthrene mineralization and the RNA/ DNA ratio (Figure 4.25).
Comparison of pyrene mineralization with and without added unlabelled pyrene, with
the nidAB RNA/DNA ratio by different cell concentration indicated that the nidAB
RNA/DNA ratio only corresponded to the amount of pyrene mineralization between
an OD600 of 0.3 to 0.5 (Figure 4.26).

82

Chapter4Results

1938bp

Figure 4.18: Ethidium bromide stained agarose gel showing the PCR product of
nidAB gene from Mycobacterium PY146. Lane L: 1KB DNA ladder; Lanes 1: nidAB.
The arrow indicates the location of the nidAB product.
L

~ 3.9 Kb
~ 1.9 Kb

Figure 4.19: Ethidium bromide stained 1 % agarose gel showing the location of the
plasmid (top arrow) and nidAB gene (bottom arrow) after digestion with EcoR1. Lane
L: 1KB DNA ladder; Lanes 1, 2 and 5: have both the plasmid and nidAB.

83

Chapter4Results

L 1

Figure 4.20: Ethidium bromide stained 1 % agarose gel showing the plasmid and
nidAB after digestion with EcoR1 and HindIII. Lane L: 1KB DNA ladder; Lanes 1
and 2 have only plasmid; Lanes 3-8: have both the plasmid and nidAB.
L

~ 1.9 Kb

Figure 4.21: Ethidium bromide stained 1 % agarose gel showing the presence of
cloned nidAB gene in transformed E. coli. Lane L: 1KB DNA ladder; Lanes 1 to 6: all
have the cloned nidAB.

84

Chapter4Results

Figure 4.22: Copies of nidAB RNA transcripts in E. coli clones grown in LB


containing ampicillin with and without IPTG.

85

Chapter4Results

Copies DNA or RNA/ml of Bacteria

1e+10

DNA copies/ml (r2 =0.90)


RNA copies/ml (r2= 0.92)
1e+9

1e+8

1e+7
0.0

0.1

0.2

0.3

0.4

0.5

O.D at 600 nm

Figure 4.23: Standard curves of DNA and RNA/ml of bacteria from different cell
densities from 0.1 to 0.5 O.D.

2.8
2.6

RNA to DNA Ratio

2.4

Ratio RNA/DNA

2.2
2.0
1.8
1.6
1.4
1.2
1.0
0.8
0.0

0.1

0.2

0.3

O.D at 600 nm

0.4

0.5

Figure 4.24: RNA to DNA ratio at different cell densities from 0.1 to 0.5 O.D.

86

Chapter4Results

2.8
2.6

60000

2.4
50000
2.2
40000

2.0
1.8

30000

1.6

RNA/DNA Ratio

Mineralization of Phenanthrene (DPM)

70000

20000
1.4

OD vs Phenanthrene
OD vs ratio

10000

1.2
1.0
0.0

0.1

0.2

0.3

0.4

0.5

0.6

O.D at 600 nm

Figure 4.25: Comparison of mineralization of phenanthrene from different cell


densities and the nidAB RNA/DNA ratio in Mycobacterium PYR146.

1e+5

2.8

2.4
1e+4

2.2
2.0
1.8

1e+3

1.6
OD vs DPM without add pyrene
OD vs DPM with add pyrene
OD vs ratio

RNA/DNA Ratio

Mineralization of Pyrene (DPM)

2.6

1.4
1.2

1e+2

1.0
0.0

0.1

0.2

0.3

0.4

0.5

O.D at 600 nm

Figure 4.26: Effect of added un-labled pyrene on the mineralization of 14C- labled
pyrene at different cell densities of Mycobacterium PYR146 and the nidAB
RNA/DNA ratio.

87

Chapter5Discussion
Polycyclic aromatic hydrocarbons (PAHs) are a ubiquitous class of organic
compounds present in the environment. Microbial degradation of PAHs by bacteria
and fungi is considered a safe and environment friendly method to remove toxic
contaminants. In the present study, bacterial isolates utilizing the PAHs (naphthalene,
anthracene, phenanthracene, and pyrene as sole carbon and energy were characterized
taxonomically and for their degradation potential for the respective PAH.
Soils contaminated with hydrocarbon are good sources for the isolation of PAHs
degrading bacteria (Jacques et al., 2009; Al-Thani et al., 2009) which can then be used
for the removal of such compounds from the environment. In this context, the
bacterial strains in this study which were isolated from the crude oil contaminated soil
utilized polycyclic hydrocarbons as sole source of carbon energy. Among these five
selected bacterial strains, KWS2, KWS4, DW3, Sol-10 and 10-1 showed the ability to
tolerate the PAHs concentration up to 1200 ppm in case of anthracene and 800 ppm in
case of pyrene (Table 4.1 & 4.2). These selected isolates were identified as; Bacillus
cereus, Bacillus licheniformis and Pseudomonas stutzeri (Table 4.6). Previously
different strains of Bacillus have been isolated from PAHs contaminated soil (Jacques
et al., 2009; Das and Mukherjee, 2007; Lin et al., 2010), which have the potential to
biodegrade and utilize organic compounds like phenanthrene, anthracene,
naphthalene, biphenyl and some other aromatic compounds. Thermophillic bacteria
belonged of genus Geobacillus have also been reported by Bubinas et al., (2008)
capable of utilizing naphthalene as a sole source of carbon. Pseudomonas species
have been well studied for their ability to degrade the different type of pollutants
including; PCBs (De et al., 2006), textile dye direct orange 102 (Pandey et al., 2010)
and crude oil components (Tang et al., 2007). Pseudomonas species have also been
reported in several studies for their degradation potential of PAHs from naphthalene
to benzo(a)pyrene

isolated from different PAHs contaminated soils. Coral and

Karagoz, (2005) isolated several Pseudomonas strains from crude oil polluted soil of
a petroleum refinery which degraded phenanthrene efficiently. Pseudomonas species
also were isolated from soil samples from located at a petrochemical treatment plants
showed degradation capability toward different PAHs (Jacques et al., 2009; Molina et
al., 2009; Aitken et al., 1998). The isolation of bacterial strains in the present study
from oil contaminated soil suggested that these isolates may be well adapted for use
88

Chapter5Discussion
because of pre-exposure and acclimatization to the contaminants and can be used
effectively in remediating the contaminated sites.
In biodegradation studies of anthracene by Bacillus cereus KWS2, a 45 % reduction
in anthracene concentration was observed at pH 7 and 30C after seven days of
incubation (Figure 4.1). These results are consistent with other reports on the
efficiency of Bacillus to degrade PAHs. For instance, Bacillus fusiformis isolated
from wastewater sludge of an oil refinery showed optimal naphthalene degradation at
pH 7 and 30C (Lin et al., 2010). Shuyu et al., (2007) reported the isolation of two
bacterial strains identified as Bacillus sp. having the ability to degrade pyrene
individually as well as in consortium and optimally at 37C at pH 7. A novel strain of
Bacillus subtilis which can utilize several PAHs including benzo(a)pyrene,
anthracene, naphthalene and dibenzothiophene as sole source of carbon energy was
isolated from a contaminated soil at an automobile workshop (Lily et al., 2009). The
pH of the medium affect the degradation rate, as Kim et al., (2005) suggested that
PAHs at pH 6.5 can get enter to the bacterial cell rapidly comparing to pH 7.5. The
degradation of naphthalene and anthracene by Pseudomonas stutzeri 10-1 was
45.35% reduction for naphthalene (in two weeks) (Figure 4.5) while 95.56 %
reduction in anthracene (three weeks time) concentration (Figure 4.6) was achieved.
Pseudomonas stutzeri P-16 and P. saccharophila P-15 isolates from creosote
contaminated soil were reported to degrade naphthalene, 2-methylnaphthalene and 1methylnaphthalene (Stringfellow and Aitken, 1995). Maximum degradation of
naphthalene by Pseudomonas sp. HOB1 was achieved in the pH and temperature
ranges of 7.5-8.5 and 35-37C, respectively (Pathak et al., 2009). The significant
reduction in total petroleum hydrocarbon (TPH) concentration in the soil after
treatment was reported by a consortium of Bacillus subtilis and two strains of
Pseudomonas aeruginosa (Das and Mukherjee, 2007).
For biodegradation the microbial potential is very critical due to the presence of
catabolic genes on plasmid and chromosomes enable them to degrade and utilize as
carbon source. In case of PAHs the compounds are not easily available, some bacteria
have the potential to produce enzymes and biosurfactant to solubilize the compounds
and make available for the degradation.
89

Chapter5Discussion
In addition to microbial potential, the right ecological conditions are also very
important for microbial activity for biodegradation. The temperature, pH, oxygen and
nutrients availability often affect the rate of biodegradation. Temperature and pH of
the medium affects the PAHs degradation by increasing the solubility of the
compound (Lau et al., 2003) and also the optimum temperature enhances the
microbial activity. As the aerobic metabolism of PAHs is efficient than anaerobic,
oxygen is required for mono and dioxygenases for the initial oxidation of PAHs.
Inorganic nutrients serve always as limiting factor and required for the microbial
degradation and growth on PAHs. The studies showed the enhanced degradation rates
of PAHs when potassium, nitrogen and phosphorous were added to the medium
(Hwang et al., 1994).
A variety of catabolic mechanisms including role of different oxygenases have been
reported in Bacillus regarding biodegradation of monocyclic aromatics and PAHs.
Such metabolic transformation abilities of Bacillus were also evaluated in the present
study by using Bacillus cereus strains KWS2 and KWS4. These results proved that
these isolates were capable of transforming and degrading naphthalene, phenanthrene
and anthracene to metabolites when observed through GC/MS. Metabolites were
detected in GC/MS chromatograph analysis when naphthalene was used as growth
substrate in resting cells of B. cereus KWS2 and B. cereus KWS4. These were 2naphthol, benzeneacetic acid, 2-hydroxy benzeneacetic acid, benzenaldehyde and
benzoic acid (Fig 5.1). One metabolite identified as 2-naphthol which have been
described as intermediate of naphthalene degradation by B. cereus (Cerniglia et al.,
1984) and Bacillus thermolrovanas (Annweiler et al., 2000; Bubinas et al., 2008). oPhthalic acid was not detected in our study but was confirmed as transformation
metabolites of 1-naphthol or 2-naphthol in studies with Geobacillus (Bubinas et al.,
2008). Benzoic acid was detected as metabolite of naphthalene degradation by B.
cereus KWS2 and KWS4 (Fig 5.1) which is also reported in the naphthalene
degradation by B. thermoleovorans (Annweiler et al., 2000).
Metabolites accumulating during resting cells incubation with three ring compound
phenanthrene with B. cereus KWS2 and B. cereus KWS4 showed monoxygenase
activity and dioxygenase activity as mono-hydroxy phenanthrene and cis-dihydroxy90

Chapter5Discussion
dihydro phenanthrene were detected showing attack on C9 and C10 position.
Degradation pathway described by Staphylococcus sp. PWY was showing initial
attack at C1 and C2 position giving 1,2-dihydroxy phenanthrene (Mallick et al.,
2007). Further metabolism of 9,10 dihydro, dihydroxy phenanthrene to 3-hydroxy
benzeneacetic acid, 2-hydroxy benzeneacetic and 4-hydroxy benzeneacetic acid
(Figure 5.1) showing ring cleavage and hydroxylation again by monoxygenase attack.
Phenanthrene trans 9-10 dihydrodiol was also detected in the growth medium of
Mycobacterium strain grown on commercial anthracene (Tongpim and Pickard, 1999)
and by Streptomyces flavovires (Sutherland et al., 1990). Luan et al., (2006) also
reported a mono and dihydroxy phenanthrene from phenanthrene degradation by a
bacterial consortium. Other gram negative bacteria known to oxidize anthracene by
dioxygenase to anthracene cis-1,2-dihdrodiol (Cerniglia and Heitkamp, 1989), while
gram positive bacteria like Coryneform bacillus (Bouchez et al., 1996) also degraded
anthracene but no metabolites were identified.
In the present study, 2-hydroxy benzoic acid (salicylic acid) was detected in the
phenanthrene biotransformation by B. cereus KWS4 (Figure 5.1). Our finding
confirms those reported by Mallick et al., (2007) who identified the salicylic acid
from phenanthrene biotransformation by Staphylococcus sp. PN/Y and proposed its
further transformation to catechol.
Over the last decade the genus Mycobacterium has been proved to be one of the best
bacterial types involved in PAHs degradation. In this context a number of
Mycobacterium species having PAHs degradation capabilities have been isolated
from river and fresh water sediments (Kageyama et al., 2007; Churchill et al., 1999).
Their metabolic successes have been associated with their cell wall constituent
mycolic acid which provides great resistant abilities to withstand recalcitrant
hydrocarbons which later help them to metabolize those compounds. Moreover, the
hydrophobic nature of mycolic acid helps developing biofilms on surface of PAHs,
which can later be help promote community biodegradation (Wick et al., 2002a).
Mycobacterium sp. has high specific affinity for PAHs and grows on the surface of
solid anthracene when provided in low amounts, attached to the crystal and form
biofilm and use as a carbon source (Wick et al., 2002b). The present study has also
used strain of Mycobacterium PY146 previously isolated from PAH contaminated
91

Chapter5Discussion
sediment of Lake Erie (Debruyn et al., 2007). This strain showed 98% homology with
Mycobacterium vanbaalenii PYR-1 which efficiently degrades and mineralizes
pyrene and phenanthrene in liquid medium as well as in microcosms. Phenanthrene
and pyrene degradation pathways also have been constructed for this strain (Heitkamp
and Cerniglia 1989; Heitkamp et al., 1988a,b; Stingley et al., 2004a,b; Kim et al.,
2004).
In PAHs biotransformation studies, the presence of metabolites in the medium
suggested that Mycobacterium PY146 has the ability to degrade phenanthrene,
anthracene and pyrene efficiently. 1,2-benzenedicarboxyic acid, also known as
phthalic acid, was detected as degradative metabolite from all the three compounds
tested.

Pyrene metabolism by Mycobacterium PY146, identified an 1-hydroxy

pyrene, showing momooxygenase attack. In addition 9, 10-anthracenedione was


detected as a metabolite from anthracene and 1-hydroxypyrene from pyrene
degradation (Table 4.9 & 4.10; Figure 5.2). Phthalic acid was also detected in all the
phenanthrene, anthracene and pyrene. Phthalic acid has also been reported in pyrene
metabolism by Mycobacterium sp. AP1 (Vila et al., 2001). Phthalic acid is also
produced in the degradation pathway of phenanthrene by bacterial consortium (Luan
et al., 2006). It is known that phthalic acid is further metabolized in aerobic bacteria
through protocatecuate as intermediate (Grifoll et al., 1994).
Phthalic acid from phenanthrene and 9,10-anthraquinone from anthracene
biodegradation along with other metabolites by Mycobacterium PYR-1 were also
reported by Moody et al., (2001). Dean-Ross and Cerniglia, (1996) reported the
phthalic acid and other metabolites from pyrene by Mycobacterium flavescens. The 1hydroxypyrene detection from pyrene biotransformation by Mycobacterium PY146 in
the present studies suggests a monoxygenase activity. Brenza et al., (2006) reported
the presence and activity of cytochrome P450 monoxygenase gene in Mycobacterium
vanbaalenii PYR-1 and detected a 1-hydroxypyrene from pyrene degradation. Zhong
et al., (2006) also reported the presence of both mono and di-hydroxypyrene
metabolites from pyrene degradation, suggested that the PAH molecule metabolized
through mono and dihydroxylation by Mycobacterium sp. strain A1-PYR.

92

Chapter5Discussion
Biodegradation of PAHs in soil by microbial action is of great importance because
soil and sediment is a sink for the accumulation of hydrophobic pollutants especially
polycyclic aromatic hydrocarbons. Potentially, the rate of biodegradation could be
changed by the addition of some nutrients. In the biodegradation studies of PAHs;
phenanthrene, anthracene and pyrene in soil revealed the removal of detectable
compounds by Mycobacterium PY146 in two weeks at pH 7 and at 30C (Figure
4.10). The addition of yeast extract did not have a significant effect on biodegradation
of PAHs in our study except on phenanthrene degradation after one week of
incubation (Figure 4.11).
Chang et al., (2001) studied the potential of a microbial consortium in the soil to
degrade phenanthrene. Phenanthrene was degraded efficiently at pH 7 and at 30C and
the degradation rate was enhanced by the addition of yeast extract and, glucose. Wan
et al., (2003) reported the removal of PAHs from spiked soil by composting and
reported a 90 % degradation of anthracene, phenanthrene and pyrene after 30 days.
The addition of pig manure slightly enhanced biodegradation soybean and sewage
sludge. Biodegradation of PAHs (phenanthrene and pyrene) also was studied in a soil
slurry reactor using an immobilized bacterial culture of Zoogloea. . This study showed
an overall degradation of pyrene and phenanthrene was 75.4 and 87 % respectively.
(Li et al., 2005). Wang et al., (2006) used immobilized bacterial consortium to
degrade the PAHs pyrene and benzo(a)pyrene in contaminated soil with a degradation
efficiency of 43.49 and 38.55 % respectively in 96 hours. Su et al., (2007) studied the
ability of three bacterial isolates Bacillus sp. SB02, Zoogloea sp.SB09, and
Flavobacterium sp.SB10 individually to degrade pyrene and benzo(a)pyrene in soil.
All three isolates degrade the PAHs, Bacillus sp. SB02 degrade 42.69 and 33.04 % of
pyrene and benzo(a)pyrene, respectively with the highest rate. The results of
biodegradation of PAHs in soil in present studies is similar and in accordance with the
data presented here from the different studies, while the effect of additional nutrients,
yeast extract had no significant effect on the overall biodegradation process, the
similar results reported by Juhasz et al., (2000) that the addition of yeast extract on the
biodegradation of high molecular weight PAHs had no effect but only affect
positively on low molecular PAHs biodegradation. The supplement of additional
carbon source such as glucose reduces the degradation of PAHs probably due to the
93

Chapter5Discussion
availability of easily degradable carbon by Burkholderia cocovenenans (BU-3)
(Wong et al., 2002) and glucose supplementation also reduced the PAHs dioxygenase
activities (Tiana et al., 2003).
The ultimate goal of biodegradation of organic pollutants is mineralization of the
problematic compound into inorganic constituents such as CO2 and H2O, because
partially degraded compounds often have more toxic effects than their parent
compounds. Mineralization of different PAHs by Bacillus cereus KWS2 and
Mycobacterium PY146 was evaluated.

Bacillus cereus KWS2 was not able to

mineralize the tested PAHs including naphthalene and phenanthrene. Mineralization


was observed for Mycobacterium PY146 and 14CO2 was evolved from phenanthrene,
pyrene and benzo(a)pyrene with yields of 45.92, 39.89 and 4.82 %, respectively
(Figure 4.12; 4.13 & 4.14). Naphthalene was not mineralized by Mycobacterium
PY146. The mineralization results from Mycobacterium PY146 were consistent with
those reported by Moody et al., (2001) which showed the mineralization capability of
Mycobacterium PYR-1 to be 45 and 52 % for anthracene and phenanthrene,
respectively. In another study, Sho et al., (2004) showed that Mycobacterium S65,
isolated from jet fuel contaminated site, was able to mineralize fluoranthene,
phenanthrene and pyrene but not fluorene, anthracene and naphthalene. Dean-Ross
and Cerniglia (1996) also showed that Mycobacterium flavescens could mineralize
phenanthrene, pyrene and fluoranthene but not other PAHs such as naphthalene,
acenaphthalene, anthracene, chrysene, fluorene and benzo(a)pyrene. Finally,
Mycobacterium strain CH1 has the ability to mineralize fluoranthene, phenanthrene
and pyrene, but was unable to mineralize anthracene, fluorene and naphthalene
(Churchill et al., 1999). These results suggest that all Mycobacterium sp. including
Mycobacterium PY146 show selective preference for mineralization of some PAHs.
Sho et al., (2004) suggested that the angularity of the PAH is the limiting factor, and
since naphthalene and anthracene are linear in structure the initial attack of the
microbes on PAH is in the K and bay region of the compound. Sundberg et al., (1997)
also suggested the bacterial attack on bay region is very common because of the open
space of the exposed carbon atoms of PAHs in this region. The intradiol cleavage at K
region diol was also reported by Rhemann et al., (1988) and Vila et al., (2001). The
Mycobacterium is very selective in the initial oxidation of PAHs. M. vanbaalenii
94

Chapter5Discussion
PYR-1 initially attacked the pyrene and phenanthrene nucleus in the 4,5 and 9,10
carbon positions, the K region of each PAH. The oxidation of PAHs at the K-region
bonds seems to be unique to Mycobacterium species (Moody et al., 2004).
Cell density and inoculum age may affect microbial metabolic activity. Thus
mineralization of

14

C pyrene and phenanthrene was evaluated using different cell

densities ranging from OD600 0.1 to 0.5 of Mycobacterium PY146 cultures. Maximum
mineralization of pyrene was observed in the culture with an OD600 of 0.2 (Figure
4.16), while in the case of phenanthrene there was an increase in mineralization in
cultures with an increasing OD600 from 0.1 to 0.2 (Figure 4.17). There were decreases
in both pyrene and phenanthrene mineralization as the cell density increased. Kanaly
et al., (2002a,b) used a microbial consortium isolated from soil to mineralize
benzo(a)pyrene and reported a positive effect on mineralization of BaP after the
addition of diesel fuel and with increasing inoculums sizes up to 10 fold, decreases
the lag phase and a little increase in the mineralization but not what they expected to
be more active mineralization. Ye et al., (1996) reported an effect of cell density of
Sphingomonas paucimobilis on the mineralization of different PAHs with an increase
in mineralization increases with an increase in cell density of the inoculums. Shin et
al., (2005) studied the effect of bacterial growth on phenanthrene degradation and
reported the positive effect of increasing growth on degradation. Semple et al., (2006)
studied the effect of additional inoculum on phenanthrene mineralizing Pseudomonas
and the addition of

14

C-phenanthrene on the mineralization of radio labeled

phenanthrene and suggested the additional phenanthrene was mineralized efficiently


but the addition of bacterial cells did not increase mineralization. The results are not
in accordance with the most of studies which showed the mineralization increases
with the increase in cell density. In our studies the maximum mineralization of PAHs
was observed by the lower cell densities cultures from OD600 0.1 to 0.2, which were
the optimal cell densities. As from the data of the growth curve of Mycobacterium
PY146 (Figure 4.7) the mid log phase is 0.2-0.3 where the cells are highly active
metabolically, and the maximum regulation of catabolic genes occurs.
Aerobic microbial degradation of PAHs is mediated by several different oxygenase
enzymes, including mono and dioxygenase. In some cases, the expression of catabolic
genes can be studied by transferring the gene into a heterologous host like E. coli
95

Chapter5Discussion
(Kurkela et al., 1988; Laurie and Lloyd Jones, 1999 and Kasuga, 2001). For instance,
Sphingomonas LH128, isolated from a PAHs contaminated soil has a dioxygenase
complex, phnA1fA2f, responsible for oxidation of different PAHs. When this gene
was cloned and over expressed in E. coli, the phnA1fA2f gene was shown to confer
the ability to oxidize a different range of PAHs from low molecular weight dibenzo-pdioxin and chlorinated biphenyls to high molecular weight pyrene, chrysene, and
benz(a)anthracene (Schuler et al., 2009). Similarly, the pyrene dioxygenase gene,
nidAB which is encoded by nidA and B ( and subunits) in Mycobacterium is
responsible for the ring hydroxylation of pyrene, which is the initial step in pyrene
metabolism (Khan et al., 2001). In this study, the attempts were carried to express the
pyrene dioxygenase gene (nidAB) from Mycobacterium PY146 in E. coli by cloning
it in an expression vector followed by transforming E. coli. The biotransformation of
naphthalene, phenanthrene, anthracene and pyrene was checked using E. coli cells
containing nidAB, but no hydroxy metabolites or derivatized metabolite were
detected.

These results differ from Khan et al., (2001) who demonstrated the

functionality of the dioxygenase of Mycobacterium PYR-1 after cloning it into E. coli.


They reported the expression of nidAB, carried on a plasmid and nidABD, carried on
a phagemid both individually by transforming the pyrene to pyrene-4,5-dihydrodiol
without reductase and ferrodoxin components. They suspected that this reaction
occurred because the cells were borrowing the reductase and ferredoxin components
from elsewhere in the cell to compliment the NidAB dioxygenase subunits. These
reductase and ferrodoxin are required by the dioxygenase systems similar to nidAB
for the electron transfer (Treadway, et al., 1999). From present study, the possible
cause of the failure in expression of the nidAB gene in E. coli may be due to the lack
of the reductase and ferrodoxin components required for the electron transfer in the
dioxygenase system. This idea is supported by Kim et al., (2006) who studied the
expression of nidA3B3 from Mycobacterium PYR-1in transformed E. coli and found
that no PAH degradation activity was observed until they introduced another plasmid
carrying the ferrodoxin reductase and ferrodoxin from Nocardioides sp. KP7 into the
E. coli. They verified the coexpression of the genes in E. coli using several PAHs
including; naphthalene, anthracene, phenanthrene, fluoranthene, pyrene and
benz(a)anthracene which were all converted to the expected cis-dihydrodiol
metabolites.
96

Chapter5Discussion
The PAH degrading microbial population can be estimated in environmental samples
by quantifying the specific catabolic gene and its activity by quantifying the RNA
transcripts and determining the RNA to DNA ratio. In our study we used the nidAB
RNA transcript to DNA ratio calculated using RT-QPCR for different cell densities of
Mycobacterium PY146. The RNA transcript to DNA ratio increased as the cell
density increased from OD600 0.1 to 0.4 and then declined when the OD600 reached 0.5
(Figure 4.24). The relationship of phenanthrene and pyrene mineralization to the RNA
transcript to DNA ratio was positive for phenanthrene between OD600 0.1 to 0.5
(Figure 4.25), but was only positive for pyrene between OD600 of 0.3 to 0.5 (Figure
4.26). Genes with catabolic activities, such as pyrene dioxygenase nidAB not only
play an important role in the contaminant removal but also serve as markers for the
degradation potential. In our studies the use of RNA transcript to DNA ratio of nidAB
to correlate with the PAHs mineralization gave the clear indication in case of
phenanthrene and pyrene mineralization capability of the bacterial isolate.
Sanseverino et al., (1993) used DNA hybridization with the nahA gene probe,
naphthalene lux reporter system and the messenger RNA transcript of nahA gene to
study PAHs degradation and the bacterial population in soil. They reported a positive
correlation of nahA transcript with the mineralization rate of naphthalene as well as
with the densities of naphthalene degrading microbes determined by colony
hybridization. Park and Crowely, (2006) reported a thousand fold increase in gene
copy number of nahAc when cells were exposed to naphthalene. They also observed
the direct relationship between the gene copy number and the naphthalene
degradation rate in soil contaminated with PAHs. Fleming et al., (1993) studied the
expression of nahA catabolic gene in soil contaminated with PAHs. They also found a
positive correlation of nahA transcript with the naphthalene mineralization and nahA
gene frequency. Tuomi et al., (2004) correlated nahAc gene abundance with the
aerobic

14

C-naphthalene mineralization, total microbial cell counts, respiration rate

and organic matter content in oxic and anoxic soil samples, they found nahAc gene
abundance was correlate with the

14

C-naphthalene mineralization in the oxic soil

layers and the total microbial cell counts were correlated with nahAc gene abundance
in both anoxic and oxic layer of soil. The results of the present study are similar to
these other studies with respect to phenanthrene in that the RNA to DNA ratio of the
97

Chapter5Discussion
nidAB relates with the mineralization of phenanthrene. However, in the case of
pyrene mineralization, the RNA transcript to DNA ratio did not relate to pyrene
mineralization. That may be due the possible activity of the monoxygenase in the
Mycobacterium PY146 in addition to the nidAB as seen from the detection of 1hydroxyryrene metabolites from pyrene degradation.
In this study, the degradation potential, including transformation and mineralization,
were characterized in different isolates from different genera and geographical
locales. From the study it is summarize, that the isolate Bacillus cereus KWS2 and
KWS4 has the potential to degrade and transform different PAHs but could not
mineralize them. In contrast, Mycobacterium PY146 has both the degradation and
mineralization potential for most of the studied PAHs. The expression of pyrene
dioxygenase nidAB in a heterologous system requires the ferrodoxin and reductase
components which was not present when the nidAB was cloned into E. coli. It is also
suggested that monoxygenase along with the nidAB dioxygenase is present and active
in the Mycobacterium PY146. Therefore the use of these isolates alone or in
combination could result in effective degradation of PAHs and related compounds in
the bioremediation process in the contaminated environment.

98

Chapter5Discussion

Naphthalene

Phenanthrene

OH
OH
HO

2-naphthol

OH

9-phenanthrenol
HO

9,10 Dihydro 9,10 dihydroxyphenanthrene


Benzeneacetic acid
OH
OH

O
O

HO

HO
HO

Benzeneacetic acid

Phthalic acid

Phthalic acid

HO
HO

HO

Benzenecarboxylic acid
(Benzoic acid)

2-hydroxy benzoic acid


(Salicylic acid)
HO

HO

HO

HO

Catechol

Catechol
Benzaldehyde

Figure 5.1: Proposed pathway for the degradation of naphthalene and phenanthrene by
B. cereus KWS2 & KWS4. Compounds in brackets are from other reported studies.

99

Chapter5Discussion

Anthracene

Phenanthrene

Pyrene
OH

O
O
HO

HO

1,2 benzenedicarboxylic acid

9,10 anthracendione
1, hydroxypyrene

O
OH

OH

HO

Benzeneacetic acid

HO

1,2 benzenedicarboxylic acid

HO

1,2 benzenedicarboxylic acid

HO

HO

Benzenecarboxylic acid

Benzenecarboxylic acid

O
O
HO

Benzeneacetic acid

HO

Benzeneacetic acid

Figure 5.2: Metabolites detected from the biotransformation of phenanthrene,


anthracene and pyrene by Mycobacterium PY146.

100

Conclusions

Conclusions
PAHs contaminated soil is the rich source of PAH degrading microbes.
Bacillus was found to be the major contributor amongst the isolated bacterial
strains from Pakistan.
From the GC/MS data Bacillus cereus KWS2 and B. cereus KWS4 are able to
transform naphthalene, phenanthrene and anthracene.
The GC/MS and mineralization data showed that, the Mycobacterium PY146
degrades phenanthrene, anthracene, pyrene and benzo(a)pyrene but cant
mineralize naphthalene.
The degradation potential of Mycobacterium PY146 in soil with and without
yeast extract was almost similar.

There is an effect of cell biomass on mineralization of phenanthrene and


pyrene.

There was negative effect of additional pyrene on radiolabled pyrene


mineralization.
Cloning of nidAB from Mycobacterium was successful but cant expressed in
E.coli in the present study. It might require its specific ferrodoxin and
reductase proteins for its expression.
There was a positive relation of RNA to DNA ratio of nidAB and
phenanthrene mineralization.

101

FutureProspects

Future prospects
PAH contaminated environments like soil and sediments can be exploited for
other bacterial strains capable of degrading PAHs.
These isolates can also be checked for their ability to degrade other organic
contaminants like pesticides, TNTs and PCBs etc.
These bacterial isolates can be applied in fields for bioremediation of PAHs
and other organic compounds.
Gene probes can be constructed to monitor PAH-degrading bacteria in
contaminated environments.
The expression of nidAB pyrene dioxygenase gene can be checked using other
competent E. coli cell lines and expression vectors.
The expression of monoxygenase gene could be checked and relate its
expression with the mineralization of PAHs.
Identification of metabolites from PAHs biotransformation to elucidate the
complete pathways by selected isolates.

102

References
Abd-Elsalam, H. E., E. E. Hafez, A. A. Hussain, A. G. Ali and A. A. El-Hanafy.
2009. Isolation and identification of three-rings polyaromatic hydrocarbons
(anthracene and phenanthrene) degrading bacteria. American-Eurasian J. Agric.
Environ. Sci. 5(1): 31-38
Aguinaga, N., N. Campillo, P. Vinas and M. H. Cordoba. 2007. Determination of 16
polycyclic aromatic hydrocarbons in milk and related products using solid-phase
microextraction coupled to gas chromatographymass spectrometry. Analytica
Chimica Acta. 596: 285-290
Aitken, M. D., W. T. Stringfellow, R. D. Nagel, C. Kazunga and S. H. Chen. 1998.
Characteristics of phenanthrene-degrading bacteria isolated from soils contaminated
with polycyclic aromatic hydrocarbons. Can. J. Microbiol. 44: 743-752
Alexander, M. 1999. Cometabolism. In: M. Alexander (Ed.), Biodegradation and
bioremediation (pp. 249267). New York: Academic Press.
Al-Thani, R. F., D. A. M. Abd-El-Haleem and M. Al-Shammri. 2009. Isolation and
characterization of polyaromatic hydrocarbons-degrading bacteria from different
Qatari soils. Afr. J. Microbiol. Res. 3(11): 761-766
Al-Turki, A. I. 2009. Microbial polycyclic aromatic hydrocarbons degradation in soil.
Res. J. Environ. Toxicol. 3(1): 1-8
Al-Turki, A. I. and W. A. Dick. 2003. Myrosinase activity in soil. Soil Sci. Soc. Am. J.
67: 139-145
Annweiler, E., H. H. Richnow, G. Antranikian, S. Hebenbrock, C. Garms, S. Franke,
W. Francke and W. Michaelis. 2000. Naphthalene degradation and incorporation of
naphthalene-derived

carbon

into

biomass

by

the

thermophile

Bacillus

thermoleovorans. Appl. Environ. Microbiol. 66(2): 518-523


Arora, P. K., M. Kumar, A. Chauhan, G. P. S. Raghava and R. K. Jain. 2009.
OxDBase: a database of oxygenases involved in biodegradation. BMC Research
Notes. 2:67.online http://www.biomedcentral.com/1756-0500/2/67
103

References
Bach, Q. D., S. J. Kim, S. C. Choi and Y. S. Oh. 2005. Enhancing the intrinsic
bioremediation of PAH-contaminated anoxic estuarine sediments with biostimulating
agents. J. Microbiol. 43(4): 319-324
Baek, K. H., H. S. Kim, H. M. Oh, B. D. Yoon, J. Kim and I. S. Leel. 2004. Effects of
crude oil, oil components, and bioremediation on plant growth. J. Environ. Sci.
Health. 39(9): 2465-2472
Bamforth, S. M. and I. Singleton. 2005. Bioremediation of polycyclic aromatic
hydrocarbons: current knowledge and future directions. A Review. J. Chem. Technol.
Biotechnol. 80: 723-736

Banat, I. M., R. S. Makkar and S. S. Cameotra. 2000. Potential commercial


applications of microbial surfactants. Appl. Microbiol. Biotechnol. 53: 495-508
Basta, N. T., J. A. Ryan and R. L. Chaney. 2005. Trace element chemistry in residualtreated soil: key concepts and metal bioavailability. J. Environ. Quality. 34: 49-63
Bezalel, L. Y. Hadar, P. P. Fu, J. P. Freeman and C. E. Cerniglia. 1996. Metabolism
of phenanthrene by the white rot fungus Pleurotus ostreatus. Appl. Environ.
Microbiol. 62(7): 2547-2553
Bezalel, L., Y. Hadar and C. E. Cerniglia. 1997. Enzymatic mechanisms involved in
phenanthrene degradation by the white-rot fungus Pleurotus ostreatus. Appl. Environ.
Microbiol. 63: 2495-2501
Bishnoi, K., R. Kumar and N. R. Bishnoi. 2008. Biodegradation of Polycyclic
aromatic hydrocarbons by white rot fungi Phanerochaete chrysosporium in sterile and
unsterile soil. J. Sci. Indust. Res. 67: 538-542
Bishoni, N. R., U. Mehta and G. G. Pandit. 2006. Quantification of polycyclic
aromatic hydrocarbons in fruits and vegetables using high performance liquid
chromatography. Ind. J. Chem. Technol. 13: 30-35
Bollag, J. M. and C. Myers. 1992. Detoxification of aquatic and terrestrial sites
through binding of pollutants to humic substances. Sci. Tot. Environ. 118: 357-366
104

References
Bollag J. M. 1992. Decontaminating soil with enzymes. Environ. Sci. Technol. 26:
1876-1881
Bouchez, M., D. Blanchet and J. P. Vandecasteele. 1996. The microbiological fate of
polycyclic aromatic hydrocarbons: carbon and oxygen balances for bacterial
degradation of model compounds. Appl. Microbiol. Biotech. 45(4): 556-561
Brezna, B., A. A. Khan and C. E. Cerniglia. 2003. Molecular characterization of
dioxygenases from polycyclic aromatic hydrocarbon-degrading Mycobacterium spp.
FEMS Microbiol. Lett. 223: 177-183
Brezna, B., O. Kweon, R. L. Stingley, J. P. Freeman, A. A. Khan, B. Polek, R. C.
Jones and C. E. Cerniglia. 2006. Molecular characterization of cytochrome P450
genes in the polycyclic aromatic hydrocarbon degrading Mycobacterium vanbaalenii
PYR-1. Appl. Microbiol. Biotechnol. 71: 522-532
Brion, D. and E. Pelletier. 2005. Modelling PAHs adsorption and sequestration in
freshwater and marine sediments. Chemosphere. 61: 867-876
Brown, J. R. and J. L. Thornton. 1957. Percivall Pott (1714-1788) and chimney
sweepers' cancer of the scrotum. Br. J. Ind. Med. 14(1): 68-70
Bubinas, A., G. Giedraityte, L. Kalediene, O. Nivinskiene and R. Butkiene. 2008.
Degradation of naphthalene by thermophilic bacteria via a pathway, through
protocatechuic acid. Cent. Eur. J. Biol. 3(1): 61-68
Bulay, O. M. and L. W. Wittenberg. 1971. Carcinogenic effects of polycyclic
hydrocarbon carcinogens administrated to mice during pregnancy on the progeny. J.
Natl. Cancer Inst. 46: 397-402
Bury, S. J. 1993. Effect of micellar solubilization on biodegradation rates of
hydrocarbons. Environ.Sci. Technol. 27: 104-110
Camargo, M. C. R. and M. C. F. Toledo. 2003. Polycyclic aromatic hydrocarbons in
Brazilian vegetables and fruits. Food Control. 14(1): 49-53
105

References
Canas, A. I., A. M. lcalde, F. Plou, M. J. Martinez, A. T. Martinez and S. Camarero.
2007. Transformation of polycyclic aromatic hydrocarbons by laccase is strongly
enhanced by phenolic compounds present in soil. Environ. Sci. Technol. 41: 29642971
Cenci, G. and G. Caldini. 1997. Catechol dioxygenase expression in a Pseudomonas
fluorescens strain exposed to different aromatic compounds. Appl. Microbiol.
Biotechnol. 47(3): 306-308
Cerniglia, C. E. 1992. Biodegradation of polycyclic aromatic hydrocarbons.
Biodegradation. 3: 351-368
Cerniglia, C. E. and M. A. Heitkamp. 1989. Microbial degradation of polycyclic
aromatic hydrcarbons (PAH) in the aquatic environment. pp.41- 68. In: U. Varanasi
(ed). Metabolism of Polycyclic Aromatic Hydrocarbons in the Aquatic Environment.
CRC Press. Boca Raton, Florida, USA.
Cerniglia, C. E., J. P. Freeman and F. E. Evans. 1984. Evidence for an arene oxideNIH shift pathway in the transformation of naphthalene to 1-naphthol by Bacillus
cereus. Arch. Microbiol. 138: 283-286
Cerniglia, C. E., J. P. Freeman, G. L. White, R. H. Heflich and D. W. Miller. 1985.
Fungal metabolism and detoxification of the nitropolycyclic aromatic hydrocarbon 1nitropyrene. Appl. Environ. Microbiol. 50(3): 649-655
Cerniglia, C. E. 1997. Fungal metabolism of polycyclic aromatic hydrocarbons: past,
present and future applications in bioremediation. J. Indus. Microbiol. Biotechnol. 19:
324-333
Cerniglia, C. E. 1984. Microbial metabolism of polycyclic aromatic hydrocarbons.
Adv. Appl. Microbiol. 30: 31-71
Chang, B. V., S. H. Wei and S. Y. Yuan. 2001. Biodegradation of phenanthrene in
soil. J. Environ. Sci. Health B. 36(2): 177-187

106

References
Chen, S. H. and M. D. Aitken. 1999. Salicylate Stimulates the Biodegradation of
High-molecular Weight Polycyclic Aromatic Hydrocarbons by Pseudomonas
saccharophila P15. Environ. Sci. Technol. 33: 435-439
Chen, S. Y., J. R. Lin, R. Darbha, P. Lin, T. Y. Liu and Y. M. Chen. 2004. Glycine Nmethyltransferase tumor susceptibility gene in the benzo[a]pyrene detoxification
pathway. Cancer Res. 64: 3617-3623
Christofi, N. and I. B. Ivshina. 2002. Microbial surfactants and their use in field
studies of soil remediation. J. Appl. Microbiol. 93: 915-929
Chulalaksananukul, S., G. M. Gadd, P. Sangvanich, P. Sihanonth, J. Piapukiew and A.
S. Vangna. 2006. Biodegradation of benzo(a)pyrene by a newly isolated Fusarium sp.
FEMS Microbiol. Lett. 262: 99-106
Chun, H. K., Y. Ohnishi, N. Misawa. K. Shindo, M. Hayashi, S. Harayama and S.
Horinouchi. 2001. Biotransformation of phenanthrene and 1-methoxynaphthalene
with Stretomyces linidans cells expressing a marine bacterial phenanthrene
dioxygenase genes. Biosci. Biotechnol. Biochem. 65(8): 1774-1781
Chung, N. J., J. Y. Cho, S. W. Park, S. A. Hwang and T. L. Park. 2008. Polycyclic
aromatic hydrocarbons in soils and crops after irrigation of wastewater discharged
from domestic sewage treatment plants. Bull. Environ. Contam. Toxicol. 81: 124-127
Chupungars, K., P. Rerngsamran and S. Thaniyavarn. 2009. Polycyclic aromatic
hydrocarbons degradation by Agrocybe sp. CU-43 and its fluorene transformation. Int.
Biodeter. Biodegrad. 63: 93-99
Churchill, P. F., A. C. Morgan and E. Kitchens. 2008. Characterization of a pyrenedegrading Mycobacterium sp. strain CH-2. J. Environ. Sci. Health B. 43(8): 698-706.
Churchill, S. A., J. P. Harper and P. F. Churchill. 1999. Isolation and characterization
of a Mycobacterium species capable of degrading three- and four-ring aromatic and
aliphatic hydrocarbons. Appl. Environ. Microbiol. 65(2): 549-552

107

References
Ciemniak, A. and L. Chrachol. 2008. Polycyclic aromatic hydrocarbons (PAH) in
cereal breakfast products. Rocz Panstw Zakl Hig. 59(3): 301-307
Cofield, N., M. K. Banks and A. P. Schwab. 2007. Evaluation of hydrophobicity in
PAH-contaminated soils during phytoremediation. Environ. Pollut. 145: 60-67
Coral, G and S. Karagoz. 2005. Isolation and characterization of phenanthrene
degrading bacteria from petroleum refinery oil. Ann. Microbiol. 55(4): 255-259
Cox, C. D., D. E. Nivens, S. A. Ripp, M. M. Wong, A. Palumbo, R. S. Burlage and G.
S.

Sayler.

2000.

An

intermediate-scale

lysimeter

facility

for

subsurface

bioremediation research. Bioremed. J. 4: 69-79


Cunningham, C. J. and J. C. Philp. 2000. Comparison of bioaugmentation and
biostimulation in ex situ treatment of diesel contaminated soil.

Land Contam.

Reclam. 8 (4): 261-269


Daane, L. L., I. Harjono, G. J. Zylstra, and M. M. Haggblom. 2001. Isolation and
characterization of polycyclic aromatic hydrocarbon-degrading bacteria associated
with the rhizosphere of salt marsh plants. Appl. Environ. Microbiol. 67(6): 2683-2691
Dabrowska, D., A. K. Wasik and J. Namiesnik. 2005. Pathways and Analytical Tools
in Degradation Studies of Organic Pollutants. Crit. Rev. Analyt. Chem. 35(2): 155-176
Das, K. and A. K. Mukherjee. 2007. Crude petroleum-oil biodegradation efficiency of
Bacillus subtilis and Pseudomonas aeruginosa strains isolated from a petroleum-oil
contaminated soil from North-East India. Biores. Technol. 98: 1339-1345
De, J., N. Ramaiah and A. Sarkar. 2006. Aerobic degradation of highly chlorinated
polychlorobiphenyls by a marine bacterium, Pseudomonas CH07. World J. Microbiol.
Biotech. 22(12): 1321-1327
Dean-Ross, D. and C. E. Cerniglia. 1996. Degradation of pyrene by Mycobacterium
flavescens. Appl. Microbiol. Biotechnol. 46: 307-312

108

References
Dean-Ross, D., J. D. Moody, J. P. Freeman, D. R. Doerge and C. E. Cerniglia. 2001.
Metabolism of anthracene by a Rhodococcus species. FEMS Microbiol. Lett. 204(1):
205-211
Debruyn, J. M., C. S. Chewning and G. S. Sayler. 2007. Comparative quantitative
prevalence of Mycobacteria and functionally abundant nidA, nahAc, and nagAc
dioxygenase genes in coal tar contaminated sediments. Environ. Sci. Technol. 41(15):
5426-5432
Derz, K., U. Klinner, I. Schuphan, E. Stackebrandt and R. M. Kroppenstedt. 2004.
Mycobacterium pyrenivorans sp. nov., a novel polycyclic-aromatic-hydrocarbondegrading species. Int. J. Syst. Evol. Microbiol. 54: 2313-2317
DouAbul, A. A. Z., H. M. A. Heba and K. H. Fareed. 1997. Polynuclear aromatic
hydrocarbons (PAHs) in fish from the Red Sea Coast of Yemen. Hydrobiologia. 352:
251-262
Feijoo-Siota, L., F. Rosa-Dos-Santos, T. de Miguel and T. G. Villa. 2008.
Biodegradation of Naphthalene by Pseudomonas stutzeri in marine environments:
Testing cells entrapment in calcium alginate for use in water detoxification.
Bioremediation Journal. 12(4): 185-192
Fleming, J. T., J. Sanseverino and G. S. Sayler. 1993. Quantitative relationship
between naphthalene catabolic gene frequency and expression in predicting PAH
degradation in soils at town gas manufacturing sites. Environ. Sci. Technol. 27: 10681074
Freeman, D. J. and F. C. R. Cattell. 1990. Wood burning as a source of atmospheric
polycyclic aromatic hydrocarbons. Environ. Sci. Technol. 24: 1581-1585
Gennaro, P. D., A. Franzetti, G. Bestetti, M. Lasagni, D. Pitea and E. Collina. 2008.
Slurry phase bioremediation of PAHs in industrial land fill samples at laboratory
scale. Waste Manage. 28: 13381345

109

References
Georgiou, G., S. Lin and M. M. Sharma. 1992. Surface-active compounds from
microorganisms. Biotechnology. 10: 60-65
Gianfreda, L. and M. A. Rao. 2004. Potential of extracellular enzymes in remediation
of polluted soils: a review. Enzyme Microb. Technol. 35: 339
Gibson, D.T. and R. E. Parales. 2000. Aromatic hydrocarbon dioxygenases in
environmental biotechnology. Curr. Opin. Biotechnol. 11: 236-243
Goldman, R., L. Enewold, E. Pellizzari, J. B. Beach, E. D. Bowman, S. S. Krishnan
and P. G. Shields. 2001. Smoking increase carcinogenic polycyclic aromatic
hydrocarbons in human lung tissue. Cancer Res. 61: 63676371
Grevenynghe, J. V., L. Sparfel, M. L. Vee, D. Gilot, B. Drenou, R. Fauchet and O.
Fardel. 2004. Cytochrome P450-dependent toxicity of environmental polycyclic
aromatic hydrocarbons towards human macrophages. Biochem. Biophys. Res. Comm.
317: 708-716
Grifoll, M., S. A. Selifanov and P. J. Chapman. 1994. Evidence for a novel pathway
in the degradation of fluorine by Pseudomonas sp. strain 2065. Appl. Environ.
Microbiol. 60: 2438-2449
Grund, E., B. Denecke and R. Eichenlaub. 1992. Naphthalene degradation via
salicylate and gentisate by Rhodococcus sp. strain B4. Appl. Environ. Microbiol.
58:1874-1877
Guerin, T. F. 1999. Bioremediation of phenols and polycyclic aromatic hydrocarbons
in creosote contaminated soil using ex-situ land treatment. J. Hazard. Mat. B. 65:
305315
Guo, C., W. Sun, J. B. Harsh and A. Ogram. 1997. Hybridization analysis of
microbial DNAfrom fuel oil-contaminated and noncontaminated soil. Microb. Ecol.
34(3): 178-187
Habe, H. O., A. Kouzuma, H. Nojiri, H. Yamane and T. Omori. 2004b. Isolation and
characterization of genes encoding polycyclic aromatic hydrocarbon dioxygenase
110

References
from acenaphthene and acenaphthylene degrading Sphingomonas sp. strain A4. FEMS
Microbiol. Lett. 238(2): 297-305
Habe, H., M. Kanemitsu, M. Nomura, T. Takemura, K. Iwata, H. Nojiri, H. Yamane
and T. Omori. 2004a. Isolation and characterization of an alkaliphilic bacterium
utilizing pyrene as a carbon source. J. Biosci. Bioeng. 98(4): 306-308
Hadibarata, T. 2009. Oxidative degradation of benzo[a]pyrene by the ligninolytic
fungi. Interdisciplinary studies on environmental chemistry. In: Environmental
Research in Asia, Eds., Y. Obayashi, T. Isobe, A. Subramanian, S. Suzuki and S.
Tanabe, pp. 309316.
Hadibarata, T. and S. Tachibana. 2009. Identification of phenanthrene metabolites
produced by Polyporus sp. S133. Interdisciplinary studies on environmental
chemistry. In: Environmental research in Asia, Eds., Y. Obayashi, T. Isobe, A.
Subramanian, S. Suzuki and S. Tanabe, pp. 293299.
Hammel, K. E., W. Z. Gai, B. Green and M. A. Moen. 1992. Oxidative degradation of
phenanthrene by the ligninolytic fungus Phanerochaete chrysosporium. Appl.
Environ. Microbiol. 58:1832-1838
Harayama, S. and M. Rekik. 1989. Bacterial aromatic ring-cleavage enzymes are
classified into two different gene families. J. Biol. Chem. 264: 15328-15333
Harayama, S., M. Kok and E. L. Neidle. 1992. Functional and evolutionary
relationships among diverse oxygenases. Ann. Rev. Microbiol. 46: 565-601
Harvey, R. G. 1996. Mechanisms of carcinogenesis of polycyclic aromatic
hydrocarbons. Polycyc. Arom. Comp. 9: 1-23
Harvey, R. G. 1997. Polycyclic Aromatic Hydrocarbons. Wiley-VCH, Inc., USA.
Heitkamp, M. A. and C. E. Cerniglia. 1989. Polycyclic aromatic hydrocarbon
degradation by a Mycobacterium sp. in microcosms containing sediment and water
from a pristine ecosystem. Appl. Environ. Microbiol. 55(8): 1968-1973
111

References
Heitkamp, M. A., J. P. Freeman, D. W. Miller and C. E. Cerniglia. 1988b. Pyrene
degradation by a Mycobacterium sp.; identification of ring oxidation and ring fission
products. Appl. Environ. Microbiol. 54(10): 25562565
Heitkamp, M. A., W. Franklin and C. E. Cerniglia. 1988a. Microbial metabolism of
polycyclic aromatic hydrocarbons: isolation and characterization of a pyrenedegrading bacterium. Appl. Environ. Microbiol. 54(10): 2549-2555
Henner, P., M. Schiavon, V. Druelle and E. Lichtfouse. 1999. Phytotoxicity of ancient
gaswork soils. Effect of polycyclic aromatic hydrocarbons (PAHs) on plant
germination. Org. Geochem. 30: 963-969
Hintner, J. P., C. Lechner, U. Riegert, A. E. Kuhm, T. Storm, T. Reemtsma and A.
Stolz. 2001. Direct ring fission of salicylate by a salicylate 1,2-dioxygenase activity
from Pseudaminobacter salicylatoxidans. J. Bacteriol. 183: 6936-6942
Hunter, S., S. Myers, P. Radmacher and C. Eno. 2010. Detection of polycyclic
aromatic hydrocarbons (PAHs) in human breast milk. Polycyc. Arom. Comp. 30: 153164
Husain, S. 2008. Identification of Pyrene-degradation pathways: Bench-scale studies
using Pseudomonas fluorescens 29L. Remediation. 18(3): 119-142
Husain, S. 2008. Literature Overview: Microbial metabolism of high molecular
weight polycyclic aromatic hydrocarbons. Remediation. 18(2): 131-161
Hutchins, S. R. 1991. Biodegradation of monoaromatic hydrocarbons by aquifer
microorganisms using oxygen, nitrate, or nitrous oxide as the terminal electron
acceptor. Appl. Environ. Microbiol. 57: 2403-2407
Hwang H-M., J. A. Loya, D. L. Perry and R. Scholze. 1994. Interactions between
subsurface microbial assemblages and mixed organic and inorganic contaminant
systems. Bull. Environ. Contam. Toxicol. 53: 771-778
Jacques, J. S., B. C. Okeke, F. M. Bento, M. C. R. Peralba and F. A. O. Camargo.
2009. Improved enrichment and isolation of polycyclic aromatic hydrocarbons
112

References
(PAH)-degrading microorganisms in soil using anthracene as a model PAH. Curr.
Microbiol. 58: 628-634
Jamroz, T., S. Ledakowicz, J. S. Miller and B. Sencio. 2003. Microbiological
evaluation of toxicity of three polycyclic aromatic hydrocarbons and their
decomposition products formed by advanced oxidation processes. Environ. Toxicol.
18(3): 187-191
Rolling, W. F., M. G. Milner, D. M. Jones, F. Daniel, R. J. Swannel and I. M. Head.
2002. Robust hydrocarbon degradation and dynamics of bacterial communities during
nutrient-enhanced oil spill bioremediation. Appl. Environ. Microbiol. 68(11): 55375548.
Janjua, N. Z., P. M. Kasi, H. Nawaz, S. Z. Farooqui, U. B. Khuwaja, N. Hassan, S. N.
Jafri, S. A. Lutfi, M. M. Kadir and N. Sathiakumar. 2006. Acute health effects of the
Tasman Spirit oil spill on residents of Karachi, Pakistan. BMC Public Health. 6: 84
http://www.biomedcentral.com/1471-2458/6/84
Jerina, D. M. 1983. Metabolism of aromatic hydrocarbons by the cytochrome P450
system and epoxide hydrolase. Drug Metab. Dispos. 11: 1-4
Juhasz, A. L., G. A. Stanley and M. L. Britz. 2000. Degradation of high molecular
weight PAHs in contaminated soil by a bacterial consortium: Effects on microtox and
mutagenicity bioassays. Bioremediation. 4(4): 271-283
Juhasz, A. L., G. A. Stanley and M. L. Britz. 2000. Microbial degradation and
detoxification of high molecular weight polycyclic aromatic hydrocarbons by
Stenotrophomonas maltophilia strain VUN 10,003. Lett. Appl. Microbiol. 30(5): 396401
Kageyama, A., Y. Takahashi, Y. Matsuo, H. Kasai, Y. Shizuri and S. Omura. 2007.
Microbacterium sediminicola sp. nov. and Microbacterium marinilacus sp. nov.,
isolated from marine environments. Int. J. Syst. Evol. Microbiol. 57: 2355-2359

113

References
Kanaly, R. A., R. Bartha, K. Watanabe and S. Harayama. 2002a. Rapid
Mineralization of Benzo[a]pyrene by a Microbial Consortium Growing on Diesel
Fuel. Appl. Environ. Microbiol. 66(10): 4205-4211
Kanaly, R., S. Harayama and K. Watanabe. 2002b. Rhodanobacter sp. strain BPC1 in
a benzo[a]pyrene-mineralizing bacterial consortium. Appl. Environ. Microbiol. 68:
5826-5833
Karimi-Loftabad, S., M. A. Pickard and M. R. Gray. 1996. Reactions of polynuclear
aromatic hydrocarbons on soil. Environ. Sci. Technol. 30: 1145-1151.
Kasai, Y., H. Kishira and S. Harayama. 2002. Bacteria belonging to the genus
Cycloclasticus play a primary role in the degradation of aromatic hydrocarbons
released in a marine environment. Appl. Environ. Microbiol. 68(11): 5625-5633
Kasuga, K., H. Habe, J. Chung, T. Yoshida, H. Nojiri, H. Yamane and T. Omori.
2001. Isolation and characterization of the genes encoding a novel oxygenase
component of angular dioxygenase from the gram-positive dibenzofuran-degrader
Terrabacter sp. strain DBF63. Biochem. Bioph. Res. Commun. 283(1): 195-204
Katagiri, M. A. K., K. Tamura, S. Yamamoto, M. Matsumoto, Y. F. Li, S. R. Cao, R.
D. Ji and C. K. Liang. 1996. Indoor and outdoor air pollution in Tokyo and Beijing
supercities. Atm. Environ. 30: 695-702
Katagiri, M., S. Takemori, K. Suzuki and H. Yasuda. 1966. Mechanism of the
salicylate hydroxylase reaction. J. Bio. Chem. 241(23): 5675-5677
Kauppi, B., K. Lee, E. Carredano, R. E. Parales, D. T. Gibson, H. Eklund and S.
Ramaswamy. 1998. Structure of an aromatic-ring-hydroxylating dioxygenasenaphthalene 1,2-dioxygenase. Structure. 6 (5): 571-586
Keith, L. H. and W. A. Telliard. 1979. Priority pollutants - a perspective view.
Environ. Sci. Technol. 13(4): 416-423
Kelley, L., J. P. Freeman and Cerniglia C. E. 1990. Identification of metabolites from
degradation of naphthalene by a Mycobacterium sp. Biodegradation. 1: 283-290
114

References
Khan, A. A., R. F. Wang, W. W. Cao, D. R. Doerge, D. Wennerstrom and C. E.
Cerniglia. 2001. Molecular cloning, nucleotide sequence, and expression of genes
encoding a polycyclic aromatic ring dioxygenase from Mycobacterium sp. strain
PYR-1. Appl. Environ. Microbiol. 67(8): 3577-3585
Khan, A., M. Ishaq and M. A. Khan. 2008. Effect of vehicle exhaust on the quantity
of polycyclic aromatic hydrocarbons (PAHs) in soil. Environ. Monit. Assess. 137:
363-369
Khan, S. A., M. Hamayun and S. Ahmed. 2006. Degradation of 4-aminophenol by
newly isolated Pseudomonas sp. strain ST-4. Enz. Microb. Technol. 38:10-13
Khurshid, R., M. A. Sheikh and S. Iqbal. 2008. Health of people working/living in the
vicinity of an oil-polluted beach near Karachi, Pakistan. East. Mediter. Health J.
14(1): 179-182
Kim, S. H., S. M. Kang, K. H. Oh, S. I. Kim, B. J. Yoon and H. Y. Kahng. 2005.
Characterization of PAH-degrading bacteria from soils of the reed rhizosphere of
Sunchon Bay using PAH consortia. Kor. J. Microbiol. 41: 208-215
Kim, S. J., O. Kweon, J. P. Freeman, R. C. Jones, M. D. Adjei, J. W. Jhoo, R. D.
Edmondson and C. E. Cerniglia. 2006. Molecular cloning and expression of genes
encoding a novel dioxygenase involved in low and high-molecular-weight polycyclic
aromatic hydrocarbon degradation in Mycobacterium vanbaalenii PYR-1. Appl.
Environ. Microbiol. 72(2): 1045-1054
Kim, S. J., R. C. Jones, C. J. Cha, O. Kweon, R. D. Edmondson and C. E. Cerniglia.
2004. Identification of proteins induced by polycyclic aromatic hydrocarbon in
Mycobacterium vanbaalenii PYR-1 using two-dimensional polyacrylamide gel
electrophoresis and de novo sequencing methods. Proteomics. 4: 3899-3908
King, J.M.H., P. M. DiGrazia, B. Applegate, R. Burlage, J. Sanseverino, P. Dunbar, F.
Larimer and G. S. Sayler. 1990. Rapid, sensitive bioluminescent reporter technology
for naphthalene exposure and biodegradation. Science. 249: 778-781

115

References
Kirk, T. K. and R. L. Farrell. 1987. Enzymatic combustion: the microbial
degradation of lignin. Annu. Rev. Microbiol. 41: 465-505
Kishikawa N., M. Wada, N. Kuroda, S. Akiyama and K. Nakashima. 2003.
Determination of polycyclic aromatic hydrocarbons in milk samples by highperformance liquid chromatography with fluorescence detection. J. Chromatog. B.
789: 257-264
Kita, A., S. Kita, I. Fujisawa, K. Inaka, T. Ishida, K. Horiike, M. Nozaki and K. Miki.
1999. An archetypical extradiol-cleaving catecholic dioxygenase: the crystal structure
of catechol 2,3-dioxygenase (metapyrocatechase) from Pseudomonas putida mt-2.
Structure. 7(1): 25-34
Kiyohara, H., S. Torigoe, N. Kaida, T. Asaki, T. Iida, H. Hayashi and N. Takizawa.
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
Klankeo, P., W. Nopcharoenkul and O. Pinyakong. 2009. Two novel pyrenedegrading Diaphorobacter sp. and Pseudoxanthomonas sp. isolated from soil. J.
Biosci. Bioeng. 108(6): 488-495
Kochany, J. and R. J. Maguire. 1994. Abiotic transformation of polynulear aromatic
hydrocarbons

and

polynuclear

aromatic

nitrogen

heterocycles

in

aquatic

environments. Sci. Tot. Environ. 144: 17-31


Kosaric, N. 2001. Biosurfactants and their application for soil bioremediation. Food
Tech. Biotechol. 39: 295-304
Kumar, G., R. Singla and R. Kumar. 2010. Plasmid associated anthracene degradation
by Pseudomonas sp. isolated from filling station site. Nature Sci. 8(4): 89-94
Kumar, M., V. Leon, A. M. De Sisto and O. A. Ilzins. 2006. A halotolerant and
thermotolerant Bacillus sp. degrades hydrocarbons and produces tensioactive
emulsifying agent. World J. Microbiol. Biotech. 23: 211-220
116

References
Kurkela, S., H. Lehvaslaiho, E. T. Palva and T. H. Teeri. 1988. Cloning, nucleotide
sequence and characterization of genes encoding naphthalene dioxygenase of
Pseudomonas putida strain NCIB9816. Gene. 73: 355-362
Lau, K. I., Y. Y. Rsang and S. W. Chiu. 2003. Use of spent mushroom compost to
bioremediate PAH-contaminated samples. Chemosphere. 52: 1539-1546
Laurie, A. D. and G. L. Jones. 1999. The phn genes of Burkholderia sp. strain RP007
constitute a divergent gene cluster for polycyclic aromatic hydrocarbon catabolism. J.
Bacteriol. 181: 531-540
Laurie, A. D. and G. L. Jones. 2000. Quantification of phnAc and nahAc in
contaminated New Zealand soils by competitive PCR. Appl. Environ. Microbiol. 66 :
1814-1817
Leahy, J. G. and R. R. Colwell. 1990. Microbial degradation of hydrocarbons in the
environment. Microbiol .Rev. 54: 305-315
Lehto, K. M., E. Vuorimaa and H. Lemmetyinen. 2000. Photolysis of polycyclic
aromatic hydrocarbons (PAHs) in dilute aqueous solutions detected by fluorescence.
J. Photochem. Photobiol. A. 136: 53-60
Leonardi, V., V. Sasek, M. Petruccioli, A. D. Annibale, P. Erbanova and T. Cajthaml.
2007. Bioavailability modification and fungal biodegradation of PAHs in aged
industrial soils. Int. Biodeter. Biodegrad. 60: 165-170
Leveau, J. H. J., A. J. B. Zehnder and J. R. van der Meer. 1998. The tfdK gene product
facilitates

uptake

of

2,4-dichlorophenoxyacetate

by

Ralstonia

eutropha

JMP134(pJP4). J. Bacteriol. 180: 2237-2243


Li, P., X. Wang, F. Stagnitti, L. Li, Z. Gong, H. Zhang, X. Xiong and C. Austin. 2005.
Degradation of phenanthrene and pyrene in soil slurry reactors with immobilized
bacteria Zoogloea sp. Environ. Eng. Sci. 22(3): 390-399

117

References
Lilia, R. C., L. G. Torres and I. Rosario. 2008. Effect of nutrient and surfactant
addition on polyaromatic hydrocarbon (PAH) biodegradation in contaminated soils.
Land Contam. Reclam. 16(1): 1-11
Lily, M. K., A. Bahuguna, K. Dangwal and V. Garg. 2009. Degradation of benzo[a]
pyrene by a novel strain Bacillus subtilis BMT4i (MTCC 9447). Braz. J. Microbiol.
40: 884-892
Lim, L. H., R. M. Harrison and S. Harrad. 1999. The contribution of traffic to
atmospheric concentrations of polycyclic aromatic hydrocarbons. Environ. Sci.
Technol. 33: 3538-3542
Lin, C., L. Gan and Z. L. Chen. 2010. Biodegradation of naphthalene by strain
Bacillus fusiformis (BFN). J. Hazard. Mat. 182(1-3): 771-777
Lin, Y. and L. X. Cai. 2008. PAH-degrading microbial consortium and its pyrenedegrading plasmids from mangrove sediment samples in Huian, China. Mar. Pollut.
Bull. 57(6-12): 703-706
Liu, H. H., M. H. Lin, C. I. Chan and H. L. Chen. 2010. Oxidative damage in foundry
workers occupationally co-exposed to PAHs and metals. Int. J. Hyg. Environ. Health.
213: 93-98
Liu, K., W. Han, W. P. Pan and J. T. Riley. 2001. Polycyclic aromatic hydrocarbon
(PAH) emissions from a coal fired pilot FBC system. J. Hazard. Mat. 84(2-3): 175188
Loehr, R. C. 1992. Bioremediation of PAH compounds in contaminated soil.
Hydrocarbon contaminated soils and ground water. In: Calabrese EJ, Kostechi PT
(eds), Proceedings 1st annual West Coast conference on hydrocarbon contaminated
soils ground water, Feb 1990, Newport Beach, CA. Vol. 1. Lewis Publishers, Chelsea
MI, pp 213222
Luan, T. G., K. S. H. Yu, Y. Zhong, H. W. Zhou, C. Y. Lan and N. F. Y. Tam. 2006.
Study of metabolites from the degradation of polycyclic aromatic hydrocarbons
118

References
(PAHs) by bacterial consortium enriched from mangrove sediments. Chemosphere.
65(11): 2289-2296
Mackay, D. 1997. Multimedia mass balance models of chemical distribution and fate,
In: Ecotoxicology edited by G. Schuurmann and B. Markert and published by John
Wiley, NY and Spektrum Press, Berlin.
Maier, M., D. Maier and B. J. Lloyd. 2000. The role of biofilms in the mobilisation of
polycyclic aromatic hydrocarbons (PAHs) from the coal-tar lining of water pipes.
Wat. Sci. Technol. 41(4-5): 279-285
Malik, Z. A. 2009. Biodegradation of crude oil by bacteria. Ph.D thesis, submitted to
Quaid-i-Azam University, Islamabad, Pakistan.
Mallick, S., S. Chatterjee and T. K. Dutta. 2007. A novel degradation pathway in the
assimilation of phenanthrene by Staphylococcus sp. strain PN/Y via meta-cleavage of
2-hydroxy-1-naphthoic acid: formation of trans-2,3-dioxo-5-(29-hydroxyphenyl)pent-4-enoic acid. Microbiology. 153: 2104-2115
Mane, S. S., D. M. Purnell and I. C. Hsu. 1990. Genotoxic effects of five polycyclic
aromatic hydrocarbons in human and rat mammary epithelial cells. Environ. Molec.
Mutag. 15(2): 78-82
Margesin, R., A. Zimmerbauer and F. Schinner. 1999. Soil lipase activity-a useful
indicator of oil biodegradation. Biotechnol. Techn. 13: 859-863
Martens, D., J. Maguhn, P. Spitzauer and A. Kettrup. 1997. Occurrence and
distribution of polycyclic aromatic hydrocarbons (PAHs) in an agricultural ecosystem.
Fresenius J. Anal. Chem. 359: 546-554
Mastrangela, G., E. Fadda and V. Marzia. 1996. Polycyclic aromatic hydrocarbons
and cancer in man. Environ. Health Perspect. 104(11): 1166-1170
Meckenstock, R. U., E. Annweiler, W. Michaelis, H. H. Richnow and B. Schink.
2000. Anaerobic naphthalene degradation by a sulphate reducing enrichment culture.
Appl. Environ. Microbiol. 66: 2743-2747
119

References
Menn, F. M., B. M. Applegate and G. S. Sayler. 1993. NAH plasmid-mediated
catabolism of anthracene and phenanthrene to naphthoic acids. Appl. Environ.
Microbiol. 59(6): 1938-1942
Mester, T. and M. Tien. 2000. Oxidation mechanism of ligninolytic enzymes involved
in the degradation of environmental pollutants. Int. Biodeterior. Biodegrad. 46: 51-59
Molina, M. C., N. Gonzalez., L. F. Bautista., R. Sanz., R. Simarro., I. Sanchez and J.
L. Sanz. 2009. Isolation and genetic identification of PAH degrading bacteria from a
microbial consortium. Biodegradation. 20: 789-800
Moody J. D., P. P. Fu, J. P. Freeman and C. E. Cerniglia. 2004. Regio- and
Stereoselective Metabolism of 7,12-Dimethylbenz[a]anthracene by Mycobacterium
vanbaalenii PYR-1. Appl. Environ. Microbiol. 69(7): 3924-3931
Moody, J. D., J. P. Freeman, D. R. Doerge and C. E. Cerniglia. 2001. Degradation of
phenanthrene and anthracene by cell suspensions of Mycobacterium sp. strain PYR-1.
Appl. Environ. Microbiol. 67: 1476-1483
Mueller, J. G., C. E. Cerniglia and P. H. Pritchard. 1996. Bioremediation of
environments contaminated by polycyclic aromatic hydrocarbons, In Bioremediation:
Principles and Applications, ed by Crawford R. L. and D. L. Crawford. Cambridge
University Press, Idaho, pp 125194
Mueller, J. G., P. J. Chapman and P. H. Pritchard. 1989. Action of a fluorantheneutilizing bacterial community on polycyclic aromatic hydrocarbon components of
creosote. Appl. Environ. Microbiol. 55: 3085-3090
Obayori, O. S., M. O. Ilori, S. A. Adebusoye, G. O. Oyetibo and O. O. Amund. 2008.
Pyrene-degradation potentials of Pseudomonas species isolated from polluted tropical
soils. World J. Microbiol. Biotechnol. 24: 2639-2646
Ohkouchi, N., K. Kawamura and H. Kawahata. 1999. Distributions of three to sevenring polynuclear aromatic hydrocarbons on the deep sea floor in the central pacific.
Environ. Sci. Technol. 33: 3086-3090
120

References
Pandey, B. V. and R.S. Upadhyay. 2010. Pseudomonas fluorescens can be used for
bioremediation of textile effluent Direct Orange-102. Tropical Ecol. 51(2S): 397-403
Papadoyannis, I. N., A. Zatou and V. F. Samanidou. 2002. Development of a solid
phase extraction protocol for the simultaneous determination of anthracene and its
oxidation products in surface waters by reversed-phase HPLC. J. Liquid Chromatog.
Related Technol. 25: 2653-2661
Parales, R. E., K. Lee, S. M. Resnick, H. Jiang, D. J. Lessner and D. T. Gibson. 2000.
Substrate specificity of naphthalene dioxygenase: effect of specific amino acids at the
active site of the enzyme. 182(6): 164-1649
Park, J. and D. E. Crowley. 2006. Dynamic changes in nahAc gene copy numbers
during degradation of naphthalene in PAH-contaminated soils. Appl. Microbiol.
Biotechnol. 72: 1322-1329
Paskova, V., K. Hilscherov, M. Feldmannov and L. Blha. 2006. Toxic effects and
oxidative stress in higher plants exposed to polycyclic aromatic hydrocarbons and
their N-heterocyclic derivatives. Environ. Toxicol. Chem. 25(12): 3238-3245
Pathak, H., D. Kantharia, A. Malpani and D. Madamwar. 2009. Naphthalene
degradation by Pseudomonas sp. HOB1: In vitro studies and assessment of
naphthalene degradation efficiency in simulated microcosms. J. Hazard. Mat. 166:
1466-1473
Patnaik, P. 1992. Hydrocarbon, aromatic. In: Patnaik P (ed) A comprehensive guide
to the hazardous properties of chemical substances. Van Nostrand Reinhold, New
York, NY, pp 425445
Peng, R. H., A. S. Xiong, Y. Xue, X. Y. Fu, F. Gao, W. Zhao, Y. S. Tian and Q. H.
Yao. 2008. Microbial biodegradation of polyaromatic hydrocarbons. FEMS
Microbiol. Rev. 32: 927-955
Perera, F., D. Tang, R. Whyatt, S. A. Lederman and W. Jedrychowski. 2005. DNA
damage from polycyclic aromatic hydrocarbons measured by benzo[a]pyrene-DNA
121

References
adducts in mothers and newborns from Northern Manhattan, The World Trade Center
Area, Poland, and China. Cancer Epidemiol. Biomarkers Prev. 14(3): 709-714
Perez-Perez, F. J. and N. D. Hanson. 2002. Detection of plasmid-mediated AmpC lactamase genes in clinical isolates by using multiplex PCR. J. Clin. Microbiol. 40(6):
2153-2162
Peterson, C. H. 2001. The Exxon Valdez oil spill in Alaska: Acute, indirect and
chronic effects on the ecosystem. Adv. Mar. Bio. 39: 1-103
Phale, P. S., A. Basu, P. D. Majhi, J. Deveryshetty, C. Vamsee-Krishna and R.
Shrivastava. 2007. Metabolic diversity in bacterial degradation of aromatic
compounds. OMICS. 11: 252-279
Pinyakong, O., H. Habe and T. Omori. 2003a. The unique aromatic catabolic genes in
sphingomonads degrading polycyclic aromatic hydrocarbons (PAHs). J. Gen. Appl.
Microbiol. 49: 119
Pinyakong, O., H. Habe, A. Kouzuma, H. Nojiri, H. Yamane and T. Omori. 2004.
Isolation and characterization of genes encoding polycyclic aromatic hydrocarbon
dioxygenase from acenaphthene and acenaphthylene degrading Sphingomonas sp.
strain A4. FEMS Microbiol. Lett. 238: 297-305
Pinyakong, O., H. Habe, T.Yoshida, H. Nojiri and T. Omori. 2003b. Identification of
three isofunctional novel salicylate 1-hydroxylases involved in the phenanthrene
degradation of Sphingobium sp. strain P2. Biochem. Biophys. Res. Commun. 201:
350-357
Pliskova, M., J. Vondracek, B. Vojtesek, A. Kozubik and M. Machala. 2005.
Deregulation of cell proliferation by polycyclic aromatic hydrocarbons in human
breast carcinoma MCF-7 cells reflects both genotoxic and nongenotoxic events.
Toxicol. Sci. 83(2): 246-256
Pointet, K. and A. Milliet. 2000. PAHs analysis of fish whole gall bladders and livers
from the Natural Reserve of Camargue by GC/MS. Chemosphere. 40: 293-299
122

References
Reddy, M. S., B. Naresh, T. Leela, M. Prashanthi, N. C. Madhusudhan, G. Dhanasri
and P. Devi. 2010. Biodegradation of phenanthrene with biosurfactant production by a
new strain of Brevibacillus sp. Bioresour. Technol. 101(20): 7980-7983
Rehmann, K., H. P. Noll, C. E. W. Steinberg and A. A. Kettrup. 1998. Pyrene
degradation by Mycobacterium sp. strain KR2. Chemosphere. 36: 2977-2992
Reinik, M., T. Tamme, M. Roasto, K. Juhkam, T. Tenno and A. Kiis. 2007.
Polycyclic aromatic hydrocarbons (PAHs) in meat products and estimated PAH intake
by children and the general population in Estonia. Food Addit. Contam. 24(4): 429437
Resnick, S. M., K. H. Lee and D. T. Gibson. 1996. Diverse reactions catalyzed by
naphthalene dioxygenase by Pseudomonas sp. strain NCCIB9816. J. Ind. Microbiol.
Biotechnol. 17: 438-458

Ripp, S., D.E. Nivens, Y. Ahn, C. Werner, J. Jarrel, J. P. Easter, C. D. Cox, R. S.


Burlage and G. S. Sayler. 2000. Controlled field release of a bioluminiscent
genetically engineered microorganisms for bioremediation process, monitoring and
control. Environ. Sci. Technol. 34: 846-853
Robinovich, M. L., A. V. Bolobova and L. G. Vasilchenko. 2004. Fungal
decomposition of natural aromatic structures and xenobiotics: A review. Appl.
Biochem. Microbiol. 40: 1-17
Rockne, K. J. and S. E. Strand. 1998. Biodegradation of bicyclic and polycyclic
aromatic hydrocarbons in anaerobic enrichments. Environ. Sci. Technol. 32: 29622967
Rockne, K. J., J. C. Chee-Sandford, R. Sanford, B. P. Hedlund, J. T. Staley and S. E.
Strand. 2000. Anaerobic naphthalene degradation by microbial pure cultures under
nitrate reducing conditions. Appl. Environ. Microbiol. 66: 1595-1601
Ron, E. Z. and E. Rosenberg. 2001. Natural roles of biosurfactants. Environ.
Microbiol. 3(4): 229-236
123

References
Sabate, J., J. M. Bayona and A. M. Solanas. 2001. Photolysis of PAHs in aqueous
phase by UV irradiation. Chemosphere. 44: 119-124
Said, S., F. M. Neves and A. J. F. Griffiths. 2004. Cinnamic acid inhibits the growth
of the fungus Neurospora crassa, but is eliminated as acetophenone. Int. Biodeter.
Biodegrad. 54: 1-6
Saifullah, S. M. and F. Chaghtai. 2005. Effect of Tasman Spirit oil spill on marine
plants in the coastal area of Karachi. Int. J. Bio. Biotechnol. 2: 299-306
Salinas, R. O., G. D. Gonzalez, B. S. Bermudez, R. G. Tolentino and S. V. Leon.
2010. Presence of polycyclic aromatic hydrocarbons (PAHs) in apple in rural terrains
from Mexico City. Bull. Environ. Contam. Toxicol. 85: 179-183
Samanta, S. K., O. V. Singh and R. K. Jain. 2002. Polycyclic aromatic hydrocarbons:
environmental pollution and bioremediation. Trends Biotechnol. 20(6): 243-248
Sanseverino, J., C. Werner , J. Fleming , B. Applegate , J. M. King and G. S. Sayler.
1993. Molecular diagnostics of polycyclic aromatic hydrocarbon biodegradation in
manufactured gas plant soils. Biodegradation. 4(4): 303-321
Sayler, G. S., M. S. Shields, E. T. Tedford, A. Breen, S. W. Hooper, K. M. Sirotkin
and J. W. Davis. 1985. Application of DNADNA colony hybridization to the
detection of catabolic genotype in environmental samples. Appl. Environ. Microbiol.
49(5): 1295-1303
Schocken, M. J. and D. T. Gibson. 1984. Bacterial oxidation of the polycyclic
aromatic hydrocarbons acenaphthene and acenaphthylene. Appl. Environ. Microbiol.
48(1): 10-16
Schuler, L., S. M. N. Chadhain, Y. Jouanneau, C. Meyer, G. J. Zylstra, P. Hols and S.
N. Agathos. 2008. Characterization of a novel angular dioxygenase from fluorenedegrading Sphingomonas sp. strain LB126. Appl. Environ. Microbiol. 74 (4): 1050
1057

124

References
Schuler, L., Y. Jouanneau, S. M. N. Chadhain, C. Meyer, M. Pouli, G. J. Zylstra, P.
Hols and S. N. Agathos. 2009. Characterization of a ring-hydroxylating dioxygenase
from phenanthrene-degrading Sphingomonas sp. strain LH128 able to oxidize
benz[a]anthracene. Appl. Microbiol. Biotechnol. 83: 465-475
Searle, C.E. 1976. Ed. Chemical Carcinogens, American Chemical Society,
Washington, DC.
Semple, K. T., N. M. Dew, K. J. Doick and A. H. Rhodes. 2006. Can microbial
mineralization be used to estimate microbial availability of organic contaminants in
soil? Environ. Pollut. 140: 164-172
Seo, J. S., Y. S. Keum, R. M. Harada and Q. X. Li. 2007. Isolation and
characterization of bacteria capable of degrading polycyclic aromatic hydrocarbons
(PAHs) and organophosphorus pesticides from PAH-contaminated soil in Hilo,
Hawaii. J. Agric. Food Chem. 55: 5383-5389
Sepic, E., M. Bricelj and H. Leskovsek. 1997. Biodegradation studies of polyaromatic
hydrocarbons in aqueous media. J. Appl. Microbiol. 83: 561-568
Sepic, E., M. Bricelj and H. Leskovsek. 1998. Degradation of fuoranthene by
Pasteurella sp. IFA and Mycobacterium sp. PYR1: isolation and identification of
metabolites. J. Appl. Microbiol. 85: 746-754
Shen, F. and L. Z. Zhu. 2007. Concentration and distribution of PAHs in vegetables
grown near an iron and steel industrial area. Huan Jing Ke Xue. 28(3): 669-672
Shin, K.H., Y. Ahn and K. W. Kim. 2005. Toxic effect of biosurfactant addition on
the biodegradation of phenanthrene. Environ. Toxicol. Chem. 24: 27682774.
Shindo, K., Y. Ohnishi, H. K. Chun, H. Takahashi, M. Hayashi, A.Saito, K. Iguchi, K.
Furukawa, S. Harayama, S. Horinouchi and N. Misawa. 2001. Oxygenation reaction
of variuos tricyclic fused aromatic compounds using E. coli and Stretomyces linidans
transformant carrying several arene dioxygenase genes. Biosci. Biotechnol. Biochem.
65(1): 2472-2481
125

References
Sho, M., C. Hamel and C. W. Greer. 2004. Two distinct gene clusters encode pyrene
degradation in Mycobacterium sp. strain S65. FEMS Microbiol. Eco. 48: 209-220
Shuyu, H., Z. Qingmin, D. Miao, Z. Yang and S. Hongwen. 2007. Optimized
cultivation of highly-efficient degradation bacterial strains and their degradation
ability towards pyrene. Front. Biol. China. 2(4): 387-390
Siddiqi H. A., F. A. Ansari and A. B. Munshi. 2009. Assessment of hydrocarbons
concentration in marine fauna due to Tasman Spirit oil spill along the Clifton beach at
Karachi coast. Environ. Monit. Assess. 148:139-148
Silva, I. S., M. Grossman and L. R. Durrant. 2009. Degradation of polycyclic
aromatic hydrocarbons (27 rings) under microaerobic and very-low-oxygen
conditions by soil fungi. Int. Biodeter. Biodegrad. 63: 224-229
Singh, A. and O. P. Ward. 2004. Biodegradation and Bioremediation: Series: Soil
Biology, vol. 2, Springer-Verlag, New York.
Singh, O. V. and R. K. Jain. 2003. Phytoremediation of toxic aromatic pollutants from
soil. Appl. Microbiol. Biotechnol. 63: 128-135
Siripong, P., B. Oraphin and P. Duanporn. 2009. The ability of five fungal isolates
from nature to degrade of polyaromatic hydrocarbons (PAHs) and polychlorinated
biphenyls (PCBs) in culture media. Aust. J. Basic Appl. Sci. 3(3): 1076-1082
Smith, D. J. T., R. M. Harrison, L. Luhana, C. A. Pio, L. M. Castro, M. N. Tariq, S.
Hayat and T. Quraishi. 1996. Concentration of particulate airborne polycyclic
aromatic hydrocarbons and metals collected in Lahore, Pakistan. Atm. Eniron. 30(23):
4031-4040
Srinivasan, C., T. M. D'Souza, K. Boominathan and C. A. Reddy. 1995.
Demonstration of laccase in the white rot basidiomycete Phanerochaete
chrysosporium BKM-F1767. Appl. Environ. Microbiol. 61: 4274-4277
Stenberg, U., T. Alsberg and R. Westerholm. 1983. Emission of carcinogenic
components with automobile exhausts. Environ. Health Perspect. 47: 53-63
126

References
Stingley, R. L., A. A. Khan and C. E. Cerniglia. 2004b. Molecular characterization of
a phenanthrene degradation pathway in Mycobacterium vanbaalenii PYR-1. Biochem.
Biophys. Res. Commun. 322: 133-146
Stingley, R. L., B. Brezna, A. A. Khan and C. E. Cerniglia. 2004a. Novel organization
of genes in a phthalate degradation operon of Mycobacterium vanbaalenii PYR-1.
Microbiology. 150: 3749-3761
Stringfellow, W. T. and M. D. Aitken. 1995. Competitive metabolism of naphthalene,
methylnaphthalenes and fluorene by phenanthrene-degrading Pseudomonads. Appl.
Environ. Microbiol. 61(1): 357-362
Su, D., P. J. Li, X. Wang and H. X. Xu. 2007. Degradation and kinetics of pyrene and
benzo(a)pyrene by three bacteria in contaminated soil. Huan Jing Ke Xue. 28(4): 913917
Sundberg, K., M. Widersten, A. Seidel, B. Mannervik and B. Jernstom. 1997.
Glutathione conjugation of bay- and fjord-region diol epoxides of polycyclic aromatic
hydrocarbons by glutathione tramsferases M1-1 and P1-1. Chem. Res. Toxicol. 10:
1221-227
Sutherland, J. B., F. Rafii, A. A. Khan and C. E. Cerniglia. 1995. Mechanisms of
polycyclic aromatic hydrocarbon degradation, In: Microbial Transformation and
Degradation of Toxic Organic Chemicals, ed by Young L. Y. and C. E. Cerniglia.
Wiley-Liss, New York, pp 269306
Sutherland, J. B., J. P. Freeman, A. L. Selby, P. P. Fu , D. W. Miller and C. E.
Cerniglia. 1990. Stereoselective formation of a K-region dihydrodiol from
phenanthrene by Streptomyces flavovirens. Arch. Microbiol. 154(3): 260-266
Tam, N. F. Y., C. L. Guo, W. Y. Yau and Y. S. Wong. 2002. Preliminary study on
biodegradation of phenanthrene by bacteria isolated from mangrove sediments in
Hong Kong. Mar. Pollut. Bull. 45: 316-324

127

References
Tang, X., Y. Zhu and Q. Meng. 2007. Enhanced crude oil biodegradability of
Pseudomonas aeruginosa ZJU after preservation in crude oil-containing medium.
World J. Microbiol. Biotechn. 23(1): 7-14
Tao, X. Q., G. N. Lu, Z. Dang. X. Y. Yi and C. Yang. 2007. Isolation of
phenanthrene-degrading bacteria and characterization of phenanthrene metabolites.
World J. Microbiol. Biotechnol. 23: 647-654
Taylor, L. T. and D. M. Jones. 2001, Bioremediation of coal tar PAH in soils using
biodiesel. Chemosphere. 44: 1131-1136
Teng, Y., Y. Luo, M. Sun, Z. Liu, Z. Li and P. Christie. 2010. Effect of
bioaugmentation by Paracoccus sp. strain HPD-2 on the soil microbial community
and removal of polycyclic aromatic hydrocarbons from an aged contaminated soil.
Biores. Technol. 101: 34373443
Thompson, J. D., T. J. Gibson and D. G. Higgins. 2003. Multiple sequence alignment
using ClustalW and ClustalX. Curr. Protocols Bioinfor. 2.3.1-2.3.22
Tiana L., P. Maa and J. Zhong. 2003. Impact of the presence of salicylate or glucose
on enzyme activity and phenanthrene degradation by Pseudomonas mendocina. Proc.
Biochem. 38(8): 1125-1132
Tiana, L., P. Ma and J. Zhong. 2002. Kinetics and key enzyme activities of
phenanthrene degradation by Pseudomonas mendocina. Process Biochem. 37: 14311437
Ting, Y., P. Zhou, Y. Zhou, L. Heng-Li and S. Liu. 2004. Saccharothrix xinjiangensis
sp. nov., a pyrene degrading actinomycetes isolated from Tianchi Lake, Xinjhiang,
China. Int. J. Syst. Evol. Microbiol. 54, 20912094.
Tongpim, S. and M. A. Pickard. 1999. Cometabolic oxidation of phenanthrene to
phenanthrene trans-9,10-dihydrodiol by Mycobacterium strain S1 growing on
anthracene in the presence of phenanthrene. Can. J. Microbiol. 45: 369-376

128

References
Treadway, S. L., K. S. Yanagimachi, E. Lankenau, P. A. Lessard, G. Stephanopoulos
and A. J. Sinskey. 1999. Isolation and characterization of indene bioconversion genes
from Rhodococcus strain I24. Appl. Microbiol. Biotechnol. 51: 786-793
Tuomi, P. M., J. M. Salminen and K. S. Jorgensen. 2004. The abundance of nahAc
genes correlates with the 14C-naphthalene mineralization potential in petroleum
hydrocarbon-contaminated oxic soil layers. FEMS Microbiol. Eco. 51: 99-107
United States Coast Guard. 1993. Federal on scene coordinators report: T/V Exxon
Valdez oil spill (Two volumes). U.S. Department of Transportation. Washington,
D.C.
Van, J. A., W. A. J. Van-Pul and F. A. De Leeuw. 1997. Modelling transport and
deposition of persistent organic pollutants in the European region. Atm. Environ.
31(7): 1011-1024
Vila, J., Z. Lopez, J. Sabate, C. Minguillon, A. M. Solanas and M. Grifoll. 2001.
Identification of a novel metabolite in the degradation of pyrene by Mycobacterium
sp. strain AP1: Actions of the isolate on two- and three-ring polycyclic aromatic
hydrocarbons. Appl. Environ. Microbiol. 67(12): 5497-5505
Vilanova, R. M., P. Fernandez, C. Martinaz and J. O. Grimalt. 2001. Polycyclic
aromatic hydrocarbons in remote mountain lake waters. Wat. Res. 35(16): 3916-3926
Villemain, D., P. Guiraud, O. Bordjiba and R. Steiman. 2006. Biotransformation of
anthracene and fluoranthene by Absidia fusca Linnemann. Electron. J. Biotechnol.
9(2): http://www.ejbiotechnology.info/content/vol9/issue2/full/10/
Wan, C. K., J. W. C. Wong, M. Fang and D. Y. Ye. 2003. Effect of organic waste
amendments on degradation of PAHs in soil using thermophillic composting.
Environ. Technol. 24(1): 23-30
Wang, X., P. Li, S. Song, Y. Zhong, H. Zhang and E. V. Verkhozina. 2006.
Immobilization of introduced bacteria and degradation of pyrene and benzo(a)pyrene
in soil by immobilized bacteria. Ying Yong Sheng Tai Xue Bao. 17(11): 2226-2228
129

References
Ward, O. P., A. Singh and J. V. Hamme. 2003.

Accelerated biodegradation of

petroleum hydrocarbon waste. J. Ind. Microbiol. Biotechnol. 30: 260-270


Watanabe, K. 2001. Microorganisms relevant to bioremediation. Curr. Opin.
Biotechnol. 12: 237-241
Wick, L. Y., A. R. de Munain, D. Springael and H. Harms. 2002b. Responses of
Mycobacterium sp. LB501T to the low bioavailability of solid anthracene. Appl.
Microbiol. Biotechnol. 58: 378-385
Wick, L. Y., P. Wattiau and H. Harms. 2002a. Influence of the growth substrate on
the mycolic acid profiles of mycobacteria. Environ. Microbiol. 4(10): 612-616
Widdle, F. and R. Rabus. 2001. Anaerobic biodegradation of saturated and aromatic
hydrocarbons. Curr. Opin. Biotechnol. 12: 259-276
Wilson, L. P. and E. J. Bouwer. 1997. Biodegradation of aromatic compounds under
mixed oxygen/denitrifying conditions: A review. J. Indust. Microbiol. Biotechnol. 18:
116-130
Wison, A. S., C. D. Davis, P. D Williams, R. A. Buckpitt, M. Pirmohamed and B. K.
Park. 1996. Characterization of the toxic metabolites of naphthalene. Toxicology. 114:
233-242
Wong, J. W. C., K. M. Lai, C. K. Wan, K. K. Ma and M. Fang. 2002. Isolation and
optimization of PAH-Degradative bacteria from contaminated soil for PAHs
bioremediation. Wat. Air Soil Poll. 139: 1-13
Wu, Y., Y. Luo, D. Zou, J. Ni, W. Liu, Y. Teng and Z. Li. 2008. Bioremediation of
polycyclic aromatic hydrocarbons contaminated soil with Monilinia sp.: degradation
and microbial community analysis. Biodegradation. 19: 247257
Yan, J., L. Wang, P. P. Fu and H. Yu. 2004. Photomutagenicity of 16 polycyclic
aromatic hydrocarbons from the US EPA priority pollutant list. Mutat. Res. 557(1):
99-108
130

References
Ye, D., M. A. Siddiqi, A. E. Maccubbin, S. Kumar and H. C. Sikka. 1996.
Degradation of polynuclear aromatic hydrocarbons by Sphingomonas paucimobilis.
Environ. Sci. Technol. 30(1): 136-142
Yu, K. S. H., A. H. Y. Wong, K. W. Y. Yau, Y. S. Wong and N. F. Y. Tam. 2005.
Natural attenuation, biostimulation and bioaugmentation on biodegradation of
polycyclic aromatic hydrocarbons (PAHs) in mangrove sediments. Mar. Poll. Bull.
51(8-12): 10711077
Zander, M. 1983. Physical and chemical properties of polycyclic aromatic
hydrocarbons. In: A. Bjorstedth, Editor, Handbook of Polycyclic Aromatic
Hydrocarbon, Marcel Dekker Inc., USA, pp. 1-25.
Zhang, Y. and S. Tao. 2009. Global atmospheric emission inventory of polycyclic
aromatic hydrocarbons (PAHs) for 2004. Atm. Environ. 43(4): 812-819
Zhao, X., L. Fu, J. Hu, J. Li, H. Wang, C. Huang and X. Wang. 2009. Analysis of
PAHs in water and fruit juice samples by DLLME combined with LC-Fluorescence
detection. Chromatographia. Published online 05 May 2009.
Zhong, Y., T. Luan, H. Zhou, C. Lan and N. F. Y. Tam. 2006. Metabolite production
in degradation of pyrene alone or in a mixture with another polycyclic aromatic
hydrocarbon by Mycobacterium Sp. Environ. Toxicol. Chem. 25(11): 2853-2859

131

Appendix

Figure I a. Phylogenetic tree of bacterial isolates

132

Appendix

Percent Identity
1
1

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

28

99.8

99.6

94.0

99.8

99.7

99.8

99.8

93.9

98.9

94.1

99.8

93.6

78.1

99.3

99.2

82.0

94.5

94.8

78.8

78.5

78.8

93.0

92.3

92.2

78.0

84.9

76.6

B. cereus AB215098

98.9

93.5

98.6

98.6

98.6

98.6

92.8

96.6

93.0

99.5

93.6

77.9

99.3

99.2

81.5

94.0

94.3

78.8

78.5

78.8

93.0

92.3

92.2

77.2

84.9

76.7

B. cereus AF227848

94.1

99.7

99.6

99.7

99.7

94.0

97.6

94.1

99.7

93.6

78.7

99.3

99.2

81.7

94.0

94.3

78.9

78.6

78.8

92.7

92.0

91.8

77.9

85.0

76.7

AJ310098.seq

94.2

94.1

94.2

94.2

99.5

93.2

99.7

94.2

99.0

78.1

94.6

94.3

81.1

98.8

99.2

78.3

78.3

78.6

92.3

91.7

91.5

78.0

83.0

75.9

B. sonorensis AJ586363

99.9 100.0 100.0 94.1

96.8

94.2 100.0 93.8

78.1

99.3

99.2

81.7

94.0

94.3

78.8

78.5

78.8

93.0

92.3

92.2

78.4

84.9

76.7

B. cereus AY224379

94.1

96.8

94.2

93.7

78.0

99.3

99.2

81.7

94.0

94.3

78.7

78.4

78.7

92.9

92.3

92.1

78.3

84.8

76.7

B. cereus AY224383

100.0 94.1

96.8

94.2 100.0 93.8

78.1

99.3

99.2

81.7

94.0

94.3

78.8

78.5

78.8

93.0

92.3

92.2

78.4

84.9

76.7

B. cereus AY224385

94.1

96.8

94.2 100.0 93.8

78.1

99.3

99.2

81.7

94.0

94.3

78.8

78.5

78.8

93.0

92.3

92.2

78.4

84.9

76.7

AY224388.seq

91.9

99.8

94.0

99.3

77.9

94.5

94.2

81.0

98.7

99.1

78.3

78.3

78.6

92.3

91.8

91.6

78.2

83.0

75.6

B. licheniformis AY728013

92.3

98.2

94.3

81.0

99.4

99.3

81.2

93.7

94.0

81.8

81.8

83.2

93.0

93.1

92.4

80.3

86.5

79.4

10

B. Cereus DQ884352.seq

94.1

99.4

78.2

94.6

94.3

81.1

98.8

99.2

78.3

78.3

78.6

92.5

91.8

91.6

78.2

83.0

75.9

11

Bacillus sp.EF471917.seq

93.8

78.2

99.3

99.2

81.7

94.0

94.3

78.8

78.5

78.8

93.0

92.3

92.2

78.3

84.9

76.7

12

B. Cereus EU373359.seq

78.2

94.6

94.2

82.4

99.3

99.7

79.0

78.9

78.7

91.9

91.6

91.3

78.2

82.9

75.8

13

B. licheniformis EU373408.

82.1

82.0

98.4

81.1

81.3

98.6

99.9

96.8

77.4

77.3

76.7

94.1

75.9

84.8

14

Staph. haemolyticus EU652064

99.4

81.8

94.0

94.2

82.0

82.0

83.4

93.1

93.2

92.6

82.1

86.8

79.7

15

Fazal-1KWS4.seq

81.7

94.1

94.4

81.9

81.9

83.3

92.9

93.0

80.7

80.8
99.7

99.0
81.4

99.3
81.4

97.1
82.7

81.1
92.8

80.7
92.7

92.4
80.2

82.0
94.6

86.6
78.6

79.6
85.7

16
17

Fazal-2KWS2.seq
Fazal-3 10_1.seq

92.1

81.4

85.0

79.6

18

Fazal-4 DW3.seq

81.7

81.7

82.8

93.1

93.0

92.4

81.4

85.3

79.9

19

Fazal-5 10.seq

98.7

97.2

78.0

78.0

77.2

94.3

76.2

84.6

20

PSU26261.seq

96.9

77.5

77.4

76.8

94.3

75.9

84.8

21

P. stutzeri U26262

77.3

77.8

76.9

95.9

75.6

83.9

22

P. putida D37923.seq

97.8

97.8

76.8

82.4

76.3

23

Staph. aureus D83358

97.6

76.8

83.5

76.3

24

Staph. haemolyticus D83367

76.3

83.1

75.7

25

Staph. hyicus D83368

75.4

84.0

26

P. fluorescens Z76662

74.8

27

Strep. mutans X58303

28

E. coli X80725

0.2

0.3

1.0

6.2

6.7

6.0

0.2

1.4

0.3

6.0

0.3

1.4

0.3

6.1

0.1

0.2

1.4

0.3

6.0

0.0

0.1

0.2

1.4

0.3

6.0

0.0

0.1

0.0

6.4

7.6

6.2

0.5

6.1

6.2

6.1

6.1

1.1

3.4

2.4

7.2

3.3

3.3

3.3

3.3

8.6

6.2

7.4

6.1

0.3

6.0

6.1

6.0

6.0

0.2

8.2

0.2

0.5

0.3

6.0

0.0

0.1

0.0

0.0

6.2

1.8

6.2

6.7

6.7

6.7

0.9

6.6

6.6

6.6

6.6

0.7

6.0

0.6

6.6

26.0

26.3

25.2

26.0

26.1

26.2

26.1

26.1

26.3

22.0

25.8

26.0

26.0

0.7

0.7

0.7

5.6

0.7

0.7

0.7

0.7

5.7

0.6

5.6

0.7

5.6

20.5

11
12
13
14
15
16
17
10

99.9

99.9

99.9

0.8

0.8

0.8

6.0

0.8

0.8

0.8

0.8

6.1

0.7

6.0

0.8

6.1

20.6

0.6

18

20.3
5.7

20.7
6.3

20.5
6.3

21.2
1.2

20.5
6.3

20.5
6.3

20.5
6.3

20.5
6.3

21.4
1.3

21.1
6.6

21.2
1.2

20.5
6.3

20.1
0.7

0.9
22.0

20.3
6.3

20.5
6.2

21.9

19

5.4

5.8

5.8

0.7

5.8

5.8

5.8

5.8

0.8

6.2

0.7

5.8

0.3

21.5

6.0

5.8

21.7

0.2

20

25.0

25.0

24.8

25.7

25.0

25.1

25.0

25.0

25.8

21.0

25.8

25.0

24.8

1.4

20.7

20.8

0.7

21.5

21.2

21

25.5

25.5

25.3

25.7

25.5

25.6

25.5

25.5

25.8

21.0

25.8

25.5

24.9

0.1

20.7

20.8

0.5

21.5

21.2

1.3

22

25.1

25.1

25.0

25.3

25.1

25.2

25.1

25.1

25.4

19.2

25.4

25.1

25.2

3.2

18.9

19.0

2.9

19.8

19.7

2.8

3.2

23

7.4

7.4

7.7

8.1

7.4

7.5

7.4

7.4

8.2

7.5

8.0

7.4

8.6

27.0

7.3

7.6

21.6

7.7

7.3

26.1

26.9

27.2

24

8.2

8.2

8.5

8.8

8.2

8.3

8.2

8.2

8.7

7.3

8.7

8.2

9.0

27.3

7.2

7.4

22.2

7.8

7.5

26.2

27.1

26.6

2.2

8.3

8.3

8.7

9.1

8.3

8.4

8.3

8.3

9.0

8.1

9.0

8.3

9.4

28.2

7.9

8.2

23.0

8.4

8.1

27.5

28.0

27.8

2.3

2.5

25.6

26.7

25.6

25.5

25.0

25.1

25.0

25.0

25.3

22.7

25.4

25.2

25.2

5.6

20.3

20.4

4.7

21.3

21.1

5.4

5.4

3.6

27.4

27.4

28.1

14.7

14.7

14.6

17.1

14.7

14.8

14.7

14.7

17.2

13.2

17.2

14.7

17.3

26.9

12.7

13.0

23.2

15.0

14.6

26.5

26.9

27.4

17.9

16.5

17.0

27.0

27.9

27.8

27.8

29.0

27.8

27.8

27.8

27.8

29.5

24.0

29.0

27.8

29.2

16.8

23.5

23.6

15.5

23.6

23.3

17.1

16.9

17.9

28.4

28.4

29.4

17.2

28.3

10

11

12

13

14

15

16

17

18

19

20

21

22

23

24

25

26

27

26
27
28
25

28

Figure I b. Percent identity of bacterial isolates with other known identified bacterial strains
133

Appendix

(Continued on next page)


134

Appendix

Figure I c. nidAB gene sequence cloned from Mycobacterium PY146 and containing
the EcoRI and HindIII restriction enzyme adaptor sequences

135

Appendix

Ab und ance

Scan 2744 (18.758 min): 0201002.D\ d ata .ms (-2724) (-)


115.0

7000000

6500000

144.0

6000000
5500000
5000000

4500000
4000000

3500000
3000000
2500000

2000000
1500000

1000000

89.0

63.0

500000

39.0

72.0

50.0

98.0

30

40

50

60

70

80

90

100

126.0135.0
110

120

130

140

m/ z-->

B
100

144

115

50
HO
89

63
39

50

72

98

30 40 50 60
(replib) 2-Naphthalenol

70

80

126

90 100 110 120 130 140

Figure II a. Mass spectra obtained by GC/MS of 2-Naphthalenol produced by Bacillus


cereus KWS2 grown with naphthalene (A) and from NIST mass spectral library (B).

136

Appendix

Ab und a nce

A
Sca n 2046 (15.025 min): 0201002.D\ d a ta .ms (-2025) (-)
91.0
200000
180000
160000
140000
120000
100000
80000

135.9

60000
40000

65.0

20000

39.0

51.0

77.0

107.1 117.9126.0

0
30

40

50

60

70

80

90

100

110

120

130

140

m/ z-->

B
91

100
O
OH
50

136

39 45 51

65
77

30 40 50 60 70
(replib)
Benzeneacetic acid

80

90 100 110 120 130 140

Figure II b. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Bacillus cereus KWS2 grown with naphthalene (A) and from NIST mass spectral
library (B).

137

Appendix

Ab und a nc e

Sc a n 1797 (13.693 min): 0201002.D \ d a ta .ms


105.0

8000

121.9

7500
7000
76.9

6500
6000
5500
5000
4500
4000
3500
3000
2500

43.0

51.0

2000
1500
59.9

1000

69.1

500

91.0
84.1

0
30

35

40

45

50

55

60

65

70

75

80

85

90

95 100105110115120125

m/ z-->

105

100

122

77
HO

51
50

39
45

94

65

30
40
50
60
70
(mainlib) Benzenecarboxylic acid

80

90

100

110

120

Figure II c. Mass spectra obtained by GC/MS of Benzenecarboxylic acid produced by


Bacillus cereus KWS2 grown with naphthalene (A) and from NIST mass spectral
library (B).

138

Appendix

Ab und ance

Sca n 1153 (10.249 min): 0201002.D\ d ata .ms (-1141) (-)

130000

106.0

120000
110000

77.0

100000
90000

80000
70000
60000

51.0

50000
40000

30000
20000
39.0

10000

63.1

45.2

0
30

35

40

45

50

55

60

65

82.7 88.9
70

75

80

85

90

95.8
95 100 105

m/ z-->

B
106

77

100

51
50

30

39

74
43

30
40
50
(replib) Benzaldehyde

63
60

84 89
70

80

90

100

Figure II d. Mass spectra obtained by GC/MS of Bezaldehyde produced by Bacillus


cereus KWS2 grown with naphthalene (A) and from NIST mass spectral library (B).

139

Appendix

A
A b und a nc e
S c a n 4002 (25.487 m in): 0201002.D \ d a ta .m s (-4009) (-)
194.0
165.0

700000

650000

600000

550000

500000

450000

400000

350000

300000

250000

200000

150000

100000
82.1

139.0

50000
69.4
39.0

97.0

115.0

51.0

127.0

0
30

40

50

60

70

80

90

151.0

181.8

100 110 120 130 140 150 160 170 180 190

m/ z-->

B
194

100

165
50

37 49

69

82

OH
97

30
50
70
90
(mainlib)
9-Phenanthrenol

115 126 139 150


110

130

150

174
170

190

Figure II e. Mass spectra obtained by GC/MS of 9-phenanthrenol produced by


Bacillus cereus KWS2 grown in phenanthrene (A) and from NIST mass spectral
library (B).

140

Appendix

A b und a nc e

S c a n 3946 (25.187 m in): 0201002.D \ d a ta .ms (-3932) (-)


165.0
170000
160000
150000
140000
130000
120000
181.0
110000
212.0

100000
90000
80000
70000
60000

193.9

50000
152.0
40000
30000
20000

76.9

10000

115.0

63.0
49.9

138.9

101.9

0
30

40

50

60

70

80

90

100 110 120 130 140 150 160 170 180 190 200 210

m/ z-->

165

100

212

181

50
82
51 63
69
0

41

194

HO
OH
115

89

151
139

30
50
70
90 110 130 150 170
(mainlib)
9,10-Dihydro-9,10-dihydroxyphenanthrene

190

210

Figure II f. Mass spectra obtained by GC/MS of 9, 10-dihydro-9, 10dihydroxyphenanthrene produced by Bacillus cereus KWS2 grown in phenanthrene
(A) and from NIST mass spectral library (B).

141

Appendix

A b und a nc e

S c a n 2058 (15.089 min): 0201002.D \ d a ta .ms (-2025) (-)


91.0
400000
380000
360000
340000
320000
300000
280000
260000
240000
220000
200000
180000
160000
140000
135.9
120000
100000
80000
65.0
60000
40000

39.0
51.0

20000

77.0
98.8 106.9

0
30

40

50

60

70

80

90

100

110

118.0
120

128.2

130

m / z-->

B
91

100
O
OH
50

136

39 45 51

65
77

30 40 50 60 70
(replib)
Benzeneacetic acid

80

90 100 110 120 130 140

Figure II g. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Bacillus cereus KWS2 grown in phenanthrene (A) and from NIST mass spectral
library (B).

142

Appendix

Ab und a nce

Sca n 2856 (19.357 min): 0301003.D \ d a ta .ms (-2816) (-)


107.0

1100000
1000000
900000
800000
700000
600000
500000
400000

151.9

300000

77.0

200000
100000

51.0
39.0

63.0

89.0 98.0

0
30

40

50

60

70

80

90

129.0

116.9

100

110

120

141.2

130

140

150

m/ z-->

107

100

O
OH
50
152
77
0

31 39

51

OH

63

86 95

121

134

30 40 50 60 70 80 90 100 110 120 130 140 150


(mainlib)
Benzeneacetic acid, 4-hydroxy

Figure II h. Mass spectra obtained by GC/MS of 4-hydroxy benzeneacetic acid


produced by Bacillus cereus KWS2 grown in phenanthrene (A) and from NIST mass
spectral library (B).

143

Appendix

A
A b und a nc e
S c a n 1862 (14.041 min): 2301024.D \ d a ta .ms (-1871) (-)
105.0

95000
90000
85000

122.0
80000
75000
70000
65000
60000

77.0

55000
50000
45000
40000
35000
30000

51.0

25000
20000
15000
10000
39.0

5000

60.0
94.1

85.0

69.0

115.0

0
25

30

35

40

45

50

55

60

65

70

75

80

85

90

95 100105110115120

m/ z-->

105

100
HO

122

77
50
51

39 45

74

65

94

0
30
40
50
60
70
(replib) Benzenecarboxylic acid

80

90

100

110

120

Figure II i. Mass spectra obtained by GC/MS of benzenecarboxylic acid produced by


Bacillus cereus KWS2 grown in anthracene (A) and from NIST mass spectral library
(B).

144

Appendix

Ab und a nce

A
Sca n 2088 (15.250 min): 2301024.D\ d a ta .ms (-2075) (-)
91.0

220000
200000
180000
160000
140000
120000
100000

136.0

80000
60000

101.0

40000

65.0
39.0

20000

116.0

75.0

51.0

128.1

0
30

40

50

60

70

80

90

100

110

120

130

m/ z-->

91

100
O
OH
50

136
65

39 46 51
56

72 77

30 40 50 60 70
(replib)
Benzeneacetic acid

80

101
90

117

100 110 120 130 140

Figure II j. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Bacillus cereus KWS2 grown in anthracene (A) and from NIST mass spectral library
(B).

145

Appendix

A b und a nc e

S c a n 2856 (19.357 min): 2301024.D \ d a ta .ms (-2826) (-)


107.0
800000
750000
700000
650000
600000
550000
500000
450000
400000
350000
300000
152.0
250000
200000
77.0
150000
100000
51.1

50000

39.1

63.0

89.0

0
30

40

50

60

70

80

90

123.1
100

110

120

141.9

130

140

150

m/ z-->

107

100

O
OH
50
152
77
0

31 39

51

63

OH
86 95

121

134

30 40 50 60 70 80 90 100 110 120 130 140 150


(mainlib)
Benzeneacetic acid, 4-hydroxy

Figure II k. Mass spectra obtained by GC/MS of 4-hydroxy benzeneacetic acid


produced by Bacillus cereus KWS2 grown in anthracene (A) and from NIST mass
spectral library (B).

146

Appendix

A b und a nc e

S c a n 2740 (18.737 m in): 0601006.D \ d a ta .ms (-2721) (-)


115.0

3200000

144.0

3000000
2800000
2600000
2400000
2200000
2000000
1800000
1600000
1400000
1200000
1000000
800000
600000
89.0

400000

63.0

200000

39.0

74.0

50.0

98.0

126.0134.8

0
30

40

50

60

70

80

90

100

110

120

130

140

m / z-->

B
100

144

115

50
HO
89

63
39

50

72

98

0
30 40 50 60
(replib) 2-Naphthalenol

70

80

126

90 100 110 120 130 140 150

Figure III a. Mass spectra obtained by GC/MS of 2-naphthalenol produced by


Bacillus cereus KWS4 grown with naphthalene (A) and from NIST mass spectral
library (B).

147

Appendix

A b und a nc e
S c a n 2045 (15.020 min): 0601006.D \ d a ta .ms (-2021) (-)
91.0

180000

170000
160000
150000
140000
130000
120000
110000
100000
90000
80000
70000

135.9

60000
50000
40000
65.0

30000
20000
39.0
51.0

10000

76.9

99.8 108.0 118.0 127.9

0
30

40

50

60

70

80

90

100

110

120

130

140

m/ z-->

B
91

100
O
OH

136

50
65
39
44

51

77

105

0
30 40 50 60 70
(mainlib) Benzeneacetic acid

80

90

118

100 110 120 130 140

Figure III b. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Bacillus cereus KWS4 grown with naphthalene (A) and from NIST mass spectral
library (B).

148

Appendix

Ab und a nce

A
Sca n 1151 (10.238 min): 0601006.D \ d a ta .ms (-1140) (-)
106.0
77.0

20000
18000
16000
14000
12000
10000
51.0
8000
6000
4000
2000

39.0
62.9

45.0

0
30

35

40

45

50

55

60

88.0

70.9

65

70

75

80

85

90

95

100 105

m/ z-->

B
106

77

100

51
50

30

39

74
43

30
40
50
(replib) Benzaldehyde

63
60

84 89
70

80

90

100

Figure III c. Mass spectra obtained by GC/MS of Bezaldehyde produced by Bacillus


cereus KWS4 grown with naphthalene (A) and from NIST mass spectral library (B).

149

Appendix

A
Ab und a nc e

S c a n 4002 (25.487 min): 0401004.D \ d a ta .ms (-4012) (-)


165.0

194.0

1050000
1000000
950000
900000
850000
800000
750000
700000
650000
600000
550000
500000
450000
400000
350000
300000
250000
200000
150000

82.2

139.0

100000
69.4

50000

97.0

39.0 51.0
0
30

40

50

60

70

80

90

115.0
127.0

150.9

179.9

100 110 120 130 140 150 160 170 180 190

m/ z-->

165

100

194

50
82
39 50

69

OH

87 97

0
30
50
70
(replib) 9-Phenanthrenol

90

115 126
110

139

130

150
150

174
170

190

Figure III d. Mass spectra obtained by GC/MS of 9-phenanthrenol produced by


Bacillus cereus KWS4 grown in phenanthrene (A) and from NIST mass spectral
library (B).

150

Appendix

A
A b und a nc e
S c a n 3947 (25.193 min): 0401004.D \ d a ta .m s (-3931) (-)
165.0

200000
190000
180000
170000
160000
150000
140000

181.0
130000
120000
212.0
110000
100000
90000
80000
70000
60000
151.9

50000
40000
30000
77.0

20000

194.9

115.0
138.9

63.0

10000
50.0
0

101.8

36.9
30

40

50

60

70

80

90

100 110 120 130 140 150 160 170 180 190 200 210

m/ z-->

B
165

100

212

181

50
82
51 63
69
0

41

194

HO

89

OH
115

151
139

30
50
70
90 110 130 150 170
(mainlib) 9,10-Dihydro-9,10-dihydroxyphenanthrene

190

210

Figure III f. Mass spectra obtained by GC/MS of 9,10-dihydro-9,10dihydroxyphenanthrene produced by Bacillus cereus KWS4 grown in phenanthrene
(A) and from NIST mass spectral library (B).

151

Appendix

A
Ab und a nc e
Sc a n 2065 (15.127 min): 0401004.D \ d a ta .ms (-2026) (-)
91.0
600000
550000
500000
450000
400000
350000
300000
250000
135.9

200000
150000
65.0

100000
39.0

50000

51.0

77.0

107.0

0
30

40

50

60

70

80

90

100

110

117.8 126.8
120

130

m/ z-->

B
91

100
O
OH
50

136

39 45 51

65
77

30 40 50 60 70
(replib) Benzeneacetic acid

80

90

100 110 120 130

Figure III g. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Bacillus cereus KWS4 grown in phenanthrene (A) and from NIST mass spectral
library (B).

152

Appendix

A
Ab und a nc e

Sc a n 2841 (19.277 min): 0501005.D \ d a ta .ms (-2810) (-)


107.0
170000
160000
150000
140000
130000
120000
110000
100000
90000
80000
70000
151.9

60000
50000
77.0
40000
30000

63.0
51.0

20000

86.9

125.9

39.0

10000

97.9

0
30

40

50

60

70

80

90

100

134.9

116.0
110

120

130

140

150

m/ z-->

B
107

100

O
OH152
50
77
39
0

30

51

HO
63

90

124 134

30 40 50 60 70 80 90 100 110 120 130 140 150


(replib)
Benzeneacetic acid, 3-hydroxy

Figure III h. Mass spectra obtained by GC/MS of 3-hydroxy benzeneacetic acid


produced by Bacillus cereus KWS4 grown in phenanthrene (A) and from NIST mass
spectral library (B).

153

Appendix

Ab und a nc e

Sca n 2218 (15.945 min): 0501005.D \ d a ta .ms (-2210) (-)


92.0
119.9
60000
55000
50000
45000
40000
137.9

35000
30000
25000
20000

64.0
43.0

15000
10000

53.0

72.0

5000
0

80.8

105.1

35.1
30

40

50

60

70

80

90

100

110

127.9
120

130

140

m/ z-->

B
92

100

120
O
OH
138

50

OH
64
39
45

53

74 81

109

0
30 40 50 60 70 80
(replib) Benzoic acid, 2-hydroxy-

90

100 110 120 130 140

Figure III i. Mass spectra obtained by GC/MS of 2-hydroxy benzoic acid produced by
Bacillus cereus KWS4 grown in phenanthrene (A) and from NIST mass spectral
library (B).

154

Appendix

Abundance

Scan 1857 (17.014 min): 2001005.D\ data.ms


104.0

600000
550000
500000
450000

76.0
400000
350000
300000
250000
200000
50.0

150000

148.0

100000
50000

38.0
61.0

85.0

0
30

40

50

60

70

80

116.0

94.8

90

100

110

120

128.9 138.0
130

140

m/ z-->

B
104

100

76

OH
50

OH

50
O

31

38

61

148

85

30 40 50 60 70 80 90 100 110 120 130 140 150


(rep
lib
) 1,2-Benzened icarb oxylic acid

Figure IV a. Mass spectra obtained by GC/MS of 1, 2-benzenecarboxylic acid


produced by resting cells of Mycobacterium PY146 with phenanthrene (A) and from
NIST mass spectral library (B).

155

Appendix

Abundance

A
Scan 1865 (17.057 min): 2201007.D\ data.ms (-1849) (-)
104.0

30000
28000
26000
24000
22000

76.0

20000
18000
16000
14000
12000
10000
8000

148.0

50.1

6000
4000
37.2

2000

60.8

30

40

50

60

116.7 126.9

93.0

0
70

80

90

100

110

120

130

140

m/ z-->

104

100

76

OH
50

OH

50
O

31

38

61

148

85

30 40 50 60 70 80 90 100 110 120 130 140 150


(rep lib ) 1,2-Benzened icarb oxylic a cid

Figure IV b. Mass spectra obtained by GC/MS of 1,2-benzenecarboxylic acid


produced by resting cells of Mycobacterium PY146 with anthracene (A) and from
NIST mass spectral library (B).

156

Appendix

Abundance

Scan 3359 (25.048 m


in): 2101006.D\ data.m
s
208.0

160000
150000
140000
180.0

130000
120000
110000
100000

152.0

90000
80000
70000
60000
50000
40000
76.0
30000
20000
50.0

10000
0

126.0
63.0

90.0

111.0

30

40

50

60

70

80

90

194.0

165.0

139.0

36.8

100 110 120 130 140 150 160 170 180 190 200

m
/ z-->

208

100
180

152

50
76
O

39

50 63

90

99

30
50
70
90
110
(replib)
9,10-Anthracenedione

126
130

163
150

170

190

210

Figure IV c. Mass spectra obtained by GC/MS of 1,2-benzenecarboxylic acid


produced by resting cells of Mycobacterium PY146 with anthracene (A) and from
NIST mass spectral library (B).

157

Appendix

Abundance

Scan 1861 (17.035 min): 2401009.D\ data.ms (-1849) (-)


104.0
40000

35000

30000
76.0

25000

20000

15000

10000

50.0

148.0

5000
38.1
61.1

87.1

113.0

0
30

40

50

60

70

80

90

100

110

124.9 133.7
120

130

140

m/ z-->

104

100

76

OH
50

OH

50
O

31

38

61

148

85

30 40 50 60 70 80 90 100 110 120 130 140 150


(replib)
1,2-Benzenedicarb oxylic acid

Figure IV d. Mass spectra obtained by GC/MS of 1,2-benzenecarboxylic acid


produced by resting cells of Mycobacterium PY146 with pyrene (A) and from NIST
mass spectral library (B).

158

Appendix

Abundance

Scan 4817 (32.846 m


in): 2301008.D\ data.m
s
218.0
24000
23000
22000
21000
20000
19000
18000
17000
16000
15000
14000
13000
12000
11000

189.0

10000
9000
8000
7000
6000
5000
4000
3000
43.0
2000
1000

57.1

72.9

94.1

111.0

130.0

147.0

163.1

203.0

0
30

40

50

60

70

80

90

100 110 120 130 140 150 160 170 180 190 200 210 220

m
/ z-->

218

100
OH
189

50

94

81

109

163

0
30
50
70
90
(rep
lib
)
1-Hyd
roxypyrene

110

130

150

170

190

210

Figure IV e. Mass spectra obtained by GC/MS of 1-hydroxypyrene produced by


resting cells of Mycobacterium PY146 with pyrene (A) and from NIST mass spectral
library (B).

159

Appendix

Ab und a nc e

2600000

Sc a n 2241 (16.068 min): 1701018.D \ d a ta .ms (-2216) (-)


104.0

2400000
2200000
76.0

2000000
1800000
1600000
1400000
1200000
1000000
50.1
800000
600000

148.0

400000
38.1

200000

61.0

85.0

0
30

40

50

60

70

80

94.9

90

118.9 129.0 139.0

100

110

120

130

140

m/ z-->

104

100

76

OH
50

OH

50
O

31

38

61

148

85

30 40 50 60 70 80 90 100 110 120 130 140 150


(replib) 1,2-Benzenedicarboxylic acid

Figure V a. Mass spectra obtained by GC/MS of 1,2-benzenedicarboxylic acid


produced by Mycobacterium PY146 grown in phenanthrene (A) and from NIST mass
spectral library (B).

160

Appendix

Ab und a nce

A
450000

Sca n 2068 (15.142 min): 1901020.D\ d a ta .ms (-2034) (-)


91.0

400000
350000
300000
250000
200000
136.0

150000
100000
65.0
50000

39.1

51.0

77.0

98.9 107.0

0
30

40

50

60

70

80

90

100

110

117.9
120

130

m/ z-->

91

100
O
OH
50

136

39 45 51

65

30 40 50 60 70
(replib)
Benzeneacetic acid

77
80

90

100 110 120 130

Figure V b. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Mycobacterium PY146 grown in phenanthrene (A) and from NIST mass spectral
library (B).

161

Appendix

A b und a nc e

Sc a n 3709 (23.920 min): 3001031.D \ d a ta .ms (-3691) (-)


180.0

90000

208.0

85000
80000
75000
70000
65000

152.0

60000
55000
50000
45000
40000
35000
30000
76.0
25000
20000
15000
50.0
10000

126.0
63.0

5000
0

97.9
111.0

37.0
30

40

50

60

70

80

90

194.9

100 110 120 130 140 150 160 170 180 190 200

m/ z-->

152

100

180

208

76

50

50
63
0

39

126
90 102

30
50
70
90 110
(replib) 9,10-Anthracenedione

130

150

170

190

210

Figure V c. Mass spectra obtained by GC/MS of 9,10-anthracenedione produced by


Mycobacterium PY146 grown in anthracene (A) and from NIST mass spectral library
(B).

162

Appendix

Ab und a nce

A
Sca n 1834 (13.891 min): 3101032.D \ d a ta .ms (-1788) (-)
105.0

50000
45000

122.0

40000
77.0

35000
30000
25000
20000
51.0
15000
10000

43.1

87.0
60.0

5000

94.0

70.0

115.1

0
25 30 35 40 45 50 55 60 65 70 75 80 85 90 95 100105110115120
m/ z-->

105

100
HO

122

77
50
51

39 45

65

74

94

0
30
40
50
60
70
(replib)
Benzenecarboxylic
acid

80

90

100

110

120

Figure V d. Mass spectra obtained by GC/MS of benzenecarboxylic acid produced by


Mycobacterium PY146 grown in anthracene (A) and from NIST mass spectral library
(B).

163

Appendix

Ab und a nc e

Sc a n 2080 (15.207 min): 3101032.D \ d a ta .ms (-2031) (-)


91.0

700000
650000
600000
550000
500000
450000
400000
350000
300000

136.0

250000
200000
150000
65.0

100000
39.1

50000

51.0
77.0

99.0 107.0

0
30

40

50

60

70

80

90

100

110

118.0 127.8
120

130

m/ z-->

91

100
O
OH
50

136

39 45 51

65

30 40 50 60 70
(replib) Benzeneacetic acid

77
80

90

100 110 120 130

Figure V e. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Mycobacterium PY146 grown in anthracene (A) and from NIST mass spectral library
(B).

164

Appendix

A b und a nc e

S c a n 2234 (16.030 min): 3101032.D \ d a ta .ms (-2218) (-)


104.0

200000
190000
180000
170000
160000
150000

76.0

140000
130000
120000
110000
100000
90000
80000
70000
50.0

60000
50000
40000

147.9

30000
20000
38.0
10000
61.0
0
20

30

40

50

60

112.9122.1

92.1
70

80

90

100

110

120

136.0

130

140

150

m/ z-->

B
104

100

76

OH
50

OH

50
O

31

38

61

148

85

30 40 50 60 70 80 90 100 110 120 130 140 150


(replib) 1,2-Benzenedicarboxylic acid

Figure V f. Mass spectra obtained by GC/MS of 1,2-benzenedicarboxylic acid


produced by Mycobacterium PY146 grown in anthracene (A) and from NIST mass
spectral library (B).

165

Appendix

Ab und a nc e

Sc a n 1614 (12.956 min): 0501005.D \ d a ta .ms (-1598) (-)


105.0

122.0

55000

50000

45000
77.0
40000

35000

30000

25000

20000
51.1

15000

87.0

10000
60.0
5000
94.0

69.0

113.0

0
45

50

55

60

65

70

75

80

85

90

95

100 105 110 115 120

m/ z-->

105

100

122

77
HO

51
50

65

74
90

0
50
60
70
80
(mainlib) Benzenecarboxylic acid

90

94
100

110

120

Figure V g. Mass spectra obtained by GC/MS of benzenecarboxylic acid produced by


Mycobacterium PY146 grown in pyrene (A) and from NIST mass spectral library (B).

166

Appendix

Ab und a nce

Sc a n 1887 (14.373 min): 0501005.D \ d a ta .ms


91.1

900000

800000

700000

600000

500000

400000
136.0
300000

200000
65.1
79.1

100000
51.1
58.0

98.1

72.1

122.0
107.1 115.0
129.1

40 45 50 55 60 65 70 75 80 85 90 95 100105110115120125130135

m/ z-->

B
91

100

O
OH136
50
65
51

63

89

77

105

0
50
60
70
80
(mainlib)
Benzeneacetic
acid

90

100

110

118
120

130

140

Figure V h. Mass spectra obtained by GC/MS of benzeneacetic acid produced by


Mycobacterium PY146 grown in pyrene (A) and from NIST mass spectral library (B).

167

Appendix

Ab und a nc e

Sc a n 2014 (15.032 min): 0501005.D \ d a ta .ms (-1993) (-)


104.0
1300000
1200000
1100000
1000000

76.0

900000
800000
700000
600000
500000
50.1
400000
300000

148.0

200000
100000
61.0

85.0

0
40

50

60

70

80

118.9127.1 136.8

93.0

90

100

110

120

130

140

m/ z-->

104

100

76

OH
50

50

OH
148
O

61 66

85

50 60 70 80 90 100 110 120


(replib) 1,2-Benzenedicarboxylic acid

130 140 150

Figure V i. Mass spectra obtained by GC/MS of 1,2-benzenedicarboxylic acid


produced by Mycobacterium PY146 grown in pyrene (A) and from NIST mass
spectral library (B).

168

You might also like